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

Near soliton evolution for 22-equivariant Schrödinger Maps in two space dimensions

Ioan Bejenaru Department of Mathematics, University of California, San Diego [email protected] Mohandas Pillai Department of Mathematics, University of California, San Diego [email protected]  and  Daniel Tataru Department of Mathematics, University of California, Berkeley [email protected]
Abstract.

We consider equivariant solutions for the Schrödinger Map equation in 2+12+1 dimensions, with values into 𝕊2\mathbb{S}^{2}. Within each equivariance class mm\in\mathbb{Z} this admits a lowest energy nontrivial steady state QmQ^{m}, which extends to a two dimensional family of steady states by scaling and rotation. If |m|3|m|\geq 3 then these ground states are known to be stable in the energy space H˙1\dot{H}^{1}, whereas instability and even finite time blow-up along the ground state family may occur if |m|=1|m|=1. In this article we consider the most delicate case |m|=2|m|=2. Our main result asserts that small H˙1\dot{H}^{1} perturbations of the ground state Q2Q^{2} yield global in time solutions, which satisfy global dispersive bounds. Unlike the higher equivariance classes, here we expect solutions to move arbitrarily far along the soliton family; however, we are able to provide a time dependent bound on the growth of the scale modulation parameter. We also show that within the equivariant class the ground state is stable in a slightly stronger topology XH˙1X\subset\dot{H}^{1}.

Key words and phrases:
Schrödinger maps, soliton stability, blow-up, local energy decay
1991 Mathematics Subject Classification:
Primary: 35Q41, 35Q55 Secondary: 35B40

1. Introduction

In this article we consider the Schrödinger map equation in 2+1\mathbb{R}^{2+1} with values into 𝕊2\mathbb{S}^{2},

(1.1) ut=u×Δu,u(0)=u0.u_{t}=u\times\Delta u,\qquad u(0)=u_{0}.

This equation admits a conserved energy,

E(u)=122|u|2𝑑x,E(u)=\frac{1}{2}\int_{\mathbb{R}^{2}}|\nabla u|^{2}dx,

and is invariant with respect to the dimensionless scaling

u(t,x)u(λ2t,λx).u(t,x)\to u(\lambda^{2}t,\lambda x).

The energy is invariant with respect to the above scaling, therefore the Schrödinger map equation in 2+1\mathbb{R}^{2+1} is energy critical.

In reviewing the literature about this problem, we note that (1.1) can be generalized in several ways. The simplest one is by considering maps u:n+1𝕊2u:\mathbb{R}^{n+1}\rightarrow\mathbb{S}^{2} (keeping in mind that n=2n=2 renders the energy critical dimension). One can also generalize the problem by replacing the target manifold 𝕊2\mathbb{S}^{2} with a Kähler manifold with a complex structure.

Local solutions for regular large initial data have been constructed in [34] and [27]. When the target manifold 𝕊2\mathbb{S}^{2} is replaced with a Kähler manifold with a compatible complex structure, local well-posedness for regular data has been established in [10], see also [8] and [18].

The global in time problem is a very difficult one, except in dimension n=1n=1, that is for maps u:1+1𝕊2u:\mathbb{R}^{1+1}\rightarrow\mathbb{S}^{2} (or some generalization of the base/target manifold), where it becomes energy subcritical; large data global well-posedness in this case has been established in [9].

The definitive result for the small data problem, for maps u:n+1𝕊2u:\mathbb{R}^{n+1}\rightarrow\mathbb{S}^{2}, with n2n\geq 2, was obtained by two of the authors and collaborators in [5], following earlier results in [4], [3], [2], [15], [16], [20], [17], [19], [30], [31], [29]. There global well-posedness and scattering were proved for initial data which is small in the critical Sobolev space H˙n2\dot{H}^{\frac{n}{2}}, which agrees with the energy space H˙1\dot{H}^{1} when n=2n=2. The counterpart of these results when 𝕊2\mathbb{S}^{2} is replaced with a more general Khäler manifold was established more recently in [24] and [25]. The same problem with the base space 2\mathbb{R}^{2} replaced with the hyperbolic plane was considered in [23], where the authors establish the asymptotic stability of (a large class of) harmonic maps under the Schrödinger maps evolution.

We now return to the setup of maps u:2+1𝕊2u:\mathbb{R}^{2+1}\rightarrow\mathbb{S}^{2} which is the energy critical one. The space of finite energy states for this problem is the space of H˙1\dot{H}^{1} maps from 2\mathbb{R}^{2} into 𝕊2\mathbb{S}^{2}, which separates into connected components according to the homotopy class mm\in\mathbb{Z}. Within each homotopy class there exists an energy minimizer, called the ground state, which is unique up to symmetries, namely scaling and isometries of the base space 2\mathbb{R}^{2} and of the target space 𝕊2\mathbb{S}^{2}. For each integer mm we denote the corresponding ground state family by 𝒬m\mathcal{Q}^{m}. Then a natural question is whether the ground states are stable with respect to the Schrödinger map flow.

The ground state family for the trivial m=0m=0 homotopy class consists of constant maps, which have energy 0. All the global results mentioned above (in the case n=2n=2) required small energy; as a consequence those maps have trivial topology and, with respect to the energy norm, they can be seen a small perturbations of these constant maps.

For m0m\neq 0, the generator QmQ^{m} for the ground state family can be taken to belong to the class of mm-equivariant maps, which satisfy

u(Rx)=Rmu(x),u(Rx)=R^{m}u(x),

where RR stands for rotations around the origin in the plane, and around the vertical axis on the sphere, with the same angle. The class of mm-equivariant maps is closed with respect to the Schrödinger map flow, therefore it is natural to restrict the above stability question of 𝒬m\mathcal{Q}^{m} to the class of mm-equivariant maps.

As it turns out, the answer to this question depends on the equivariance class mm:

  • If |m|3|m|\geq 3, then the ground state is stable in H˙1\dot{H}^{1} within the equivariant class; this was proved in [13] for |m|4|m|\geq 4, and in [14] for |m|=3|m|=3.

  • If |m|=1|m|=1, then the ground state is unstable in H˙1\dot{H}^{1} within the equivariant class; this was proved in [6]. Furthermore, finite time blow-up may also occur along the ground state family, see [28] and [32].

This left open the case |m|=2|m|=2, which is in some sense borderline and does not fit either of the two patterns above.

The objective of the present article is to investigate the ground state stability exactly in this remaining case |m|=2|m|=2 within the ±2\pm 2 equivariance class. In brief, our main results are as follows:

  • Small H˙1\dot{H}^{1} perturbations of the ground state yield global in time solutions.

  • The solutions satisfy universal dispersive and local energy bounds.

  • The solutions can move along the ground state manifold, but we prove quantitative time dependent bounds on the modulation parameters.

  • The ground state is stable in a slightly stronger topology.

1.1. Homotopy classes and the ground state

Given a H˙1\dot{H}^{1} map u:2𝕊2u:\mathbb{R}^{2}\to\mathbb{S}^{2}, its homotopy class mm\in\mathbb{Z} is defined as

m=14π2u(xu×yu)𝑑xm=\frac{1}{4\pi}\int_{\mathbb{R}^{2}}u\cdot(\partial_{x}u\times\partial_{y}u)\,dx

and roughly speaking counts the number of times the plane wraps around the sphere, taking also the orientation into account.

The homotopy classes generate a partition of the family of all H˙1\dot{H}^{1} maps into connected components. Within each homotopy class, one may look for energy minimizers, which are called ground states. The Euler-Lagrange equation for ground states is the harmonic map equation,

Δu=u|u|2,-\Delta u=u|\nabla u|^{2},

so the ground states are in particular harmonic maps and also steady states for the Schrödinger map equation.

For each nonzero integer mm there exists a family 𝒬m\mathcal{Q}^{m} of ground states. To describe these families we begin with the maps QmQ^{m} defined in polar coordinates by

(1.2) Qm(r,θ)=emθRQ¯m(r),Q¯m(r)=(h1m(r)0h3m(r)),m{0}Q^{m}(r,\theta)=e^{m\theta R}\bar{Q}^{m}(r),\qquad\bar{Q}^{m}(r)=\left(\begin{array}[]{ccc}h_{1}^{m}(r)\\ 0\\ h_{3}^{m}(r)\end{array}\right),\qquad m\in\mathbb{Z}\setminus\{0\}

with

h1m(r)=2rmr2m+1,h3m(r)=r2m1r2m+1.h_{1}^{m}(r)=\frac{2r^{m}}{r^{2m}+1},\qquad h_{3}^{m}(r)=\frac{r^{2m}-1}{r^{2m}+1}.

Here RR is the generator of horizontal rotations, which can be interpreted as a matrix or, equivalently, as the operator below

R=(010100000),Ru=k×u.R=\left(\begin{array}[]{ccc}0&-1&0\\ 1&0&0\\ 0&0&0\end{array}\right),\qquad Ru=\overrightarrow{k}\times u.

The families 𝒬m\mathcal{Q}^{m} are constructed from QmQ^{m} via the symmetries of the problem, namely scaling and orientation preserving isometries of the base space 2\mathbb{R}^{2} and of the target space 𝕊2\mathbb{S}^{2}. 𝒬m\mathcal{Q}^{-m} is equivalent to 𝒬m\mathcal{Q}^{m} modulo one reflection. The elements of 𝒬m\mathcal{Q}^{m} are harmonic maps from 2\mathbb{R}^{2} into 𝕊2\mathbb{S}^{2}, and admit a variational characterization as the unique energy minimizers, up to symmetries, among all maps u:2𝕊2u:\mathbb{R}^{2}\to\mathbb{S}^{2} within their homotopy class.

In the above context, a natural question is to study Schrödinger maps for which the initial data is close in H˙1\dot{H}^{1} to one of the 𝒬m\mathcal{Q}^{m} families. One may try to think of this as a small data problem, but in some aspects it turns out to be closer to a large data problem. Studying this in full generality is very difficult. In this article we confine ourselves to a class of maps which have some extra symmetry properties, namely the equivariant Schrödinger maps. These are indexed by an integer mm called the equivariance class, and consist of maps of the form

(1.3) u(r,θ)=emθRu¯(r).u(r,\theta)=e^{m\theta R}\bar{u}(r).

In particular the maps QmQ^{m} above are mm-equivariant. The case m=0m=0 would correspond to spherical symmetry. Restricted to equivariant functions the energy has the form

(1.4) E(u)=π0(|ru¯(r)|2+m2r2(u¯12(r)+u¯22(r)))r𝑑r.E(u)=\pi\int_{0}^{\infty}\left(|\partial_{r}\bar{u}(r)|^{2}+\frac{m^{2}}{r^{2}}(\bar{u}_{1}^{2}(r)+\bar{u}_{2}^{2}(r))\right)rdr.

In the case m0m\neq 0 this implies that the mm-equivariant maps have the property that u¯(0):=limr0u¯(r)\bar{u}(0):=\lim_{r\rightarrow 0}\bar{u}(r) and u¯():=limru¯(r)\bar{u}(\infty):=\lim_{r\rightarrow\infty}\bar{u}(r) are well-defined distinct elements of the set {±k}\{\pm\vec{k}\}. In particular, if m0m\neq 0 this implies that m-equivariant maps have a limit at 0 and at infinity, both in the set {±k}\{\pm\vec{k}\}. The homotopy class of mm-equivariant maps depends on whether the two limits are equal:

  1. (1)

    If u¯(0)=u¯()\bar{u}(0)=\bar{u}(\infty) then the map uu is topologically trivial, i.e. has homotopy 0.

  2. (2)

    If u¯(0)=k\bar{u}(0)=-\vec{k} and u¯()=k\bar{u}(\infty)=\vec{k} then the map uu has homotopy mm.

  3. (3)

    If u¯(0)=k\bar{u}(0)=\vec{k} and u¯()=k\bar{u}(\infty)=-\vec{k} then the map uu has homotopy m-m.

To fix the notations in the sequel we will work with the connected component of mm-equivariant maps with homotopy mm, i.e. with mm-equivariant maps satisfying

(1.5) u¯(0)=k,u¯()=k.\bar{u}(0)=-\vec{k},\qquad\bar{u}(\infty)=\vec{k}.

Intersecting the full set 𝒬m\mathcal{Q}^{m} with the mm-equivariant class and with the homotopy class of QmQ^{m} we obtain the two parameter family 𝒬em\mathcal{Q}^{m}_{e} generated from QmQ^{m} by rotations and scaling,

𝒬em={Qα,λm;α/2π,λ+},Qα,λm(r,θ)=emαRQm(λr,θ)\mathcal{Q}^{m}_{e}=\{Q^{m}_{\alpha,\lambda};\alpha\in\mathbb{R}/2\pi\mathbb{Z},\lambda\in\mathbb{R}^{+}\},\qquad Q^{m}_{\alpha,\lambda}(r,\theta)=e^{m\alpha R}Q^{m}(\lambda r,\theta)

Here Q0,1m=QmQ^{m}_{0,1}=Q^{m}. Their energy depends on mm as follows:

E(Qα,λm)=4π|m|:=E(𝒬m).E(Q^{m}_{\alpha,\lambda})=4\pi|m|:=E(\mathcal{Q}^{m}).

1.2. A brief history of the problem

The study of equivariant Schrödinger maps for mm-equivariant initial data close to 𝒬em\mathcal{Q}^{m}_{e} was initiated by Gustafson, Kang, Tsai in [12], [13], and continued by Gustafson, Nakanishi, Tsai in [14]. The energy conservation suffices to confine solutions to a small neighborhood of 𝒬em\mathcal{Q}^{m}_{e} due to the inequality (see [12])

(1.6) distH˙1(u,𝒬em)2=infα,λQα,λmuH˙12E(u)E(𝒬m),\text{dist}_{\dot{H}^{1}}(u,\mathcal{Q}^{m}_{e})^{2}=\inf_{\alpha,\lambda}\|Q^{m}_{\alpha,\lambda}-u\|_{\dot{H}^{1}}^{2}\lesssim E(u)-E(\mathcal{Q}^{m}),

which holds for all mm-equivariant maps u:2𝕊2u:\mathbb{R}^{2}\to\mathbb{S}^{2} in the homotopy class of 𝒬em\mathcal{Q}^{m}_{e} with 0E(u)E(𝒬m)10\leq E(u)-E(\mathcal{Q}^{m})\ll 1. One can interpret this as an orbital stability result for 𝒬em\mathcal{Q}^{m}_{e}. However, this does not say much about the global behavior of solutions since these soliton families are noncompact; thus one might have even finite time blow-up while staying close to a soliton family.

To track the evolution of an mm-equivariant Schrödinger map u(t)u(t) along 𝒬em\mathcal{Q}^{m}_{e} we use functions (α(t),λ(t))(\alpha(t),\lambda(t)) describing trajectories in 𝒬em\mathcal{Q}^{m}_{e}. A natural choice is to choose them as minimizers for the infimum in (1.6); while this is feasible and useful for the local theory, see [12], it is much less helpful for the purpose of the global theory. Instead, we will allow ourselves more freedom, and be content with any choice (α(t),λ(t))(\alpha(t),\lambda(t)) satisfying

(1.7) uQα(t),λ(t)mH˙12E(u)E(𝒬m).\|u-Q_{\alpha(t),\lambda(t)}^{m}\|_{\dot{H}^{1}}^{2}\lesssim E(u)-E(\mathcal{Q}^{m}).

An important preliminary step in this analysis is the next result concerning both the local wellposedness in H˙1\dot{H}^{1} and the persistence of higher regularity:

Theorem 1.1.

Let |m|1|m|\geq 1. There exists δ>0\delta>0 and σ>0\sigma>0 such that the following holds true. Given an mm-equivariant initial data u0u_{0} in the homotopy class of 𝒬em\mathcal{Q}^{m}_{e} and with E(u0)4π|m|+δ2E(u_{0})\leq 4\pi|m|+\delta^{2}, let λ0=λ(u0)\lambda_{0}=\lambda(u_{0}) be the choice of λ\lambda which (in combination with α\alpha) achieves the infimum in (1.6). Then the equation (1.1) with initial data u0u_{0} is locally well-posed in H˙1\dot{H}^{1} on a time interval I=[σλ02,σλ02]I=[-\frac{\sigma}{\lambda_{0}^{2}},\frac{\sigma}{\lambda_{0}^{2}}], and λλ0\lambda\approx\lambda_{0} in II.

If, in addition, u0H˙2u_{0}\in\dot{H}^{2} then uCt(I:H˙2)u\in C_{t}(I:\dot{H}^{2}). Furthermore, the H˙1H˙2\dot{H}^{1}\cap\dot{H}^{2} regularity persists for as long as the function λ(t)\lambda(t) in (1.7) remains in a compact set. If TmaxT_{max} is the maximal time for which there is a unique solution u(t)C([0,Tmax):H˙2)u(t)\in C([0,T_{max}):\dot{H}^{2}), then we have that Tmaxσλ02T_{\max}\geq\frac{\sigma}{\lambda_{0}^{2}}. In addition, if TmaxT_{max} is finite then limtTmaxλ(t)=\lim_{t\nearrow T_{max}}\lambda(t)=\infty.

This follows from Theorem 1.1 in [12] and Theorem 1.41.4 in [13]. Given the above result, the main remaining problem is to understand whether the steady states Qα,λmQ_{\alpha,\lambda}^{m} are stable or not; in the latter case, one would like to understand the dynamics of the motion of the solutions along the soliton family. The steady states turn out to be stable in the case of large mm, which was considered in prior work:

Theorem 1.2 ([13] for |m|4|m|\geq 4, [14] for |m|=3|m|=3).

The solitons Qα,λmQ_{\alpha,\lambda}^{m} are asymptotically stable in the H˙1\dot{H}^{1} topology within the mm-equivariant class.

However, in the cases |m|=1,2|m|=1,2 the behavior is expected to be different in nature. The |m|=1|m|=1 case was considered by two of the authors in [6]. There it is shown that the ground states are in general unstable in the energy topology, but on the other hand that they are stable with respect to perturbations which are small in a stronger topology. Further, Merle, Raphael and Rodnianski constructed solutions which blow up in finite time along the ground state family, see [28]; soon after, Perelman provided an alternative construction of singularity formation in [32].

In this article we consider the case m=2m=2, where there are no analogue results to the ones just mentioned for in the case m=1m=1 or m3m\geq 3; in other words no blow-up construction is known, and no stability regimes have been established.

Our first result will be to prove the global in time well-posedness of the equivariant Schrödinger Maps with data near Q2Q^{2}; as part of the proof, we establish global bounds on the potential growth of the noncompact modulation parameter λ(t)\lambda(t), and thus on the growth of higher regularity norms. Our result leaves open the possibility of blow-up/relaxation in infinite time, which will be considered in a forthcoming paper.

Our second result identifies stability regimes and it is similar in spirit to the one obtained in [6] in the case m=1m=1. It essentially says that if the original data u0u_{0} is close to Q2Q^{2} in a slightly stronger topology than the one induced by the energy, then we obtain uniform global bounds and rule out the possibility of blow-up/relaxation in infinite time. There are two main differences between our result here and the one in [6] for m=1m=1. The first one is that here the slightly stronger topology is rather close to the one induced by the energy. The second one is that while the smallness of u0Q2u_{0}-Q^{2} in the stronger norm guarantees stability, in effect our result still applies for any finite size of u0Q2u_{0}-Q^{2} in the stronger norm, in which case it guarantees global bounds (no infinite-time blow-up) in this scenario.

1.3. The main results

In order to state our first main result, we need to recall two important concepts: the modulation parameters α\alpha and λ\lambda, and the reduced field ψ\psi.

The modulation parameters describe the motion of our solution along the ground state family, and have been already discussed in the paragraph leading to (1.7). As noted there, one has quite a bit of flexibility on what are good choices for these parameters. In practice one imposes some orthogonality conditions which uniquely identify these parameters in the regime described in (1.7). Our orthogonality condition is defined by the relation (4.29) in Section 4.

Next, a common strategy in analyzing the SM equation (1.1) is to write uTu𝕊2\nabla u\in T_{u}\mathbb{S}^{2} in an appropriate gauge, that is a choice of an orthonormal frame in Tu𝕊2T_{u}\mathbb{S}^{2}. In complex notation, this representation gives rise to the complex valued differentiated field ψ\psi, which is small in L2L^{2} and solves a nonlinear Schrödinger type evolution. This construction is developed in Section 4; our gauge choice is the classical Coulomb gauge. The L2L^{2} size of ψ\psi is provided by a very simple formula,

(1.8) ψ(0)L22=E(u0)8π,\|\psi(0)\|_{L^{2}}^{2}=E(u_{0})-8\pi,

which in particular shows it is a conserved quantity. One advantage of working with the differentiated field ψ\psi is that it is more amenable to a dispersive type analysis, that is ψ\psi can be measured in function spaces which contain information such as Strichartz estimates and local energy decay estimates.

With these objects at hand, the analysis of the dynamics of the Schrödinger map uu near 𝒬e2\mathcal{Q}^{2}_{e} (and more generally near 𝒬em\mathcal{Q}^{m}_{e}) can be essentially reduced to two major tasks:

  1. i)

    a dispersive analysis of the nonlinear Schrödinger type PDE which governs the evolution of the differentiated field ψ\psi;

  2. ii)

    an analysis of the modulation equation, i.e. the nonlinear ODE that governs the evolution of the modulation parameters α\alpha and λ\lambda.

One has to keep in mind that the PDE and the ODE mentioned above are coupled and thus the two tasks are not performed independently. In most of the prior works, the variation of λ(t)\lambda(t) was small and one could essentially freeze its value for the PDE analysis; at the same time, once the PDE yields enough information of dispersive type, one could show that λ\lambda^{\prime} is small in Lt1L^{1}_{t}. While this description is a bit oversimplified, it highlights the light coupling in the prior works between the PDE and ODE analysis discussed above. In our context, the coupling between the PDE and the ODE analysis is significantly more involved at all levels.

We also emphasize another novel feature of our analysis: in all prior works on the Schrödinger Map equation (1.1) in this context, the results seem to indicate that if blow-up occurs, that is limtTmaxλ(t)=+\lim_{t\nearrow T_{max}}\lambda(t)=+\infty, then the dispersive properties of the field ψ\psi are lost on the (forward) maximal interval of existence Imax=[0,Tmax)I_{max}=[0,T_{max}); see for instance Lemma 3.1 in [12] and Lemma 2.6 in [13], as well as Proposition 5.2 in [14] where one assumes that λ\lambda stays close to 11 in order to recover dispersive estimates.

Surprisingly, our first result in this paper establishes uniform dispersive estimates (measured by Strichartz and local energy decay norms) on the full interval time of existence even in the case of potential blow-up (whether in finite or infinite time).

Theorem 1.3.

Assume that we have a ±2\pm 2-equivariant initial data u0H˙1u_{0}\in\dot{H}^{1} in the homotopy class of Q±2Q^{\pm 2}, and with energy

(1.9) E(u0)8π+δ2,δ1.E(u_{0})\leq 8\pi+\delta^{2},\qquad\delta\ll 1.

Let Imax=(Tmin,Tmax)I_{max}=(T_{min},T_{max}) be the maximal time of existence of the solution uu to the Schrödinger map flow (1.1) with initial data u0u_{0}. Then the associated Coulomb gauge field ψ\psi and the associated modulation parameters λ\lambda and α\alpha satisfy

(1.10) λλ2Lt2(Imax)+αλLt2(Imax)δ,\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}_{t}(I_{max})}+\|\frac{\alpha^{\prime}}{\lambda}\|_{L^{2}_{t}(I_{max})}\lesssim\delta,

respectively

(1.11) ψl2S(Imax)δ.\|\psi\|_{l^{2}S(I_{max})}\lesssim\delta.

Importantly, the implicit constants in the above bounds do not depend on the length of the time interval. The space l2Sl^{2}S, which is defined later in Section 7, contains standard dispersive estimates, namely Strichartz bounds and local energy decay. In particular we have the embedding l2SLx,t4l^{2}S\subset L^{4}_{x,t}.

The above theorem brings a new insight into the stability theory: even if blow-up occurs, the dispersive properties of the gauge field (as measured by Strichartz and local energy decay norms) hold true uniformly all the way up to the blow up time. However, we point out that the estimate in (1.10) is not strong enough to preclude the scenario that Tmax<+T_{\max}<+\infty and limtTmaxλ(t)=\lim_{t\nearrow T_{max}}\lambda(t)=\infty which corresponds to blow-up in finite time; nor does it preclude a similar scenario happening at Tmin>T_{min}>-\infty.

The l2l^{2} summation in (1.11) corresponds to a Littlewood-Paley dyadic frequency decomposition for ψ0L2\psi_{0}\in L^{2} or equivalently for u0Qα,λ±2H˙1u_{0}-Q^{\pm 2}_{\alpha,\lambda}\in\dot{H}^{1} with (α,λ)(\alpha,\lambda) as in (1.7). It will also be of interest to consider the situation where the l2l^{2} Besov bound is supplemented with an l1l^{1} Besov bound. For this purpose we introduce a stronger topology

(1.12) X=B˙2,11H˙1.X=\dot{B}^{1}_{2,1}\subset\dot{H}^{1}.

This is a space of functions in 2\mathbb{R}^{2}, which, when restricted to the class of equivariant functions u(x)=u¯(r)eimθu(x)=\bar{u}(r)e^{im\theta}, yields a space for the radial profile u¯\bar{u}, denoted by X¯=B˙2,1,e1\bar{X}=\dot{B}^{1}_{2,1,e}, with norm

(1.13) u¯B˙2,1,e1=uB˙2,11\|\bar{u}\|_{\dot{B}^{1}_{2,1,e}}=\|u\|_{\dot{B}^{1}_{2,1}}

Correspondingly for ψ\psi we should use the space LX¯:=B˙2,1,e0{L\bar{X}}:=\dot{B}^{0}_{2,1,e} at fixed time; the notation here is motivated in Section 6. In this setting, as we prove later in Proposition 6.1, we have the norm relation

(1.14) ψLX¯uQα,λ±2X.\|\psi\|_{{L\bar{X}}}\approx\|u-Q^{\pm 2}_{\alpha,\lambda}\|_{X}.

Also to measure the space-time regularity of ψ\psi we will use the l1Sl^{1}S norm, which is the l1l^{1} Besov analogue of the earlier l2Sl^{2}S space. We refer the reader to Section 2 for more details on function spaces. Our main bounds in the l1l^{1} Besov class are as follows:

Theorem 1.4.

Let uu be a solution to the Schrödinger map equation as in Theorem 1.3. Assume in addition that the initial data u0u_{0} satisfies

(1.15) u0Qα(0),λ(0)±2X=B˙2,11.u_{0}-Q^{\pm 2}_{\alpha(0),\lambda(0)}\in X=\dot{B}^{1}_{2,1}.

Then we have

(1.16) ψl1S(Imax)u0Qα(0),λ(0)±2X.\|\psi\|_{l^{1}S(I_{max})}\lesssim\|u_{0}-Q^{\pm 2}_{\alpha(0),\lambda(0)}\|_{X}.

We emphasize that the implicit constant is universal, despite the fact that no smallness is assumed for the l1l^{1} norm on the right. There one may choose any (α,λ)(\alpha,\lambda) as in (1.7), but the norm on the left potentially depends on our exact choice of the modulation parameters.

So far, our results have only considered the evolution of the reduced field ψ\psi. The next natural step is to also consider the time evolution of the modulation parameters (α(t),λ(t))(\alpha(t),\lambda(t)), of which λ(t)\lambda(t) plays the leading role as it governs the potential blow-up behavior.

Our first result in this direction establishes that there is no finite time blow-up, but potentially allows for an infinite time blow-up:

Theorem 1.5.

Assume that we have a ±2\pm 2-equivariant initial data u0H˙1u_{0}\in\dot{H}^{1} in the homotopy class of Q±2Q^{\pm 2}, and with energy as in (1.9). Then the Schrödinger Maps equation (1.1) has a unique global in time solution u(t)u(t). Furthermore, if we assume the normalization λ(0)=1\lambda(0)=1, then the modulation parameter λ(t)\lambda(t) satisfies

(1.17) t12λ(t)etCδ2,t,\langle t\rangle^{-\frac{1}{2}}\lesssim\lambda(t)\lesssim e^{\langle t\rangle^{C\delta^{2}}},\forall t\in\mathbb{R},

for some universal constant CC.

Here we recall that, as a consequence of Theorem 1.3, the field associated with this solution satisfies uniform global dispersive bounds as stated in (1.11).

In (1.17) the bound on the right on λ(t)\lambda(t) is the interesting one. The one on the left is simply the self-similar scale which follows directly from the local well-posedness result in Theorem 1.1. The normalization λ(0)\lambda(0) may be removed by rescaling, in which case the bound (1.17) becomes

(1.18) tλ(0)212λ(t)λ(0)etλ(0)2Cδ2,t,\langle t\lambda(0)^{2}\rangle^{-\frac{1}{2}}\lesssim\frac{\lambda(t)}{\lambda(0)}\lesssim e^{\langle t\lambda(0)^{2}\rangle^{C\delta^{2}}},\forall t\in\mathbb{R},

In particular, the above theorem allows for an infinite time blow-up. In forthcoming work, we aim to show that this can actually happen for well chosen initial data.

On the other hand, if in addition we assume l1l^{1} dyadic summability for the data, then we can prevent this from happening:

Theorem 1.6.

Let uu be a solution to the Schrödinger map equation as in Theorem 1.5. Assume in addition that the modulation parameters α(0),λ(0)\alpha(0),\lambda(0) corresponding to the initial data u0u_{0} satisfy

(1.19) u0Qα(0),λ(0)±2XM<+.\|u_{0}-Q^{\pm 2}_{\alpha(0),\lambda(0)}\|_{X}\leq M<+\infty.

Then we have

(1.20) |lnλ(t)lnλ(0)|+|α(t)α(0)|M+M2,t.|\ln\lambda(t)-\ln\lambda(0)|+|\alpha(t)-\alpha(0)|\lesssim M+M^{2},\qquad\forall t\in\mathbb{R}.

Remark. The above condition (1.19) in which we use the specific choice of modulation parameters dictated by (4.29) can be easily replaced by any good choice of modulation parameters subject to (1.7); this simply changes MM to M+CδM+C\delta, where δ\delta was defined in Theorem 1.3, but by definition δM\delta\lesssim M.


Our final result below asserts that the soliton Q±2Q^{\pm 2} is stable in the XX topology and is a direct consequence of the previous Theorem.

Theorem 1.7.

Assume that we have a ±2\pm 2-equivariant initial data u0H˙1u_{0}\in\dot{H}^{1} in the homotopy class of Q±2Q^{\pm 2}, whose initial data satisfies

(1.21) u0Q±2Xγ1.\|{u}_{0}-Q^{\pm 2}\|_{X}\leq\gamma\ll 1.

Then there exists a unique global solution uu for the Schrödinger map equation (1.1), which satisfies uQ±2C(;X){u}-Q^{\pm 2}\in C(\mathbb{R};X) and

(1.22) uQ±2C(;X)γ.\|u-Q^{\pm 2}\|_{C(\mathbb{R};X)}\lesssim\gamma.

One consequence of the above theorem is that classical perturbations of Q±2Q^{\pm 2} that are small in the energy class yield global solutions which do not blow-up in infinite time; in particular they will obey classical bounds uniformly in time. By classical perturbations we mean perturbations that are smooth and decay at infinity, enough to control the B˙2,11\dot{B}^{1}_{2,1} norm (which does not have to be small).

Another consequence of the above theorem is that any potential infinite time blow-up construction would not be stable in the energy class. Indeed, if an initial data u0u_{0}, with u0Q±2H˙11\|u_{0}-Q^{\pm 2}\|_{\dot{H}^{1}}\ll 1, leads to a solution which blows up in an infinite time, then for each δ>0\delta>0 one can construct another data v0v_{0} with u0v0H˙1δ\|u_{0}-v_{0}\|_{\dot{H}^{1}}\leq\delta and with v0Q±2B˙2,11<+\|v_{0}-Q^{\pm 2}\|_{\dot{B}^{1}_{2,1}}<+\infty by cutting off low and high frequencies of u0Q±2u_{0}-Q^{\pm 2}. In the equivariant class of functions this is easily achieved by performing a direct Littlewood-Paley truncation for u0Q±2u_{0}-Q^{\pm 2}, followed by a geometric projection of the new u0u_{0} back on the sphere. The instability of the blow-up in the energy class then follows from the fact that the solution with data v0v_{0} does not blow-up in infinite time (by Theorem 1.7).

1.4. Acknowledgements

The first author was supported by the NSF grant DMS-1900603. The second author was supported by the NSF grant DMS-2103106. The third author was supported by the NSF grant DMS-2054975 as well as by a Simons Investigator grant from the Simons Foundation.

2. Definitions and notations.

In this section we cover some definitions and notations which will be used extensively thoghout the paper. The section does not intend to exhaust all of them, as many objects are defined later as the paper progresses.

While at fixed time our maps into the sphere are functions defined on 2\mathbb{R}^{2}, the equivariance condition allows us to reduce our analysis to functions of a single variable |x|=r[0,)|x|=r\in[0,\infty). One such instance is exhibited in (1.3) where to each equivariant map uu we naturally associate its radial component u¯\bar{u}. Some other functions will turn out to be radial by definition, see, for instance, all the gauge elements in Section 4. We agree to identify such radial functions with the corresponding one dimensional functions of rr. Some of these functions are complex valued, and this convention allows us to use the bar notation with the standard meaning, i.e. the complex conjugate.

Even though we work mainly with functions of a single spatial variable rr, they originate in two dimensions. Therefore, it is natural to make the convention that for the one dimensional functions all the Lebesgue integrals and spaces are with respect to the rdrrdr measure, unless otherwise specified.

We define ABA\lesssim B, if there is a universal constant C>0C>0 such that ACBA\leq CB, and ABA\gtrsim B iff BAB\lesssim A. Further, we define ABA\approx B iff both ABA\lesssim B and BAB\lesssim A. Also, we define ABA\ll B if AcBA\leq cB for some constant c>0c>0 that can be chosen very small, depending only on some universal constants. Also, ABA\gg B iff BAB\ll A. For y>0y>0, we write x=O(y)x=O(y) to mean |x|Cy|x|\leq Cy for some universal constant C>0C>0.

Given a positive parameter λ\lambda, and a function f:(0,)f:(0,\infty)\rightarrow\mathbb{C}, we define its dilation by λ\lambda as follows

fλ(r)=f(λr),r(0,).f^{\lambda}(r)=f(\lambda r),\quad\forall r\in(0,\infty).

On the physical space side, we will use a dyadic partition of (0,)(0,\infty) (which has the natural extension to 2\mathbb{R}^{2}) into sets {Aj}j\{A_{j}\}_{j\in\mathbb{Z}} given by

Aj={2j1<r<2j+1}.A_{j}=\{2^{j-1}<r<2^{j+1}\}.

We also use the notation A<k=j<kAjA_{<k}=\cup_{j<k}A_{j} as well as A>k,AkA_{>k},A_{\geq k} which are similarly defined. Throughout the paper we involve smooth approximations of the characteristic functions of these sets. We construct in the usual manner functions χj,j\chi_{j},\ j\in\mathbb{Z} with the following properties

χj(r)=χ(r2j),χCc((14,2)),χ(r)=1,12r1,\chi_{j}(r)=\chi(\frac{r}{2^{j}}),\quad\chi\in C^{\infty}_{c}((\frac{1}{4},2)),\quad\chi(r)=1,\quad\frac{1}{2}\leq r\leq 1,

and which satisfy the summation property

jχj(r)=1,r(0,).\sum_{j}\chi_{j}(r)=1,\quad\forall r\in(0,\infty).

Finally we also define

χ<k=j<kχj,χk=jkχjχ>k=j>kχj,χk=jkχj.\chi_{<k}=\sum_{j<k}\chi_{j},\quad\chi_{\leq k}=\sum_{j\leq k}\chi_{j}\quad\chi_{>k}=\sum_{j>k}\chi_{j},\quad\chi_{\geq k}=\sum_{j\geq k}\chi_{j}.

In Section 1.1 we encountered the following functions

h1m(r)=2rmr2m+1,h3m(r)=r2m1r2m+1;h_{1}^{m}(r)=\frac{2r^{m}}{r^{2m}+1},\qquad h_{3}^{m}(r)=\frac{r^{2m}-1}{r^{2m}+1};

as detailed in (1.2) they were the components of the basic ground state QmQ^{m} and they will play a crucial role in the analysis in this paper. These functions are continuously differentiable on their domain [0,)[0,\infty), have limits at \infty, limrh1m(r)=0,limrh3m(r)=1\lim_{r\rightarrow\infty}h_{1}^{m}(r)=0,\lim_{r\rightarrow\infty}h_{3}^{m}(r)=1, and satisfy the following obvious property:

(h1m(r))2+(h3m(r))2=1,r[0,);(h_{1}^{m}(r))^{2}+(h_{3}^{m}(r))^{2}=1,\quad\forall\ r\in[0,\infty);

their derivatives are given by

(2.1) rh1m=mh1mh3mr,rh3m=m(h1m)2r.\partial_{r}h_{1}^{m}=-m\frac{h_{1}^{m}h_{3}^{m}}{r},\quad\partial_{r}h_{3}^{m}=m\frac{(h_{1}^{m})^{2}}{r}.

We point out that fairly early in the paper we specialize in the case m=2m=2 and drop the index m=2m=2 from notation, that is we will use h1:=h12,h3:=h32h_{1}:=h_{1}^{2},\ h_{3}:=h_{3}^{2}.


We will use several modified Fourier transforms, adapted to our setting, and use Greek letters such as ξ,η+\xi,\eta\in\mathbb{R}^{+} to denote frequencies. For a positive ξ\xi, we use the notation ξ^=min{ξ,1}\hat{\xi}=\min\{\xi,1\}.

In the context of various Fourier transforms, we involve Fourier projectors which are defined in the standard fashion.We start with the functions mj,jm_{j},j\in\mathbb{Z} satisfying the following properties

mj(ξ)=m(ξ2j),mCc((14,2)),m(ξ)=1,12ξ1.m_{j}(\xi)=m(\frac{\xi}{2^{j}}),\quad m\in C^{\infty}_{c}((\frac{1}{4},2)),\quad m(\xi)=1,\quad\frac{1}{2}\leq\xi\leq 1.

and which satisfy the summation property

jmj(ξ)=1,ξ(0,).\sum_{j}m_{j}(\xi)=1,\forall\xi\in(0,\infty).

We also define m<k,m>k,mkm_{<k},m_{>k},m_{\geq k} in the standard way, that is m<k:=j<kmjm_{<k}:=\sum_{j<k}m_{j} etc. The reader may notice that these functions are not different from the previously defined χj\chi_{j} functions; the only reason we use different letters (for the same onject) is that they are used in different context: the χ\chi’s are used on the physical side, while the mm’s are used on the Fourier side.

We also make use of the following functions: given m~Cc((18,4)),m~(x)=1\widetilde{m}\in C^{\infty}_{c}((\frac{1}{8},4)),\widetilde{m}(x)=1 in the support of mm then, we let m~j(x)=m~(x2j)\widetilde{m}_{j}(x)=\widetilde{m}(\frac{x}{2^{j}}). m~j\widetilde{m}_{j} are very similar to mjm_{j}, except that they do not enjoy the partition of unity property from above; in addition the following holds true: m~jmj=mj\widetilde{m}_{j}\cdot m_{j}=m_{j}.

As mentioned earlier we will employ several Fourier transforms and we list them below for convenience.

  • The standard Fourier transform. This will be briefly used in 2\mathbb{R}^{2} in the section just below. In sections 10 and 11 it will be used in the context of time-dependent functions.

  • The Hankel transform (decomposition into Bessel frames). This will be used in Section 6, see in particular Section 6.2. The use of the Hankel transform is natural in the context of equivariant setup.

  • The Fourier transform in the frames associated to the operators HH and H~\tilde{H}; this theory is fully developed in Section 5. The operators HH and H~\tilde{H} (see their precise definition in (3.7)) occur naturally in our problem once we linearize near the solitons.

We use the following function χi=j=2|ij|2\chi_{i=j}=2^{-\frac{|i-j|}{2}} defined for i,ji,j\in\mathbb{Z}; this can be seen as a smoother version of the Dirac mass δi=j\delta_{i=j}.

For λ>0\lambda>0 and jj\in\mathbb{Z} we also use the notation

χλ=2j=(2j/2λ𝟙{2jλ4}+λ2j𝟙{2jλ1})\chi_{\lambda=2^{j}}=\left(\frac{2^{j/2}}{\sqrt{\lambda}}\mathbbm{1}_{\{\frac{2^{j}}{\lambda}\leq 4\}}+\frac{\lambda}{2^{j}}\mathbbm{1}_{\{\frac{2^{j}}{\lambda}\geq 1\}}\right)

which is essentially χi=j\chi_{i=j}, except with ii formally replaced by log2(λ)\log_{2}(\lambda), and has slightly better decay as 2jλ\frac{2^{j}}{\lambda}\rightarrow\infty.

2.1. Sobolev and Besov spaces

Since equivariant functions are easily reduced to their one-dimensional companion via (1.3), we introduce the one dimensional equivariant version of H˙1\dot{H}^{1},

(2.2) fH˙e12=rfL2(rdr)2+4r1fL2(rdr)2,fHe12=fH˙e12+fL2(rdr)2.\|f\|_{\dot{H}^{1}_{e}}^{2}=\|\partial_{r}f\|_{L^{2}(rdr)}^{2}+4\|r^{-1}f\|_{L^{2}(rdr)}^{2},\quad\|f\|_{H^{1}_{e}}^{2}=\|f\|_{\dot{H}^{1}_{e}}^{2}+\|f\|_{L^{2}(rdr)}^{2}.

This is natural since for functions u:22u:\mathbb{R}^{2}\to\mathbb{R}^{2} with u(r,θ)=e2θRu¯(r)u(r,\theta)=e^{2\theta R}\bar{u}(r) we have

uH˙1=(2π)12u¯H˙e1,uH1=(2π)12u¯He1.\|u\|_{\dot{H}^{1}}=(2\pi)^{\frac{1}{2}}\|\bar{u}\|_{\dot{H}^{1}_{e}},\qquad\|u\|_{H^{1}}=(2\pi)^{\frac{1}{2}}\|\bar{u}\|_{H^{1}_{e}}.

In a similar fashion we define H˙e2\dot{H}^{2}_{e} and He2H^{2}_{e} by the norms

fH˙e22=r2fL22+r1rfL22+r2fL22,fHe22=fH˙e22+fL22\|f\|^{2}_{\dot{H}^{2}_{e}}=\|\partial_{r}^{2}f\|^{2}_{L^{2}}+\|r^{-1}\partial_{r}f\|_{L^{2}}^{2}+\|r^{-2}f\|_{L^{2}}^{2},\qquad\|f\|^{2}_{H^{2}_{e}}=\|f\|^{2}_{\dot{H}^{2}_{e}}+\|f\|_{L^{2}}^{2}

as the natural substitute for H˙2\dot{H}^{2} and H2H^{2}.

We will also use the dual space H˙e1\dot{H}^{-1}_{e}, defined by

fH˙e1=supϕH˙e1;ϕH˙e1=1f,ϕ.\|f\|_{\dot{H}^{-1}_{e}}=\sup_{\phi\in\dot{H}^{1}_{e};\|\phi\|_{\dot{H}^{1}_{e}}=1}\langle f,\phi\rangle.

We can think of the spaces H˙e1\dot{H}^{1}_{e}, H˙e2\dot{H}^{2}_{e} as the ”H1H^{1}”, respectively the ”H2H^{2}” space associated to the equivariant Laplacian

Δm=r21rr+m2r2,-\Delta_{m}=-\partial_{r}^{2}-\frac{1}{r}\partial_{r}+\frac{m^{2}}{r^{2}},

precisely

H˙e1=D((Δm)12)):={f:(Δm)12fL2}.\dot{H}^{1}_{e}=D((-\Delta_{m})^{\frac{1}{2}})):=\{f:(-\Delta_{m})^{\frac{1}{2}}f\in L^{2}\}.

and

H˙e2=D(Δm):={f:ΔmfL2}.\dot{H}^{2}_{e}=D(-\Delta_{m}):=\{f:-\Delta_{m}f\in L^{2}\}.

In our problem the case m=2m=2 is relevant, but to define the Sobolev spaces the choice of |m|2|m|\geq 2 is not important. The above Sobolev spaces can be characterized using spectral projectors. If PkeP_{k}^{e} are the spectral projectors associated to Δm-\Delta_{m}, (which can be defined by restricting the corresponding projectors associated to Δ-\Delta in 2\mathbb{R}^{2} to the mm-equivariant class, or equivalently by using Bessel frames, as discussed later in Section 6.2), then we have

fH˙es2k22ksPkefL22,\|f\|^{2}_{\dot{H}^{s}_{e}}\approx\sum_{k\in\mathbb{Z}}2^{2ks}\|P_{k}^{e}f\|_{L^{2}}^{2},

where ss can be any of the values used above.

In some of the analysis carried in the paper, using these spectral projectors and sharp Fourier decompositions turns out to be counterproductive; instead we can use more robust decompositions. Precisely we have the following

(2.3) fH˙e12inff=fkk(2kfkL2+2kfkH˙e2)2\|f\|^{2}_{\dot{H}^{1}_{e}}\approx\inf_{f=\sum f_{k}}\sum_{k}(2^{k}\|f_{k}\|_{L^{2}}+2^{-k}\|f_{k}\|_{\dot{H}^{2}_{e}})^{2}

and

(2.4) fL22inff=fkk(2kfkH˙e1+2kfkH˙e1)2.\|f\|_{L^{2}}^{2}\approx\inf_{f=\sum f_{k}}\sum_{k}(2^{k}\|f_{k}\|_{\dot{H}^{-1}_{e}}+2^{-k}\|f_{k}\|_{\dot{H}^{1}_{e}})^{2}.

It is a straightforward exercise to establish (2.3) and (2.4).

Correspondingly we define associated Besov spaces. The ones we use in this paper are B˙2,11\dot{B}^{1}_{2,1}, B˙2,1,e0\dot{B}^{0}_{2,1,e} and B˙2,1,e0\dot{B}^{0}_{2,1,e} with norms

uB˙2,11=k2kPkfL2,fB˙2,1,e0=kPkefL2,fB˙2,1,e1=k2kPkefL2.\|u\|_{\dot{B}^{1}_{2,1}}=\sum_{k}2^{k}\|P_{k}f\|_{L^{2}},\qquad\|f\|_{\dot{B}^{0}_{2,1,e}}=\sum_{k}\|P^{e}_{k}f\|_{L^{2}},\qquad\|f\|_{\dot{B}^{1}_{2,1,e}}=\sum_{k}2^{k}\|P^{e}_{k}f\|_{L^{2}}.

In the first norm PkP_{k} represent the standard Littlewood-Paley projectors in 2\mathbb{R}^{2}, while in the latter norms, just as above, PkeP_{k}^{e} are the spectral projectors associated to Δm-\Delta_{m}. For a complete definition, we let B˙2,1,e1\dot{B}^{1}_{2,1,e} be the subspace of functions fH˙e1f\in\dot{H}^{1}_{e} which have the property that fB˙2,1,e1<\|f\|_{\dot{B}^{1}_{2,1,e}}<\infty; similarly B˙2,1,e0\dot{B}^{0}_{2,1,e} is the subspace of functions fL2f\in L^{2} with the property that fB˙2,1,e0<\|f\|_{\dot{B}^{0}_{2,1,e}}<\infty. The key upgrade that the spaces B˙2,1,e1\dot{B}^{1}_{2,1,e}, and B˙2,1,e0\dot{B}^{0}_{2,1,e}, have over their counterparts H˙e1\dot{H}^{1}_{e}, respectively L2L^{2}, is that they bring an l1L2l^{1}L^{2} structure versus the standard l2L2l^{2}L^{2} one; here l1l^{1} and l2l^{2} indicate the norm used for the sequence (PkefL2)k(\|P^{e}_{k}f\|_{L^{2}})_{k\in\mathbb{Z}}.

For brevity we also introduce some shorter notations, which will be suggestive in different contexts:

X:=B˙2,11,LX:=B˙2,10,X:=\dot{B}^{1}_{2,1},\qquad LX:=\dot{B}^{0}_{2,1},

which will be used for functions defined on 2\mathbb{R}^{2}, and, the corresponding for one dimensional functions,

X¯:=B˙2,1,e1,LX¯:=B˙2,1,e0,\bar{X}:=\dot{B}^{1}_{2,1,e},\qquad{L\bar{X}}:=\dot{B}^{0}_{2,1,e},

where we restrict XX to functions uu of the form u(r,θ)=e2θRu¯(r)u(r,\theta)=e^{2\theta R}\bar{u}(r), so that we have the algebraic and topological identification uXu\in X iff u¯X¯\bar{u}\in\bar{X}.

Just as we did above for the Sobolev spaces, we can alternatively define these Besov spaces as follows:

  • for j=1j=1 we can set

    (2.5) uB˙2,1,e1=infu=ukk2kukL2+2kukH˙e2;\|u\|_{\dot{B}^{1}_{2,1,e}}=\inf_{u=\sum u_{k}}\sum_{k}2^{k}\|u_{k}\|_{L^{2}}+2^{-k}\|u_{k}\|_{\dot{H}^{2}_{e}};
  • for j=0j=0 we have

    (2.6) uB˙2,1,e0=infu=ukk2kukH˙e1+2kukH˙e1.\|u\|_{\dot{B}^{0}_{2,1,e}}=\inf_{u=\sum u_{k}}\sum_{k}2^{k}\|u_{k}\|_{\dot{H}^{-1}_{e}}+2^{-k}\|u_{k}\|_{\dot{H}^{1}_{e}}.

2.2. Integration operators on radial functions

Two operators which are often used on radial functions are [r]1[\partial_{r}]^{-1} and [rr]1[r\partial_{r}]^{-1} defined as

(2.7) [r]1f(r)=rf(s)𝑑s,[rr]1f(r)=r1sf(s)𝑑s.[\partial_{r}]^{-1}f(r)=-\int_{r}^{\infty}f(s)ds,\qquad[r\partial_{r}]^{-1}f(r)=-\int_{r}^{\infty}\frac{1}{s}f(s)ds.

A direct argument shows that we have the Hardy type inequality

(2.8) [rr]1fLppfLp,1p<.\|[r\partial_{r}]^{-1}f\|_{L^{p}}\lesssim_{p}\|f\|_{L^{p}},\qquad 1\leq p<\infty.

We also have a weighted version

(2.9) w[rr]1fLppwfLp,1p<,\|w[r\partial_{r}]^{-1}f\|_{L^{p}}\lesssim_{p}\|wf\|_{L^{p}},\qquad 1\leq p<\infty,

under the assumption that g(r)=w(r)r2pg(r)=w(r)r^{\frac{2}{p}} is an increasing function satisfying

g(r)(1ϵ)g(2r),g(r)\leq(1-\epsilon)g(2r),

for some ϵ>0\epsilon>0. The proof is straightforward.

Given a function fL2f\in L^{2} we note that the following holds true:

fL22=mfL2(Am)2.\|f\|_{L^{2}}^{2}=\sum_{m\in\mathbb{Z}}\|f\|^{2}_{L^{2}(A_{m})}.

Thus we have an lm2l^{2}_{m} structure for the sequence (fL2(Am)2)mZ(\|f\|^{2}_{L^{2}(A_{m})})_{m\in Z}. The following slightly stronger norm (improving the l2l^{2} summation to an l1l^{1} one)

fl1L2=mfL2(Am),\|f\|_{l^{1}L^{2}}=\sum_{m\in\mathbb{Z}}\|f\|_{L^{2}(A_{m})},

will play an important role in our analysis; one can simply define l1L2l^{1}L^{2} as the subspace of L2L^{2} for which the above norm is finite.

Just as above, the operator [r]1[\partial_{r}]^{-1} is used to define the space [r]1l1L2[\partial_{r}]^{-1}l^{1}L^{2}, as the completion of of Hcomp1([0,))H^{1}_{comp}([0,\infty)) with respect to the following norm

f[r]1l1L2=rfl1L2=mrfL2(Am).\|f\|_{[\partial_{r}]^{-1}l^{1}L^{2}}=\|\partial_{r}f\|_{l^{1}L^{2}}=\sum_{m}\|\partial_{r}f\|_{L^{2}(A_{m})}.

Since rfL1(dr)f[r]1l1L2\|\partial_{r}f\|_{L^{1}(dr)}\lesssim\|f\|_{[\partial_{r}]^{-1}l^{1}L^{2}}, it follows that ff has limits both at 0 and \infty; and since it is approximated by functions in Hcomp1([0,))H^{1}_{comp}([0,\infty)), it follows that limrf(r)=0\lim_{r\rightarrow\infty}f(r)=0. We also have the following inequality

(2.10) rfL2+fLf[r]1l1L2.\|\partial_{r}f\|_{L^{2}}+\|f\|_{L^{\infty}}\lesssim\|f\|_{[\partial_{r}]^{-1}l^{1}L^{2}}.

We will also work with space H˙e1+[r]1l1L2\dot{H}^{1}_{e}+[\partial_{r}]^{-1}l^{1}L^{2} and note that r(H˙e1+[r]1l1L2)=rH˙e1+l1L2\partial_{r}(\dot{H}^{1}_{e}+[\partial_{r}]^{-1}l^{1}L^{2})=\partial_{r}\dot{H}^{1}_{e}+l^{1}L^{2}.

We seek to understand a little better the structure of the space [r]1(l1L2)[\partial_{r}]^{-1}(l^{1}L^{2}). Since rfl1L2\partial_{r}f\in l^{1}L^{2}, it suffices to consider the case when rf=χmgL2,gL2\partial_{r}f=\chi_{m}g\in L^{2},g\in L^{2}. Then we let

f=χ<m4cm+χm4rχmg(s)𝑑s,cm=χmg(s)𝑑s;f=\chi_{<m-4}c_{m}+\chi_{\geq m-4}\int_{r}^{\infty}\chi_{m}g(s)ds,\quad c_{m}=\int\chi_{m}g(s)ds;

The first observation is that

|cm|χmgL2.|c_{m}|\lesssim\|\chi_{m}g\|_{L^{2}}.

The second component is obviously in H˙e1\dot{H}^{1}_{e}, thus we have obtained that a generic function in [r]1(l1L2)[\partial_{r}]^{-1}(l^{1}L^{2}) decomposes as follows:

(2.11) f=f1+mcmχm,f1H˙e1+(cm)l1f[r]1(l1L2).f=f_{1}+\sum_{m}c_{m}\chi_{\leq m},\quad\|f_{1}\|_{\dot{H}^{1}_{e}}+\|(c_{m})\|_{l^{1}}\lesssim\|f\|_{[\partial_{r}]^{-1}(l^{1}L^{2})}.

3. An outline of the paper

Due to the complexity of the paper, an overview of the ideas and the organization of the paper is helpful before an in-depth reading.

3.1. The frame method and the Coulomb gauge

At first sight the Schrödinger Map equation has little to do with the Schrödinger equation. A good way to bring in the Schrödinger structure is by using the frame method. Precisely, at each point (x,t)2+1(x,t)\in\mathbb{R}^{2+1} one introduces an orthonormal frame (v,w)(v,w) in Tu(x,t)𝕊2T_{u(x,t)}\mathbb{S}^{2}. This frame is used to measure the derivatives of uu, and reexpress them as the complex valued radial differentiated fields

ψ1=ruv+iruw,ψ2=θuv+iθuw.\psi_{1}=\partial_{r}u\cdot v+i\partial_{r}u\cdot w,\qquad\psi_{2}=\partial_{\theta}u\cdot v+i\partial_{\theta}u\cdot w.

Here the use of polar coordinates is motivated by the equivariance condition. Thus instead of working with the equation for uu, one writes the evolution equations for the differentiated fields. The frame (v,w)(v,w) does not appear directly there, but only via the real valued radial connection coefficients

A1=rvw,A2=θvw,A0=tvw.A_{1}=\partial_{r}v\cdot w,\qquad A_{2}=\partial_{\theta}v\cdot w,\qquad A_{0}=\partial_{t}v\cdot w.

A-priori the frame is not uniquely determined. To fix it one first asks that the frame be equivariant, and then that it satisfies an appropriate condition. Here it is convenient to use the Coulomb gauge; due to the equivariance this takes a very simple form, A1=0A_{1}=0. The construction of the Coulomb gauge is the first goal in the next section. In Proposition 4.2 we prove that for H˙1\dot{H}^{1} equivariant maps into 𝕊2\mathbb{S}^{2} close to QQ there exists an unique Coulomb frame (v,w)(v,w) which satisfies appropriate boundary conditions at infinity, see (4.11). In addition, this frame has a C1C^{1} dependence on the map uu in a suitable topology.

In the Coulomb gauge the other spatial connection coefficient A2A_{2}, while nonzero, has a very simple form A2=u3A_{2}=u_{3}. We also compute A0A_{0} in terms of ψ1\psi_{1}, ψ2\psi_{2} and A2A_{2},

(3.1) A0=12(|ψ1|21r2|ψ2|2)+[rr1](|ψ1|21r2|ψ2|2).A_{0}=-\frac{1}{2}\left(|\psi_{1}|^{2}-\frac{1}{r^{2}}|\psi_{2}|^{2}\right)+[r\partial_{r}^{-1}]\left(|\psi_{1}|^{2}-\frac{1}{r^{2}}|\psi_{2}|^{2}\right).

3.2. The reduced field ψ\psi

The two fields ψ1\psi_{1} and ψ2\psi_{2} are not independent. Hence it is convenient to work with a single field

ψ=ψ1ir1ψ2,\psi=\psi_{1}-ir^{-1}\psi_{2},

which we will call the reduced field. The relevance of the variable ψ\psi comes from the following reinterpretation. If 𝒲\mathcal{W} is defined as the vector

𝒲=rumru×RuTu(𝕊2),\mathcal{W}=\partial_{r}u-\frac{m}{r}u\times Ru\in T_{u}(\mathbb{S}^{2}),

then ψ\psi is the representation of 𝒲\mathcal{W} with respect to the frame (v,w)(v,w). On the other hand, a direct computation, see for instance [13], leads to

E(u)=π0(|ru¯|2+m2r2|u¯×Ru¯|2)r𝑑r=π𝒲¯L2(rdr)2+4π|m|E(u)=\pi\int_{0}^{\infty}\left(|\partial_{r}\bar{u}|^{2}+\frac{m^{2}}{r^{2}}|\bar{u}\times R\bar{u}|^{2}\right)rdr=\pi\|\bar{\mathcal{W}}\|_{L^{2}(rdr)}^{2}+4\pi|m|

where we recall that u(r,θ)=emθRu¯(r)u(r,\theta)=e^{m\theta R}\bar{u}(r) and, similarly, let 𝒲(r,θ)=emθR𝒲¯(r)\mathcal{W}(r,\theta)=e^{m\theta R}\bar{\mathcal{W}}(r). Therefore ψ=0\psi=0 is a complete characterization of uu being a harmonic map. Moreover the mass of ψ\psi is directly related to the energy of uu via

(3.2) ψL22=𝒲¯L22=E(u)4π|m|π.\|\psi\|_{L^{2}}^{2}=\|\bar{\mathcal{W}}\|_{L^{2}}^{2}=\frac{E(u)-4\pi|m|}{\pi}.

A second goal of the next section is to derive an equation for the time evolution of ψ\psi. This is governed by a cubic NLS type equation,

(3.3) (it+Δ2r2)ψ=(A02A2r21r(ψ2ψ¯))ψ.(i\partial_{t}+\Delta-\frac{2}{r^{2}})\psi=\left(A_{0}-2\frac{A_{2}}{r^{2}}-\frac{1}{r}\Im({\psi}_{2}\bar{\psi})\right)\psi.

In addition, we show that ψ\psi is connected back to (ψ2,A2)(\psi_{2},A_{2}) via the ODE system

(3.4) rA2=(ψψ¯2)+1r|ψ2|2,rψ2=iA2ψ1rA2ψ2\partial_{r}A_{2}=\Im{(\psi\bar{\psi}_{2})}+\frac{1}{r}|\psi_{2}|^{2},\qquad\partial_{r}\psi_{2}=iA_{2}\psi-\frac{1}{r}A_{2}\psi_{2}

with the conservation law A22+|ψ2|2=1A_{2}^{2}+|\psi_{2}|^{2}=1.

3.3. The modulation parameters (α,λ)(\alpha,\lambda)

The system (3.3) is not self-contained, as the reduced field ψ\psi does not uniquely determine (ψ2,A2)(\psi_{2},A_{2}) via the system (3.4). What is missing is a suitable initial condition.

The initial condition for (3.4) is closely related to a suitable choice of a ”closest soliton” to the map uu. This is denoted by Qα,λQ_{\alpha,\lambda}, where the soliton parameters (α,λ)(\alpha,\lambda) are viewed as modulation parameters for our Schrödinger map. The modulation parameters are uniquely identified with a solution to the compatibility ode (3.4) via a suitable orthogonality condition, see (4.29). The one-to-one correspondence between the modulation parameters (λ,α)(\lambda,\alpha) and the solutions to (3.4) is established in Proposition 4.5.

3.4. Linearizations and the operators HH, H~\tilde{H}

A key role in our analysis is played by the linearization of the Schrödinger Map equation around the soliton QQ. A solution to the linearized flow is a function

ulin:2+1TQS2.u_{lin}:\mathbb{R}^{2+1}\to T_{Q}S^{2}.

The Coulomb frame associated to QQ has the form

vQ(θ,r)=emθRv¯Q(r),wQ(θ,r)=emθRw¯Q(r)v_{Q}(\theta,r)=e^{m\theta R}\bar{v}_{Q}(r),\qquad w_{Q}(\theta,r)=e^{m\theta R}\bar{w}_{Q}(r)

with

v¯Q(r)=(h3m(r)0h1m(r)),w¯Q(r)=(010).\bar{v}_{Q}(r)=\left(\begin{array}[]{ccc}h_{3}^{m}(r)\\ 0\\ -h_{1}^{m}(r)\end{array}\right),\qquad\bar{w}_{Q}(r)=\left(\begin{array}[]{ccc}0\\ 1\\ 0\end{array}\right).\qquad

Expressing ulinu_{lin} in this frame,

ϕlin=ulin,vQ+iulin,wQ\phi_{lin}=\langle u_{lin},v_{Q}\rangle+i\langle u_{lin},w_{Q}\rangle

one obtains the Schrödinger type equation

(3.5) (itH)ϕlin=0(i\partial_{t}-H)\phi_{lin}=0

where the operator HH acting on radial functions has the form

H=Δ+V,V(r)=m2r2(12(h1m)2).H=-\Delta+V,\qquad V(r)=\frac{m^{2}}{r^{2}}(1-2(h_{1}^{m})^{2}).

On the other hand linearizing the equation (3.3) around the soliton QQ, we obtain a linear Schrödinger equation of the form

(3.6) (itH~)ψlin=0(i\partial_{t}-\tilde{H})\psi_{lin}=0

where the operator H~\tilde{H} acting on radial functions has the form

H~=Δ+V~,V~(r)=1+m22mh3mr2.\tilde{H}=-\Delta+\tilde{V},\qquad\tilde{V}(r)=\frac{1+m^{2}-2mh_{3}^{m}}{r^{2}}.

The operators HH and H~\tilde{H} are conjugate operators and admit the factorizations

(3.7) H=LL,H~=LLH=L^{*}L,\qquad\tilde{H}=LL^{*}

where

L=h1mr1h1m=r+mrh3m,L=1h1mrh1m1r=r+mh3m1r.L=h_{1}^{m}\partial_{r}\frac{1}{h_{1}^{m}}=\partial_{r}+\frac{m}{r}h_{3}^{m},\qquad L^{*}=-\frac{1}{h_{1}^{m}}\partial_{r}h_{1}^{m}-\frac{1}{r}=-\partial_{r}+\frac{mh_{3}^{m}-1}{r}.

The linearized variables ϕlin\phi_{lin} and ψlin\psi_{lin} are also conjugated variables,

(3.8) ψlin=Lϕlin.\psi_{lin}=L\phi_{lin}.

The operator HH is nonnegative and bounded from H˙1\dot{H}^{1} to H˙1\dot{H}^{-1}, but it is not positive definite; instead it has a zero mode ϕ0\phi_{0}, solving Lϕ0=0L\phi_{0}=0, namely

ϕ0(r)=h1m.\phi_{0}(r)=h_{1}^{m}.

This corresponds to the solution ϕlin\phi_{lin} for (3.5) obtained by differentiating the soliton family with respect to either parameter. A consequence of this is that the linear Schrödinger evolution (3.5) does not have good dispersive properties, a fact which is at the heart of our instability result.

If |m|=1|m|=1 then the zero mode is a resonance, while if |m|2|m|\geq 2 then the zero resonance is replaced by a zero eigenvalue. If |m|3|m|\geq 3 then this eigenvalue belongs to H˙1\dot{H}^{-1}, which allows for a clean splitting of the H˙1\dot{H}^{1} space into an eigenvalue mode, which is stationary, and an orthogonal component, which has good dispersive properties. This leads to the stability results in [13], [14]. As already mentioned in the introduction, if m=1m=1, it was shown in [6] that instability occurs, and construction of solutions which blow-up in finite time were provided in [28] and [32]. In this paper we address the case |m|=2|m|=2 and show that instability still occurs. Hence, despite the fact that h12h_{1}^{2} is an eigenvalue, the result for |m|=2|m|=2 is closer to the |m|=1|m|=1 case.

If QQ is replaced by Qα,λQ_{\alpha,\lambda} then HH and H~\tilde{H} are replaced by their rescaled versions HλH_{\lambda} and H~λ\tilde{H}_{\lambda} where VV and V~\tilde{V} are replaced by

(3.9) Vλ=λ2V(λr),V~λ=λ2V~(λr).V_{\lambda}=\lambda^{2}V(\lambda r),\qquad\tilde{V}_{\lambda}=\lambda^{2}\tilde{V}(\lambda r).

Correspondingly, the operator LL in the factorization (3.7) is replaced by

(3.10) Lλ=r+mrh3m(λr).L_{\lambda}=\partial_{r}+\frac{m}{r}h_{3}^{m}(\lambda r).

From this point on our analysis becomes specialized to the case m=2m=2; the case m=2m=-2 is identical. Hence for the reminder of this section we assume that m=2m=2.

3.5. Spectral theory for HH and H~\tilde{H}.

A first objective of Section 5 is to describe the spectral theory for the linear operators HH and H~\tilde{H}. The analysis in the case of HH has already been done in [21], and it is easily obtained via the LL conjugation in the case of H~\tilde{H}. The normalized generalized eigenfunctions for HH and H~\tilde{H} are denoted by ϕξ\phi_{\xi}, respectively ψξ\psi_{\xi}, and satisfy

Hϕξ=ξ2ϕξ,H~ψξ=ξ2ψξ,Lϕξ=ξψξ.H\phi_{\xi}=\xi^{2}\phi_{\xi},\qquad\tilde{H}\psi_{\xi}=\xi^{2}\psi_{\xi},\qquad L\phi_{\xi}=\xi\psi_{\xi}.

Correspondingly we have a generalized Fourier transform H{\mathcal{F}}_{H} associated to HH and a generalized Fourier transform H~{\mathcal{F}}_{\tilde{H}} associated to H~\tilde{H}.

This quickly leads to generalized eigenfunctions for the rescaled operators HλH_{\lambda} and H~λ\tilde{H}_{\lambda}. A considerable effort is devoted to the study of the transition from one H~λ\tilde{H}_{\lambda} frame to another. This is closely related to the transference operator introduced in [21].

One reason we prefer to work with the ψ\psi variable is that the operator H~\tilde{H} has a good spectral behavior at frequency zero, therefore we have favorable dispersive decay estimates for the corresponding linear Schrödinger evolution (3.6).

3.6. Stronger 1\ell^{1} Besov topologies: the spaces XX, LX¯{L\bar{X}}.

The bulk of our analysis is done for solutions uu in the energy space H˙1\dot{H}^{1}, which corresponds to ψ\psi in L2L^{2}. But in these topologies the solitons turn out to be unstable, which motivates us to also seek stronger topologies where the solitons are instead stable. In the 22-equivariant setting studied in this article, these spaces, denoted by XX for the map uu, respectively by LX¯{L\bar{X}} for the for the reduced field ψ\psi, turn out to be simply the corresponding 1\ell^{1} Besov type spaces, namely X=B˙2,11X=\dot{B}^{1}_{2,1}, respectively LX¯=B˙2,1,e0{L\bar{X}}=\dot{B}^{0}_{2,1,e} (see Section 2.1 for precise defintions).

These spaces and their properties are discussed in detail in Section 6, where we establish the one-to-one elliptic correspondence between uXu\in X and ψLX¯\psi\in{L\bar{X}}, see Proposition 6.1. In the same section we also establish equivalent characterizations of X¯\bar{X} and LX¯{L\bar{X}} relative to the Littlewood-Paley decompositions associated to the operators HH, respectively H~\tilde{H}. This in particular justifies our notations, by showing that

L:X¯LX¯,L:\bar{X}\to{L\bar{X}},

as a surjective map with a one dimensional kernel.

3.7. The linear H~\tilde{H} flow

This represents the main component in the evolution equation for the reduced field ψ\psi. In Section 7 we study this linear flow, with the aim of proving local energy decay and Strichartz bounds.

One added difficulty is that we need to allow the scaling parameter λ\lambda to vary as a function of time. But an arbitrary time dependence of λ\lambda cannot be allowed; instead, we assume that we have a bound on the scale invariant quantity

=λλ2Lt2.\mathcal{B}=\left\|\frac{\lambda^{\prime}}{\lambda^{2}}\right\|_{L^{2}_{t}}.

This turns out to be the crucial quantity, whose finiteness guarantees global, universal bounds for the linear H~\tilde{H} flow.

3.8. The nonlinear source term N(ψ)N(\psi)

Once the modulation parameters (α,λ)(\alpha,\lambda) have been selected using our orthogonality condition, the ψ\psi equation (3.3) can be recast in the form

(3.11) (itH~)ψ=N(ψ).(i\partial_{t}-\tilde{H})\psi=N(\psi).

where N(ψ)N(\psi) is at least quadratic and contains all nonlinear contributions.

In Section 8 we prove that N(ψ)N(\psi) can be estimated perturbatively within the framework of the linear H~\tilde{H} equation. This implies global, uniform Strichartz an local energy bounds for the reduced field ψ\psi associated to a Schrödinger map uu in terms of the initial data size in L2L^{2}, under the sole assumption that we control the quantity \mathcal{B} above. In the same section we also show that for initial data ψ(0)\psi(0) in the smaller space LX¯{L\bar{X}}, the 1\ell^{1} Besov structure carries over to the solution ψ\psi, relative to the (time-dependent) Littlewood-Paley decomposition associated to H~λ\tilde{H}_{\lambda}.

3.9. The modulation equations

Following our study of the ψ\psi equation, the second main step in the proof of our results is the study of the modulation equations for the modulation parameters (λ(t),α(t))(\lambda(t),\alpha(t)). This analysis begins in Section 9, where we derive the modulation equations as a nonlinear ode system with source terms which depend on ψ\psi. These source terms can be divided as follows:

  • linear in ψ\psi; these are the most troublesome ones, which at best can be estimated in Lt2L^{2}_{t} using the local energy decay bounds.

  • quadratic and higher, which can be estimated in Lt1L^{1}_{t} and thus play a perturbative role.

Nevertheless, the L2L^{2} bounds turn out to be sufficient in order to close the bound for \mathcal{B} under the sole assumption that ψ(0)\psi(0) is small in L2L^{2}. This is a critical step in our analysis, as it implies that the bounds for ψ\psi are universal on the existence time for the solutions, even though by itself it does not preclude finite time blow-up.

A second, more refined step is undertaken in Section 10, where we show that the linear in ψ\psi source terms can be in effect placed also in the space H˙t12\dot{H}_{t}^{-\frac{1}{2}}, for which perturbative analysis fails but only in a borderline fashion.

3.10. ODE analysis for the modulation parameters

The final steps in our study of the modulation equation are carried out in Section 11 and Section 12.

In a first step we begin by assuming a stronger Besov bound B˙2,112\dot{B}^{-\frac{1}{2}}_{2,1} for the source terms. In that case the source terms can be treated perturbatively, but the difficulty is that we cannot assume smallness so a careful control of the constants is needed. This step suffices for the case when the initial data ψ(0)\psi(0) is in the smaller space LX¯{L\bar{X}}.

In the second and final step, we consider source terms which are small in H˙t12\dot{H}_{t}^{-\frac{1}{2}} and also with small \mathcal{B}, which gives an L2L^{2} bound at sufficiently high frequencies. Over compact time intervals, such source terms may be placed in the stronger Besov space B˙2,112\dot{B}^{-\frac{1}{2}}_{2,1}, but with a large norm. From here, a careful balancing of the time interval [0,T][0,T] and of the scale of λ\lambda in [0,T][0,T] leads to the proof of the bound

λ(t)TCδ2,δ=ψ(0)L2\lambda(t)\lesssim T^{C\delta^{2}},\qquad\delta=\|\psi(0)\|_{L^{2}}

which in turn implies our main result, asserting that no finite time blow-up is possible.

4. The Coulomb gauge representation of the equation

Our first goal in this section is to introduce the Coulomb gauge, which allows us to rewrite the Schrödinger map equation for equivariant solutions as a semilinear Schrödinger system for a differentiated field ψ\psi, with a nonlinearity that contains no derivatives; this is an important feature that makes the nonlinear analysis much simpler. However, the recovery of a Schrödinger Map state from its Coulomb gauge representation ψ\psi is not unique, instead it retains two degrees of freedom. We fix these two degrees of freedom via an orthogonality condition relative to a nearby reference soliton Qα,λQ_{\alpha,\lambda}. The soliton parameters (α,λ)(\alpha,\lambda) will vary as a function a time, and will be referred to as modulation parameters. Later in the paper, we will use the orthogonality condition in order to derive a set of ordinary differential equations, called the modulation equations, for the two modulation parameters. This will allow us to view the Schrödinger map evolution as a coupled system with two components:

  1. (i)

    A differentiated field ψ\psi which solves a semilinear Schrödinger type equation, and

  2. (ii)

    Two modulation parameters (α,λ)(\alpha,\lambda) which solve an appropriate ODE system.

Given that we need to introduce and develop an array of notations along the way, it is difficult to summarize the full conclusion of this section here; instead, we will do so at the end, see Section 4.5.

For reference, we note that the use of the Coulomb gauge in the context of the Schrödinger map equation originates in the work of Chang, Shatah, Uhlenbeck [7]. This strategy was particularly successful in later works on the Schrödinger map equation with data (hence solutions) that have radial or, more general, equivariant symmetries; see for instance [1, 6, 12, 13, 14]. The Coulomb gauge has been also used for general data (without symmetries) in [2] in high dimensions n4n\geq 4, but fails to be the efficient gauge in low dimensions n=2,3n=2,3.

The gauge representation theory in this section is entirely based on the setup developed in [6] by the first and last authors, see Chapter 3 there. The orthogonality condition is also similar to the one we have used in [6], but we should mention that it first appeared in the work of Gustafson, Nakanishi, Tsai [14] in a different form, namely at the level of maps rather than gauge components. Indirectly we are using some of the analysis in [14].

4.1. The differentiated maps

We let the differentiation operators 0,1,2\partial_{0},\partial_{1},\partial_{2} stand for t,r,θ\partial_{t},\partial_{r},\partial_{\theta} respectively. Our strategy will be to replace the equation for the Schrödinger map uu with equations for its derivatives 1u\partial_{1}u, 2u\partial_{2}u expressed in an orthonormal frame v,wTu𝕊2v,w\in T_{u}\mathbb{S}^{2}. To fix the sign in the choice of ww, we assume that

u×v=w.u\times v=w.

Since uu is mm-equivariant it is natural to work with mm-equivariant frames, i.e.

v=emθRv¯(r),w=emθRw¯(r).v=e^{m\theta R}\bar{v}(r),\qquad w=e^{m\theta R}\bar{w}(r).

Given such a frame we introduce the differentiated fields ψk\psi_{k} and the connection coefficients AkA_{k} by

(4.1) ψk=kuv+ikuw,Ak=kvw,k=0,1,2.\begin{split}\psi_{k}=\partial_{k}u\cdot v+i\partial_{k}u\cdot w,\qquad A_{k}=\partial_{k}v\cdot w,\quad k=0,1,2.\end{split}

Due to the equivariance of (u,v,w)(u,v,w) it follows that both ψk\psi_{k} and AkA_{k} are radially symmetric (therefore subject to the conventions made in Section 2). Conversely, given ψk\psi_{k} and AkA_{k}, we can return to the frame (u,v,w)(u,v,w) via the ODE system:

(4.2) {ku=(ψk)v+(ψk)wkv=(ψk)u+Akwkw=(ψk)uAkv.\left\{\begin{array}[]{l}\partial_{k}u=(\Re{\psi_{k}})v+(\Im{\psi_{k}})w\cr\partial_{k}v=-(\Re{\psi_{k}})u+A_{k}w\cr\partial_{k}w=-(\Im{\psi_{k}})u-A_{k}v.\end{array}\right.

If we introduce the covariant differentiation

Dk=k+iAk,k{0,1,2},D_{k}=\partial_{k}+iA_{k},\ \ k\in\{0,1,2\},

then it is a straightforward computation to check the compatibility conditions:

(4.3) Dlψk=Dkψl,l,k=0,1,2.D_{l}\psi_{k}=D_{k}\psi_{l},\ \ \ l,k=0,1,2.

The curvature of this connection is given by

(4.4) DlDkDkDl=i(lAkkAl)=i(ψlψ¯k),l,k=0,1,2.D_{l}D_{k}-D_{k}D_{l}=i(\partial_{l}A_{k}-\partial_{k}A_{l})=i\Im{(\psi_{l}\overline{\psi}_{k})},\ \ \ l,k=0,1,2.

An important geometric feature is that ψ2,A2\psi_{2},A_{2} are closely related to the original map. Precisely, for A2A_{2} we have:

(4.5) A2=m(k×v)w=mk(v×w)=mku=mu3A_{2}=m(k\times v)\cdot w=mk\cdot(v\times w)=mk\cdot u=mu_{3}

and, in a similar manner,

(4.6) ψ2=m(w3iv3).\psi_{2}=m(w_{3}-iv_{3}).

Since the (u,v,w)(u,v,w) frame is orthonormal, the following relations also follow:

(4.7) |ψ2|2=m2(u12+u22),|ψ2|2+A22=m2.|\psi_{2}|^{2}=m^{2}(u_{1}^{2}+u_{2}^{2}),\qquad|\psi_{2}|^{2}+A_{2}^{2}=m^{2}.

4.2. The Coulomb gauge

Now we turn our attention to the choice of the (v,w)(v,w) at θ=0\theta=0, that is to (v¯,w¯)(\bar{v},\bar{w}). Here we have the gauge freedom of an arbitrary rotation depending on tt and rr. Our gauge choice aims to remove this freedom. In this article we will use the Coulomb gauge, which for general maps uu has the form

div A=0.\text{div }A=0.

In polar coordinates this is written as

1r1(rA1)+1r22A2=0.\frac{1}{r}\partial_{1}(rA_{1})+\frac{1}{r^{2}}\partial_{2}A_{2}=0.

However, in the equivariant case A2A_{2} is radial, so we are left with a simpler formulation A1=0A_{1}=0, or equivalently

(4.8) rv¯w¯=0\partial_{r}\bar{v}\cdot\bar{w}=0

which can be rearranged into a convenient ODE as follows

(4.9) rv¯=(v¯u¯)ru¯(v¯ru¯)u¯.\partial_{r}\bar{v}=(\bar{v}\cdot\bar{u})\partial_{r}\bar{u}-(\bar{v}\cdot\partial_{r}\bar{u})\bar{u}.

The first term on the right vanishes and could be omitted, but it is convenient to add it so that the above linear ODE is solved not only by vv and ww, but also by uu. Then we can write an equation for the matrix 𝒪=(v¯,w¯,u¯)\mathcal{O}=(\bar{v},\bar{w},\bar{u}):

(4.10) r𝒪=M𝒪,M=ru¯u¯:=ru¯u¯u¯ru¯\partial_{r}\mathcal{O}=M\mathcal{O},\qquad M=\partial_{r}\bar{u}\wedge\bar{u}:=\partial_{r}\bar{u}\otimes\bar{u}-\bar{u}\otimes\partial_{r}\bar{u}

with an antisymmetric matrix MM.

The ODE (4.9) needs to be initialized at some point. A change in the initialization leads to a multiplication of all of the ψk\psi_{k} by a unit sized complex number. This is irrelevant at fixed time, but as the time varies we need to be careful and choose this initialization uniformly with respect to tt, in order to avoid introducing a constant time dependent potential into the equations via A0A_{0}. Since in our results we start with data which converges asymptotically to k\vec{k} as rr\to\infty, and the solutions continue to have this property, it is natural to fix the choice of v¯\bar{v} and w¯\bar{w} at infinity,

(4.11) limrv¯(r,t)=i,limrw¯(r,t)=j.\lim_{r\to\infty}\bar{v}(r,t)=\vec{i},\qquad\lim_{r\to\infty}\bar{w}(r,t)=\vec{j}.

The justification of the fact that we can impose these conditions at \infty and find the gauge, or equivalently uniquely solve (4.9) with this boundary condition for v¯\bar{v}, is provided in the proof of Theorem 3.2 in [6].

Before considering the general case we begin with the solitons. The simplest case is when u=Qmu=Q^{m}, when the triplet (v¯,w¯,u¯)(\bar{v},\bar{w},\bar{u}) is given by

(4.12) (V¯m,W¯m,Q¯m)=(h3m(r)0h1m(r)010h1m(r)0h3m(r)).\left(\bar{V}^{m},\bar{W}^{m},\bar{Q}^{m}\right)=\left(\begin{array}[]{ccc}h_{3}^{m}(r)&0&h_{1}^{m}(r)\cr 0&1&0\cr-h_{1}^{m}(r)&0&h_{3}^{m}(r)\end{array}\right).

More generally, if u=Qα,λmu=Q^{m}_{\alpha,\lambda} then from the above formula, by rescaling and rotation, we obtain the corresponding triplet (V¯α,λm,W¯α,λm,Q¯α,λm)\left(\bar{V}^{m}_{\alpha,\lambda},\bar{W}^{m}_{\alpha,\lambda},\bar{Q}^{m}_{\alpha,\lambda}\right) of the form

(h3m(λr)cos2mα+sin2mα(h3m(λr)1)sinmαcosmαh1m(λr)cosmα(h3m(λr)1)sinmαcosmαh3m(λr)sin2mα+cos2mαh1m(λr)sinmαh1m(λr)cosmαh1m(λr)sinmαh3m(λr)).\!\left(\!\!\!\begin{array}[]{ccc}\!h_{3}^{m}(\lambda r)\cos^{2}m\alpha+\sin^{2}m\alpha&\!(h_{3}^{m}(\lambda r)-1)\sin m\alpha\cos m\alpha&h_{1}^{m}(\lambda r)\cos m\alpha\cr\!(h_{3}^{m}(\lambda r)-1)\sin m\alpha\cos m\alpha\!&\!h_{3}^{m}(\lambda r)\sin^{2}m\alpha+\cos^{2}m\alpha&h_{1}^{m}(\lambda r)\sin m\alpha\cr\!-h_{1}^{m}(\lambda r)\cos m\alpha&\!-h_{1}^{m}(\lambda r)\sin m\alpha&h_{3}^{m}(\lambda r)\end{array}\!\!\!\right).\!

For later reference we also note the values of ψ1\psi_{1}, ψ2\psi_{2} and A2A_{2} in this case:

(4.13) ψα,λ,1m=mr1h1m(λr)eimα,ψα,λ,2m=imh1m(λr)eimα,Aα,λ,2m=mh3m(λr).\begin{split}\psi_{\alpha,\lambda,1}^{m}=-mr^{-1}h_{1}^{m}(\lambda r)e^{im\alpha},&\quad\psi_{\alpha,\lambda,2}^{m}=imh_{1}^{m}(\lambda r)e^{im\alpha},\\ A_{\alpha,\lambda,2}^{m}=&mh_{3}^{m}(\lambda r).\end{split}

In this article we work with maps u:2𝕊2u:\mathbb{R}^{2}\to\mathbb{S}^{2} which are near a soliton Qα,λmQ^{m}_{\alpha,\lambda} in the sense that

(4.14) uQα,λmH˙1δ1.\|u-Q_{\alpha,\lambda}^{m}\|_{\dot{H}^{1}}\leq\delta\ll 1.

The following result had been established in [6].

Lemma 4.1.

Let m1m\geq 1 and u:2𝕊2u:\mathbb{R}^{2}\to\mathbb{S}^{2} be an mm-equivariant map which satisfies (4.14). Then

(4.15) limr0u(r,θ)=k,limru(r,θ)=k\lim_{r\to 0}u(r,\theta)=-\vec{k},\qquad\lim_{r\to\infty}u(r,\theta)=\vec{k}

and

(4.16) r1(uQα,λm)L2+uQα,λmLδ.\|r^{-1}(u-Q_{\alpha,\lambda}^{m})\|_{L^{2}}+\|u-Q_{\alpha,\lambda}^{m}\|_{L^{\infty}}\lesssim\delta.

To measure the regularity of the frame (v,w)(v,w) we use the Sobolev type space H˙C1\dot{H}^{1}_{C} of functions f:23f:\mathbb{R}^{2}\to\mathbb{R}^{3}, with norm

fH˙C1=rf¯L2+f¯L+r1f¯3L2,f(r,θ)=emθRf¯(r)\|f\|_{\dot{H}^{1}_{C}}=\|\partial_{r}\bar{f}\|_{L^{2}}+\|\bar{f}\|_{L^{\infty}}+\|r^{-1}\bar{f}_{3}\|_{L^{2}},\qquad f(r,\theta)=e^{m\theta R}\bar{f}(r)

The next proposition, which was also established in [6], shows that the initialization (4.11) is well-defined for arbitrary maps uu close to the soliton family:

Proposition 4.2.

a) For each mm-equivariant map u:2𝕊2u:\mathbb{R}^{2}\to\mathbb{S}^{2} satisfying (4.14) there exists an unique mm-equivariant orthonormal frame (v,w)(v,w) which satisfies the Coulomb gauge condition (4.8) and the boundary condition (4.11). This frame satisfies the bounds

(4.17) vVα,λmH˙C1+wWα,λmH˙C1δ.\|v-V^{m}_{\alpha,\lambda}\|_{\dot{H}^{1}_{C}}+\|w-W^{m}_{\alpha,\lambda}\|_{\dot{H}^{1}_{C}}\lesssim\delta.

b) Furthermore, the maps uv,wu\to v,w are C1C^{1} from H1˙\dot{H^{1}} into H˙C1\dot{H}^{1}_{C} as well as from L2L2L^{2}\to L^{2}.

As a direct consequence of part (a) of the above proposition, we can describe the regularity and properties of the differentiated fields ψ1\psi_{1}, ψ2\psi_{2} and the connection coefficient A2A_{2} at fixed time:

Corollary 4.3.

Let u:2𝕊2u:\mathbb{R}^{2}\to\mathbb{S}^{2} be an mm-equivariant map as in (4.14). Then ψ1\psi_{1}, ψ2\psi_{2} and A2A_{2} satisfy (4.3), (4.4) for k,l=1,2k,l=1,2 as well as the bounds

ψ1ψα,λ,1mL2+ψ2ψα,λ,2mH˙e1+A2Aα,λ,2mH˙e1δ.\|\psi_{1}-\psi_{\alpha,\lambda,1}^{m}\|_{L^{2}}+\|\psi_{2}-\psi_{\alpha,\lambda,2}^{m}\|_{\dot{H}^{1}_{e}}+\|A_{2}-A_{\alpha,\lambda,2}^{m}\|_{\dot{H}^{1}_{e}}\lesssim\delta.

In addition, the map u(ψ1,ψ2,A2)u\to(\psi_{1},\psi_{2},A_{2}) from H˙1\dot{H}^{1} into the above spaces is C1C^{1}.

A second step is to consider Schrödinger maps with more regularity; this is particularly useful in order to justify formal computations. As a consequence of part (b) of Proposition 4.2 we have:

Corollary 4.4.

Let II be a compact interval, and u:I×2𝕊2u:I\times\mathbb{R}^{2}\to\mathbb{S}^{2} be an mm-equivariant map satisfying (4.14) uniformly in II and which has the additional regularity

uC(I;H˙2),tuC(I;L2).u\in C(I;\dot{H}^{2}),\qquad\partial_{t}u\in C(I;L^{2}).

Then ψ0\psi_{0}, ψ1\psi_{1}, ψ2\psi_{2} and A0A_{0}, A2A_{2} satisfy the relations (4.3), (4.4) for k,l=0,1,2k,l=0,1,2 and have the additional regularity

(4.18) ψ0,A0C(I;L2),ψ1ψα,λ,1mC(I;H˙e1),ψ2ψα,λ,2m,A2Aα,λ,2mC(I;H˙e2).\begin{split}&\psi_{0},A_{0}\in C(I;L^{2}),\psi_{1}-\psi_{\alpha,\lambda,1}^{m}\in C(I;\dot{H}^{1}_{e}),\\ &\psi_{2}-\psi^{m}_{\alpha,\lambda,2},A_{2}-A^{m}_{\alpha,\lambda,2}\in C(I;\dot{H}^{2}_{e}).\end{split}

This last result is almost identical to the corresponding one in [6], except that we remove the (low frequency) hypothesis uQα,λmL2u-Q^{m}_{\alpha,\lambda}\in L^{2} and adjust the conclusion in (4.18) by removing the associated (low frequency) information ψ2ψα,λ,2m,A2Aα,λ,2mC(I;L2)\psi_{2}-\psi^{m}_{\alpha,\lambda,2},A_{2}-A^{m}_{\alpha,\lambda,2}\in C(I;L^{2}).

In practice, in order to ensure the additional regularity in the corollary above (namely uC(I;H˙2),tuC(I;L2)u\in C(I;\dot{H}^{2}),\partial_{t}u\in C(I;L^{2})) for solutions to the Schrödinger map equation, it suffices to assume that uC(I;H˙1H˙2)u\in C(I;\dot{H}^{1}\cap\dot{H}^{2}), since the second part follows from the equation and the LL^{\infty} embedding.

4.3. Schrödinger maps in the Coulomb gauge

At this point we are ready to write the evolution equations for the differentiated fields ψ1\psi_{1} and ψ2\psi_{2} in (4.1) computed with respect to the Coulomb gauge.

Writing the Laplacian in polar coordinates, a direct computation using the formulas (4.1) shows that we can rewrite the Schrödinger Map equation (1.1) in the form

ψ0=i(D1ψ1+1rψ1+1r2D2ψ2).\psi_{0}=i\left(D_{1}\psi_{1}+\frac{1}{r}\psi_{1}+\frac{1}{r^{2}}D_{2}\psi_{2}\right).

Applying the operators D1D_{1} and D2D_{2} to both sides of this equation and using the relations (4.3) and (4.4), we can derive the evolution equations for ψm\psi_{m}, m=1,2m=1,2:

(4.19) tψ1+iA0ψ1=iΔψ12A11ψ11A1ψ11rA1ψ1iA12ψ1i1r2A22ψ1i1r2ψ1+2r3A2ψ21r2(ψ1ψ¯2)ψ2,tψ2+iA0ψ2=iΔψ22A11ψ21A1ψ21rA1ψ2iA12ψ2i1r2A22ψ2(ψ2ψ¯1)ψ1.\begin{split}\partial_{t}\psi_{1}+iA_{0}\psi_{1}=&\ i\Delta\psi_{1}-2A_{1}\partial_{1}\psi_{1}-\partial_{1}A_{1}\psi_{1}-\frac{1}{r}A_{1}\psi_{1}\\ &\ -iA_{1}^{2}\psi_{1}-i\frac{1}{r^{2}}A_{2}^{2}\psi_{1}-i\frac{1}{r^{2}}\psi_{1}+\frac{2}{r^{3}}A_{2}\psi_{2}-\frac{1}{r^{2}}\Im{(\psi_{1}\bar{\psi}_{2})}\psi_{2},\\ \partial_{t}\psi_{2}+iA_{0}\psi_{2}=&\ i\Delta\psi_{2}-2A_{1}\partial_{1}\psi_{2}-\partial_{1}A_{1}\psi_{2}-\frac{1}{r}A_{1}\psi_{2}\\ &\ -iA_{1}^{2}\psi_{2}-i\frac{1}{r^{2}}A_{2}^{2}\psi_{2}-\Im{(\psi_{2}\bar{\psi}_{1})}\psi_{1}.\end{split}

Under the Coulomb gauge A1=0A_{1}=0 these equations become

tψ1+iA0ψ1=iΔψ1i1r2A22ψ1i1r2ψ1+2r3A2ψ21r2(ψ1ψ¯2)ψ2,tψ2+iA0ψ2=iΔψ2i1r2A22ψ2(ψ2ψ¯1)ψ1.\begin{split}\partial_{t}\psi_{1}+iA_{0}\psi_{1}=&i\Delta\psi_{1}-i\frac{1}{r^{2}}A_{2}^{2}\psi_{1}-i\frac{1}{r^{2}}\psi_{1}+\frac{2}{r^{3}}A_{2}\psi_{2}-\frac{1}{r^{2}}\Im{(\psi_{1}\bar{\psi}_{2})}\psi_{2},\\ \partial_{t}\psi_{2}+iA_{0}\psi_{2}=&i\Delta\psi_{2}-i\frac{1}{r^{2}}A_{2}^{2}\psi_{2}-\Im{(\psi_{2}\bar{\psi}_{1})}\psi_{1}.\end{split}

while the relations (4.3) and (4.4) become an ode system for (A2,ψ2)(A_{2},\psi_{2}), namely

(4.20) rA2=(ψ1ψ¯2),rψ2=iA2ψ1.\partial_{r}A_{2}=\Im{(\psi_{1}\bar{\psi}_{2})},\qquad\partial_{r}\psi_{2}=iA_{2}\cdot\psi_{1}.

On the other hand from the compatibility relations involving A0A_{0} we obtain

(4.21) rA0=12r2r(r2|ψ1|2|ψ2|2),\partial_{r}A_{0}=-\frac{1}{2r^{2}}\partial_{r}(r^{2}|\psi_{1}|^{2}-|\psi_{2}|^{2}),

which determines A0A_{0} modulo constants. This allows us to derive an expression for A0A_{0},

(4.22) A0=12(|ψ1|21r2|ψ2|2)+[rr]1(|ψ1|21r2|ψ2|2)A_{0}=-\frac{1}{2}\left(|\psi_{1}|^{2}-\frac{1}{r^{2}}|\psi_{2}|^{2}\right)+[r\partial_{r}]^{-1}\left(|\psi_{1}|^{2}-\frac{1}{r^{2}}|\psi_{2}|^{2}\right)

where we recall the definition of [rr]1f(r)[r\partial_{r}]^{-1}f(r) from (2.7). To justify (4.22), we are using the fact that [rr]1[r\partial_{r}]^{-1} is bounded on L1L^{1} and that |ψ1|21r2|ψ2|2L1|\psi_{1}|^{2}-\frac{1}{r^{2}}|\psi_{2}|^{2}\in L^{1}; this guarantees that the right-hand side of (4.22) belongs to L1L^{1} (and also to r1Lr^{-1}L^{\infty}). To dispense with the additional possible constant in A0A_{0} we simply note that an additional regularity assumption on the map uu in this problem gives us that A0L2A_{0}\in L^{2}, see Corollary 4.4. Since a constant is not integrable in any sense, it follows that the only solution of (4.21) which has some decay at \infty must be the one in (4.22).

There is quite a bit of redundancy in the equations for ψ1\psi_{1} and ψ2\psi_{2}; we eliminate this by introducing a single differentiated field ψ\psi by

(4.23) ψ=ψ1iψ2r.\psi=\psi_{1}-i\frac{\psi_{2}}{r}.

The size of ψ\psi is closely related to the energy of the original map uu, precisely

(4.24) πψL22=E(u)4πm,\pi\|\psi\|_{L^{2}}^{2}=E(u)-4\pi m,

and in particular ψ\psi vanishes iff uu is a soliton.

A direct computation yields the Schrödinger type equation for ψ\psi:

itψ+Δψ=A0ψ2A2r2ψ+1r2ψ+A22r2ψ1r(ψ2ψ¯1)ψ.i\partial_{t}\psi+\Delta\psi=A_{0}\psi-2\frac{A_{2}}{r^{2}}\psi+\frac{1}{r^{2}}\psi+\frac{A_{2}^{2}}{r^{2}}\psi-\frac{1}{r}\Im{(\psi_{2}\bar{\psi}_{1})}\psi.

By replacing ψ1=ψ+ir1ψ2\psi_{1}=\psi+ir^{-1}\psi_{2} and using A22+|ψ2|2=m2A_{2}^{2}+|\psi_{2}|^{2}=m^{2}, we obtain the key evolution equation we work with in this paper, namely

(4.25) itψ+Δψ1+m2r2ψ=A0ψ2A2r2ψ1r(ψ2ψ¯)ψ.i\partial_{t}\psi+\Delta\psi-\frac{1+m^{2}}{r^{2}}\psi=A_{0}\psi-2\frac{A_{2}}{r^{2}}\psi-\frac{1}{r}\Im{(\psi_{2}\bar{\psi})}\psi.

Our strategy will be to use this equation in order to obtain estimates for ψ\psi. The functions A2A_{2} and ψ2\psi_{2} are related to ψ\psi via the system of ODE’s

(4.26) {rA2=(ψψ¯2)+1r|ψ2|2,rψ2=iA2ψ1rA2ψ2\left\{\begin{array}[]{l}\partial_{r}A_{2}=\Im{(\psi\bar{\psi}_{2})}+\dfrac{1}{r}|\psi_{2}|^{2},\cr\cr\partial_{r}\psi_{2}=iA_{2}\psi-\dfrac{1}{r}A_{2}\psi_{2}\end{array}\right.

which is derived from (4.20), together with the compatibility condition from (4.7),

(4.27) A22+|ψ2|2=m2.A_{2}^{2}+|\psi_{2}|^{2}=m^{2}.

However, they are not uniquely determined by ψ\psi, instead we have two additional degree of freedom, which correspond to prescribing initial data in the above ode subject to the constraint provided by (4.27). This will lead us to the modulation parameters described in the next subsection.

Once (A2,ψ2)(A_{2},\psi_{2}) are given, the A0A_{0} connection coefficient is given by (4.22) which now becomes

(4.28) A0(r)=12|ψ|2+1r(ψ2ψ¯)+[rr]1(|ψ|22r(ψ2ψ¯)).A_{0}(r)=-\frac{1}{2}|\psi|^{2}+\frac{1}{r}\Im(\psi_{2}\bar{\psi})+[r\partial_{r}]^{-1}(|\psi|^{2}-\frac{2}{r}\Im(\psi_{2}\bar{\psi})).

Finally, given ψ\psi, A2A_{2} and ψ2\psi_{2}, we can return to the Schrödinger map uu via the system (4.2) with the boundary condition at infinity given by (4.11).

4.4. The modulation parameters λ(t),α(t)\lambda(t),\alpha(t)

Our paper is concerned with the behaviour of maps near the Q±2Q^{\pm 2} soliton (this is meant to include the more general objects Qα,λ±2Q^{\pm 2}_{\alpha,\lambda}). The cases m=2m=2 and m=2m=-2 can be identified via a reflection, therefore, beginning with this subsection, we simply set m=2m=2.

This allows us to drop the upper script mm from h1mh_{1}^{m} and h3mh_{3}^{m} and simply use h1,h3h_{1},h_{3}. This allows us to introduce another upper script convention

h1λ(r)=h1(λr),h3λ(r)=h3(λr),h_{1}^{\lambda}(r)=h_{1}(\lambda r),\qquad h_{3}^{\lambda}(r)=h_{3}(\lambda r),

which is very useful due to the key role the scaling parameter λ\lambda plays in our analysis.

Given a 22-equivariant map u:2𝕊2u:\mathbb{R}^{2}\to\mathbb{S}^{2} in the homotopy class of 𝒬e2\mathcal{Q}^{2}_{e} and satisfying 0E(u)E(𝒬2)10\leq E(u)-E(\mathcal{Q}^{2})\ll 1, we recall that, by (1.6), uu must be close in H1H^{1} to one of the solitons Qα,λ2Q^{2}_{\alpha,\lambda}:

distH˙1(u,𝒬e2)2=infα,λQα,λ2uH˙12E(u)E(𝒬2)=E(u)8π.\text{dist}_{\dot{H}^{1}}(u,\mathcal{Q}^{2}_{e})^{2}=\inf_{\alpha,\lambda}\|Q^{2}_{\alpha,\lambda}-u\|_{\dot{H}^{1}}^{2}\lesssim E(u)-E(\mathcal{Q}^{2})=E(u)-8\pi.

For such uu, it is important to identify a specific soliton Qα,λ2Q^{2}_{\alpha,\lambda} to be the ”closest” soliton to uu. We will think of the chosen (α,λ)(\alpha,\lambda) as the modulation parameters associated to uu, and later on we will study their evolution as uu evolves along the Schrödinger map flow. At the same time, we want to use (α,λ)(\alpha,\lambda) to remove the two remaining degrees of freedom in the choice of (ψ2,A2)(\psi_{2},A_{2}) in the previous section.

While the soliton Qα,λ2Q^{2}_{\alpha,\lambda} does not need to be the minimizer of the above distance, a natural condition to impose is that

Qα,λ2uH˙12E(u)8π.\|Q^{2}_{\alpha,\lambda}-u\|_{\dot{H}^{1}}^{2}\lesssim E(u)-8\pi.

This still leaves us with infinitely many choices for the parameters α\alpha and λ\lambda and it is important to seek an efficient rule that selects a unique choice for the two parameters. We will select our parameters based on the following orthogonality condition

(4.29) ψ22ie2iα(t)h1(λr),ϰ(λr)=0,\langle\psi_{2}-2ie^{2i\alpha(t)}h_{1}(\lambda r),\varkappa(\lambda r)\rangle=0,

where ϰ\varkappa can be subject to various choices. Next we discuss how these choices were made in prior works on this problem, and then what is our choice here.

One natural choice would have been to choose ϰ=h1\varkappa=h_{1} (or better h12h_{1}^{2} in the more general context of mm-equivariant setup which we discuss here), which is the eigenvalue of the linearized operator HλH_{\lambda}. Such a choice (that is, setting ϰ=h1m\varkappa=h_{1}^{m}) works well for m4m\geq 4, see [13]. However the finite energy bound for uu corresponds to r1(ψ22ie2iα(t)h1(λr))L2r^{-1}(\psi_{2}-2ie^{2i\alpha(t)}h_{1}(\lambda r))\in L^{2}, therefore in order to make sense of (4.29) we would need rϰL2r\varkappa\in L^{2}. But rh1mL2rh_{1}^{m}\notin L^{2} when m=2m=2, therefore the choice ϰ=h12\varkappa=h_{1}^{2} is unsuitable for us.

Another choice would be a point-type condition by simply setting ϰ=δr=1\varkappa=\delta_{r=1}. This has been used in the case m=1m=1 by the first and third author, see [6]. However there are several elements of the analysis in [6] that are not reproducible in the case m=2m=2, and this is why, although the point condition will be partially used in our analysis, it will not be the one we ultimately choose to determine our parameters α\alpha and λ\lambda.

Our choice here (that is in our context with m=2m=2) is similar to the one used by Gustafson, Nakanishi, Tsai [14]: ϰ\varkappa is a smooth function, compactly supported in [12,2][\frac{1}{2},2] and subject to the following two nondegeneracy conditions:

(4.30) ϰ,h10,ϰ,h1h30.\langle\varkappa,h_{1}\rangle\neq 0,\qquad\langle\varkappa,h_{1}h_{3}\rangle\neq 0.

Thus there is a lot of freedom in choosing ϰ\varkappa. However we note that the actual orthogonality condition used in [14] is different than ours - in [14] the orthogonality involves a linearization of the actual map, while our condition (4.29) involves the linearization of the gauge components.

To keep formulas shorter in the remaining of this section and the rest of the paper, we introduce the notation

δλ,αψ2=ψ22ie2iαh1λ,δλA2=A22h3λ.\delta^{\lambda,\alpha}\psi_{2}=\psi_{2}-2ie^{2i\alpha}h_{1}^{\lambda},\quad\delta^{\lambda}A_{2}=A_{2}-2h_{3}^{\lambda}.

The first issue that needs to be addressed is the existence of modulation parameters α\alpha and λ\lambda satisfying the condition (4.29). For this purpose we prove the following

Proposition 4.5 (Modulation parameters).

Assume that ϰ\varkappa satisfies the nondegeneracy conditions (4.30). Then the following properties hold for δ>0\delta>0 small enough:

i) Given any 2-equivariant map u:2𝕊2u:\mathbb{R}^{2}\rightarrow\mathbb{S}^{2}, uH˙1u\in\dot{H}^{1} satisfying (1.5) and such that E(u)8πδ2E(u)-8\pi\leq\delta^{2}, there exist a unique pair of modulation parameters (α,λ)[0,2π)×+(\alpha,\lambda)\in[0,2\pi)\times\mathbb{R}^{+} such that

(4.31) δλ,αψ2,ϰλ=0\langle\delta^{\lambda,\alpha}\psi_{2},\varkappa^{\lambda}\rangle=0

and

(4.32) δλ,αψ2Lδ.\|\delta^{\lambda,\alpha}\psi_{2}\|_{L^{\infty}}\lesssim\delta.

Further, (α,λ)(\alpha,\lambda) have a Lipschitz dependence on ψ2\psi_{2} in the following sense:

(4.33) |lnλ(ψ2)lnλ(ψ~2)|+|α(ψ2)α(ψ~2)|min{λmax1sψ2ψ~2H˙es;s=1,0,1},|\ln\lambda(\psi_{2})-\ln\lambda(\tilde{\psi}_{2})|+|\alpha(\psi_{2})-\alpha(\tilde{\psi}_{2})|\lesssim\min\{\lambda_{max}^{1-s}\|\psi_{2}-\tilde{\psi}_{2}\|_{\dot{H}^{s}_{e}};s=1,0,-1\},

where λmax=max(λ(ψ2),λ(ψ~2))\lambda_{max}=\max(\lambda(\psi_{2}),\lambda(\tilde{\psi}_{2})).

ii) Assume that the modulation parameters (α,λ)(\alpha,\lambda) are chosen as in (4.31) and (4.32). Then the following holds true

(4.34) δλ,αψ2H˙e1+δλA2H˙e1ψL2.\|\delta^{\lambda,\alpha}\psi_{2}\|_{\dot{H}^{1}_{e}}+\|\delta^{\lambda}A_{2}\|_{\dot{H}^{1}_{e}}\lesssim\|\psi\|_{L^{2}}.

As a consequence these modulation parameters are a good choice in the sense of (1.7), that is

(4.35) uQα,λ2H˙1ψL2.\|u-Q_{\alpha,\lambda}^{2}\|_{\dot{H}^{1}}\lesssim\|\psi\|_{L^{2}}.

Furthermore, under additional hypothesis, we have the following estimates

  • if ψL4\psi\in L^{4} (not necessarily small) then the following holds true

    (4.36) δλ,αψ2rL4+δλA2rL4ψL4.\|\frac{\delta^{\lambda,\alpha}\psi_{2}}{r}\|_{L^{4}}+\|\frac{\delta^{\lambda}A_{2}}{r}\|_{L^{4}}\lesssim\|\psi\|_{L^{4}}.
  • if ψrL2\dfrac{\psi}{r}\in L^{2} (not necessarily small) then the following holds true

    (4.37) δλ,αψ2r2L2+δλA2r2L2ψrL2.\|\frac{\delta^{\lambda,\alpha}\psi_{2}}{r^{2}}\|_{L^{2}}+\|\frac{\delta^{\lambda}A_{2}}{r^{2}}\|_{L^{2}}\lesssim\|\frac{\psi}{r}\|_{L^{2}}.
Remark 1.

The pointwise smallness condition in (4.32) is needed in order to guarantee that we choose the correct pair (α,λ)(\alpha,\lambda), in particular they satisfy (4.35).

We also have a natural converse for the proposition above:

Proposition 4.6 (Modulation parameters - part 2).

Assume that ϰ\varkappa satisfies the nondegeneracy conditions (4.30) and δ>0\delta>0 is small enough. Then given any radial map ψL2\psi\in L^{2} such that ψL2δ\|\psi\|_{L^{2}}\leq\delta, and any pair of modulation parameters (α,λ)[0,2π)×+(\alpha,\lambda)\in[0,2\pi)\times\mathbb{R}^{+}, there exists a unique solution (ψ2,A2)(\psi_{2},A_{2}) for the system (4.20) so that both the orthogonality relation (4.31) and the bound (4.32) hold. In addition there exists an unique 22-equivariant map u:2𝕊2u:\mathbb{R}^{2}\rightarrow\mathbb{S}^{2}, uH˙1u\in\dot{H}^{1} satisfying (1.5) and E(u)8πδ2E(u)-8\pi\leq\delta^{2}, and with the property that ψ,ψ2,A2\psi,\psi_{2},A_{2} are its Coulomb gauge components as constructed in Section 4.3 and α,λ\alpha,\lambda are its correct modulation parameters as described in Proposition 4.5.

Proof of Proposition 4.5.

i) Following [14], the strategy here is to actually establish the existence of the parameters subject to the point-type condition (when ϰ=δr=1\varkappa=\delta_{r=1}) and then use the inverse function theorem to find the parameters for the smooth choice of ϰ\varkappa. In addition to the intrinsic technical advantages of this approach, this will also provide us with an easy way to prove the estimates claimed in ii).

Given the above strategy, the first step is to observe that we can find parameters α0,λ0\alpha_{0},\lambda_{0} so that

ψ2(λ01)=2ie2iα0h1(1)=2ie2iα0.\psi_{2}(\lambda_{0}^{-1})=2ie^{2i\alpha_{0}}h_{1}(1)=2ie^{2i\alpha_{0}}.

To see this, we note that limr0A2(r)=2\lim_{r\rightarrow 0}A_{2}(r)=-2 and limrA2(r)=2\lim_{r\rightarrow\infty}A_{2}(r)=2 and the continuity of A2A_{2} imply that there exists λ0\lambda_{0} with A2(λ01)=0A_{2}(\lambda_{0}^{-1})=0. Then the compatibility relation (4.7) implies that |ψ2(λ01)|=2|\psi_{2}(\lambda_{0}^{-1})|=2, from which it follows that there exists α0\alpha_{0} such that ψ2(λ01)=2ie2iα0\psi_{2}(\lambda_{0}^{-1})=2ie^{2i\alpha_{0}}.

The next step is to examine δλ,αψ2\delta^{\lambda,\alpha}\psi_{2} and δλA2\delta^{\lambda}A_{2} corresponding to this pointwise choice of α\alpha and λ\lambda. For further use, we phrase the result in a slightly more general fashion:

Lemma 4.7.

Consider the system (4.26) with ψL2\psi\in L^{2}, small, and initial data

(4.38) δλ,αψ2(r0)=0,δλA2(r0)=0.\delta^{\lambda,\alpha}\psi_{2}(r_{0})=0,\qquad\delta^{\lambda}A_{2}(r_{0})=0.

for some r0λ1r_{0}\approx\lambda^{-1}. Then the system admits a unique solution (ψ2,A2)(\psi_{2},A_{2}) with regularity (δλ,αψ2,δλA2)H˙e1×H˙e1(\delta^{\lambda,\alpha}\psi_{2},\delta^{\lambda}A_{2})\in\dot{H}^{1}_{e}\times\dot{H}^{1}_{e} and which satisfies (4.27). Furthermore, this solution satisfies the following H˙e1\dot{H}^{1}_{e} bound:

(4.39) δλ,αψ2H˙e1+δλA2H˙e1ψL2.\|\delta^{\lambda,\alpha}\psi_{2}\|_{\dot{H}^{1}_{e}}+\|\delta^{\lambda}A_{2}\|_{\dot{H}^{1}_{e}}\lesssim\|\psi\|_{L^{2}}.

If in addition ψL4\psi\in L^{4} (not necessarily small) then the following holds true:

(4.40) δλ,αψ2rL4+δλA2rL4ψL4.\|\frac{\delta^{\lambda,\alpha}\psi_{2}}{r}\|_{L^{4}}+\|\frac{\delta^{\lambda}A_{2}}{r}\|_{L^{4}}\lesssim\|\psi\|_{L^{4}}.

If in addition ψrL2\frac{\psi}{r}\in L^{2} (not necessarily small) then the following holds true:

(4.41) δλ,αψ2r2L2+δλA2r2L2ψrL2.\|\frac{\delta^{\lambda,\alpha}\psi_{2}}{r^{2}}\|_{L^{2}}+\|\frac{\delta^{\lambda}A_{2}}{r^{2}}\|_{L^{2}}\lesssim\|\frac{\psi}{r}\|_{L^{2}}.
Proof.

We first establish (4.39). Using the compatibility condition (4.27), the equations (4.26) become

(4.42) {Lλδλ,αψ2= 2ih3λψ+δλA2ψ1rδλA2(2ie2iαh1λ+δλ,αψ2),LλδλA2=2h1λ(e2iαψ)+(ψδλ,αψ¯2)1r(δλA2)2.\left\{\begin{aligned} L_{\lambda}\delta^{\lambda,\alpha}\psi_{2}=&\ 2ih_{3}^{\lambda}\psi+\delta^{\lambda}A_{2}\psi-\frac{1}{r}\delta^{\lambda}A_{2}(2ie^{2i\alpha}h_{1}^{\lambda}+\delta^{\lambda,\alpha}\psi_{2}),\\ L_{\lambda}\delta^{\lambda}A_{2}=&\ -2h_{1}^{\lambda}\Re{(e^{2i\alpha}\psi)}+\Im(\psi\overline{\delta^{\lambda,\alpha}\psi}_{2})-\frac{1}{r}(\delta^{\lambda}A_{2})^{2}.\end{aligned}\right.

with LλL_{\lambda} as in (3.10). The solution to the homogeneous LλL_{\lambda} equation is given by h1λh_{1}^{\lambda}, so the equations above can be rewritten in the integral form:

δλ,αψ2(r)=h1λ(r)r0r1h1λ(2ih3λψ+δλA2ψ1sδλA2(2ie2iαh1λ+δλ,αψ2))𝑑sδλA2=h1λ(r)r0r1h1λ(2h1λ(e2iαψ)+(ψδλ,αψ¯2)(δλA2)2s)𝑑s.\begin{split}&\delta^{\lambda,\alpha}\psi_{2}(r)=h^{\lambda}_{1}(r)\int_{r_{0}}^{r}\frac{1}{h_{1}^{\lambda}}\left(2ih_{3}^{\lambda}\psi+\delta^{\lambda}A_{2}\psi-\frac{1}{s}\delta^{\lambda}A_{2}(2ie^{2i\alpha}h_{1}^{\lambda}+\delta^{\lambda,\alpha}\psi_{2})\right)ds\\ &\delta^{\lambda}A_{2}=h^{\lambda}_{1}(r)\int_{r_{0}}^{r}\frac{1}{h_{1}^{\lambda}}\left(-2h^{\lambda}_{1}\Re{(e^{2i\alpha}\psi)}+\Im(\psi\overline{\delta^{\lambda,\alpha}\psi}_{2})-\frac{(\delta^{\lambda}A_{2})^{2}}{s}\right)ds.\end{split}

As a tool to obtain bounds on the solutions, we record the following simple inequality,

(4.43) 1rh1λr0r1h1λf𝑑sLpfLp,1p,\|\frac{1}{r}h_{1}^{\lambda}\int_{r_{0}}^{r}\frac{1}{h_{1}^{\lambda}}fds\|_{L^{p}}\lesssim\|f\|_{L^{p}},\qquad 1\leq p\leq\infty,

which in the particular case p=2p=2 can be augmented to

(4.44) 1rh1λr0r1h1λf𝑑sL2+h1λr0r1h1λf𝑑sLfL2,1p.\|\frac{1}{r}h_{1}^{\lambda}\int_{r_{0}}^{r}\frac{1}{h_{1}^{\lambda}}fds\|_{L^{2}}+\|h_{1}^{\lambda}\int_{r_{0}}^{r}\frac{1}{h_{1}^{\lambda}}fds\|_{L^{\infty}}\lesssim\|f\|_{L^{2}},\qquad 1\leq p\leq\infty.

The only role that r0r_{0} plays here is that is sits at the ”peak” of h1λh_{1}^{\lambda}, that is h1λ(r0)1h_{1}^{\lambda}(r_{0})\approx 1 and h1λh_{1}^{\lambda} decays away from a neighborhood of size λ1\lambda^{-1} of r0r_{0}.

Under the assumption that ψL21\|\psi\|_{L^{2}}\ll 1, the above inequality allows us to obtain the solutions (δλ,αψ2,δλA2)(\delta^{\lambda,\alpha}\psi_{2},\delta^{\lambda}A_{2}) to the above integral system using the contraction principle in the space YY with norm

fY=fL+r1fL2.\|f\|_{Y}=\|f\|_{L^{\infty}}+\|r^{-1}f\|_{L^{2}}.

Returning to (4.42), the L2L^{2} bound for r(δλ,αψ2,δλA2)\partial_{r}(\delta^{\lambda,\alpha}\psi_{2},\delta^{\lambda}A^{2}) also follows, completing the proof of the H˙e1\dot{H}^{1}_{e} bound in the lemma. The uniqueness of the (small) solution (δλ,αψ2H˙e1×H˙e1,δλA2)(\delta^{\lambda,\alpha}\psi_{2}\dot{H}^{1}_{e}\times\dot{H}^{1}_{e},\delta^{\lambda}A_{2}) in H˙e1×H˙e1\dot{H}^{1}_{e}\times\dot{H}^{1}_{e} follows from the above fixed point argument.

The argument for (4.40) is entirely similar: it uses (4.43) with p=4p=4 and the already established smallness of δλ,αψ2L+δλA2L\|\delta^{\lambda,\alpha}\psi_{2}\|_{L^{\infty}}+\|\delta^{\lambda}A_{2}\|_{L^{\infty}}.

To obtain (4.41), we rely instead on the following estimate in the special case p=2p=2,

(4.45) 1r2h1λr0r1h1λf𝑑sLpfrLp,1p<,\|\frac{1}{r^{2}}h_{1}^{\lambda}\int_{r_{0}}^{r}\frac{1}{h_{1}^{\lambda}}fds\|_{L^{p}}\lesssim\|\frac{f}{r}\|_{L^{p}},\qquad 1\leq p<\infty,

and proceed just as above.

Now we return to our initial guess (α0,λ0)(\alpha_{0},\lambda_{0}). From Lemma 4.7 above it follows that the functions (ψ2,A2)(\psi_{2},A_{2}) satisfy

(4.46) ψ22ie2iα0h1λ0H˙e1+A2h3λ0H˙e1ψL2.\|\psi_{2}-2ie^{2i\alpha_{0}}h_{1}^{\lambda_{0}}\|_{\dot{H}^{1}_{e}}+\|A_{2}-h_{3}^{\lambda_{0}}\|_{\dot{H}^{1}_{e}}\lesssim\|\psi\|_{L^{2}}.

We use this as as starting point in order to establish the existence of modulation parameters satisfying the orthogonality condition (4.31) with the smooth choice of ϰ\varkappa. Here it is convenient to work on an exponential scale for λ\lambda, so we set λ=eγλ0\lambda=e^{\gamma}\lambda_{0}. Then it is natural to consider the following function

F(ψ2,α,γ)=λ2ψ22ie2iαh1(λr),ϰ(λr).F(\psi_{2},\alpha,\gamma)=\lambda^{2}\langle\psi_{2}-2ie^{2i\alpha}h_{1}(\lambda r),\varkappa(\lambda r)\rangle.

and seek a zero of FF near (α0,0)(\alpha_{0},0). From (4.46) it follows that

ψ22ie2iα0h1(λ0r)rL2ψL2,\|\frac{\psi_{2}-2ie^{2i\alpha_{0}}h_{1}(\lambda_{0}r)}{r}\|_{L^{2}}\lesssim\|\psi\|_{L^{2}},

and this implies that

|F(ψ2,α0,0)|ψL2δ.|F(\psi_{2},\alpha_{0},0)|\lesssim\|\psi\|_{L^{2}}\lesssim\delta.

Further, FF is uniformly C2C^{2} with respect to α,γ\alpha,\gamma and

Dα,γF|α0,0=4e2iα0(h1,χdαih1h3,ϰdγ)+O(δdγ).D_{\alpha,\gamma}F|_{\alpha_{0},0}=4e^{2i\alpha_{0}}(\langle h_{1},\chi\rangle d\alpha-i\langle h_{1}h_{3},\varkappa\rangle d\gamma)+O(\delta d\gamma).

Assuming that both h1,ϰ0\langle h_{1},\varkappa\rangle\neq 0 and h1h3,ϰ0\langle h_{1}h_{3},\varkappa\rangle\neq 0 and that δ\delta is small enough, by the inverse function theorem the function FF has a unique zero in a small neighbourhood of (α0,0)(\alpha_{0},0), so that

|αα0|+|σ|ψL2δ.|\alpha-\alpha_{0}|+|\sigma|\lesssim\|\psi\|_{L^{2}}\lesssim\delta.

This in turn implies both the pointwise smallness condition

ψ22ie2iαh1(λr)LψL2δ,\|\psi_{2}-2ie^{2i\alpha}h_{1}(\lambda r)\|_{L^{\infty}}\lesssim\|\psi\|_{L^{2}}\lesssim\delta,

and the bound (4.34) as a consequence of the similar bound for (α0,λ0)(\alpha_{0},\lambda_{0}).

To obtain the slightly more general uniqueness result in the proposition, we observe that if we had another solution (α1,λ1)(\alpha_{1},\lambda_{1}) which also satisfies the second bound in (4.31) then we must have

e2iαh1λe2iα1h1λ1Lδ,\|e^{2i\alpha}h_{1}^{\lambda}-e^{2i\alpha_{1}}h_{1}^{\lambda_{1}}\|_{L^{\infty}}\lesssim\delta,

which in turn implies that

|αα1|+|lnλlnλ1|δ,|\alpha-\alpha_{1}|+|\ln\lambda-\ln\lambda_{1}|\lesssim\delta,

which places (α1,λ1)(\alpha_{1},\lambda_{1}) within the range of applicability of the previous uniqueness statement.


Next we consider the dependence of (α,γ)(\alpha,\gamma) with respect to ψ2\psi_{2}. For this we note the following dependence of FF with respect to ψ2\psi_{2},

|F(ψ2,α,γ)F(ψ~2,α,γ)|ψ2ψ~2H˙e1,|F(\psi_{2},\alpha,\gamma)-F(\tilde{\psi}_{2},\alpha,\gamma)|\lesssim\|\psi_{2}-\tilde{\psi}_{2}\|_{\dot{H}^{1}_{e}},

This implies that we have Lipschitz dependence of the parameters α\alpha and γ\gamma with respect to the H˙e1\dot{H}^{1}_{e} norm of ψ2\psi_{2}, just as claimed in (4.33). In addition we can establish a similar Lipschitz dependence with respect to rougher norms of ψ2\psi_{2} - this is possible because we test with a smooth χ\chi. A straightforward computation shows that

|F(ψ2,α,γ)F(ψ~2,α,γ)|λψ2ψ~2L2,|F(\psi_{2},\alpha,\gamma)-F(\tilde{\psi}_{2},\alpha,\gamma)|\lesssim\lambda\|\psi_{2}-\tilde{\psi}_{2}\|_{L^{2}},

and this establishes the second part of the claim in (4.33).

Finally the last part of (4.33) follows from the straightforward inequality

|F(ψ2,α,γ)F(ψ~2,α,γ)|λ2ψ2ψ~2H˙e1.|F(\psi_{2},\alpha,\gamma)-F(\tilde{\psi}_{2},\alpha,\gamma)|\lesssim\lambda^{2}\|\psi_{2}-\tilde{\psi}_{2}\|_{\dot{H}^{-1}_{e}}.

ii) Lemma 4.7 above provides the desired estimates when the choice of parameters is made with the point-type condition versus the smooth one that we work with. Thus our goal here is to prove that the correction coming from the change in parameters can still be controlled in the same way. By definition,

F(ψ2,α0,λ0)=λ02ψ22ie2iα0h1(λ0r),ϰ(λ0r).F(\psi_{2},\alpha_{0},\lambda_{0})=\lambda_{0}^{2}\langle\psi_{2}-2ie^{2i\alpha_{0}}h_{1}(\lambda_{0}r),\varkappa(\lambda_{0}r)\rangle.

Here (ψ22ie2iα0h1λ0)(λ01)=0(\psi_{2}-2ie^{2i\alpha_{0}}h_{1}^{\lambda_{0}})(\lambda_{0}^{-1})=0 while the inner product depends only on ψ2\psi_{2} in the region {rλ01}\{r\approx\lambda_{0}^{-1}\}. To determine ψ2\psi_{2} in this region from ψ\psi via the ode (4.26), it suffices to know ψ\psi in the same region. By a local ode stability analysis it follows that

ψ22ie2iα0h1(λ0r)L(rλ01)ψL2(rλ1).\|\psi_{2}-2ie^{2i\alpha_{0}}h_{1}(\lambda_{0}r)\|_{L^{\infty}(r\approx\lambda_{0}^{-1})}\lesssim\|\psi\|_{L^{2}(r\approx\lambda^{-1})}.

This implies that

|F(ψ2,α0,λ0)|ψL2(rλ1).|F(\psi_{2},\alpha_{0},\lambda_{0})|\lesssim\|\psi\|_{L^{2}(r\approx\lambda^{-1})}.

from which the inverse function theorem used above gives the improved estimate

|αα0|+|lnλlnλ0|ψL2(rλ1).|\alpha-\alpha_{0}|+|\ln\lambda-\ln\lambda_{0}|\lesssim\|\psi\|_{L^{2}(r\approx\lambda^{-1})}.

From this last bound it follows that

ψα,λ,2ψα0,λ0,2rL2+Aα,λ,2Aα0,λ0,2rL2ψL2,\|\frac{\psi_{\alpha,\lambda,2}-\psi_{\alpha_{0},\lambda_{0},2}}{r}\|_{L^{2}}+\|\frac{A_{\alpha,\lambda,2}-A_{\alpha_{0},\lambda_{0},2}}{r}\|_{L^{2}}\lesssim\|\psi\|_{L^{2}},
ψα,λ,2ψα0,λ0,2rL4+Aα,λ,2Aα0,λ0,2rL4ψL4\|\frac{\psi_{\alpha,\lambda,2}-\psi_{\alpha_{0},\lambda_{0},2}}{r}\|_{L^{4}}+\|\frac{A_{\alpha,\lambda,2}-A_{\alpha_{0},\lambda_{0},2}}{r}\|_{L^{4}}\lesssim\|\psi\|_{L^{4}}

and

ψα,λ,2ψα0,λ0,2r2L2+Aα,λ,2Aα0,λ0,2r2L2ψrL2.\|\frac{\psi_{\alpha,\lambda,2}-\psi_{\alpha_{0},\lambda_{0},2}}{r^{2}}\|_{L^{2}}+\|\frac{A_{\alpha,\lambda,2}-A_{\alpha_{0},\lambda_{0},2}}{r^{2}}\|_{L^{2}}\lesssim\|\frac{\psi}{r}\|_{L^{2}}.

Combined with (4.40) and (4.41), this finishes the proof of the two claims (4.36) and (4.37).

The first inequality combined with (4.39) gives

δλ,αψ2rL2+δλA2rL2ψL2.\|\frac{\delta^{\lambda,\alpha}\psi_{2}}{r}\|_{L^{2}}+\|\frac{\delta^{\lambda}A_{2}}{r}\|_{L^{2}}\lesssim\|\psi\|_{L^{2}}.

Then we use this estimate and the system (4.42) to conclude that

rδλ,αψ2L2+rδλA2L2ψL2;\|\partial_{r}\delta^{\lambda,\alpha}\psi_{2}\|_{L^{2}}+\|\partial_{r}\delta^{\lambda}A_{2}\|_{L^{2}}\lesssim\|\psi\|_{L^{2}};

this concludes the proof of (4.34).


The proof of Proposition 4.5 is concluded once we establish (4.35). For this purpose we need the following result.

Lemma 4.8.

Consider the system of ODE

(4.47) rZ=NZ+F,limrZ(r)=0.\partial_{r}Z=NZ+F,\qquad\lim_{r\rightarrow\infty}Z(r)=0.

If N,FN,F are in in rH˙e1+l1L2\partial_{r}\dot{H}^{1}_{e}+l^{1}L^{2}, with NrH˙e1+l1L2\|N\|_{\partial_{r}\dot{H}^{1}_{e}+l^{1}L^{2}} small, then the above equation has a unique solution ZH˙e1+[r]1l1L2Z\in\dot{H}^{1}_{e}+[\partial_{r}]^{-1}l^{1}L^{2} satisfying

(4.48) ZH˙e1+[r]1l1L2FrH˙e1+l1L2.\|Z\|_{\dot{H}^{1}_{e}+[\partial_{r}]^{-1}l^{1}L^{2}}\lesssim\|F\|_{\partial_{r}\dot{H}^{1}_{e}+l^{1}L^{2}}.

Furthermore, the map from N,FrH˙e1+l1L2N,F\in\partial_{r}\dot{H}^{1}_{e}+l^{1}L^{2} to ZH˙e1+[r]1l1L2Z\in\dot{H}^{1}_{e}+[\partial_{r}]^{-1}l^{1}L^{2} is analytic.

Some remarks are in order here:

  • We did not specify the size of our system, but we will be mainly interested in the following two cases: i) Z,N,FZ,N,F are 3×33\times 3 matrices and ii) Z,FZ,F are 3×13\times 1, NN is 3×33\times 3.

  • A similar result holds true for systems of type rZ=ZN+F\partial_{r}Z=ZN+F.

Proof.

We claim the following basic inequality:

(4.49) fgrH˙e1+l1L2frH˙e1+l1L2gH˙e1+[r]1l1L2\|f\cdot g\|_{\partial_{r}\dot{H}^{1}_{e}+l^{1}L^{2}}\lesssim\|f\|_{\partial_{r}\dot{H}^{1}_{e}+l^{1}L^{2}}\|g\|_{\dot{H}^{1}_{e}+[\partial_{r}]^{-1}l^{1}L^{2}}

We write f=rf1+f2,f1H˙e1,f2l1L2f=\partial_{r}f_{1}+f_{2},f_{1}\in\dot{H}^{1}_{e},f_{2}\in l^{1}L^{2} and g=g1+g2,g1H˙e1,rg2l1L2g=g_{1}+g_{2},g_{1}\in\dot{H}^{1}_{e},\partial_{r}g_{2}\in l^{1}L^{2} with g2()=0g_{2}(\infty)=0. Then from the simple bound

(4.50) hl2LhH˙e1,\|h\|_{l^{2}L^{\infty}}\lesssim\|h\|_{\dot{H}^{1}_{e}},

it follows that rf1g1l1L2f1H˙e1g1H˙e1\|\partial_{r}f_{1}\cdot g_{1}\|_{l^{1}L^{2}}\lesssim\|f_{1}\|_{\dot{H}^{1}_{e}}\|g_{1}\|_{\dot{H}^{1}_{e}}. Next we write rf1g2=r(f1g2)f1rg2\partial_{r}f_{1}\cdot g_{2}=\partial_{r}(f_{1}g_{2})-f_{1}\partial_{r}g_{2}; f1rg2l1L2f_{1}\partial_{r}g_{2}\in l^{1}L^{2} is obvious since f1Lf1H˙e1\|f_{1}\|_{L^{\infty}}\lesssim\|f_{1}\|_{\dot{H}^{1}_{e}}. We claim that f1g2H˙e1f_{1}g_{2}\in\dot{H}^{1}_{e}; indeed this follows from the representation (2.11) applied for g2g_{2}, the algebra property of H˙e1\dot{H}^{1}_{e} and the fact that H˙e1\dot{H}^{1}_{e} is stable under multiplication by χAm\chi_{A_{\leq m}}. We also have the trivial inequality f2g1l1L2f2l1L2g1Lf2l1L2g1H˙e1\|f_{2}g_{1}\|_{l^{1}L^{2}}\lesssim\|f_{2}\|_{l^{1}L^{2}}\|g_{1}\|_{L^{\infty}}\lesssim\|f_{2}\|_{l^{1}L^{2}}\|g_{1}\|_{\dot{H}^{1}_{e}}. Finally f2g2l1L2f_{2}g_{2}\in l^{1}L^{2} given the bound (2.10) that places g2g_{2} in LL^{\infty}.

Back to our problem, the solution ZZ is obtained via a Picard iteration in the space H˙e1+[r]1l1L2\dot{H}^{1}_{e}+[\partial_{r}]^{-1}l^{1}L^{2} as follows: from (4.49) we obtain

NZ+FrH˙e1+l1L2NrH˙e1+l1L2ZH˙e1+[r]1l1L2+FrH˙e1+l1L2,\|NZ+F\|_{\partial_{r}\dot{H}^{1}_{e}+l^{1}L^{2}}\lesssim\|N\|_{\partial_{r}\dot{H}^{1}_{e}+l^{1}L^{2}}\|Z\|_{\dot{H}^{1}_{e}+[\partial_{r}]^{-1}l^{1}L^{2}}+\|F\|_{\partial_{r}\dot{H}^{1}_{e}+l^{1}L^{2}},

and [r]1[\partial_{r}]^{-1} (the operator that gives the solution to the linear inhomogeneous ODE) is bounded from rH˙e1+l1L2\partial_{r}\dot{H}^{1}_{e}+l^{1}L^{2} to H˙e1+[r]1l1L2\dot{H}^{1}_{e}+[\partial_{r}]^{-1}l^{1}L^{2}; finally the convergence of the iterations is insured by the smallness of NrH˙e1+l1L2\|N\|_{\partial_{r}\dot{H}^{1}_{e}+l^{1}L^{2}}. ∎

We return to the proof of (4.35). We use the system (4.2), which we recast in a matrix form as an equation for 𝒪=(v¯,w¯,u¯)\mathcal{O}=(\bar{v},\bar{w},\bar{u}) as follows

(4.51) r𝒪=𝒪R(ψ),𝒪()=I3.\partial_{r}\mathcal{O}=\mathcal{O}R(\psi),\quad\mathcal{O}(\infty)=I_{3}.

with

R(ψ)=(00ψ100ψ1ψ1ψ10)R(\psi)=\left(\begin{array}[]{lll}0&0&\Re\psi_{1}\\ 0&0&\Im\psi_{1}\\ -\Re\psi_{1}&-\Im\psi_{1}&0\end{array}\right)

If ψ=0\psi=0 then ψ2=2ie2iαh1λ\psi_{2}=2ie^{2i\alpha}h_{1}^{\lambda}, which yields ψ1=2e2iαh1λr\psi_{1}=-2e^{2i\alpha}\frac{h_{1}^{\lambda}}{r}, hence

(4.52) R(0)=2h1λr(00cos2α00sin2αcos2αsin2α0).R(0)=2\frac{h_{1}^{\lambda}}{r}\left(\begin{array}[]{lll}0&0&-\cos{2\alpha}\\ 0&0&-\sin{2\alpha}\\ \cos{2\alpha}&\sin{2\alpha}&0\end{array}\right).

The solution is given by (see the generalization of (4.12))

𝒪0=(h3λcos22α+sin22α(h3λ1)sin2αcos2αh1λcos2α(h3λ1)sin2αcos2αh3λsin22α+cos22αh1λsin2αh1λcos2αh1λsin2αh3λ).\mathcal{O}_{0}=\!\left(\!\!\!\begin{array}[]{ccc}\!h_{3}^{\lambda}\cos^{2}2\alpha+\sin^{2}2\alpha&\!(h_{3}^{\lambda}-1)\sin 2\alpha\cos 2\alpha&h_{1}^{\lambda}\cos 2\alpha\cr\!(h_{3}^{\lambda}-1)\sin 2\alpha\cos 2\alpha\!&\!h_{3}^{\lambda}\sin^{2}2\alpha+\cos^{2}2\alpha&h_{1}^{\lambda}\sin 2\alpha\cr\!-h_{1}^{\lambda}\cos 2\alpha&\!-h_{1}^{\lambda}\sin 2\alpha&h_{3}^{\lambda}\end{array}\!\!\!\right).\!

We note that 𝒪01=𝒪0t\mathcal{O}_{0}^{-1}=\mathcal{O}_{0}^{t}. We will prove that

(4.53) R(ψ)R(0)rH˙e1+l1L2ψL2.\|R(\psi)-R(0)\|_{\partial_{r}\dot{H}^{1}_{e}+l^{1}L^{2}}\lesssim\|\psi\|_{L^{2}}.

Suppose this is done. Then we write the solution to (4.51) is of the form

(4.54) 𝒪(r)=(I+Y(r))𝒪0(r)\mathcal{O}(r)=(I+Y(r))\mathcal{O}_{0}(r)

where YY solves the differential equation

(4.55) rY=YN+G,Y()=0N=G=𝒪0(R(ψ)R(0))𝒪01.\partial_{r}Y=YN+G,\qquad Y(\infty)=0\qquad N=G=\mathcal{O}_{0}(R(\psi)-R(0))\mathcal{O}_{0}^{-1}.

We apply Lemma 4.8 to solve this system and conclude that

u¯Q¯α,λ2H˙e1+[r]1l1L2+v¯V¯α,λ2H˙e1+[r]1l1L2+w¯W¯α,λ2H˙e1+[r]1l1L2ψL2.\|\bar{u}-\bar{Q}^{2}_{\alpha,\lambda}\|_{\dot{H}^{1}_{e}+[\partial_{r}]^{-1}l^{1}L^{2}}+\|\bar{v}-\bar{V}^{2}_{\alpha,\lambda}\|_{\dot{H}^{1}_{e}+[\partial_{r}]^{-1}l^{1}L^{2}}+\|\bar{w}-\bar{W}^{2}_{\alpha,\lambda}\|_{\dot{H}^{1}_{e}+[\partial_{r}]^{-1}l^{1}L^{2}}\lesssim\|\psi\|_{L^{2}}.

To finish our claim we need an additional bound for r1(u¯Q¯α,λ)L2\|r^{-1}(\bar{u}-\bar{Q}_{\alpha,\lambda})\|_{L^{2}}. We first remark that the last row of 𝒪\mathcal{O} is a-priori known, namely (v¯3,w¯3,u¯3)=12(ψ2,ψ2,A2)(\bar{v}_{3},\bar{w}_{3},\bar{u}_{3})=\frac{1}{2}(-\Im\psi_{2},\Re\psi_{2},A_{2}); this already shows that

r1(v¯3+h1λcos2α)L2+r1(w¯3+h1λsin2α)L2+r1(u¯3h3λ)L2ψL2.\|r^{-1}(\bar{v}_{3}+h_{1}^{\lambda}\cos{2\alpha})\|_{L^{2}}+\|r^{-1}(\bar{w}_{3}+h_{1}^{\lambda}\sin{2\alpha})\|_{L^{2}}+\|r^{-1}(\bar{u}_{3}-h_{3}^{\lambda})\|_{L^{2}}\lesssim\|\psi\|_{L^{2}}.

To transfer this information to u¯1\bar{u}_{1} and u¯2\bar{u}_{2} we use again the orthogonality of 𝒪\mathcal{O}. To keep the computations below compact we assume that α=0\alpha=0; this does not restrict the generality of the argument. For u¯1\bar{u}_{1} we have

u¯1=v¯2w¯3v¯3w¯2=v¯2w¯3(v¯3+h1λ)w¯2+h1λw¯2,\bar{u}_{1}=\bar{v}_{2}\bar{w}_{3}-\bar{v}_{3}\bar{w}_{2}=\bar{v}_{2}\bar{w}_{3}-(\bar{v}_{3}+h_{1}^{\lambda})\bar{w}_{2}+h_{1}^{\lambda}\bar{w}_{2},

from which it follows that

u¯1h1λr=v¯2w¯3rv¯3+h1λrw¯2+h1λr(w¯21);\frac{\bar{u}_{1}-h_{1}^{\lambda}}{r}=\bar{v}_{2}\frac{\bar{w}_{3}}{r}-\frac{\bar{v}_{3}+h_{1}^{\lambda}}{r}\bar{w}_{2}+\frac{h_{1}^{\lambda}}{r}(\bar{w}_{2}-1);

From the above estimates it follows that u¯1h1λrL2ψL2\|\frac{\bar{u}_{1}-h_{1}^{\lambda}}{r}\|_{L^{2}}\lesssim\|\psi\|_{L^{2}}. A similar argument shows that u¯2rL2ψL2\|\frac{\bar{u}_{2}}{r}\|_{L^{2}}\lesssim\|\psi\|_{L^{2}}. This concludes the proof of (4.35).

It remains to prove the bound (4.53). Using the second relation in (4.26) we have

ψ1=ψ+iψ2r=14(iA2rψ2+|ψ2|2ψ+iψ2r|ψ2|2)=2e2iαh1riA2rδλ,αψ2+δλA22e2iαrh1λ+|ψ2|2ψ+14r(iψ2|ψ2|2+8e2iα(h1λ)3)\begin{split}\psi_{1}=&\psi+i\frac{\psi_{2}}{r}=\frac{1}{4}(-iA_{2}\partial_{r}\psi_{2}+|\psi_{2}|^{2}\psi+i\frac{\psi_{2}}{r}|\psi_{2}|^{2})\\ =&-2e^{2i\alpha}\frac{h_{1}}{r}-iA_{2}\partial_{r}\delta^{\lambda,\alpha}\psi_{2}+\delta^{\lambda}A_{2}\cdot 2e^{2i\alpha}\partial_{r}h_{1}^{\lambda}+|\psi_{2}|^{2}\psi+\frac{1}{4r}(i\psi_{2}|\psi_{2}|^{2}+8e^{2i\alpha}(h_{1}^{\lambda})^{3})\end{split}

From this we obtain:

(4.56) ψ1+2e2iαh1r=2ih3λrδλ,αψ2+B,\psi_{1}+2e^{2i\alpha}\frac{h_{1}}{r}=-2ih_{3}^{\lambda}\partial_{r}\delta^{\lambda,\alpha}\psi_{2}+B,

where

B=iδλA2rδλ,αψ2+δλA22e2iαrh1λ+|ψ2|2ψ+14r(iψ2|ψ2|2+8e2iα(h1λ)3).B=-i\delta^{\lambda}A_{2}\partial_{r}\delta^{\lambda,\alpha}\psi_{2}+\delta^{\lambda}A_{2}\cdot 2e^{2i\alpha}\partial_{r}h_{1}^{\lambda}+|\psi_{2}|^{2}\psi+\frac{1}{4r}(i\psi_{2}|\psi_{2}|^{2}+8e^{2i\alpha}(h_{1}^{\lambda})^{3}).

Using (4.34), (4.50) and the simple observation that h1λl2L1\|h_{1}^{\lambda}\|_{l^{2}L^{\infty}}\lesssim 1, it is a straightforward exercise to check that

Bl1L2ψL2.\|B\|_{l^{1}L^{2}}\lesssim\|\psi\|_{L^{2}}.

Finally, we write

h3λrδλ,αψ2=r(h3λδλ,αψ2)+rh3λδλ,αψ2,h_{3}^{\lambda}\partial_{r}\delta^{\lambda,\alpha}\psi_{2}=\partial_{r}(h_{3}^{\lambda}\delta^{\lambda,\alpha}\psi_{2})+\partial_{r}h_{3}^{\lambda}\cdot\delta^{\lambda,\alpha}\psi_{2},

and note that h3λδλ,αψ2H˙e1δλ,αψ2H˙e1ψL2\|h_{3}^{\lambda}\delta^{\lambda,\alpha}\psi_{2}\|_{\dot{H}^{1}_{e}}\lesssim\|\delta^{\lambda,\alpha}\psi_{2}\|_{\dot{H}^{1}_{e}}\lesssim\|\psi\|_{L^{2}}, while

rh3λδλ,αψ2l1L2δλ,αψ2rL2ψL2\|\partial_{r}h_{3}^{\lambda}\cdot\delta^{\lambda,\alpha}\psi_{2}\|_{l^{1}L^{2}}\lesssim\|\frac{\delta^{\lambda,\alpha}\psi_{2}}{r}\|_{L^{2}}\lesssim\|\psi\|_{L^{2}}

(just as one does for the terms in BB). This concludes the proof of (4.53), and, in turn, the proof of our Proposition.

Proof of Proposition 4.6.

The argument is similar to the proof of Proposition 4.5. To emphasize this similarity and avoid cluttering the notations we redenote (α,λ)(\alpha,\lambda) in the proposition by (α0,λ0)(\alpha_{0},\lambda_{0}).

We fix some r0λ01r_{0}\approx\lambda_{0}^{-1}. Then for (α,λ=ϵγλ0)(\alpha,\lambda=\epsilon^{\gamma}\lambda_{0}) close to (α0,λ0)(\alpha_{0},\lambda_{0}) we use Lemma 4.7 to solve the system (4.26) with initial data as in (4.38). Denoting

G(α,γ)=δλ0,α0ψ2,ϰλ0,G(\alpha,\gamma)=\langle\delta^{\lambda_{0},\alpha_{0}}\psi_{2},\varkappa^{\lambda_{0}}\rangle,

we look for (α,γ)(\alpha,\gamma) so that G(α,γ)=0G(\alpha,\gamma)=0. Here GG is uniformly smooth, and by (4.34) we have

|G(α0,λ0)|ψL2|G(\alpha_{0},\lambda_{0})|\lesssim\|\psi\|_{L^{2}}

Further, a direct computation shows that the differential DG(α0,λ0)DG(\alpha_{0},\lambda_{0}) is nondegenerate at (α0,0)(\alpha_{0},0). Then the inverse function theorem yields (α,γ)(\alpha,\gamma) so that G(α,γ)=0G(\alpha,\gamma)=0, and

|αα0|+|γ|ψL2.|\alpha-\alpha_{0}|+|\gamma|\lesssim\|\psi\|_{L^{2}}.

This in turn allows us to switch from (α,λ)(\alpha,\lambda) to (α0,λ0)(\alpha_{0},\lambda_{0}) thus completing the argument for the first part of the Proposition, concerning the recovery of ψ2,A2\psi_{2},A_{2} and the verification of (4.31) and (4.32).

Concerning the second part, which requires the reconstruction of an actual map uu, we note that this has been essentially done in the argument for (4.35). Indeed, there we start with the analysis of the system (4.51) which has ψ1=ψ+iψ2r\psi_{1}=\psi+i\frac{\psi_{2}}{r} as its main input (see R(ψ)R(\psi)), and where we have already recovered ψ2\psi_{2} in the first part of the Proposition. The analysis of the the system (4.51) essentially reconstructs uu (along with its gauge components v,wv,w). At this point we need to establish the following:

  1. (i)

    the gauge elements corresponding to the map uu are indeed ψ1,ψ2,A1=0,A2\psi_{1},\psi_{2},A_{1}=0,A_{2}, and the gauge field is ψ\psi;

  2. (ii)

    the modulation parameters subject to (4.30) are indeed (α,λ)(\alpha,\lambda) from above.

The fact that u¯,v¯,w¯\bar{u},\bar{v},\bar{w} satisfy the system (4.51) shows that (v,w)(v,w) is the correct gauge corresponding to A1=0A_{1}=0; this is better seen in the formulation (4.2) of the system where we take k=1k=1. It also follows that ψ1\psi_{1} is the correct representation of ru\partial_{r}u in the frame (v,w)(v,w).

We need to do the same for the representation of θu\partial_{\theta}u, the angular derivative of uu. At this point we have the pair (A2,ψ2)(A_{2},\psi_{2}) which was constructed from ψ\psi in the first part of the Proposition, and the modulation parameters in the first part of the Proposition, and we have also the pair (A2true,ψ2true)(A_{2}^{true},\psi_{2}^{true}) which are the ”true” gauge components of the map uu in the Coulomb gauge as described in Section 4.2. Our goal is to establish that they are the same, that is (A2,ψ2)=(A2true,ψ2true)(A_{2},\psi_{2})=(A_{2}^{true},\psi_{2}^{true}). if we achieve this then we are done since we have already established that ψ1=ψ1true\psi_{1}=\psi_{1}^{true}; thus ψ=ψ1iψ2r\psi=\psi_{1}-i\frac{\psi_{2}}{r} is the correct gauge field, and the modulation parameters satisfy the correct orthogonality condition (4.30), so they are also the correct modulation parameters for uu.

To establish the equality (A2,ψ2)=(A2true,ψ2true)(A_{2},\psi_{2})=(A_{2}^{true},\psi_{2}^{true}) we note the following:

  1. (i)

    The vector v=(v3,w3,u3)\textbf{v}=(v_{3},w_{3},u_{3}) satisfies the system

    rv=vR(ψ),v()=(0,0,1),\partial_{r}\textbf{v}=\textbf{v}R(\psi),\quad\textbf{v}(\infty)=(0,0,1),

    which is a consequence of (4.51);

  2. (ii)

    The vector v’=12(ψ2,ψ2,A2)\textbf{v'}=\frac{1}{2}(-\Im\psi_{2},\Re\psi_{2},A_{2}) satisfies the same exact system

    rv’=v’R(ψ),v’()=(0,0,1).\partial_{r}\textbf{v'}=\textbf{v'}R(\psi),\quad\textbf{v'}(\infty)=(0,0,1).

Indeed ψ2,A2\psi_{2},A_{2} obey the system (4.42) which, using the compatibility condition (4.27), implies that they obey (4.26) and in turn (4.20); the later system (4.20) implies the claim above that v’ obeys the system, while the condition at \infty follows from the fact that δλA2,δλ,αψ2\delta^{\lambda}A_{2},\delta^{\lambda,\alpha}\psi_{2} are both in H˙e1\dot{H}^{1}_{e}, hence have zero limit at \infty, and thus the limit is given by the limit of 12((2ie2iαh1λ),(2ie2iαh1λ),2h3λ)\frac{1}{2}(-\Im(2ie^{2i\alpha}h_{1}^{\lambda}),\Re(2ie^{2i\alpha}h_{1}^{\lambda}),2h_{3}^{\lambda}) which is precisely (0,0,1)(0,0,1). Just as in the analysis of (4.51), we recast this so that it fits into the framework of Lemma 4.8; precisely, we write the system for v(0,0,1)\textbf{v}-(0,0,1) and v’(0,0,1)\textbf{v'}-(0,0,1), so that we have the zero data at infinity. To invoke the uniqueness part from Lemma 4.8, we need establish that v(0,0,1)\textbf{v}-(0,0,1) and v’(0,0,1)\textbf{v'}-(0,0,1) belong to H˙e1+[r]1l1L2\dot{H}^{1}_{e}+[\partial_{r}]^{-1}l^{1}L^{2}; this is clear for v’(0,0,1)\textbf{v'}-(0,0,1) since this is where we solved the system (4.51). As for v(0,0,1)=12(ψ2,ψ2,A22)\textbf{v}-(0,0,1)=\frac{1}{2}(-\Im\psi_{2},\Re\psi_{2},A_{2}-2), we note that from the above arguments we already know that

δλ,αψ2H˙e1+δλA2H˙e1ψL2.\|\delta^{\lambda,\alpha}\psi_{2}\|_{\dot{H}^{1}_{e}}+\|\delta^{\lambda}A_{2}\|_{\dot{H}^{1}_{e}}\lesssim\|\psi\|_{L^{2}}.

This is enough to place the first two components of v(0,0,1)\textbf{v}-(0,0,1) in H˙e1\dot{H}^{1}_{e}. Next we write A22=δλA2+2(h3λ1)A_{2}-2=\delta^{\lambda}A_{2}+2(h_{3}^{\lambda}-1) and note that δλA2H˙e1\delta^{\lambda}A_{2}\in\dot{H}^{1}_{e}, while h3λ1[r]1l1L2h_{3}^{\lambda}-1\in[\partial_{r}]^{-1}l^{1}L^{2} (this follows from r(h31)=2(h1)2rl1L2\partial_{r}(h_{3}-1)=2\frac{(h_{1})^{2}}{r}\in l^{1}L^{2} when λ=1\lambda=1). Therefore we conclude that v=v’\textbf{v}=\textbf{v'}, which completes the proof of Proposition 4.6.

As we consider solutions for the Schrödinger map equation, the modulation parameters will vary as functions of time. Later we will study this dependence in much greater detail, but for now we are content with proving that they are C1C^{1} functions of time:

Corollary 4.9.

The modulation parameters λ(t)\lambda(t) and α(t)\alpha(t) associated to a map uu based on (4.31) and (4.32) are continuously differentiable on the maximal interval of existence ImaxI_{max} and obey the pointwise bounds

|α|+|λλ|λ2.|\alpha^{\prime}|+|\frac{\lambda^{\prime}}{\lambda}|\lesssim\lambda^{2}.
Proof.

We simply invoke (4.33) and seek an estimate for tψ2\partial_{t}\psi_{2} in appropriate norms. Recall from (4.19) with A1=0A_{1}=0 that

tψ2=iA0ψ2+iΔψ2i1r2A22ψ2(ψ2ψ¯1)ψ1.\partial_{t}\psi_{2}=-iA_{0}\psi_{2}+i\Delta\psi_{2}-i\frac{1}{r^{2}}A_{2}^{2}\psi_{2}-\Im{(\psi_{2}\bar{\psi}_{1})}\psi_{1}.

Based on (4.33), it suffices to obtain a uniform bound in H˙e1\dot{H}^{-1}_{e} for the right-hand side. The main observation here is that we have the uniform in time bound

(4.57) ψ2(t)H˙e11.\|\psi_{2}(t)\|_{\dot{H}^{1}_{e}}\lesssim 1.

Indeed, from the relation |ψ2|2=m2(u12+u22)|\psi_{2}|^{2}=m^{2}(u_{1}^{2}+u_{2}^{2}) and the expression of the energy in (1.4), it follows that ψ2rL21\|\frac{\psi_{2}}{r}\|_{L^{2}}\lesssim 1. From the compatibility condition (4.20), we have rψ2=iA2ψ1\partial_{r}\psi_{2}=iA_{2}\cdot\psi_{1}, thus

rψ2L2ψ1L2ruL2E(u)1.\|\partial_{r}\psi_{2}\|_{L^{2}}\lesssim\|\psi_{1}\|_{L^{2}}\lesssim\|\partial_{r}u\|_{L^{2}}\lesssim E(u)\lesssim 1.

Based on (4.57), we estimate as follows:

|Δψ2,ϕ|rψ2L2rϕL2ϕH˙e1,|\langle\Delta\psi_{2},\phi\rangle|\lesssim\|\partial_{r}\psi_{2}\|_{L^{2}}\cdot\|\partial_{r}\phi\|_{L^{2}}\lesssim\|\phi\|_{\dot{H}^{1}_{e}},
|1r2A2ψ2,ϕ|ψ2rL2ϕrL2ϕH˙e1,|\langle\frac{1}{r^{2}}A_{2}\psi_{2},\phi\rangle|\lesssim\|\frac{\psi_{2}}{r}\|_{L^{2}}\cdot\|\frac{\phi}{r}\|_{L^{2}}\lesssim\|\phi\|_{\dot{H}^{1}_{e}},
|(ψ2ψ¯1)ψ1,ϕ|ψ2Lψ1L22ϕLψ2H˙e1ϕH˙e1ϕH˙e1,|\langle\Im{(\psi_{2}\bar{\psi}_{1})}\psi_{1},\phi\rangle|\lesssim\|\psi_{2}\|_{L^{\infty}}\cdot\|\psi_{1}\|_{L^{2}}^{2}\cdot\|\phi\|_{L^{\infty}}\lesssim\|\psi_{2}\|_{\dot{H}^{1}_{e}}\|\phi\|_{\dot{H}^{1}_{e}}\lesssim\|\phi\|_{\dot{H}^{1}_{e}},
|A0ψ2,ϕ|ψ2LA0L1ϕLψ2H˙e1ϕH˙e1ψL2ψ2rL2ϕH˙e1.|\langle A_{0}\psi_{2},\phi\rangle|\lesssim\|\psi_{2}\|_{L^{\infty}}\cdot\|A_{0}\|_{L^{1}}\cdot\|\phi\|_{L^{\infty}}\lesssim\|\psi_{2}\|_{\dot{H}^{1}_{e}}\|\phi\|_{\dot{H}^{1}_{e}}\|\psi\|_{L^{2}}\|\frac{\psi_{2}}{r}\|_{L^{2}}\lesssim\|\phi\|_{\dot{H}^{1}_{e}}.

The above estimates suffice to conclude the proof.

4.5. The complete setup of the problem

Here we summarize the conclusions of this section. Each 22-equivariant solution to the Schrödinger map equation with homotopy degree 22 and energy just above the ground state energy is uniquely represented by two components, which vary as functions of time:

  1. (1)

    the differentiated field ψL2\psi\in L^{2};

  2. (2)

    the modulation parameters (α,λ)(\alpha,\lambda).

In order to identify the reference soliton associated to a map and thus fix the modulation parameters, we use the orthogonality conditions (4.31) and (4.32), which we recall here for convenience:

(4.58) δλ,αψ2,ϰλ=0,δλ,αψ2Lδ.\langle\delta^{\lambda,\alpha}\psi_{2},\varkappa^{\lambda}\rangle=0,\qquad\|\delta^{\lambda,\alpha}\psi_{2}\|_{L^{\infty}}\lesssim\delta.

Having made this choice, the PDE driving the evolution of ψ\psi, namely the equation (4.25), becomes

(4.59) itψH~λψ=N(ψ),i\partial_{t}\psi-\tilde{H}_{\lambda}\psi=N(\psi),

where we have used the notations

H~λ:=Δ+54h3λr2,N(ψ):=Wλψ,Wλ:=A02δλA2r21r(ψ2ψ¯).\tilde{H}_{\lambda}:=-\Delta+\frac{5-4h_{3}^{\lambda}}{r^{2}},\quad N(\psi):=W_{\lambda}\cdot\psi,\quad W_{\lambda}:=A_{0}-2\frac{\delta^{\lambda}A_{2}}{r^{2}}-\frac{1}{r}\Im{(\psi_{2}\bar{\psi})}.

The initial data of (4.59) satisfies ψ(0)L2δ\|\psi(0)\|_{L^{2}}\lesssim\delta. Our analysis shows that the auxiliary functions (ψ2,A2)(\psi_{2},A_{2}) and then A0A_{0} are uniquely determined by ψ\psi together with (α,λ)(\alpha,\lambda).

It is easy to see that all terms in NN are at least quadratic in ψ\psi, or can be estimated by quantities that are at least quadratic in ψ\psi. We place it in right-hand side of (4.59) in order to emphasize our goal to treat NN as a nonlinear perturbation.

The linear part of (4.59) shows that the following linear PDE

(4.60) (itH~λ)Ψ=0(i\partial_{t}-\tilde{H}_{\lambda})\Psi=0

should play crucial role in our analysis. It is important to note here that λ=λ(t)\lambda=\lambda(t), and thus we are dealing with a variable coefficient linear PDE. In Section 7 we carry a full analysis of (4.60) (including its inhomogeneous counterpart), and in Section 8 we prove an array of estimates fo N(ψ)N(\psi).

The analysis of the time evolution of the modulation parameters (α,λ)(\alpha,\lambda), on the other hand, begins in Section 9, where we use the orthogonality condition (4.58) in order to derive the modulation equations governing the evolution of λ\lambda and α\alpha as a function of time. This analysis is further refined in Section 10.

At this point we are also able to explain the role of the parameter δ\delta involved in this paper. Recall that we consider 22-equivariant maps uu satisfying (1.5) and whose energy satisfies E(u)8π+δ2E(u)\leq 8\pi+\delta^{2}; this translates into the mass constraint ψ(0)Lr2δ\|\psi(0)\|_{L^{2}_{r}}\leq\delta, which is propagated along the flow on the maximal interval of existence. The motivation for the smallness condition δ1\delta\ll 1 comes from two different sources:

i) the fixed time theory developed in this section, more precisely the constraint imposed in Proposition 4.5;

ii) the dynamic (in time) analysis of the PDE (4.59) described above, as well as the dynamics (in time) of the ODE system describing the evolution of the modulation parameters α(t)\alpha(t) and λ(t)\lambda(t) (to be detailed in Section 9).

5. Spectral analysis for the operators HH, H~\tilde{H}

As discussed in the previous section, see in particular (4.60), the linearization of the Schrödinger map problem in its gauge representation near the gauge components of the soliton Qα,λ2Q^{2}_{\alpha,\lambda} reveals that the operator H~λ\tilde{H}_{\lambda} plays an important role in our problem. If we were instead to directly linearize the Schrödinger map problem near the the soliton Qα,λ2Q^{2}_{\alpha,\lambda}, then the corresponding linearization in (4.60) would replace H~λ\tilde{H}_{\lambda} with HλH_{\lambda}, see [14] for details. This second linearization plays no role in our analysis, so apriori there is no reason to study the spectral properties of HλH_{\lambda}. Indeed, our main goal in this section is to develop the spectral theory for H~λ\tilde{H}_{\lambda}. But for technical reasons, it is convenient to do so first for HλH_{\lambda} and then use this to derive the one for H~λ\tilde{H}_{\lambda}; in particular we heavily rely on the analysis developed by Krieger, Schlag and Tataru in Section 4 of [22], where the spectral theory is developed for an operator related to HλH_{\lambda} rather than H~λ\tilde{H}_{\lambda}.

The spectral theory for H~λ\tilde{H}_{\lambda} will play a key role in the study of the dispersive properties of the linear Schrödinger equation (4.60), which in turn will be critical in the study of (4.59), the main PDE governing the evolution of the gauge field ψ\psi.

Let us recall from Section 3.4 the definition of the two operators

Hλ=\displaystyle H_{\lambda}= Δ+Vλ,Vλ(r)=4r2(12(h1λ)2),\displaystyle\ -\Delta+V_{\lambda},\quad V_{\lambda}(r)=\frac{4}{r^{2}}(1-2(h_{1}^{\lambda})^{2}),
H~λ=\displaystyle\tilde{H}_{\lambda}= Δ+V~λ,V~λ(r)=54h3λr2.\displaystyle\ -\Delta+\tilde{V}_{\lambda},\quad\tilde{V}_{\lambda}(r)=\frac{5-4h_{3}^{\lambda}}{r^{2}}.

The operators HλH_{\lambda} and H~λ\tilde{H}_{\lambda} are conjugate operators and admit the factorizations

Hλ=LλLλ,H~λ=LλLλ,H_{\lambda}=L^{*}_{\lambda}L_{\lambda},\qquad\tilde{H}_{\lambda}=L_{\lambda}L^{*}_{\lambda},

where

(5.1) Lλ=h1λr1h1λ=r+2rh3λ,Lλ=1h1λrh1λ1r=r+2h3λ1r.L_{\lambda}=h_{1}^{\lambda}\partial_{r}\frac{1}{h_{1}^{\lambda}}=\partial_{r}+\frac{2}{r}h_{3}^{\lambda},\qquad L^{*}_{\lambda}=-\frac{1}{h_{1}^{\lambda}}\partial_{r}h_{1}^{\lambda}-\frac{1}{r}=-\partial_{r}+\frac{2h_{3}^{\lambda}-1}{r}.

In these expressions, λ\lambda plays the role of a scaling parameter, therefore for the spectral theory it suffices to consider the case λ=1\lambda=1. For convenience we denote the corresponding operators by H=H1,H~=H~1,L=L1,L=L1H=H_{1},\tilde{H}=\tilde{H}_{1},L=L_{1},L^{*}=L^{*}_{1}. Note that we also have the scaling relations:

Lλ(v)(r)=λL(v(λ))(rλ),Lλ(v)(r)=λL(v(λ))(rλ).L_{\lambda}(v)(r)=\lambda L(v(\frac{\cdot}{\lambda}))(r\lambda),\quad L^{*}_{\lambda}(v)(r)=\lambda L^{*}(v(\frac{\cdot}{\lambda}))(r\lambda).

In the first part of the section we develop the spectral theory for the operators HH and H~\tilde{H}, and at the end we use it to derive the spectral theory for the rescaled operators HλH_{\lambda} and H~λ\tilde{H}_{\lambda}, considering in particular the transference operator which describes the λ\lambda dependence in the spectral theory.

We note that the spectral theory for HH in the case m=1m=1 was studied in detail by Krieger-Schlag-Tataru in [21]. Then it was adapted to a setup similar setup to ours (in particular extending the theory to H~\tilde{H}) in [6].

In the current paper we need the spectral theory for HH and H~\tilde{H} in the case m=2m=2. Just as in prior works, see [21, 22], one shows that HH fits into the theory developed by Gesztesy-Zinchenko in [11], which provides the abstract framework for the spectral theory. We first characterize the generalized eigenfunctions for HH; once the spectral theory for HH is complete, we use it together with the conjugation described above in order to derive the one for H~\tilde{H}.

To show that HH fits into the theory developed by Gesztesy-Zinchenko in [11], we proceed as in Section 4 of [22]: if we define the operator \mathcal{L} by conjugating HH with respect to the weight r\sqrt{r}, so that it is selfadjoint in L2(dr)L^{2}(dr),

(5.2) =rHr12=r2+(154r232r2(1+r4)2):=r2+V(r).\mathcal{L}=\sqrt{r}Hr^{-\frac{1}{2}}=-\partial_{r}^{2}+\left(\frac{15}{4r^{2}}-32\frac{r^{2}}{(1+r^{4})^{2}}\right):=-\partial_{r}^{2}+V(r).

Then, by Example 3.10 in [11], the potential VV satisfies hypothesis 3.1 in [11]; the interested reader may find some more details in [22], which studies the spectral theory for the operator KST\mathcal{L}_{KST} defined by

KST:=r2+(154r224(r2+1)2).\mathcal{L}_{KST}:=-\partial_{r}^{2}+\left(\frac{15}{4r^{2}}-\frac{24}{\left(r^{2}+1\right)^{2}}\right).

This provides us with the abstract spectral theory described below.

Precisely, we consider HH acting as an unbounded selfadjoint operator in L2(rdr)L^{2}(rdr). Then HH is nonnegative, and its spectrum [0,)[0,\infty) is absolutely continuous. HH has a zero eigenvalue, namely ϕ0=h1\phi_{0}=h_{1},

Hh1=0,h1(r)=2r2r4+1.Hh_{1}=0,\qquad h_{1}(r)=\frac{2r^{2}}{r^{4}+1}.

For each ξ>0\xi>0 one can choose a normalized generalized eigenfunction ϕξ\phi_{\xi}, smooth at r=0r=0,

(5.3) Hϕξ=ξ2ϕξ.H\phi_{\xi}=\xi^{2}\phi_{\xi}.

These generalized eigenfunctions are unique up to a ξ\xi dependent multiplicative factor, which is chosen as described below. To these one associates a generalized Fourier transform {\mathcal{F}}.

Because of the zero eigenvalue, we regard {\mathcal{F}} as a two-component vector, defined by

(5.4) :L2(rdr)×L2(dξ),(f)(ξ)=[ag(ξ)]{\mathcal{F}}:L^{2}(rdr)\to\mathbb{R}\times L^{2}(d\xi),\qquad{\mathcal{F}}(f)(\xi)=\begin{bmatrix}a\\ g(\xi)\end{bmatrix}

and

a=f,ϕ0L2(rdr)ϕ0L2(rdr)1,a=\langle f,\phi_{0}\rangle_{L^{2}(rdr)}\|\phi_{0}\|^{-1}_{L^{2}(rdr)},
g(ξ)=0ϕξ(r)f(r)r𝑑r,ξ>0,g(\xi)=\int_{0}^{\infty}\phi_{\xi}(r)f(r)rdr,\quad\xi>0,

where the integral above is considered in the singular sense. This is the same notation as in [22]. On the image we will use the norm

[ag]L2,02:=|a|2+0|g(ξ)|2𝑑ξ.\|\begin{bmatrix}a\\ g\end{bmatrix}\|_{L^{2,0}}^{2}:=|a|^{2}+\int_{0}^{\infty}|g(\xi)|^{2}d\xi.

This makes the generalized Fourier transform an L2L^{2} isometry, and we have the inversion formula

(5.5) 1([ag()])(r)=aϕ0(r)ϕ0L2(rdr)+0g(ξ)ϕξ(r)𝑑ξ.{\mathcal{F}}^{-1}(\begin{bmatrix}a\\ g(\cdot)\end{bmatrix})(r)=a\frac{\phi_{0}(r)}{\|\phi_{0}\|_{L^{2}(rdr)}}+\int_{0}^{\infty}g(\xi)\phi_{\xi}(r)d\xi.

Compared to the setup of [22], there are several differences in our setting here, stemming from the different choice of the spectral parameter in (5.3), namely ξ2\xi^{2} rather than ξ\xi, as well as our use of the rdrrdr measure rather than drdr. For quick reference, we provide a brief dictionary.

The counterparts, in our setting, of ϕ(r,ξ)\phi(r,\xi) from Proposition 4.5 of [22] are ϕKST(r,ξ)\phi_{KST}(r,\xi), which we define, for ξ>0\xi>0, so that

(5.6) ρ(ξ)ϕKST(r,ξ)=r2ξ14ϕξ(r)\sqrt{\rho(\xi)}\phi_{KST}(r,\xi)=\frac{\sqrt{r}}{\sqrt{2}\xi^{\frac{1}{4}}}\phi_{\sqrt{\xi}}(r)

where ρ\rho is the density of the continuous part of the spectral measure of \mathcal{L}, which exactly corresponds to the function ρ\rho from Theorem 4.3 of [22]. For the zero mode, we have

ϕKST(r,0)=r2ϕ0(r).\phi_{KST}(r,0)=\frac{\sqrt{r}}{2}\phi_{0}(r).

The counterpart of the function ψ+\psi^{+} from Theorem 4.3 of [22] is

ψKST+(r,ξ):=rξ1/4ϕξ+(r).\psi_{KST}^{+}(r,\xi):=\frac{\sqrt{r}}{\xi^{1/4}}\phi_{\sqrt{\xi}}^{+}(r).

where ϕξ+\phi_{\xi}^{+} is defined in (5.48).

The functions ϕξ\phi_{\xi} are smooth with respect to both rr and ξ\xi. To describe them one considers two distinct regions, rξ1r\xi\lesssim 1 and rξ1r\xi\gtrsim 1. The main properties of ϕξ\phi_{\xi} are summarized in the next theorem.

Theorem 5.1.

a) In the inner region {rξ1}\{r\xi\lesssim 1\} the functions ϕξ\phi_{\xi} admit a power series expansion of the form

(5.7) ϕξ(r)=q(ξ)(ϕ0(r)+j=1(rξ)2jϕj(r2))\phi_{\xi}(r)=q(\xi)\left(\phi_{0}(r)+\sum_{j=1}^{\infty}(r\xi)^{2j}\phi_{j}(r^{2})\right)

where ϕ0=h1\phi_{0}=h_{1} and the functions ϕj\phi_{j} are analytic and satisfy

(5.8) |ϕj(u)|C1((j1)!)2(323/2)j1uu,u>0,j1,|\phi_{j}(u)|\leq\frac{C_{1}}{((j-1)!)^{2}}\left(\frac{3}{2^{3/2}}\right)^{j-1}\frac{u}{\langle u\rangle},\quad u>0,\ j\geq 1,

as well as111 Here ϕj\phi_{j} are also analytic at 0; we only wrote the bounds in this manner in order to gain uniformity in jj.

(5.9) |ϕj(n)(u)|Mnj2n1((j1)!)2(323/2)j1uunu,u>0,j1,n1,|\phi_{j}^{(n)}(u)|\leq\frac{M_{n}j^{2n-1}}{((j-1)!)^{2}}\left(\frac{3}{2^{3/2}}\right)^{j-1}\frac{u}{u^{n}\langle u\rangle},\quad u>0,\ j\geq 1,\ n\geq 1,

for some Mn>0M_{n}>0. The smooth positive weight qq satisfies

(5.10) q(ξ){ξ12,ξ1ξ52,ξ1,|(ξξ)αq|αq.q(\xi)\approx\left\{\begin{array}[]{ll}\displaystyle\xi^{\frac{1}{2}},&\xi\ll 1\cr\cr\xi^{\frac{5}{2}},&\xi\gg 1\end{array}\right.,\qquad|(\xi\partial_{\xi})^{\alpha}q|\lesssim_{\alpha}q.

In addition, if

(5.11) ωk1(r)={min{1,22kr2,22kr4},k<0min{1,24kr4},k0\omega_{k}^{1}(r)=\left\{\begin{array}[]{ll}\min\{1,2^{2k}r^{2},2^{2k}r^{4}\},&\ k<0\\ \\ \min\{1,2^{4k}r^{4}\},&\ k\geq 0\end{array}\right.

then, we have

(5.12) |(ξξ)α(rr)β(ϕξ(r)q(ξ)ϕ0(r))|αβ2k2ωk1(r),ξ2k,rξ1.|(\xi\partial_{\xi})^{\alpha}(r\partial_{r})^{\beta}\left(\phi_{\xi}(r)-q(\xi)\phi_{0}(r)\right)|\lesssim_{\alpha\beta}2^{\frac{k}{2}}\omega_{k}^{1}(r),\qquad\xi\approx 2^{k},\ r\xi\lesssim 1.

b) In the outer region, {rξ1}\{r\xi\gtrsim 1\}, we have the representation

(5.13) ϕξ(r)=a(ξ)ϕξ+(r)+a(ξ)ϕξ+(r)¯\phi_{\xi}(r)=a(\xi)\phi^{+}_{\xi}(r)+\overline{a(\xi)\phi^{+}_{\xi}(r)}

where the complex valued weight aa satisfies

(5.14) |a(ξ)|=12π,|(ξξ)αa(ξ)|α1|a(\xi)|=\frac{1}{\sqrt{2\pi}},\qquad|(\xi\partial_{\xi})^{\alpha}a(\xi)|\lesssim_{\alpha}1

and

(5.15) ϕξ+(r)=r12eirξσ(rξ,r),rξ1.\phi^{+}_{\xi}(r)=r^{-\frac{1}{2}}e^{ir\xi}\sigma(r\xi,r),\qquad r\xi\gtrsim 1.

Moreover, σ\sigma above admits the asymptotic expansion

(5.16) σ(y,r)j=0yjϕj+(r),ϕ0+=1\sigma(y,r)\approx\sum_{j=0}^{\infty}y^{-j}\phi^{+}_{j}(r),\qquad\phi_{0}^{+}=1

with smooth functions ϕj+\phi_{j}^{+} satisfying symbol type bounds

supr>0|(rr)kϕj+|<\sup_{r>0}|(r\partial_{r})^{k}\phi^{+}_{j}|<\infty

and ϕ1+\phi_{1}^{+} with asymptotic behavior

(5.17) ϕ1+(r)=15i816i5r4+O(1r8),r.\begin{split}\phi_{1}^{+}(r)=\frac{15i}{8}-\frac{16i}{5r^{4}}+O\left(\frac{1}{r^{8}}\right),\quad r\rightarrow\infty.\end{split}

Here the expansion (5.16) holds in in the sense that for all n,m0n,m\geq 0, there exists Nn,m>0N_{n,m}>0 such that, for all NNn,mN\geq N_{n,m}, there exists CN,n,mC_{N,n,m} such that

(5.18) |(rr)n(yy)m(σ(y,r)k=0Nykϕk+(r))|CN,n,myN+1,y>1,r>0.\begin{split}|(r\partial_{r})^{n}(y\partial_{y})^{m}\left(\sigma(y,r)-\sum_{k=0}^{N}y^{-k}\phi_{k}^{+}(r)\right)|\leq\frac{C_{N,n,m}}{y^{N+1}},\quad y>1,r>0.\end{split}

We remark that the theorem in particular implies that

(5.19) |ϕξ(r)q(ξ)ϕ0(r)|Cq(ξ)r2ξ2r2r2,rξ1,|\phi_{\xi}(r)-q(\xi)\phi_{0}(r)|\leq Cq(\xi)r^{2}\xi^{2}\frac{r^{2}}{\langle r^{2}\rangle},\quad r\xi\lesssim 1,
(5.20) |ϕξ(r)|Cr,1rξ.|\phi_{\xi}(r)|\leq\frac{C}{\sqrt{r}},\quad 1\lesssim r\xi.

The leading role in this paper the spectral theory for the operator H~\tilde{H}. This is derived from the spectral theory for HH due to the conjugate representations

H=LL,H~=LL.H=L^{*}L,\qquad\tilde{H}=LL^{*}.

This allows us to define generalized eigenfunctions ψξ\psi_{\xi} for H~\tilde{H} using the generalized eigenfunctions ϕξ\phi_{\xi} for HH,

(5.21) ψξ=ξ1Lϕξ,Lψξ=ξϕξ.\psi_{\xi}=\xi^{-1}L\phi_{\xi},\qquad L^{*}\psi_{\xi}=\xi\phi_{\xi}.

It is easy to see that ψξ\psi_{\xi} are real, smooth, vanish at r=0r=0 and solve

H~ψξ=ξ2ψξ.\tilde{H}\psi_{\xi}=\xi^{2}\psi_{\xi}.

With respect to this frame we can define the generalized Fourier transform adapted to H~\tilde{H} by

~f(ξ)=0ψξ(r)f(r)r𝑑r{\tilde{\mathcal{F}}}{f}(\xi)=\int_{0}^{\infty}\psi_{\xi}(r)f(r)rdr

where the integral above is again considered in the singular sense. This is again an L2L^{2} isometry, and we have the inversion formula

(5.22) f(r)=0ψξ(r)~(f)(ξ)𝑑ξ.f(r)=\int_{0}^{\infty}\psi_{\xi}(r){\tilde{\mathcal{F}}}(f)(\xi)d\xi.

To see this we compute, for a Schwartz function ff:

(5.23) ~(Lf)(ξ)=0ψξ(r)Lf(r)r𝑑r=0Lψξ(r)f(r)r𝑑r=0ξϕξ(r)f(r)r𝑑r=ξ(f)(ξ).\begin{split}{\tilde{\mathcal{F}}}(Lf)(\xi)&=\!\int_{0}^{\infty}\psi_{\xi}(r)Lf(r)rdr=\!\int_{0}^{\infty}L^{*}\psi_{\xi}(r)f(r)rdr\\ &=\!\int_{0}^{\infty}\xi\phi_{\xi}(r)f(r)rdr=\xi{\mathcal{F}}(f)(\xi).\end{split}

Hence

~(Lf)L22=ξf(ξ)L22=Hf,fL2(rdr)=LfL22\|{\tilde{\mathcal{F}}}(Lf)\|_{L^{2}}^{2}=\|\xi{\mathcal{F}}{f}(\xi)\|_{L^{2}}^{2}=\langle Hf,f\rangle_{L^{2}(rdr)}=\|Lf\|_{L^{2}}^{2}

which suffices since LfLf spans a dense subset of L2L^{2}.

The representation of ψξ\psi_{\xi} in the two regions rξ1r\xi\lesssim 1 and rξ1r\xi\gtrsim 1 is obtained from the similar representation of ϕξ\phi_{\xi}. The main properties of ψξ\psi_{\xi} which we use are summarized below.

Theorem 5.2.

a) In the region rξ1r\xi\lesssim 1 the functions ψξ\psi_{\xi} admit a power series expansion of the form

ψξ=ξq(ξ)(ψ0(r)+rj1(rξ)2jψj(r2)),\psi_{\xi}=\xi q(\xi)\left(\psi_{0}(r)+r\sum_{j\geq 1}(r\xi)^{2j}{\psi}_{j}(r^{2})\right),

where the coefficients ψj\psi_{j} are given by

ψj(r)=(2h3(r)+2+2j)ϕj+1(r)+2rrϕj+1(r),{\psi}_{j}(r)=(2h_{3}(\sqrt{r})+2+2j)\phi_{j+1}(r)+2r\partial_{r}\phi_{j+1}(r),

and satisfy

|(rr)αψj|αCj(j1)!r2r2,|(r\partial_{r})^{\alpha}\psi_{j}|\lesssim_{\alpha}\frac{C^{j}}{(j-1)!}\frac{r^{2}}{\langle r\rangle^{2}},
ψ0(r)=L(r2ϕ1(r2))=r2(1+r4)arctan(r2)2r3.\psi_{0}(r)=L(r^{2}\phi_{1}(r^{2}))=\frac{r^{2}-(1+r^{4})\arctan(r^{2})}{2r^{3}}.

In particular,

ψ0(r)=r33+r715+O(r11),r0\psi_{0}(r)=-\frac{r^{3}}{3}+\frac{r^{7}}{15}+O(r^{11}),\quad r\rightarrow 0

and

ψ0(r)=πr4+1r+O(r3),r.\psi_{0}(r)=-\frac{\pi r}{4}+\frac{1}{r}+O(r^{-3}),\quad r\rightarrow\infty.

In addition, if

(5.24) ωk(r)={min(1,r2r22kr),ifk<0min(1,23kr3),ifk0\omega_{k}(r)=\left\{\begin{array}[]{ll}&\min(1,\frac{r^{2}}{\langle r^{2}\rangle}2^{k}r),\ \ \mbox{if}\ k<0\\ \\ &\min(1,2^{3k}r^{3}),\ \ \mbox{if}\ k\geq 0\end{array}\right.

then, we have

(5.25) |(rr)α(ξξ)βψξ(r)|αβ2k2ωk(r),ξ2k,rξ1.|(r\partial_{r})^{\alpha}(\xi\partial_{\xi})^{\beta}\psi_{\xi}(r)|\lesssim_{\alpha\beta}2^{\frac{k}{2}}\omega_{k}(r),\qquad\xi\approx 2^{k},\ r\xi\lesssim 1.

b) In the region rξ1r\xi\gtrsim 1 we have the representation

(5.26) ψξ(r)=a(ξ)ψξ+(r)+a(ξ)ψξ+(r)¯\psi_{\xi}(r)=a(\xi)\psi^{+}_{\xi}(r)+\overline{a(\xi)\psi^{+}_{\xi}(r)}

where

(5.27) ψξ+(r)=r12eirξσ~(rξ,r),rξ1\psi^{+}_{\xi}(r)=r^{-\frac{1}{2}}e^{ir\xi}\tilde{\sigma}(r\xi,r),\qquad r\xi\gtrsim 1

and σ~\tilde{\sigma} is given by

σ~(y,r)=iσ(y,r)12y1σ(y,r)+yσ(y,r)+ryLσ(y,r)\tilde{\sigma}(y,r)=i\sigma(y,r)-\frac{1}{2}y^{-1}\sigma(y,r)+\frac{\partial}{\partial y}\sigma(y,r)+\frac{r}{y}L\sigma(y,r)

and has exactly the same properties as σ\sigma. In particular, for fixed ξ\xi we have

(5.28) σ~(rξ,r)=i38r1ξ1+O(r2).\tilde{\sigma}(r\xi,r)=i-\frac{3}{8}r^{-1}\xi^{-1}+O(r^{-2}).

The last two theorems describe the spectral theory for the operators HλH_{\lambda} and H~λ\tilde{H}_{\lambda} in the case λ=1\lambda=1. For general λ\lambda, it is easy to see that the following functions

ϕξλ(r)=λ12ϕλ1ξ(λr),ψξλ(r)=λ12ψλ1ξ(λr).\phi_{\xi}^{\lambda}(r)=\lambda^{\frac{1}{2}}\phi_{\lambda^{-1}\xi}(\lambda r),\qquad\psi_{\xi}^{\lambda}(r)=\lambda^{\frac{1}{2}}\psi_{\lambda^{-1}\xi}(\lambda r).

are the eigenvalues of the operators HλH_{\lambda}, respectively H~λ\tilde{H}_{\lambda}; obviously for HλH_{\lambda} we also have the zero-eigenvalue ϕ0λ=h1λ\phi_{0}^{\lambda}=h_{1}^{\lambda}. Using these new eigenfunctions, the spectral theory for HλH_{\lambda} and H~λ\tilde{H}_{\lambda} is similar to the corresponding one for λ=1\lambda=1.

For instance, we have

(5.29) ~λ(u)(ξ)=0ψξλ(r)u(r)r𝑑r=1λ3/2~(u(λ))(ξλ).{\tilde{\mathcal{F}}}_{\lambda}(u)(\xi)=\int_{0}^{\infty}\psi^{\lambda}_{\xi}(r)u(r)rdr=\frac{1}{\lambda^{3/2}}{\tilde{\mathcal{F}}}(u(\frac{\cdot}{\lambda}))(\frac{\xi}{\lambda}).

Then, the Fourier inversion formula for ~{\tilde{\mathcal{F}}} gives

(5.30) ~λ1(u)(r)=0ψξλ(r)u(ξ)𝑑ξ.{\tilde{\mathcal{F}}}^{-1}_{\lambda}(u)(r)=\int_{0}^{\infty}\psi^{\lambda}_{\xi}(r)u(\xi)d\xi.

Both integrals above are considered in the singular sense.

5.1. The generalized eigenfunctions of HH: Proof of Theorem 5.1

We will carry out the proof in three steps:

STEP 1:

We construct ϕξ\phi_{\xi} in the region rξ1r\xi\lesssim 1, modulo the determination of the normalization coefficient q(ξ)q(\xi). We also complete ϕξ\phi_{\xi} to a basis of solutions for (5.3) by constructing a second solution θξ\theta_{\xi} satisfying a Wronskian condition.

STEP II:

We construct ϕξ+\phi_{\xi}^{+} in the region rξ1r\xi\gtrsim 1, modulo the determination of the normalization coefficient a(ξ)a(\xi).

STEP III:

We determine the choice of the normalization coefficients qq and aa so that they match, and so that we have the L2L^{2} isometry property.

STEP I: Generalized eigenfunctions in the region rξ1r\xi\lesssim 1. We consider a basis (ϕ0,θ0)(\phi_{0},\theta_{0}) in the null space of HH,

Hϕ0=Hθ0=0,H\phi_{0}=H\theta_{0}=0,

where θ0\theta_{0} is unbounded at r=0r=0, and given by

(5.31) θ0(r)=r8+r48r4log(r)+18(r6+r2)=18r2+r28(1+11+r4)log(r)2ϕ0(r).\theta_{0}(r)=\frac{-r^{8}+r^{4}-8r^{4}\log(r)+1}{8\left(r^{6}+r^{2}\right)}=\frac{1}{8r^{2}}+\frac{r^{2}}{8}\left(-1+\frac{1}{1+r^{4}}\right)-\frac{\log(r)}{2}\phi_{0}(r).

With this choice of θ0\theta_{0}, we have the Wronskian normalization

r(rϕ0(r)θ0(r)ϕ0(r)rθ0(r))=1.r\left(\partial_{r}\phi_{0}(r)\theta_{0}(r)-\phi_{0}(r)\partial_{r}\theta_{0}(r)\right)=1.

We use these two functions to define a solution operator for the inhomogeneous problem

Hψ=g.H\psi=g.

Precisely, for gg vanishing quadratically at 0, the unique solution ψ\psi which also vanishes of order four at r=0r=0 is given by the variation of parameters formula

(5.32) ψ(r)=𝒯(f)(r):=ϕ0(r)0rF(s)θ0(s)s𝑑s+θ0(r)0rF(s)ϕ0(s)s𝑑s.\psi(r)=\mathcal{T}(f)(r):=-\phi_{0}(r)\int_{0}^{r}F(s)\theta_{0}(s)sds+\theta_{0}(r)\int_{0}^{r}F(s)\phi_{0}(s)sds.

We rewrite this in the form

(5.33) 𝒯(f)(r)=0rT(r,s)f(s)s𝑑s,\mathcal{T}(f)(r)=\int_{0}^{r}T(r,s)f(s)sds,

where the kernel TT is given by

(5.34) T(r,s)=(θ0(r)ϕ0(s)ϕ0(r)θ0(s))=(s4r4)(1+r4s4)+8r4s4(log(sr))4r2s2(r4+1)(s4+1),T(r,s)=\left(\theta_{0}(r)\phi_{0}(s)-\phi_{0}(r)\theta_{0}(s)\right)=\frac{(s^{4}-r^{4})(1+r^{4}s^{4})+8r^{4}s^{4}(\log(\frac{s}{r}))}{4r^{2}s^{2}\left(r^{4}+1\right)\left(s^{4}+1\right)},

and has the property

T(r,s)0,0<sr.T(r,s)\leq 0,\quad 0<s\leq r.

We can use this to iteratively compute the functions ϕj\phi_{j} in (5.7). Given the equation (5.3), these functions must satisfy the recurrence relations

(5.35) {H(r2ϕ1(r2))=ϕ0(r)H(r2jϕj(r2))=r2(j1)ϕj1(r2),j2.\begin{cases}H(r^{2}\phi_{1}(r^{2}))=\phi_{0}(r)\\ H(r^{2j}\phi_{j}(r^{2}))=r^{2(j-1)}\phi_{j-1}(r^{2}),\quad j\geq 2.\end{cases}

This implies that we must have

(5.36) {r2ϕ1(r2)=𝒯(ϕ0)(r)r2jϕj(r2)=𝒯(()2(j1)ϕj1(()2))(r),j2,\begin{cases}r^{2}\phi_{1}(r^{2})=\mathcal{T}(\phi_{0})(r)\\ r^{2j}\phi_{j}(r^{2})=\mathcal{T}((\cdot)^{2(j-1)}\phi_{j-1}((\cdot)^{2}))(r),\quad j\geq 2,\end{cases}

which we take as the inductive definition of the functions ϕj\phi_{j}. Our next objective is to show that these functions satisfy the bounds (5.8) and (5.9) in the theorem. Once this is done, we can conclude that ϕξ\phi_{\xi} given by (5.7) solves indeed the eigenfunction equation (5.3).


For j=1j=1 we have

(5.37) ϕ1(u)=2u20u2tan1(x)x𝑑x+(u41)tan1(u)3u3+u8(u4+u2).\phi_{1}(u)=-\frac{2u^{2}\int_{0}^{u}\frac{2\tan^{-1}(x)}{x}dx+\left(u^{4}-1\right)\tan^{-1}(u)-3u^{3}+u}{8\left(u^{4}+u^{2}\right)}\ .

For j2j\geq 2, we have

r2jϕj(r2)=0rs2(j1)ϕj1(s2)T(r,s)s𝑑s,r^{2j}\phi_{j}(r^{2})=\int_{0}^{r}s^{2(j-1)}\phi_{j-1}(s^{2})T(r,s)sds,

which gives

ujϕj(u)=0uyj1ϕj1(y)T(u,y)2𝑑y.u^{j}\phi_{j}(u)=\int_{0}^{u}y^{j-1}\phi_{j-1}(y)\frac{T(\sqrt{u},\sqrt{y})}{2}dy.

By (5.37), there exists C1>0C_{1}>0 such that

|ϕ1(u)|C1uu,u0,|\phi_{1}(u)|\leq C_{1}\frac{u}{\langle u\rangle},\quad u\geq 0,

which gives (5.8) for j=1j=1. For j1j\geq 1, we let Cj=C1((j1)!)2(323/2)j1C_{j}=\frac{C_{1}}{((j-1)!)^{2}}\left(\frac{3}{2^{3/2}}\right)^{j-1} and prove (5.8) inductively. If we suppose, for some j1j\geq 1, that

|ϕj(u)|Cjuu,u>0,|\phi_{j}(u)|\leq C_{j}\frac{u}{\langle u\rangle},\qquad u>0,

then

uj+1ϕj+1(u)=0uyjϕj(y)((y2u2)(1+u2y2)+4u2y2log(yu)8uy(u2+1)(y2+1))𝑑y.u^{j+1}\phi_{j+1}(u)=\int_{0}^{u}y^{j}\phi_{j}(y)\left(\frac{(y^{2}-u^{2})(1+u^{2}y^{2})+4u^{2}y^{2}\log(\frac{y}{u})}{8uy(u^{2}+1)(y^{2}+1)}\right)dy.

If u1u\leq 1, we estimate this by

(5.38) |uj+1ϕj+1(u)|0uyjCjy(u2y28uy+uylog(uy)2)𝑑yCj(18u2uj+3(3+4j+j2)+uj+401xj+22log(1x)𝑑x)34j2Cjuj+2,\begin{split}|u^{j+1}\phi_{j+1}(u)|&\leq\int_{0}^{u}y^{j}C_{j}y\left(\frac{u^{2}-y^{2}}{8uy}+\frac{uy\log(\frac{u}{y})}{2}\right)dy\\ &\leq C_{j}\left(\frac{1}{8u}\frac{2u^{j+3}}{(3+4j+j^{2})}+u^{j+4}\int_{0}^{1}\frac{x^{j+2}}{2}\log(\frac{1}{x})dx\right)\\ &\leq\frac{3}{4j^{2}}C_{j}u^{j+2},\end{split}

from which it follows that

|ϕj+1(u)|324j2Cjuu,u1.|\phi_{j+1}(u)|\leq\frac{3\sqrt{2}}{4j^{2}}C_{j}\frac{u}{\langle u\rangle},\quad u\leq 1.

On the other hand, if u1u\geq 1, then we have

(5.39) |uj+1ϕj+1(u)|0uyjCj((u2y2)8uy+12uylog(uy))𝑑yCjuj+14j2+Cjuj12j2,u1.\begin{split}|u^{j+1}\phi_{j+1}(u)|&\leq\int_{0}^{u}y^{j}C_{j}\left(\frac{(u^{2}-y^{2})}{8uy}+\frac{1}{2uy}\log(\frac{u}{y})\right)dy\\ &\leq\frac{C_{j}u^{j+1}}{4j^{2}}+\frac{C_{j}u^{j-1}}{2j^{2}},\quad u\geq 1.\end{split}

Combining the two cases, we get

|ϕj+1(u)|3Cj24j2uu=Cj+1uu,u>0,|\phi_{j+1}(u)|\leq\frac{3C_{j}\sqrt{2}}{4j^{2}}\frac{u}{\langle u\rangle}=C_{j+1}\frac{u}{\langle u\rangle},\quad u>0,

which completes the inductive proof of (5.8).

Next, we consider higher order derivatives of ϕj\phi_{j} in order to prove (5.9). From (5.37) we have

|ϕ1(u)|Cu,u>0|\phi_{1}^{\prime}(u)|\leq\frac{C}{\langle u\rangle},\quad u>0

For j2j\geq 2 we compute

(5.40) u(ujϕj(u))=0uyj1ϕj1(y)h1(y,u)𝑑y\partial_{u}\left(u^{j}\phi_{j}(u)\right)=-\int_{0}^{u}y^{j-1}\phi_{j-1}(y)h^{1}(y,u)\,dy

where

h1(y,u)=y2+u2(1u2+(7+7u2+u4)y2+(1+u2)y4)+4u2(1+u2)y2log(yu)8u2(1+u2)2y(y2+1).h^{1}(y,u)=\frac{y^{2}+u^{2}(1-u^{2}+(7+7u^{2}+u^{4})y^{2}+(-1+u^{2})y^{4})+4u^{2}(-1+u^{2})y^{2}\log(\frac{y}{u})}{8u^{2}(1+u^{2})^{2}y(y^{2}+1)}.

Then, we use (5.8) to directly estimate the integral. This yields

|u(ujϕj(u))|CC1((j2)!)2(j1)(323/2)j2uju.|\partial_{u}\left(u^{j}\phi_{j}(u)\right)|\leq\frac{CC_{1}}{((j-2)!)^{2}(j-1)}\left(\frac{3}{2^{3/2}}\right)^{j-2}\frac{u^{j}}{\langle u\rangle}.

Hence, for all j1j\geq 1 we have

|ϕj(u)|CC1((j1)!)2(323/2)j1ju|\phi_{j}^{\prime}(u)|\leq\frac{CC_{1}}{((j-1)!)^{2}}\left(\frac{3}{2^{3/2}}\right)^{j-1}\frac{j}{\langle u\rangle}

which is the case n=1n=1 of (5.9).

For larger nn we denote

fj(r)=r2jϕj(r2).f_{j}(r)=r^{2j}\phi_{j}(r^{2}).

Then by (5.35) we have

fj′′(r)=1rfj(r)+4(16r4+r8)fj(r)r2(1+r4)2fj1(r),f_{j}^{\prime\prime}(r)=\frac{-1}{r}f_{j}^{\prime}(r)+\frac{4(1-6r^{4}+r^{8})f_{j}(r)}{r^{2}(1+r^{4})^{2}}-f_{j-1}(r),

which implies

|fj′′(r)|CC1j2((j1)!)2(323/2)j1r2jr2,|f_{j}^{\prime\prime}(r)|\leq\frac{CC_{1}j^{2}}{((j-1)!)^{2}}\left(\frac{3}{2^{3/2}}\right)^{j-1}\frac{r^{2j}}{\langle r^{2}\rangle},

thereby completing the proof of (5.9) for n=2n=2.

For larger nn, if we inductively assume that

|fj(n)(r)|Mnjn((j1)!)2(323/2)j1r2j+2r2rn|f_{j}^{(n)}(r)|\leq\frac{M_{n}j^{n}}{((j-1)!)^{2}}\left(\frac{3}{2^{3/2}}\right)^{j-1}\frac{r^{2j+2}}{\langle r^{2}\rangle r^{n}}

for all 0nk+10\leq n\leq k+1 for some k1k\geq 1, then, using

fj(k+2)(r)=l=0k(kl)rkl(1r)fj(l+1)(r)+l=0k(kl)rkl(4(16r4+r8)r2(1+r4)2)fj(l)(r)fj1(k)(r)f_{j}^{(k+2)}(r)=-\sum_{l=0}^{k}\binom{k}{l}\partial_{r}^{k-l}\left(\frac{1}{r}\right)f_{j}^{(l+1)}(r)+\sum_{l=0}^{k}\binom{k}{l}\partial_{r}^{k-l}\left(\frac{4(1-6r^{4}+r^{8})}{r^{2}(1+r^{4})^{2}}\right)f_{j}^{(l)}(r)-f_{j-1}^{(k)}(r)

we obtain

|fj(k+2)(r)|Mk+2jk+2rk+2((j1)!)2(323/2)j1r2j+2r2,|f_{j}^{(k+2)}(r)|\leq\frac{M_{k+2}j^{k+2}}{r^{k+2}((j-1)!)^{2}}\left(\frac{3}{2^{3/2}}\right)^{j-1}\frac{r^{2j+2}}{\langle r^{2}\rangle},

and this completes the proof of (5.9) by induction.

This finishes the analysis in part (a) of the theorem, except for the choice of the normalization coefficient q(ξ)q(\xi). This choice will only be considered in the last step of the proof, where we connect the solutions of (5.3) near r=0r=0 with the solutions near r=r=\infty. In order to accomplish this, we supplement the above solution ϕξ\phi_{\xi} with a second solution θξ\theta_{\xi},

H(θξ)=ξ2θξH(\theta_{\xi})=\xi^{2}\theta_{\xi}

with the Wronskian normalization

(5.41) r(rϕξ(r)θξ(r)ϕξ(r)rθξ(r))=1r\left(\partial_{r}\phi_{\xi}(r)\theta_{\xi}(r)-\phi_{\xi}(r)\partial_{r}\theta_{\xi}(r)\right)=1

which determines θξ\theta_{\xi} uniquely up to a multiple of ϕξ\phi_{\xi}:

Lemma 5.3.

In the region rξ1r\xi\lesssim 1 one could choose θξ\theta_{\xi} of the form

θξ(r)=18q(ξ)r2+ξ232q(ξ)+18q(ξ)r2k=0θk~(r)r2kξ2k1q(ξ)2(12+ξ4256)log(r)ϕξ(r)\theta_{\xi}(r)=\frac{1}{8q(\xi)r^{2}}+\frac{\xi^{2}}{32q(\xi)}+\frac{1}{8q(\xi)r^{2}}\sum_{k=0}^{\infty}\widetilde{\theta_{k}}(r)r^{2k}\xi^{2k}-\frac{1}{q(\xi)^{2}}\left(\frac{1}{2}+\frac{\xi^{4}}{256}\right)\log(r)\phi_{\xi}(r)

where we have

θ0~(r)=r81+r4\widetilde{\theta_{0}}(r)=\frac{-r^{8}}{1+r^{4}}

and the bounds

|θk~(r)|{Cr4,k=1,2Tkk!(1+r4),k3.|\widetilde{\theta_{k}}(r)|\leq\begin{cases}Cr^{4},\quad k=1,2\\ \frac{T^{k}}{k!}(1+r^{4}),\quad k\geq 3\end{cases}.

Note that our definition of θ0\theta_{0} is such that

limξ0+(θξ(r)q(ξ))=θ0(r)\lim_{\xi\rightarrow 0^{+}}(\theta_{\xi}(r)q(\xi))=\theta_{0}(r)

The logr\log r factor serves to provide a constant term in HθξH\theta_{\xi}; all other contributions should be analytic in rr at r=0r=0. This exactly corresponds to a similar term in θ0\theta_{0}, see (5.31).

Proof.

It is natural to try the ansatz

θξ(r)=c0(ξ)r2+c1(ξ)+c0(ξ)r2k=0dk(r)r2kξ2k+c2(ξ)log(r)ϕξ(r)\theta_{\xi}(r)=\frac{c_{0}(\xi)}{r^{2}}+c_{1}(\xi)+c_{0}(\xi)r^{2}\sum_{k=0}^{\infty}d_{k}(r)r^{2k}\xi^{2k}+c_{2}(\xi)\log(r)\phi_{\xi}(r)

where dkd_{k} are analytic in r2r^{2} and all contributions from the sum are at least O(r4)O(r^{4}) at r=0r=0. We have

(5.42) H(θξ(r))=32(1+r4)2c0(ξ)+c1(ξ)4(16r4+r8)r2(1+r4)2+H(r2k=0dk(r)r2kξ2k)c0(ξ)+c2(ξ)(log(r)ξ2ϕξ(r)2rrϕξ(r))\begin{split}H(\theta_{\xi}(r))&=\frac{-32}{(1+r^{4})^{2}}c_{0}(\xi)+c_{1}(\xi)\cdot\frac{4(1-6r^{4}+r^{8})}{r^{2}(1+r^{4})^{2}}+H\left(r^{2}\sum_{k=0}^{\infty}d_{k}(r)r^{2k}\xi^{2k}\right)c_{0}(\xi)\\ &+c_{2}(\xi)\left(\log(r)\xi^{2}\phi_{\xi}(r)-\frac{2}{r}\partial_{r}\phi_{\xi}(r)\right)\end{split}

and

(5.43) ξ2θξ(r)=ξ2c0(ξ)r2+ξ2c1(ξ)+r2ξ2k=0dk(r)r2kξ2kc0(ξ)+ξ2c2(ξ)log(r)ϕξ(r)\xi^{2}\theta_{\xi}(r)=\frac{\xi^{2}c_{0}(\xi)}{r^{2}}+\xi^{2}c_{1}(\xi)+r^{2}\xi^{2}\sum_{k=0}^{\infty}d_{k}(r)r^{2k}\xi^{2k}c_{0}(\xi)+\xi^{2}c_{2}(\xi)\log(r)\phi_{\xi}(r)

Then the equation Hθξ=ξ2θξH\theta_{\xi}=\xi^{2}\theta_{\xi} implies

(5.44) 32c0(ξ)(1+r4)2+c1(ξ)(4r232r2(1+r4)2)+c0(ξ)H(r2k=0dk(r)r2kξ2k)=ξ2c0(ξ)r2+ξ2c1(ξ)+r2ξ2k=0dk(r)r2kξ2kc0(ξ)+c2(ξ)2rrϕξ(r)\begin{split}&-\frac{32c_{0}(\xi)}{(1+r^{4})^{2}}+c_{1}(\xi)\left(\frac{4}{r^{2}}-\frac{32r^{2}}{(1+r^{4})^{2}}\right)+c_{0}(\xi)H\left(r^{2}\sum_{k=0}^{\infty}d_{k}(r)r^{2k}\xi^{2k}\right)\\ &=\frac{\xi^{2}c_{0}(\xi)}{r^{2}}+\xi^{2}c_{1}(\xi)+r^{2}\xi^{2}\sum_{k=0}^{\infty}d_{k}(r)r^{2k}\xi^{2k}c_{0}(\xi)+c_{2}(\xi)\frac{2}{r}\partial_{r}\phi_{\xi}(r)\end{split}

If we look for a solution for which

limr0H(r2k=0dk(r)r2kξ2k)=0\lim_{r\rightarrow 0}H\left(r^{2}\sum_{k=0}^{\infty}d_{k}(r)r^{2k}\xi^{2k}\right)=0

then, by examining the 1r2\dfrac{1}{r^{2}} terms on each side, we obtain

c1(ξ)=ξ2c0(ξ)4.c_{1}(\xi)=\frac{\xi^{2}c_{0}(\xi)}{4}.

By considering the limit as r0+r\rightarrow 0^{+} of both sides, we get

32c0(ξ)2limr0(c2(ξ)q(ξ)ϕ0(r)r)=ξ2c1(ξ)-32c_{0}(\xi)-2\lim_{r\rightarrow 0}\left(c_{2}(\xi)q(\xi)\frac{\phi_{0}^{\prime}(r)}{r}\right)=\xi^{2}c_{1}(\xi)

and this gives

c2(ξ)=(4+ξ432)c0(ξ)q(ξ).c_{2}(\xi)=-\left(4+\frac{\xi^{4}}{32}\right)\frac{c_{0}(\xi)}{q(\xi)}.

Finally, the coefficient c0c_{0} is determined by examining the Wronskian relation (5.41) in the limit as rr approaches 0. This gives

c0(ξ)=18q(ξ).c_{0}(\xi)=\frac{1}{8q(\xi)}.

Then we can factor out all the c0c_{0} factors, and the equation Hθξ=ξ2θξH\theta_{\xi}=\xi^{2}\theta_{\xi} reduces to

(5.45) 32r4(2+r4)(1+r4)28ξ2r24(1+r4)2+H(r2k=0dk(r)r2kξ2k)=r2ξ2k=0dk(r)r2kξ2k(4+ξ432)(2rq(ξ)rϕξ(r)8).\begin{split}&\frac{32r^{4}(2+r^{4})}{(1+r^{4})^{2}}-\frac{8\xi^{2}r^{2}}{4(1+r^{4})^{2}}+H\left(r^{2}\sum_{k=0}^{\infty}d_{k}(r)r^{2k}\xi^{2k}\right)\\ &=r^{2}\xi^{2}\sum_{k=0}^{\infty}d_{k}(r)r^{2k}\xi^{2k}-\left(4+\frac{\xi^{4}}{32}\right)\left(\frac{2}{rq(\xi)}\partial_{r}\phi_{\xi}(r)-8\right).\end{split}

At this point we use the expansion (5.7) for ϕξ\phi_{\xi}, to obtain

(5.46) H(r2k=0dk(r)r2kξ2k)=32r4(1+r4)2+8ξ2r2(1+r4)2+ξ4r4(3+r4)4(1+r4)2+r2ξ2k=0dk(r)r2kξ2k2r(4+ξ432)k=1r(r2kϕk(r2))ξ2k.\begin{split}H\left(r^{2}\sum_{k=0}^{\infty}d_{k}(r)r^{2k}\xi^{2k}\right)=&\ \frac{32r^{4}}{(1+r^{4})^{2}}+\frac{8\xi^{2}r^{2}}{(1+r^{4})^{2}}+\frac{\xi^{4}r^{4}(3+r^{4})}{4(1+r^{4})^{2}}\\ &+r^{2}\xi^{2}\sum_{k=0}^{\infty}d_{k}(r)r^{2k}\xi^{2k}-\frac{2}{r}\left(4+\frac{\xi^{4}}{32}\right)\sum_{k=1}^{\infty}\partial_{r}\left(r^{2k}\phi_{k}(r^{2})\right)\xi^{2k}.\end{split}

Now we simply match the coefficients of ξ2j\xi^{2j}. By considering the ξ0\xi^{0} terms, we get

H(r2d0(r))=32r4(3+r4)(1+r4)232r4(2+r4)(1+r4)2H\left(r^{2}d_{0}(r)\right)=\frac{32r^{4}(3+r^{4})}{(1+r^{4})^{2}}-\frac{32r^{4}(2+r^{4})}{(1+r^{4})^{2}}

which gives

d0(r)=1+11+r4.d_{0}(r)=-1+\frac{1}{1+r^{4}}.

By examining the ξ2\xi^{2} terms, we obtain

H(r4d1(r))=8r2(1+r4)28rr(r2ϕ1(r2))+r2d0(r).H\left(r^{4}d_{1}(r)\right)=\frac{8r^{2}}{(1+r^{4})^{2}}-\frac{8}{r}\partial_{r}\left(r^{2}\phi_{1}(r^{2})\right)+r^{2}d_{0}(r).

The right hand side has size O(r2)O(r^{2}) at zero, so the above equation has an unique solution of size O(r4)O(r^{4}), namely the one given by (5.32),

r4d1(r)=0r(s2(8+s4+s8)(1+s4)28ss(s2ϕ1(s2)))T(r,s)s𝑑s.r^{4}d_{1}(r)=\int_{0}^{r}\left(\frac{-s^{2}(-8+s^{4}+s^{8})}{(1+s^{4})^{2}}-\frac{8}{s}\partial_{s}\left(s^{2}\phi_{1}(s^{2})\right)\right)T(r,s)sds.

The integrand is an analytic, odd function of ss, except for the log factor. But the log factor can be eliminated integrating by parts, for instance

0rs2k+1log(s/r)𝑑s=12k+20rs2k+1𝑑s=1(2k+2)2r2k+2.\int_{0}^{r}s^{2k+1}\log(s/r)ds=-\frac{1}{2k+2}\int_{0}^{r}s^{2k+1}ds=-\frac{1}{(2k+2)^{2}}r^{2k+2}.

So we obtain an analytic, even function of rr. Using the estimates on ϕ1\phi_{1} we get

|d1(r)|1,r>0.|d_{1}(r)|\lesssim 1,\quad r>0.

Similarly, we consider the ξ4\xi^{4} terms, to get

H(r6d2(r))=r4d1(r)8rr(r4ϕ2(r2))+r44(3+r4)(1+r4)2H\left(r^{6}d_{2}(r)\right)=r^{4}d_{1}(r)-\frac{8}{r}\partial_{r}\left(r^{4}\phi_{2}(r^{2})\right)+\frac{r^{4}}{4}\frac{(3+r^{4})}{(1+r^{4})^{2}}

which is solved again using (5.32) with the kernel TT as in (5.34) to obtain

r6d2(r)=0r(s4d1(s)8ss(s4ϕ2(s2))+s44(3+s4)(1+s4)2)T(r,s)s𝑑s.r^{6}d_{2}(r)=\int_{0}^{r}\left(s^{4}d_{1}(s)-\frac{8}{s}\partial_{s}\left(s^{4}\phi_{2}(s^{2})\right)+\frac{s^{4}}{4}\frac{(3+s^{4})}{(1+s^{4})^{2}}\right)T(r,s)sds.

We obtain a function d2d_{2} which is even, analytic, and using the estimates on ϕj\phi_{j} and d1d_{1} above, we get

|d2(r)|1,r>0.|d_{2}(r)|\lesssim 1,\quad r>0.

Finally, we let

gk(r)=r2k+1dk(r),k2,g_{k}(r)=r^{2k+1}d_{k}(r),\quad k\geq 2,

and note that Hθ=ξ2θH\theta=\xi^{2}\theta is completely solved if, for all k3k\geq 3,

(5.47) rgk(r)=0r(sgk1(s)116ss(s2k4ϕk2(s2))8ss(s2kϕk(s2)))T(r,s)s𝑑s.\begin{split}rg_{k}(r)=\int_{0}^{r}\left(sg_{k-1}(s)-\frac{1}{16s}\partial_{s}\left(s^{2k-4}\phi_{k-2}(s^{2})\right)-\frac{8}{s}\partial_{s}\left(s^{2k}\phi_{k}(s^{2})\right)\right)T(r,s)sds.\end{split}

Using the ϕj\phi_{j} estimates and the d2d_{2} estimate, we obtain

|g3(r)|C3~r3(1+r4).|g_{3}(r)|\leq\widetilde{C_{3}}r^{3}(1+r^{4}).

If we suppose, for some k13k-1\geq 3, that

|gk1(r)|Ck1~r2k5(1+r4).|g_{k-1}(r)|\leq\widetilde{C_{k-1}}r^{2k-5}(1+r^{4}).

then, (5.47) gives the following. There exists C>0C>0 so that

|gk(r)|Cr2k3(1+r4)(Ckk!+ck1~k).|g_{k}(r)|\leq Cr^{2k-3}(1+r^{4})\left(\frac{C^{k}}{k!}+\frac{\widetilde{c_{k-1}}}{k}\right).

Therefore, there exists TT sufficiently large so that, by induction, for all k3k\geq 3, and r>0r>0,

|gk(r)|Tkk!r2k3(1+r4).|g_{k}(r)|\leq\frac{T^{k}}{k!}r^{2k-3}(1+r^{4}).

Setting θk~(r)=dk(r)r4\widetilde{\theta_{k}}(r)=d_{k}(r)r^{4} yields

θ0~(r)=r81+r4,|θk~(r)|{Cr4,k=1,2Tkk!(1+r4),k3\widetilde{\theta_{0}}(r)=\frac{-r^{8}}{1+r^{4}},\quad|\widetilde{\theta_{k}}(r)|\leq\begin{cases}Cr^{4},\quad k=1,2\\ \frac{T^{k}}{k!}(1+r^{4}),\quad k\geq 3\end{cases}

which finishes the proof of the claim.

This completes the first step of the proof, i.e. the analysis of the rξ1r\xi\lesssim 1 region.


STEP II: Generalized eigenfunctions in the region rξ1r\xi\gtrsim 1. Here our aim is to construct the generalized eigenfunctions

(5.48) ϕξ+(r)=r12eirξσ(rξ,r),rξ1\phi^{+}_{\xi}(r)=r^{-\frac{1}{2}}e^{ir\xi}\sigma(r\xi,r),\qquad r\xi\gtrsim 1

solving

Hϕξ+=ξ2ϕξ+.H\phi^{+}_{\xi}=\xi^{2}\phi^{+}_{\xi}.

Substituting, this yields a second order equation for the amplitude σ\sigma:

(5.49) 0=\displaystyle 0= eirξr(Hξ2)(eirξσ(rξ,r)r)\displaystyle\ e^{-ir\xi}\sqrt{r}\left(H-\xi^{2}\right)\left(\frac{e^{ir\xi}\sigma(r\xi,r)}{\sqrt{r}}\right)
=\displaystyle= r2(σ(rξ,r))2iξr(σ(rξ,r))+V(r)σ(rξ,r),\displaystyle\ -\partial_{r}^{2}\left(\sigma(r\xi,r)\right)-2i\xi\partial_{r}\left(\sigma(r\xi,r)\right)+V(r)\sigma(r\xi,r),

where we recall, see (5.2), that

V(r)=154r232r2(1+r4)2.V(r)=\frac{15}{4r^{2}}-32\frac{r^{2}}{(1+r^{4})^{2}}.

We first look for a WKB type asymptotic series expansion of σ(rξ,r)\sigma(r\xi,r) in the region of large rξr\xi. This is derived by considering the system of equations arising from formally substituting

σ(rξ,r)=j=0fj(r)ξj,f0(r)=1\sigma(r\xi,r)=\sum_{j=0}^{\infty}f_{j}(r)\xi^{-j},\qquad f_{0}(r)=1

into (5.49). This yields

fk+1(r)=12i(V(r)fk(r)fk′′(r)),k0,f_{k+1}^{\prime}(r)=\frac{1}{2i}\left(V(r)f_{k}(r)-f_{k}^{\prime\prime}(r)\right),\quad k\geq 0,

where we look for the unique solutions fkf_{k} satisfying a boundary condition at infinity,

limrfk(r)=0,k1.\lim_{r\rightarrow\infty}f_{k}(r)=0,\quad k\geq 1.

Setting

(5.50) ϕk+(r)=rkfk(r)\phi_{k}^{+}(r)=r^{k}f_{k}(r)

we compute directly

ϕ1+(r)=4ir4+1ir(log(r2+2r+1r22r+1)+2(πtan1(2r1)tan1(2r+1)))2+47i8.\phi_{1}^{+}(r)=\frac{-4i}{r^{4}+1}-\frac{ir\left(\log\left(\frac{r^{2}+\sqrt{2}r+1}{r^{2}-\sqrt{2}r+1}\right)+2(\pi-\tan^{-1}\left(\sqrt{2}r-1\right)-\tan^{-1}\left(\sqrt{2}r+1\right))\right)}{\sqrt{2}}+\frac{47i}{8}.

In particular we note that

ϕ1+(r)=15i816i5r4+O(1r8),r.\phi_{1}^{+}(r)=\frac{15i}{8}-\frac{16i}{5r^{4}}+O\left(\frac{1}{r^{8}}\right),\quad r\rightarrow\infty.

By inspection one easily sees that there exist constants Cm>0C_{m}>0 such that

|(rr)mϕ1+(r)|Cm,r>0.|(r\partial_{r})^{m}\phi_{1}^{+}(r)|\leq C_{m},\quad r>0.

Then, by induction, it easily follows for all m0m\geq 0, j0j\geq 0 there exist constants Cm,j>0C_{m,j}>0 such that

|(rr)mϕj+(r)|Cm,j,r>0.|(r\partial_{r})^{m}\phi_{j}^{+}(r)|\leq C_{m,j},\quad r>0.

Overall, we thus have the asymptotic expansion

σ(y,r)j=0yjϕj+(r).\sigma(y,r)\sim\sum_{j=0}^{\infty}y^{-j}\phi_{j}^{+}(r).

The difficulty at this point is that we do not know that this series converges. To rectify this, as usual in WKB approximations, we construct an approximate sum by truncating the terms in the series at well chosen points. Precisely, we define σap\sigma_{ap} by

σap(y,r)=k=0ykχ1(1yδk)ϕk+(r),\sigma_{ap}(y,r)=\sum_{k=0}^{\infty}y^{-k}\chi_{\leq 1}(\frac{1}{y\delta_{k}})\phi_{k}^{+}(r),

where

χ1(x)={1,x120,x1,χ1C(),\chi_{\leq 1}(x)=\begin{cases}1,\quad x\leq\frac{1}{2}\\ 0,\quad x\geq 1\end{cases},\quad\chi_{\leq 1}\in C^{\infty}(\mathbb{R}),

and δk0\delta_{k}\rightarrow 0 rapidly enough in order to ensure convergence.

Comparing the approximate sum with the exact partial sums, we have that for each N,m,n0N,m,n\geq 0,

(5.51) |(yy)m(rr)n(k=0ykχ1(1yδk)ϕk+(r)k=0Nykϕk+(r))|CN,n,myN+1,r>0,y1.\left|(y\partial_{y})^{m}\left(r\partial_{r}\right)^{n}\left(\sum_{k=0}^{\infty}y^{-k}\chi_{\leq 1}(\frac{1}{y\delta_{k}})\phi_{k}^{+}(r)-\sum_{k=0}^{N}y^{-k}\phi_{k}^{+}(r)\right)\right|\leq\frac{C_{N,n,m}}{y^{N+1}},\quad r>0,\ y\geq 1.

This implies that σap\sigma_{ap} is a good approximate solution for (5.49) near infinity, in the sense that its associated source term

e(rξ,r)=(r2+2iξrV(r))(σap(rξ,r))e(r\xi,r)=-\left(\partial_{r}^{2}+2i\xi\partial_{r}-V(r)\right)\left(\sigma_{ap}(r\xi,r)\right)

is rapidly decreasing at infinity, i.e. for all N,m,n0N,m,n\geq 0,

(5.52) |(yy)m(rr)ne(y,r)|CN,n,mr2yN,r>0,y1.\left|(y\partial_{y})^{m}(r\partial_{r})^{n}e(y,r)\right|\leq\frac{C_{N,n,m}}{r^{2}y^{N}},\quad r>0,\ y\geq 1.

We then need to complete σap\sigma_{ap} to an exact solution to (5.49). Writing

σ(rξ,r)=σap(rξ,r)+f(rξ,r),\sigma(r\xi,r)=\sigma_{ap}(r\xi,r)+f(r\xi,r),

the correction ff should solve

(5.53) (r22iξr+V(r))f(rξ,r)=e(rξ,r).\left(-\partial_{r}^{2}-2i\xi\partial_{r}+V(r)\right)f(r\xi,r)=e(r\xi,r).

We will show that, given ee as in (5.52), there exists a unique solution ff satisfying the bounds

(5.54) |(yy)m(rr)nf(y,r)|CN,n,myN,r>0,y1.\left|(y\partial_{y})^{m}(r\partial_{r})^{n}f(y,r)\right|\leq\frac{C_{N,n,m}}{y^{N}},\quad r>0,\ y\geq 1.

To achieve this we interpret the ordinary differential equation (5.53) as a Cauchy problem with zero Cauchy data at infinity. Here it is convenient to use the variable ξ\xi as a parameter, instead of yy which also contains some rr dependence. By the chain rule we can interchange the operators (rr,yy)(r\partial_{r},y\partial_{y}) with (rr,ξξ)(r\partial_{r},\xi\partial_{\xi}). Then, changing notations to view ff and ee as functions of rr and ξ\xi, it remains to prove the following

Proposition 5.4.

Consider the equation

(5.55) (r22iξr+V(r))f(ξ,r)=e(ξ,r)\left(-\partial_{r}^{2}-2i\xi\partial_{r}+V(r)\right)f(\xi,r)=e(\xi,r)

Assume that ee satisfies the bounds

(5.56) |(ξξ)m(rr)ne(ξ,r)|CN,n,mr2(rξ)N,r>0,rξ1.\left|(\xi\partial_{\xi})^{m}(r\partial_{r})^{n}e(\xi,r)\right|\leq\frac{C_{N,n,m}}{r^{2}(r\xi)^{N}},\quad r>0,\ r\xi\geq 1.

Then there exists a unique solution satisfying the bound

(5.57) |(ξξ)m(rr)nf(ξ,r)|CN,n,m(rξ)N,r>0,rξ1.\left|(\xi\partial_{\xi})^{m}(r\partial_{r})^{n}f(\xi,r)\right|\leq\frac{C_{N,n,m}}{(r\xi)^{N}},\quad r>0,\ r\xi\geq 1.
Proof.

We first drop the vector fields (rr,ξξ)(r\partial_{r},\xi\partial_{\xi}) and consider our problem for fixed ξ\xi. We summarize the result in the following Lemma:

Lemma 5.5.

Suppose k>k0k>k_{0} and

|e(r)|rk2,rξ>1.|e(r)|\leq r^{-k-2},\qquad r\xi>1.

Then the equation (5.55) admits a unique solution ff which satisfies the bound

(5.58) |f(r)|krk,|rf(r)|krk1,|r2f(r)|krk2|f(r)|\lesssim_{k}r^{-k},\qquad|\partial_{r}f(r)|\lesssim_{k}r^{-k-1},\qquad|\partial_{r}^{2}f(r)|\lesssim_{k}r^{-k-2}
Proof.

The heart of the proof is an energy estimate which holds for any local solution to (5.55). This has the form

(5.59) r((|f|2+r2|rf|2)rC)Cr3+C|e(r)|2\partial_{r}\left((|f|^{2}+r^{2}|\partial_{r}f|^{2})r^{C}\right)\geq-Cr^{3+C}|e(r)|^{2}

with a universal CC. This directly gives uniqueness of solutions as in (5.58). Existence is obtained in a standard manner, by solving a truncated problem

(5.60) (r22iξr+V(r))fr0(r)=1{r<r0}e(r)\left(-\partial_{r}^{2}-2i\xi\partial_{r}+V(r)\right)f_{r_{0}}(r)=1_{\{r<r_{0}\}}e(r)

The desired solution ff is then obtained as the limit of f0f_{{}_{0}} as r0r_{0}\to\infty.

For later use we remark that we could also replace the sharp truncation with a smooth truncation on the dyadic scale. This has the advantage that also preserves the vector field bounds (5.56). ∎

Now we use the lemma to conclude the proof of the proposition. Since we restrict ourselves to the range rξ1r\xi\geq 1, it suffices to work with NN large enough. This provides sufficient decay for ee to allow us to apply the lemma. Then the uniqueness of ee is immediate. Further, the discussion above, it suffices to prove it under the qualitative assumption that ee has compact support. Then the smoothness of ff as a function of both rr and ξ\xi directly follows, and it remains to establish the bounds (5.57).

We prove these bounds by induction on mm. For m=0m=0 we start with the bounds in the lemma and get the higher rr derivatives directly from the equation. Suppose now, by induction, that we have the bounds (5.57) for m=km=k, and prove them for m=k+1m=k+1. Denoting

fm=(ξξ)mf,em=(ξξ)mef_{m}=(\xi\partial_{\xi})^{m}f,\qquad e_{m}=(\xi\partial_{\xi})^{m}e

we write an equation for fk+1f_{k+1}:

(r22iξr+V(r))fk+1(ξ,r)=ek+1(ξ,r)+2i(k+1)ξrfk\left(-\partial_{r}^{2}-2i\xi\partial_{r}+V(r)\right)f_{k+1}(\xi,r)=e_{k+1}(\xi,r)+2i(k+1)\xi\partial_{r}f_{k}

By the induction hypothesis we can include the second term on the right into ek+1e_{k+1}. This allows us to conclude by directly applying the m=0m=0 result. ∎

STEP 3: The normalization coefficients aa and qq. Now, we compute the asymptotic behavior of aa and qq. For this, it is convenient to follow the procedure of [22]. For comparison purposes we note that the theory developed in [22], which uses also parts from [21], has three normalization features that differ from our setup:

  • the corresponding eigenfunctions ϕξ\phi_{\xi} there are used for the eigenvalues ξ\xi versus ξ2\xi^{2};

  • the calculus on the physical side is with respect to the measure drdr versus rdrrdr in our case; this corresponds to the conjugation of HH by r\sqrt{r} in [22], which yields the operator KST\mathcal{L}_{KST}.

  • the calculus on the Fourier side is with respect to a measure ρ(ξ)dξ\rho(\xi)d\xi versus dξd\xi in our case; here we compensate this by using the additional coefficient q(ξ)q(\xi) in the expression of our eigenfunctions ϕξ(r)\phi_{\xi}(r).

Below we modify the elements of our spectral theory so as to fit the setup developed in [22] and [21] (which in part was based on the setup in [11]). This allows us to use that theory and then by reversing the process we can recover a(ξ)a(\xi) and q(ξ)q(\xi) in our setup.

Therefore, we define the following functions, which are counterparts of the eigenfunctions in Theorem 4.3 of [22]. In other words, repeating the identical procedure of [22] for our equation yields the functions ϕKST(r,ξ),θKST(r,ξ),ψKST+(r,ξ)\phi_{KST}(r,\xi),\theta_{KST}(r,\xi),\psi^{+}_{KST}(r,\xi) given below, which correspond to the functions ϕ(r,ξ),θ(r,ξ),ψ+(r,ξ)\phi(r,\xi),\theta(r,\xi),\psi^{+}(r,\xi) from Theorem 4.3 of [22]. We let

ϕKST(r,ξ)=ϕ0,KST(r)+r32j=1(r2ξ)jϕj,KST(r2)\phi_{KST}(r,\xi)=\phi_{0,KST}(r)+r^{-\frac{3}{2}}\sum_{j=1}^{\infty}(r^{2}\xi)^{j}\phi_{j,KST}(r^{2})

with

ϕ0,KST(r)=rϕ0(r)2=r5/21+r4.\phi_{0,KST}(r)=\frac{\sqrt{r}\phi_{0}(r)}{2}=\frac{r^{5/2}}{1+r^{4}}.

We also set

ϕj,KST(u)=u2ϕj(u),θKST(r,ξ)=2q(ξ)rθξ(r)\phi_{j,KST}(u)=\frac{u}{2}\phi_{j}(u),\quad\theta_{KST}(r,\xi)=2q(\sqrt{\xi})\sqrt{r}\theta_{\sqrt{\xi}}(r)

and

ψKST+(r,ξ)=rξ1/4ϕξ+(r)=eirξξ1/4σ(rξ,r).\psi_{KST}^{+}(r,\xi)=\frac{\sqrt{r}}{\xi^{1/4}}\phi_{\sqrt{\xi}}^{+}(r)=\frac{e^{ir\sqrt{\xi}}}{\xi^{1/4}}\sigma(r\sqrt{\xi},r).

With these notations we have

(ϕKST(,ξ))(r)=ξϕKST(r,ξ)\mathcal{L}(\phi_{KST}(\cdot,\xi))(r)=\xi\phi_{KST}(r,\xi)
(θKST(,ξ))=ξθKST(r,ξ)\mathcal{L}(\theta_{KST}(\cdot,\xi))=\xi\theta_{KST}(r,\xi)
(ψKST+(,ξ))=ξψKST+(r,ξ),\mathcal{L}(\psi^{+}_{KST}(\cdot,\xi))=\xi\psi_{KST}^{+}(r,\xi),

where \mathcal{L} is given in (5.2). From our functions θ\theta and ϕ\phi we inherit the Wronskian normalization

W(θKST(,ξ),ϕKST(,ξ))=1.W(\theta_{KST}(\cdot,\xi),\phi_{KST}(\cdot,\xi))=1.

On the other hand, using the above formulae for ψKST+\psi^{+}_{KST} and ψKST=ψKST+¯\psi^{-}_{KST}=\overline{\psi^{+}_{KST}}, we obtain

W(ψKST+,ψKST)=2i.W(\psi^{+}_{KST},\psi^{-}_{KST})=-2i.

Finally, since

ϕKST(r,ξ)=W(ψKST,ϕKST)W(ψKST,ψKST+)ψKST+(r,ξ)+W(ϕKST,ψKST+)W(ψKST,ψKST+)ψKST(r,ξ)\phi_{KST}(r,\xi)=\frac{W(\psi^{-}_{KST},\phi_{KST})}{W(\psi^{-}_{KST},\psi^{+}_{KST})}\psi^{+}_{KST}(r,\xi)+\frac{W(\phi_{KST},\psi^{+}_{KST})}{W(\psi^{-}_{KST},\psi^{+}_{KST})}\psi_{KST}^{-}(r,\xi)

and ϕKST\phi_{KST} is real-valued, while ψKST=ψKST+¯\psi^{-}_{KST}=\overline{\psi^{+}_{KST}}, we obtain

ϕKST(r,ξ)=aKST(ξ)ψKST+(r,ξ)+aKST(ξ)ψKST+(r,ξ)¯\phi_{KST}(r,\xi)=a_{KST}(\xi)\psi^{+}_{KST}(r,\xi)+\overline{a_{KST}(\xi)\psi^{+}_{KST}(r,\xi)}

with

aKST(ξ)=iW(ϕKST,ψKST)2.a_{KST}(\xi)=\frac{iW(\phi_{KST},\psi^{-}_{KST})}{2}.

By using, for instance, Faa di Bruno’s formula and the symbol-type estimates on σ\sigma, we obtain that, for all α0\alpha\geq 0, and rξ1r\sqrt{\xi}\sim 1,

|ξαϕKST|Cαr5/2ξαr2,|ξαψKST(r,ξ)|Cαξα+14,|ξαrϕKST|Cαr3/2r2ξα,|ξαrψKST|Cαξα14|\partial_{\xi}^{\alpha}\phi_{KST}|\leq\frac{C_{\alpha}r^{5/2}}{\xi^{\alpha}\langle r^{2}\rangle},\quad|\partial_{\xi}^{\alpha}\psi^{-}_{KST}(r,\xi)|\leq\frac{C_{\alpha}}{\xi^{\alpha+\frac{1}{4}}},\quad|\partial_{\xi}^{\alpha}\partial_{r}\phi_{KST}|\leq\frac{C_{\alpha}r^{3/2}}{\langle r^{2}\rangle\xi^{\alpha}},\quad|\partial_{\xi}^{\alpha}\partial_{r}\psi^{-}_{KST}|\leq\frac{C_{\alpha}}{\xi^{\alpha-\frac{1}{4}}}

and these give, for all α0\alpha\geq 0

|aKST(α)(ξ)|Cαξα+11ξCαξα{1,ξ11ξ,ξ1.|a_{KST}^{(\alpha)}(\xi)|\leq\frac{C_{\alpha}}{\xi^{\alpha+1}\langle\frac{1}{\xi}\rangle}\leq\frac{C_{\alpha}}{\xi^{\alpha}}\begin{cases}1,\quad\xi\lesssim 1\\ \frac{1}{\xi},\quad\xi\gtrsim 1\end{cases}.

On the other hand, from (5.7), we have

ϕKST(r,ξ)=r5/21+r4+r2ξ2rϕ1(r2)+j=2(r2ξ)j2rϕj(r2).\phi_{KST}(r,\xi)=\frac{r^{5/2}}{1+r^{4}}+\frac{r^{2}\xi}{2}\sqrt{r}\phi_{1}(r^{2})+\sum_{j=2}^{\infty}\frac{(r^{2}\xi)^{j}}{2}\sqrt{r}\phi_{j}(r^{2}).

Therefore,

(5.61) rϕKST(r,ξ)=r3/2(53r4)2(1+r4)2+54r3/2ξϕ1(r2)+r7/2ξϕ1(r2)+j=2(2j+12)r2j122ξjϕj(r2)+j=2r2j+12ξjrϕj(r2).\begin{split}\partial_{r}\phi_{KST}(r,\xi)&=\frac{r^{3/2}(5-3r^{4})}{2(1+r^{4})^{2}}+\frac{5}{4}r^{3/2}\xi\phi_{1}(r^{2})+r^{7/2}\xi\phi_{1}^{\prime}(r^{2})\\ &+\sum_{j=2}^{\infty}\frac{(2j+\frac{1}{2})r^{2j-\frac{1}{2}}}{2}\xi^{j}\phi_{j}(r^{2})+\sum_{j=2}^{\infty}r^{2j+\frac{1}{2}}\xi^{j}r\phi_{j}^{\prime}(r^{2}).\end{split}

Now, we use the same argument as the proof of Lemma 4.7 of [22]. In particular, using the formula for ϕ1\phi_{1}, namely (5.37), we get, for some small, fixed δ>0\delta>0,

|1ϕKST(δξ,ξ)|{Cξ3/4,ξ1Cξ1/4,ξ1.|\partial_{1}\phi_{KST}(\frac{\delta}{\sqrt{\xi}},\xi)|\geq\begin{cases}\frac{C}{\xi^{3/4}},\quad\xi\gtrsim 1\\ C\xi^{1/4},\quad\xi\lesssim 1\end{cases}.

Although (5.18) was stated for y>1y>1, we still have that

|1ψKST(δξ,ξ)|Cξ1/4.|\partial_{1}\psi^{-}_{KST}(\frac{\delta}{\sqrt{\xi}},\xi)|\leq C\xi^{1/4}.

Then,

(5.62) |aKST(ξ)|C|rϕKST||rψKST+|C{1ξ,ξ11,ξ1.\begin{split}|a_{KST}(\xi)|\geq C\frac{|\partial_{r}\phi_{KST}|}{|\partial_{r}\psi^{+}_{KST}|}\geq C\begin{cases}\frac{1}{\xi},\quad\xi\gtrsim 1\\ 1,\quad\xi\lesssim 1\end{cases}.\end{split}

In particular,

(5.63) ξα|aKST(α)(ξ)|Cα|aKST(ξ)|,α0.\xi^{\alpha}|a_{KST}^{(\alpha)}(\xi)|\leq C_{\alpha}|a_{KST}(\xi)|,\quad\alpha\geq 0.

As in Theorem 4.3 of [22], the density of the continuous part of the spectral measure is given by

ρ(ξ)=1πIm(m(ξ))|ξ>0=1πIm(W(θ,ψ+)W(ψ+,ϕ)),ξ>0.\rho(\xi)=\frac{1}{\pi}Im(m(\xi))\Bigr{|}_{\xi>0}=\frac{1}{\pi}Im\left(\frac{W(\theta,\psi^{+})}{W(\psi^{+},\phi)}\right),\quad\xi>0.

As in the proof of Proposition 5.7 of [21], we get

(5.64) ρ(ξ)=14π|aKST(ξ)|2{ξ2,1ξ1,ξ1.\rho(\xi)=\frac{1}{4\pi|a_{KST}(\xi)|^{2}}\sim\begin{cases}\xi^{2},\quad 1\lesssim\xi\\ 1,\quad\xi\lesssim 1\end{cases}.

Also, we have

(5.65) q(u)=uρ(u2)2q(u)=\sqrt{\frac{u\rho(u^{2})}{2}}

and thus

(5.66) q(ξ){ξ5/2,1ξξ,ξ1.q(\xi)\sim\begin{cases}\xi^{5/2},\quad 1\lesssim\xi\\ \sqrt{\xi},\quad\xi\lesssim 1\end{cases}.

Since |aKST(ξ)|2>0|a_{KST}(\xi)|^{2}>0 (which, for instance follows from the above lower bounds on |aKST||a_{KST}|), we get that, for all α0\alpha\geq 0,

(5.67) |ξα(1|aKST(ξ)|)|=|ξα(1|aKST(ξ)|2)|Cαξα{ξ,1ξ1,ξ1.|\partial_{\xi}^{\alpha}\left(\frac{1}{|a_{KST}(\xi)|}\right)|=|\partial_{\xi}^{\alpha}\left(\frac{1}{\sqrt{|a_{KST}(\xi)|^{2}}}\right)|\leq\frac{C_{\alpha}}{\xi^{\alpha}}\begin{cases}\xi,\quad 1\lesssim\xi\\ 1,\quad\xi\lesssim 1.\end{cases}

Using Faa di Bruno’s formula, we obtain that

|ξαq(ξ)|Cαξαq(ξ),α0.|\partial_{\xi}^{\alpha}q(\xi)|\leq\frac{C_{\alpha}}{\xi^{\alpha}}q(\xi),\quad\alpha\geq 0.

Then we have the representation

(5.68) ϕξ(r)=a(ξ)ϕξ+(r)+a(ξ)ϕξ+(r)¯.\phi_{\xi}(r)=a(\xi)\phi^{+}_{\xi}(r)+\overline{a(\xi)\phi^{+}_{\xi}(r)}.

where the complex valued function aa satisfies

(5.69) a(ξ)=2q(ξ)aKST(ξ2)ξ.a(\xi)=\frac{2q(\xi)a_{KST}(\xi^{2})}{\sqrt{\xi}}.

Using (5.65) and (5.64), we get that

a(ξ)=aKST(ξ2)2π|aKST(ξ2)|.a(\xi)=\frac{a_{KST}(\xi^{2})}{\sqrt{2\pi}|a_{KST}(\xi^{2})|}.

Then, using (5.63) and (5.67), we get that

(5.70) |a(ξ)|=12π,|ξα(a(ξ))|Cαξα,α0.|a(\xi)|=\frac{1}{\sqrt{2\pi}},\qquad|\partial_{\xi}^{\alpha}(a(\xi))|\leq\frac{C_{\alpha}}{\xi^{\alpha}},\quad\alpha\geq 0.

5.2. The generalized eigenfunctions of H~\tilde{H}: Proof of Theorem 5.2

We begin with the region rξ1r\xi\lesssim 1. Here, we use (5.21) to get

ψξ=ξq(ξ)(ψ0(r)+rj1(rξ)2jψj(r2))\psi_{\xi}=\xi q(\xi)\left(\psi_{0}(r)+r\sum_{j\geq 1}(r\xi)^{2j}{\psi}_{j}(r^{2})\right)

where

ψj(r)=(2h3(r)+2+2j)ϕj+1(r)+2rrϕj+1(r).{\psi}_{j}(r)=(2h_{3}(\sqrt{r})+2+2j)\phi_{j+1}(r)+2r\partial_{r}\phi_{j+1}(r).

Next,

(5.71) |(rr)αψj|αCj(j1)!r2r2|(r\partial_{r})^{\alpha}\psi_{j}|\lesssim_{\alpha}\frac{C^{j}}{(j-1)!}\frac{r^{2}}{\langle r\rangle^{2}}

follows from (5.9) In addition, ψ0\psi_{0} solves Lψ0=ϕ0L^{*}\psi_{0}=\phi_{0} therefore a direct computation shows that

ψ0(r)=L(r2ϕ1(r2))=r2(1+r4)arctan(r2)2r3.\psi_{0}(r)=L(r^{2}\phi_{1}(r^{2}))=\frac{r^{2}-(1+r^{4})\arctan(r^{2})}{2r^{3}}.

The estimates (5.25) follow from a direct estimation using (5.10) and (5.71).
On the other hand in the region rξ1r\xi\gtrsim 1 we define

ψξ+=ξ1Lϕξ+\psi_{\xi}^{+}=\xi^{-1}L\phi_{\xi}^{+}

and therefore get (5.26). The theorem now follows from the properties of ϕξ+\phi^{+}_{\xi} from Theorem 5.1.

5.3. The transference Identity

The operators HH and H~\tilde{H} have variable coefficients so they do not admit a scaling symmetry. However, the behavior of the generalized Fourier transform under scaling is important, as the scale parameter λ\lambda varies significantly along our Schrödinger map flow. The commutator of the Fourier transform with scaling is captured by the so-called ”transference operator” 𝒦\mathcal{K}, following [22]. The aim of this section will be to define and study the transference operator, which provides a convenient expression for (rr)\mathcal{F}(r\partial_{r}) (see (5.72)) called the ”transference identity”.

We use the same notation as in [22], regarding the elements in the range of \mathcal{F} as a two-component vector:

Definition 5.6.

The transference operator 𝒦\mathcal{K} is defined (a priori, for uCc((0,))u\in C^{\infty}_{c}((0,\infty))) by

(5.72) (rru)(ξ)=[100ξξ32](u)+𝒦((u))(ξ).\mathcal{F}(r\partial_{r}u)(\xi)=\begin{bmatrix}-1&0\\ 0&-\xi\partial_{\xi}-\frac{3}{2}\end{bmatrix}\mathcal{F}(u)+\mathcal{K}(\mathcal{F}(u))(\xi).

We remark that for the standard Fourier transform applied to radial functions in 2\mathbb{R}^{2} one has 𝒦=32I\mathcal{K}=-\frac{3}{2}I. Precisely, the 3/23/2 factor arises due to the different measures used in the physical space (rdrrdr) and in the Fourier space (dξd\xi).

Since the generalized Fourier transform is an isometry, it is easily verified that 𝒦\mathcal{K} is a skew-adjoint operator. Hence, in light of (5.4), we can regard 𝒦\mathcal{K} as a matrix:

𝒦=(0𝒦12𝒦21𝒦22)\mathcal{K}=\left(\begin{array}[]{cc}0&\mathcal{K}_{12}\\ \mathcal{K}_{21}&\mathcal{K}_{22}\\ \end{array}\right)

where the entries satisfy

𝒦12=𝒦21.𝒦22=𝒦22\mathcal{K}_{12}^{*}=-\mathcal{K}_{21}.\qquad\mathcal{K}_{22}^{*}=-\mathcal{K}_{22}

and can be interpreted as integral operators with kernels as follows:

Proposition 5.7.

a) The entries 𝒦ij\mathcal{K}_{ij} of the transference operator are integral operators of the form

(5.73) 𝒦12f=0K(ξ)f(ξ)𝑑ξ,\displaystyle\mathcal{K}_{12}f=-\int_{0}^{\infty}K(\xi)f(\xi)\,d\xi,
𝒦21(ξ)=K(ξ),\displaystyle\mathcal{K}_{21}(\xi)=K(\xi),
𝒦22f(ξ)=0K22(ξ,η)f(η)𝑑η,\displaystyle\mathcal{K}_{22}f(\xi)=\int_{0}^{\infty}K_{22}(\xi,\eta)f(\eta)\,d\eta,

where the kernels KK and K22K_{22} are given by

(5.74) K(ξ)=ϕ0L2(rdr)10r2ϕ0(r)ϕξ(r)𝑑r,K(\xi)=\|\phi_{0}\|_{L^{2}(rdr)}^{-1}\int_{0}^{\infty}r^{2}\phi_{0}^{\prime}(r)\phi_{\xi}(r)dr,
(5.75) K22(ξ,η)=p.v.F(η,ξ)ξ2η2,F(ξ,η)=0ϕη(r)(128r2(r41)(r4+1)3)ϕξ(r)r𝑑r.K_{22}(\xi,\eta)=\text{p.v.}\frac{F(\eta,\xi)}{\xi^{2}-\eta^{2}},\qquad F(\xi,\eta)=\int_{0}^{\infty}\phi_{\eta}(r)\left(\frac{-128r^{2}(r^{4}-1)}{(r^{4}+1)^{3}}\right)\phi_{\xi}(r)rdr.

b) These kernels satisfy the bounds

(5.76) |K(ξ)|{ξ,ξ1ξ2,ξ1,|K(\xi)|\lesssim\begin{cases}\sqrt{\xi},\quad\xi\leq 1\\ \xi^{-2},\quad\xi\geq 1\end{cases},

respectively, with ξ^=min{ξ,1}\hat{\xi}=\min\{\xi,1\},

(5.77) |F(ξ,η)|(ξ^2+η^2)ξ^12η^12|F(\xi,\eta)|\lesssim(\hat{\xi}^{2}+\hat{\eta}^{2})\hat{\xi}^{\frac{1}{2}}\hat{\eta}^{\frac{1}{2}}

and

(5.78) ξ^|ξF(ξ,η)|(ξ^2+η^2)ξ^12η^12\hat{\xi}|\partial_{\xi}F(\xi,\eta)|\lesssim(\hat{\xi}^{2}+\hat{\eta}^{2})\hat{\xi}^{\frac{1}{2}}\hat{\eta}^{\frac{1}{2}}

c) In particular. we have the L2L^{2} bound

(5.79) 𝒦L2,0L2,01.\|\mathcal{K}\|_{L^{2,0}\to L^{2,0}}\lesssim 1.

We remark that the bounds (5.77) and (5.78) above are not sharp as written. One could gain off-diagonal decay, i.e. ξη1\langle\xi-\eta\rangle^{-1} factors. We do not pursue this here because it is not needed, but we will obtain such an improvement later for the operator 𝒦~\tilde{\mathcal{K}} associated to H~\tilde{H}.

Proof.

We start by describing each 𝒦jk\mathcal{K}_{jk}. For 𝒦12\mathcal{K}_{12} we have

𝒦12f=0rϕ0(r)ϕ0L2(rdr)rr(0ϕξ(r)f(ξ)𝑑ξ)dr=20ϕ0(r)ϕ0L2(rdr)0ϕξ(r)f(ξ)𝑑ξr𝑑r00ϕξ(r)f(ξ)ϕ0(r)ϕ0L2(rdr)r2𝑑ξ𝑑r=00ϕξ(r)f(ξ)ϕ0(r)ϕ0L2(rdr)r2𝑑ξ𝑑r.\begin{split}\mathcal{K}_{12}f&=\int_{0}^{\infty}r\frac{\phi_{0}(r)}{\|\phi_{0}\|_{L^{2}(rdr)}}r\partial_{r}\left(\int_{0}^{\infty}\phi_{\xi}(r)f(\xi)d\xi\right)dr\\ &=-2\int_{0}^{\infty}\frac{\phi_{0}(r)}{\|\phi_{0}\|_{L^{2}(rdr)}}\int_{0}^{\infty}\phi_{\xi}(r)f(\xi)d\xi\,rdr-\int_{0}^{\infty}\int_{0}^{\infty}\phi_{\xi}(r)f(\xi)\frac{\phi_{0}^{\prime}(r)}{\|\phi_{0}\|_{L^{2}(rdr)}}r^{2}d\xi dr\\ &=-\int_{0}^{\infty}\int_{0}^{\infty}\phi_{\xi}(r)f(\xi)\frac{\phi_{0}^{\prime}(r)}{\|\phi_{0}\|_{L^{2}(rdr)}}r^{2}d\xi dr.\end{split}

We obtain

𝒦12f=0K(ξ)f(ξ)𝑑ξ,\mathcal{K}_{12}f=-\int_{0}^{\infty}K(\xi)f(\xi)d\xi,

where KK is as in (5.74).

Next we prove (5.76), which in particular implies that KL2K\in L^{2}. We separate two cases. When ξ1\xi\geq 1, we write H(ϕξ)=ξ2ϕξH(\phi_{\xi})=\xi^{2}\phi_{\xi} and integrate by parts. On the other hand, when ξ1\xi\leq 1, we directly estimate KK, as follows.

|K(ξ)|01ξr𝑑r|ϕξ(r)q(ξ)ϕ0(r)||ϕ0(r)|r+01ξr𝑑rq(ξ)ϕ0(r)|ϕ0(r)|r+1ξr𝑑r|ϕξ(r)||ϕ0(r)|rξ,\begin{split}|K(\xi)|&\lesssim\int_{0}^{\frac{1}{\xi}}rdr|\phi_{\xi}(r)-q(\xi)\phi_{0}(r)||\phi_{0}^{\prime}(r)|r+\int_{0}^{\frac{1}{\xi}}rdrq(\xi)\phi_{0}(r)|\phi_{0}^{\prime}(r)|r\\ &+\int_{\frac{1}{\xi}}^{\infty}rdr|\phi_{\xi}(r)||\phi_{0}^{\prime}(r)|r\\ &\lesssim\sqrt{\xi},\end{split}

where we have used (5.19), (5.20), (5.10). Then, Cauchy-Schwartz implies the L2L^{2} bound

|𝒦12(f)|fL2.|\mathcal{K}_{12}(f)|\lesssim\|f\|_{L^{2}}.

Having 𝒦12\mathcal{K}_{12}, we also directly obtain its adjoint

𝒦21(ξ)=K(ξ).\mathcal{K}_{21}(\xi)=K(\xi).

Finally we consider 𝒦22f\mathcal{K}_{22}f, for which we compute integrating by parts

𝒦22f(ξ)=0r𝑑rϕξ(r)0rϕη(r)f(η)𝑑η+0r𝑑rϕξ(r)0ηf(η)ϕη(r)𝑑η+32f(ξ)=0r𝑑rϕξ(r)0rϕη(r)f(η)𝑑η0r𝑑rϕξ(r)0𝑑ηf(η)(ϕη(r)+ηηϕη(r))+32f(ξ)=0r𝑑rϕξ(r)0𝑑η(rϕη(r)ηηϕη(r))f(η)+12f(ξ):=𝒦220f(ξ)+12f(ξ).\begin{split}&\mathcal{K}_{22}f(\xi)\\ &=\int_{0}^{\infty}rdr\phi_{\xi}(r)\int_{0}^{\infty}r\phi_{\eta}^{\prime}(r)f(\eta)d\eta+\int_{0}^{\infty}rdr\phi_{\xi}(r)\int_{0}^{\infty}\eta f^{\prime}(\eta)\phi_{\eta}(r)d\eta+\frac{3}{2}f(\xi)\\ &=\int_{0}^{\infty}rdr\phi_{\xi}(r)\int_{0}^{\infty}r\phi_{\eta}^{\prime}(r)f(\eta)d\eta-\int_{0}^{\infty}rdr\phi_{\xi}(r)\int_{0}^{\infty}d\eta f(\eta)(\phi_{\eta}(r)+\eta\partial_{\eta}\phi_{\eta}(r))+\frac{3}{2}f(\xi)\\ &=\int_{0}^{\infty}rdr\phi_{\xi}(r)\int_{0}^{\infty}d\eta(r\phi_{\eta}^{\prime}(r)-\eta\partial_{\eta}\phi_{\eta}(r))f(\eta)+\frac{1}{2}f(\xi)\\ &:=\mathcal{K}_{220}f(\xi)+\frac{1}{2}f(\xi).\end{split}

In order to capture the off-diagonal behavior of 𝒦220\mathcal{K}_{220} we commute the HH operator inside this representation, using the relation

(Hη2)ηϕη=2ηϕη.(H-\eta^{2})\partial_{\eta}\phi_{\eta}=2\eta\phi_{\eta}.

This yields

ξ2𝒦220f(ξ)=0r𝑑rH(ϕξ)(r)0𝑑η(rϕη(r)ηηϕη(r))f(η)=0𝑑rϕξ(r)r0𝑑ηH(rϕη(r)ηηϕη(r))f(η)=0𝑑rϕξ(r)r0𝑑ηη2(rrηη)(ϕη(r))f(η)+0𝑑rϕξ(r)r0𝑑η(128r2(r41)ϕη(r)(r4+1)3f(η)).\begin{split}\xi^{2}\mathcal{K}_{220}f(\xi)&=\int_{0}^{\infty}rdrH(\phi_{\xi})(r)\int_{0}^{\infty}d\eta(r\phi_{\eta}^{\prime}(r)-\eta\partial_{\eta}\phi_{\eta}(r))f(\eta)\\ &=\int_{0}^{\infty}dr\phi_{\xi}(r)r\int_{0}^{\infty}d\eta H(r\phi_{\eta}^{\prime}(r)-\eta\partial_{\eta}\phi_{\eta}(r))f(\eta)\\ &=\int_{0}^{\infty}dr\phi_{\xi}(r)r\int_{0}^{\infty}d\eta\eta^{2}(r\partial_{r}-\eta\partial_{\eta})(\phi_{\eta}(r))f(\eta)\\ &+\int_{0}^{\infty}dr\phi_{\xi}(r)r\int_{0}^{\infty}d\eta\left(\frac{-128r^{2}(r^{4}-1)\phi_{\eta}(r)}{(r^{4}+1)^{3}}f(\eta)\right).\end{split}

The kernel of the last term is a smooth function of ξ\xi and η\eta, which we denote by F(ξ,η)F(\xi,\eta) as in (5.75). Hence we have proved that

(5.80) ξ2𝒦22f(ξ)𝒦22(()2f)(ξ)=0F(ξ,η)f(η)𝑑η.\xi^{2}\mathcal{K}_{22}f(\xi)-\mathcal{K}_{22}((\cdot)^{2}f)(\xi)=\int_{0}^{\infty}F(\xi,\eta)f(\eta)d\eta.

This shows that away from the diagonal, the kernel of 𝒦22\mathcal{K}_{22} is given by

(5.81) K22(ξ,η)=F(ξ,η)ξ2η2.K_{22}(\xi,\eta)=\frac{F(\xi,\eta)}{\xi^{2}-\eta^{2}}.

It remains to determine the diagonal behavior of 𝒦22\mathcal{K}_{22}. In the region 1rξ1\lesssim r\xi we have

ϕξ(r)=2Re(a(ξ)reirξ(1+15i8rξ))+O(1r11/2ξ)+O(1r5/2ξ2),\phi_{\xi}(r)=2\text{Re}\left(\frac{a(\xi)}{\sqrt{r}}e^{ir\xi}(1+\frac{15i}{8r\xi})\right)+O\left(\frac{1}{r^{11/2}\xi}\right)+O\left(\frac{1}{r^{5/2}\xi^{2}}\right),

which yields the asymptotic

(rrξξ)ϕξ(r)=12ϕξ(r)+O(1r11/2ξ)+O(1r5/2ξ2)2Re(ξa(ξ)eirξr(1+15i8rξ)),\left(r\partial_{r}-\xi\partial_{\xi}\right)\phi_{\xi}(r)=-\frac{1}{2}\phi_{\xi}(r)+O\left(\frac{1}{r^{11/2}\xi}\right)+O\left(\frac{1}{r^{5/2}\xi^{2}}\right)-2\text{Re}\left(\frac{\xi a^{\prime}(\xi)e^{ir\xi}}{\sqrt{r}}(1+\frac{15i}{8r\xi})\right),

where the decay of the tail limits the singular behavior on the diagonal. This is similar to the analysis in [22], so we do not repeat it here. Using also the antisymmetry of 𝒦22\mathcal{K}_{22}, we obtain that its kernel is given by the principal value expression in (5.81), which concludes the proof of (5.75).


To conclude the proof of the proposition we need to study the kernel

F(ξ,η)=0(128r2(r41)(r4+1)3)ϕξ(r)ϕη(r)r𝑑r.F(\xi,\eta)=\int_{0}^{\infty}\left(\frac{-128r^{2}(r^{4}-1)}{(r^{4}+1)^{3}}\right)\phi_{\xi}(r)\phi_{\eta}(r)rdr.

For this, we recall (5.19), (5.20), and note that

|ξϕξ(r)q(ξ)ϕ0(r)|q(ξ)r4ξr2,rξ1|\partial_{\xi}\phi_{\xi}(r)-q^{\prime}(\xi)\phi_{0}(r)|\lesssim\frac{q(\xi)r^{4}\xi}{\langle r^{2}\rangle},\quad r\xi\lesssim 1
|ξϕξ(r)|r,1rξ.|\partial_{\xi}\phi_{\xi}(r)|\lesssim\sqrt{r},\quad 1\lesssim r\xi.

In addition, we observe the cancellation

0ϕ0(r)(128r2(r41)(r4+1)3)ϕ0(r)r𝑑r=0.\int_{0}^{\infty}\phi_{0}(r)\left(\frac{-128r^{2}(r^{4}-1)}{(r^{4}+1)^{3}}\right)\phi_{0}(r)\,rdr=0.

This can be seen by direct computation, or one could note that

0ϕ0(r)(128r2(r41)(r4+1)3)ϕ0(r)r𝑑r=0ϕ0(r)([H,rr]ϕ02H(ϕ0))r𝑑r\int_{0}^{\infty}\phi_{0}(r)\left(\frac{-128r^{2}(r^{4}-1)}{(r^{4}+1)^{3}}\right)\phi_{0}(r)\,rdr=\int_{0}^{\infty}\phi_{0}(r)\left([H,r\partial_{r}]\phi_{0}-2H(\phi_{0})\right)\,rdr

and use H(ϕ0)=0H(\phi_{0})=0 and the fact that HH is self-adjoint on L2(rdr)L^{2}(rdr). Combining these properties, it is a direct computation to prove the bounds (5.77) and (5.78).

It remains to prove the L2L^{2} bound for the operator 𝒦22\mathcal{K}_{22}, which is expressed in terms of FF as in (5.75). For this we decompose the kernel into a leading near diagonal part, and two milder terms

p.v.FH(ξ,η)η2ξ2=p.v.𝟙|ξη|1ηξFH(η,η)2η+𝟙|ξη|1ηξ(FH(ξ,η)η+ξFH(η,η)2η)+𝟙|ξη|>1FH(ξ,η)(ηξ)(η+ξ)=TFH(η,η)2η+K221+K222\begin{split}\text{p.v.}\frac{F_{H}(\xi,\eta)}{\eta^{2}-\xi^{2}}=&\ \text{p.v.}\frac{\mathbbm{1}_{|\xi-\eta|\leq 1}}{\eta-\xi}\cdot\frac{F_{H}(\eta,\eta)}{2\eta}+\frac{\mathbbm{1}_{|\xi-\eta|\leq 1}}{\eta-\xi}\left(\frac{F_{H}(\xi,\eta)}{\eta+\xi}-\frac{F_{H}(\eta,\eta)}{2\eta}\right)+\frac{\mathbbm{1}_{|\xi-\eta|>1}F_{H}(\xi,\eta)}{(\eta-\xi)(\eta+\xi)}\\ =&\ T\frac{F_{H}(\eta,\eta)}{2\eta}+K_{22}^{1}+K_{22}^{2}\end{split}

where we have factored the first term on the right into a multiplication with a bounded function and a frequency localized Hilbert transform,

(5.82) T(g)(η)=p.v.𝟙|ξη|1ηξg(ξ)𝑑ξ.T(g)(\eta)=\text{p.v.}\int_{\mathbb{R}}\frac{\mathbbm{1}_{|\xi-\eta|\leq 1}}{\eta-\xi}g(\xi)d\xi.

This is a Fourier multiplier with bounded symbol, so is bounded on L2()L^{2}(\mathbb{R}). The remaining terms are nonsingular, and we can simply estimate their kernels in L2L^{2}. This yields the L2L^{2} bound

(5.83) 𝒦22fL2fL2(dξ)(FH(η,η)2ηLη+K221L2+K222L2).\begin{split}\|\mathcal{K}_{22}f\|_{L^{2}}&\lesssim\|f\|_{L^{2}(d\xi)}\left(\left\|\frac{F_{H}(\eta,\eta)}{2\eta}\right\|_{L^{\infty}_{\eta}}+\|K_{22}^{1}\|_{L^{2}}+\|K_{22}^{2}\|_{L^{2}}\right).\end{split}

It remains to bound the norms in the last sum.

Using (5.77), we directly see that FH(η,η)2ηLη\dfrac{F_{H}(\eta,\eta)}{2\eta}\in L^{\infty}_{\eta}.

Next, by the fundamental theorem of calculus, we get

(5.84) |FH(ξ,η)(η+ξ)FH(η,η)2η|min{η,ξ}max{η,ξ}(|1FH(ω,η)|(η+ω)+|FH(ω,η)|(η+ω)2)𝑑ω\begin{split}&\left|\frac{F_{H}(\xi,\eta)}{(\eta+\xi)}-\frac{F_{H}(\eta,\eta)}{2\eta}\right|\lesssim\int_{\min\{\eta,\xi\}}^{\max\{\eta,\xi\}}\left(\frac{|\partial_{1}F_{H}(\omega,\eta)|}{(\eta+\omega)}+\frac{|F_{H}(\omega,\eta)|}{(\eta+\omega)^{2}}\right)d\omega\end{split}

Using (5.77) and (5.78) we can bound this by

(5.85) |FH(ξ,η)(η+ξ)FH(η,η)2η||ξη|ξ^2+η^2ξ+η,\begin{split}&\left|\frac{F_{H}(\xi,\eta)}{(\eta+\xi)}-\frac{F_{H}(\eta,\eta)}{2\eta}\right|\lesssim|\xi-\eta|\frac{\hat{\xi}^{2}+\hat{\eta}^{2}}{\xi+\eta},\end{split}

which shows that

K221(ξ,η)=1ηξ𝟙|ξη|1(FH(ξ,η)(η+ξ)FH(η,η)2η)L2(dξ)L2(dη).K_{22}^{1}(\xi,\eta)=\frac{1}{\eta-\xi}\mathbbm{1}_{|\xi-\eta|\leq 1}\left(\frac{F_{H}(\xi,\eta)}{(\eta+\xi)}-\frac{F_{H}(\eta,\eta)}{2\eta}\right)\in L^{2}(d\xi)L^{2}(d\eta).

Finally, (5.77) gives that |FH(ξ,η)|1|F_{H}(\xi,\eta)|\lesssim 1 in the region ξ+η1\xi+\eta\gtrsim 1. This suffices in order to show that

K222(ξ,η)=𝟙|ξη|1(ηξ)FH(ξ,η)(η+ξ)L2(dξ)L2(dη).K_{22}^{2}(\xi,\eta)=\frac{\mathbbm{1}_{|\xi-\eta|\geq 1}}{(\eta-\xi)}\frac{F_{H}(\xi,\eta)}{(\eta+\xi)}\in L^{2}(d\xi)L^{2}(d\eta).

In light of (5.83), this completes the proof of the L2L^{2} boundedness of 𝒦22\mathcal{K}_{22}, and thus the proof of the proposition.

In this article we will also use even more the transference operator associated the the H~\widetilde{H} operator, which we will denote by 𝒦~\tilde{\mathcal{K}}. Compared with 𝒦\mathcal{K} this is simpler as we no longer need to separate the zero mode:

Definition 5.8.

The transference operator 𝒦~\tilde{\mathcal{K}} associated to H~\tilde{H} is defined (a priori, for uCc((0,))u\in C^{\infty}_{c}((0,\infty))) by

(5.86) ~(rru)(ξ)=(ξξ32)~(u)+𝒦~(~(u))(ξ),{\tilde{\mathcal{F}}}(r\partial_{r}u)(\xi)=(-\xi\partial_{\xi}-\frac{3}{2}){\tilde{\mathcal{F}}}(u)+\tilde{\mathcal{K}}({\tilde{\mathcal{F}}}(u))(\xi),

Here the 32\frac{3}{2} correction is again chosen in order to ensure that 𝒦~\tilde{\mathcal{K}} is antisymmetric.

One may think of 𝒦~\tilde{\mathcal{K}} as a scaling derivative of the Fourier transform ~λ{\tilde{\mathcal{F}}}_{\lambda}, in the sense that

(5.87) ddλ~λ|λ=1=𝒦~~.{\frac{d}{d\lambda}{{\tilde{\mathcal{F}}}_{\lambda}}}_{|\lambda=1}=-\tilde{\mathcal{K}}{\tilde{\mathcal{F}}}.

With this interpretation the antisymmetry requirement is clear, since the Fourier transforms are isometries. This is immediately seen by differentiating with respect to λ\lambda in (5.29), and applying the chain rule. For the inverse Fourier transform this directly gives

(5.88) ddλ~λ1|λ=1=~1𝒦~.{\frac{d}{d\lambda}{{\tilde{\mathcal{F}}}_{\lambda}^{-1}}}_{|\lambda=1}={\tilde{\mathcal{F}}}^{-1}\tilde{\mathcal{K}}.

We summarize the properties of this operator in the next proposition, which is the direct counterpart of Proposition 5.7:

Proposition 5.9.

a) The transference operator 𝒦~\tilde{\mathcal{K}} is an integral operator of the form

(5.89) 𝒦~f(ξ)=0K~(ξ,η)f(η)𝑑η,\displaystyle\tilde{\mathcal{K}}f(\xi)=\int_{0}^{\infty}\tilde{K}(\xi,\eta)f(\eta)\,d\eta,

where the kernel K~\tilde{K} is given by

(5.90) K~(ξ,η)=p.v.F~(η,ξ)ξ2η2,F~(ξ,η)=0ψη(r)ψξ(r)32r2(1+r4)2r𝑑r.\tilde{K}(\xi,\eta)=\text{p.v.}\,\frac{\tilde{F}(\eta,\xi)}{\xi^{2}-\eta^{2}},\qquad\tilde{F}(\xi,\eta)=-\int_{0}^{\infty}\frac{\psi_{\eta}(r)\psi_{\xi}(r)32r^{2}}{(1+r^{4})^{2}}rdr.

b) This kernel satisfies the bounds

(5.91) |F~(ξ,η)|ξ^32η^32min{ξ,η}ξ+ηξη,|\tilde{F}(\xi,\eta)|\lesssim\hat{\xi}^{\frac{3}{2}}\hat{\eta}^{\frac{3}{2}}\frac{\min\{\langle\xi\rangle,\langle\eta\rangle\}}{\langle\xi+\eta\rangle\langle\xi-\eta\rangle},

and

(5.92) ξ^|ξF~(ξ,η)|ξ^32η^32.\hat{\xi}|\partial_{\xi}\tilde{F}(\xi,\eta)|\lesssim\hat{\xi}^{\frac{3}{2}}\hat{\eta}^{\frac{3}{2}}.

c) In particular. we have the L2L^{2} bound

(5.93) 𝒦~L2L21.\|\tilde{\mathcal{K}}\|_{L^{2}\to L^{2}}\lesssim 1.
Proof.

a) From the definition of 𝒦~\tilde{\mathcal{K}} we have

𝒦~f(ξ)=~(rr~1(f))(ξ)+ξf(ξ)+32f(ξ),fCc((0,)).\tilde{\mathcal{K}}f(\xi)={\tilde{\mathcal{F}}}(r\partial_{r}{\tilde{\mathcal{F}}}^{-1}(f))(\xi)+\xi f^{\prime}(\xi)+\frac{3}{2}f(\xi),\quad f\in C^{\infty}_{c}((0,\infty)).

Proceeding as for 𝒦\mathcal{K}, we rewrite this in the form

𝒦~f(ξ)=0r𝑑rψξ(r)0f(η)(rrψη(r)ηηψη(r))𝑑η+12f(ξ).\tilde{\mathcal{K}}f(\xi)=\int_{0}^{\infty}rdr\psi_{\xi}(r)\int_{0}^{\infty}f(\eta)\left(r\partial_{r}\psi_{\eta}(r)-\eta\partial_{\eta}\psi_{\eta}(r)\right)\,d\eta+\frac{1}{2}f(\xi).

Commuting the operator H~\tilde{H} inside this representation we arrive at

ξ2𝒦~(f)(ξ)𝒦~(()2f())(η)=F~(ξ,η)f(η)𝑑η,\xi^{2}\tilde{\mathcal{K}}(f)(\xi)-\tilde{\mathcal{K}}((\cdot)^{2}f(\cdot))(\eta)=\int\tilde{F}(\xi,\eta)f(\eta)d\eta,

where the smooth symmetric function F~\tilde{F} is given by

(5.94) F~(ξ,η)=0ψη(r)ψξ(r)32r2(1+r4)2r𝑑r.\tilde{F}(\xi,\eta)=-\int_{0}^{\infty}\frac{\psi_{\eta}(r)\psi_{\xi}(r)32r^{2}}{(1+r^{4})^{2}}rdr.

This yields the off-diagonal behavior for the kernel K~\tilde{K} as in (5.90). For its diagonal behavior, we start with

ψξ(r)=2Re(a(ξ)reirξσ~(rξ,r)),rξC\psi_{\xi}(r)=2\text{Re}\left(\frac{a(\xi)}{\sqrt{r}}e^{ir\xi}\widetilde{\sigma}(r\xi,r)\right),\quad r\xi\geq C

and

(rrξξ)(ψξ(r))=12ψξ(r)+O(1r3/2ξ)2Re(ξra(ξ)eirξσ~(rξ,r))(r\partial_{r}-\xi\partial_{\xi})(\psi_{\xi}(r))=-\frac{1}{2}\psi_{\xi}(r)+O\left(\frac{1}{r^{3/2}\xi}\right)-2\text{Re}\left(\frac{\xi}{\sqrt{r}}a^{\prime}(\xi)e^{ir\xi}\widetilde{\sigma}(r\xi,r)\right)

where, for fixed ξ\xi,

σ~(rξ,r)=i38r1ξ1+O(r2).\tilde{\sigma}(r\xi,r)=i-\frac{3}{8}r^{-1}\xi^{-1}+O(r^{-2}).

Arguing as in [22] and using the antisymmetry, we obtain the p.v. diagonal kernel behavior as in (5.90).

b) The estimates on ψξ\psi_{\xi} from Theorem 5.2 directly give

(5.95) |F~(ξ,η)|ξ^3/2η^3/2|\tilde{F}(\xi,\eta)|\lesssim\hat{\xi}^{3/2}\hat{\eta}^{3/2}

and

(5.96) ξ^|ξF~(ξ,η)|ξ^3/2η^3/2\hat{\xi}|\partial_{\xi}\tilde{F}(\xi,\eta)|\lesssim\hat{\xi}^{3/2}\hat{\eta}^{3/2}

Here (5.96) is what we need, but (5.95) only suffices in the region |ξη|1|\xi-\eta|\lesssim 1. To prove the better bound (5.91) we need an off-diagonal refinement of (5.95) when ξ,η1\xi,\eta\gtrsim 1. For this, we recall that

F~(ξ,η)=320r2(1+r4)2ψξ(r)ψη(r)r𝑑r.\tilde{F}(\xi,\eta)=-32\int_{0}^{\infty}\frac{r^{2}}{(1+r^{4})^{2}}\psi_{\xi}(r)\psi_{\eta}(r)\,rdr.

Then,

ξ2F~(ξ,η)=320r3dr(1+r4)2H~(ψξ)(r)ψη(r)=32r2(1+r4)2η2ψη+[H~,r2(1+r4)2]ψη,ψξL2(rdr)=F~(ξ,η)η2+320ψξ(r)[H~,r2(1+r4)2]ψηr𝑑r.\begin{split}\xi^{2}\tilde{F}(\xi,\eta)&=-32\int_{0}^{\infty}\frac{r^{3}dr}{(1+r^{4})^{2}}\widetilde{H}(\psi_{\xi})(r)\psi_{\eta}(r)\\ &=-32\langle\frac{r^{2}}{(1+r^{4})^{2}}\eta^{2}\psi_{\eta}+[\widetilde{H},\frac{r^{2}}{(1+r^{4})^{2}}]\psi_{\eta},\psi_{\xi}\rangle_{L^{2}(rdr)}\\ &=\tilde{F}(\xi,\eta)\eta^{2}+32\int_{0}^{\infty}\psi_{\xi}(r)[\widetilde{H},\frac{r^{2}}{(1+r^{4})^{2}}]\psi_{\eta}\,rdr.\end{split}

Computing the commutator, for ξη\xi\neq\eta we obtain

F~(ξ,η)=32(ξ2η2)0ψξ(r)4r(1+r4)4((114r4+9r8)ψη(r)+r(12r43r8)ψη(r))𝑑r.\tilde{F}(\xi,\eta)=\frac{32}{(\xi^{2}-\eta^{2})}\int_{0}^{\infty}\frac{\psi_{\xi}(r)\cdot 4r}{(1+r^{4})^{4}}\left(\left(1-14r^{4}+9r^{8}\right)\psi_{\eta}(r)+r\left(1-2r^{4}-3r^{8}\right)\psi_{\eta}^{\prime}(r)\right)dr.

Hence, using the bounds on ψξ\psi_{\xi} from from Theorem 5.2, and the symmetry of F~\tilde{F}, we obtain

|F~(ξ,η)|min{ξ,η}|ξ2η2|,ξ,η1|\tilde{F}(\xi,\eta)|\lesssim\frac{\text{min}\{\xi,\eta\}}{|\xi^{2}-\eta^{2}|},\quad\xi,\eta\gtrsim 1
|F~(ξ,η)|Cmin{ξ,η}3/2|ξ2η2|,0<η1ξ or 0<ξ1η,|\tilde{F}(\xi,\eta)|\leq\frac{C\min\{\xi,\eta\}^{3/2}}{|\xi^{2}-\eta^{2}|},\quad 0<\eta\lesssim 1\lesssim\xi\text{ or }0<\xi\lesssim 1\lesssim\eta,

which completes the proof of (5.91).

c) Now, we study the operator

fp.v.(0f(ξ)F~(ξ,η)dξξ2η2).f\mapsto\text{p.v.}\left(\int_{0}^{\infty}\frac{f(\xi)\tilde{F}(\xi,\eta)d\xi}{\xi^{2}-\eta^{2}}\right).

As for 𝒦\mathcal{K}, we start with the decomposition

(5.97) F~(ξ,η)ξ2η2=𝟙|ξη|>1F~(ξ,η)(ξη)(η+ξ)+𝟙|ξη|1(F~(ξ,η)η+ξF~(η,η)2η)ξη+𝟙|ξη|1F~(η,η)2η(ξη)\frac{\tilde{F}(\xi,\eta)}{\xi^{2}-\eta^{2}}=\frac{\mathbbm{1}_{|\xi-\eta|>1}\tilde{F}(\xi,\eta)}{(\xi-\eta)(\eta+\xi)}+\frac{\mathbbm{1}_{|\xi-\eta|\leq 1}\left(\frac{\tilde{F}(\xi,\eta)}{\eta+\xi}-\frac{\tilde{F}(\eta,\eta)}{2\eta}\right)}{\xi-\eta}+\frac{\mathbbm{1}_{|\xi-\eta|\leq 1}\tilde{F}(\eta,\eta)}{2\eta(\xi-\eta)}

and estimate the contributions of each of the three kernels separately. We use (5.95) and (5.91) to get

𝟙|ξη|>1|F~(ξ,η)|C{ξ+η,η,ξ11,otherwise.\mathbbm{1}_{|\xi-\eta|>1}|\tilde{F}(\xi,\eta)|\leq C\begin{cases}\xi+\eta,\quad\eta,\xi\lesssim 1\\ 1,\quad\text{otherwise.}\end{cases}

This implies that

F~(ξ,η)𝟙|ξη|>1(ξη)(η+ξ)L2(dξ)L2(dη),\frac{\tilde{F}(\xi,\eta)\mathbbm{1}_{|\xi-\eta|>1}}{(\xi-\eta)(\eta+\xi)}\in L^{2}(d\xi)L^{2}(d\eta),

which suffices for the first term in (5.97). Next, we note that

|F~(ξ,η)η+ξF~(η,η)2η|Cmin{η,ξ}max{η,ξ}(|1F~(ω,η)|η+ω+|F~(ω,η)|(η+ω)2)𝑑ω.\left|\frac{\tilde{F}(\xi,\eta)}{\eta+\xi}-\frac{\tilde{F}(\eta,\eta)}{2\eta}\right|\leq C\int_{\min\{\eta,\xi\}}^{\max\{\eta,\xi\}}\left(\frac{|\partial_{1}\tilde{F}(\omega,\eta)|}{\eta+\omega}+\frac{|\tilde{F}(\omega,\eta)|}{(\eta+\omega)^{2}}\right)d\omega.

From (5.95),

|F~(ω,η)|(η+ω)2C(𝟙{4η}+1η𝟙{η4}).\frac{|\tilde{F}(\omega,\eta)|}{(\eta+\omega)^{2}}\leq C\left(\mathbbm{1}_{\{4\geq\eta\}}+\frac{1}{\eta}\mathbbm{1}_{\{\eta\geq 4\}}\right).

Also, from (5.96), we get

|1F~(ω,η)|η+ωC(𝟙{4η}+1η𝟙{η4}).\frac{|\partial_{1}\tilde{F}(\omega,\eta)|}{\eta+\omega}\leq C\left(\mathbbm{1}_{\{4\geq\eta\}}+\frac{1}{\eta}\mathbbm{1}_{\{\eta\geq 4\}}\right).

Therefore,

|F~(ξ,η)η+ξF~(η,η)2η|C|ηξ|(𝟙{4η}+1η𝟙{η4})\left|\frac{\tilde{F}(\xi,\eta)}{\eta+\xi}-\frac{\tilde{F}(\eta,\eta)}{2\eta}\right|\leq C|\eta-\xi|\left(\mathbbm{1}_{\{4\geq\eta\}}+\frac{1}{\eta}\mathbbm{1}_{\{\eta\geq 4\}}\right)

which gives

𝟙|ξη|1|ξη|(F~(ξ,η)(η+ξ)F~(η,η)2η)L2(dξ)L2(dη),\frac{\mathbbm{1}_{|\xi-\eta|\leq 1}}{|\xi-\eta|}\left(\frac{\tilde{F}(\xi,\eta)}{(\eta+\xi)}-\frac{\tilde{F}(\eta,\eta)}{2\eta}\right)\in L^{2}(d\xi)L^{2}(d\eta),

as needed for the second term in (5.97).

Next, as in (5.83), we have

F~(η,η)2ηp.v.0dξ𝟙|ξη|1F~(η,η)ξη𝑑ξL2(dη)CT(L2(),L2())fL2(dξ)F~(η,η)2ηLη\begin{split}&\left\|\frac{\tilde{F}(\eta,\eta)}{2\eta}\text{p.v.}\int_{0}^{\infty}\frac{d\xi\mathbbm{1}_{|\xi-\eta|\leq 1}\tilde{F}(\eta,\eta)}{\xi-\eta}d\xi\right\|_{L^{2}(d\eta)}\leq C\|T\|_{\mathcal{L}(L^{2}(\mathbb{R}),L^{2}(\mathbb{R}))}\|f\|_{L^{2}(d\xi)}\|\frac{\tilde{F}(\eta,\eta)}{2\eta}\|_{L^{\infty}_{\eta}}\end{split}

where we recall the definition of TT in (5.82). Finally, (5.95) shows that

F~(η,η)ηLηC\|\frac{\tilde{F}(\eta,\eta)}{\eta}\|_{L^{\infty}_{\eta}}\leq C

which suffices for the third term in (5.97), and finally shows that the operator 𝒦~\tilde{\mathcal{K}} is bounded on L2((0,))L^{2}((0,\infty)). ∎

The transference operators 𝒦\mathcal{K} and 𝒦~\tilde{\mathcal{K}} were defined above in the context of the operators HH and H~\tilde{H}. However, we will also need them in the rescaled setting, associated to HλH_{\lambda} and H~λ\tilde{H}_{\lambda}. Their rescaled versions are denoted by 𝒦λ\mathcal{K}_{\lambda} and 𝒦~λ\tilde{\mathcal{K}}_{\lambda}, and are still defined by (5.72), respectively (5.86), but with {\mathcal{F}} and ~{\tilde{\mathcal{F}}} replaced by λ{\mathcal{F}}_{\lambda} , respectively ~λ{\tilde{\mathcal{F}}}_{\lambda}. These are obtained from 𝒦\mathcal{K} and 𝒦~\tilde{\mathcal{K}} by rescaling, for instance

(5.98) 𝒦~λf(ξ)=1λ𝒦~(f(λ))(ξλ).\tilde{\mathcal{K}}_{\lambda}f(\xi)=\frac{1}{\lambda}\tilde{\mathcal{K}}(f(\cdot\lambda))(\frac{\xi}{\lambda}).

The representation (5.90) remains valid, but with F~\tilde{F} replaced by its rescaled version F~λ\tilde{F}_{\lambda} given by

(5.99) F~λ(ξ,η)=F~(ξλ,ηλ).\tilde{F}_{\lambda}(\xi,\eta)=\tilde{F}(\frac{\xi}{\lambda},\frac{\eta}{\lambda}).

Finally, the relations (5.87) and (5.88) also carry through, and give

(5.100) ddλ~λ=𝒦~λ~λ,ddλ~λ1=~λ1𝒦~λ.{\frac{d}{d\lambda}{{\tilde{\mathcal{F}}}_{\lambda}}}=-\tilde{\mathcal{K}}_{\lambda}{\tilde{\mathcal{F}}}_{\lambda},\qquad\frac{d}{d\lambda}{{\tilde{\mathcal{F}}}_{\lambda}^{-1}}={\tilde{\mathcal{F}}}^{-1}_{\lambda}\tilde{\mathcal{K}}_{\lambda}.

5.4. The operator r~1r{\tilde{\mathcal{F}}}^{-1}

Our goal here is to study the operator of multiplication by rr on the Fourier side. This will allow us later on to obtain more refined, spatially localized bounds for the solutions to the linear H~λ\tilde{H}_{\lambda} flow.

Lemma 5.10.

For all fCc((0,))f\in C^{\infty}_{c}((0,\infty)) we have

r~1fL2(rdr)f(ξ)ξL2(dξ)+f(ξ)L2(dξ).\|r{\tilde{\mathcal{F}}}^{-1}f\|_{L^{2}(rdr)}\lesssim\|\frac{f(\xi)}{\xi}\|_{L^{2}(d\xi)}+\|f^{\prime}(\xi)\|_{L^{2}(d\xi)}.

In particular, the above inequality is true for all ff in the closure of Cc((0,))C^{\infty}_{c}((0,\infty)) under the norm

f2:=f(ξ)L2(dξ)2+f(ξ)ξL2(dξ)2+f(ξ)L2(dξ)2\|f\|^{2}:=\|f(\xi)\|_{L^{2}(d\xi)}^{2}+\|\frac{f(\xi)}{\xi}\|_{L^{2}(d\xi)}^{2}+\|f^{\prime}(\xi)\|_{L^{2}(d\xi)}^{2}
Proof.

We start with an analog of

(5.101) rJ1(rξ)=ξJ2(rξ)+2J2(rξ)ξrJ_{1}(r\xi)=\partial_{\xi}J_{2}(r\xi)+\frac{2J_{2}(r\xi)}{\xi}

in our setting. (Recall that J1(rξ)J_{1}(r\xi) are eigenfunctions of the large rr part of H~\widetilde{H}, and J2(rξ)J_{2}(r\xi) are eigenfunctions of the large rr part of HH). More precisely, our starting point is the identity

ψξ(r)=ξ1Lϕξ,\psi_{\xi}(r)=\xi^{-1}L\phi_{\xi},

where we recall that

Lf=rf+2rh3(r)f,h3(r)=r41r4+1.Lf=\partial_{r}f+\frac{2}{r}h_{3}(r)f,\quad h_{3}(r)=\frac{r^{4}-1}{r^{4}+1}.

This allows us to obtain the identity

(5.102) rψξ(r)=1ξ(rrϕξ+2h3ϕξ)=1ξ(rrξξ)ϕξ+ξϕξ+2ϕξξ+2(h3(r)1)ξϕξ.\begin{split}r\psi_{\xi}(r)&=\frac{1}{\xi}\left(r\partial_{r}\phi_{\xi}+2h_{3}\phi_{\xi}\right)\\ &=\frac{1}{\xi}(r\partial_{r}-\xi\partial_{\xi})\phi_{\xi}+\partial_{\xi}\phi_{\xi}+\frac{2\phi_{\xi}}{\xi}+\frac{2(h_{3}(r)-1)}{\xi}\phi_{\xi}.\end{split}

This is the analog of (5.101), with the only error terms either decaying quickly at infinity (e.g. |h3(r)1|Cr4|h_{3}(r)-1|\leq\frac{C}{r^{4}}) or involving an operator which annihilates functions of the form f(rξ)f(r\xi).

Therefore, for fCc((0,))f\in C^{\infty}_{c}((0,\infty)) we have

(5.103) r~1(f)(r)=r0ψξ(r)f(ξ)𝑑ξ=0(rrξξ)(ϕξ(r))f(ξ)ξ𝑑ξ+0ξϕξ(r)f(ξ)dξ+02ϕξξf(ξ)𝑑ξ+02(h3(r)1)ξϕξ(r)f(ξ)𝑑ξ.\begin{split}r{\tilde{\mathcal{F}}}^{-1}(f)(r)&=r\int_{0}^{\infty}\psi_{\xi}(r)f(\xi)d\xi\\ &=\int_{0}^{\infty}(r\partial_{r}-\xi\partial_{\xi})(\phi_{\xi}(r))\frac{f(\xi)}{\xi}d\xi+\int_{0}^{\infty}\partial_{\xi}\phi_{\xi}(r)f(\xi)d\xi\\ &+\int_{0}^{\infty}\frac{2\phi_{\xi}}{\xi}f(\xi)d\xi+\int_{0}^{\infty}\frac{2(h_{3}(r)-1)}{\xi}\phi_{\xi}(r)f(\xi)d\xi.\end{split}

We use the transference operator 𝒦\mathcal{K} to rewrite the first term on the right-hand side of (5.103) in the form

(5.104) 0(rrξξ)ϕξ(r)f(ξ)ξ𝑑ξ=rr1[0f/ξ]+1[0f()]=1[0ξξ(f/ξ)]+1(𝒦32)[0f/ξ]+1[0f()]=1(𝒦12)[0f/ξ].\begin{split}\int_{0}^{\infty}(r\partial_{r}-\xi\partial_{\xi})\phi_{\xi}(r)\frac{f(\xi)}{\xi}d\xi&=r\partial_{r}{\mathcal{F}}^{-1}\begin{bmatrix}0\\ f/\xi\end{bmatrix}+{\mathcal{F}}^{-1}\begin{bmatrix}0\\ f^{\prime}(\cdot)\end{bmatrix}\\ &={\mathcal{F}}^{-1}\begin{bmatrix}0\\ -\xi\partial_{\xi}\left(f/\xi\right)\end{bmatrix}+{\mathcal{F}}^{-1}(\mathcal{K}-\frac{3}{2})\begin{bmatrix}0\\ f/\xi\end{bmatrix}+{\mathcal{F}}^{-1}\begin{bmatrix}0\\ f^{\prime}(\cdot)\end{bmatrix}\\ &={\mathcal{F}}^{-1}(\mathcal{K}-\frac{1}{2})\begin{bmatrix}0\\ f/\xi\end{bmatrix}.\end{split}

Overall, (5.103) becomes

(5.105) r~1f=1(𝒦+32)[0f/ξ]1[0f()]+2(h3(r)1)1[0f/ξ]\begin{split}r{\tilde{\mathcal{F}}}^{-1}f&={\mathcal{F}}^{-1}\left(\mathcal{K}+\frac{3}{2}\right)\begin{bmatrix}0\\ f/\xi\end{bmatrix}-{\mathcal{F}}^{-1}\begin{bmatrix}0\\ f^{\prime}(\cdot)\end{bmatrix}+2(h_{3}(r)-1){\mathcal{F}}^{-1}\begin{bmatrix}0\\ f/\xi\end{bmatrix}\end{split}

which, by the L2L^{2} boundedness of the Fourier transform and of the transference operator (see Proposition 5.7) gives

r~1(f)L2(rdr)f(ξ)ξL2(dξ)+f(ξ)L2(dξ),\|r{\tilde{\mathcal{F}}}^{-1}(f)\|_{L^{2}(rdr)}\lesssim\|\frac{f(\xi)}{\xi}\|_{L^{2}(d\xi)}+\|f^{\prime}(\xi)\|_{L^{2}(d\xi)},

as needed. ∎

5.5. Littlewood-Paley projectors in the H~λ\tilde{H}_{\lambda} frame

In this section we seek to understand the properties of the Littlewood-Paley projectors in the H~λ\tilde{H}_{\lambda} frame. Here one could proceed as we did in the previous subsections, with λ=1\lambda=1, and then rescale. But for reference purposes we preferred to state the results with the parameter λ\lambda included.

We recall from Section 2 the functions mj,jm_{j},j\in\mathbb{Z} with the properties

mj(x)=m(x2j),mCc((14,2)),m(x)=1,12x1.m_{j}(x)=m(\frac{x}{2^{j}}),\quad m\in C^{\infty}_{c}((\frac{1}{4},2)),\quad m(x)=1,\quad\frac{1}{2}\leq x\leq 1.

and

jmj=1.\sum_{j}m_{j}=1.

Based on these functions we define our projectors

(5.106) Pjλ=~λ1mj~λ.P^{\lambda}_{j}={\tilde{\mathcal{F}}}_{\lambda}^{-1}m_{j}{\tilde{\mathcal{F}}}_{\lambda}.

We note that the notation PjλP_{j}^{\lambda} does not carry the ~\tilde{\cdot} symbol which we used for the operator H~λ\tilde{H}_{\lambda}. The reasons we do so is because throughout most of this paper we work with projectors only in the H~λ\tilde{H}_{\lambda} calculus, and not with the ones in the HλH_{\lambda} calculus; the only exception is in Section 6 where we will carefully differentiate between the notation used for the projectors in the two different calculi. This allows us to reserve the notation P~jλ\tilde{P}_{j}^{\lambda} for the purpose described below (which is standard in the literature).

We also need the projectors P~jλ\tilde{P}_{j}^{\lambda} which also localize at frequency 2j2^{j} and enjoy the property P~jλPjλ=Pjλ\tilde{P}_{j}^{\lambda}P_{j}^{\lambda}=P_{j}^{\lambda}; they are constructed using the functions m~j\widetilde{m}_{j} by

P~jλ=~λ1m~j~λ.\tilde{P}^{\lambda}_{j}={\tilde{\mathcal{F}}}_{\lambda}^{-1}{\tilde{m}}_{j}{\tilde{\mathcal{F}}}_{\lambda}.

We recall from Section 2 that m~j(x)=m~(x2j)\widetilde{m}_{j}(x)=\widetilde{m}(\frac{x}{2^{j}}) and that m~Cc((0,)),m~(x)=1\widetilde{m}\in C^{\infty}_{c}((0,\infty)),\widetilde{m}(x)=1 in the support of mm.

If uCc((0,))u\in C^{\infty}_{c}((0,\infty)), then we can represent PjλuP_{j}^{\lambda}u using an integral kernel,

Pjλ(u)(r)=0Kjλ(r,s)u(s)s𝑑s,P_{j}^{\lambda}(u)(r)=\int_{0}^{\infty}K_{j}^{\lambda}(r,s)u(s)sds,

where the kernel of the projector is given by

Kjλ(r,s)=0ψξλ(r)mj(ξ)ψξλ(s)𝑑ξ=λ0ψλ1ξ(rλ)mj(ξ)ψλ1ξ(sλ)𝑑ξ.K_{j}^{\lambda}(r,s)=\int_{0}^{\infty}\psi^{\lambda}_{\xi}(r)m_{j}(\xi)\psi^{\lambda}_{\xi}(s)d\xi=\lambda\int_{0}^{\infty}\psi_{\lambda^{-1}\xi}(r\lambda)m_{j}(\xi)\psi_{\lambda^{-1}\xi}(s\lambda)d\xi.

This kernel and its properties will play an important role in this paper.

There is another kernel whose characterization we need, which arises when we seek to express a function ψ\psi in terms of LψL^{*}\psi in an elliptic fashion, for a frequency localized function ψ\psi. This is done based on the identity

Pjλψ=Pjλ~(Pjλψ)=Pjλ~H~λ1LλLλ(Pjλψ).P^{\lambda}_{j}\psi=\widetilde{P_{j}^{\lambda}}(P_{j}^{\lambda}\psi)=\widetilde{P_{j}^{\lambda}}\widetilde{H}_{\lambda}^{-1}L_{\lambda}L^{*}_{\lambda}(P_{j}^{\lambda}\psi).

which shows that the operator we need to consider is Pjλ~H~λ1Lλ\widetilde{P_{j}^{\lambda}}\widetilde{H}_{\lambda}^{-1}L_{\lambda}. We also represent this operator via an integral kernel, writing for uCc((0,))u\in C^{\infty}_{c}((0,\infty))

(5.107) Pj~λ(H~λ1Lλ(u))(r)=0ψξλ(r)m~j(ξ)~λ(H~λ1Lλu)(ξ)𝑑ξ=0Kj1,λ(s,r)u(s)s𝑑s.\begin{split}\widetilde{P_{j}}^{\lambda}(\widetilde{H}_{\lambda}^{-1}L_{\lambda}(u))(r)=\int_{0}^{\infty}\psi^{\lambda}_{\xi}(r)\widetilde{m}_{j}(\xi){\tilde{\mathcal{F}}}_{\lambda}(\widetilde{H}_{\lambda}^{-1}L_{\lambda}u)(\xi)d\xi=\int_{0}^{\infty}K_{j}^{1,\lambda}(s,r)u(s)sds.\end{split}

Here we recall that ξψξλ=Lλϕξ\xi\psi^{\lambda}_{\xi}=L_{\lambda}\phi_{\xi}. This allows us to use the HλH_{\lambda} associated Fourier representation of uu in order to write

Kj1,λ(s,r)=0ψξλ(r)m~j(ξ)ϕξλ(s)ξ𝑑ξ=λ0ψξλ(rλ)m~j(ξ)ϕξλ(sλ)ξ𝑑ξ.K_{j}^{1,\lambda}(s,r)=\int_{0}^{\infty}\psi^{\lambda}_{\xi}(r)\widetilde{m}_{j}(\xi)\frac{\phi^{\lambda}_{\xi}(s)}{\xi}d\xi=\lambda\int_{0}^{\infty}\psi_{\frac{\xi}{\lambda}}(r\lambda)\widetilde{m}_{j}(\xi)\frac{\phi_{\frac{\xi}{\lambda}}(s\lambda)}{\xi}d\xi.

Another related operator that will be used in this paper is Lλ1Pjλ~L^{-1}_{\lambda}\widetilde{P_{j}^{\lambda}}, which can be defined as

Lλ1Pjλ~:=LλH~λ1Pjλ~.L^{-1}_{\lambda}\widetilde{P_{j}^{\lambda}}:=L^{*}_{\lambda}\widetilde{H}_{\lambda}^{-1}\widetilde{P_{j}^{\lambda}}.

We remark that its adjoint is given by

(Lλ1Pjλ~)=Pjλ~(Lλ)1:=Pjλ~H~λ1Lλ,(L^{-1}_{\lambda}\widetilde{P_{j}^{\lambda}})^{*}=\widetilde{P_{j}^{\lambda}}(L_{\lambda}^{*})^{-1}:=\widetilde{P_{j}^{\lambda}}\tilde{H}_{\lambda}^{-1}L_{\lambda},

therefore the kernel of Lλ1Pjλ~L^{-1}_{\lambda}\widetilde{P_{j}^{\lambda}} is simply Kj1,λ(r,s)K_{j}^{1,\lambda}(r,s).

Our aim in this section is to characterize the two kernels introduced above, Kjλ(r,s)K_{j}^{\lambda}(r,s) (the kernel of the projector PjλP_{j}^{\lambda}) and Kj1,λ(r,s)K_{j}^{1,\lambda}(r,s). For this purpose we introduce two auxiliary weight functions

(5.108) ωj,λ(r)={min{1,r323j},2jλmin{1,2jrr2λ21+r2λ2},2jλ.\omega_{j,\lambda}(r)=\begin{cases}\text{min}\{1,r^{3}2^{3j}\},\quad 2^{j}\geq\lambda\\ \text{min}\{1,2^{j}r\dfrac{r^{2}\lambda^{2}}{1+r^{2}\lambda^{2}}\},\quad 2^{j}\leq\lambda.\end{cases}

and

(5.109) ω~j,λ(r)={min{1,r2λ2r2λ22+r2λ2(2jr)2r2λ2},2jλmin{1,r222j},2jλ.\tilde{\omega}_{j,\lambda}(r)=\begin{cases}\text{min}\{1,\dfrac{r^{2}\lambda^{2}}{\langle r^{2}\lambda^{2}\rangle^{2}}+\dfrac{r^{2}\lambda^{2}(2^{j}r)^{2}}{\langle r^{2}\lambda^{2}\rangle}\},\quad 2^{j}\leq\lambda\\ \text{min}\{1,r^{2}2^{2j}\},\quad 2^{j}\geq\lambda\end{cases}.

With this notation in place, we state our main result in this section.

Proposition 5.11.

The kernels KjλK^{\lambda}_{j} and Kj1,λK_{j}^{1,\lambda} satisfy the following bounds:

(5.110) |Kjλ(r,s)|CN22jωj,λ(r)ωj,λ(s)(1+2j(s+r))(1+2j|rs|)N,|K^{\lambda}_{j}(r,s)|\leq\frac{C_{N}2^{2j}\omega_{j,\lambda}(r)\omega_{j,\lambda}(s)}{(1+2^{j}(s+r))(1+2^{j}|r-s|)^{N}},
(5.111) |rKjλ(r,s)|CN23j(1+2j(r+s))ωj,λ(s)(1+2j|rs|)N(ωj,λ(r)r2j+ω~j,λ(r)),|\partial_{r}K^{\lambda}_{j}(r,s)|\leq\frac{C_{N}2^{3j}}{(1+2^{j}(r+s))}\frac{\omega_{j,\lambda}(s)}{(1+2^{j}|r-s|)^{N}}\left(\frac{\omega_{j,\lambda}(r)}{r2^{j}}+\tilde{\omega}_{j,\lambda}(r)\right),
(5.112) |Kj1,λ(r,s)|CN2j(1+2j(s+r))ωj,λ(s)(1+2j|sr|)Nω~j,λ(r).|K_{j}^{1,\lambda}(r,s)|\leq\frac{C_{N}2^{j}}{(1+2^{j}(s+r))}\frac{\omega_{j,\lambda}(s)}{(1+2^{j}|s-r|)^{N}}\tilde{\omega}_{j,\lambda}(r).

In addition, if 2jr12^{j}r\leq 1 and 2jλ2^{j}\leq\lambda, then Kj1,λK_{j}^{1,\lambda} can be decomposed as

Kj1,λ(r,s)=Kj,res1,λ(r,s)+Kj,reg1,λ(r,s)K_{j}^{1,\lambda}(r,s)=K_{j,res}^{1,\lambda}(r,s)+K_{j,reg}^{1,\lambda}(r,s)

where

Kj,res1,λ(r,s)=λϕ0(rλ)0q(ξ)mj~(ξ)ξψλ1ξ(sλ)𝑑ξK_{j,res}^{1,\lambda}(r,s)=\lambda\phi_{0}(r\lambda)\int_{0}^{\infty}\frac{q(\xi)\widetilde{m_{j}}(\xi)}{\xi}\psi_{\lambda^{-1}\xi}(s\lambda)d\xi

and we have

(5.113) |Kj,res1,λ(r,s)|CNϕ0(rλ)𝟙{r2j1}2jmin{1,s2js2λ2s2λ2}(s2j+1)N,|Kj,reg1,λ(r,s)|CN2j((r+s)2j+1)ωj,λ(r)(1+2j|rs|)Nωj,λ(s).\begin{split}&|K_{j,res}^{1,\lambda}(r,s)|\leq C_{N}\phi_{0}(r\lambda)\mathbbm{1}_{\{r2^{j}\leq 1\}}2^{j}\frac{\min\{1,\dfrac{s2^{j}s^{2}\lambda^{2}}{\langle s^{2}\lambda^{2}\rangle}\}}{(s2^{j}+1)^{N}},\\ &|K_{j,reg}^{1,\lambda}(r,s)|\leq\frac{C_{N}2^{j}}{((r+s)2^{j}+1)}\frac{\omega_{j,\lambda}(r)}{(1+2^{j}|r-s|)^{N}}\omega_{j,\lambda}(s).\end{split}

We remark that in these statements λ\lambda is just a scale parameter, and all the bounds are equivalent to the ones for λ=1\lambda=1. The same applies to the proof below.

We also note that the estimates on KjλK_{j}^{\lambda} in the proposition above are also true for the kernel of Pjλ~\widetilde{P_{j}^{\lambda}}.

Proof.

We begin with (5.110), for which we consider several cases:

(i) 2jr12^{j}r\leq 1 and 2js12^{j}s\leq 1. From the formulae for ψξ(r)\psi_{\xi}(r), we obtain the bound

|ψξλ(rλ)|C2j/2λωj,λ(r),ξ2j,rξ1.|\psi_{\frac{\xi}{\lambda}}(r\lambda)|\leq\frac{C2^{j/2}}{\sqrt{\lambda}}\omega_{j,\lambda}(r),\quad\xi\sim 2^{j},\quad r\xi\lesssim 1.

Then we directly estimate using the definition of KjλK_{j}^{\lambda} to get

|Kjλ(r,s)|C22jωj,λ(r)ωj,λ(s),r,s2j.|K_{j}^{\lambda}(r,s)|\leq C2^{2j}\omega_{j,\lambda}(r)\omega_{j,\lambda}(s),\quad r,s\leq 2^{-j}.

(ii) r2jsr\leq 2^{-j}\leq s (and the symmetric case) Here we have

Kjλ(r,s)=λ2sλRe(0mj(ξ)ψξλ(rλ)a(ξλ)eisξσ~(sξ,s)𝑑ξ).K_{j}^{\lambda}(r,s)=\lambda\frac{2}{\sqrt{s\lambda}}\text{Re}\left(\int_{0}^{\infty}m_{j}(\xi)\psi_{\frac{\xi}{\lambda}}(r\lambda)a(\frac{\xi}{\lambda})e^{is\xi}\widetilde{\sigma}(s\xi,s)d\xi\right).

Then we integrate by parts, using the symbol-type estimates on σ~\widetilde{\sigma} and aa, to obtain

|Kjλ(r,s)|CN22j(2js)Nωj,λ(r).|K_{j}^{\lambda}(r,s)|\leq\frac{C_{N}2^{2j}}{(2^{j}s)^{N}}\omega_{j,\lambda}(r).

This suffices, since 2js12^{j}s\geq 1, so ωj,λ(s)1\omega_{j,\lambda}(s)\sim 1.

(iii) 2jrs2^{-j}\leq r\leq s. Then we have

(5.114) Kjλ(r,s)=0𝑑ξmj(ξ)rs(ei(r+s)ξ(a¯(ξλ))2σ~¯(rξ,rλ)σ~¯(sξ,sλ)+ei(rs)ξ|a(ξλ)|2σ~¯(rξ,rλ)σ~(sξ,sλ)+ei(sr)ξ|a(ξλ)|2σ~(rξ,rλ)σ~¯(sξ,sλ)+ei(r+s)ξ(a(ξλ))2σ~(rξ,rλ)σ~(sξ,sλ)).K_{j}^{\lambda}(r,s)=\int_{0}^{\infty}d\xi\frac{m_{j}(\xi)}{\sqrt{r}\sqrt{s}}\begin{aligned} &\left(e^{-i(r+s)\xi}(\overline{a}(\frac{\xi}{\lambda}))^{2}\overline{\widetilde{\sigma}}(r\xi,r\lambda)\overline{\widetilde{\sigma}}(s\xi,s\lambda)\right.\\ &\left.+e^{-i(r-s)\xi}|a(\frac{\xi}{\lambda})|^{2}\overline{\widetilde{\sigma}}(r\xi,r\lambda)\widetilde{\sigma}(s\xi,s\lambda)\right.\\ &\left.+e^{-i(s-r)\xi}|a(\frac{\xi}{\lambda})|^{2}\widetilde{\sigma}(r\xi,r\lambda)\overline{\widetilde{\sigma}}(s\xi,s\lambda)\right.\\ &\left.+e^{i(r+s)\xi}(a(\frac{\xi}{\lambda}))^{2}\widetilde{\sigma}(r\xi,r\lambda)\widetilde{\sigma}(s\xi,s\lambda)\right).\end{aligned}

If 2j|rs|12^{j}|r-s|\lesssim 1, then we estimate the above integral directly to get

|Kjλ(r,s)|C22jrs2jC22jωj,λ(r)ωj,λ(s)(1+2j(r+s)).|K_{j}^{\lambda}(r,s)|\leq\frac{C2^{2j}}{\sqrt{r}\sqrt{s}2^{j}}\leq\frac{C2^{2j}\omega_{j,\lambda}(r)\omega_{j,\lambda}(s)}{(1+2^{j}(r+s))}.

If s2rs\frac{s}{2}\leq r\leq s and 2j|rs|12^{j}|r-s|\geq 1 then we integrate by parts to obtain

|Kjλ(r,s)|C2jrs12jN1|rs|NC22jωj,λ(r)ωj,λ(s)(1+2j(r+s))(1+2j|rs|)N.|K_{j}^{\lambda}(r,s)|\leq\frac{C2^{j}}{\sqrt{rs}}\frac{1}{2^{jN}}\frac{1}{|r-s|^{N}}\leq\frac{C2^{2j}\omega_{j,\lambda}(r)\omega_{j,\lambda}(s)}{(1+2^{j}(r+s))(1+2^{j}|r-s|)^{N}}.

Finally, if rs2r\leq\frac{s}{2} then we integrate by parts, to obtain instead

|Kjλ(r,s)|C22j(2js)Nωj,λ(r)ωj,λ(s).|K_{j}^{\lambda}(r,s)|\leq\frac{C2^{2j}}{(2^{j}s)^{N}}\omega_{j,\lambda}(r)\omega_{j,\lambda}(s).

The estimates for Kj1,λK^{1,\lambda}_{j} are obtained using the same procedure. More precisely, the expression for Kj1,λK_{j}^{1,\lambda} is exactly that of KjλK_{j}^{\lambda}, except that ψξλ(t)(sλ(t))\psi_{\frac{\xi}{\lambda(t)}}(s\lambda(t)) is replaced by ξ1ϕξλ(t)(sλ(t))\xi^{-1}\phi_{\frac{\xi}{\lambda(t)}}(s\lambda(t)). This explains the similar form of the corresponding estimates, and the extra factor of 2j2^{-j} in the Kj1,λK_{j}^{1,\lambda} estimate, compared to that of KjλK_{j}^{\lambda} . To estimate rKjλ\partial_{r}K_{j}^{\lambda}, we use the fact that

Lλ=r+2h3λ1rL^{*}_{\lambda}=-\partial_{r}+\frac{2h_{3}^{\lambda}-1}{r}

and estimate LλKjλ(r,s)L^{*}_{\lambda}K^{\lambda}_{j}(r,s). We have

LλKjλ(r,s)=λ0ϕξλ(rλ)ξmj(ξ)ψξλ(sλ)𝑑ξ.L^{*}_{\lambda}K_{j}^{\lambda}(r,s)=\lambda\int_{0}^{\infty}\phi_{\frac{\xi}{\lambda}}(r\lambda)\xi m_{j}(\xi)\psi_{\frac{\xi}{\lambda}}(s\lambda)d\xi.

which has a similar form as Kj1,λ(r,s)K_{j}^{1,\lambda}(r,s), except for an extra factor of ξ2\xi^{2} in the integrand.

For (5.113) we write

ϕξλ(rλ)=ϕ0(rλ)q(ξλ)+ϕξλ(rλ)ϕ0(rλ)q(ξλ),\phi_{\frac{\xi}{\lambda}}(r\lambda)=\phi_{0}(r\lambda)q(\frac{\xi}{\lambda})+\phi_{\frac{\xi}{\lambda}}(r\lambda)-\phi_{0}(r\lambda)q(\frac{\xi}{\lambda}),

and note that Kj,res1,λK_{j,res}^{1,\lambda} is obtained by inserting the first term of the above decomposition into the definition of Kj1,λK_{j}^{1,\lambda}. Then we have

|Kj,res1,λ(r,s)|Cλϕ0(rλ)𝟙{r2j1}0q(ξλ)|mj~(ξ)|ξ|ψξλ(sλ)|𝑑ξ.|K_{j,res}^{1,\lambda}(r,s)|\leq C\lambda\phi_{0}(r\lambda)\mathbbm{1}_{\{r2^{j}\leq 1\}}\int_{0}^{\infty}\frac{q(\frac{\xi}{\lambda})|\widetilde{m_{j}}(\xi)|}{\xi}|\psi_{\frac{\xi}{\lambda}}(s\lambda)|\,d\xi.

If 2js12^{j}s\lesssim 1 then we directly estimate the integral, while for 2js12^{j}s\gtrsim 1 we write

Kj,res1,λ=λϕ0(rλ)𝟙{r2j1}0𝑑ξq(ξλ)mj~(ξ)ξ2Re(a(ξλ)ψξλ+(sλ))K_{j,res}^{1,\lambda}=\lambda\phi_{0}(r\lambda)\mathbbm{1}_{\{r2^{j}\leq 1\}}\int_{0}^{\infty}d\xi\frac{q(\frac{\xi}{\lambda})\widetilde{m_{j}}(\xi)}{\xi}\cdot 2\text{Re}\left(a(\frac{\xi}{\lambda})\psi_{\frac{\xi}{\lambda}}^{+}(s\lambda)\right)

and then we integrate by parts. Next, we have

Kj,reg1,λ=λ𝟙{r2j1}l=10𝑑ξq(ξλ)(rξ)2lψl(r2λ2)mj~(ξ)ξψξλ(sλ).K_{j,reg}^{1,\lambda}=\lambda\mathbbm{1}_{\{r2^{j}\leq 1\}}\sum_{l=1}^{\infty}\int_{0}^{\infty}d\xi q(\frac{\xi}{\lambda})(r\xi)^{2l}\psi_{l}(r^{2}\lambda^{2})\frac{\widetilde{m_{j}}(\xi)}{\xi}\psi_{\frac{\xi}{\lambda}}(s\lambda).

We again directly estimate the integral in the case 2js12^{j}s\lesssim 1 and integrate by parts as above for the case 2js12^{j}s\gtrsim 1. ∎

Using the above kernel type estimate we derive the following Bernstein-type estimate.

Lemma 5.12.

Assume that fL2(rdr)f\in L^{2}(rdr) is localized at frequency 2k2^{k} in the H~λ\tilde{H}_{\lambda} calculus; then for any 1qp1\leq q\leq p\leq\infty the following holds true:

(5.115) fLp(rdr)22k(1q1p)fLq(rdr).\|f\|_{L^{p}(rdr)}\lesssim 2^{2k(\frac{1}{q}-\frac{1}{p})}\|f\|_{L^{q}(rdr)}.
Proof.

Without restricting the generality of the argument we can assume that f=P~kλff=\tilde{P}_{k}^{\lambda}f, thus we have

f(r)=K~kλ(r,s)f(s)s𝑑s.f(r)=\int\tilde{K}_{k}^{\lambda}(r,s)f(s)sds.

The bounds on the kernel Kkλ(r,s)K_{k}^{\lambda}(r,s) are provided in (5.110). We write

f=χAkf+j1χ[2kj,2k(j+1)]f:=f0+j1fj.f=\chi_{A_{\leq-k}}f+\sum_{j\geq 1}\chi_{[2^{-k}j,2^{-k}(j+1)]}f:=f_{0}+\sum_{j\geq 1}f_{j}.

For the first term we simply use

|K~kλ(r,s)f0(s)s𝑑s|22k(1+2kr)N|f0(s)|s𝑑s,|\int\tilde{K}_{k}^{\lambda}(r,s)f_{0}(s)sds|\lesssim\int 2^{2k}(1+2^{k}r)^{-N}|f_{0}(s)|sds,

from which we obtain

K~kλ(r,s)f0(s)s𝑑sLp22k(1q1p)f0Lq(rdr)\|\int\tilde{K}_{k}^{\lambda}(r,s)f_{0}(s)sds\|_{L^{p}}\lesssim 2^{2k(\frac{1}{q}-\frac{1}{p})}\|f_{0}\|_{L^{q}(rdr)}

with improved decay away from AkA_{-k}, that is

K~kλ(r,s)f0(s)s𝑑sLp[2kj,2k(j+1)]jN22k(1q1p)f0Lq(rdr),j1.\|\int\tilde{K}_{k}^{\lambda}(r,s)f_{0}(s)sds\|_{L^{p}[2^{-k}j,2^{-k}(j+1)]}\lesssim\langle j\rangle^{-N}2^{2k(\frac{1}{q}-\frac{1}{p})}\|f_{0}\|_{L^{q}(rdr)},\quad\forall j\geq 1.

For the terms in the sum, we use

|K~kλ(r,s)fj(s)s𝑑s|22k(1+2k(r+s))(1+2k|rs|)Nf(s)s𝑑s,|\int\tilde{K}_{k}^{\lambda}(r,s)f_{j}(s)sds|\lesssim\int 2^{2k}(1+2^{k}(r+s))(1+2^{k}|r-s|)^{-N}f(s)sds,

and then rely on Schur’s test to obtain

K~kλ(r,s)fj(s)s𝑑sLp22k(1q1p)fjLq(rdr).\|\int\tilde{K}_{k}^{\lambda}(r,s)f_{j}(s)sds\|_{L^{p}}\lesssim 2^{2k(\frac{1}{q}-\frac{1}{p})}\|f_{j}\|_{L^{q}(rdr)}.

Improved decay is away from [2kj,2k(j+1)][2^{-k}j,2^{-k}(j+1)] is also available, and this allows us to retrieve the estimate (5.115) from the estimates above.

5.6. The time dependence of the Littlewood-Paley projectors

Throughout this section, we have studied the spectral properties of the operators HλH_{\lambda} and H~λ\tilde{H}_{\lambda} where λ(0,)\lambda\in(0,\infty) is a scaling parameter. As noted in Section 4.5, this scale parameter is needed because the linear operator H~λ=Δ+V~λ\tilde{H}_{\lambda}=-\Delta+\tilde{V}_{\lambda}, that naturally occurs in the PDE governing the dynamics of the field ψ\psi, has in fact a time dependent potential coming precisely from the fact that λ=λ(t)\lambda=\lambda(t).

In Section 7 we will seek to establish estimates for the linear equation (4.60)

(itH~λ)Ψ=0,(i\partial_{t}-\tilde{H}_{\lambda})\Psi=0,

when λ=λ(t)\lambda=\lambda(t). In doing so, we will use the appropriate form of a Littlewood-Paley decomposition, at which stage we encounter the operator [t,Pkλ][\partial_{t},P_{k}^{\lambda}], whose various operator norms need to be estimated.

For convenience, we found it more efficient to prepare the study of this operator in this section. Thus, in what follows, we assume λ:I\lambda:I\rightarrow\mathbb{R}, where II\subset\mathbb{R} can be any interval. In terms of regularity, we will eventually work with λH˙12L\lambda\in\dot{H}^{\frac{1}{2}}\cap L^{\infty}, although for all practical purposes one can assume that λ\lambda is continuously differentiable, by Corollary 4.9.

Recall that the Fourier transform associated to H~λ\tilde{H}_{\lambda} is obtained by rescaling from the Fourier transform associated to H~\tilde{H}, see (5.29).

We also recall, from (5.106), the spectral projectors PjλP^{\lambda}_{j} associated to H~λ\tilde{H}_{\lambda}. We are now interested in the time derivative of the spectral projectors. The first step will be to express it in terms of the transference operator.

Lemma 5.13.

The commutator of t\partial_{t} with Pkλ(t)P_{k}^{\lambda(t)} is given by the following:

(5.116) [t,Pkλ(t)]u=λ(t)~λ(t)1[𝒦~λ(t),mk]~λ(t)u[\partial_{t},P_{k}^{\lambda(t)}]u=\lambda^{\prime}(t){\tilde{\mathcal{F}}}^{-1}_{\lambda(t)}[\tilde{\mathcal{K}}_{\lambda(t)},m_{k}]{\tilde{\mathcal{F}}}_{\lambda(t)}u

where 𝒦~λ\tilde{\mathcal{K}}_{\lambda} is as in (5.98).

Proof.

The proof is a direct computation using the relations (5.100):

tPkλ(t)=λ(t)λ(~λ1mk~λ)=λ(t)(~λ1𝒦~λmk~λ~λ1mk𝒦~λ~λ)\partial_{t}P_{k}^{\lambda(t)}=\lambda^{\prime}(t)\partial_{\lambda}({\tilde{\mathcal{F}}}^{-1}_{\lambda}m_{k}{\tilde{\mathcal{F}}}_{\lambda})=\lambda^{\prime}(t)({\tilde{\mathcal{F}}}^{-1}_{\lambda}\tilde{\mathcal{K}}_{\lambda}m_{k}{\tilde{\mathcal{F}}}_{\lambda}-{\tilde{\mathcal{F}}}^{-1}_{\lambda}m_{k}\tilde{\mathcal{K}}_{\lambda}{\tilde{\mathcal{F}}}_{\lambda})

and (5.116) follows.

Given the above commutator formula, it is natural to seek estimates for the commutator [𝒦~λ,mj][\tilde{\mathcal{K}}_{\lambda},m_{j}], where we recall the notation (5.98). Heuristically, we expect this commutator to be largest when λ\lambda and jj are matched, i.e. λ2j\lambda\approx 2^{j}. To measure the decay away from this region we will use the weight

χλ=2j=(2j/2λ𝟙{2jλ4}+λ2j𝟙{2jλ1}).\chi_{\lambda=2^{j}}=\left(\frac{2^{j/2}}{\sqrt{\lambda}}\mathbbm{1}_{\{\frac{2^{j}}{\lambda}\leq 4\}}+\frac{\lambda}{2^{j}}\mathbbm{1}_{\{\frac{2^{j}}{\lambda}\geq 1\}}\right).

While mjm_{j} provided frequency localization, 𝒦~λ\tilde{\mathcal{K}}_{\lambda} does not, so neither does the commutator. However, we also expect to have decay at frequencies 2k2^{k} when kk is away from jj. For this we will use the weight

χk=j=2|kj|2.\chi_{k=j}=2^{-\frac{|k-j|}{2}}.

Note that if |kj|>2|k-j|>2 then we have

mk[𝒦~λ,mj]=mk𝒦~λmj=[mk,𝒦~λ]mj,m_{k}[\tilde{\mathcal{K}}_{\lambda},m_{j}]=m_{k}\tilde{\mathcal{K}}_{\lambda}m_{j}=[m_{k},\tilde{\mathcal{K}}_{\lambda}]m_{j},

so it does not matter if we place the mkm_{k} localization on the left or on the right. For symmetry we will include both. Our main bounds in this section are collected in the following Lemma.

Lemma 5.14.

The following fixed time estimates hold true:

(5.117) ml[𝒦~λ,mj]mkuL2(dξ)2jλ2χλ=2kχj=kχj=luL2(dξ)\begin{split}\|m_{l}[\tilde{\mathcal{K}}_{\lambda},m_{j}]m_{k}u\|_{L^{2}(d\xi)}\lesssim\frac{2^{j}}{\lambda^{2}}\chi_{\lambda=2^{k}}\chi_{j=k}\chi_{j=l}\|u\|_{L^{2}(d\xi)}\end{split}

respectively

(5.118) r~λ1ml[𝒦~λ,mj]mkuL2(rdr)1λ2χλ=2jχj=kχj=luL2(dξ).\begin{split}&\|r{\tilde{\mathcal{F}}}_{\lambda}^{-1}m_{l}[\tilde{\mathcal{K}}_{\lambda},m_{j}]m_{k}u\|_{L^{2}(rdr)}\lesssim\frac{1}{\lambda^{2}}\chi_{\lambda=2^{j}}\chi_{j=k}\chi_{j=l}\|u\|_{L^{2}(d\xi)}.\end{split}
Proof.

We first remark that λ\lambda is a scaling parameter in this Lemma. Hence, without any loss in generality we can simply set λ=1\lambda=1 in the proof. Using the representation in (5.90) for the kernel of 𝒦~\tilde{\mathcal{K}}, it is easily seen that the operator ml[mj,𝒦~]mkm_{l}[m_{j},\tilde{\mathcal{K}}]m_{k} can be expressed in the form

(5.119) ml[mj,𝒦~]mkf(ξ):=0mj(ξ)mj(η)ξ2η2ml(ξ)mk(η)F~(ξ,η)f(η)𝑑η\begin{split}&m_{l}[m_{j},\tilde{\mathcal{K}}]m_{k}f(\xi):=-\int_{0}^{\infty}\frac{m_{j}(\xi)-m_{j}(\eta)}{\xi^{2}-\eta^{2}}m_{l}(\xi)m_{k}(\eta)\tilde{F}(\xi,\eta)f(\eta)d\eta\end{split}

with the integral kernel

Kjkl(ξ,η)=mj(ξ)mj(η)ξ2η2ml(ξ)mk(η)F~(ξ,η),K_{jkl}(\xi,\eta)=\frac{m_{j}(\xi)-m_{j}(\eta)}{\xi^{2}-\eta^{2}}m_{l}(\xi)m_{k}(\eta)\tilde{F}(\xi,\eta),

where we note that one consequence of taking the commutator is that the kernel no longer has a singularity on the diagonal. This kernel vanishes unless we have either |lj|<4|l-j|<4 or |kj|<4|k-j|<4. Assuming this holds, we note that Kjkl(ξ,η)K_{jkl}(\xi,\eta) is supported in the region {ξ2l,η2k}\{\xi\approx 2^{l},\ \eta\approx 2^{k}\}, and satisfies

|Kjkl(ξ,η)||F~(ξ,η)|(ξ+η)2,|ξKjkl(ξ,η)|+ξ1|Kjkl(ξ,η)||F~(ξ,η)|ξ(ξ+η)2+|ξF~(ξ,η)|(ξ+η)2.|K_{jkl}(\xi,\eta)|\lesssim\frac{|\tilde{F}(\xi,\eta)|}{(\xi+\eta)^{2}},\qquad|\partial_{\xi}K_{jkl}(\xi,\eta)|+\xi^{-1}|K_{jkl}(\xi,\eta)|\lesssim\frac{|\tilde{F}(\xi,\eta)|}{\xi(\xi+\eta)^{2}}+\frac{|\partial_{\xi}\tilde{F}(\xi,\eta)|}{(\xi+\eta)^{2}}.

To prove (5.117) it suffices to have the

Kjkl(ξ,η)Lξ,η22jχk=0χj=kχj=l\|K_{jkl}(\xi,\eta)\|_{L^{2}_{\xi,\eta}}\lesssim 2^{j}\chi_{k=0}\chi_{j=k}\chi_{j=l}

which is easily verified using the above kernel bound and (5.91).

For (5.118) we use Lemma 5.10, which reduces it to the bounds

ξKjkl(ξ,η)Lξ,η2+ξ1Kjkl(ξ,η)Lξ,η2χk=0χj=kχj=l\|\partial_{\xi}K_{jkl}(\xi,\eta)\|_{L^{2}_{\xi,\eta}}+\|\xi^{-1}K_{jkl}(\xi,\eta)\|_{L^{2}_{\xi,\eta}}\lesssim\chi_{k=0}\chi_{j=k}\chi_{j=l}

which is again easily verified using the above kernel bound together with both (5.91) and (5.92). ∎

6. Elliptic analysis and 1\ell^{1} Besov structures

Most of the analysis in this paper is concentrated around the gauge field ψL2\psi\in L^{2} and the modulation parameters α\alpha and λ\lambda. In Section 4 we have shown why this is essential in order to understand the dynamics of the original map u(t)u(t), which is uniquely determined by (ψ,α,λ)(\psi,\alpha,\lambda). In particular one may think of Qα(t),λ(t)Q_{\alpha(t),\lambda(t)} as the closest soliton to uu, and the orthogonality condition (4.29) ensures that our choice of α,λ\alpha,\lambda satisfies the condition (see (4.35))

uQα,λ2H˙1ψL2.\|u-Q^{2}_{\alpha,\lambda}\|_{\dot{H}^{1}}\approx\|\psi\|_{L^{2}}.

Our results assert that the smallness of ψ\psi in L2L^{2} suffices in order to guarantee global well-posedness for the Schrödinger map flow, but likely this does not preclude blow-up at infinity. Then a natural question becomes whether there exists some slightly stronger topology for the initial data u(0)u(0), and correspondingly for ψ(0)\psi(0), where we have soliton stability, and in particular no blow-up can happen.

This turns out to be indeed the case, and the appropriate spaces for u(0)u(0), respectively ψ(0)\psi(0) are Besov spaces with 1\ell^{1} summation, precisely

(6.1) X=B˙2,11,X¯=B˙2,1,e1,LX¯=B˙2,1,e0,X=\dot{B}^{1}_{2,1},\qquad\bar{X}=\dot{B}^{1}_{2,1,e},\qquad{L\bar{X}}=\dot{B}^{0}_{2,1,e},

see Theorem 1.4, Theorem 1.6 and Theorem 1.7. The definitions of these spaces, along with alternative characterizations, are provided in Section 2. The notations XX and LX¯{L\bar{X}} are inherited from [6], and justified by the equivalent characterization of these spaces in the next subsection.

Our analysis is done primarily at the level of (ψ,λ,α)(\psi,\lambda,\alpha), while the original problem is stated at the level of the map uu. The goal of this section is to describe the transition between the two settings, and in particular to establish the norm equivalence

(6.2) uQα,λ2XψLX¯\|u-Q^{2}_{\alpha,\lambda}\|_{X}\approx\|\psi\|_{{L\bar{X}}}

The main result of this section is the following:

Proposition 6.1.

We assume the setup from Proposition 4.5 part i). Then

i) If uQα,λ2X<\|u-Q^{2}_{\alpha,\lambda}\|_{X}<\infty, then the associated field obeys ψLX¯uQα,λ2X\|\psi\|_{{L\bar{X}}}\lesssim\|u-Q^{2}_{\alpha,\lambda}\|_{X}.

ii) Vice versa, if the field ψLX¯\psi\in{L\bar{X}}, then uQα,λ2XψLX¯\|u-Q^{2}_{\alpha,\lambda}\|_{X}\lesssim\|\psi\|_{{L\bar{X}}}.

The rest of this section is devoted to the proof of the above proposition, and is organized as follows. First we cover some basic properties of these spaces including the mapping properties of the operator LL in this context and some algebra properties. Then we introduce a companion space X~\tilde{X} which, while morally at the level of X¯\bar{X}, allows for slightly more general functions which arise in our analysis. This space is well suited for characterizing the solutions of an ODE which plays a key role in the analysis of the transfer information between uQα,λ2u-Q^{2}_{\alpha,\lambda} and ψ\psi; in fact this system was analyzed earlier in the context of energy setup, see Lemma 4.8. With all these at hand, we finish the section with the proof of the Proposition above.

6.1. An equivalent characterization of the X¯\bar{X} and LX¯{L\bar{X}} spaces

Above (and in more detail in Section 2) we have defined X¯\bar{X} and LX¯{L\bar{X}} as classical Besov spaces restricted to equivariant functions. On the other hand in our earlier work on the 11-equivariant case in [6] we have defined the counterparts of these spaces based on the spectral decomposition associated to the HλH_{\lambda} and H~λ\tilde{H}_{\lambda} operators. It is then a natural question whether the same can be done here, which we shall answer in the affirmative.

To both justify the notation LX¯{L\bar{X}} and as a starting point for the subsequent discussion, we first show that the spaces X¯\bar{X} and LX¯{L\bar{X}} can be connected via the operator LL:

Lemma 6.2.

a) The operator LL maps X¯\bar{X} to LX¯{L\bar{X}}, and

(6.3) LuLX¯uX¯\|Lu\|_{{L\bar{X}}}\lesssim\|u\|_{\bar{X}}

b) Conversely, given fLX¯f\in{L\bar{X}}, there exists a solution ww to Lw=fLw=f with

(6.4) wX¯fLX¯\|w\|_{\bar{X}}\lesssim\|f\|_{{L\bar{X}}}

This holds in particular if ww is chosen either so that w(1)=0w(1)=0, or if ww is chosen to be orthogonal to h1h_{1}.

In this result the spaces X¯\bar{X} and LX¯{L\bar{X}} are homogeneous, while the operator LλL_{\lambda} satisfies the scaling relation

Lλuλ=λ(Lu)λ.L_{\lambda}u^{\lambda}=\lambda(Lu)^{\lambda}.

Hence by rescaling one may easily replace LL by LλL_{\lambda}, without affecting the implicit constants in the estimates in the lemma.

Proof.

a) It is easily seen that

L:H˙e2H˙e1,L:L2H˙e1,L:\dot{H}^{2}_{e}\to\dot{H}^{1}_{e},\qquad L:L^{2}\to\dot{H}^{-1}_{e},

where the second can also be seen as the dual property to

L:H˙e1L2.L^{*}:\dot{H}^{1}_{e}\to L^{2}.

Then the property (6.3) directly follows from the equivalent characterization of the two norms in (2.5), (2.6).


b) Given fLX¯f\in{L\bar{X}} we split it into f=f1+f2f=f_{1}+f_{2} where the two terms are supported in r<2r<2, respectively r>1r>1. In the two regions we replace the equation

(6.5) Lw=f,Lw=f,

where

L=r+2rh3=r+2rr41r4+1,L=\partial_{r}+\frac{2}{r}h_{3}=\partial_{r}+\frac{2}{r}\frac{r^{4}-1}{r^{4}+1},

with

(r2r)w1=f1,w1([2,))=0,(\partial_{r}-\frac{2}{r})w_{1}=f_{1},\qquad w_{1}([2,\infty))=0,

respectively

(r+2r)w2=f2.w2[(0,1])=0.(\partial_{r}+\frac{2}{r})w_{2}=f_{2}.\qquad w_{2}[(0,1])=0.

Here the solutions w1,w2w_{1},w_{2} have a similar support, and are easily seen to satisfy good scale invariant Besov bounds

w1B˙2,1,e1f1B˙2,1,e0,w2B˙2,1,e1f2B˙2,1,e0.\|w_{1}\|_{\dot{B}^{1}_{2,1,e}}\lesssim\|f_{1}\|_{\dot{B}^{0}_{2,1,e}},\qquad\|w_{2}\|_{\dot{B}^{1}_{2,1,e}}\lesssim\|f_{2}\|_{\dot{B}^{0}_{2,1,e}}.

These are proved by interpolating between the corresponding H˙e1H˙e2\dot{H}^{1}_{e}\to\dot{H}^{2}_{e} and H˙e1Le2\dot{H}^{-1}_{e}\to L^{2}_{e} bounds, which in turn reduce to one dimensional Hardy type inequalities.

Then we get an approximate solution

wapp=w1+w2w_{app}=w_{1}+w_{2}

for (6.5), so that

Lwapp=f+g,Lw_{app}=f+g,

where the extra source term gg has the form

gr3w1+r5w2,g\approx r^{3}w_{1}+r^{-5}w_{2},

which has regularity gHe1g\in H^{1}_{e} and is also localized near r=1r=1.

Now it remains to solve Lz=gLz=g, where it is convenient to start with the initial condition g(1)=0g(1)=0, using the fundamental solution for LL. This yields a solution zHe2z\in H^{2}_{e}, which is much better than needed.

The solution w=wappzw=w_{app}-z which we have constructed satisfies the bound (6.4), which in particular implies the bounds

|w(1)|fLX,w,h1fLX|w(1)|\lesssim\|f\|_{LX},\qquad\langle w,h_{1}\rangle\lesssim\|f\|_{LX}

This allows us to correct ww with a well chosen multiple of h1h_{1} to insure that either w(1)=0w(1)=0 or w,h1=0\langle w,h_{1}\rangle=0.

An immediate consequence of the above Lemma and its proof is the following:

(6.6) rfLX¯+frLX¯fX¯,fX¯.\|\partial_{r}f\|_{{L\bar{X}}}+\|\frac{f}{r}\|_{{L\bar{X}}}\lesssim\|f\|_{\bar{X}},\quad\forall f\in\bar{X}.

This can be easily seen from the arguments used in the proof of part a) of the Lemma.


To set the stage for what follows, we note that throughout this paper we use the notation PkλP_{k}^{\lambda} for the Fourier projectors in the H~λ\tilde{H}_{\lambda} frame; this is because most of the analysis is carried out at the level of the gauge field ψ\psi whose dynamics uses the H~λ\tilde{H}_{\lambda} operator. This is the only section in the paper where we need to use projectors in both frames HλH_{\lambda} and H~λ\tilde{H}_{\lambda}, and we need to differentiate between them at the notation level. Thus we use the notation PkHλP_{k}^{H_{\lambda}} for the Fourier projectors in the HλH_{\lambda} frame, and PkH~λP_{k}^{\tilde{H}_{\lambda}} for the Fourier projectors in the H~λ\tilde{H}_{\lambda}. If λ=1\lambda=1, then we drop the index λ\lambda from the operators and simply use PkHP_{k}^{H} and PkH~P_{k}^{\tilde{H}} instead.

Now we are ready to state our equivalent characterizations of X¯\bar{X} and LX¯{L\bar{X}}:

Proposition 6.3.

a) The space LX¯{L\bar{X}} can be equivalently characterized as the space of functions fL2f\in L^{2} for which the following sum is finite, with equivalent norms:

(6.7) fLX¯kPkH~λfL2.\|f\|_{{L\bar{X}}}\approx\sum_{k\in\mathbb{Z}}\|P^{\tilde{H}_{\lambda}}_{k}f\|_{L^{2}}.

b) The space X¯\bar{X} can be equivalently characterized as the space of functions uH˙e1u\in\dot{H}^{1}_{e} for which the following sum is finite, with equivalent norms:

(6.8) uX¯uH˙e1+k2kPkHλuL2.\|u\|_{\bar{X}}\approx\|u\|_{\dot{H}^{1}_{e}}+\sum_{k\in\mathbb{Z}}2^{k}\|P^{H_{\lambda}}_{k}u\|_{L^{2}}.
Proof.

To start with, we note that the left hand side in both (6.7) and (6.8) does not depend on λ\lambda, but the right hand side apriori does. Hence λ\lambda only plays the role of a scaling parameter, which we can harmlessly set to λ=1\lambda=1.

We begin with the relation (6.7), for which we use the equivalent LX¯{L\bar{X}} norm given by (2.6). On the other hand for the right hand side we have a similar equivalent norm but using the H˙1\dot{H}^{1} and H˙1\dot{H}^{-1} spaces associated to the operator H~\tilde{H}.

But then (6.7) is straightforward due to the fact that the standard H˙e1\dot{H}^{1}_{e} norm and the one associated to H~\tilde{H} are equivalent (which is to say, H~\tilde{H} is coercive in H˙e1\dot{H}^{1}_{e}), and correspondingly the associated H˙1\dot{H}^{-1} norms are equivalent.

The bound (6.8) is more interesting. Using the relation

(H~PkH~Lf)(ξ)=mk(ξ)ξ(Hf)(ξ)=ξ(HPkHf)(ξ)({\mathcal{F}}^{\tilde{H}}P_{k}^{\tilde{H}}Lf)(\xi)=m_{k}(\xi)\xi({\mathcal{F}}^{H}f)(\xi)=\xi({\mathcal{F}}^{H}P_{k}^{H}f)(\xi)

and (6.7) we immediately see that the right hand side of (6.8) may be equivalently written as

uH˙e1+LuLX¯.\|u\|_{\dot{H}^{1}_{e}}+\|Lu\|_{{L\bar{X}}}.

By part (a) of Lemma 6.2 we immediately get the bound

uH˙e1+LuLX¯uX¯.\|u\|_{\dot{H}^{1}_{e}}+\|Lu\|_{{L\bar{X}}}\lesssim\|u\|_{\bar{X}}.

For the opposite inequality let wX¯w\in\bar{X} be, as in Lemma 6.2, the unique solution to Lw=LuLw=Lu which is orthogonal to h1h_{1}. Then we have L(uw)=0L(u-w)=0 which implies that uwu-w is a multiple of h1h_{1}. Hence, using part (b) of Lemma 6.2,

uX¯uwX¯+wX¯uwH˙1+wX¯uH˙1+LwLX¯,\|u\|_{\bar{X}}\lesssim\|u-w\|_{\bar{X}}+\|w\|_{\bar{X}}\approx\|u-w\|_{\dot{H}^{1}}+\|w\|_{\bar{X}}\lesssim\|u\|_{\dot{H}^{1}}+\|Lw\|_{{L\bar{X}}},

as desired. ∎

In connection with (6.8), we remark that PkHuP_{k}^{H}u is orthogonal to h1h_{1}, therefore by Lemma 6.2 we have the equivalence

PkHuX¯PkH~LuL22kPkHuL2.\|P_{k}^{H}u\|_{\bar{X}}\approx\|P_{k}^{\tilde{H}}Lu\|_{L^{2}}\approx 2^{k}\|P_{k}^{H}u\|_{L^{2}}.

This implies that for uX¯u\in\bar{X} the sum

kPkHu\sum_{k\in\mathbb{Z}}P_{k}^{H}u

converges in X¯\bar{X} and is orthogonal to h1h_{1}. This in turn yields the representation

(6.9) u=ch1+kPkHu,u=ch_{1}+\sum_{k\in\mathbb{Z}}P_{k}^{H}u,

where (6.4) can be equivalently interpreted as

(6.10) uX¯|c|+k2kPkHλfL2.\|u\|_{\bar{X}}\approx|c|+\sum_{k\in\mathbb{Z}}2^{k}\|P^{H_{\lambda}}_{k}f\|_{L^{2}}.

The above analysis highlights a key property that functions in the space X¯\bar{X} enjoy, and which functions in H˙e1\dot{H}^{1}_{e} do not. In the above we have established that for any uX¯u\in\bar{X} we are able to meaningfully project uu on h1h_{1}. This cannot be done for fH˙e1f\in\dot{H}^{1}_{e} because in the classical sense we would need to make sense of the quantity f,h1=fr,rh1\langle f,h_{1}\rangle=\langle\frac{f}{r},rh_{1}\rangle; the problem is that while frH˙e1fH˙e1\|\frac{f}{r}\|_{\dot{H}^{1}_{e}}\lesssim\|f\|_{\dot{H}^{1}_{e}}, rh1L2rh_{1}\notin L^{2} and there is no other way to fix this, that is to make sense of f,h1\langle f,h_{1}\rangle.

6.2. Spectral analysis in the Bessel frame

This is needed in order to work with classical Littlewood-Paley projectors restricted to the equivariant class of functions. We recall the definition of the Bessel function of the first kind (for instance, from page 511 of [33])

(6.11) Jm(x)=k=0(1)k(x2)m+2kk!(k+m)!,m.J_{m}(x)=\sum_{k=0}^{\infty}\frac{(-1)^{k}\left(\frac{x}{2}\right)^{m+2k}}{k!(k+m)!},\quad m\in\mathbb{N}.

From page 52-53 of [33], we recall the definition of the Hankel transform of order mm of ff:

m(f)(ξ)=0f(r)Jm(rξ)r𝑑r.\mathcal{F}_{m}(f)(\xi)=\int_{0}^{\infty}f(r)J_{m}(r\xi)rdr.

The inversion formula is

f(r)=0m(f)(ξ)Jm(rξ)ξ𝑑ξ.f(r)=\int_{0}^{\infty}\mathcal{F}_{m}(f)(\xi)J_{m}(r\xi)\xi d\xi.

Page 60 of [33] gives the L2L^{2} isometry property of m\mathcal{F}_{m}:

0f(r)g(r)r𝑑r=0m(f)(ξ)m(g)(ξ)ξ𝑑ξ.\int_{0}^{\infty}f(r)g(r)rdr=\int_{0}^{\infty}\mathcal{F}_{m}(f)(\xi)\mathcal{F}_{m}(g)(\xi)\xi d\xi.

We will also use the following asymptotic expansion of Bessel functions of the first kind (for instance, when mm\in\mathbb{N}) from page 199 of [36]

(6.12) Jm(x)(2πx)1/2(cos(xπ2mπ4)k=0ak(m)x2ksin(xπ2mπ4)k=0bk(m)x2k+1).J_{m}(x)\sim\left(\frac{2}{\pi x}\right)^{1/2}\left(\cos(x-\frac{\pi}{2}m-\frac{\pi}{4})\sum_{k=0}^{\infty}\frac{a_{k}(m)}{x^{2k}}-\sin(x-\frac{\pi}{2}m-\frac{\pi}{4})\sum_{k=0}^{\infty}\frac{b_{k}(m)}{x^{2k+1}}\right).

If fk=Pkf=21(mk2f)f_{k}=P_{k}f=\mathcal{F}_{2}^{-1}(m_{k}\mathcal{F}_{2}f) (this is the projector at frequency 2k\approx 2^{k} in the J2J_{2} frames), then we record the following pointwise bound:

Lemma 6.4.

If 2f\mathcal{F}_{2}f is localized at frequencies 2k\approx 2^{k}, then,

(6.13) |fk(r)|CfkL22k{(2kr)1/2,r2k+21(2kr)2,r2k+2<1.|f_{k}(r)|\leq C\|f_{k}\|_{L^{2}}2^{k}\begin{cases}(2^{k}r)^{-1/2},\quad r2^{k+2}\geq 1\\ (2^{k}r)^{2},\qquad r2^{k+2}<1.\end{cases}

This follows from directly estimating the definition of fkf_{k}, using the following pointwise estimates on J2J_{2} (which follow from (6.11) and (6.12)):

|J2(x)|C{x2,0<x11x,x>1|J_{2}(x)|\leq C\begin{cases}x^{2},\quad 0<x\leq 1\\ \frac{1}{\sqrt{x}},\quad x>1\end{cases}

We will also make use of the following lemma, which describes the action of r\partial_{r} (conjugated by a power of rr) on frequency localized functions.

Lemma 6.5.

Assume 2gL2\mathcal{F}_{2}g\in L^{2} is localized at ξ2k\xi\approx 2^{k}. Then we can write

g=(r+3r)g1,g=(\partial_{r}+\frac{3}{r})g^{1},

where 3g1\mathcal{F}_{3}g^{1} is localized at frequency ξ2k\xi\approx 2^{k}. In addition the following holds true

(6.14) 2kg1L2gL2,g1H˙e1gL2.2^{k}\|g^{1}\|_{L^{2}}\lesssim\|g\|_{L^{2}},\quad\|g^{1}\|_{\dot{H}^{1}_{e}}\lesssim\|g\|_{L^{2}}.
Proof.

The conclusion of the lemma can be obtained by replacing mm by 2 everywhere in the proof below (which is true for any m1m\geq 1). Using the Hankel transform of order mm, we write:

(6.15) ΔΔ1g=(r+m+1r)(rmr)(0Jm(rξ)1ξ2m(g)(ξ)ξ𝑑ξ)=(r+m+1r)(0Jm+1(rξ)1ξm(g)(ξ)ξ𝑑ξ):=(r+m+1r)g1(r)\begin{split}\Delta\Delta^{-1}g&=\left(\partial_{r}+\frac{m+1}{r}\right)\left(\partial_{r}-\frac{m}{r}\right)\left(\int_{0}^{\infty}J_{m}(r\xi)\frac{1}{-\xi^{2}}\mathcal{F}_{m}(g)(\xi)\xi d\xi\right)\\ &=\left(\partial_{r}+\frac{m+1}{r}\right)\left(\int_{0}^{\infty}J_{m+1}(r\xi)\frac{1}{\xi}\mathcal{F}_{m}(g)(\xi)\xi d\xi\right)\\ &:=\left(\partial_{r}+\frac{m+1}{r}\right)g^{1}(r)\end{split}

Taking m=2m=2 leads to our specific choice of g1g^{1}. It is also clear from the above that g1g^{1} has the support of its Hankel transform (of order 33) compactly supported in the region ξ2k\xi\approx 2^{k}. Next, using the above formulas, we have

g1L2m(g)ξL22kgL2.\|g^{1}\|_{L^{2}}\approx\|\frac{\mathcal{F}_{m}(g)}{\xi}\|_{L^{2}}\approx 2^{-k}\|g\|_{L^{2}}.

We obtain the second bound in (6.14) by noting that g1g^{1} is naturally identified with an m+1m+1-equivariant function on 2\mathbb{R}^{2}, so that we have

g1H˙e1=ξm+1(g1)L2(ξdξ)=ξ1ξm(g)(ξ)L2(ξdξ)=m(g)(ξ)L2(ξdξ)=gL2(rdr)\|g^{1}\|_{\dot{H}^{1}_{e}}=\|\xi\mathcal{F}_{m+1}(g^{1})\|_{L^{2}(\xi d\xi)}=\|\xi\cdot\frac{1}{\xi}\mathcal{F}_{m}(g)(\xi)\|_{L^{2}(\xi d\xi)}=\|\mathcal{F}_{m}(g)(\xi)\|_{L^{2}(\xi d\xi)}=\|g\|_{L^{2}(rdr)}

6.3. Algebra properties for X¯\bar{X} and LX¯{L\bar{X}}.

Here we use the Besov characterization of X¯\bar{X} and LX¯{L\bar{X}} in order to study multiplicative properties in these spaces:

Lemma 6.6.

The spaces X¯,LX¯\bar{X},{L\bar{X}} satisfy the following properties:

i) X¯\bar{X} is an algebra. Further, if f,gX¯f,g\in\bar{X} then following refined bilinear estimate holds true:

(6.16) fgX¯fH˙e1gX¯+fX¯gH˙e1;\|fg\|_{\bar{X}}\lesssim\|f\|_{\dot{H}^{1}_{e}}\|g\|_{\bar{X}}+\|f\|_{\bar{X}}\|g\|_{\dot{H}^{1}_{e}};

ii) LX¯{L\bar{X}} is stable under multiplication by functions in X¯\bar{X}; moreover, the following refined bilinear estimate holds true:

(6.17) fgLX¯fX¯gL2+fH˙e1gLX¯,fX¯,gLX¯.\|fg\|_{{L\bar{X}}}\lesssim\|f\|_{\bar{X}}\|g\|_{L^{2}}+\|f\|_{\dot{H}^{1}_{e}}\|g\|_{{L\bar{X}}},\qquad\forall f\in\bar{X},\ g\in{L\bar{X}}.
Proof.

We first prove the X¯\bar{X} algebra property (6.16). This is most easily done using equivariant extensions in 2\mathbb{R}^{2}, where we can use classical dyadic frequency localization. If for f,gf,g below we use 2-equivariant extensions then fgfg is 4-equivariant, but this is not a problem.

We have

fgX\displaystyle\|fg\|_{X}\lesssim k2kPk(fg)L2\displaystyle\ \sum_{k}2^{k}\|P_{k}(fg)\|_{L^{2}}
\displaystyle\lesssim k2k(P<kfLPkgL2+P<kgLPkfL2+2kjkPjfL2PjgL2)\displaystyle\ \sum_{k}2^{k}(\|P_{<k}f\|_{L^{\infty}}\|P_{k}g\|_{L^{2}}+\|P_{<k}g\|_{L^{\infty}}\|P_{k}f\|_{L^{2}}+2^{k}\sum_{j\geq k}\|P_{j}f\|_{L^{2}}\|P_{j}g\|_{L^{2}})
\displaystyle\lesssim fH˙1gX+fXgH˙1.\displaystyle\ \|f\|_{\dot{H}^{1}}\|g\|_{X}+\|f\|_{X}\|g\|_{\dot{H}^{1}}.

Next we prove the product estimate in (6.17), recalling the notation LXLX for the 22-equivariant two dimensional lift of LX¯{L\bar{X}} functions. We have

fgLX\displaystyle\|fg\|_{LX}\lesssim kPk(fg)L2\displaystyle\ \sum_{k}\|P_{k}(fg)\|_{L^{2}}
\displaystyle\lesssim k(P<kfLPkgL2+P<kgLPkfL2+jk2kPjfL2PjgL2)\displaystyle\ \sum_{k}(\|P_{<k}f\|_{L^{\infty}}\|P_{k}g\|_{L^{2}}+\|P_{<k}g\|_{L^{\infty}}\|P_{k}f\|_{L^{2}}+\sum_{j\geq k}2^{k}\|P_{j}f\|_{L^{2}}\|P_{j}g\|_{L^{2}})
\displaystyle\lesssim fH˙e1gLX+gL2fX.\displaystyle\ \|f\|_{\dot{H}^{1}_{e}}\|g\|_{LX}+\|g\|_{L^{2}}\|f\|_{X}.

6.4. The companion space X~\tilde{X} and an ODE result

The spaces X¯\bar{X} and LX¯{L\bar{X}} are used to measure uQXu-Q\in X, respectively ψLX¯\psi\in{L\bar{X}}, but unfortunately they cannot be used in the context of the gauge elements v,wv,w. This issue was already highlighted in the analysis in Section 4, see the statement of Proposition 4.2. In that context, we could not use the native energy space H˙1\dot{H}^{1} for the frame, but instead we used H˙C1\dot{H}^{1}_{C}.

In our context, we measure the frame elements (or better their one-dimensional reduced version) in X~\tilde{X} - this a Sobolev type companion X~\tilde{X} for X¯\bar{X}; thus it is meant to measure functions defined on (0,+)(0,+\infty). We introduce the following atoms. Given kk\in\mathbb{Z} we say that the function φ:[0,+),φC2\varphi:[0,+\infty)\rightarrow\mathbb{C},\varphi\in C^{2} (that is φ\varphi is twice continuously differentiable) is an 𝒞k2\mathcal{C}^{2}_{k} atom provided that it satisfies the following properties

1) φ\varphi is supported in the interval [0,2k+10)[0,2^{-k+10});

2) rφ(0)=0\partial_{r}\varphi(0)=0;

3) it obeys the following φ𝒞k2=α=022kαrαφL<+\|\varphi\|_{\mathcal{C}^{2}_{k}}=\sum_{\alpha=0}^{2}\|2^{-k\alpha}\partial_{r}^{\alpha}\varphi\|_{L^{\infty}}<+\infty.

It is obvious that χ<k\chi_{<-k} is an 𝒞k2\mathcal{C}^{2}_{k} atom; in fact the 𝒞k2\mathcal{C}^{2}_{k} atoms are meant to act as generalization of this basic function χ<k\chi_{<-k}. We record the following basic inequality for the atoms in 𝒞k2\mathcal{C}^{2}_{k}:

(6.18) 2krφH˙e1+2krφH˙e1φ𝒞k2;2^{k}\|\partial_{r}\varphi\|_{\dot{H}_{e}^{-1}}+2^{-k}\|\partial_{r}\varphi\|_{\dot{H}^{1}_{e}}\lesssim\|\varphi\|_{\mathcal{C}^{2}_{k}};

its proof is straightforward and left as an exercise.

We define 𝒞2\mathcal{C}^{2} to be the space of functions φ:[0,+)\varphi:[0,+\infty)\rightarrow\mathbb{C} which admit an atomic decomposition

φ=kφk,φk𝒞k2andkφk𝒞k2<;\varphi=\sum_{k\in\mathbb{Z}}\varphi_{k},\quad\varphi_{k}\in\mathcal{C}^{2}_{k}\ \mbox{and}\ \sum_{k\in\mathbb{Z}}\|\varphi_{k}\|_{\mathcal{C}^{2}_{k}}<\infty;

we endow this space with a norm in the standard fashion, that is

φ𝒞2=infrep{kφk𝒞k2;φ=kφk,φk𝒞k2},\|\varphi\|_{\mathcal{C}^{2}}=\inf_{rep}\{\sum_{k\in\mathbb{Z}}\|\varphi_{k}\|_{\mathcal{C}^{2}_{k}};\varphi=\sum_{k\in\mathbb{Z}}\varphi_{k},\varphi_{k}\in\mathcal{C}^{2}_{k}\},

where the infimum is taken over all possible representations of φ\varphi as a sum of atoms.

We record here a basic inequality:

(6.19) φLφ𝒞2,\|\varphi\|_{L^{\infty}}\lesssim\|\varphi\|_{\mathcal{C}^{2}},

which follows from the atomic structure of φ\varphi and the LL^{\infty} bounds available for its atoms.

The space X~\tilde{X} is the sum of X¯\bar{X} and 𝒞2\mathcal{C}^{2}; its precise definition is as follows. A function ff is in X~\tilde{X} if it admits a decomposition

(6.20) f=f1+f2,f1X¯andf2𝒞2;f=f_{1}+f_{2},\qquad f_{1}\in\bar{X}\ \mbox{and}\ f_{2}\in\mathcal{C}^{2};

the norm in X~\tilde{X} is defined as follows

fX~=inf{f1X¯+f2𝒞2;f=f1+f2,f1X¯andf2𝒞2}.\|f\|_{\tilde{X}}=\inf\{\|f_{1}\|_{\bar{X}}+\|f_{2}\|_{\mathcal{C}^{2}};f=f_{1}+f_{2},f_{1}\in\bar{X}\ \mbox{and}\ f_{2}\in\mathcal{C}^{2}\}.

A natural question to ask is how different are these two components. An element in φ𝒞2\varphi\in\mathcal{C}^{2} with an atomic decomposition φ=kφk,φk𝒞k2\varphi=\sum_{k}\varphi_{k},\varphi_{k}\in\mathcal{C}^{2}_{k} contains an lk1l^{1}_{k} structure just as X¯\bar{X}. The characterization of elements in 𝒞k2\mathcal{C}^{2}_{k}, shows that φk\varphi_{k} is morally at frequency 2k2^{k} and (6.18) establishes good bounds for rφkL2\|\partial_{r}\varphi_{k}\|_{L^{2}}; what is missing is the equivalent bound for φkrL2\|\frac{\varphi_{k}}{r}\|_{L^{2}} and this is why we cannot place the atoms in H˙e1\dot{H}^{1}_{e}. This failure is simply related to the fact that the atoms φk\varphi_{k} are not required to satisfy the condition φk(0)=0\varphi_{k}(0)=0; indeed note that we can easily estimate φkφk(0)rχk+10L2\|\frac{\varphi_{k}-\varphi_{k}(0)}{r}\cdot\chi_{\leq-k+10}\|_{L^{2}} and this would place φkφk(0)χk+10\varphi_{k}-\varphi_{k}(0)\cdot\chi_{\leq-k+10} in H˙e1\dot{H}^{1}_{e}. The need to augment the structure X¯\bar{X} with these atoms, and arrive at the X~\tilde{X} structure, stems from the analysis below where integrals of type r\int_{r}^{\infty} fail the basic cancellation property that 0=0\int_{0}^{\infty}=0, thus producing functions which do not have zero limit at r=0r=0, but otherwise have properties that are very similar to those of elements in X¯\bar{X}. The result in (6.22) in Lemma 6.7 highlights again how close these elements are to X¯\bar{X}, as we establish that rφLX¯\partial_{r}\varphi\in{L\bar{X}} for any φ𝒞2\varphi\in\mathcal{C}^{2}.

We will also need to bound the two components on different scales, for which we use the subset BM,mX~B^{\tilde{X}}_{M,m} of X~\tilde{X} defined as the functions which admit at least a representation of type (6.20) and for which

(6.21) f1X¯M,f1H˙e1+f2𝒞2m.\|f_{1}\|_{\bar{X}}\leq M,\quad\|f_{1}\|_{\dot{H}^{1}_{e}}+\|f_{2}\|_{\mathcal{C}^{2}}\leq m.

Our first result seeks to understand some more of the properties of functions in 𝒞2\mathcal{C}^{2}, and in particular how they interact with functions in the more classical spaces introduced earlier, such as L2,H˙e1,LX¯,X¯L^{2},\dot{H}^{1}_{e},{L\bar{X}},\bar{X}.

Lemma 6.7.

If c𝒞2c\in\mathcal{C}^{2} then the following hold true:

(6.22) rcLX¯c𝒞2.\|\partial_{r}c\|_{{L\bar{X}}}\lesssim\|c\|_{\mathcal{C}^{2}}.

In addition the following multiplicative estimates hold true

(6.23) cfH˙e1c𝒞2fH˙e1,cgL2c𝒞2gL2,\|c\cdot f\|_{\dot{H}^{1}_{e}}\lesssim\|c\|_{\mathcal{C}^{2}}\|f\|_{\dot{H}^{1}_{e}},\quad\|c\cdot g\|_{L^{2}}\lesssim\|c\|_{\mathcal{C}^{2}}\|g\|_{L^{2}},

for any fH˙e1f\in\dot{H}^{1}_{e} and gL2g\in L^{2}, and

(6.24) cfX¯c𝒞2fX¯,cgLX¯c𝒞2gLX¯,\|c\cdot f\|_{\bar{X}}\lesssim\|c\|_{\mathcal{C}^{2}}\|f\|_{\bar{X}},\quad\|c\cdot g\|_{{L\bar{X}}}\lesssim\|c\|_{\mathcal{C}^{2}}\|g\|_{{L\bar{X}}},

for any fX¯f\in\bar{X} and gLX¯g\in{L\bar{X}}.

Proof.

Throughout this argument we work with a decomposition of cc as follows

c=kck,kck𝒞k2c𝒞2.c=\sum_{k}c_{k},\quad\sum_{k}\|c_{k}\|_{\mathcal{C}^{2}_{k}}\lesssim\|c\|_{\mathcal{C}^{2}}.

The estimate (6.22) is a direct consequence of the definition of LX¯{L\bar{X}} and the estimates on rck\partial_{r}c_{k} in H˙e1\dot{H}^{-1}_{e} and H˙e1\dot{H}^{1}_{e} from (6.18).

The proof of (6.23) is straightforward; using (6.22) (which implies an L2L^{2} bound for rc\partial_{r}c) and (6.19) we obtain the first estimate in (6.23) as follows

r(cf)L2+cfrL2rcL2fL+rfL2cL+cLfrL2c𝒞2fH˙e1.\|\partial_{r}(cf)\|_{L^{2}}+\|\frac{cf}{r}\|_{L^{2}}\lesssim\|\partial_{r}c\|_{L^{2}}\|f\|_{L^{\infty}}+\|\partial_{r}f\|_{L^{2}}\|c\|_{L^{\infty}}+\|c\|_{L^{\infty}}\|\frac{f}{r}\|_{L^{2}}\lesssim\|c\|_{\mathcal{C}^{2}}\|f\|_{\dot{H}^{1}_{e}}.

The second estimate in (6.23) follows from (6.19).

The estimate (6.24) requires a bit more work. For the estimate in X¯\bar{X} it suffices to consider a single component ckfjc_{k}f_{j} where ck𝒞k2c_{k}\in\mathcal{C}^{2}_{k} and fjf_{j} is localized at frequency 2j2^{j} in the sense that its 2-equivariant extension to 2\mathbb{R}^{2} is as described in Section 2; the most important thing is that we have control on the quantity

2jfjL2+2jfjH˙e2,2^{j}\|f_{j}\|_{L^{2}}+2^{-j}\|f_{j}\|_{\dot{H}^{2}_{e}},

as highlighted in (2.5)). The term ckfjc_{k}f_{j} is essentially treated as being at frequency max(2k,2j)\max(2^{k},2^{j}). We consider two cases.

a) If kjk\leq j, then we estimate

2jckfjL2ckL2jfjL2ck𝒞k22jfjL2,2^{j}\|c_{k}f_{j}\|_{L^{2}}\lesssim\|c_{k}\|_{L^{\infty}}2^{j}\|f_{j}\|_{L^{2}}\lesssim\|c_{k}\|_{\mathcal{C}^{2}_{k}}2^{j}\|f_{j}\|_{L^{2}},

and

2jckfjH˙e2ckL2jfjH˙e2+22jr2ckL2jfjL2+2jrckLrfjL2+2jrckLfjrL2ck𝒞k2(2jfjL2+2jfjH˙e2).\begin{split}2^{-j}\|c_{k}f_{j}\|_{\dot{H}^{2}_{e}}&\lesssim\|c_{k}\|_{L^{\infty}}2^{-j}\|f_{j}\|_{\dot{H}^{2}_{e}}+2^{-2j}\|\partial_{r}^{2}c_{k}\|_{L^{\infty}}2^{j}\|f_{j}\|_{L^{2}}\\ &+2^{-j}\|\partial_{r}c_{k}\|_{L^{\infty}}\|\partial_{r}f_{j}\|_{L^{2}}+2^{-j}\|\partial_{r}c_{k}\|_{L^{\infty}}\|\frac{f_{j}}{r}\|_{L^{2}}\\ &\lesssim\|c_{k}\|_{\mathcal{C}^{2}_{k}}(2^{j}\|f_{j}\|_{L^{2}}+2^{-j}\|f_{j}\|_{\dot{H}^{2}_{e}}).\end{split}

This we conclude with the bound

2jckfjL2+2jckfjH˙e2ck𝒞k2(2jfjL2+2jfjH˙e2),2^{j}\|c_{k}f_{j}\|_{L^{2}}+2^{-j}\|c_{k}f_{j}\|_{\dot{H}^{2}_{e}}\lesssim\|c_{k}\|_{\mathcal{C}^{2}_{k}}(2^{j}\|f_{j}\|_{L^{2}}+2^{-j}\|f_{j}\|_{\dot{H}^{2}_{e}}),

which suffices in this case.

b) If kjk\geq j, on the other hand, then we estimate as follows

2kckfjL22kckL2fjLck𝒞k2fjH˙e1,2^{k}\|c_{k}f_{j}\|_{L^{2}}\lesssim 2^{k}\|c_{k}\|_{L^{2}}\|f_{j}\|_{L^{\infty}}\lesssim\|c_{k}\|_{\mathcal{C}^{2}_{k}}\|f_{j}\|_{\dot{H}^{1}_{e}},

and

2kckfjH˙e22kckLfjH˙e2+2kr2ckL2fjL+2krckLrfjL2+2krckLfjrL2ck𝒞k2(2jfjL2+2jfjH˙e2).\begin{split}2^{-k}\|c_{k}f_{j}\|_{\dot{H}^{2}_{e}}&\lesssim 2^{-k}\|c_{k}\|_{L^{\infty}}\|f_{j}\|_{\dot{H}^{2}_{e}}+2^{-k}\|\partial_{r}^{2}c_{k}\|_{L^{2}}\|f_{j}\|_{L^{\infty}}\\ &+2^{-k}\|\partial_{r}c_{k}\|_{L^{\infty}}\|\partial_{r}f_{j}\|_{L^{2}}+2^{-k}\|\partial_{r}c_{k}\|_{L^{\infty}}\|\frac{f_{j}}{r}\|_{L^{2}}\\ &\lesssim\|c_{k}\|_{\mathcal{C}^{2}_{k}}(2^{j}\|f_{j}\|_{L^{2}}+2^{-j}\|f_{j}\|_{\dot{H}^{2}_{e}}).\end{split}

From the two estimates above we obtain

2kckfjL2+2kckfjH˙e2ck𝒞k2(2jfjL2+2jfjH˙e2).2^{k}\|c_{k}f_{j}\|_{L^{2}}+2^{-k}\|c_{k}f_{j}\|_{\dot{H}^{2}_{e}}\lesssim\|c_{k}\|_{\mathcal{C}^{2}_{k}}(2^{j}\|f_{j}\|_{L^{2}}+2^{-j}\|f_{j}\|_{\dot{H}^{2}_{e}}).

The two bounds above suffice in order to conclude the proof of the first estimate in (6.24).

Similarly, for the estimate in LX¯{L\bar{X}} it suffices to consider a single component ckgjc_{k}g_{j} where ckc_{k} is an atom in 𝒞k2\mathcal{C}^{2}_{k} and gjg_{j} is just as fjf_{j} above except that it sits at the L2L^{2} level and we have control on the following quantity

2jgjH˙e1+2jgjH˙e1.2^{j}\|g_{j}\|_{\dot{H}^{-1}_{e}}+2^{-j}\|g_{j}\|_{\dot{H}^{1}_{e}}.

Just as above, this component is essentially treated as being at frequency max(2k,2j)\max(2^{k},2^{j}). We consider the same two cases:

a) If kjk\leq j, then we test

2j|ckgj,φ|=|gj,ckφ|gjH˙e1ckφH˙e12jgjH˙e1ck𝒞k2φH˙e1,2^{j}|\langle c_{k}g_{j},\varphi\rangle|=|\langle g_{j},c_{k}\varphi\rangle|\lesssim\|g_{j}\|_{\dot{H}_{e}^{-1}}\|c_{k}\varphi\|_{\dot{H}^{1}_{e}}\lesssim 2^{j}\|g_{j}\|_{\dot{H}_{e}^{-1}}\|c_{k}\|_{\mathcal{C}^{2}_{k}}\|\varphi\|_{\dot{H}^{1}_{e}},

where we have used that

ckφH˙e1ck𝒞k2φH˙e1,\|c_{k}\varphi\|_{\dot{H}^{1}_{e}}\lesssim\|c_{k}\|_{\mathcal{C}^{2}_{k}}\|\varphi\|_{\dot{H}^{1}_{e}},

which is essentially contained in (6.23). We also have

2jckgjH˙e1ckL2jgjH˙e1+2jrckLgjL2ck𝒞k2(2jgjH˙e1+2jgjH˙e1).\begin{split}2^{-j}\|c_{k}g_{j}\|_{\dot{H}^{1}_{e}}&\lesssim\|c_{k}\|_{L^{\infty}}2^{-j}\|g_{j}\|_{\dot{H}^{1}_{e}}+2^{-j}\|\partial_{r}c_{k}\|_{L^{\infty}}\|g_{j}\|_{L^{2}}\\ &\lesssim\|c_{k}\|_{\mathcal{C}^{2}_{k}}(2^{j}\|g_{j}\|_{\dot{H}^{-1}_{e}}+2^{-j}\|g_{j}\|_{\dot{H}^{1}_{e}})\end{split}.

From the above estimates we obtain

2jckgjH˙e1+2jckgjH˙e1ckCk(2jgjH˙e1+2jgjH˙e1),2^{j}\|c_{k}g_{j}\|_{\dot{H}_{e}^{-1}}+2^{-j}\|c_{k}g_{j}\|_{\dot{H}^{1}_{e}}\lesssim\|c_{k}\|_{C_{k}}(2^{j}\|g_{j}\|_{\dot{H}^{-1}_{e}}+2^{-j}\|g_{j}\|_{\dot{H}^{1}_{e}}),

which suffices.

b) If kjk\geq j, on the other hand, then we estimate

|ckgj,φ|rckLgjL2φrL22kck𝒞k2gjL2φH˙e1.|\langle c_{k}g_{j},\varphi\rangle|\lesssim\|rc_{k}\|_{L^{\infty}}\|g_{j}\|_{L^{2}}\|\frac{\varphi}{r}\|_{L^{2}}\lesssim 2^{-k}\|c_{k}\|_{\mathcal{C}^{2}_{k}}\|g_{j}\|_{L^{2}}\|\varphi\|_{\dot{H}^{1}_{e}}.

This implies

ckgjH˙e12kck𝒞k2gjL2.\|c_{k}g_{j}\|_{\dot{H}_{e}^{-1}}\lesssim 2^{-k}\|c_{k}\|_{\mathcal{C}^{2}_{k}}\|g_{j}\|_{L^{2}}.

Also, just like in the previous case, we have

2kckgjH˙e12kckLgjH˙e1+2krckLgjL2ck𝒞k2(2jgjH˙e1+2jgjH˙e1).\begin{split}2^{-k}\|c_{k}g_{j}\|_{\dot{H}^{1}_{e}}\lesssim 2^{-k}\|c_{k}\|_{L^{\infty}}\|g_{j}\|_{\dot{H}^{1}_{e}}+2^{-k}\|\partial_{r}c_{k}\|_{L^{\infty}}\|g_{j}\|_{L^{2}}\lesssim\|c_{k}\|_{\mathcal{C}^{2}_{k}}(2^{j}\|g_{j}\|_{\dot{H}^{-1}_{e}}+2^{-j}\|g_{j}\|_{\dot{H}^{1}_{e}}).\end{split}

Thus we conclude with the estimate

2kckgjH˙e1+2kckgjH˙e1ck𝒞k2(2jgjH˙e1+2jgjH˙e1).2^{k}\|c_{k}g_{j}\|_{\dot{H}^{1}_{e}}+2^{-k}\|c_{k}g_{j}\|_{\dot{H}^{1}_{e}}\lesssim\|c_{k}\|_{\mathcal{C}^{2}_{k}}(2^{j}\|g_{j}\|_{\dot{H}^{-1}_{e}}+2^{-j}\|g_{j}\|_{\dot{H}^{1}_{e}}).

The bounds above for the two cases suffice in order to conclude the proof of the second estimate in (6.24).

As a consequence of the above results we also claim the following bounds:

(6.25) h1fX¯+h3fX¯fX¯,h1gLX¯+h3gLX¯gLX¯.\|h_{1}f\|_{\bar{X}}+\|h_{3}f\|_{\bar{X}}\lesssim\|f\|_{\bar{X}},\quad\|h_{1}g\|_{{L\bar{X}}}+\|h_{3}g\|_{{L\bar{X}}}\lesssim\|g\|_{{L\bar{X}}}.

It is straightforward to check that h1X¯h_{1}\in\bar{X}, so Lemma 6.6 justifies the the estimates for the components involving multiplication with h1h_{1}. It is a straightforward exercise to check that h31=2r4+1𝒞2h_{3}-1=\frac{-2}{r^{4}+1}\in\mathcal{C}^{2} and then Lemma 6.7 justifies the estimates for the components involving multiplication with h3h_{3}; here we simply split h31=χ0(h31)+(1χ0)(h31)h_{3}-1=\chi_{\leq 0}(h_{3}-1)+(1-\chi_{\leq 0})(h_{3}-1) and notice that the first term is an atom in 𝒞02\mathcal{C}^{2}_{0}, while the second term belongs to X¯\bar{X}.

The next Lemma highlights the context in which the X~\tilde{X} structure comes in handy; the integrals considered below appear naturally when solving some ODEs in the following section.

Lemma 6.8.

Assume fX¯f\in\bar{X} and gLX¯g\in{L\bar{X}}. Then we have the following representation:

(6.26) rf(s)g(s)𝑑s=l+c,c=kZck,\int_{r}^{\infty}f(s)g(s)ds=l+c,\quad c=\sum_{k\in Z}c_{k},

where lXl\in X and ckc_{k} are supported in [0,2k+10)[0,2^{-k+10}). Moreover the following holds true:

(6.27) lX¯fX¯gL2+fH˙e1gLX¯,\|l\|_{\bar{X}}\lesssim\|f\|_{\bar{X}}\|g\|_{L^{2}}+\|f\|_{\dot{H}^{1}_{e}}\|g\|_{{L\bar{X}}},

and

(6.28) lH˙e1+kck𝒞k2fH˙e1gL2.\begin{split}\|l\|_{\dot{H}^{1}_{e}}+\sum_{k}\|c_{k}\|_{\mathcal{C}^{2}_{k}}\lesssim\|f\|_{\dot{H}^{1}_{e}}\|g\|_{L^{2}}.\end{split}

In addition rck(0)=0\partial_{r}c_{k}(0)=0, hence cc has the atomic structure in 𝒞2\mathcal{C}^{2}.

The integral rf(s)g(s)𝑑s\int_{r}^{\infty}f(s)g(s)ds appears in the process of solving an ODE in the following subsection. Our result essentially states that the integral yields a component lX¯l\in\bar{X} and a second component cc which, while not in X¯\bar{X} nor in H˙e1\dot{H}^{1}_{e}, retains some of their features, in particular rcLX¯\partial_{r}c\in{L\bar{X}} ( which is a consequence of (6.28) and (6.22)).

Proof.

We define ll by

(6.29) l=klk=klk1+lk2:=kχk(r)(rf<k(s)gk(s)𝑑s+rfk(s)gk(s)𝑑s),\begin{split}l&=\sum_{k}l_{k}=\sum_{k}l^{1}_{k}+l_{k}^{2}\\ &\ :=\sum_{k}\chi_{\geq-k}(r)\left(\int_{r}^{\infty}f_{<k}(s)g_{k}(s)ds+\int_{r}^{\infty}f_{k}(s)g_{\leq k}(s)ds\right),\end{split}

and

ck=(1χk(r))(ck1+ck2);ck1:=rf<k(s)gk(s)𝑑s,ck2:=rfk(s)gk(s)𝑑s.c_{k}=(1-\chi_{\geq-k}(r))(c^{1}_{k}+c_{k}^{2});\quad c^{1}_{k}:=\int_{r}^{\infty}f_{<k}(s)g_{k}(s)ds,\quad c_{k}^{2}:=\int_{r}^{\infty}f_{k}(s)g_{\leq k}(s)ds.

This provides a decomposition as follows:

rf(s)g(s)𝑑s=l+kck.\int_{r}^{\infty}f(s)g(s)ds=l+\sum_{k}c_{k}.

It remains to prove that this decomposition satisfies the claims in the Lemma.

Concerning ll, in order to conclude the proof of (6.27) it suffices to prove the following

(6.30) 2klk1L2+2klk1H˙e2fH˙e1gkL2,2klk2L2+2klk2H˙e2fkH˙e1gL2,2^{k}\|l^{1}_{k}\|_{L^{2}}+2^{-k}\|l^{1}_{k}\|_{\dot{H}^{2}_{e}}\lesssim\|f\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}},\quad 2^{k}\|l^{2}_{k}\|_{L^{2}}+2^{-k}\|l^{2}_{k}\|_{\dot{H}^{2}_{e}}\lesssim\|f_{k}\|_{\dot{H}^{1}_{e}}\|g\|_{L^{2}},

We start with a pointwise bound. For gkg_{k} we use Lemma 6.5 to write gk=(s+3s)gk1g_{k}=(\partial_{s}+\frac{3}{s})g_{k}^{1} and then estimate

(6.31) |rf<k(s)gk(s)𝑑s|=|rf<k(s)(s+3s)gk1(s)𝑑s|=|f<k(r)gk1(r)+r(gk1(s)f<k(s)+f<k(s)(3)sgk1(s))𝑑s||f<k(r)||gk1(r)|+0𝟙{sr}(|gk1(s)||f<k(s)|+|f<k(s)|s|gk1(s)|)𝑑s.\begin{split}&|\int_{r}^{\infty}f_{<k}(s)g_{k}(s)ds|=|\int_{r}^{\infty}f_{<k}(s)\left(\partial_{s}+\frac{3}{s}\right)g_{k}^{1}(s)ds|\\ =&\ |-f_{<k}(r)g_{k}^{1}(r)+\int_{r}^{\infty}(-g_{k}^{1}(s)f_{<k}^{\prime}(s)+f_{<k}(s)\frac{(3)}{s}g_{k}^{1}(s))ds|\\ \lesssim&\ |f_{<k}(r)||g_{k}^{1}(r)|+\int_{0}^{\infty}\mathbbm{1}_{\{s\geq r\}}(|g_{k}^{1}(s)||f_{<k}^{\prime}(s)|+\frac{|f_{<k}(s)|}{s}|g_{k}^{1}(s)|)ds.\end{split}

Therefore,

(6.32) χk(r)rf<k(s)gk(s)𝑑sL2(rdr)f<k(r)gk1(r)L2(rdr)+0𝟙{sr}L2(rdr)(|gk1(s)||f<k(s)|+|f<k(s)|s|gk1(s)|)𝑑sf<kLgk1L2+0(0sr𝑑r)1/2(|gk1(s)||f<k(s)|+|f<k(s)|s|gk1(s)|)𝑑sf<kLgk1L2+0s(|gk1(s)||f<k(s)|+|f<k(s)|s|gk1(s)|)𝑑s 2kf<kH˙e1gkL2+gk1L2(sds)f<k(s)L2(sds)+f<k(s)sL2(sds)gk1L2(sds) 2kf<kH˙e1gkL2,\begin{split}&\|\chi_{\geq-k}(r)\int_{r}^{\infty}f_{<k}(s)g_{k}(s)ds\|_{L^{2}(rdr)}\\ \lesssim&\ \|f_{<k}(r)g_{k}^{1}(r)\|_{L^{2}(rdr)}+\int_{0}^{\infty}\|\mathbbm{1}_{\{s\geq r\}}\|_{L^{2}(rdr)}(|g_{k}^{1}(s)||f_{<k}^{\prime}(s)|+\frac{|f_{<k}(s)|}{s}|g_{k}^{1}(s)|)ds\\ \lesssim&\ \|f_{<k}\|_{L^{\infty}}\|g_{k}^{1}\|_{L^{2}}+\int_{0}^{\infty}\left(\int_{0}^{s}rdr\right)^{1/2}(|g_{k}^{1}(s)||f_{<k}^{\prime}(s)|+\frac{|f_{<k}(s)|}{s}|g_{k}^{1}(s)|)ds\\ \lesssim&\ \|f_{<k}\|_{L^{\infty}}\|g_{k}^{1}\|_{L^{2}}+\int_{0}^{\infty}s(|g_{k}^{1}(s)||f_{<k}^{\prime}(s)|+\frac{|f_{<k}(s)|}{s}|g_{k}^{1}(s)|)ds\\ \lesssim&\ 2^{-k}\|f_{<k}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}}+\|g_{k}^{1}\|_{L^{2}(sds)}\|f_{<k}^{\prime}(s)\|_{L^{2}(sds)}+\|\frac{f_{<k}(s)}{s}\|_{L^{2}(sds)}\|g_{k}^{1}\|_{L^{2}(sds)}\\ \lesssim&\ 2^{-k}\|f_{<k}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}},\end{split}

where we have used the bound gk1L22kgkL2\|g_{k}^{1}\|_{L^{2}}\lesssim 2^{-k}\|g_{k}\|_{L^{2}} from Lemma 6.5. This provides the desired estimate for lkl_{k} in L2L^{2}, namely

(6.33) lk1L22kf<kH˙e1gkL2.\|l^{1}_{k}\|_{L^{2}}\lesssim 2^{-k}\|f_{<k}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}}.

We now turn to estimating lkl_{k} in H˙e2\dot{H}^{2}_{e}. From the above it follows that

(6.34) lk1r2L222klk1L22kf<kH˙e1gkL2.\|\frac{l_{k}^{1}}{r^{2}}\|_{L^{2}}\lesssim 2^{2k}\|l_{k}^{1}\|_{L^{2}}\lesssim 2^{k}\|f_{<k}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}}.

Next, we begin the estimate for the r2lk1\partial_{r}^{2}l_{k}^{1} as follows:

(6.35) r2lk1L2χ>k′′(r)22krf<k(s)gk(s)𝑑sL2+χ>k(r)2kf<k(r)gk(r)L2+χ>k(r)f<k(r)gk(r)L2+χ>k(r)f<k(r)gk(r)L2.\begin{split}\|\partial_{r}^{2}l_{k}^{1}\|_{L^{2}}\lesssim&\ \|\frac{\chi_{>-k}^{\prime\prime}(r)}{2^{-2k}}\int_{r}^{\infty}f_{<k}(s)g_{k}(s)ds\|_{L^{2}}+\|\frac{\chi_{>-k}^{\prime}(r)}{2^{-k}}\cdot f_{<k}(r)g_{k}(r)\|_{L^{2}}\\ +&\ \|\chi_{>-k}(r)f_{<k}^{\prime}(r)g_{k}(r)\|_{L^{2}}+\|\chi_{>-k}(r)f_{<k}(r)g_{k}^{\prime}(r)\|_{L^{2}}.\end{split}

Below we estimate each of the four terms on the right-hand side in the order in which they appear. From the estimates above for lk1l_{k}^{1} in L2L^{2}, it follows that

χ>k′′(r)22krf<k(s)gk(s)𝑑sL222k2kf<kH˙e1gkL2=2kf<kH˙e1gkL2.\|\frac{\chi_{>-k}^{\prime\prime}(r)}{2^{-2k}}\int_{r}^{\infty}f_{<k}(s)g_{k}(s)ds\|_{L^{2}}\lesssim 2^{-2k}2^{k}\|f_{<k}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}}=2^{k}\|f_{<k}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}}.

Next we have the straightforward estimate

χ>k2kf<kgkL22kf<kLgkL22kf<kH˙e1gkL2.\|\frac{\chi_{>-k}^{\prime}}{2^{-k}}\cdot f_{<k}g_{k}\|_{L^{2}}\lesssim 2^{-k}\|f_{<k}\|_{L^{\infty}}\|g_{k}\|_{L^{2}}\lesssim 2^{k}\|f_{<k}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}}.

A consequence of (2.4) is that gkL22kgkL2\|g_{k}^{\prime}\|_{L^{2}}\lesssim 2^{k}\|g_{k}\|_{L^{2}}, thus

χ>k(r)f<k(r)gk(r)L2f<kLgkL22kf<kH˙e1gkL2.\|\chi_{>-k}(r)f_{<k}(r)g_{k}^{\prime}(r)\|_{L^{2}}\lesssim\|f_{<k}\|_{L^{\infty}}\|g_{k}^{\prime}\|_{L^{2}}\lesssim 2^{k}\|f_{<k}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}}.

Similarly, we obtain

f<kLj<kf<kH˙e1j<kfjH˙e1j<kfjH˙e2j<k2jfjH˙e12kj<k2jkfjH˙e12k(j<kfjH˙e12)122kf<kH˙e1.\begin{split}\|f^{\prime}_{<k}\|_{L^{\infty}}&\lesssim\sum_{j<k}\|f^{\prime}_{<k}\|_{\dot{H}^{1}_{e}}\lesssim\sum_{j<k}\|f_{j}^{\prime}\|_{\dot{H}^{1}_{e}}\lesssim\sum_{j<k}\|f_{j}\|_{\dot{H}^{2}_{e}}\lesssim\sum_{j<k}2^{j}\|f_{j}\|_{\dot{H}^{1}_{e}}\\ &\lesssim 2^{k}\sum_{j<k}2^{j-k}\|f_{j}\|_{\dot{H}^{1}_{e}}\lesssim 2^{k}\left(\sum_{j<k}\|f_{j}\|_{\dot{H}^{1}_{e}}^{2}\right)^{\frac{1}{2}}\lesssim 2^{k}\|f_{<k}\|_{\dot{H}^{1}_{e}}.\end{split}

Using this we can estimate

χ>k(r)f<kgkL2f<kLgkL22kf<kH˙e1gkL2.\|\chi_{>-k}(r)f_{<k}^{\prime}g_{k}\|_{L^{2}}\lesssim\|f_{<k}^{\prime}\|_{L^{\infty}}\|g_{k}\|_{L^{2}}\lesssim 2^{k}\|f_{<k}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}}.

As a consequence of all the bounds above we obtain

(6.36) r2lk1L22kf<kH˙e1gkL2.\|\partial_{r}^{2}l_{k}^{1}\|_{L^{2}}\lesssim 2^{k}\|f_{<k}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}}.

We also claim that

(6.37) rlk1rL22kf<kH˙e1gkL2;\|\frac{\partial_{r}l_{k}^{1}}{r}\|_{L^{2}}\lesssim 2^{k}\|f_{<k}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}};

the argument is similar to the one provided for (6.36) and the details are left as an exercise.

From (6.34), (6.36) and (6.37) we obtain that

lk1H˙e22kf<kH˙e1gkL2.\|l^{1}_{k}\|_{\dot{H}^{2}_{e}}\lesssim 2^{k}\|f_{<k}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}}.

Together with (6.33) this provides the correct contribution to (6.30) for lk1l^{1}_{k}.

The bound for lk2l_{k}^{2} is similar but simpler since the integration by parts is not needed. We start with

(6.38) |rfk(s)gk(s)𝑑s|0𝟙{sr}|fk(s)||gk(s)|𝑑s.\begin{split}\left|\int_{r}^{\infty}f_{k}(s)g_{\leq k}(s)ds\right|\lesssim\int_{0}^{\infty}\mathbbm{1}_{\{s\geq r\}}|f_{k}(s)||g_{\leq k}(s)|ds.\end{split}

Based on this we estimate

lk2L2(rdr)0s|fk(s)||gk(s)|𝑑sfkL2g<kL22kfkH˙e1gkL2,\|l_{k}^{2}\|_{L^{2}(rdr)}\lesssim\int_{0}^{\infty}s|f_{k}(s)||g_{\leq k}(s)|ds\lesssim\|f_{k}\|_{L^{2}}\|g_{<k}\|_{L^{2}}\lesssim 2^{-k}\|f_{k}\|_{\dot{H}^{1}_{e}}\|g_{\leq k}\|_{L^{2}},

as needed for the L2L^{2} bound for lk2l_{k}^{2} in (6.30). From this it follows that

(6.39) lk2r2L222klk2L22kfkH˙e1gkL2.\|\frac{l_{k}^{2}}{r^{2}}\|_{L^{2}}\lesssim 2^{2k}\|l_{k}^{2}\|_{L^{2}}\lesssim 2^{k}\|f_{k}\|_{\dot{H}^{1}_{e}}\|g_{\leq k}\|_{L^{2}}.

Finally, the estimate

(6.40) r2lk2L2+rlk2r2kfkH˙e1gkL2,\|\partial_{r}^{2}l_{k}^{2}\|_{L^{2}}+\|\frac{\partial_{r}l^{2}_{k}}{r}\|\lesssim 2^{k}\|f_{k}\|_{\dot{H}^{1}_{e}}\|g_{\leq k}\|_{L^{2}},

follows in a similar manner, using the same steps as in the corresponding estimate for lk1l^{1}_{k}.

We now turn our attention to the ckc_{k} terms and prove (6.28). For ck1c_{k}^{1} term we use the same integration by parts used earlier for lk1l_{k}^{1},

(6.41) ck1=j<krfj(s)gk(s)𝑑s=j<k(fj(r)gk1(r)r(gk1(s)fj(s)fj(s)3sgk1(s))𝑑s).\begin{split}c_{k}^{1}=\sum_{j<k}\int_{r}^{\infty}f_{j}(s)g_{k}(s)ds=\sum_{j<k}\left(-f_{j}(r)g_{k}^{1}(r)-\int_{r}^{\infty}\left(g_{k}^{1}(s)f_{j}^{\prime}(s)-f_{j}(s)\frac{3}{s}g_{k}^{1}(s)\right)ds\right).\end{split}

From (6.13) we have

|fj(r)|(2jr)2fjH˙e1;|f_{j}(r)|\lesssim(2^{j}r)^{2}\|f_{j}\|_{\dot{H}^{1}_{e}};

which allows us to bound the first term by

(6.42) |fj(r)gk1(r)|(2jr)2fjH˙e1gn1H˙e122(jk)fjH˙e1gkL2,r2k.\begin{split}|f_{j}(r)g_{k}^{1}(r)|\lesssim(2^{j}r)^{2}\|f_{j}\|_{\dot{H}^{1}_{e}}\|g_{n}^{1}\|_{\dot{H}^{1}_{e}}\lesssim 2^{2(j-k)}\|f_{j}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}},\quad r\lesssim 2^{-k}.\end{split}

We recall from Section 2 that

(6.43) s2fjL2+sfj(s)sL2+fj(s)s2L22jfjH˙e1.\|\partial_{s}^{2}f_{j}\|_{L^{2}}+\|\frac{\partial_{s}f_{j}(s)}{s}\|_{L^{2}}+\|\frac{f_{j}(s)}{s^{2}}\|_{L^{2}}\lesssim 2^{j}\|f_{j}\|_{\dot{H}^{1}_{e}}.

Using the two estimates above we obtain the following for r2kr\lesssim 2^{-k},

(6.44) |ck1(r)|j<k(|fj(r)||gk1(r)|+r|gk1(s)|(|fj(s)|s+|fj(s)|s2)s𝑑s)j<k(|fj(r)||gk1(r)|+gk1(s)L2(fj(s)sL2+fj(s)s2L2))j<k2jkfjH˙e1gkL2.\begin{split}|c_{k}^{1}(r)|&\lesssim\sum_{j<k}\left(|f_{j}(r)||g_{k}^{1}(r)|+\int_{r}^{\infty}|g_{k}^{1}(s)|\left(\frac{|f_{j}^{\prime}(s)|}{s}+\frac{|f_{j}(s)|}{s^{2}}\right)sds\right)\\ &\lesssim\sum_{j<k}\left(|f_{j}(r)||g_{k}^{1}(r)|+\|g_{k}^{1}(s)\|_{L^{2}}\left(\|\frac{f_{j}^{\prime}(s)}{s}\|_{L^{2}}+\|\frac{f_{j}(s)}{s^{2}}\|_{L^{2}}\right)\right)\\ &\lesssim\sum_{j<k}2^{j-k}\|f_{j}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}}.\end{split}

For ck2c_{k}^{2} we estimate as follows

(6.45) |ck2(r)|jkr|fk(s)||gj(s)|𝑑sjkr|fk(s)gj(s)s|s𝑑sjkfkL2gjsL2jk2jkfkH˙e1gjL2,\begin{split}|c_{k}^{2}(r)|&\lesssim\sum_{j\leq k}\int_{r}^{\infty}|f_{k}(s)|\cdot|g_{j}(s)|ds\lesssim\sum_{j\leq k}\int_{r}^{\infty}|f_{k}(s)\frac{g_{j}(s)}{s}|sds\\ &\lesssim\sum_{j\leq k}\|f_{k}\|_{L^{2}}\|\frac{g_{j}}{s}\|_{L^{2}}\lesssim\sum_{j\leq k}2^{j-k}\|f_{k}\|_{\dot{H}^{1}_{e}}\|g_{j}\|_{L^{2}},\end{split}

where we have used the bounds

fkL22kfkH˙e1,gj(s)sL2gjH˙e12jgjL2\|f_{k}\|_{L^{2}}\lesssim 2^{-k}\|f_{k}\|_{\dot{H}^{1}_{e}},\quad\|\frac{g_{j}(s)}{s}\|_{L^{2}}\lesssim\|g_{j}\|_{\dot{H}^{1}_{e}}\lesssim 2^{j}\|g_{j}\|_{L^{2}}

from Section 2. We combine the estimates above to conclude that

(6.46) kckL(r2k+10)kck1L+ck3Lkjk2jk(fkH˙e1gjL2+fjH˙e1gkL2)(kfkH˙e12)12(kgkL22)12fH˙e1gL2;\begin{split}\sum_{k}\|c_{k}\|_{L^{\infty}(r\leq 2^{-k+10})}&\lesssim\sum_{k}\|c_{k}^{1}\|_{L^{\infty}}+\|c_{k}^{3}\|_{L^{\infty}}\lesssim\sum_{k}\sum_{j\leq k}2^{j-k}(\|f_{k}\|_{\dot{H}^{1}_{e}}\|g_{j}\|_{L^{2}}+\|f_{j}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{L^{2}})\\ &\lesssim\left(\sum_{k}\|f_{k}\|_{\dot{H}^{1}_{e}}^{2}\right)^{\frac{1}{2}}\left(\sum_{k}\|g_{k}\|^{2}_{L^{2}}\right)^{\frac{1}{2}}\lesssim\|f\|_{\dot{H}^{1}_{e}}\|g\|_{L^{2}};\end{split}

Next, we compute

rck1=f<kgk=j<kfjgk,\partial_{r}c_{k}^{1}=-f_{<k}g_{k}=-\sum_{j<k}f_{j}g_{k},

and estimate, for r2kr\lesssim 2^{-k},

|rck1(r)|j<k|fj(r)||gk(r)|j<k(2jr)2fjH˙e1gkH˙e1j<k22(jk)fjH˙e12kgkL2.|\partial_{r}c_{k}^{1}(r)|\lesssim\sum_{j<k}|f_{j}(r)||g_{k}(r)|\lesssim\sum_{j<k}(2^{j}r)^{2}\|f_{j}\|_{\dot{H}^{1}_{e}}\|g_{k}\|_{\dot{H}^{1}_{e}}\lesssim\sum_{j<k}2^{2(j-k)}\|f_{j}\|_{\dot{H}^{1}_{e}}2^{k}\|g_{k}\|_{L^{2}}.

where we have used again (6.13). In a similar way

rck2=fkgk=jkfkgj,\partial_{r}c_{k}^{2}=-f_{k}g_{\leq k}=-\sum_{j\leq k}f_{k}g_{j},

and

|rck2(r)|jk|fk(r)||gj(r)|jk2j(2jr)2gjL2fkH˙e12kjk22(jk)gjL2fkH˙e1.|\partial_{r}c_{k}^{2}(r)|\lesssim\sum_{j\leq k}|f_{k}(r)||g_{j}(r)|\lesssim\sum_{j\leq k}2^{j}(2^{j}r)^{2}\|g_{j}\|_{L^{2}}\|f_{k}\|_{\dot{H}^{1}_{e}}\lesssim 2^{k}\sum_{j\leq k}2^{2(j-k)}\|g_{j}\|_{L^{2}}\|f_{k}\|_{\dot{H}^{1}_{e}}.

Based on the two estimates above we conclude with the following estimates

k2krckL(r2k+10)fH˙e1gL2;\sum_{k}2^{-k}\|\partial_{r}c_{k}\|_{L^{\infty}(r\leq 2^{-k+10})}\lesssim\|f\|_{\dot{H}^{1}_{e}}\|g\|_{L^{2}};

we note that the bound above are stable under multiplication by 1χk1-\chi_{\geq-k} (recall that this is how ckc_{k} is obtained from ck1+ck2c_{k}^{1}+c_{k}^{2}).

Finally we can differentiate cki,i=1,2c_{k}^{i},i=1,2 twice, argue as above and conclude with

k22kr2ckL((r2k+10)fH˙e1gL2.\sum_{k}2^{-2k}\|\partial^{2}_{r}c_{k}\|_{L^{\infty}((r\leq 2^{-k+10})}\lesssim\|f\|_{\dot{H}^{1}_{e}}\|g\|_{L^{2}}.

We also have the obvious properties that ckc_{k} is supported in [0,2k+10)[0,2^{-k+10}) and rck(0)=0\partial_{r}c_{k}(0)=0 is obvious from its definition and the fact that fk(0)=f<k(0)=0f_{k}(0)=f_{<k}(0)=0 since they are elements in H˙e1\dot{H}^{1}_{e}.

This completes the proof of the claims made for cc, and in turn it concludes the proof of the Lemma.

Using these spaces we now revisit the result in Lemma 4.8 and provide a similar result in the context of the refined structures X~\tilde{X}.

Lemma 6.9.

Consider the vector valued ODE

(6.47) rZ=NZ+N,limrZ(r)=0,\partial_{r}Z=NZ+N,\qquad\lim_{r\rightarrow\infty}Z(r)=0,

where the coefficients NN have the form

N=rH+F,HX¯,Fij=lfinitefl,ijgl,ijN=\partial_{r}H+F,\qquad H\in\bar{X},\quad F_{ij}=\sum_{l}^{finite}f_{l,ij}g_{l,ij}

with entries satisfying the bounds

HX¯+lgl,ijLX¯M,lfl,ijX¯C+M,\|H\|_{\bar{X}}+\sum_{l}\|g_{l,ij}\|_{{L\bar{X}}}\leq M,\quad\sum_{l}\|f_{l,ij}\|_{\bar{X}}\leq C+M,
HH˙e1+lgl,ijL2m,lfl,ijH˙e1C+m,\|H\|_{\dot{H}^{1}_{e}}+\sum_{l}\|g_{l,ij}\|_{L^{2}}\leq m,\quad\sum_{l}\|f_{l,ij}\|_{\dot{H}^{1}_{e}}\leq C+m,

where CC is a universal constant. Then, assuming that mm is small enough222The smallness of mm is universal and independent on MM, the above equation has a unique solution in B2(C+M),2mX~B^{\tilde{X}}_{2(C+M),2m} (see Definition 6.21). Furthermore, the map from the coefficients NN in the norms above to ZX~Z\in\tilde{X} is analytic.

We make several remarks concerning this result:

  1. (i)

    The statement above can be easily adapted to systems of type rZ=ZN+N\partial_{r}Z=ZN+N.

  2. (ii)

    The statement above (including the one made in Remark 1 above) can be generalized to systems of the form rZ=N1Z+N2\partial_{r}Z=N_{1}Z+N_{2} (or rZ=ZN1+N2\partial_{r}Z=ZN_{1}+N_{2}), where N1,N2N_{1},N_{2} satisfy similar bounds as NN.

  3. (iii)

    The universal constant CC may be very well thought of as C=1C=1; in practice though it depends on h1h_{1} and h3h_{3} and it is larger than 11, but it is independent on MM used in the statement.

Proof.

The system (6.47)is iterated as follows:

rZ1=N,Z1()=0\partial_{r}Z_{1}=N,\quad Z_{1}(\infty)=0

and

rZn+1=NZn,Zn+1()=0,\partial_{r}Z_{n+1}=NZ_{n},\quad Z_{n+1}(\infty)=0,

where the final solution is Z=n=1ZnZ=\sum_{n=1}^{\infty}Z_{n}, provided we establish the convergence of this series in X~\tilde{X}. We note that the first iteration is given by

Z1=H(r)rF(s)𝑑s=H+L1+C1,Z_{1}=H(r)-\int_{r}^{\infty}F(s)ds=H+L_{1}+C_{1},

where L1L_{1} and C1C_{1} are given by Lemma 6.8. In particular

L1H˙e1m,L1X¯(1+m)(1+M),C1𝒞2m(1+m),\|L_{1}\|_{\dot{H}^{1}_{e}}\lesssim m,\quad\|L_{1}\|_{\bar{X}}\lesssim(1+m)(1+M),\quad\|C_{1}\|_{\mathcal{C}^{2}}\lesssim m(1+m),

where the 1+m1+m factors can be omitted as m1m\ll 1. We note that although the estimate for L1X¯\|L_{1}\|_{\bar{X}} does not yet contain a factor of mm (this is the small parameter), the bound for L1H˙e1\|L_{1}\|_{\dot{H}^{1}_{e}} does so; the point is that all further iterates will add factors of mm in the estimates in both H˙e1\dot{H}^{1}_{e} and X¯\bar{X}.

Next, we solve for the second iterate

Z2(r)=r(F(s)+sH(s))(H(s)+L1(s)+C1(s))𝑑s=r(F(s)+sH(s))(H(s)+L1(s))𝑑srF(s)C1(s)𝑑s+H(r)C1(r)+rH(s)sC1(s)ds.\begin{split}Z_{2}(r)&=-\int_{r}^{\infty}(F(s)+\partial_{s}H(s))(H(s)+L_{1}(s)+C_{1}(s))ds\\ &=-\int_{r}^{\infty}(F(s)+\partial_{s}H(s))(H(s)+L_{1}(s))ds\\ &-\int_{r}^{\infty}F(s)C_{1}(s)ds+H(r)C_{1}(r)+\int_{r}^{\infty}H(s)\partial_{s}C_{1}(s)ds.\end{split}

Based on the bounds for the components of Z1Z_{1}, by the results in Lemma 6.7 and Lemma 6.6 we obtain the following representation:

Z2=L2+C2,Z_{2}=L_{2}+C_{2},
L2H˙e1m2,L2X¯m(1+m)(1+M),C2𝒞2m2(1+m).\|L_{2}\|_{\dot{H}^{1}_{e}}\lesssim m^{2},\quad\|L_{2}\|_{\bar{X}}\lesssim m(1+m)(1+M),\quad\|C_{2}\|_{\mathcal{C}^{2}}\lesssim m^{2}(1+m).

One notices that at this point even the estimate for L2X¯\|L_{2}\|_{\bar{X}} contains a factor of mm. For general nn we obtain a similar representation Zn=Ln+CnZ_{n}=L_{n}+C_{n} obeying the following bounds

LnH˙e1Dnmn,LnX¯Dnmn1(1+M),Cn𝒞2Dnmn.\|L_{n}\|_{\dot{H}^{1}_{e}}\leq D^{n}m^{n},\quad\|L_{n}\|_{\bar{X}}\leq D^{n}m^{n-1}(1+M),\quad\|C_{n}\|_{\mathcal{C}^{2}}\leq D^{n}m^{n}.

Here DD is the universal constant that appears in the use of \lesssim above. It is clear that if mm is sufficiently small, depending on DD, but not on MM, then the above iteration scheme converges.

At this point we are ready to start proving Proposition 6.1 (i). To keep things streamlined, it is preferable to work with the reduced map u¯\bar{u} and the corresponding reduced gauge v¯,w¯\bar{v},\bar{w} (recall that u=emθRu¯(r)u=e^{m\theta R}\bar{u}(r) and similarly for v,wv,w), since these are functions of rr only.

In this context, it suffices to show that the differentiated field ψ\psi satisfies the bound

(6.48) ψLX¯u¯Q¯α,λ2X¯.\|\psi\|_{{L\bar{X}}}\lesssim\|\bar{u}-\bar{Q}^{2}_{\alpha,\lambda}\|_{\bar{X}}.

This is done in two stages: first we transfer information from u¯\bar{u} to the gauge elements v¯,w¯\bar{v},\bar{w} and then we transfer all the information we have (including the one on the gauge elements) to ψ\psi. By scaling and rotation we can assume the parameter choice α=0,λ=1\alpha=0,\lambda=1, and drop the indices for Q2,V2,W2Q^{2},V^{2},W^{2} in the arguments that follow.

6.5. The transition from uu to (v,w)(v,w)

Our aim in this first step is to show that we have

(6.49) v¯V¯X~+w¯W¯X~u¯Q¯X¯,\|\bar{v}-\bar{V}\|_{\tilde{X}}+\|\bar{w}-\bar{W}\|_{\tilde{X}}\lesssim\|\bar{u}-\bar{Q}\|_{\bar{X}},

as well as

(6.50) δψ2X~+δA2X~u¯Q¯X¯.\|\delta\psi_{2}\|_{\tilde{X}}+\|\delta A_{2}\|_{\tilde{X}}\lesssim\|\bar{u}-\bar{Q}\|_{\bar{X}}.

We use the equation (4.10) for the matrix 𝒪=(v¯,w¯,u¯)\mathcal{O}=(\bar{v},\bar{w},\bar{u}), namely

(6.51) r𝒪=M(u¯)𝒪,𝒪()=I3,M(u¯)=ru¯u¯.\partial_{r}\mathcal{O}=M(\bar{u})\mathcal{O},\qquad\mathcal{O}(\infty)=I_{3},\qquad M(\bar{u})=\partial_{r}\bar{u}\wedge\bar{u}.

If u=Qu=Q then M(u¯)M(\bar{u}) has the form

(6.52) M(Q¯)=2(00h1r000h1r00).M(\bar{Q})=2\left(\begin{array}[]{lll}0&0&-\frac{h_{1}}{r}\\ 0&0&0\\ \frac{h_{1}}{r}&0&0\end{array}\right).

We start with the following identity:

M(u¯)M(Q¯)=r(u¯Q¯)(u¯Q¯)+2r(u¯Q¯)Q¯+r(Q¯(u¯Q¯))=r(u¯Q¯)(u¯Q¯)+r(Q¯(u¯Q¯))2(u¯Q¯)rQ¯=r(u¯Q¯+2Q¯)(u¯Q¯)+r(Q¯(u¯Q¯)).\begin{split}M(\bar{u})-M(\bar{Q})&=\partial_{r}(\bar{u}-\bar{Q})\wedge(\bar{u}-\bar{Q})+2\partial_{r}(\bar{u}-\bar{Q})\wedge\bar{Q}+\partial_{r}\left(\bar{Q}\wedge(\bar{u}-\bar{Q})\right)\\ &=\partial_{r}(\bar{u}-\bar{Q})\wedge(\bar{u}-\bar{Q})+\partial_{r}\left(\bar{Q}\wedge(\bar{u}-\bar{Q})\right)-2(\bar{u}-\bar{Q})\wedge\partial_{r}\bar{Q}\\ &=\partial_{r}(\bar{u}-\bar{Q}+2\bar{Q})\wedge(\bar{u}-\bar{Q})+\partial_{r}\left(\bar{Q}\wedge(\bar{u}-\bar{Q})\right).\end{split}

We notice that we can write

(6.53) M(u¯)M(Q¯)=rH+F,M(\bar{u})-M(\bar{Q})=\partial_{r}H+F,

where

HX¯u¯Q¯X¯,HH˙e1u¯Q¯H˙e1,\|H\|_{\bar{X}}\lesssim\|\bar{u}-\bar{Q}\|_{\bar{X}},\quad\|H\|_{\dot{H}^{1}_{e}}\lesssim\|\bar{u}-\bar{Q}\|_{\dot{H}^{1}_{e}},

and the entries in FF are of the form lfl,ijgl,ij,i,j=1,3\sum_{l}f_{l,ij}g_{l,ij},i,j=1,3 (l=2l=2 in this case) and, in view of (6.25), obey the estimates

lfl,ijH˙e11+u¯Q¯H˙e1,lfl,ijX¯1+u¯Q¯X¯,\sum_{l}\|f_{l,ij}\|_{\dot{H}^{1}_{e}}\lesssim 1+\|\bar{u}-\bar{Q}\|_{\dot{H}^{1}_{e}},\quad\sum_{l}\|f_{l,ij}\|_{\bar{X}}\lesssim 1+\|\bar{u}-\bar{Q}\|_{\bar{X}},
lgl,ijL2u¯Q¯H˙e1,lgl,ijX¯u¯Q¯X¯.\sum_{l}\|g_{l,ij}\|_{L^{2}}\lesssim\|\bar{u}-\bar{Q}\|_{\dot{H}^{1}_{e}},\quad\sum_{l}\|g_{l,ij}\|_{\bar{X}}\lesssim\|\bar{u}-\bar{Q}\|_{\bar{X}}.

Returning to (6.51), we start with the solution 𝒪0\mathcal{O}_{0} for the case u=Qu=Q, which is given by

(6.54) 𝒪0=(h30h1010h10h3),𝒪01=𝒪0t.\mathcal{O}_{0}=\left(\begin{array}[]{ccc}h_{3}&0&h_{1}\\ 0&1&0\\ -h_{1}&0&h_{3}\end{array}\right),\qquad\mathcal{O}_{0}^{-1}=\mathcal{O}_{0}^{t}.

Then we express the solution to (6.51) in the form

(6.55) 𝒪(r)=𝒪0(r)(I+Y(r)),\mathcal{O}(r)=\mathcal{O}_{0}(r)(I+Y(r)),

where YY solves the differential equation

(6.56) rY=NY+N,Y()=0N=𝒪01(M(u¯)M(Q¯))𝒪0.\partial_{r}Y=NY+N,\qquad Y(\infty)=0\qquad N=\mathcal{O}_{0}^{-1}(M(\bar{u})-M(\bar{Q}))\mathcal{O}_{0}.

By invoking (6.25), it is clear that NN has a similar representation as in Lemma 6.9 with mu¯Q¯H˙e1m\approx\|\bar{u}-\bar{Q}\|_{\dot{H}^{1}_{e}} and Mu¯Q¯X¯M\approx\|\bar{u}-\bar{Q}\|_{\bar{X}}. Thus we obtain a solution YY for this system satisfying Y=L+CY=L+C with LH˙e1u¯Q¯H˙e1,LX¯u¯Q¯X¯,C𝒞2u¯Q¯H˙e1\|L\|_{\dot{H}^{1}_{e}}\lesssim\|\bar{u}-\bar{Q}\|_{\dot{H}^{1}_{e}},\|L\|_{\bar{X}}\lesssim\|\bar{u}-\bar{Q}\|_{\bar{X}},\|C\|_{\mathcal{C}^{2}}\lesssim\|\bar{u}-\bar{Q}\|_{\dot{H}^{1}_{e}}. This information is then easily transferred to v¯V¯,w¯W¯\bar{v}-\bar{V},\bar{w}-\bar{W}, and (6.49) follows.

It is useful to note that these are precisely the same type of bounds that have been used in the proof of (4.35), see the analysis of (4.51) and the particular setup in (4.55). Here we essentially upgrade that theory from the H˙e1\dot{H}^{1}_{e} framework to the X¯\bar{X} framework.

We can further improve the bounds for v¯3,w¯3\bar{v}_{3},\bar{w}_{3}; for instance

w¯3=u¯1v¯2u¯2v¯1=(u¯1h1)v¯2+h1v¯2u¯2v¯1.\bar{w}_{3}=\bar{u}_{1}\bar{v}_{2}-\bar{u}_{2}\bar{v}_{1}=(\bar{u}_{1}-h_{1})\bar{v}_{2}+h_{1}\bar{v}_{2}-\bar{u}_{2}\bar{v}_{1}.

We have just proved that v¯2=f+c\bar{v}_{2}=f+c where fX¯u¯Q¯X¯\|f\|_{\bar{X}}\lesssim\|\bar{u}-\bar{Q}\|_{\bar{X}} and fH˙e1+cu¯Q¯H˙e1\|f\|_{\dot{H}^{1}_{e}}+\|c\|\lesssim\|\bar{u}-\bar{Q}\|_{\dot{H}^{1}_{e}}; also we have the trivial estimate u¯1h1X¯u¯Q¯X¯\|\bar{u}_{1}-h_{1}\|_{\bar{X}}\lesssim\|\bar{u}-\bar{Q}\|_{\bar{X}} and u1h1H˙e1u¯Q¯H˙e1\|u_{1}-h_{1}\|_{\dot{H}^{1}_{e}}\lesssim\|\bar{u}-\bar{Q}\|_{\dot{H}^{1}_{e}}. Then using (6.16) we obtain

(u¯1h1)fX¯u¯Q¯X¯u¯Q¯H˙e1,\|(\bar{u}_{1}-h_{1})f\|_{\bar{X}}\lesssim\|\bar{u}-\bar{Q}\|_{\bar{X}}\|\bar{u}-\bar{Q}\|_{\dot{H}^{1}_{e}},

while using (6.24) gives

(u¯1h1)cX¯u¯Q¯X¯u¯Q¯H˙e1.\|(\bar{u}_{1}-h_{1})c\|_{\bar{X}}\lesssim\|\bar{u}-\bar{Q}\|_{\bar{X}}\|\bar{u}-\bar{Q}\|_{\dot{H}^{1}_{e}}.

In a similar manner, using the fact that h1Xh_{1}\in X, we obtain

h1v¯2X¯u¯Q¯X¯.\|h_{1}\bar{v}_{2}\|_{\bar{X}}\lesssim\|\bar{u}-\bar{Q}\|_{\bar{X}}.

The term u¯2v¯1\bar{u}_{2}\bar{v}_{1} is estimated in the same fashion and the final conclusion of this analysis is w¯3Xu¯Q¯X\|\bar{w}_{3}\|_{X}\lesssim\|\bar{u}-\bar{Q}\|_{X}. A similar analysis shows that v¯3+h1X¯u¯Q¯X¯\|\bar{v}_{3}+h_{1}\|_{\bar{X}}\lesssim\|\bar{u}-\bar{Q}\|_{\bar{X}}; in particular this bound implies that δψ2X¯u¯Q¯X¯\|\delta\psi_{2}\|_{\bar{X}}\lesssim\|\bar{u}-\bar{Q}\|_{\bar{X}}; since A2=2u3A_{2}=2u_{3}, we also automatically obtain δA2X¯u¯Q¯X¯\|\delta A_{2}\|_{\bar{X}}\lesssim\|\bar{u}-\bar{Q}\|_{\bar{X}}, concluding the proof of (6.50).

6.6. The transition from (u,v,w)(u,v,w) to ψ\psi

Here we consider ψ\psi, which is represented as

ψ=𝒲¯v¯+i𝒲¯w¯,𝒲¯=ru¯1ru¯×Ru¯.\psi=\overline{\mathcal{W}}\cdot\bar{v}+i\overline{\mathcal{W}}\cdot\bar{w},\qquad\overline{\mathcal{W}}=\partial_{r}\bar{u}-\frac{1}{r}\bar{u}\times R\bar{u}.

The structure for (v¯,w¯)(\bar{v},\bar{w}), including the representation in the space X~\tilde{X}, and the multiplicative estimates for LX¯{L\bar{X}}, that is (6.17), (6.25) and (6.24), show that, in order to complete the proof of the result in part i) of Proposition 6.1, precisely the bound (6.48), it suffices to establish that

(6.57) 𝒲¯LX¯u¯Q¯X¯.\|\overline{\mathcal{W}}\|_{{L\bar{X}}}\lesssim\|\bar{u}-\bar{Q}\|_{\bar{X}}.

Since 𝒲¯\overline{\mathcal{W}} vanishes if u=Qu=Q, we can write

𝒲¯=r(u¯Q¯)1r(u¯Q¯)×R(u¯Q¯)1rQ¯×R(u¯Q¯)1r(u¯Q¯)×RQ¯=L(u¯Q¯)1r(u¯Q¯)×R(u¯Q¯)+𝒲~.\begin{split}\overline{\mathcal{W}}\!=&\partial_{r}(\bar{u}-\bar{Q})-\frac{1}{r}(\bar{u}-\bar{Q})\times\!R(\bar{u}-\bar{Q})-\frac{1}{r}\bar{Q}\times R(\bar{u}-\bar{Q})-\frac{1}{r}(\bar{u}-\bar{Q})\times R\bar{Q}\\ =&L(\bar{u}-\bar{Q})-\frac{1}{r}(\bar{u}-\bar{Q})\times R(\bar{u}-\bar{Q})+\tilde{\mathcal{W}}.\end{split}

The first term is in LX¯{L\bar{X}} by definition and the second belongs to LX¯{L\bar{X}} by using (6.6) and (6.17). It remains to consider the last component

𝒲~=h3r(u¯Q¯)1rQ¯×R(u¯Q¯)1r(u¯Q¯)×RQ¯\tilde{\mathcal{W}}=-\frac{h_{3}}{r}(\bar{u}-\bar{Q})-\frac{1}{r}\bar{Q}\times R(\bar{u}-\bar{Q})-\frac{1}{r}(\bar{u}-\bar{Q})\times R\bar{Q}

This is easily estimated in LX¯{L\bar{X}} using (6.6) and (6.25). This concludes the proof of (6.48), completing the argument for part i) of Proposition 6.1.

6.7. The transition from ψ\psi to uu

In this subsection we establish part ii) of Proposition 6.1, that is recovering information on the map uu from information on the differentiated field ψ\psi. Since the value of α\alpha plays no particular role in the arguments and λ\lambda is just a scaling parameter, we do not restrict the generality of the argument by assuming α=0\alpha=0 and λ=1\lambda=1; also, to keep notation compact, we simply write δψ2=δ1,0ψ2\delta\psi_{2}=\delta^{1,0}\psi_{2} and δA2=δ1A2\delta A_{2}=\delta^{1}A_{2}.

Our goal here is to establish the following

(6.58) u¯Q¯X¯ψLX¯.\|\bar{u}-\bar{Q}\|_{\bar{X}}\lesssim\|\psi\|_{{L\bar{X}}}.

This is also done in two steps: first we transfer the LX¯{L\bar{X}} information from ψ\psi to information at the level of X¯\bar{X} for δψ2,δA2\delta\psi_{2},\delta A_{2}, and then we transfer the information to u¯\bar{u}.

For the first step, our main claim here is the following:

(6.59) δψ2X¯+δA2X¯ψLX¯.\|\delta\psi_{2}\|_{\bar{X}}+\|\delta A_{2}\|_{\bar{X}}\lesssim\|\psi\|_{{L\bar{X}}}.

This is done as in the Lemma 4.7, where we transferred L2L^{2} information on ψ\psi to H˙e1\dot{H}^{1}_{e} on δψ2,δA2\delta\psi_{2},\delta A_{2}. We rely on the system (4.42) which we recall here:

(6.60) {Lδψ2= 2ih3ψ+δA2ψ1rδA2(2ih1+δψ2),LδA2=2h1ψ+(ψδψ¯2)1r(δA2)2.\left\{\begin{aligned} L\delta\psi_{2}=&\ 2ih_{3}\psi+\delta A_{2}\psi-\frac{1}{r}\delta A_{2}(2ih_{1}+\delta\psi_{2}),\\ L\delta A_{2}=&\ -2h_{1}\Re{\psi}+\Im(\psi\overline{\delta\psi}_{2})-\frac{1}{r}(\delta A_{2})^{2}.\end{aligned}\right.

The strategy is to produce an iteration scheme in X¯\bar{X} for the above system with an appropriate initialization. The easiest initialization that we can impose is in the form of an initial data, just as we did in Lemma 4.7; to be more precise, we let λ0,α0\lambda_{0},\alpha_{0} be the parameters obtained by imposing the pointwise orthogonality condition δλ0,α0ψ2(r0)=0,δλ0A2(r0)=0\delta^{\lambda_{0},\alpha_{0}}\psi_{2}(r_{0})=0,\delta^{\lambda_{0}}A_{2}(r_{0})=0 for some r01r_{0}\approx 1; recall that our true parameters were set to λ=1,α=0\lambda=1,\alpha=0. The proof of Proposition 4.5 establishes that

|lnλlnλ0|+|αα0|ψL2,|\ln\lambda-\ln\lambda_{0}|+|\alpha-\alpha_{0}|\lesssim\|\psi\|_{L^{2}},

which in turn gives

δψ2δλ0,α0ψ2X¯h1e2iα0h1λ0X¯|lnλlnλ0|+|αα0|ψL2,\begin{split}\|\delta\psi_{2}-\delta^{\lambda_{0},\alpha_{0}}\psi_{2}\|_{\bar{X}}&\approx\|h_{1}-e^{2i\alpha_{0}}h_{1}^{\lambda_{0}}\|_{\bar{X}}\\ &\lesssim|\ln\lambda-\ln\lambda_{0}|+|\alpha-\alpha_{0}|\lesssim\|\psi\|_{L^{2}},\end{split}

together with

δA2δλ0A2X¯h3h3λ0X¯|lnλlnλ0|ψL2.\begin{split}\|\delta A_{2}-\delta^{\lambda_{0}}A_{2}\|_{\bar{X}}\approx\|h_{3}-h_{3}^{\lambda_{0}}\|_{\bar{X}}\lesssim|\ln\lambda-\ln\lambda_{0}|\lesssim\|\psi\|_{L^{2}}.\end{split}

Thus it suffices to establish XX bounds for δλ0,α0ψ2\delta^{\lambda_{0},\alpha_{0}}\psi_{2} and δλ0A2\delta^{\lambda_{0}}A_{2}. In Lemma 4.7 we have established that these were the unique solutions to (6.60) with the above initial conditions, which were obtained via an iteration scheme. Our strategy here is simple: assuming the additional structure ψLX¯\psi\in{L\bar{X}}, we show that the same iteration scheme improves the structure of the solution to the desired XX norms

(6.61) δλ0,α0ψ2X¯+δλ0A2X¯ψLX¯.\|\delta^{\lambda_{0},\alpha_{0}}\psi_{2}\|_{\bar{X}}+\|\delta^{\lambda_{0}}A_{2}\|_{\bar{X}}\lesssim\|\psi\|_{{L\bar{X}}}.

The improvement in the iteration scheme relies on two basic ingredients:

  1. (1)

    the bilinear estimates in Lemma 6.6, namely the estimates in (6.25) and (6.6); it is important for the convergence of the iteration scheme that the bilinear estimates in (6.6) allow us to gain smallness from the quantities δλ0,α0ψ2H˙e1\|\delta^{\lambda_{0},\alpha_{0}}\psi_{2}\|_{\dot{H}^{1}_{e}} and δλ0A2H˙e1\|\delta^{\lambda_{0}}A_{2}\|_{\dot{H}^{1}_{e}}, as we do not have any smallness in X¯\bar{X}.

  2. (2)

    the result in Lemma 6.2 (b), which allows us to solve the linear LL equation in X¯\bar{X}.

This completes the proof of (6.59).

The next step is to transfer the LXLX bounds from ψ\psi together with the XX bounds on δψ2\delta\psi_{2} and δA2\delta A_{2} to u¯Q¯\bar{u}-\bar{Q} and conclude the proof of (6.58). The strategy here follows the same steps as in the proof of (4.35). We recall the system (4.51) for 𝒪=(v¯,w¯,u¯)\mathcal{O}=(\bar{v},\bar{w},\bar{u})

(6.62) r𝒪=𝒪R(ψ),𝒪()=I3.\partial_{r}\mathcal{O}=\mathcal{O}R(\psi),\quad\mathcal{O}(\infty)=I_{3}.

with

R(ψ)=(00ψ100ψ1ψ1ψ10,)R(\psi)=\left(\begin{array}[]{lll}0&0&\Re\psi_{1}\\ 0&0&\Im\psi_{1}\\ -\Re\psi_{1}&-\Im\psi_{1}&0\end{array},\right)

as well as the fact that if ψ=0\psi=0 then ψ2=2ih1\psi_{2}=2ih_{1}, which yields ψ1=2h1r\psi_{1}=-2\frac{h_{1}}{r}, hence

(6.63) R(0)=2h1r(001000100).R(0)=2\frac{h_{1}}{r}\left(\begin{array}[]{lll}0&0&-1\\ 0&0&0\\ 1&0&0\end{array}\right).

The solution is given by (see the generalization of (4.12))

𝒪0=(h30h1010h10h3).\mathcal{O}_{0}=\!\left(\!\!\!\begin{array}[]{ccc}\!h_{3}&\!0&h_{1}\cr\!0\!&\!1&0\cr\!-h_{1}&\!0&h_{3}\end{array}\!\!\!\right).\!

We note that 𝒪01=𝒪0t\mathcal{O}_{0}^{-1}=\mathcal{O}_{0}^{t}. We will prove that

(6.64) R(ψ)R(0)=(lfl,ijgl,ij)i,j=1,3+rH,R(\psi)-R(0)=(\sum_{l}f_{l,ij}g_{l,ij})_{i,j=1,3}+\partial_{r}H,

where

HH˙e1+lgl,ijL2ψL2,HX¯+lgl,ijLX¯ψLX¯\|H\|_{\dot{H}^{1}_{e}}+\sum_{l}\|g_{l,ij}\|_{L^{2}}\lesssim\|\psi\|_{L^{2}},\quad\|H\|_{\bar{X}}+\sum_{l}\|g_{l,ij}\|_{{L\bar{X}}}\lesssim\|\psi\|_{{L\bar{X}}}

and

lfl,ijH˙e11+ψL2,lfl,ijX¯1+ψLX¯.\sum_{l}\|f_{l,ij}\|_{\dot{H}^{1}_{e}}\lesssim 1+\|\psi\|_{L^{2}},\quad\sum_{l}\|f_{l,ij}\|_{\bar{X}}\lesssim 1+\|\psi\|_{{L\bar{X}}}.

Suppose this is done. Then we write the solution to (4.51) in the form

(6.65) 𝒪(r)=(I+Y(r))𝒪0(r)\mathcal{O}(r)=(I+Y(r))\mathcal{O}_{0}(r)

where YY solves the differential equation

(6.66) rY=YN+N,Y()=0N=𝒪0(R(ψ)R(0))𝒪01.\partial_{r}Y=YN+N,\qquad Y(\infty)=0\qquad N=\mathcal{O}_{0}(R(\psi)-R(0))\mathcal{O}_{0}^{-1}.

The matrix NN inherits a similar representation to the one in (6.64) and with similar bounds. Thus we can apply Lemma 6.9 to solve this system and conclude that

Y=L+CY=L+C

with

(6.67) LX¯ψX¯,LH˙e1ψL2,CψL2.\|L\|_{\bar{X}}\lesssim\|\psi\|_{\bar{X}},\quad\|L\|_{\dot{H}^{1}_{e}}\lesssim\|\psi\|_{L^{2}},\quad\|C\|\lesssim\|\psi\|_{L^{2}}.

This provides a similar representation for all columns of YY, in particular for u¯Q¯\bar{u}-\bar{Q}. To finish the proof of our claim (6.64) we need to upgrade the information on u¯Q¯\bar{u}-\bar{Q} to X¯\bar{X}. We first remark that the last row of 𝒪\mathcal{O} is a-priori known, namely (v¯3,w¯3,u¯3)=12(ψ2,ψ2,A2)(\bar{v}_{3},\bar{w}_{3},\bar{u}_{3})=\frac{1}{2}(-\Im\psi_{2},\Re\psi_{2},A_{2}); this shows that from (6.59) we obtain

v¯3+h1X¯+w¯3X¯+u¯3h3X¯ψLX¯.\|\bar{v}_{3}+h_{1}\|_{\bar{X}}+\|\bar{w}_{3}\|_{\bar{X}}+\|\bar{u}_{3}-h_{3}\|_{\bar{X}}\lesssim\|\psi\|_{{L\bar{X}}}.

To transfer this information to u¯1\bar{u}_{1} and u¯2\bar{u}_{2} we use again the orthogonality of 𝒪\mathcal{O}. For u¯1\bar{u}_{1} we have

u¯1h1=v¯2w¯3v¯3w¯2=v¯2w¯3(v¯3+h1)w¯2+h1(w¯21),\bar{u}_{1}-h_{1}=\bar{v}_{2}\bar{w}_{3}-\bar{v}_{3}\bar{w}_{2}=\bar{v}_{2}\bar{w}_{3}-(\bar{v}_{3}+h_{1})\bar{w}_{2}+h_{1}(\bar{w}_{2}-1),

It is important that in each product above we have one term which is apriori bounded in X¯\bar{X} (in order of appearance, w¯3,v¯3+h1,h1\bar{w}_{3},\bar{v}_{3}+h_{1},h_{1}), while the other component of each product is one of the entries in Y=L+CY=L+C and inherits the bounds in (6.67). Thus we can we use (6.24) and (6.6) and the bounds in (6.67) (or better for entries in those matrices) to conclude that

u¯1h1X¯ψLX¯.\|\bar{u}_{1}-h_{1}\|_{\bar{X}}\lesssim\|\psi\|_{{L\bar{X}}}.

A similar argument shows that u¯2X¯ψLX¯\|\bar{u}_{2}\|_{\bar{X}}\lesssim\|\psi\|_{{L\bar{X}}}. This concludes the proof of (6.58).

It remains to prove the bounds claimed for the elements in the decomposition (6.64). From (4.56) we have:

ψ1+2h1r=2ih3rδψ2+B,\psi_{1}+2\frac{h_{1}}{r}=-2ih_{3}\partial_{r}\delta\psi_{2}+B,

where

B=iδA2rδψ2+δA22rh1+|ψ2|2ψ+14r(iψ2|ψ2|2+8(h1)3).B=-i\delta A_{2}\cdot\partial_{r}\delta\psi_{2}+\delta A_{2}\cdot 2\partial_{r}h_{1}+|\psi_{2}|^{2}\psi+\frac{1}{4r}(i\psi_{2}|\psi_{2}|^{2}+8(h_{1})^{3}).

Using (6.59) and the apriori bounds in the energy spaces from Lemma 4.7, it is a straightforward exercise to check that

B=jfjgj,fjX¯1+ψLX¯,fjH˙e11+ψL2,gjLX¯ψLX¯,gjL2ψL2.B=\sum_{j}f_{j}g_{j},\quad\|f_{j}\|_{\bar{X}}\lesssim 1+\|\psi\|_{{L\bar{X}}},\|f_{j}\|_{\dot{H}^{1}_{e}}\lesssim 1+\|\psi\|_{L^{2}},\|g_{j}\|_{{L\bar{X}}}\lesssim\|\psi\|_{{L\bar{X}}},\|g_{j}\|_{L^{2}}\lesssim\|\psi\|_{L^{2}}.

Finally, we write

h3rδψ2=r(h3δψ2)+rh3δψ2,h_{3}\partial_{r}\delta\psi_{2}=\partial_{r}(h_{3}\delta\psi_{2})+\partial_{r}h_{3}\cdot\delta\psi_{2},

and note that h3δψ2X¯δψ2X¯ψLX¯\|h_{3}\delta\psi_{2}\|_{\bar{X}}\lesssim\|\delta\psi_{2}\|_{\bar{X}}\lesssim\|\psi\|_{{L\bar{X}}}, while rh3LX¯1\|\partial_{r}h_{3}\|_{{L\bar{X}}}\lesssim 1. This completes the proof of the claims regarding the decomposition in (6.64), and in turn of (6.58).

This completes the proof of Proposition 6.1.

7. The linear H~\tilde{H} Schrödinger equation

The main goal of this section is the study of estimates for the linear H~\tilde{H} evolution,

(7.1) {itψH~λψ=fψ(0)=ψ0\begin{cases}i\partial_{t}\psi-\tilde{H}_{\lambda}\psi=f\\ \psi(0)=\psi_{0}\end{cases}

where we recall that

H~λ=r21rr+(1r2+8r2(1+λ(t)4r4)).\tilde{H}_{\lambda}=-\partial_{r}^{2}-\frac{1}{r}\partial_{r}+\left(\frac{1}{r^{2}}+\frac{8}{r^{2}(1+\lambda(t)^{4}r^{4})}\right).

One of the important features to highlight is that λ\lambda is allowed to depend on tt. These linear estimates will be crucial in the study of the nonlinear equation (4.59) and the evolution of the modulation parameters λ\lambda and α\alpha.

Our analysis will provide estimates in two categories of spaces. The first one includes the classical energy, Strichartz and local energy decay norms. The second category is a refinement of the first, considering each of the above elements at the dyadic level in the context of our time dependent Littlewood-Paley decomposition.

The first spaces are straightforward; with the local energy norm and its dual defined by

ψLE=ψrLt2Lr2(rdr),ψLE=rψLt2Lr2(rdr),\|\psi\|_{LE}=\|\frac{\psi}{r}\|_{L^{2}_{t}L^{2}_{r}(rdr)},\quad\|\psi\|_{LE^{*}}=\|r\psi\|_{L^{2}_{t}L^{2}_{r}(rdr)},

we define the space SS for solutions to (7.1) and the dual type space NN (precisely, S=NS=N^{*}) for the inhomogeneous term in (7.1) as follows:

S=LtLr2Ltr4LE,N=Lt1Lr2+Ltr43+LE.S=L^{\infty}_{t}L^{2}_{r}\cap L^{4}_{tr}\cap LE,\qquad N=L^{1}_{t}L^{2}_{r}+L^{\frac{4}{3}}_{tr}+LE^{*}.

For instance the SS and NN structures are robust enough to close the main result in Section 9, where we get a first insight into the dispersive behaviour of the problem and the crucial control on λλ2Lt2\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}_{t}}; the relevance of this quantity will become apparent in Proposition 7.8 below.

On the other hand, for many of our estimates we need to be more precise and work with a dyadic Littlewood-Paley decomposition in the H~λ(t)\tilde{H}_{\lambda(t)}-frequency, Pkλ(t)ψP_{k}^{\lambda(t)}\psi; in fact even our strategy to derive the linear estimates involving the spaces SS and NN uses these finer structures. To measure frequency 2k2^{k} waves we define a local energy space LEkLE_{k},

ψLEk=2kψLt2Lr2(A<k)+supm>k2km2ψLt2Lr2(Am),\|\psi\|_{LE_{k}}=2^{k}\|\psi\|_{L^{2}_{t}L^{2}_{r}(A_{<-k})}+\sup_{m>-k}2^{\frac{k-m}{2}}\|\psi\|_{L^{2}_{t}L^{2}_{r}(A_{m})},

as well as the dual space LEkLE_{k}^{*}. Here we note that these norms vary slowly with kk,

(7.2) ψLEkψLEk,if|kk|10.\|\psi\|_{LE_{k}}\approx\|\psi\|_{LE_{k^{\prime}}},\quad\mbox{if}\ |k^{\prime}-k|\leq 10.

Verification is straightforward and left as an exercise.

We aggregate these norms in an 2\ell^{2}-Besov fashion, and set

ψ2LE2=kPkψLL2LEk2,f2LE2=kPkfL1L2+LEk2.\|\psi\|_{\ell^{2}LE}^{2}=\sum_{k}\|P_{k}\psi\|_{L^{\infty}L^{2}\cap LE_{k}}^{2},\qquad\|f\|_{\ell^{2}LE^{*}}^{2}=\sum_{k}\|P_{k}f\|_{L^{1}L^{2}+LE_{k}^{*}}^{2}.

Following [6], we also define an adapted Lk4L^{4}_{k} norm, which is allowed due to the radial symmetry:

ψLk4=supmmax{2m+k2,2m+k8}ψLt4Lr4(Am).\|\psi\|_{L^{4}_{k}}=\sup_{m}\max\{2^{-\frac{m+k}{2}},2^{\frac{m+k}{8}}\}\|\psi\|_{L^{4}_{t}L^{4}_{r}(A_{m})}.

The dual norm is denoted by Lk43L^{\frac{4}{3}}_{k}. The frequency adapted versions of the SS and NN norms are

Sk=LtLr2Lk4LEk,Nk=Lt1Lr2+Lk43+LEk,Sk=Nk.S_{k}=L^{\infty}_{t}L^{2}_{r}\cap L^{4}_{k}\cap LE_{k},\qquad N_{k}=L^{1}_{t}L^{2}_{r}+L^{\frac{4}{3}}_{k}+LE_{k}^{*},\qquad S_{k}=N_{k}^{*}.

Square summing these norms we obtain the spaces l2Sl^{2}S and l2Nl^{2}N with norms

(7.3) ψl2S2=kPkλψSk2,fl2N2=kPkλfNk2.\|\psi\|_{l^{2}S}^{2}=\sum_{k\in\mathbb{Z}}\|P_{k}^{\lambda}\psi\|_{S_{k}}^{2},\qquad\|f\|_{l^{2}N}^{2}=\sum_{k\in\mathbb{Z}}\|P_{k}^{\lambda}f\|_{N_{k}}^{2}.

Given the nice bound (5.110) on the kernel of the projectors PkP_{k}, it is easy to see that these are dual spaces, thus justifying our notation. We recall from Section 5.5 that PkλP_{k}^{\lambda} are the standard projectors in the H~λ\tilde{H}_{\lambda} calculus, or, equivalently, in the ~λ{\tilde{\mathcal{F}}}_{\lambda} frame; also, P~kλ\tilde{P}_{k}^{\lambda} are similar projectors with the additional property that P~kλPkλ=Pkλ\tilde{P}_{k}^{\lambda}P_{k}^{\lambda}=P_{k}^{\lambda}.

We will establish in the following subsection the following relation between the two structure introduced above

(7.4) l2SS,Nl2N.l^{2}S\subset S,\qquad N\subset l^{2}N.

The main result of this section is the following linear estimate.

Theorem 7.1.

Assume the time dependent function λ\lambda satisfies

(7.5) λλ2L2(0,T)1.\|\lambda^{\prime}\lambda^{-2}\|_{L^{2}(0,T)}\lesssim 1.

Then the evolution (7.1) is well-posed in L2L^{2}, and the following estimate holds in [0,T][0,T]:

(7.6) ψl2Sψ0L2+fl2N.\|\psi\|_{l^{2}S}\lesssim\|\psi_{0}\|_{L^{2}}+\|f\|_{l^{2}N}.

In particular, we obtain the following estimates in the SS and NN spaces.

Corollary 7.2.

If ψ\psi is as in the above Theorem, then the following estimate holds in [0,T][0,T]:

ψSψ0L2+fN\|\psi\|_{S}\lesssim\|\psi_{0}\|_{L^{2}}+\|f\|_{N}

Theorem 7.1 provides a bound corresponding to 2\ell^{2} dyadic summation, which is natural when working in the finite energy setting, i.e. ψ(0)L2\psi(0)\in L^{2}. However, later in the article we also investigate the more restrictive case of data with 1\ell^{1} dyadic summation, namely ψ(0)LX¯\psi(0)\in{L\bar{X}}. Our result is as follows:

Theorem 7.3.

Assume that λ\lambda satisfies (7.5). Then the following bound holds in [0,T][0,T] for solutions to the equation (7.1):

(7.7) ψ1Sψ(0)LX¯+f1N.\begin{split}\|\psi\|_{\ell^{1}S}&\lesssim\|\psi(0)\|_{{L\bar{X}}}+\|f\|_{\ell^{1}N}.\end{split}

Here the 1S\ell^{1}S and 1N\ell^{1}N norms are defined as in (7.3) but with 1\ell^{1} summation instead.

We remark that one may also write a frequency envelope version of the above bound. This would assert that if ψ(0)\psi(0) and ff can be placed under a slowly varying frequency envelope ckc_{k} in either 1\ell^{1} or 2\ell^{2}, then the SkS_{k} norm of PkψP_{k}\psi can be placed under a similar frequency envelope.

The proof of the two main theorems above uses Littlewood-Paley decomposition associated to the time dependent operator H~λ\tilde{H}_{\lambda} and begins in the next subsection with some elliptic estimates for frequency localized functions. The main building block of our analysis is the proof of local energy decay for frequency localized functions. This is carried out in Section 7.2 under smallness assumption for our control norm λ/λ2L2\|\lambda^{\prime}/\lambda^{2}\|_{L^{2}}, and then expanded in the next subsection to solutions with a control norm which is large but finite. The Strichartz component of the SS and NN norms is added in Section 7.4. The last step of the analysis is to assemble the dyadic bounds into the full bounds in the theorems, which is achieved by perturbatively estimating the frequency localization errors, which are related to the transference operator studied in Section 5.

In order to keep notation compact, in what follows, whenever a space-time is involved, the time interval is assumed to be restricted to the interval [0,T][0,T] in the condition (7.5); the estimates are uniform with respect to TT.

7.1. Properties of function spaces

In this section we establish the basic relation l2SSl^{2}S\subset S and its dual Nl2NN\subset l^{2}N, along with some other properties of the spaces defined at the beginning of this section.

We begin with some simple estimates which are helpful for later arguments. For convenience we recall the definition of the function ωj,λ\omega_{j,\lambda} from (5.108):

ωj,λ(r)={min{1,r323j},2jλmin{1,2jrr2λ21+r2λ2},2jλ.\omega_{j,\lambda}(r)=\begin{cases}\text{min}\{1,r^{3}2^{3j}\},\quad 2^{j}\geq\lambda\\ \text{min}\{1,2^{j}r\dfrac{r^{2}\lambda^{2}}{1+r^{2}\lambda^{2}}\},\quad 2^{j}\leq\lambda.\end{cases}

Our first result is the following.

Lemma 7.4.

For any k,jk,j\in\mathbb{Z} and λ:I(0,+)\lambda:I\rightarrow(0,+\infty), the following holds true:

(7.8) ωk,λ(t)(r)1Pkλ(t)ψ(t,r)Lt2Lr(Aj)Pkλ(t)ψLEk,\|\omega_{k,\lambda(t)}(r)^{-1}P_{k}^{\lambda(t)}\psi(t,r)\|_{L^{2}_{t}L^{\infty}_{r}(A_{j})}\lesssim\|P_{k}^{\lambda(t)}\psi\|_{LE_{k}},
(7.9) (1+2kr)1/2ωk,λ(t)1Pkλ(t)ψLtLr2kPkλ(t)ψLtLr2,\|(1+2^{k}r)^{1/2}\omega_{k,\lambda(t)}^{-1}P_{k}^{\lambda(t)}\psi\|_{L^{\infty}_{t}L^{\infty}_{r}}\lesssim 2^{k}\|P_{k}^{\lambda(t)}\psi\|_{L^{\infty}_{t}L^{2}_{r}},

with universal implicit constants, independent of k,jk,j and the function λ\lambda.

Proof.

Representing

Pkλ(t)ψ(r)=Pkλ(t)~(Pkλ(t)(ψ))=0Kkλ(t)~(r,s)Pkλ(t)ψ(s)s𝑑s,P_{k}^{\lambda(t)}\psi(r)=\widetilde{P_{k}^{\lambda(t)}}(P_{k}^{\lambda(t)}(\psi))=\int_{0}^{\infty}\widetilde{K_{k}^{\lambda(t)}}(r,s)P_{k}^{\lambda(t)}\psi(s)sds,

we use the spectral projector kernel bound (5.110) with N=1N=1 to estimate at fixed time

|ωk,λ(t)(r)1Pkλ(t)ψ(r)|022kωk,λ(t)(s)|Pkλ(t)(ψ)(s)|(1+2k(s+r))(1+2k|rs|)s𝑑s22k0|Pkλ(t)ψ(s)|(1+2k(s+r))(1+2k|rs|)s𝑑s.\begin{split}|\omega_{k,\lambda(t)}(r)^{-1}P_{k}^{\lambda(t)}\psi(r)|&\lesssim\int_{0}^{\infty}\frac{2^{2k}\omega_{k,\lambda(t)}(s)|P_{k}^{\lambda(t)}(\psi)(s)|}{(1+2^{k}(s+r))(1+2^{k}|r-s|)}sds\\ &\lesssim 2^{2k}\int_{0}^{\infty}\frac{|P_{k}^{\lambda(t)}\psi(s)|}{(1+2^{k}(s+r))(1+2^{k}|r-s|)}sds.\end{split}

We bound the right hand side using the dyadic spatial decomposition in the LEkLE_{k} norm and applying the Cauchy-Schwarz inequality. For rAmr\in A_{m} this gives

|ωk,λ(r)1Pkλ(t)ψ(r)|2kPkλ(t)ψLr2(A<k)+22kj=k+1Pkλ(t)ψL2(Aj)ckjm,|\omega_{k,\lambda}(r)^{-1}P_{k}^{\lambda(t)}\psi(r)|\lesssim 2^{k}\|P_{k}^{\lambda(t)}\psi\|_{L^{2}_{r}(A_{<-k})}+2^{2k}\sum_{j=-k+1}^{\infty}\|P_{k}^{\lambda(t)}\psi\|_{L^{2}(A_{j})}c_{kjm},

where

ckjm2=suprAmAj2jds(1+2k+j)2(1+2k|rs|)2{22k2jk,j3mj+322k22(j+k),otherwise.c_{kjm}^{2}=\sup_{r\in A_{m}}\int_{A_{j}}\frac{2^{j}ds}{(1+2^{k+j})^{2}(1+2^{k}|r-s|)^{2}}\approx\begin{cases}2^{-2k}2^{-j-k},\quad{j-3}\leq m\leq{j+3}\\ 2^{-2k}2^{-2(j+k)},\quad\text{otherwise}.\end{cases}

Taking the Lt2LrL^{2}_{t}L^{\infty}_{r} norm we arrive at

ωk,λ(t)1Pkλ(t)ψLt2Lr(Am)\displaystyle\|\omega_{k,\lambda(t)}^{-1}P_{k}^{\lambda(t)}\psi\|_{L^{2}_{t}L^{\infty}_{r}(A_{m})}\lesssim 2kPkλ(t)ψLt2Lr2(A<k)+22kj=k+1Pkλ(t)ψLt2Lr2(Aj)ckjm\displaystyle\ 2^{k}\|P_{k}^{\lambda(t)}\psi\|_{L^{2}_{t}L^{2}_{r}(A_{<-k})}+2^{2k}\sum_{j=-k+1}^{\infty}\|P_{k}^{\lambda(t)}\psi\|_{L^{2}_{t}L^{2}_{r}(A_{j})}c_{kjm}
\displaystyle\lesssim Pkλ(t)ψLEkj=k22kckjm2j+k2\displaystyle\ \|P_{k}^{\lambda(t)}\psi\|_{LE_{k}}\sum_{j=-k}^{\infty}2^{2k}c_{kjm}2^{-\frac{j+k}{2}}
\displaystyle\lesssim Pkλ(t)ψLEk.\displaystyle\ \|P_{k}^{\lambda(t)}\psi\|_{LE_{k}}.

Similarly, we have the fixed time bound

|(1+2kr)1/2ωk,λ(t)1Pkλ(t)ψ(r)|(1+2kr)1/2022kωk,λ(t)(s)|Pkλ(t)ψ(s)|(1+2k(s+r))(1+2k|rs|)s𝑑s,|(1+2^{k}r)^{1/2}\omega_{k,\lambda(t)}^{-1}P_{k}^{\lambda(t)}\psi(r)|\lesssim(1+2^{k}r)^{1/2}\int_{0}^{\infty}\frac{2^{2k}\omega_{k,\lambda(t)}(s)|P_{k}^{\lambda(t)}\psi(s)|}{(1+2^{k}(s+r))(1+2^{k}|r-s|)}sds,

which gives, by Cauchy-Schwarz,

|(1+2kr)1/2ωk,λ(t)1Pkλ(t)ψ|Pkλ(t)ψL2ck(r)12,|(1+2^{k}r)^{1/2}\omega_{k,\lambda(t)}^{-1}P_{k}^{\lambda(t)}\psi|\lesssim\|P_{k}^{\lambda(t)}\psi\|_{L^{2}}c_{k}(r)^{\frac{1}{2}},

where

ck(r)=0(1+2kr)24k(1+2k(s+r))2(1+2k|rs|)2s𝑑s023kds(1+|2k(rs)|)222k.\begin{split}c_{k}(r)=\int_{0}^{\infty}\frac{(1+2^{k}r)2^{4k}}{(1+2^{k}(s+r))^{2}(1+2^{k}|r-s|)^{2}}sds\lesssim\int_{0}^{\infty}\frac{2^{3k}\,ds}{(1+|2^{k}(r-s)|)^{2}}\lesssim 2^{2k}.\end{split}

Hence (7.9) follows.

We are now ready to proceed with the arguments for the inclusions l2SSl^{2}S\subset S and Nl2NN\subset l^{2}N. First we prove that the LELE norm is controlled by the square sum of the LEkLE_{k} norms and the corresponding dual estimate.

Lemma 7.5.

The following estimates hold true:

(7.10) ψLE2kPkλ(t)ψLEk2,\|\psi\|_{LE}^{2}\lesssim\sum_{k\in\mathbb{Z}}\|P_{k}^{\lambda(t)}\psi\|_{LE_{k}}^{2},
(7.11) kPkλ(t)fLEk2fLE2.\sum_{k}\|P_{k}^{\lambda(t)}f\|_{LE_{k}^{*}}^{2}\lesssim\|f\|_{LE^{*}}^{2}.
Proof.

In each spatial dyadic region we decompose in frequency and estimate as follows:

ψrLr2(Aj)12jψLr2(Aj)12jkPkλ(t)ψLr2(Aj).\|\frac{\psi}{r}\|_{L^{2}_{r}(A_{j})}\lesssim\frac{1}{2^{j}}\|\psi\|_{L^{2}_{r}(A_{j})}\lesssim\frac{1}{2^{j}}\sum_{k\in\mathbb{Z}}\|P_{k}^{\lambda(t)}\psi\|_{L^{2}_{r}(A_{j})}.

To estimate the terms in the last sum we consider two cases. If k+j0k+j\geq 0 we estimate directly

(7.12) 2jPkλ(t)ψLr2(Aj)2j+k2Pkλ(t)ψLEk.2^{-j}\|P_{k}^{\lambda(t)}\psi\|_{L^{2}_{r}(A_{j})}\lesssim 2^{-\frac{j+k}{2}}\|P_{k}^{\lambda(t)}\psi\|_{LE_{k}}.

If k+j0k+j\leq 0, we use the L2LL^{2}L^{\infty} bound from (7.8), and note that in AjA_{j} we have ωk,λ(t)2j+k\omega_{k,\lambda(t)}\lesssim 2^{j+k}; thus we obtain

(7.13) 2jPkλ(t)ψLr2(Aj)Pkλ(t)ψL2L(Aj)2j+kPkλ(t)ψLEk.2^{-j}\|P_{k}^{\lambda(t)}\psi\|_{L^{2}_{r}(A_{j})}\lesssim\|P_{k}^{\lambda(t)}\psi\|_{L^{2}L^{\infty}(A_{j})}\lesssim 2^{j+k}\|P_{k}^{\lambda(t)}\psi\|_{LE_{k}}.

Combining the bounds (7.12) and (7.13) we arrive at

(7.14) ψrLt2Lr2(Aj)k=2|j+k|2Pkλ(t)ψLEk.\|\frac{\psi}{r}\|_{L^{2}_{t}L^{2}_{r}(A_{j})}\lesssim\sum_{k=-\infty}^{\infty}2^{\frac{-|j+k|}{2}}\|P_{k}^{\lambda(t)}\psi\|_{LE_{k}}.

Finally, using a discrete convolution estimate, we have

ψLE2jψrLt2Lr2(Aj)2kPkλ(t)ψLEk2,\|\psi\|_{LE}^{2}\leq\sum_{j\in\mathbb{Z}}\|\frac{\psi}{r}\|_{L^{2}_{t}L^{2}_{r}(A_{j})}^{2}\lesssim\sum_{k\in\mathbb{Z}}\|P_{k}^{\lambda(t)}\psi\|_{LE_{k}}^{2},

as needed. This finishes the proof of (7.11); the estimate (7.10) follows by duality. ∎

Next we prove that the Ltr4L^{4}_{tr} norm is controlled by the square sum of the Lk4L^{4}_{k} norms and the corresponding dual estimate.

Lemma 7.6.

We have the following estimates:

(7.15) ψLt,r4(mPmλ(t)ψLm42)1/2,\|\psi\|_{L^{4}_{t,r}}\lesssim\left(\sum_{m}\|P_{m}^{\lambda(t)}\psi\|_{L^{4}_{m}}^{2}\right)^{1/2},
(7.16) (mPmλ(t)fLm4/32)1/2fLt,r4/3.\left(\sum_{m}\|P_{m}^{\lambda(t)}f\|_{L_{m}^{4/3}}^{2}\right)^{1/2}\lesssim\|f\|_{L^{4/3}_{t,r}}.
Proof.

We have

(7.17) ψLt,r4(Am)jPjλ(t)ψLt,r4(Am)j>mPjλ(t)ψLt,r4(Am)+jmPjλ(t)ψLt,r4(Am)j>m2(m+j)82(m+j)8Pjλ(t)ψLt,r4(Am)+jm2(m+j)22(m+j)2Pjλ(t)ψLt,r4(Am)j2|m+j|8Pjλ(t)ψLj4.\begin{split}\|\psi\|_{L^{4}_{t,r}(A_{m})}&\leq\sum_{j}\|P_{j}^{\lambda(t)}\psi\|_{L^{4}_{t,r}(A_{m})}\\ &\leq\sum_{j>-m}\|P_{j}^{\lambda(t)}\psi\|_{L^{4}_{t,r}(A_{m})}+\sum_{j\leq-m}\|P_{j}^{\lambda(t)}\psi\|_{L^{4}_{t,r}(A_{m})}\\ &\leq\sum_{j>-m}2^{\frac{-(m+j)}{8}}2^{\frac{(m+j)}{8}}\|P_{j}^{\lambda(t)}\psi\|_{L^{4}_{t,r}(A_{m})}+\sum_{j\leq-m}2^{\frac{-(m+j)}{2}}2^{\frac{(m+j)}{2}}\|P_{j}^{\lambda(t)}\psi\|_{L^{4}_{t,r}(A_{m})}\\ &\leq\sum_{j\in\mathbb{Z}}2^{\frac{-|m+j|}{8}}\|P_{j}^{\lambda(t)}\psi\|_{L^{4}_{j}}.\end{split}

Then, we use Young’s inequality for discrete convolutions to obtain

mψLt4Lr4(Am)2m|j2|m+j|8Pjλ(t)ψLj4|2nPnλ(t)ψLn42.\sum_{m\in\mathbb{Z}}\|\psi\|_{L^{4}_{t}L^{4}_{r}(A_{m})}^{2}\lesssim\sum_{m}|\sum_{j\in\mathbb{Z}}2^{\frac{-|m+j|}{8}}\|P_{j}^{\lambda(t)}\psi\|_{L^{4}_{j}}|^{2}\lesssim\sum_{n\in\mathbb{Z}}\|P_{n}^{\lambda(t)}\psi\|_{L^{4}_{n}}^{2}.

Based on this we obtain

ψLt,r42(mψLt,r4(Am)4)1/2m(ψLt,r4(Am)4)1/2nPnλ(t)ψLn42,\|\psi\|_{L^{4}_{t,r}}^{2}\lesssim\left(\sum_{m}\|\psi\|_{L^{4}_{t,r}(A_{m})}^{4}\right)^{1/2}\lesssim\sum_{m}(\|\psi\|_{L^{4}_{t,r}(A_{m})}^{4})^{1/2}\lesssim\sum_{n\in\mathbb{Z}}\|P_{n}^{\lambda(t)}\psi\|_{L^{4}_{n}}^{2},

which implies (7.15). We also have the dual estimate (7.16).

It is clear that (7.4) follows from the previous two Lemmas.

7.2. Local energy decay for frequency localized solutions.

A first step in the analysis of the linear Schrödinger equation (7.1) is to derive local energy decay estimates for frequency localized functions. We recall the equation (7.1) here for convenience:

itψH~λ(t)ψ=f,ψ(0)=ψ0.i\partial_{t}\psi-\tilde{H}_{\lambda(t)}\psi=f,\qquad\psi(0)=\psi_{0}.

Throughout the rest of this section we say that a time dependent function ψ\psi is localized at a dyadic frequency 2j2^{j} if ~λ(t)ψ(t){\tilde{\mathcal{F}}}_{\lambda(t)}\psi(t) is supported in the region |ξ|2j|\xi|\approx 2^{j}. Notably, the scale function λ\lambda is allowed to depend on time.

In the following Lemma we establish the basic linear estimates in the spaces LEjLE_{j} for solutions to (7.1) which are frequency localized.

Lemma 7.7.

Assume that ψ\psi is a solution to (7.1) which is localized at frequency 2j2^{j}. Then the following holds true:

(7.18) ψLEjLtLr2+2jrψLEjψ(0)L2+fLt1Lr2+LEj.\|\psi\|_{LE_{j}\cap L^{\infty}_{t}L^{2}_{r}}+2^{-j}\|\partial_{r}\psi\|_{LE_{j}}\lesssim\|\psi(0)\|_{L^{2}}+\|f\|_{L^{1}_{t}L^{2}_{r}+LE_{j}^{*}}.

We remark that the frequency localization of uu does not guarantee a similar frequency localization of ff. This is due to the time dependence of λ\lambda.

Proof.

We first establish a weaker version of (7.18), namely

(7.19) ψLEj+2jrψLEjψLtLr2+fLt1Lr2+LEj.\|\psi\|_{LE_{j}}+2^{-j}\|\partial_{r}\psi\|_{LE_{j}}\lesssim\|\psi\|_{L^{\infty}_{t}L^{2}_{r}}+\|f\|_{L^{1}_{t}L^{2}_{r}+LE_{j}^{*}}.

Then we show that a direct energy estimate allows us to replace the uniform energy bound by the initial data size and conclude with with (7.18).

Our approach is in the spirit of the one used by the third author in [35], see also [26], using the positive commutator method. First we say that a sequence {αn}n\{\alpha_{n}\}_{n\in\mathbb{Z}} of positive numbers is slowly varying if

|lnαjlnαj1|210,j.|\ln{\alpha_{j}}-\ln{\alpha_{j-1}}|\leq 2^{-10},\qquad\forall j\in\mathbb{Z}.

Based on such a sequence we introduce the normed space Xk,αX_{k,\alpha} and its dual Xk,αX_{k,\alpha}^{\prime} as follows

uXk,α2=22kuL2(A<k)2+2klkαl2luL2(Al)2uXk,α2=22kuL2(A<k)2+2klkαl12luL2(Al)2.\begin{split}\|u\|_{X_{k,\alpha}}^{2}&=2^{2k}\|u\|^{2}_{L^{2}(A_{<-k})}+2^{k}\sum_{l\geq-k}\alpha_{l}2^{-l}\|u\|^{2}_{L^{2}(A_{l})}\\ \|u\|_{X_{k,\alpha}^{\prime}}^{2}&=2^{-2k}\|u\|^{2}_{L^{2}(A_{<-k})}+2^{-k}\sum_{l\geq-k}\alpha_{l}^{-1}2^{l}\|u\|^{2}_{L^{2}(A_{l})}.\end{split}

For all slowly varying sequences {αn}n\{\alpha_{n}\}_{n\in\mathbb{Z}} with nαn=1\sum_{n}\alpha_{n}=1, we claim that

(7.20) ψXj,α+2jrψXj,αψLtLr2+fXj,α+Lt1Lr2.\|\psi\|_{X_{j,\alpha}}+2^{-j}\|\partial_{r}\psi\|_{X_{j,\alpha}}\lesssim\|\psi\|_{L^{\infty}_{t}L^{2}_{r}}+\|f\|_{X_{j,\alpha}^{\prime}+L^{1}_{t}L^{2}_{r}}.

If we assume that (7.20) is true, then, we can let β\beta be a slowly varying sequence with nβn=1\sum_{n}\beta_{n}=1. Then, n(12(αn+βn))=1\sum_{n}(\frac{1}{2}(\alpha_{n}+\beta_{n}))=1, and we apply (7.20) for 12(αn+βn)\frac{1}{2}(\alpha_{n}+\beta_{n}) to obtain, for instance, that

(7.21) ψXj,α+β+2jrψXj,α+βψLtLr2+fXj,α+β+Lt1Lr2,\|\psi\|_{X_{j,\alpha+\beta}}+2^{-j}\|\partial_{r}\psi\|_{X_{j,\alpha+\beta}}\lesssim\|\psi\|_{L^{\infty}_{t}L^{2}_{r}}+\|f\|_{X_{j,\alpha+\beta}^{\prime}+L^{1}_{t}L^{2}_{r}},

from which we derive the weaker estimate

(7.22) ψXj,α+2jrψXj,αψLtLr2+fXj,β+Lt1Lr2.\|\psi\|_{X_{j,\alpha}}+2^{-j}\|\partial_{r}\psi\|_{X_{j,\alpha}}\lesssim\|\psi\|_{L^{\infty}_{t}L^{2}_{r}}+\|f\|_{X_{j,\beta}^{\prime}+L^{1}_{t}L^{2}_{r}}.

Since any l1l^{1} sequence can be dominated by a slowly varying sequence with a comparable l1l^{1} size, we can drop the assumption in (7.22) that α\alpha and β\beta are slowly varying. By maximizing the left-hand side with respect to αl1\alpha\in l^{1} and by minimizing the right-hand side with respect to βl1\beta\in l^{1}, we obtain (7.19).

The remaining part of this step is devoted to the proof of (7.20). Let

Qj(u)=χ(2jr)rru+(2+rr)(χ(2jr)u),Q_{j}(u)=\chi(2^{j}r)r\partial_{r}u+(2+r\partial_{r})(\chi(2^{j}r)u),

where χ\chi will be chosen to be a smooth function related to the slowly varying sequence {αn}\{\alpha_{n}\}. A straightforward computation shows that QjQ_{j} is antisymmetric; based on this and the equation for ψ\psi we obtain

0TQjψ,f𝑑t=0TQjψ,(itH~λ(t))ψ𝑑t=0TQjψ,tψ𝑑t0TQjψ,H~λ(t)ψ𝑑t=120TtQjψ,ψdt0TQjψ,H~λ(t)ψ𝑑t,\begin{split}\Re\int_{0}^{T}\langle Q_{j}\psi,f\rangle dt&=\Re\int_{0}^{T}\langle Q_{j}\psi,(i\partial_{t}-\tilde{H}_{\lambda(t)})\psi\rangle dt\\ &=\Im\int_{0}^{T}\langle Q_{j}\psi,\partial_{t}\psi\rangle dt-\Re\int_{0}^{T}\langle Q_{j}\psi,\tilde{H}_{\lambda(t)}\psi\rangle dt\\ &=\frac{1}{2}\Im\int_{0}^{T}\partial_{t}\langle Q_{j}\psi,\psi\rangle dt-\Re\int_{0}^{T}\langle Q_{j}\psi,\tilde{H}_{\lambda(t)}\psi\rangle dt,\end{split}

which, by rearranging terms, becomes

(7.23) 0TQjψ,f𝑑t+12Qjψ,ψ|0T=0TQjψ,H~λ(t)ψ𝑑t.-\Re\int_{0}^{T}\langle Q_{j}\psi,f\rangle dt+\frac{1}{2}\Im\langle Q_{j}\psi,\psi\rangle|_{0}^{T}=\Re\int_{0}^{T}\langle Q_{j}\psi,\tilde{H}_{\lambda(t)}\psi\rangle dt.

The right hand side can be expanded as follows

0TQjψ,H~λ(t)ψ𝑑t=0TQjψ,Δψ+V~λ(t)ψ𝑑t=0TQjrψ,rψ𝑑t+0T[r,Qj]ψ,rψ𝑑t+0TQjψ,V~λ(t)ψ𝑑t=0T[r,Qj]ψ,rψ𝑑t+120T[V~λ(t),Qj]ψ,ψ𝑑t,\begin{split}\Re\int_{0}^{T}\langle Q_{j}\psi,\tilde{H}_{\lambda(t)}\psi\rangle dt&=\Re\int_{0}^{T}\langle Q_{j}\psi,-\Delta\psi+\tilde{V}_{\lambda(t)}\psi\rangle dt\\ &=\Re\int_{0}^{T}\langle Q_{j}\partial_{r}\psi,\partial_{r}\psi\rangle dt+\Re\int_{0}^{T}\langle[\partial_{r},Q_{j}]\psi,\partial_{r}\psi\rangle dt+\Re\int_{0}^{T}\langle Q_{j}\psi,\tilde{V}_{\lambda(t)}\psi\rangle dt\\ &=\Re\int_{0}^{T}\langle[\partial_{r},Q_{j}]\psi,\partial_{r}\psi\rangle dt+\frac{1}{2}\int_{0}^{T}\langle[\tilde{V}_{\lambda(t)},Q_{j}]\psi,\psi\rangle dt,\end{split}

where we have used twice the antisymmetry of QjQ_{j}. We now compute the commutators and start with the easier one,

12[V~λ(t),Qj]=rχ(2jr)rV~λ(t)=χ(2jr)2r2(1+81+3λ4(t)r4(1+λ(t)4r4)2)=2χ(2jr)(9+26r4λ(t)4+r8λ(t)8r2(1+r4λ(t)4)2>0.\begin{split}\frac{1}{2}[\tilde{V}_{\lambda(t)},Q_{j}]=-r\chi(2^{j}r)\partial_{r}\tilde{V}_{\lambda(t)}&=\chi(2^{j}r)\frac{2}{r^{2}}\left(1+8\frac{1+3\lambda^{4}(t)r^{4}}{(1+\lambda(t)^{4}r^{4})^{2}}\right)\\ &=\frac{2\chi(2^{j}r)(9+26r^{4}\lambda(t)^{4}+r^{8}\lambda(t)^{8}}{r^{2}(1+r^{4}\lambda(t)^{4})^{2}}>0.\end{split}

The other commutator is

[r,Qj]=2(2jrχ(2jr)+χ(2jr))r+(32jχ(2jr)+22jrχ′′(2jr)).[\partial_{r},Q_{j}]=2(2^{j}r\chi^{\prime}(2^{j}r)+\chi(2^{j}r))\partial_{r}+(3\cdot 2^{j}\chi^{\prime}(2^{j}r)+2^{2j}r\chi^{\prime\prime}(2^{j}r)).

From these we obtain

(7.24) 0TQjψ,H~λ(t)ψ𝑑t=120T[V~λ(t),Qj]ψ,ψ𝑑t+0T0(2[χ(2jr)+2jrχ(2jr)]|rψ|2+[32jχ(2jr)+r4jχ′′(2jr)]r|ψ|22)r𝑑r𝑑t.\begin{split}&\Re\int_{0}^{T}\langle Q_{j}\psi,\tilde{H}_{\lambda(t)}\psi\rangle dt=\frac{1}{2}\int_{0}^{T}\langle[\tilde{V}_{\lambda(t)},Q_{j}]\psi,\psi\rangle dt\\ +&\int_{0}^{T}\int_{0}^{\infty}\left(2[\chi(2^{j}r)+2^{j}r\chi^{\prime}(2^{j}r)]|\partial_{r}\psi|^{2}+[3\cdot 2^{j}\chi^{\prime}(2^{j}r)+r4^{j}\chi^{\prime\prime}(2^{j}r)]\frac{\partial_{r}|\psi|^{2}}{2}\right)rdrdt.\end{split}

We impose the following condition on χ\chi

(7.25) |χ(r)|+|rχ′′(r)|δ(χ(r)+rχ(r))|\chi^{\prime}(r)|+|r\chi^{\prime\prime}(r)|\leq\delta(\chi(r)+r\chi^{\prime}(r))

for some sufficiently small δ>0\delta>0. This leads to

0TQjψ,H~λ(t)ψ𝑑t120T[V~λ(t),Qj]ψ,ψ𝑑t+120T0(|rψ|2322jδ|ψ|2)aj(r)r𝑑r𝑑t,\Re\int_{0}^{T}\langle Q_{j}\psi,\tilde{H}_{\lambda(t)}\psi\rangle dt\geq\frac{1}{2}\int_{0}^{T}\langle[\tilde{V}_{\lambda(t)},Q_{j}]\psi,\psi\rangle dt+\frac{1}{2}\int_{0}^{T}\int_{0}^{\infty}(|\partial_{r}\psi|^{2}-3\cdot 2^{2j}\delta|\psi|^{2})a_{j}(r)rdrdt,

where aj(r)=χ(2jr)+2jrχ(2jr)a_{j}(r)=\chi(2^{j}r)+2^{j}r\chi^{\prime}(2^{j}r). From (7.23) and the previous estimates, we obtain

(7.26) 0T[V~λ(t),Qj]ψ,ψ𝑑t+0T0(|rψ|2322jδ|ψ|2)aj(r)r𝑑r𝑑t20TQjψ,f𝑑t+Qjψ,ψ|0T.\begin{split}&\int_{0}^{T}\langle[\tilde{V}_{\lambda(t)},Q_{j}]\psi,\psi\rangle dt+\int_{0}^{T}\int_{0}^{\infty}(|\partial_{r}\psi|^{2}-3\cdot 2^{2j}\delta|\psi|^{2})a_{j}(r)rdrdt\\ \leq&-2\Re\int_{0}^{T}\langle Q_{j}\psi,f\rangle dt+\Im\langle Q_{j}\psi,\psi\rangle|_{0}^{T}.\end{split}

The main idea of what follows next is to show that there is an appropriate choice of χ\chi, depending on the sequence {αn}\{\alpha_{n}\} such that we can derive (7.20) from (7.26). We first increase αn\alpha_{n} the so that it remains slowly varying and, in addition, satisfies

(7.27) αn=1,fornn0j;nn0jαn1.\alpha_{n}=1,\ \ \ \mbox{for}\ n\leq n_{0}-j;\qquad\sum_{n\geq n_{0}-j}\alpha_{n}\approx 1.

Here n0n_{0} is a positive number to be chosen later.

We claim that given a slowly varying sequence αn\alpha_{n}, satisfying (7.27), and δ>0\delta>0 we can find χ\chi satisfying (7.25), so that

(7.28) aj(r)αn1+2n+j,r2na_{j}(r)\gtrsim\frac{\alpha_{n}}{1+2^{n+j}},\qquad r\approx 2^{n}

and the following three fixed time bounds hold:

(7.29) QjψLr2δ1ψLr2,QjψXk,αδ1ψXk,α,\|Q_{j}\psi\|_{L^{2}_{r}}\lesssim\delta^{-1}\|\psi\|_{L^{2}_{r}},\quad\|Q_{j}\psi\|_{X_{k,\alpha}}\lesssim\delta^{-1}\|\psi\|_{X_{k,\alpha}},
(7.30) 22j0aj(r)|ψ|2r𝑑r0aj(r)|rψ|2r𝑑r+120[V~,Q]|ψ|2r𝑑r.2^{2j}\int_{0}^{\infty}a_{j}(r)|\psi|^{2}rdr\lesssim\int_{0}^{\infty}a_{j}(r)|\partial_{r}\psi|^{2}rdr+\frac{1}{2}\int_{0}^{\infty}[\tilde{V},Q]|\psi|^{2}rdr.

These estimates are claimed for functions ψ\psi which are localized at frequency 2j2^{j} in the frame H~λ(t)\tilde{H}_{\lambda(t)} and the constants involved do not depend on the value of λ(t)(0,)\lambda(t)\in(0,\infty) or the sequence {αn}\{\alpha_{n}\} or δ\delta.

Going back to the estimate (7.26) and making use of (7.28), (7.29) and (7.30), we obtain

(7.31) ψXj,α2+22jrψXj,α2ψLtLr22+fXj,α+Lt1Lr22,\|\psi\|_{X_{j,\alpha}}^{2}+2^{-2j}\|\partial_{r}\psi\|_{X_{j,\alpha}}^{2}\lesssim\|\psi\|_{L^{\infty}_{t}L^{2}_{r}}^{2}+\|f\|_{X_{j,\alpha}^{\prime}+L^{1}_{t}L^{2}_{r}}^{2},

when all terms are restricted to the time interval [0,T][0,T], but with the a constant independent of TT. This implies (7.20), which in turn was shown to imply our main claim (7.19).

It would be helpful to explain the role of the small constant δ\delta in the above. δ\delta needs to be small enough so that when taking into account (7.30), the following holds true:

0T[V~λ(t),Qj]ψ,ψ𝑑t+0T0(|rψ|2322jδ|ψ|2)aj(r)r𝑑r𝑑t 22j0aj(r)|ψ|2r𝑑r+0aj(r)|rψ|2r𝑑r.\begin{split}&\int_{0}^{T}\langle[\tilde{V}_{\lambda(t)},Q_{j}]\psi,\psi\rangle dt+\int_{0}^{T}\int_{0}^{\infty}(|\partial_{r}\psi|^{2}-3\cdot 2^{2j}\delta|\psi|^{2})a_{j}(r)rdrdt\\ \gtrsim&\ 2^{2j}\int_{0}^{\infty}a_{j}(r)|\psi|^{2}rdr+\int_{0}^{\infty}a_{j}(r)|\partial_{r}\psi|^{2}rdr.\end{split}

Once this is achieved, δ\delta simply becomes just another constant whose quantification is not necessary; in particular we could have ignored its quantification in (7.29) or, as we already have done, its quantification in (7.31).

We now proceed with the construction of χ\chi satisfying (7.25), (7.28), (7.29) and (7.30). Based on this, we construct a slowly varying function α\alpha such that

α(s)αnifs2n\alpha(s)\approx\alpha_{n}\ \ \mbox{if}\ \ s\approx 2^{n}

and with symbol regularity

|lα(s)|lslα(s),l.|\partial^{l}\alpha(s)|\lesssim_{l}s^{-l}\alpha(s),\qquad l\in\mathbb{N}.

Due to the first condition in (7.27) we can take α\alpha such that α(s)=1\alpha(s)=1 for s2n0js\leq 2^{n_{0}-j}. We then construct the function χ\chi by

sχ(s)=0sα(2jτ)h(τ)𝑑τ,s\chi(s)=\int_{0}^{s}\alpha(2^{-j}\tau)h(\tau)d\tau,

where hh is a smooth adapted variant of r1r^{-1}, namely h(s)=1h(s)=1 for s2n0s\leq 2^{n_{0}} and h(s)2n0s1h(s)\approx 2^{n_{0}}s^{-1} for s2n0+1s\geq 2^{n_{0}+1}; in particular hh has symbol type estimates |h(k)(s)|k2n0sk1,x2n0|h^{(k)}(s)|\lesssim_{k}2^{n_{0}}s^{-k-1},x\geq 2^{n_{0}}. With the observation that χ(s)=1\chi(s)=1 for s2n0s\leq 2^{n_{0}}, one easily verifies the pointwise bounds

(7.32) χ(s)(1+2n0s)1,|χ(l)(s)|2ln0(1+2n0s)l1,1l4.\chi(s)\approx(1+2^{-n_{0}}s)^{-1},\qquad|\chi^{(l)}(s)|\lesssim 2^{-ln_{0}}(1+2^{-n_{0}}s)^{-l-1},\quad 1\leq l\leq 4.

Furthermore, we have

(sχ(s))=α(2js)h(s)(1+2n0s)1.1,|(sχ(s))|2n0(1+2n0s)2.(s\chi(s))^{\prime}=\alpha(2^{-j}s)h(s)\gtrsim(1+2^{-n_{0}}s)^{-1.1},\qquad|(s\chi^{\prime}(s))^{\prime}|\lesssim 2^{-n_{0}}(1+2^{-n_{0}}s)^{-2}.

It is a straightforward exercise to verify that χ\chi satisfies (7.28). A direct computation shows that

|χ(r)|+|rχ′′(r)|2n0(1+2n0r)22n0(rχ(r)),\begin{split}|\chi^{\prime}(r)|+|r\chi^{\prime\prime}(r)|&\lesssim\frac{2^{-n_{0}}}{(1+2^{-n_{0}}r)^{2}}\lesssim 2^{-n_{0}}(r\chi(r))^{\prime},\end{split}

where the constants involved are independent of n0n_{0}; thus by choosing 2n0δ2^{-n_{0}}\sim\delta we obtain that χ\chi satisfies (7.25).

Next we seek to establish the first estimate in (7.29) which requires an estimate on QjψLr2\|Q_{j}\psi\|_{L^{2}_{r}}. We have

(7.33) Qjψ=2rχ(2jr)rψ+2χ(2jr)ψ+rr(χ(2jr))ψ.Q_{j}\psi=2r\chi(2^{j}r)\partial_{r}\psi+2\chi(2^{j}r)\psi+r\partial_{r}(\chi(2^{j}r))\psi.

From this we obtain

(7.34) QjψLr2rχ(2jr)rψLr2+ψLr2.\|Q_{j}\psi\|_{L^{2}_{r}}\lesssim\|r\chi(2^{j}r)\partial_{r}\psi\|_{L^{2}_{r}}+\|\psi\|_{L^{2}_{r}}.

Due to the frequency localization of ψ\psi, we can write

ψ(t)=Pjλ(t)~ψ(t)=0Kj~λ(t)(r,s)ψ(t,s)s𝑑s,\psi(t)=\widetilde{P_{j}^{\lambda(t)}}\psi(t)=\int_{0}^{\infty}\widetilde{K_{j}}^{\lambda(t)}(r,s)\psi(t,s)sds,

from which it follows that

rψ(t,r)=0rKj~λ(t)(r,s)ψ(s)sds.\partial_{r}\psi(t,r)=\int_{0}^{\infty}\partial_{r}\widetilde{K_{j}}^{\lambda(t)}(r,s)\psi(s)sds.

Moreover, from Lemma 5.11 we obtain the following estimate:

(7.35) |rχ(2jr)rK~jλ(t)(r,s)|N22j2n0(1+2j(r+s))(1+2j|rs|)N.|r\chi(2^{j}r)\partial_{r}\widetilde{K}_{j}^{\lambda(t)}(r,s)|\lesssim_{N}\frac{2^{2j}2^{n_{0}}}{(1+2^{j}(r+s))(1+2^{j}|r-s|)^{N}}.

This allows us to estimate as follows:

(7.36) rχ(2jr)rψLr2supsrχ(2jr)rK~jλ(t)(r,s)Lr112suprrχ(2jr)rK~jλ(t)(r,s)Lr112ψL22n0ψL2.\begin{split}\|r\chi(2^{j}r)\partial_{r}\psi\|_{L^{2}_{r}}&\lesssim\sup_{s}\|r\chi(2^{j}r)\partial_{r}\widetilde{K}_{j}^{\lambda(t)}(r,s)\|_{L^{1}_{r}}^{\frac{1}{2}}\cdot\sup_{r}\|r\chi(2^{j}r)\partial_{r}\widetilde{K}_{j}^{\lambda(t)}(r,s)\|_{L^{1}_{r}}^{\frac{1}{2}}\cdot\|\psi\|_{L^{2}}\\ &\lesssim 2^{n_{0}}\|\psi\|_{L^{2}}.\end{split}

Inserting this estimate into (7.34) leads to:

Qjψ(t)Lr2δ1ψ(t)Lr2,\|Q_{j}\psi(t)\|_{L^{2}_{r}}\lesssim\delta^{-1}\|\psi(t)\|_{L^{2}_{r}},

which concludes the first part of (7.29). We now turn to the second part of (7.29), namely establishing that

Qjψ(t)Xj,αδ1ψ(t)Xj,α.\|Q_{j}\psi(t)\|_{X_{j,\alpha}}\lesssim\delta^{-1}\|\psi(t)\|_{X_{j,\alpha}}.

We start from the formula (7.33) for QjψQ_{j}\psi. Using the rapid decay away from the diagonal in (7.35), and the fact that the weights αk\alpha_{k} are slowly varying, we obtain

rχ(2jr)rψXj,α222n0ψXj,α2,\|r\chi(2^{j}r)\partial_{r}\psi\|_{X_{j,\alpha}}^{2}\lesssim 2^{2n_{0}}\|\psi\|_{X_{j,\alpha}}^{2},

from which it follows that

QjψXj,α222n0PjψXj,α2.\|Q_{j}\psi\|_{X_{j,\alpha}}^{2}\lesssim 2^{2n_{0}}\|P_{j}\psi\|_{X_{j,\alpha}}^{2}.

Recalling that δ2n0\delta\approx 2^{-n_{0}}, this concludes our argument for (7.29).

We are then left with verifying (7.30); we start with

ψ(t)=P~jλ(t)ψ(t)=P~jλ(t)H~λ(t)1Lλ(t)Lλ(t)ψ=0Kj1,λ(t)(r,s)Lλ(t)ψ(s)s𝑑s.\psi(t)=\widetilde{P}_{j}^{\lambda(t)}\psi(t)=\widetilde{P}_{j}^{\lambda(t)}\widetilde{H}_{\lambda(t)}^{-1}L_{\lambda(t)}L^{*}_{\lambda(t)}\psi=\int_{0}^{\infty}K_{j}^{1,\lambda(t)}(r,s)\cdot L^{*}_{\lambda(t)}\psi(s)sds.

From this we obtain:

aj(r)2jψ(t,r)=0K7λ(t)(s,r)Lλ(t)ψ(s)aj(s)s𝑑s,\sqrt{a_{j}(r)}2^{j}\psi(t,r)=\int_{0}^{\infty}K_{7}^{\lambda(t)}(s,r)L^{*}_{\lambda(t)}\psi(s)\sqrt{a_{j}(s)}sds,

where

K7λ(t)(r,s)=aj(r)aj(s)2jKj1,λ(t)(r,s).K_{7}^{\lambda(t)}(r,s)=\frac{\sqrt{a_{j}(r)}}{\sqrt{a_{j}(s)}}2^{j}K_{j}^{1,\lambda(t)}(r,s).

Using the symbol-type estimates on α\alpha and hh, and noting that

aj(x)=χ(2jx)+2jxχ(2j(x))=α(x)h(2jx),a_{j}(x)=\chi(2^{j}x)+2^{j}x\chi^{\prime}(2^{j}(x))=\alpha(x)h(2^{j}x),

we obtain |aj(s)|aj(s)s1\frac{|a_{j}^{\prime}(s)|}{a_{j}(s)}\lesssim s^{-1} which implies |aj(r)||aj(s)||a_{j}(r)|\approx|a_{j}(s)| in the regime rsr\sim s. Based on this and the estimates on Kj1,λ(t)K_{j}^{1,\lambda(t)} from (5.112), we obtain

|K7λ(t)(s,r)|N22j(1+2j(r+s))(1+2j|rs|)N,rs.|K_{7}^{\lambda(t)}(s,r)|\lesssim_{N}\frac{2^{2j}}{(1+2^{j}(r+s))(1+2^{j}|r-s|)^{N}},\quad r\sim s.

If r>2sr>2s or rs2r\leq\frac{s}{2}, then,

aj(r)aj(s)=χ(2jr)+2jrχ(2jr)aj(s)1(1+2n0+jr)aj(s)1(yχ)(2js)(1+2js)1.1.\frac{a_{j}(r)}{a_{j}(s)}=\frac{\chi(2^{j}r)+2^{j}r\chi^{\prime}(2^{j}r)}{a_{j}(s)}\lesssim\frac{1}{(1+2^{-n_{0}+j}r)a_{j}(s)}\lesssim\frac{1}{(y\chi)^{\prime}(2^{j}s)}\lesssim(1+2^{j}s)^{1.1}.

Using this we can estimate

|K7λ(t)(r,s)|Naj(r)aj(s)22j(1+2j(r+s))(1+2j|rs|)N22j(1+2j(r+s))(1+2jmax{r,s})N2.|K_{7}^{\lambda(t)}(r,s)|\lesssim_{N}\frac{\sqrt{a_{j}(r)}}{\sqrt{a_{j}(s)}}\frac{2^{2j}}{(1+2^{j}(r+s))(1+2^{j}|r-s|)^{N}}\lesssim\frac{2^{2j}}{(1+2^{j}(r+s))(1+2^{j}\text{max}\{r,s\})^{N-2}}.

From this it follows that

0aj(r)22j|ψ|2r𝑑r0|Lλ(t)ψ|2(r)aj(r)r𝑑r.\int_{0}^{\infty}a_{j}(r)2^{2j}|\psi|^{2}rdr\lesssim\int_{0}^{\infty}|L^{*}_{\lambda(t)}\psi|^{2}(r)a_{j}(r)rdr.

From the definition of Lλ(t)L^{*}_{\lambda(t)}, see (5.1), it follows that

|Lλ(t)u|21r2|u(r)|2+|ru(r)|2.|L^{*}_{\lambda(t)}u|^{2}\lesssim\frac{1}{r^{2}}|u(r)|^{2}+|\partial_{r}u(r)|^{2}.

Since we have the straightforward inequality |aj(r)|(1+2n0+jr)1χ(2jr)|a_{j}(r)|\lesssim(1+2^{-n_{0}+j}r)^{-1}\lesssim\chi(2^{j}r), we can estimate

aj(r)|Lλ(t)ψ(r)||ψ(t,r)|2r2aj(r)+|rψ(t,r)|2aj(r)(χ(2jr)|ψ|2r2+|rψ|2)aj(r).\begin{split}a_{j}(r)|L^{*}_{\lambda(t)}\psi(r)|&\lesssim\frac{|\psi(t,r)|^{2}}{r^{2}}a_{j}(r)+|\partial_{r}\psi(t,r)|^{2}a_{j}(r)\\ &\lesssim(\frac{\chi(2^{j}r)|\psi|^{2}}{r^{2}}+|\partial_{r}\psi|^{2})a_{j}(r).\end{split}

Therefore, we have

0aj(r)22j|ψ(r)|2r𝑑r0|ψ|24χ(2jr)r2(9+26r4λ(t)4+r8λ(t)8)(1+r4λ(t)4)2r𝑑r+0|2ψ|2aj(r)r𝑑r,\begin{split}\int_{0}^{\infty}a_{j}(r)2^{2j}|\psi(r)|^{2}rdr&\lesssim\int_{0}^{\infty}|\psi|^{2}\frac{4\chi(2^{j}r)}{r^{2}}\frac{(9+26r^{4}\lambda(t)^{4}+r^{8}\lambda(t)^{8})}{(1+r^{4}\lambda(t)^{4})^{2}}rdr\\ &+\int_{0}^{\infty}|\partial_{2}\psi|^{2}a_{j}(r)rdr,\end{split}

which is precisely (7.30). This finishes the proof of (7.19). To conclude the proof of the Lemma, we involve a simple energy argument: multiplying (7.1) by ψ¯\bar{\psi}, integrating with respect to rr and taking the imaginary part, gives

12ddtψ(t)Lr22=f,ψ.\frac{1}{2}\frac{d}{dt}\|\psi(t)\|^{2}_{L^{2}_{r}}=\Im\langle f,\psi\rangle.

Integrating this on an arbitrary interval [0,t0][0,t_{0}] gives

ψ(t0)Lr22ψ(0)Lr22+20t0|f,ψ|𝑑tψ(0)Lr22+2M1ψLEk2+2MfLEk2.\|\psi(t_{0})\|^{2}_{L^{2}_{r}}\leq\|\psi(0)\|^{2}_{L^{2}_{r}}+2\int_{0}^{t_{0}}|\Im\langle f,\psi\rangle|dt\leq\|\psi(0)\|^{2}_{L^{2}_{r}}+2M^{-1}\|\psi\|_{LE_{k}}^{2}+2M\|f\|_{LE_{k}^{*}}^{2}.

Thus we obtain

ψLtLr22ψ(0)Lr22+2M1ψLEk2+2MfLEk2.\|\psi\|^{2}_{L^{\infty}_{t}L^{2}_{r}}\leq\|\psi(0)\|^{2}_{L^{2}_{r}}+2M^{-1}\|\psi\|_{LE_{k}}^{2}+2M\|f\|_{LE_{k}^{*}}^{2}.

Using (7.19) for the term ψLEk\|\psi\|_{LE_{k}} and choosing MM large enough gives (7.18) and this finishes the proof of our Lemma.

7.3. Full local energy decay

Here we assemble the dyadic local energy bounds of the previous section into a full local energy bound. This does not require smallness of λ/λ2\lambda^{\prime}/\lambda^{2} in L2L^{2}. Instead, we will be able to track the dependence of the implicit constants on the above L2L^{2} norm. This justifies introducing the notation (already mentioned in section 3.7)

(7.37) (T):=λλ2Lt2[0,T].\mathcal{B}(T):=\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}_{t}[0,T]}.

For some of the dyadic estimates we will also use the more refined quantities

(7.38) j(T)=χλ=2jλλ2L2[0,T],\mathcal{B}_{j}(T)=\|\chi_{\lambda=2^{j}}\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}[0,T]},

which measures the same L2L^{2} norm but with a weight concentrated around the dyadic region {t[0,T];λ(t)2j}\{t\in[0,T];\lambda(t)\approx 2^{j}\}. The two are related by

2(T)jj2(T),\mathcal{B}^{2}(T)\approx\sum_{j}\mathcal{B}_{j}^{2}(T),

where the constants used in \approx are independent of TT.

Our main well-posedness result concerning the linear H~\tilde{H} equation is as follows:

Proposition 7.8.

Assume that λ/λ2L2[0,T]\lambda^{\prime}/\lambda^{2}\in L^{2}[0,T]. Then the equation (7.1) is well-posed in L2L^{2}, and the solution ψ\psi satisfies the following bound:

(7.39) ψ2(LELtLr2)[0,T](1+(T))ψ(0)L2+(1+(T))2f2(LE+Lt1Lr2)[0,T].\begin{split}\|\psi\|_{\ell^{2}(LE\cap L^{\infty}_{t}L^{2}_{r})[0,T]}&\lesssim(1+\mathcal{B}(T))\|\psi(0)\|_{L^{2}}+(1+\mathcal{B}(T))^{2}\|f\|_{\ell^{2}(LE^{*}+L^{1}_{t}L^{2}_{r})[0,T]}.\end{split}

For clarity we also write this bound in an expanded form,

(7.40) jPjψLEjLL22+22jrPjψLEjLL22(1+(T))2ψ(0)L22+(1+(T))4jPjfLt1Lr2+LEj2.\begin{split}&\!\!\sum_{j}\|P_{j}\psi\|_{LE_{j}\cap L^{\infty}L^{2}}^{2}+2^{-2j}\|\partial_{r}P_{j}\psi\|_{LE_{j}\cap L^{\infty}L^{2}}^{2}\\ \lesssim&\ (1+\mathcal{B}(T))^{2}\|\psi(0)\|_{L^{2}}^{2}+(1+\mathcal{B}(T))^{4}\sum_{j}\|P_{j}f\|_{L^{1}_{t}L^{2}_{r}+LE_{j}^{*}}^{2}.\end{split}

Here we have added also the bounds for rPjψ\partial_{r}P_{j}\psi, which follow almost freely from the argument due to the frequency localization.

Proof.

Since the equation (7.1) coincides with its adjoint equation and the time is reversible, a standard duality argument shows that L2L^{2} well-posedness follows from the bound (7.39). Hence we turn our attention to the proof of this bound.

For each jj, the functions PjλψP_{j}^{\lambda}\psi solve the Cauchy problem

(7.41) {itPjλψH~λ(t)Pjλψ=Pjλf+gjPjλψ(0)=Pjλψ0\begin{cases}i\partial_{t}P_{j}^{\lambda}\psi-\widetilde{H}_{\lambda(t)}P_{j}^{\lambda}\psi=P_{j}^{\lambda}f+g_{j}\\ P_{j}^{\lambda}\psi(0)=P_{j}^{\lambda}\psi_{0}\end{cases}

where gjg_{j} arises from the commutator of PjP_{j} with t\partial_{t},

gj=i[t,Pjλ]ψ.g_{j}=i[\partial_{t},P_{j}^{\lambda}]\psi.

Hence our first task is to obtain good estimates for gjg_{j}. The above commutator is described in Lemma 5.13 in terms of the transference operator 𝒦~\tilde{\mathcal{K}}, which is in turn estimated in Lemma 5.14. Here we will use these building blocks to prove the following

Lemma 7.9.

The above commutators gjg_{j} satisfy the bounds

(7.42) gjLEjj(T)kχk=jψkLtLr2\|g_{j}\|_{LE_{j}^{*}}\lesssim\mathcal{B}_{j}(T)\sum_{k}\chi_{k=j}\|\psi_{k}\|_{L^{\infty}_{t}L^{2}_{r}}

with ψk:=Pkλψ\psi_{k}:=P_{k}^{\lambda}\psi, respectively

(7.43) gjLt1Lr2j(T)kχk=jψkLEk.\|g_{j}\|_{L^{1}_{t}L^{2}_{r}}\lesssim\mathcal{B}_{j}(T)\sum_{k}\chi_{k=j}\|\psi_{k}\|_{LE_{k}}.

In the above, all the space-time norms are restricted to the time interval [0,T][0,T].

Proof.

Just as in the statement of the Lemma, in what follows below all the space-time norms are restricted to the time interval [0,T][0,T]; in order to keep the formualas compact we skip this from notation.

From Lemma 5.13 we have

gj=iλ~λ1[𝒦~λ,mj]~λψ.g_{j}=i\lambda^{\prime}{\tilde{\mathcal{F}}}_{\lambda}^{-1}[\tilde{\mathcal{K}}_{\lambda},m_{j}]{\tilde{\mathcal{F}}}_{\lambda}\psi.

We split this into

gj=kgjk,gjk=iλ~λ1[𝒦~λ,mj]m~k~λψk,g_{j}=\sum_{k}g_{jk},\qquad g_{jk}=i\lambda^{\prime}{\tilde{\mathcal{F}}}_{\lambda}^{-1}[\tilde{\mathcal{K}}_{\lambda},m_{j}]{\tilde{m}}_{k}{\tilde{\mathcal{F}}}_{\lambda}\psi_{k},

and then it remains to show that

(7.44) gjkLEjj(T)χk=jψkLtLr2,gjkLt1Lr2j(T)ψkLEk.\|g_{jk}\|_{LE_{j}^{*}}\lesssim\mathcal{B}_{j}(T)\chi_{k=j}\|\psi_{k}\|_{L^{\infty}_{t}L^{2}_{r}},\qquad\|g_{jk}\|_{L^{1}_{t}L^{2}_{r}}\lesssim\mathcal{B}_{j}(T)\|\psi_{k}\|_{LE_{k}}.

These are dual bounds so it remains to prove the first one. For that we need to estimate gjkg_{jk} in Lt2Lr2(Am)L^{2}_{t}L^{2}_{r}(A_{m}), first for m<jm<j and then for mjm\geq j. In the first case we use (5.117) to obtain the fixed time bound

gjk(t)Lr2(A<j)2jχλ=2jχj=k|λ|λ2ψk(t)L2,\|g_{jk}(t)\|_{L^{2}_{r}(A_{<j})}\lesssim 2^{j}\chi_{\lambda=2^{j}}\chi_{j=k}\frac{|\lambda^{\prime}|}{\lambda^{2}}\|\psi_{k}(t)\|_{L^{2}},

which after time integration yields

gjkLt2Lr2(A<j)2jjψkLtLr2.\|g_{jk}\|_{L^{2}_{t}L^{2}_{r}(A_{<j})}\lesssim 2^{j}\mathcal{B}_{j}\|\psi_{k}\|_{L^{\infty}_{t}L^{2}_{r}}.

In the second case we use (5.118) to obtain the fixed time bound

gjk(t)Lr2(Am)2mχλ=2jχj=k|λ|λ2ψ(t)L2,mj,\|g_{jk}(t)\|_{L^{2}_{r}(A_{m})}\lesssim 2^{-m}\chi_{\lambda=2^{j}}\chi_{j=k}\frac{|\lambda^{\prime}|}{\lambda^{2}}\|\psi(t)\|_{L^{2}},\qquad m\geq j,

which after time integration yields

gjkLt2Lr2(Am)2mj(T)ψLL2.\|g_{jk}\|_{L^{2}_{t}L^{2}_{r}(A_{m})}\lesssim 2^{-m}\mathcal{B}_{j}(T)\|\psi\|_{L^{\infty}L^{2}}.

Combining the two cases we arrive at

gjkLEjj(T)χj=kψkLL2mj2m+j2j(T)χj=kψkLtLr2,\|g_{jk}\|_{LE_{j}^{*}}\lesssim\mathcal{B}_{j}(T)\chi_{j=k}\|\psi_{k}\|_{L^{\infty}L^{2}}\sum_{m\geq-j}2^{-\frac{m+j}{2}}\lesssim\mathcal{B}_{j}(T)\chi_{j=k}\|\psi_{k}\|_{L^{\infty}_{t}L^{2}_{r}},

as desired.

Now we return to the proof of the proposition, and apply Lemma 7.7 to Pjλ(t)ψP_{j}^{\lambda(t)}\psi, using Lemma 7.9 for gjg_{j}. This yields

Pjλ(t)ψLEjLtLr2\displaystyle\|P_{j}^{\lambda(t)}\psi\|_{LE_{j}\cap L^{\infty}_{t}L^{2}_{r}}\lesssim Pjλ(0)ψ(0)L2+Pjλ(t)f+gjL1L2+LEj\displaystyle\ \|P_{j}^{\lambda(0)}\psi(0)\|_{L^{2}}+\|P_{j}^{\lambda(t)}f+g_{j}\|_{L^{1}L^{2}+LE_{j}^{*}}
\displaystyle\lesssim Pjλ(0)ψ(0)L2+Pjλ(t)fL1L2+LEj+gjLEj\displaystyle\ \|P_{j}^{\lambda(0)}\psi(0)\|_{L^{2}}+\|P_{j}^{\lambda(t)}f\|_{L^{1}L^{2}+LE_{j}^{*}}+\|g_{j}\|_{LE_{j}^{*}}
\displaystyle\lesssim Pjλ(0)ψ(0)L2+Pjλ(t)fL1L2+LEj+j(T)ψLL2.\displaystyle\ \|P_{j}^{\lambda(0)}\psi(0)\|_{L^{2}}+\|P_{j}^{\lambda(t)}f\|_{L^{1}L^{2}+LE_{j}^{*}}+\mathcal{B}_{j}(T)\|\psi\|_{L^{\infty}L^{2}}.

We square this bound and sum over jj\in\mathbb{Z} to arrive at

(7.45) jPjλ(t)ψLEjLL22jPjλ(t)fLt1Lr2+LEj2+(1+2(T))ψLL22.\begin{split}\sum_{j}\|P_{j}^{\lambda(t)}\psi\|_{LE_{j}\cap L^{\infty}L^{2}}^{2}&\lesssim\sum_{j}\|P_{j}^{\lambda(t)}f\|_{L^{1}_{t}L^{2}_{r}+LE_{j}^{*}}^{2}+(1+\mathcal{B}^{2}(T))\|\psi\|_{L^{\infty}L^{2}}^{2}.\end{split}

It remains to supplement this with an energy estimate for ψ\psi. We have

ddtψ(t)L22=2ψ,if=2j,kPkλ(t)ψ,iPjλ(t)f.\frac{d}{dt}\|\psi(t)\|_{L^{2}}^{2}=-2\Re\langle\psi,if\rangle=-2\sum_{j,k\in\mathbb{Z}}\Re\langle P_{k}^{\lambda(t)}\psi,iP_{j}^{\lambda(t)}f\rangle.

The terms in the last sum vanish unless |kj|3|k-j|\leq 3, in which case the LEkLE_{k} and LEjLE_{j} norms are equivalent by (7.2). We integrate in time the last relation and use Cauchy-Schwarz to obtain

ψLL22\displaystyle\|\psi\|_{L^{\infty}L^{2}}^{2}\lesssim ψ(0)L22+|jk|3Pkλ(t)ψLL2LEkPjλ(t)fL1L2+LEj\displaystyle\ \|\psi(0)\|_{L^{2}}^{2}+\sum_{|j-k|\leq 3}\|P_{k}^{\lambda(t)}\psi\|_{L^{\infty}L^{2}\cap LE_{k}}\|P_{j}^{\lambda(t)}f\|_{L^{1}L^{2}+LE_{j}^{*}}
\displaystyle\lesssim ψ(0)L22+(kPkλ(t)ψLL2LEk2)12(jPjλ(t)fL1L2+LEj2)12\displaystyle\ \|\psi(0)\|_{L^{2}}^{2}+\left(\sum_{k}\|P_{k}^{\lambda(t)}\psi\|_{L^{\infty}L^{2}\cap LE_{k}}^{2}\right)^{\frac{1}{2}}\left(\sum_{j}\|P_{j}^{\lambda(t)}f\|_{L^{1}L^{2}+LE_{j}^{*}}^{2}\right)^{\frac{1}{2}}

Finally, we insert this in (7.45) and apply Cauchy-Schwarz one more time to arrive at (7.39).

7.4. Adding the Strichartz norms to the mix

Here we start with the estimate (7.39), written in the expanded form (7.40), and show that we can add in the weighted L4L^{4}, respectively L43L^{\frac{4}{3}} norms to arrive at the main bound (7.6) in Theorem 7.1.

We do this in several steps, beginning with a Lemma which captures the essential one dimensional Strichartz estimate:

Lemma 7.10.

Let ψ\psi be a solution to (7.1). Then the following estimate holds:

(7.46) supjk2j2ψLt4Lr(Aj)ψLtLr2LEk+2krψLtLr2LEk+fNk.\sup_{j\geq-k}2^{\frac{j}{2}}\|\psi\|_{L^{4}_{t}L^{\infty}_{r}(A_{j})}\lesssim\|\psi\|_{L^{\infty}_{t}L^{2}_{r}\cap LE_{k}}+2^{-k}\|\partial_{r}\psi\|_{L^{\infty}_{t}L^{2}_{r}\cap LE_{k}}+\|f\|_{N_{k}}.

This lemma is designed for functions at frequency 2k2^{k}, but does not actually assumes any frequency localization. We disregard for now the region A<kA_{<k}, where matters will be simpler but will involve the frequency localization.

Proof.

We begin by localizing the problem to each region AjA_{j}, using a suitable cutoff function χj\chi_{j}. For ψj=χjψ\psi_{j}=\chi_{j}\psi we can write a one dimensional Schrödinger equation

(it+r2)ψj=fj:=gj+χjf(i\partial_{t}+\partial_{r}^{2})\psi_{j}=f_{j}:=g_{j}+\chi_{j}f

where the additional source term gjg_{j} is given by

gj=2χjrψ+χj′′ψχj1rrψχjV~λψ,g_{j}=2\chi^{\prime}_{j}\partial_{r}\psi+\chi^{\prime\prime}_{j}\psi-{\chi_{j}}\frac{1}{r}\partial_{r}\psi-{\chi_{j}}\tilde{V}_{\lambda}\psi,

and can be readily estimated by

gjLEkψLEk+2krψLEk.\|g_{j}\|_{LE_{k}^{*}}\lesssim\|\psi\|_{LE_{k}}+2^{-k}\|\partial_{r}\psi\|_{LE_{k}}.

Then (7.46) reduces to proving

(7.47) 2j2ψjLt4LrψjLEk+fjNk2^{\frac{j}{2}}\|\psi_{j}\|_{L^{4}_{t}L^{\infty}_{r}}\lesssim\|\psi_{j}\|_{LE_{k}}+\|f_{j}\|_{N_{k}}

for ψj\psi_{j} localized in AjA_{j}.

We show that this last bound is a direct consequence of the one dimensional Strichartz estimate for ψ\psi, which implies that

(7.48) ψjLt4LrLrLr2(dr)2kj2ψjLt2Lr2(dr)+fj2kj2Lt2Lr2(dr)+Lt1Lr2(dr)+Lt43Lr1(dr).\|\psi_{j}\|_{L^{4}_{t}L^{\infty}_{r}\cap L^{\infty}_{r}L^{2}_{r}(dr)}\lesssim 2^{\frac{k-j}{2}}\|\psi_{j}\|_{L^{2}_{t}L^{2}_{r}(dr)}+\|f_{j}\|_{2^{\frac{k-j}{2}}L^{2}_{t}L^{2}_{r}(dr)+L^{1}_{t}L^{2}_{r}(dr)+L^{\frac{4}{3}}_{t}L^{1}_{r}(dr)}.

This is not exactly the classical Strichartz estimate, but it is derived from it in a straightforward manner. Precisely, the norms involved on the right-hand side terms only involve LtpL^{p}_{t} spaces with p2p\leq 2, so they are square summable with respect to time interval decompositions. This allows one to apply a second localization, now with respect to time, on the time scale 2k+j2^{-k+j}, with an error that may be included in the L2L^{2} component of fjf_{j}. This reduces (7.48) to the case when ψj\psi_{j} is localized in a time interval of size 2k+j2^{-k+j}. But on this time scale the L2L^{2} norm of ψ\psi bounds the averaged energy, while the L2L2L^{2}L^{2} component of fjf_{j} may be included in L1L2L^{1}L^{2} by Hölder’s inequality.

The last inequality uses the one dimensional measure drdr. Converting to the rdrrdr measure with r2jr\approx 2^{j} we arrive at

(7.49) 2j2ψjLt4Lr2kj2ψjLt,r2+fj2kj2Lt,r2+L1Lr2+2j2Lt43Lr1.2^{\frac{j}{2}}\|\psi_{j}\|_{L^{4}_{t}L^{\infty}_{r}}\lesssim 2^{\frac{k-j}{2}}\|\psi_{j}\|_{L^{2}_{t,r}}+\|f_{j}\|_{2^{\frac{k-j}{2}}L^{2}_{t,r}+L^{1}L^{2}_{r}+2^{-\frac{j}{2}}L^{\frac{4}{3}}_{t}L^{1}_{r}}.

To see that this implies (7.47), it suffices to verify that restricted to AjA_{j} we have the uniform inclusion

Nk2kj2Lt,r2+Lt1Lr2+2j2Lt43Lr1.N_{k}\subset 2^{\frac{k-j}{2}}L^{2}_{t,r}+L^{1}_{t}L^{2}_{r}+2^{-\frac{j}{2}}L^{\frac{4}{3}}_{t}L^{1}_{r}.

It suffices to consider the L43L^{\frac{4}{3}} component of NkN_{k}, for which this is a straightforward interpolation computation.

The corresponding bound inside A<kA_{<-k} is instead a Bernstein type inequality:

Lemma 7.11.

Let ψ\psi be a function which is localized at frequency 2k2^{k}. Then we have

(7.50) 2k2ψLt4Lr(A<k)ψLtLr2LEk.2^{-\frac{k}{2}}\|\psi\|_{L^{4}_{t}L^{\infty}_{r}(A_{<-k})}\lesssim\|\psi\|_{L^{\infty}_{t}L^{2}_{r}\cap LE_{k}}.
Proof.

To capture the frequency localization we rewrite the above inequality in the form

2k2PkλψLt4Lr(A<k)ψLtLr2LEk.2^{-\frac{k}{2}}\|P_{k}^{\lambda}\psi\|_{L^{4}_{t}L^{\infty}_{r}(A_{<-k})}\lesssim\|\psi\|_{L^{\infty}_{t}L^{2}_{r}\cap LE_{k}}.

Then the kernel bounds for PkλP_{k}^{\lambda} from (5.110) yield the following two bounds:

PkλψLt,r(A<k)2kψLtLr2,PkλψLt2Lr(A<k)ψLEk.\|P_{k}^{\lambda}\psi\|_{L^{\infty}_{t,r}(A_{<-k})}\lesssim 2^{k}\|\psi\|_{L^{\infty}_{t}L^{2}_{r}},\quad\|P_{k}^{\lambda}\psi\|_{L^{2}_{t}L^{\infty}_{r}(A_{<-k})}\lesssim\|\psi\|_{LE_{k}}.

Interpolating these to bound leads to the desired Lt4LrL^{4}_{t}L^{\infty}_{r} bound. ∎

The next step is to combine the above two Lemmas in order to prove a frequency localized Lk4L^{4}_{k} bound:

Lemma 7.12.

Let ψ\psi be a solution to (7.1) which is localized at frequency 2k2^{k}. Then we have

(7.51) ψSkψLtLr2LEk+fNk.\|\psi\|_{S_{k}}\lesssim\|\psi\|_{L^{\infty}_{t}L^{2}_{r}\cap LE_{k}}+\|f\|_{N_{k}}.
Proof.

We need to estimate ψ\psi in Lk4L^{4}_{k}, which requires Ltr4L^{4}_{tr} bounds in each of the sets AjA_{j}. We have two cases:

i) If j<kj<-k then we use (7.50) and Hölder’s inequality in rr.

ii) If j>kj>k, then we first use the bounds in (5.111) for the kernel of rPkλ\partial_{r}P_{k}^{\lambda} to bound rψ\partial_{r}\psi,

2krψLtLr2LEkψLtLr2LEk.2^{-k}\|\partial_{r}\psi\|_{L^{\infty}_{t}L^{2}_{r}\cap LE_{k}}\lesssim\|\psi\|_{L^{\infty}_{t}L^{2}_{r}\cap LE_{k}}.

It then follows from (7.46) that

supjk2j2ψLt4Lr(Aj)ψLtLr2LEk+fNk.\sup_{j\geq-k}2^{\frac{j}{2}}\|\psi\|_{L^{4}_{t}L^{\infty}_{r}(A_{j})}\lesssim\|\psi\|_{L^{\infty}_{t}L^{2}_{r}\cap LE_{k}}+\|f\|_{N_{k}}.

This allows us to bound the Lt4LrL^{4}_{t}L^{\infty}_{r} norm in the left hand of (7.46). Then we interpolate this estimate with the LtLr2L^{\infty}_{t}L^{2}_{r} energy bound to obtain an estimate in Lt,r6L^{6}_{t,r}, which is then interpolated with the Lt,r2L^{2}_{t,r} bound from the local energy to obtain the desired conclusion; tracking the correct powers in this interpolation is left as an exercise. ∎

Finally, we begin to assemble the dyadic bounds so that we complete the proof of our main result in this section, Theorem 7.1. Applying (7.51) to PkψP_{k}\psi and then Lemma 7.9 we obtain

PkλψSk\displaystyle\|P_{k}^{\lambda}\psi\|_{S_{k}}\lesssim PkλψLtLr2LEk+PkλfNk+gkLEk\displaystyle\ \|P_{k}^{\lambda}\psi\|_{L^{\infty}_{t}L^{2}_{r}\cap LE_{k}}+\|P_{k}^{\lambda}f\|_{N_{k}}+\|g_{k}\|_{LE_{k}^{*}}
\displaystyle\lesssim PkλψLtLr2LEk+PkλfNk+k(T)ψLtLr2.\displaystyle\ \|P_{k}^{\lambda}\psi\|_{L^{\infty}_{t}L^{2}_{r}\cap LE_{k}}+\|P_{k}^{\lambda}f\|_{N_{k}}+\mathcal{B}_{k}(T)\|\psi\|_{L^{\infty}_{t}L^{2}_{r}}.

Square summing these bounds yields

(7.52) ψ2S(1+(T))ψ2(LELtLr2)+f2N.\|\psi\|_{\ell^{2}S}\lesssim(1+\mathcal{B}(T))\|\psi\|_{\ell^{2}(LE\cap L^{\infty}_{t}L^{2}_{r})}+\|f\|_{\ell^{2}N}.

To continue we consider first the case when f2(LE+L1L2)f\in\ell^{2}(LE^{*}+L^{1}L^{2}). Then we can apply Proposition 7.8 to conclude that

(7.53) ψ2S(1+(T))2ψ0L2+(1+(T))3f2(LE+L1L2).\|\psi\|_{\ell^{2}S}\lesssim(1+\mathcal{B}(T))^{2}\|\psi_{0}\|_{L^{2}}+(1+\mathcal{B}(T))^{3}\|f\|_{\ell^{2}(LE^{*}+L^{1}L^{2})}.

From here, we claim that a duality argument yields

(7.54) ψ2(LELL2)(1+(T))2ψ0L2+(1+(T))3f2N.\|\psi\|_{\ell^{2}(LE\cap L^{\infty}L^{2})}\lesssim(1+\mathcal{B}(T))^{2}\|\psi_{0}\|_{L^{2}}+(1+\mathcal{B}(T))^{3}\|f\|_{\ell^{2}N}.

To see this, we pair ψ\psi and ff with a second solution ψ~\tilde{\psi}, f~\tilde{f}. The duality relation between the two in a time interval [0,T][0,T] yields

ψ,ψ~|0T=0Tψ,if~f,iψ~dt.\langle\psi,\tilde{\psi}\rangle|_{0}^{T}=\int_{0}^{T}\langle\psi,i\tilde{f}\rangle-\langle f,i\tilde{\psi}\rangle\,dt.

This implies

ψ(T),ψ~(T)0Tψ,if~𝑑tψ(0)L2ψ~(0)L2+f2Nψ~2S,\langle\psi(T),\tilde{\psi}(T)\rangle-\int_{0}^{T}\langle\psi,i\tilde{f}\rangle\,dt\lesssim\|\psi(0)\|_{L^{2}}\|\tilde{\psi}(0)\|_{L^{2}}+\|f\|_{\ell^{2}N}\|\tilde{\psi}\|_{\ell^{2}S},

which combined with (7.53) applied for the backward problem yields

ψ(T),ψ~(T)0Tψ,if~𝑑t(ψ(0)L2+f2N)(ψ~(T)L2+f~2(LE+L1L2)).\langle\psi(T),\tilde{\psi}(T)\rangle-\int_{0}^{T}\langle\psi,i\tilde{f}\rangle\,dt\lesssim(\|\psi(0)\|_{L^{2}}+\|f\|_{\ell^{2}N})(\|\tilde{\psi}(T)\|_{L^{2}}+\|\tilde{f}\|_{\ell^{2}(LE^{*}+L^{1}L^{2})}).

This in turn gives (7.54).

Finally, combining (7.54) with (7.52) we obtain the conclusion of Theorem 7.1.

7.5. The 1\ell^{1} Strichartz and local energy bounds

Here we turn our attention to the l1l^{1} bounds in Theorem 7.3, which we now prove. Precisely, we will show that

Proposition 7.13.

The following bound holds for solutions to the equation (7.1):

(7.55) ψ1S(1+(T))2ψ(0)LX¯+(1+(T))4f1N.\begin{split}\|\psi\|_{\ell^{1}S}&\lesssim(1+\mathcal{B}(T))^{2}\|\psi(0)\|_{{L\bar{X}}}+(1+\mathcal{B}(T))^{4}\|f\|_{\ell^{1}N}.\end{split}

In the above the space-time norms are restricted to the time interval [0,T][0,T].

Proof.

As in the case of the 2\ell^{2} bound, our starting point is Lemma 7.7, which applied to PjψP_{j}\psi yields the dyadic bound

(7.56) PjψSjψj(0)L2+PjfNj+gjLEj.\begin{split}\|P_{j}\psi\|_{S_{j}}&\lesssim\|\psi_{j}(0)\|_{L^{2}}+\|P_{j}f\|_{N_{j}}+\|g_{j}\|_{LE^{*}_{j}}.\end{split}

For gjg_{j} we use Lemma 7.9, which implies that

(7.57) gjLEjj(T)kψkLtLr2χk=j.\|g_{j}\|_{LE^{*}_{j}}\lesssim\mathcal{B}_{j}(T)\sum_{k}\|\psi_{k}\|_{L^{\infty}_{t}L^{2}_{r}}\chi_{k=j}.

Hence we arrive at

(7.58) PjψSjψj(0)L2+PjfNj+j(T)kχk=jPkψLtLr2.\begin{split}\|P_{j}\psi\|_{S_{j}}&\lesssim\|\psi_{j}(0)\|_{L^{2}}+\|P_{j}f\|_{N_{j}}+\mathcal{B}_{j}(T)\sum_{k}\chi_{k=j}\|P_{k}\psi\|_{L^{\infty}_{t}L^{2}_{r}}.\end{split}

Here we sum up with respect to jj. For j\mathcal{B}_{j} we know that

jj2(T)2(T).\sum_{j}\mathcal{B}_{j}^{2}(T)\lesssim\mathcal{B}^{2}(T).

Then, using Cauchy-Schwarz and a convolution bound for the last term on the right, we obtain

ψ1Sψ(0)LX¯+f1N+(T)ψ2LE\|\psi\|_{\ell^{1}S}\lesssim\|\psi(0)\|_{{L\bar{X}}}+\|f\|_{\ell^{1}N}+\mathcal{B}(T)\|\psi\|_{\ell^{2}LE}

Finally, using (7.39) for the last term we obtain (7.55).

8. Bounds for the nonlinearity N(ψ)N(\psi)

An important intermediate step in the proof of our results is to have good estimates on the nonlinear term N(ψ)N(\psi) appearing in the nonlinear equation (4.59) for ψ\psi. This is our task in this section. In a nutshell, the estimates in this section will allow us to treat N(ψ)N(\psi) in a perturbative fashion in two places: in the nonlinear analysis of the Schrödinger evolution (4.59) and in quantifying its effect in the dynamics of the modulation parameters λ\lambda and α\alpha. To serve these multiple goals, we include several estimates on N(ψ)N(\psi) in this section as follows:

  • Lemma 8.1 provides a simple estimate on N(ψ)N(\psi) which suffices to treat the effect of N(ψ)N(\psi) in (4.59) as perturbative when closing the theory in l2Sl^{2}S.

  • Lemma 8.3 refines the earlier result, showing that even when ψl2S\psi\in l^{2}S, the nonlinearity NN(\psi) belongs to l1Nl^{1}N; this indicates that the nonlinear effect on the dynamic variable ψ\psi is perturbative in the stronger structure l1Sl^{1}S, and allows us to close the l1Sl^{1}S bounds for ψ\psi even for large l1L2=LX¯l^{1}L^{2}={L\bar{X}} data.

  • The other results in Lemma 8.2 and Lemma 8.4 are needed in order to quantify the effect of N(ψ)N(\psi) in the dynamics of the modulation parameters λ\lambda and α\alpha.

For context, we begin by recalling the nonlinear equation for ψ\psi in (4.59),

(8.1) (itH~λ)ψ=N(ψ),ψ(0)=ψ0,(i\partial_{t}-\tilde{H}_{\lambda})\psi=N(\psi),\qquad\psi(0)=\psi_{0},

where

N(ψ)=Wλψ,Wλ=A02δλA2r21r(ψ2ψ¯)N(\psi)=W_{\lambda}\psi,\qquad W_{\lambda}=A_{0}-2\frac{\delta^{\lambda}A_{2}}{r^{2}}-\frac{1}{r}\Im({\psi}_{2}\bar{\psi})

with A2A_{2} and ψ2\psi_{2} uniquely determined by ψ\psi, see Proposition 4.7, and δλA2=A22h3λ\delta^{\lambda}A_{2}=A_{2}-2h_{3}^{\lambda}. A0A_{0} is given by (4.28) which we recall for convenience

A0(r)=12|ψ|2+1r(ψ2ψ¯)+[rr]1(|ψ|22r(ψ2ψ¯)).A_{0}(r)=-\frac{1}{2}|\psi|^{2}+\frac{1}{r}\Im(\psi_{2}\bar{\psi})+[r\partial_{r}]^{-1}(|\psi|^{2}-\frac{2}{r}\Im(\psi_{2}\bar{\psi})).

Thus we can rewrite

Wλ=12|ψ|22δλA2r2+[rr]1(|ψ|22r(ψ2ψ¯)).W_{\lambda}=-\frac{1}{2}|\psi|^{2}-2\frac{\delta^{\lambda}A_{2}}{r^{2}}+[r\partial_{r}]^{-1}(|\psi|^{2}-\frac{2}{r}\Im(\psi_{2}\bar{\psi})).

Our goal here is to estimate the nonlinear term N(ψ)=WλψN(\psi)=W_{\lambda}\psi in the above equation, which we will do in two different ways.

We make the following convention: in the statements below whenever a space-time is involved, the time interval is restricted to either [0,T][0,T] for some T>0T>0 or to [0,+)[0,+\infty); the estimates are uniform with respect to any such choice. We will not indicate this restriction at the level of notation, in order to keep notation compact.

The first result provides an estimate for WλψW_{\lambda}\psi in the admissible dual Strichartz space L43L^{\frac{4}{3}}, and will help us to interpret this term perturbatively when studying the evolution of ψ\psi in l2Sl^{2}S.

Lemma 8.1.

The nonlinearity N(ψ)N(\psi) satisfies

(8.2) WλψL43t,rψL4t,r(ψrL2t,r+ψLt,r42).\|W_{\lambda}\psi\|_{L^{\frac{4}{3}}_{t,r}}\lesssim\|\psi\|_{L^{4}_{t,r}}(\|\frac{\psi}{r}\|_{L^{2}_{t,r}}+\|\psi\|_{L_{t,r}^{4}}^{2}).
Proof.

To keep the arguments below compact we use the notation Lp:=Lpt,rL^{p}:=L^{p}_{t,r}. The non-integral terms are estimated as follows:

12|ψ|2ψL43ψ3L4,δλA2r2ψL43δλA2r2L2ψL4ψrL2ψL4,\|-\frac{1}{2}|\psi|^{2}\psi\|_{L^{\frac{4}{3}}}\lesssim\|\psi\|^{3}_{L^{4}},\quad\|\frac{\delta^{\lambda}A_{2}}{r^{2}}\psi\|_{L^{\frac{4}{3}}}\lesssim\|\frac{\delta^{\lambda}A_{2}}{r^{2}}\|_{L^{2}}\|\psi\|_{L^{4}}\lesssim\|\frac{\psi}{r}\|_{L^{2}}\|\psi\|_{L^{4}},

where in the last inequality we used (4.41). The integral terms are estimated using (2.8) and (4.40) as follows:

ψ[rr]1(|ψ|2)L43ψL4[rr]1(|ψ|2)L2ψL4|ψ|2L2ψ3L4,\|\psi\cdot[r\partial_{r}]^{-1}(|\psi|^{2})\|_{L^{\frac{4}{3}}}\lesssim\|\psi\|_{L^{4}}\|[r\partial_{r}]^{-1}(|\psi|^{2})\|_{L^{2}}\lesssim\|\psi\|_{L^{4}}\||\psi|^{2}\|_{L^{2}}\lesssim\|\psi\|^{3}_{L^{4}},
ψ[rr]1(δψ2rψ¯)L43ψL4[rr]1(δψ2rψ¯)L2ψL4δψ2rψ¯L2ψL42δψ2rL4ψL43,\|\psi\cdot[r\partial_{r}]^{-1}(\frac{\delta\psi_{2}}{r}\bar{\psi})\|_{L^{\frac{4}{3}}}\lesssim\|\psi\|_{L^{4}}\|[r\partial_{r}]^{-1}(\frac{\delta\psi_{2}}{r}\bar{\psi})\|_{L^{2}}\lesssim\|\psi\|_{L^{4}}\|\frac{\delta\psi_{2}}{r}\bar{\psi}\|_{L^{2}}\lesssim\|\psi\|_{L^{4}}^{2}\|\frac{\delta\psi_{2}}{r}\|_{L^{4}}\lesssim\|\psi\|_{L^{4}}^{3},
ψ[rr]1(h1λrψ¯)L43ψL4[rr]1(h1λψ¯r)L2ψL4h1λψ¯rL2ψL4ψrL2.\|\psi\cdot[r\partial_{r}]^{-1}(\frac{h_{1}^{\lambda}}{r}\bar{\psi})\|_{L^{\frac{4}{3}}}\lesssim\|\psi\|_{L^{4}}\|[r\partial_{r}]^{-1}(h_{1}^{\lambda}\frac{\bar{\psi}}{r})\|_{L^{2}}\lesssim\|\psi\|_{L^{4}}\|h_{1}^{\lambda}\frac{\bar{\psi}}{r}\|_{L^{2}}\lesssim\|\psi\|_{L^{4}}\|\frac{\psi}{r}\|_{L^{2}}.

Adding all of the above estimates gives us (8.2).

The next estimate for WλψW_{\lambda}\psi will be useful when considering its indirect contribution to the modulation equations, and show that it yields L1tL^{1}_{t} source terms there.

Lemma 8.2.

The following bounds hold true:

(8.3) WλL2t,rψrL2t,r+ψL4t,r2;\|W_{\lambda}\|_{L^{2}_{t,r}}\lesssim\|\frac{\psi}{r}\|_{L^{2}_{t,r}}+\|\psi\|_{L^{4}_{t,r}}^{2};
(8.4) WλψrL1t,rψrL2t,r(ψrL2t,r+ψL4t,r2).\|\frac{W_{\lambda}\psi}{r}\|_{L^{1}_{t,r}}\lesssim\|\frac{\psi}{r}\|_{L^{2}_{t,r}}(\|\frac{\psi}{r}\|_{L^{2}_{t,r}}+\|\psi\|_{L^{4}_{t,r}}^{2}).
Proof.

The estimate (8.4) is a direct consequence of (8.3), so we focus on the latter. The proof is fairly simple and it follows the same approach as the previous one, just that it relies more on the energy decay estimate. Indeed, the non-integral components are estimated as follows, using (4.41):

|ψ|2L2ψ2L4,δλA2r2L2ψrL2,\||\psi|^{2}\|_{L^{2}}\lesssim\|\psi\|^{2}_{L^{4}},\quad\|\frac{\delta^{\lambda}A_{2}}{r^{2}}\|_{L^{2}}\lesssim\|\frac{\psi}{r}\|_{L^{2}},

while the integral terms are estimated, using (2.8) and (4.40), as follows

[rr]1(|ψ|2)L2ψ2L4,[rr]1(δψ2rψ¯)L2ψrL2ψ2L4,\|[r\partial_{r}]^{-1}(|\psi|^{2})\|_{L^{2}}\lesssim\|\psi\|^{2}_{L^{4}},\quad\|[r\partial_{r}]^{-1}(\frac{\delta\psi_{2}}{r}\bar{\psi})\|_{L^{2}}\lesssim\|\frac{\psi}{r}\|_{L^{2}}\|\psi\|^{2}_{L^{4}},
[rr]1(h1λψ¯r)L2ψrL2.\|[r\partial_{r}]^{-1}(h_{1}^{\lambda}\frac{\bar{\psi}}{r})\|_{L^{2}}\lesssim\|\frac{\psi}{r}\|_{L^{2}}.

Obviously these estimates suffice to conclude with (8.3) and (8.4) follows from it.

The following result two results provide frequency localized refinement estimates for N(ψ)N(\psi). The first one highlights an improvement for the nonlinearity N(ψ)N(\psi) in that it gains the l1l^{1} summability even in the context of the problem with data with l2l^{2} structure; the precise statement is as follows.

Lemma 8.3.

The following holds true:

(8.5) N(ψ)l1Nψl2S2.\|N(\psi)\|_{l^{1}N}\lesssim\|\psi\|_{l^{2}S}^{2}.
Proof.

It is obvious that (8.5) follows from the following frequency localized version:

(8.6) PkλN(ψ)L43jk1,k3χ~k1=kχ~k3=kPk1λψSk1Pk3λψSk3,\|P_{k}^{\lambda}N(\psi)\|_{L^{\frac{4}{3}}_{j}}\lesssim\sum_{k_{1},k_{3}}\tilde{\chi}_{k_{1}=k}\tilde{\chi}_{k_{3}=k}\|P_{k_{1}}^{\lambda}\psi\|_{S_{k_{1}}}\|P_{k_{3}}^{\lambda}\psi\|_{S_{k_{3}}},

where χ~i=j=2|ij|10\tilde{\chi}_{i=j}=2^{-\frac{|i-j|}{10}}. The rest of the proof is concerned with (8.6). Recall that

N(ψ)=Wλψ,Wλ=12|ψ|22δλA2r2+[rr]1(|ψ|22r(ψ2ψ¯)).N(\psi)=W_{\lambda}\cdot\psi,\quad W_{\lambda}=-\frac{1}{2}|\psi|^{2}-2\frac{\delta^{\lambda}A_{2}}{r^{2}}+[r\partial_{r}]^{-1}(|\psi|^{2}-\frac{2}{r}\Im(\psi_{2}\bar{\psi})).

We start with the cubic term. It follows directly from the definition of the spaces and the characterization of the kernel of PkλP_{k}^{\lambda} in (5.110) that

Pkλ(Pk1λψPk2λψ¯Pk3λψ)L43kχ~k1=kχ~k2=kχ~k3=kPk1λψL4k1Pk2λψL4k2Pk3λψL4k3.\|P_{k}^{\lambda}(P_{k_{1}}^{\lambda}\psi P_{k_{2}}^{\lambda}\overline{\psi}P_{k_{3}}^{\lambda}\psi)\|_{L^{\frac{4}{3}}_{k}}\lesssim\tilde{\chi}_{k_{1}=k}\tilde{\chi}_{k_{2}=k}\tilde{\chi}_{k_{3}=k}\|P_{k_{1}}^{\lambda}\psi\|_{L^{4}_{k_{1}}}\|P_{k_{2}}^{\lambda}\psi\|_{L^{4}_{k_{2}}}\|P_{k_{3}}^{\lambda}\psi\|_{L^{4}_{k_{3}}}.

Next we use (2.9) with the choice w(r)=(2k1r)19,r2k1w(r)=(2^{k_{1}}r)^{-\frac{1}{9}},r\leq 2^{-k_{1}} and (2k1r)19,r2k1(2^{k_{1}}r)^{\frac{1}{9}},r\geq 2^{-k_{1}} to obtain

w[rr]1(Pk1λψψ¯)L2Pk1λψL4k1ψL4.\|w[r\partial_{r}]^{-1}(P_{k_{1}}^{\lambda}\psi\cdot\overline{\psi})\|_{L^{2}}\leq\|P_{k_{1}}^{\lambda}\psi\|_{L^{4}_{k_{1}}}\|\psi\|_{L^{4}}.

From this, the definition of the spaces and the characterization of the kernel of PkλP_{k}^{\lambda} in (5.110) it follows that

Pkλ([rr]1(Pk1λψψ¯)Pk3λψ)L43kχ~k1=kχ~k3=kPk1λψL4k1ψL4Pk3λψL4k3.\|P_{k}^{\lambda}([r\partial_{r}]^{-1}(P_{k_{1}}^{\lambda}\psi\cdot\overline{\psi})\cdot P_{k_{3}}^{\lambda}\psi)\|_{L^{\frac{4}{3}}_{k}}\lesssim\tilde{\chi}_{k_{1}=k}\tilde{\chi}_{k_{3}=k}\|P_{k_{1}}^{\lambda}\psi\|_{L^{4}_{k_{1}}}\|\psi\|_{L^{4}}\|P_{k_{3}}^{\lambda}\psi\|_{L^{4}_{k_{3}}}.

From the definition of the space LEkLE_{k} and (7.8), we obtain

PjλψrL2tL2(Am)2k+m2PjλψLEj.\|\frac{P_{j}^{\lambda}\psi}{r}\|_{L^{2}_{t}L^{2}(A_{m})}\lesssim 2^{-\frac{k+m}{2}}\|P_{j}^{\lambda}\psi\|_{LE_{j}}.

From this it follows that with w1(r)=(2jr)13,r2jw_{1}(r)=(2^{j}r)^{-\frac{1}{3}},r\leq 2^{-j} and (2jr)19,r2j(2^{j}r)^{\frac{1}{9}},r\geq 2^{-j} we have

w1PjλψrL2t,rPjλψLEj.\|w_{1}\frac{P_{j}^{\lambda}\psi}{r}\|_{L^{2}_{t,r}}\lesssim\|P_{j}^{\lambda}\psi\|_{LE_{j}}.

Invoking (2.9) (and the trivial fact that ψ2L1\|\psi_{2}\|_{L^{\infty}}\leq 1) gives us

(8.7) w1[rr]1(ψ¯2Pjλψr)L2t,rPjλψLEj.\|w_{1}[r\partial_{r}]^{-1}(\overline{\psi}_{2}\frac{P_{j}^{\lambda}\psi}{r})\|_{L^{2}_{t,r}}\lesssim\|P_{j}^{\lambda}\psi\|_{LE_{j}}.

Using this, the definition of the spaces and the characterization of the kernel of PkλP_{k}^{\lambda} in (5.110), we obtain the following estimate

Pkλ([rr]1(ψ¯2Pk1λψr)Pk3λψ)L43kχ~k1=kχ~k3=kPk1λψLEk1Pk3λψL4k3.\|P_{k}^{\lambda}([r\partial_{r}]^{-1}(\overline{\psi}_{2}\frac{P_{k_{1}}^{\lambda}\psi}{r})P_{k_{3}}^{\lambda}\psi)\|_{L^{\frac{4}{3}}_{k}}\lesssim\tilde{\chi}_{k_{1}=k}\tilde{\chi}_{k_{3}=k}\|P_{k_{1}}^{\lambda}\psi\|_{LE_{k_{1}}}\|P_{k_{3}}^{\lambda}\psi\|_{L^{4}_{k_{3}}}.

We note that in all terms estimated above we took advantage of the fact that they contain (at least) two factors of ψ\psi which are easily amenable to frequency localization since it was important that we could use localization for at least two terms. This makes the last term in the nonlinearity δλA2r2ψ\frac{\delta^{\lambda}A_{2}}{r^{2}}\cdot\psi more challenging; to be more precise, the challenging aspect is to bring some sort of frequency localization information on the term δλA2\delta^{\lambda}A_{2}. The transition from ψ\psi to δλ,αψ2\delta^{\lambda,\alpha}\psi_{2} and δλA2\delta^{\lambda}A_{2} is governed by the system (4.42) (which was studied in detail in the proof of Lemma 4.7); for convenience we recall the system here

(8.8) {Lλδλ,αψ2=2ih3λψ+δλA2ψ1rδλA2(2ie2iαh1λ+δλ,αψ2),LλδλA2=2h1λ(e2iαψ)+(ψδλ,αψ¯2)(δλA2)2r.\left\{\begin{array}[]{l}L_{\lambda}\delta^{\lambda,\alpha}\psi_{2}=2ih_{3}^{\lambda}\psi+\delta^{\lambda}A_{2}\psi-\frac{1}{r}\delta^{\lambda}A_{2}(2ie^{2i\alpha}h_{1}^{\lambda}+\delta^{\lambda,\alpha}\psi_{2}),\cr\cr L_{\lambda}\delta^{\lambda}A_{2}=-2h_{1}^{\lambda}\Re{(e^{2i\alpha}\psi)}+\Im(\psi\overline{\delta^{\lambda,\alpha}\psi}_{2})-\frac{(\delta^{\lambda}A_{2})^{2}}{r}.\end{array}\right.

The idea is to localize ψ\psi in frequency and solve for the corresponding δλ,αψ2\delta^{\lambda,\alpha}\psi_{2} and δλA2\delta^{\lambda}A_{2}; however the system is nonlinear, and this requires some care. A simple alternative is to force linearity in this system by letting δkλ,αψ2\delta_{k}^{\lambda,\alpha}\psi_{2} and δλkA2\delta^{\lambda}_{k}A_{2} be the solutions of the following system

(8.9) {Lλδkλ,αψ2=2ih3λPkλψ+δλA2Pkλψ1rδλkA2(2ie2iαh1λ+δλ,αψ2),LλδkλA2=2h1λ(e2iαPkλψ)+(Pkλψδλ,αψ¯2)δλA2δkλA2r,\left\{\begin{array}[]{l}L_{\lambda}\delta_{k}^{\lambda,\alpha}\psi_{2}=2ih_{3}^{\lambda}P_{k}^{\lambda}\psi+\delta^{\lambda}A_{2}P_{k}^{\lambda}\psi-\frac{1}{r}\delta^{\lambda}_{k}A_{2}(2ie^{2i\alpha}h_{1}^{\lambda}+\delta^{\lambda,\alpha}\psi_{2}),\cr\cr L_{\lambda}\delta_{k}^{\lambda}A_{2}=-2h_{1}^{\lambda}\Re{(e^{2i\alpha}P_{k}^{\lambda}\psi)}+\Im(P_{k}^{\lambda}\psi\overline{\delta^{\lambda,\alpha}\psi}_{2})-\frac{\delta^{\lambda}A_{2}\cdot\delta_{k}^{\lambda}A_{2}}{r},\end{array}\right.

with the same set of initial data as in Lemma 4.7

δkλ,αψ2(r0)=0,δkλ,αA2(r0)=0.\delta_{k}^{\lambda,\alpha}\psi_{2}(r_{0})=0,\qquad\delta_{k}^{\lambda,\alpha}A_{2}(r_{0})=0.

Here we recall that r0λ1r_{0}\approx\lambda^{-1}. It is clear that this provides a linear decomposition of δλ,αψ2\delta^{\lambda,\alpha}\psi_{2} and δλA2\delta^{\lambda}A_{2} as follows

δλ,αψ2=kδkλ,αψ2,δλA2=kδkλ,αA2.\delta^{\lambda,\alpha}\psi_{2}=\sum_{k\in\mathbb{Z}}\delta_{k}^{\lambda,\alpha}\psi_{2},\qquad\delta^{\lambda}A_{2}=\sum_{k\in\mathbb{Z}}\delta_{k}^{\lambda,\alpha}A_{2}.

Given that we already have the apriori knowledge δλ,αψ2L+δλA2LψL21\|\delta^{\lambda,\alpha}\psi_{2}\|_{L^{\infty}}+\|\delta^{\lambda}A_{2}\|_{L^{\infty}}\lesssim\|\psi\|_{L^{2}}\ll 1 from Lemma 4.7, it suffices to analyse only the equation for δλkA2\delta^{\lambda}_{k}A_{2}. This can be rewritten in the integral form:

δλkA2(r)=hλ1(r)r0r1h1λ(2hλ1(e2iαPkλψ)+(Pkλψδλ,αψ¯2)δλA2δkλA2s)ds.\delta^{\lambda}_{k}A_{2}(r)=h^{\lambda}_{1}(r)\int_{r_{0}}^{r}\frac{1}{h_{1}^{\lambda}}\left(-2h^{\lambda}_{1}\Re{(e^{2i\alpha}P_{k}^{\lambda}\psi)}+\Im(P_{k}^{\lambda}\psi\overline{\delta^{\lambda,\alpha}\psi}_{2})-\frac{\delta^{\lambda}A_{2}\cdot\delta_{k}^{\lambda}A_{2}}{s}\right)ds.

We record the following simple variant of (4.43):

(8.10) wrh1λr0r1h1λfdsLpwfLp,1p,\|\frac{w}{r}h_{1}^{\lambda}\int_{r_{0}}^{r}\frac{1}{h_{1}^{\lambda}}fds\|_{L^{p}}\lesssim\|wf\|_{L^{p}},\forall 1\leq p\leq\infty,

under the following assumptions on ww:

- w:++w:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+} is slowly varying, that is w(r1)w(r2)w(r_{1})\approx w(r_{2}) for 21r2r12r22^{-1}r_{2}\leq r_{1}\leq 2r_{2};

- r1w(r)r^{-1}w(r) is decreasing and r12w(r)r^{\frac{1}{2}}w(r) is increasing.

The second condition implies the slowly varying one, but we have stated them separately in order to also emphasize the first. The only role that r0r_{0} plays in the above inequality is that is sits at the ”height” of h1λh_{1}^{\lambda}, that is h1λ(r0)1h_{1}^{\lambda}(r_{0})\approx 1 and h1λh_{1}^{\lambda} decays away from a neighborhood of size λ1\lambda^{-1} of r0r_{0}.

Next we use (8.10) with the choice wk1(r)=(2k1r)19,r2k1w_{k_{1}}(r)=(2^{k_{1}}r)^{-\frac{1}{9}},r\leq 2^{-k_{1}} and (2k1r)19,r2k1(2^{k_{1}}r)^{\frac{1}{9}},r\geq 2^{-k_{1}} to obtain

wk1rhλ1(r)r0r1h1λ(2hλ1(e2iαPk1λψ)+(Pk1λψδλ,αψ¯2))dsL4wk1Pk1λψL4Pk1λψL4k1.\|\frac{w_{k_{1}}}{r}h^{\lambda}_{1}(r)\int_{r_{0}}^{r}\frac{1}{h_{1}^{\lambda}}\left(-2h^{\lambda}_{1}\Re{(e^{2i\alpha}P_{k_{1}}^{\lambda}\psi)}+\Im(P_{k_{1}}^{\lambda}\psi\overline{\delta^{\lambda,\alpha}\psi}_{2})\right)ds\|_{L^{4}}\lesssim\|w_{k_{1}}P_{k_{1}}^{\lambda}\psi\|_{L^{4}}\lesssim\|P_{k_{1}}^{\lambda}\psi\|_{L^{4}_{k_{1}}}.

This suggests that, at least at the linear level, the following holds true

(8.11) wk1δλk1A2rL4Pk1λψL4k1.\|w_{k_{1}}\frac{\delta^{\lambda}_{k_{1}}A_{2}}{r}\|_{L^{4}}\lesssim\|P_{k_{1}}^{\lambda}\psi\|_{L^{4}_{k_{1}}}.

The quadratic contributions (coming from the term δλA2δk1λA2s\frac{\delta^{\lambda}A_{2}\cdot\delta_{k_{1}}^{\lambda}A_{2}}{s}) are then estimated using the apriori assumption that δλA2LψL21\|\delta^{\lambda}A_{2}\|_{L^{\infty}}\lesssim\|\psi\|_{L^{2}}\ll 1; the actual justification for (8.11) can be done using a continuity/bootstrap argument on intervals [r0,R][r_{0},R] and [r,r0][r,r_{0}] for any rr0Rr\leq r_{0}\leq R.

Using (8.11) and (8.7), as well as the definition of L43kL^{\frac{4}{3}}_{k} and the characterizayion of the lernel of PkλP_{k}^{\lambda} in (5.110), allows us to estimate

Pkλ(δλk1A2rPk3ψr)L43kχ~k1=kχ~k3=kPk1λψL4k1Pk3λψLEk3.\|P_{k}^{\lambda}(\frac{\delta^{\lambda}_{k_{1}}A_{2}}{r}\cdot\frac{P_{k_{3}}\psi}{r})\|_{L^{\frac{4}{3}}_{k}}\lesssim\tilde{\chi}_{k_{1}=k}\tilde{\chi}_{k_{3}=k}\|P_{k_{1}}^{\lambda}\psi\|_{L^{4}_{k_{1}}}\|P_{k_{3}}^{\lambda}\psi\|_{LE_{k_{3}}}.

This finishes the proof of (8.6) and in turn, the proof of the Lemma. ∎

For technical reasons we also need the following result (which will be used in the proof of Lemma 10.2, precisely in estimating the term ek2e_{k}^{2} in Step 2).

Lemma 8.4.

The following holds true:

(8.12) PkN(ψ)rL2tL1r2kk2|kk|10PkλψSk.\|\frac{P_{k}N(\psi)}{r}\|_{L^{2}_{t}L^{1}_{r}}\lesssim 2^{k}\sum_{k^{\prime}}2^{-\frac{|k-k^{\prime}|}{10}}\|P_{k^{\prime}}^{\lambda}\psi\|_{S_{k^{\prime}}}.
Proof.

It is obvious that it suffices to establish the following estimate

(8.13) Pkλ(WλPkλψ)rL2tL1r2k2|kk|10PkλψSkψS(1+ψS).\|\frac{P_{k}^{\lambda}(W_{\lambda}\cdot P_{k^{\prime}}^{\lambda}\psi)}{r}\|_{L^{2}_{t}L^{1}_{r}}\lesssim 2^{k}2^{-\frac{|k-k^{\prime}|}{10}}\|P_{k^{\prime}}^{\lambda}\psi\|_{S_{k^{\prime}}}\|\psi\|_{S}(1+\|\psi\|_{S}).

In order to prove (8.13), we need the following technical result:

(8.14) PkλfrL2tL1r2km2|m+k|9fL2tL1r(Am).\|\frac{P_{k}^{\lambda}f}{r}\|_{L^{2}_{t}L^{1}_{r}}\lesssim 2^{k}\sum_{m\in\mathbb{Z}}2^{-\frac{|m+k|}{9}}\|f\|_{L^{2}_{t}L^{1}_{r}(A_{m})}.

The argument is fairly straightforward; we start with

PkλfrL2tL1r(An)2nPkλfL2tL1r(An)2nmK~λk(,s)χAm(s)f(s)sdsL2tL1r(An)N2nmωk,λ(2n)22kωk,λ(s)χAm(s)f(s)(1+2k(2m+2n))(1+2k|rs|)NsdsL2tL1r(An),\begin{split}\|\frac{P_{k}^{\lambda}f}{r}\|_{L^{2}_{t}L^{1}_{r}(A_{n})}&\approx 2^{-n}\|P_{k}^{\lambda}f\|_{L^{2}_{t}L^{1}_{r}(A_{n})}\\ &\lesssim 2^{-n}\sum_{m\in\mathbb{Z}}\|\int\tilde{K}^{\lambda}_{k}(\cdot,s)\chi_{A_{m}}(s)f(s)sds\|_{L^{2}_{t}L^{1}_{r}(A_{n})}\\ &\lesssim_{N}2^{-n}\sum_{m\in\mathbb{Z}}\omega_{k,\lambda}(2^{n})\|\int\frac{2^{2k}\omega_{k,\lambda}(s)\chi_{A_{m}}(s)f(s)}{(1+2^{k}(2^{m}+2^{n}))(1+2^{k}|r-s|)^{-N}}sds\|_{L^{2}_{t}L^{1}_{r}(A_{n})},\end{split}

where in passing to the last line we have used the bound (5.110) on the kernel K~λk\tilde{K}^{\lambda}_{k}; in what follows below we heavily rely on the bounds on ωk,λ\omega_{k,\lambda} from (5.108).

If n+k0n+k\leq 0, then we can further bound the above quantity by

2n+2kmkωk,λ(2n)ωk,λ(2m)fL2tL1r(Am)+2n+2kmkωk,λ(2n)2N(k+m)fL2tL1r(Am)2n+2km2|k+m|fL2tL1r(Am).\begin{split}&2^{n+2k}\sum_{m\leq-k}\omega_{k,\lambda}(2^{n})\omega_{k,\lambda}(2^{m})\|f\|_{L^{2}_{t}L^{1}_{r}(A_{m})}+2^{n+2k}\sum_{m\geq-k}\omega_{k,\lambda}(2^{n})2^{-N(k+m)}\|f\|_{L^{2}_{t}L^{1}_{r}(A_{m})}\\ \lesssim&2^{n+2k}\sum_{m\in\mathbb{Z}}2^{-|k+m|}\|f\|_{L^{2}_{t}L^{1}_{r}(A_{m})}.\end{split}

If n+k0n+k\geq 0, then we can further bound the above quantity by

2n+2kmkωk,λ(2m)2N(k+n)fL2tL1r(Am)+2kmk(2k+m+2k+n)1fL2tL1r(Am)2km2|k+m|+|k+n|2fL2tL1r(Am).\begin{split}&2^{n+2k}\sum_{m\leq-k}\omega_{k,\lambda}(2^{m})2^{-N(k+n)}\|f\|_{L^{2}_{t}L^{1}_{r}(A_{m})}+2^{k}\sum_{m\geq-k}(2^{k+m}+2^{k+n})^{-1}\|f\|_{L^{2}_{t}L^{1}_{r}(A_{m})}\\ \lesssim&2^{k}\sum_{m\in\mathbb{Z}}2^{-\frac{|k+m|+|k+n|}{2}}\|f\|_{L^{2}_{t}L^{1}_{r}(A_{m})}.\end{split}

Wrapping things up, we have succeeded to prove

PkλfrL2tL1r(An)2km2|k+m|+|k+n|2fL2tL1r(Am),\|\frac{P_{k}^{\lambda}f}{r}\|_{L^{2}_{t}L^{1}_{r}(A_{n})}\lesssim 2^{k}\sum_{m\in\mathbb{Z}}2^{-\frac{|k+m|+|k+n|}{2}}\|f\|_{L^{2}_{t}L^{1}_{r}(A_{m})},

from which (8.14) follows.

Now we proceed with the proof of (8.13). For convenience recall the formula for WλW_{\lambda},

Wλ=12|ψ|22δλA2r2+[rr]1(|ψ|22r(ψ2ψ¯)).W_{\lambda}=-\frac{1}{2}|\psi|^{2}-2\frac{\delta^{\lambda}A_{2}}{r^{2}}+[r\partial_{r}]^{-1}(|\psi|^{2}-\frac{2}{r}\Im(\psi_{2}\bar{\psi})).

For the cubic component we note the following estimate

supm2|m+k|8|ψ|2PkλψL2tL1r(Am)PkλψL4kψL4ψLtL2r,\sup_{m\in\mathbb{Z}}2^{\frac{|m+k^{\prime}|}{8}}\||\psi|^{2}P_{k^{\prime}}^{\lambda}\psi\|_{L^{2}_{t}L^{1}_{r}(A_{m})}\lesssim\|P_{k^{\prime}}^{\lambda}\psi\|_{L^{4}_{k^{\prime}}}\|\psi\|_{L^{4}}\|\psi\|_{L^{\infty}_{t}L^{2}_{r}},

followed by

m2|m+k|9|ψ|2PkλψL2tL1r(Am)PkλψL4kψL4ψLtL2r.\sum_{m\in\mathbb{Z}}2^{\frac{|m+k^{\prime}|}{9}}\||\psi|^{2}P_{k^{\prime}}^{\lambda}\psi\|_{L^{2}_{t}L^{1}_{r}(A_{m})}\lesssim\|P_{k^{\prime}}^{\lambda}\psi\|_{L^{4}_{k^{\prime}}}\|\psi\|_{L^{4}}\|\psi\|_{L^{\infty}_{t}L^{2}_{r}}.

At this point we invoke (8.14) to conclude that

Pkλ(|ψ|2Pkλψ)rL2tL1r2k2|kk|10PkλψSkψS2,\|\frac{P_{k}^{\lambda}(|\psi|^{2}P_{k^{\prime}}^{\lambda}\psi)}{r}\|_{L^{2}_{t}L^{1}_{r}}\lesssim 2^{k}2^{-\frac{|k-k^{\prime}|}{10}}\|P_{k^{\prime}}^{\lambda}\psi\|_{S_{k^{\prime}}}\|\psi\|_{S}^{2},

which implies the claim made in (8.13) for the corresponding cubic component in N(ψ)N(\psi). In order to extend the arguments above to the other components of N(ψ)N(\psi), it is important to note that for the quadratic term |ψ|2|\psi|^{2} all that we have used was an estimate in L4tL43rL^{4}_{t}L^{\frac{4}{3}}_{r} (by invoking the L4t,rL^{4}_{t,r} information on one of the ψ\psi’s and LtL2rL^{\infty}_{t}L^{2}_{r} on the other ψ\psi).

Now we turn to the other components in WλW_{\lambda}. We first deal with the component which contains the operator [rr]1[r\partial_{r}]^{-1}; from (2.8) we know that the operator [rr]1[r\partial_{r}]^{-1} is bounded on L43rL^{\frac{4}{3}}_{r}, and, as a consequence, it is bounded on L4tL43rL^{4}_{t}L^{\frac{4}{3}}_{r} - as pointed earlier this is all that was needed for the term |ψ|2|\psi|^{2} in the analysis of the cubic term above. From these considerations it follows that the term Pkλψ[rr]1(|ψ|2)P_{k^{\prime}}^{\lambda}\psi\cdot[r\partial_{r}]^{-1}(|\psi|^{2}) can be treated in a similar fashion to the cubic one above.

In analyzing the term [rr]1(2r(ψ2ψ¯))[r\partial_{r}]^{-1}(\frac{2}{r}\Im(\psi_{2}\bar{\psi})) we write ψ2=δλ,αψ2+2e2iαh1λ\psi_{2}=\delta^{\lambda,\alpha}\psi_{2}+2e^{2i\alpha}h_{1}^{\lambda}. The term Pkλψ[rr]1(δλ,αψ2rψ¯)P_{k^{\prime}}^{\lambda}\psi\cdot[r\partial_{r}]^{-1}(\frac{\delta^{\lambda,\alpha}\psi_{2}}{r}\overline{\psi}) is treated as above (similar to the cubic component) by using the LtL2rL^{\infty}_{t}L^{2}_{r} bound on δλ,αψ2r\frac{\delta^{\lambda,\alpha}\psi_{2}}{r} from (4.39) and the mapping properties of the operator [rr]1[r\partial_{r}]^{-1}; the term Pkλψ[rr]1(h1λrψ¯)P_{k^{\prime}}^{\lambda}\psi\cdot[r\partial_{r}]^{-1}(\frac{h_{1}^{\lambda}}{r}\overline{\psi}) is entirely similar since h1λrL2r=h1rL2r1\|\frac{h_{1}^{\lambda}}{r}\|_{L^{2}_{r}}=\|\frac{h_{1}}{r}\|_{L^{2}_{r}}\lesssim 1 and this bound is independent of time (thus uniform in time).

From the above we conclude with the following estimate

Pkλ(Pkλψ[rr]1(|ψ|22r(ψ2ψ¯)))rL2tL1r2k2|kk|10PkλψSkψS(1+ψS),\|\frac{P_{k}^{\lambda}(P_{k^{\prime}}^{\lambda}\psi\cdot[r\partial_{r}]^{-1}(|\psi|^{2}-\frac{2}{r}\Im(\psi_{2}\bar{\psi})))}{r}\|_{L^{2}_{t}L^{1}_{r}}\lesssim 2^{k}2^{-\frac{|k-k^{\prime}|}{10}}\|P_{k^{\prime}}^{\lambda}\psi\|_{S_{k^{\prime}}}\|\psi\|_{S}(1+\|\psi\|_{S}),

Finally we deal with the δλA2r2\frac{\delta^{\lambda}A_{2}}{r^{2}} component in WλW_{\lambda}. From (7.12) and (7.13), it follows that

m2|k+m|4PkλψrL2t,r(Am)PkλψLEk.\sum_{m\in\mathbb{Z}}2^{\frac{|k^{\prime}+m|}{4}}\|\frac{P_{k^{\prime}}^{\lambda}\psi}{r}\|_{L^{2}_{t,r}(A_{m})}\lesssim\|P_{k^{\prime}}^{\lambda}\psi\|_{LE_{k^{\prime}}}.

From this and (4.39) we obtain that

m2|k+m|4δλA2rPkλψrL2tL1r(Am)δλA2rLtL2rPkλψLEkψLtL2rPkλψLEk.\sum_{m\in\mathbb{Z}}2^{\frac{|k^{\prime}+m|}{4}}\|\frac{\delta^{\lambda}A_{2}}{r}\cdot\frac{P_{k^{\prime}}^{\lambda}\psi}{r}\|_{L^{2}_{t}L^{1}_{r}(A_{m})}\lesssim\|\frac{\delta^{\lambda}A_{2}}{r}\|_{L^{\infty}_{t}L^{2}_{r}}\|P_{k^{\prime}}^{\lambda}\psi\|_{LE_{k^{\prime}}}\lesssim\|\psi\|_{L^{\infty}_{t}L^{2}_{r}}\|P_{k^{\prime}}^{\lambda}\psi\|_{LE_{k^{\prime}}}.

Combining this with (8.14) gives the following:

Pkλ(PkλψδλA2r2)rL2tL1r2k2|kk|10PkλψSkψS(1+ψS).\|\frac{P_{k}^{\lambda}(P_{k^{\prime}}^{\lambda}\psi\cdot\frac{\delta^{\lambda}A_{2}}{r^{2}})}{r}\|_{L^{2}_{t}L^{1}_{r}}\lesssim 2^{k}2^{-\frac{|k-k^{\prime}|}{10}}\|P_{k^{\prime}}^{\lambda}\psi\|_{S_{k^{\prime}}}\|\psi\|_{S}(1+\|\psi\|_{S}).

This finishes the proof of (8.13), and in turn of our Lemma. ∎

9. The modulation parameter dynamics - Part 1

The main goal of this section is to prove Theorems 1.3 and 1.4. Those theorems are mostly stated at the level of the map uu, although the field ψ\psi appears in their conclusions, see (1.11) and (1.16). Below we restate those Theorems fully in terms of ψ\psi, see Theorems 9.1 and 9.2, and prove the latter ones; Theorem 1.3 and Theorem 9.1 are identical, while the equivalence between Theorem 1.4 and Theorem 9.2 follows directly from Proposition 6.1.

As described in Section 4.4, H˙1\dot{H}^{1} states for the Schrödinger Map flow are described via the differentiated L2L^{2} field ψ\psi and the modulation parameters α\alpha and λ\lambda, which are chosen via the orthogonality condition (4.31) (in conjunction with (4.32) in order to ensure uniqueness).

Dynamically, we expect the Schrödinger maps evolution to be governed by a coupled system consisting of a Schrödinger type evolution for ψ\psi coupled with appropriate modulation equations for α\alpha and λ\lambda. The equation for ψ\psi has already been derived in (4.59), and contains two types of nonlinear effects due to (i) the nonlinear terms on the right hand side of (4.59), and (ii) the time dependence of λ\lambda in the left hand side of (4.59). The former are at least quadratic, and will play a perturbative role in our analysis. The effect of λ\lambda is stronger, and was investigated in the last two sections; as seen there, the crucial control is that of λλ2L2\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}}.

In this section we begin investigating the other half of the story, namely the evolution of the modulation parameters α\alpha and λ\lambda. Our goal is twofold:

  1. (1)

    We derive the modulation equation for λ\lambda and α\alpha as an ode system, separating the contributions of ψ\psi into a leading order linear part and a perturbative quadratic part. In this analysis the essential control is that of the local energy decay of ψ\psi, primarily ψrL2t,r\|\frac{\psi}{r}\|_{L^{2}_{t,r}}.

  2. (2)

    We combine the bounds for the ψ\psi equation with the bounds for the modulation equation in order to simultaneously close the local energy bounds for ψ\psi and the modulation equation bound for λλ2L2\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}}.

The second part above is carried out on the full time of existence of the solutions, regardless of whether this time is finite or infinite. The result is stated as follows:

Theorem 9.1.

Assume that we have a 22-equivariant initial data u0H˙1u_{0}\in\dot{H}^{1} in the homotopy class of Q2Q^{2}, and with energy below 8π+δ28\pi+\delta^{2}, and with δ\delta sufficiently small. Let Imax=(Tmin,Tmax)I_{max}=(T_{min},T_{max}) be the maximal time of existence of the solution uu to the Schrödinger map flow (1.1) with initial data u0u_{0}. Then the associated Coulomb gauge field ψ\psi and the associated parameter λ\lambda (by the rule (4.31)) satisfy

(9.1) λλ2L2t(Imax)+αλL2t(Imax)ψ(0)L2,\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}_{t}(I_{max})}+\|\frac{\alpha^{\prime}}{\lambda}\|_{L^{2}_{t}(I_{max})}\lesssim\|\psi(0)\|_{L^{2}},

and

(9.2) ψl2S(Imax)ψ(0)L2.\|\psi\|_{l^{2}S(I_{max})}\lesssim\|\psi(0)\|_{L^{2}}.

Later on it will also be useful to have an l1l^{1} version of the above theorem, more precisely of the estimate (9.2):

Theorem 9.2.

Let uu be the solution uu to the Schrödinger map flow (1.1) as in Theorem 9.1. Assume in addition that ψ(0)LX¯\psi(0)\in{L\bar{X}}. Then we have

(9.3) ψl1S(Imax)ψ(0)LX¯.\|\psi\|_{l^{1}S(I_{max})}\lesssim\|\psi(0)\|_{{L\bar{X}}}.

A density argument using the local well-posedness result in Theorem 1.1 shows that it suffices to prove these results for H˙1H˙2\dot{H}^{1}\cap\dot{H}^{2} solutions. The forward in time and backward in time problems are similar, so for simplicity we drop the superscripts and we will work only on [0,Tmax)[0,T_{max}) in what follows.

This theorem provides two important pieces of information. Concerning the field ψ\psi, it asserts that its dispersive properties (measured by local energy decay and Strichartz norms) persist on the maximal time of existence; in other words potential blow-up in finite or infinite time does not interfere with the dispersion.

Concerning the parameter λ\lambda, the estimate (9.1) prevents the scenario that λ0\lambda\rightarrow 0 in finite time (this can also be seen as a consequence of the local result in Theorem 1.1). However, it does not prevent the scenarios λ0\lambda\rightarrow 0 in infinite time, or λ\lambda\rightarrow\infty in finite or infinite time. In the next section we will refine our analysis of the ODE system so that we can investigate these other potential scenarios.

9.1. The modulation equation

In this section we derive the modulation equation for α,λ\alpha,\lambda from the orthogonality condition (4.29). From Corollary (4.9) we already know that λ\lambda and α\alpha are differentiable in time on the maximal interval of existence. Hence we can differentiate (4.29) with respect to time. The computations below are formally justified for more regular solutions as in Theorem  1.1; such solutions satisfy in particular ψLH1\psi\in L^{\infty}H^{1}. However, the final outcome, namely the modulation equations, are valid for all finite energy solutions via a density argument.

We differentiate (4.29) with respect to time and use (2.1) to obtain the following

0=tψ22ie2iα(t)h1(λr),ϰ(λr)=tψ2,ϰ(λr)+4αe2iα(t)h1(λr),ϰ(λr)+4iλe2iα(t)rh1(λr)h3(λr)λr,ϰ(λr)+λψ22ie2iα(t)h1(λr),rϰ(λr)=tψ2,ϰ(λr)+4e2iα(t)(αλ2h1,χ+iλλ3h1h3,ϰ)+λλ3ψ2(λ1r)2ie2iα(t)h1(r),rϰ(r).\begin{split}0=&\ \partial_{t}\langle\psi_{2}-2ie^{2i\alpha(t)}h_{1}(\lambda r),\varkappa(\lambda r)\rangle\\ =&\ \langle\partial_{t}\psi_{2},\varkappa(\lambda r)\rangle+4\alpha^{\prime}\langle e^{2i\alpha(t)}h_{1}(\lambda r),\varkappa(\lambda r)\rangle\\ &+4i\lambda^{\prime}\langle e^{2i\alpha(t)}r\frac{h_{1}(\lambda r)h_{3}(\lambda r)}{\lambda r},\varkappa(\lambda r)\rangle+\lambda^{\prime}\langle\psi_{2}-2ie^{2i\alpha(t)}h_{1}(\lambda r),r\varkappa^{\prime}(\lambda r)\rangle\\ =&\ \langle\partial_{t}\psi_{2},\varkappa(\lambda r)\rangle+4e^{2i\alpha(t)}\left(\frac{\alpha^{\prime}}{\lambda^{2}}\langle h_{1},\chi\rangle+i\frac{\lambda^{\prime}}{\lambda^{3}}\langle h_{1}h_{3},\varkappa\rangle\right)\\ &+\frac{\lambda^{\prime}}{\lambda^{3}}\langle\psi_{2}(\lambda^{-1}r)-2ie^{2i\alpha(t)}h_{1}(r),r\varkappa^{\prime}(r)\rangle.\end{split}

A closer look at the last expression in the sequence of equalities reveals the following:

  • the second term 4e2iα(t)(αλ2h1,ϰ+iλλ3h1h3,ϰ)4e^{2i\alpha(t)}\left(\frac{\alpha^{\prime}}{\lambda^{2}}\langle h_{1},\varkappa\rangle+i\frac{\lambda^{\prime}}{\lambda^{3}}\langle h_{1}h_{3},\varkappa\rangle\right) contains the primary dynamical information about the two real parameters α\alpha and λ\lambda due to the nondegeneracy conditions (4.30)

    (9.4) c1:=h1,ϰ0,c2:=h1h3,ϰ0;c_{1}:=\langle h_{1},\varkappa\rangle\neq 0,\qquad c_{2}:=\langle h_{1}h_{3},\varkappa\rangle\neq 0;

    in fact it is precisely the above computation (and later consideration) that justfies imposing these specific non-degenaracy conditions;

  • the last term λλ3ψ2(λ1r)2ie2iα(t)h1(r),rϰ(r)\frac{\lambda^{\prime}}{\lambda^{3}}\langle\psi_{2}(\lambda^{-1}r)-2ie^{2i\alpha(t)}h_{1}(r),r\varkappa^{\prime}(r)\rangle also contains some dynamical information, but in quadratic form, and will be shown to be negligible/perturbative;

  • the first term tψ2,ϰ(λr)\langle\partial_{t}\psi_{2},\varkappa(\lambda r)\rangle will provide the leading linear source term in the modulation equations. Understanding this require additional analysis, which we present below.

Using the covariant rules of calculus detailed in Section 4, we have

(9.5) tψ2=D0ψ2iA0ψ2=D2ψ0iA0ψ2=iA2ψ0iA0ψ2.\partial_{t}\psi_{2}=D_{0}\psi_{2}-iA_{0}\psi_{2}=D_{2}\psi_{0}-iA_{0}\psi_{2}=iA_{2}\psi_{0}-iA_{0}\psi_{2}.

Our goal is now to reexpress the right hand side in terms of ψ\psi and the quantities δλA2\delta^{\lambda}A_{2} and δλ,αψ2\delta^{\lambda,\alpha}\psi_{2}, which measure the deviation from the soliton manifold. For ψ\psi, we will seek to write its radial derivative in terms of LλψL_{\lambda}^{*}\psi; this is because the operator LλL_{\lambda}^{*} acts as a derivative in the H~λ\tilde{H}_{\lambda} calculus, and thus has a favourable behavior in the low frequency regime.

Using the computations in Section 4, we have

iψ0=D1ψ1+1rψ1+1r2D2ψ2=1(ψ+iψ2r)+1r(ψ+iψ2r)+i1r2A2ψ2=1ψ+1rψ+irψ2riψ2r2+iψ2r2+i1r2A2ψ2=rψ+1rψ+ir(iA2ψ)=rψ+1A2rψ=Lλψ+2h3λA2rψ.\begin{split}-i\psi_{0}&=D_{1}\psi_{1}+\frac{1}{r}\psi_{1}+\frac{1}{r^{2}}D_{2}\psi_{2}\\ &=\partial_{1}(\psi+i\frac{\psi_{2}}{r})+\frac{1}{r}(\psi+i\frac{\psi_{2}}{r})+i\frac{1}{r^{2}}A_{2}\psi_{2}\\ &=\partial_{1}\psi+\frac{1}{r}\psi+i\frac{\partial_{r}\psi_{2}}{r}-i\frac{\psi_{2}}{r^{2}}+i\frac{\psi_{2}}{r^{2}}+i\frac{1}{r^{2}}A_{2}\psi_{2}\\ &=\partial_{r}\psi+\frac{1}{r}\psi+\frac{i}{r}(iA_{2}\psi)\\ &=\partial_{r}\psi+\frac{1-A_{2}}{r}\psi\\ &=-L_{\lambda}^{*}\psi+\frac{2h_{3}^{\lambda}-A_{2}}{r}\psi.\end{split}

Therefore

iA2ψ0=2h3λLλψ+(A22hλ3)LλψA22h3λA2rψ=2h3λLλψ+δλA2Lλψ+A2δλA2rψ.iA_{2}\psi_{0}=2h_{3}^{\lambda}L^{*}_{\lambda}\psi+(A_{2}-2h^{\lambda}_{3})L_{\lambda}^{*}\psi-A_{2}\frac{2h_{3}^{\lambda}-A_{2}}{r}\psi=2h_{3}^{\lambda}L^{*}_{\lambda}\psi+\delta^{\lambda}A_{2}L_{\lambda}^{*}\psi+A_{2}\frac{\delta^{\lambda}A_{2}}{r}\psi.

We now consider the second term on the right in (9.5). Here we do not directly use the derivation in (4.22) for A0A_{0}, but instead we seek to highlight again the component LλψL_{\lambda}^{*}\psi in the expression that gives A0A_{0}. For this purpose we rearrange the formula for A0A_{0} as follows (this being a consequence of (4.4) and the gauge condition A1=0A_{1}=0):

rA0=(ψ0ψ¯1)=(iψ0ψ¯1)=((LλψδλA2rψ)(ψ¯iψ2¯r))=(iLλψψ2¯r)(Lλψψ¯)(δλA2rψψ¯1)=(2e2iα(t)h1λrLλψ)+(iLλψδλ,αψ2¯r)(Lλψψ¯)(δλA2rψψ¯1);\begin{split}-\partial_{r}A_{0}&=\Im(\psi_{0}\overline{\psi}_{1})=-\Re(i\psi_{0}\overline{\psi}_{1})=\Re\left((-L_{\lambda}^{*}\psi-\frac{\delta^{\lambda}A_{2}}{r}\psi)(\overline{\psi}-i\frac{\overline{\psi_{2}}}{r})\right)\\ &=\Re(iL_{\lambda}^{*}\psi\frac{\overline{\psi_{2}}}{r})-\Re(L_{\lambda}^{*}\psi\cdot\overline{\psi})-\Re(\frac{\delta^{\lambda}A_{2}}{r}\psi\overline{\psi}_{1})\\ &=\Re(\frac{2e^{-2i\alpha(t)}h_{1}^{\lambda}}{r}L_{\lambda}^{*}\psi)+\Re(iL_{\lambda}^{*}\psi\frac{\overline{\delta^{\lambda,\alpha}\psi_{2}}}{r})-\Re(L_{\lambda}^{*}\psi\cdot\overline{\psi})-\Re(\frac{\delta^{\lambda}A_{2}}{r}\psi\overline{\psi}_{1});\end{split}

the third equality is justified by the computation above for iψ0-i\psi_{0}.

From this we obtain the following representation for A0A_{0}:

(9.6) A0(r)=r((2e2iα(t)h1λrLλψ)+(iLλψδλ,αψ2¯r)(Lλψψ¯)(δλA2rψψ¯1))dr,A_{0}(r)=\int_{r}^{\infty}\left(\Re(\frac{2e^{-2i\alpha(t)}h_{1}^{\lambda}}{r}L_{\lambda}^{*}\psi)+\Re(iL_{\lambda}^{*}\psi\frac{\overline{\delta^{\lambda,\alpha}\psi_{2}}}{r})-\Re(L_{\lambda}^{*}\psi\cdot\overline{\psi})-\Re(\frac{\delta^{\lambda}A_{2}}{r}\psi\overline{\psi}_{1})\right)dr,

where the first term is linear in ψ\psi and the remaining terms are at least quadratic.

Bringing all the above computations together into the previous ODE for λ\lambda and α\alpha (derived at the beginning of this section) gives us a first form of the modulation equation,

(9.7) 4e2iα(t)(αλ2c1+iλλ3c2)=Lin+Nlin,4e^{2i\alpha(t)}\left(\frac{\alpha^{\prime}}{\lambda^{2}}c_{1}+i\frac{\lambda^{\prime}}{\lambda^{3}}c_{2}\right)=Lin+Nlin,

where LinLin collects all the linear terms:

Lin=2e2iαh1λr(2e2iαh1λrLλψ)dr,ϰλ2h3λLλψ,ϰλ,Lin=-\langle 2e^{2i\alpha}h_{1}^{\lambda}\int_{r}^{\infty}\Re(\frac{2e^{-2i\alpha}h_{1}^{\lambda}}{r}L_{\lambda}^{*}\psi)dr,\varkappa^{\lambda}\rangle-\langle 2h_{3}^{\lambda}L_{\lambda}^{*}\psi,\varkappa^{\lambda}\rangle,

and NlinNlin collects all nonlinear (quadratic or higher order) terms

Nlin=λλ3ψ2(λ1r)2ie2iα(t)h1(r),rϰ(r)δλA2Lλψ+A2δλA2rψ,ϰλ2e2iα(t)r(iLλψδλ,αψ2¯r)(Lλψψ¯)(δλA2rψψ¯1)dr,(h1ϰ)λ+iA0δλ,αψ2,ϰλ.\begin{split}Nlin&=-\frac{\lambda^{\prime}}{\lambda^{3}}\langle\psi_{2}(\lambda^{-1}r)-2ie^{2i\alpha(t)}h_{1}(r),r\varkappa^{\prime}(r)\rangle-\langle\delta^{\lambda}A_{2}\cdot L^{*}_{\lambda}\psi+A_{2}\frac{\delta^{\lambda}A_{2}}{r}\psi,\varkappa^{\lambda}\rangle\\ &-2e^{2i\alpha(t)}\langle\int_{r}^{\infty}\Re(iL^{*}_{\lambda}\psi\frac{\overline{\delta^{\lambda,\alpha}\psi_{2}}}{r})-\Re(L^{*}_{\lambda}\psi\overline{\psi})-\Re(\frac{\delta^{\lambda}A_{2}}{r}\psi\overline{\psi}_{1})dr,(h_{1}\varkappa)^{\lambda}\rangle\\ &+\langle iA_{0}\delta^{\lambda,\alpha}\psi_{2},\varkappa^{\lambda}\rangle.\end{split}

We have obtained the ODE system (9.7) that governs the evolution of the two parameters α\alpha and λ\lambda; this is a first version from which we derive an alternative version which is more amenable to estimates. We will prove that the contributions in NlinNlin to the dynamics of λ\lambda and α\alpha are perturbative. However the first term in NlinNlin is special in that it contains the expression λλ3\frac{\lambda^{\prime}}{\lambda^{3}}; this is why this term is treated slightly differently, and justifies the following separation:

Nlin=λλ3ψ2(λ1r)2ie2iα(t)h1(r),rϰ(r)+Nlin~:=λλ3e+Nlin~.Nlin=-\frac{\lambda^{\prime}}{\lambda^{3}}\langle\psi_{2}(\lambda^{-1}r)-2ie^{2i\alpha(t)}h_{1}(r),r\varkappa^{\prime}(r)\rangle+\widetilde{Nlin}:=-\frac{\lambda^{\prime}}{\lambda^{3}}e+\widetilde{Nlin}.

It will soon become clear that in (9.7), the variable λ\lambda is not scaled properly. For this reason we rewrite it as follows

(9.8) 4e2iα(t)(αc1+iλλc2)=l(t)+q(t).4e^{2i\alpha(t)}\left(\alpha^{\prime}c_{1}+i\frac{\lambda^{\prime}}{\lambda}c_{2}\right)=l(t)+q(t).

where

l(t)=λ2Lin,q(t)=λ2Nlin.l(t)=\lambda^{2}Lin,\quad q(t)=\lambda^{2}Nlin.

This is the main formulation of the ODE describing the parameter dynamics. Similar to NlinNlin, we split the qq terms as follows

q=λλe+q~,q=-\frac{\lambda^{\prime}}{\lambda}e+\tilde{q},

where q~(t)=λ2Nlin~\tilde{q}(t)=\lambda^{2}\widetilde{Nlin}.

We note that the form (9.8) of the modulation equations will be used in Section 10 in order to provide the most accurate dynamical information on α\alpha and lnλ\ln\lambda. However for the main result of this section, namely Theorem 9.1, we will use a slight modification of it.

9.2. Estimates for the modulation equations

In this subsection we provide the main bounds for the source terms in the modulation equation, which will be used in the proof of Theorem 9.1. For this purpose, we rewrite (9.8) as follows:

(9.9) 4e2iα(t)(c1αλ+ic2λλ2+e2iα(t)λλ2e(t))=λ1l(t)+λ1q~(t).4e^{2i\alpha(t)}\left(c_{1}\frac{\alpha^{\prime}}{\lambda}+ic_{2}\frac{\lambda^{\prime}}{\lambda^{2}}+e^{-2i\alpha(t)}\frac{\lambda^{\prime}}{\lambda^{2}}e(t)\right)=\lambda^{-1}l(t)+\lambda^{-1}\tilde{q}(t).

We bound the entries e,l,q~e,l,\tilde{q} in this equation as follows:

Lemma 9.3.

The functions e,l,q~e,l,\tilde{q} in (9.9) satisfy the following fixed time estimates:

(9.10) |e(t)|ψ(t)L2,λ|e(t)|ψ(t)rL2r,λ1|l(t)|ψ(t)rL2r,|q~(t)|ψ(t)rL2r2,λ1|q~(t)|ψ(t)L2rψ(t)rL2r.\begin{split}&|e(t)|\lesssim\|\psi(t)\|_{L^{2}},\quad\lambda|e(t)|\lesssim\|\frac{\psi(t)}{r}\|_{L^{2}_{r}},\quad\lambda^{-1}|l(t)|\lesssim\|\frac{\psi(t)}{r}\|_{L^{2}_{r}},\\ &|\tilde{q}(t)|\lesssim\|\frac{\psi(t)}{r}\|_{L^{2}_{r}}^{2},\quad\lambda^{-1}|\tilde{q}(t)|\lesssim\|\psi(t)\|_{L^{2}_{r}}\|\frac{\psi(t)}{r}\|_{L^{2}_{r}}.\end{split}
Proof of Lemma 9.3.

We successively consider the bounds for ee, ll and q~\tilde{q}.

a) The bounds for ee. From (4.39) it follows that ψ22ie2iα(t)h1λLψL2\|\psi_{2}-2ie^{2i\alpha(t)}h_{1}^{\lambda}\|_{L^{\infty}}\lesssim\|\psi\|_{L^{2}}, which directly implies that |e(t)|ψ(t)L21|e(t)|\lesssim\|\psi(t)\|_{L^{2}}\ll 1. On the other hand we can bound ee using the local energy norm of ψ\psi,

λe(t)=λψ2(λ1r)2ie2iα(t)h1(r),rϰ(r)=λ4δλ,αψ2,rϰ(λr)=λ3δλ,αψ2,(rϰ)λ=λδλ,αψ2r2,(r3ϰ)λ.\begin{split}\lambda e(t)&=\lambda\langle\psi_{2}(\lambda^{-1}r)-2ie^{2i\alpha(t)}h_{1}(r),r\varkappa^{\prime}(r)\rangle\\ &=\lambda^{4}\langle\delta^{\lambda,\alpha}\psi_{2},r\varkappa^{\prime}(\lambda r)\rangle=\lambda^{3}\langle\delta^{\lambda,\alpha}\psi_{2},(r\varkappa^{\prime})^{\lambda}\rangle\\ &=\lambda\langle\frac{\delta^{\lambda,\alpha}\psi_{2}}{r^{2}},(r^{3}\varkappa^{\prime})^{\lambda}\rangle.\end{split}

Then using (4.37) we obtain

λ|e(t)|δλ,αψ2r2L2λ(r3ϰ)λL2ψrL2t,r,\lambda|e(t)|\lesssim\|\frac{\delta^{\lambda,\alpha}\psi_{2}}{r^{2}}\|_{L^{2}}\|\lambda(r^{3}\varkappa^{\prime})^{\lambda}\|_{L^{2}}\lesssim\|\frac{\psi}{r}\|_{L^{2}_{t,r}},

as desired.

b) The bounds for ll. We write l(t)=l1(t)+l2(t)l(t)=l^{1}(t)+l^{2}(t) where

(9.11) l1=2λ2h3λLλψ,ϰλ,l2=4λ2e2iα(t)h1λ(t)r(e2iα(t)h1λ(t)sLλψ)ds,ϰλ.l^{1}=-2\lambda^{2}\langle h_{3}^{\lambda}L_{\lambda}^{*}\psi,\varkappa^{\lambda}\rangle,\quad l^{2}=-4\lambda^{2}\langle e^{2i\alpha(t)}h_{1}^{\lambda(t)}\int_{r}^{\infty}\Re(\frac{e^{-2i\alpha(t)}h_{1}^{\lambda(t)}}{s}L_{\lambda}^{*}\psi)ds,\varkappa^{\lambda}\rangle.

A direct computation gives

12λ1l1(t)=λLλψ,h3λϰλ=λψ,Lλ(h3λϰλ)=λ2ψ,(L(h3ϰ))λ=ψ(λ1),L(h3ϰ).-\frac{1}{2}\lambda^{-1}l^{1}(t)=\lambda\langle L_{\lambda}^{*}\psi,h_{3}^{\lambda}\varkappa^{\lambda}\rangle=\lambda\langle\psi,L_{\lambda}(h_{3}^{\lambda}\varkappa^{\lambda})\rangle=\lambda^{2}\langle\psi,(L(h_{3}\varkappa))^{\lambda}\rangle=\langle\psi(\lambda^{-1}\cdot),L(h_{3}\varkappa)\rangle.

This allows us to estimate

|λ1l1(t)|ψ(λ1)L2(r1)λ1ψ(λ1)λ1L2(r1)ψrL2(rλ1)ψrL2.|\lambda^{-1}l^{1}(t)|\lesssim\|\psi(\lambda^{-1}\cdot)\|_{L^{2}(r\approx 1)}\approx\lambda^{-1}\|\frac{\psi(\lambda^{-1}\cdot)}{\lambda^{-1}\cdot}\|_{L^{2}(r\approx 1)}\approx\|\frac{\psi}{r}\|_{L^{2}(r\approx\lambda^{-1})}\lesssim\|\frac{\psi}{r}\|_{L^{2}}.

Similarly, we compute:

e2iα4λ1l2=λr(e2iαh1λsLλψ)ds,h1λϰλ=(e2iα(t)λrh1λ(t)sLλψds,h1λϰλ).\begin{split}-\frac{e^{-2i\alpha}}{4}\lambda^{-1}l^{2}=\lambda\langle\int_{r}^{\infty}\Re(\frac{e^{-2i\alpha}h_{1}^{\lambda}}{s}L_{\lambda}^{*}\psi)ds,h_{1}^{\lambda}\varkappa^{\lambda}\rangle=\Re\left(e^{-2i\alpha(t)}\lambda\langle\int_{r}^{\infty}\frac{h_{1}^{\lambda(t)}}{s}L_{\lambda}^{*}\psi ds,h_{1}^{\lambda}\varkappa^{\lambda}\rangle\right).\end{split}

Defining

(9.12) 𝔤1(r)=0rh1(s)ϰ(s)sds,𝔤2=𝔤1h1r2,\mathfrak{g}_{1}(r)=\int_{0}^{r}h_{1}(s)\varkappa(s)sds,\qquad\mathfrak{g}_{2}=\frac{\mathfrak{g}_{1}h_{1}}{r^{2}},

this is further written as

e2iα4λ1l2=(e2iα(t)λLλψ,𝔤2λ),-\frac{e^{-2i\alpha}}{4}\lambda^{-1}l_{2}=\Re(e^{-2i\alpha(t)}\lambda\langle L_{\lambda}^{*}\psi,\mathfrak{g}_{2}^{\lambda}\rangle),

thus resembling the previous expression modulo the phase plus a change in the test function 𝔤2\mathfrak{g}_{2}. Since 𝔤1(r)=0\mathfrak{g}_{1}(r)=0 for r12r\leq\frac{1}{2} and 𝔤1(r)=c1\mathfrak{g}_{1}(r)=c_{1} for r2r\geq 2, it follows that 𝔤2(r)=0\mathfrak{g}_{2}(r)=0 for r12r\leq\frac{1}{2} and 𝔤2(r)=c1h1r2\mathfrak{g}_{2}(r)=\frac{c_{1}h_{1}}{r^{2}} for r2r\geq 2; therefore the 𝔤2\mathfrak{g}_{2} factor is similar to the h3ϰh_{3}\varkappa factor appearing in the expression for l1l^{1}), the only difference being that it decays like r4r^{-4} for large rr, as opposed to being 0. Thus we can estimate

|λLλψ,𝔤2λ|=|λ2ψ,(L𝔤2)λ|=|ψ(λ1),L𝔤2|λ1ψ(λ1r)λ1rL2rL𝔤2L2ψrL2.|\lambda\langle L_{\lambda}^{*}\psi,\mathfrak{g}_{2}^{\lambda}\rangle|=|\lambda^{2}\langle\psi,(L\mathfrak{g}_{2})^{\lambda}\rangle|=|\langle\psi(\lambda^{-1}\cdot),L\mathfrak{g}_{2}\rangle|\lesssim\lambda^{-1}\|\frac{\psi(\lambda^{-1}r)}{\lambda^{-1}r}\|_{L^{2}}\|rL\mathfrak{g}_{2}\|_{L^{2}}\lesssim\|\frac{\psi}{r}\|_{L^{2}}.

Based on the two estimates above on λ1l1\lambda^{-1}l^{1} and λ1l2\lambda^{-1}l^{2}, we obtain the estimate on λ1l\lambda^{-1}l as claimed in (9.10).

c) The bounds for q~\tilde{q}. There are many terms to be estimated here and we deal with them in the order in which they appear in the above expression of NlinNlin; recall that q~\tilde{q} skips the first term in the expression of NlinNlin since it collects only the terms which appear in Nlin~\widetilde{Nlin}.

c1) The expression δλA2Lλψ+A2δλA2rψ,ϰλ\langle\delta^{\lambda}A_{2}\cdot L^{*}_{\lambda}\psi+A_{2}\frac{\delta^{\lambda}A_{2}}{r}\psi,\varkappa^{\lambda}\rangle. Using (4.42), the first term in this expression is

δλA2Lλψ,χλ=Lλψ,δλA2ϰλ=ψ,Lλ(δλA2ϰλ)=ψ,LλδλA2ϰλ+ψ,δλA2rϰλ=ψ,(2h1λ(e2iαψ)+(ψδλ,αψ¯2)(δλA2)2r)ϰλ+ψ,δλA21r(rϰ)λ.\begin{split}\langle\delta^{\lambda}A_{2}\cdot L_{\lambda}^{*}\psi,\chi^{\lambda}\rangle&=\langle L_{\lambda}^{*}\psi,\delta^{\lambda}A_{2}\cdot\varkappa^{\lambda}\rangle=\langle\psi,L_{\lambda}(\delta^{\lambda}A_{2}\cdot\varkappa^{\lambda})\rangle\\ &=\langle\psi,L_{\lambda}\delta^{\lambda}A_{2}\cdot\varkappa^{\lambda}\rangle+\langle\psi,\delta^{\lambda}A_{2}\cdot\partial_{r}\varkappa^{\lambda}\rangle\\ &=\langle\psi,\left(-2h_{1}^{\lambda}\Re{(e^{2i\alpha}\psi)}+\Im(\psi\overline{\delta^{\lambda,\alpha}\psi}_{2})-\frac{(\delta^{\lambda}A_{2})^{2}}{r}\right)\varkappa^{\lambda}\rangle+\langle\psi,\delta^{\lambda}A_{2}\cdot\frac{1}{r}(r\varkappa^{\prime})^{\lambda}\rangle.\\ \end{split}

Using (4.37) and (4.39) (of which we use the consequence δλ,αψ2L+δλA2LψL21\|\delta^{\lambda,\alpha}\psi_{2}\|_{L^{\infty}}+\|\delta^{\lambda}A_{2}\|_{L^{\infty}}\lesssim\|\psi\|_{L^{2}}\ll 1), this leads us to the following estimates:

|δλA2Lλψ,ϰλ||ψ|2r2,(1+|δλ,αψ2|)r2ϰλ+|ψ|r|δλA2|r2|δλA2|,r2ϰλ+|ψ|r|δλA2|r2,r2|(rϰ)λ|λ2(1+ψL2)ψr2L2,\begin{split}|\langle\delta^{\lambda}A_{2}\cdot L_{\lambda}^{*}\psi,\varkappa^{\lambda}\rangle|&\lesssim\langle\frac{|\psi|^{2}}{r^{2}},(1+|\delta^{\lambda,\alpha}\psi_{2}|)r^{2}\varkappa^{\lambda}\rangle+\langle\frac{|\psi|}{r}\cdot\frac{|\delta^{\lambda}A_{2}|}{r^{2}}\cdot|\delta^{\lambda}A_{2}|,r^{2}\varkappa^{\lambda}\rangle\\ &+\langle\frac{|\psi|}{r}\cdot\frac{|\delta^{\lambda}A_{2}|}{r^{2}},r^{2}|(r\varkappa^{\prime})^{\lambda}|\rangle\lesssim\lambda^{-2}(1+\|\psi\|_{L^{2}})\|\frac{\psi}{r}\|^{2}_{L^{2}},\end{split}

and

|δλA2Lλψ,ϰλ|λ1(1+ψL2)ψL2ψrL2.|\langle\delta^{\lambda}A_{2}\cdot L_{\lambda}^{*}\psi,\varkappa^{\lambda}\rangle|\lesssim\lambda^{-1}(1+\|\psi\|_{L^{2}})\|\psi\|_{L^{2}}\cdot\|\frac{\psi}{r}\|_{L^{2}}.

These are the estimates we were seeking for the first term; it is clear that the second term A2δλA2rψ,ϰλ\langle A_{2}\frac{\delta^{\lambda}A_{2}}{r}\psi,\varkappa^{\lambda}\rangle is estimated in a similar manner (recall that |A2|2|A_{2}|\leq 2).

c2) The expression r(iLλψδλ,αψ2¯r)dr,(h1χ)λ\langle\int_{r}^{\infty}\Re(iL^{*}_{\lambda}\psi\frac{\overline{\delta^{\lambda,\alpha}\psi_{2}}}{r})dr,(h_{1}\chi)^{\lambda}\rangle. Note that here we can harmlessly ignore the additional phase factor 2e2iα2e^{2i\alpha}. We compute:

rLλψ¯δλ,αψ2s2sds=ψ(r)¯δλ,αψ2r+rψ¯Lλ(δλ,αψ2r2)rdr=ψ(r)¯δλ,αψ2r+rψ¯Lλδλ,αψ2s2sds2rψ¯δλ,αψ2s3sds=ψ(r)¯δλ,αψ2r2rψ¯δλ,αψ2s3sds+rψ¯2ih3λψ+δλA2ψ1sδλA2(2ie2iαh1λ+δλ,αψ2)s2sds,\begin{split}\int_{r}^{\infty}L^{*}_{\lambda}\overline{\psi}\frac{\delta^{\lambda,\alpha}\psi_{2}}{s^{2}}sds&=-\overline{\psi(r)}\frac{\delta^{\lambda,\alpha}\psi_{2}}{r}+\int_{r}^{\infty}\overline{\psi}\cdot L_{\lambda}(\frac{\delta^{\lambda,\alpha}\psi_{2}}{r^{2}})rdr\\ &=-\overline{\psi(r)}\frac{\delta^{\lambda,\alpha}\psi_{2}}{r}+\int_{r}^{\infty}\overline{\psi}\cdot\frac{L_{\lambda}\delta^{\lambda,\alpha}\psi_{2}}{s^{2}}sds-2\int_{r}^{\infty}\overline{\psi}\cdot\frac{\delta^{\lambda,\alpha}\psi_{2}}{s^{3}}sds\\ &=-\overline{\psi(r)}\frac{\delta^{\lambda,\alpha}\psi_{2}}{r}-2\int_{r}^{\infty}\overline{\psi}\cdot\frac{\delta^{\lambda,\alpha}\psi_{2}}{s^{3}}sds\\ &+\int_{r}^{\infty}\overline{\psi}\cdot\frac{2ih_{3}^{\lambda}\psi+\delta^{\lambda}A_{2}\psi-\frac{1}{s}\delta^{\lambda}A_{2}(2ie^{2i\alpha}h_{1}^{\lambda}+\delta^{\lambda,\alpha}\psi_{2})}{s^{2}}sds,\\ \end{split}

where in passing to the last line we have used the system (4.42).

We estimate the term ψ(r)δλ,αψ2r,ϰλ\langle\psi(r)\frac{\delta^{\lambda,\alpha}\psi_{2}}{r},\varkappa^{\lambda}\rangle just as we estimated the ψ,δλA21r(rϰ)λ\langle\psi,\delta^{\lambda}A_{2}\cdot\frac{1}{r}(r\varkappa^{\prime})^{\lambda}\rangle term above. A direct estimate gives us:

1r2rψsδλ,αψ2s2sdsL1(rλ1)ψrL2δλ,αψ2r2L2,\|\frac{1}{r^{2}}\int_{r}^{\infty}\frac{\psi}{s}\cdot\frac{\delta^{\lambda,\alpha}\psi_{2}}{s^{2}}sds\|_{L^{1}(r\approx\lambda^{-1})}\lesssim\|\frac{\psi}{r}\|_{L^{2}}\cdot\|\frac{\delta^{\lambda,\alpha}\psi_{2}}{r^{2}}\|_{L^{2}},

and then using (4.37) we obtain:

|rψδλ,αψ2s3sds,ϰλ|λ2ψrL22.|\langle\int_{r}^{\infty}\psi\cdot\frac{\delta^{\lambda,\alpha}\psi_{2}}{s^{3}}sds,\varkappa^{\lambda}\rangle|\lesssim\lambda^{-2}\|\frac{\psi}{r}\|_{L^{2}}^{2}.

Alternatively, we use (2.9) (with p=1p=1 and w=r1w=r^{-1}) and estimate

1rrψsδλ,αψ2s1ssdsL1ψrL2δλ,αψ2rL2,\|\frac{1}{r}\int_{r}^{\infty}\frac{\psi}{s}\cdot\frac{\delta^{\lambda,\alpha}\psi_{2}}{s}\frac{1}{s}sds\|_{L^{1}}\lesssim\|\frac{\psi}{r}\|_{L^{2}}\cdot\|\frac{\delta^{\lambda,\alpha}\psi_{2}}{r}\|_{L^{2}},

from which, by using (4.39), we obtain

|rψδλ,αψ2s3sds,ϰλ|λ1ψL2ψrL2.|\langle\int_{r}^{\infty}\psi\cdot\frac{\delta^{\lambda,\alpha}\psi_{2}}{s^{3}}sds,\varkappa^{\lambda}\rangle|\lesssim\lambda^{-1}\|\psi\|_{L^{2}}\|\frac{\psi}{r}\|_{L^{2}}.

These are the expected estimates for the second term; finally the third term

rψ2ih3λψ+δλA2ψ1sδλA2(2ie2iαh1+δλ,αψ2)s2sds,ϰλ\langle\int_{r}^{\infty}\psi\cdot\frac{2ih_{3}^{\lambda}\psi+\delta^{\lambda}A_{2}\psi-\frac{1}{s}\delta^{\lambda}A_{2}(2ie^{2i\alpha}h_{1}+\delta^{\lambda,\alpha}\psi_{2})}{s^{2}}sds,\varkappa^{\lambda}\rangle

is estimated in a similar fashion; the details are left as an exercise for the interested reader.

c3) The expression r(Lλψψ¯),(h1ϰ)λ\langle\int_{r}^{\infty}\Re(L^{*}_{\lambda}\psi\overline{\psi}),(h_{1}\varkappa)^{\lambda}\rangle. Here we write:

r(Lλψψ¯)ds,(h1ϰ)λ=r((s+2h3λ1s)ψψ¯)ds,(h1ϰ)λ=r12s|ψ|2+2h3λ1s|ψ|2ds,(h1ϰ)λ=12|ψ|2,(h1χ)λ+r2h3λ1s|ψ|2ds,(h1ϰ)λ.\begin{split}\langle\int_{r}^{\infty}\Re(L^{*}_{\lambda}\psi\overline{\psi})ds,(h_{1}\varkappa)^{\lambda}\rangle&=\langle\int_{r}^{\infty}\Re((-\partial_{s}+\frac{2h_{3}^{\lambda}-1}{s})\psi\cdot\overline{\psi})ds,(h_{1}\varkappa)^{\lambda}\rangle\\ &=\langle\int_{r}^{\infty}-\frac{1}{2}\partial_{s}|\psi|^{2}+\frac{2h_{3}^{\lambda}-1}{s}|\psi|^{2}ds,(h_{1}\varkappa)^{\lambda}\rangle\\ &=\frac{1}{2}\langle|\psi|^{2},(h_{1}\chi)^{\lambda}\rangle+\langle\int_{r}^{\infty}\frac{2h_{3}^{\lambda}-1}{s}|\psi|^{2}ds,(h_{1}\varkappa)^{\lambda}\rangle.\end{split}

The first term above is estimated as follows:

|ψ|2,(h1ϰ)λrλ1|ψ|2rdrλ2rλ1|ψ|2r2rdr,\langle|\psi|^{2},(h_{1}\varkappa)^{\lambda}\rangle\approx\int_{r\approx\lambda^{-1}}|\psi|^{2}rdr\approx\lambda^{-2}\int_{r\approx\lambda^{-1}}\frac{|\psi|^{2}}{r^{2}}rdr,

and this gives the correct contribution for the first q~\tilde{q} bound. Alternatively, the same term is estimated as follows:

|ψ|2,(h1ϰ)λλ1rλ1|ψ||ψ|rrdrλ1ψ(t)L2rψ(t)rL2r2,\langle|\psi|^{2},(h_{1}\varkappa)^{\lambda}\rangle\approx\lambda^{-1}\int_{r\approx\lambda^{-1}}|\psi|\frac{|\psi|}{r}rdr\lesssim\lambda^{-1}\|\psi(t)\|_{L^{2}_{r}}\|\frac{\psi(t)}{r}\|_{L^{2}_{r}}^{2},

and this gives the correct contribution to the second q~\tilde{q} bound.

The second term above is estimated as follows:

|r2h3λ1s|ψ|2ds,(h1ϰ)λ|=|r(2h3λ1)|ψ|2s2sds,(h1ϰ)λ||ψ|2r2L1(h1ϰ)(λr)rdrλ2ψr2L2r,\begin{split}|\langle\int_{r}^{\infty}\frac{2h_{3}^{\lambda}-1}{s}|\psi|^{2}ds,(h_{1}\varkappa)^{\lambda}\rangle|&=|\langle\int_{r}^{\infty}(2h_{3}^{\lambda}-1)\frac{|\psi|^{2}}{s^{2}}sds,(h_{1}\varkappa)^{\lambda}\rangle|\\ &\lesssim\|\frac{|\psi|^{2}}{r^{2}}\|_{L^{1}}\int(h_{1}\varkappa)(\lambda r)rdr\\ &\lesssim\lambda^{-2}\|\frac{\psi}{r}\|^{2}_{L^{2}_{r}},\end{split}

and this gives the correct contribution to the first q~\tilde{q} bound. We can also estimate this term as follows

|r|ψ|s|ψ|ds,(h1ϰ)λ||ψ|r|ψ|L1r(h1ϰ)(λr)rrdrλ1ψ(t)rL2rψ(t)L2,\begin{split}&\lesssim|\langle\int_{r}^{\infty}\frac{|\psi|}{s}|\psi|ds,(h_{1}\varkappa)^{\lambda}\rangle|\lesssim\|\frac{|\psi|}{r}\cdot|\psi|\|_{L^{1}_{r}}\int\frac{(h_{1}\varkappa)(\lambda r)}{r}rdr\lesssim\lambda^{-1}\|\frac{\psi(t)}{r}\|_{L^{2}_{r}}\|\psi(t)\|_{L^{2}},\end{split}

to obtain the correct contribution to the second q~\tilde{q} bound.

c4) The expression r(δλA2rψψ¯1)ds,ϰλ\langle\int_{r}^{\infty}\Re(\frac{\delta^{\lambda}A_{2}}{r}\psi\overline{\psi}_{1})ds,\varkappa^{\lambda}\rangle. We write

δλA2rψψ¯1=δλA2rψ(ψ+iψ2r¯)=δλA2rψ(ψ¯iψ2¯r),\frac{\delta^{\lambda}A_{2}}{r}\psi\overline{\psi}_{1}=\frac{\delta^{\lambda}A_{2}}{r}\cdot\psi\cdot(\overline{\psi+i\frac{\psi_{2}}{r}})=\frac{\delta^{\lambda}A_{2}}{r}\cdot\psi\cdot(\overline{\psi}-i\frac{\overline{\psi_{2}}}{r}),

and note that the previous estimates provide a template on how to obtain similar estimates for this term as well. The details are left as an exercise.

c5) The expression iA0δλ,αψ2,ϰλ\langle iA_{0}\delta^{\lambda,\alpha}\psi_{2},\varkappa^{\lambda}\rangle. Based on the expression (9.6) for A0A_{0}, this term has two components:

i) the contribution of the nonlinear component of A0A_{0}, which combined with the trivial inequality δλ,αψ2LψL2\|\delta^{\lambda,\alpha}\psi_{2}\|_{L^{\infty}}\lesssim\|\psi\|_{L^{2}} is estimated exactly as in the previous steps (c2)-(c4);

ii) the combination of the linear component of A0A_{0} with the δλ,αψ2\delta^{\lambda,\alpha}\psi_{2} factor, which in this context is also viewed as a nonlinear contribution:

r(2e2iαh1λsLλψ)dsδλ,αψ2,ϰλ=(e2iαr2h1λsLλψds)δλ,αψ2,ϰλ.\langle\int_{r}^{\infty}\Re(\frac{2e^{-2i\alpha}h_{1}^{\lambda}}{s}L_{\lambda}^{*}\psi)ds\cdot\delta^{\lambda,\alpha}\psi_{2},\varkappa^{\lambda}\rangle=\langle\Re(e^{-2i\alpha}\int_{r}^{\infty}\frac{2h_{1}^{\lambda}}{s}L_{\lambda}^{*}\psi ds)\cdot\delta^{\lambda,\alpha}\psi_{2},\varkappa^{\lambda}\rangle.

We first integrate by parts

rh1λsLλψds=h1λrψ+rψLλ(h1λs2)sds=h1λrψ2rψh1λs3sds.\int_{r}^{\infty}\frac{h_{1}^{\lambda}}{s}L_{\lambda}^{*}\psi ds=-\frac{h_{1}^{\lambda}}{r}\psi+\int_{r}^{\infty}\psi L_{\lambda}(\frac{h_{1}^{\lambda}}{s^{2}})sds=-\frac{h_{1}^{\lambda}}{r}\psi-2\int_{r}^{\infty}\psi\frac{h_{1}^{\lambda}}{s^{3}}sds.

which is a-priori justified for regular ψ\psi, and then we estimate

|h1λrψδλ,αψ2,ϰλ|=|h1λrψδλ,αψ2r2,r2ϰλ|λ2ψrL2δλ,αψ2r2L2λ2ψrL22,|\langle\frac{h_{1}^{\lambda}}{r}\psi\cdot\delta^{\lambda,\alpha}\psi_{2},\varkappa^{\lambda}\rangle|=|\langle\frac{h_{1}^{\lambda}}{r}\psi\cdot\frac{\delta^{\lambda,\alpha}\psi_{2}}{r^{2}},r^{2}\varkappa^{\lambda}\rangle|\lesssim\lambda^{-2}\|\frac{\psi}{r}\|_{L^{2}}\|\frac{\delta^{\lambda,\alpha}\psi_{2}}{r^{2}}\|_{L^{2}}\lesssim\lambda^{-2}\|\frac{\psi}{r}\|_{L^{2}}^{2},

as well as

|h1λrψδλ,αψ2,ϰλ|=|h1λrψδλ,αψ2r,rϰλ|λ1ψrL2δλ,αψ2rL2λ1ψrL2ψL2.|\langle\frac{h_{1}^{\lambda}}{r}\psi\cdot\delta^{\lambda,\alpha}\psi_{2},\varkappa^{\lambda}\rangle|=|\langle\frac{h_{1}^{\lambda}}{r}\psi\cdot\frac{\delta^{\lambda,\alpha}\psi_{2}}{r},r\varkappa^{\lambda}\rangle|\lesssim\lambda^{-1}\|\frac{\psi}{r}\|_{L^{2}}\|\frac{\delta^{\lambda,\alpha}\psi_{2}}{r}\|_{L^{2}}\lesssim\lambda^{-1}\|\frac{\psi}{r}\|_{L^{2}}\|\psi\|_{L^{2}}.

The second component above is estimated similarly in combination with the straightforward inequality:

rψh1λs3sdsL2ψrL2.\|\int_{r}^{\infty}\psi\frac{h_{1}^{\lambda}}{s^{3}}sds\|_{L^{2}}\lesssim\|\frac{\psi}{r}\|_{L^{2}}.

This finishes the proof of all our claims in (9.10).

9.3. Proof of Theorem 9.1

The idea here is to carefully combine the bounds in Lemma 9.3 for the terms in the modulation equation with the local energy bounds for ψ\psi.

Without restricting the generality of the argument, it suffices to establish the result on the forward in time maximal time of existence Imax+=[0,Tmax)I_{max}^{+}=[0,T_{max}).

a) Modulation equation bounds. While the bounds in Lemma 9.3 are fixed time bounds, here we will primarily use them in an integrated fashion. Assume that we have space-time control of the local energy decay, i.e. that is we control the quantity ψ(t)rL2tL2r(I×)\|\frac{\psi(t)}{r}\|_{L^{2}_{t}L^{2}_{r}(I\times\mathbb{R})} for some time interval I=[0,T]I=[0,T] where 0<T<Tmax0<T<T_{max}. All the estimates involving space-time norms below are restricted to the time interval II.

The first bound on ee in (9.10) combined with the smallness of ψL2\|\psi\|_{L^{2}} shows that ee is uniformly small,

eLψLL21.\|e\|_{L^{\infty}}\lesssim\|\psi\|_{L^{\infty}L^{2}}\ll 1.

This shows that we can interpret ee perturbatively in (9.9) and estimate using (9.10)

(9.13) αλL2t(I)+λλ2L2t(I)(1+ψ0L2r)ψ(t)rL2tL2r(I).\|\frac{\alpha^{\prime}}{\lambda}\|_{L^{2}_{t}(I)}+\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}_{t}(I)}\lesssim(1+\|\psi_{0}\|_{L^{2}_{r}})\|\frac{\psi(t)}{r}\|_{L^{2}_{t}L^{2}_{r}(I)}.

Incidentally, we remark, for later use, that this allows us to capture the perturbative nature of the term λλe(t)\frac{\lambda^{\prime}}{\lambda}e(t) in (9.9), and thus justify why it can be included in the perturbative term qq. Indeed, the above with the second ee bound in (9.10) yields

(9.14) λλe(t)L1tλλ2L2tλe(t)L2t(1+ψ0L2r)ψr2L2,\|\frac{\lambda^{\prime}}{\lambda}e(t)\|_{L^{1}_{t}}\lesssim\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}_{t}}\|\lambda e(t)\|_{L^{2}_{t}}\lesssim(1+\|\psi_{0}\|_{L^{2}_{r}})\|\frac{\psi}{r}\|^{2}_{L^{2}},

while using the first bound for ee in (9.10) gives

(9.15) λλ2e(t)L2tλλ2L2te(t)Lt(1+ψ0L2r)ψLtL2rψrL2.\|\frac{\lambda^{\prime}}{\lambda^{2}}e(t)\|_{L^{2}_{t}}\lesssim\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}_{t}}\|e(t)\|_{L^{\infty}_{t}}\lesssim(1+\|\psi_{0}\|_{L^{2}_{r}})\|\psi\|_{L^{\infty}_{t}L^{2}_{r}}\|\frac{\psi}{r}\|_{L^{2}}.

This allows us to add the term λλe(t)\frac{\lambda^{\prime}}{\lambda}e(t) to the quadratic component q(t)q(t) and extend the estimates on q~\tilde{q} in (9.10) to qq, namely

(9.16) qL1tψ(t)rL2tL2r2,λ1qL2tψLtL2rψrL2t,r.\|q\|_{L^{1}_{t}}\lesssim\|\frac{\psi(t)}{r}\|_{L^{2}_{t}L^{2}_{r}}^{2},\quad\|\lambda^{-1}q\|_{L^{2}_{t}}\lesssim\|\psi\|_{L^{\infty}_{t}L^{2}_{r}}\|\frac{\psi}{r}\|_{L^{2}_{t,r}}.

This justifies the use of (9.8) as the main formulation of the modulation equations, which will be used in all the analysis below.

Applied in the modulation equation (9.8), the second bound above, combined with the second bound for ll in (9.10), yields

(9.17) λλ2L2t+αλL2tψrL2t,r.\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}_{t}}+\|\frac{\alpha^{\prime}}{\lambda}\|_{L^{2}_{t}}\lesssim\|\frac{\psi}{r}\|_{L^{2}_{t,r}}.

b) Bounds for ψ\psi. Here we start with the local energy bound in Corollary 7.2, which yields

ψrL2t,r+ψL4t,r(1+λλ2L2t)(ψ0L2+WλψL43).\|\frac{\psi}{r}\|_{L^{2}_{t,r}}+\|\psi\|_{L^{4}_{t,r}}\lesssim(1+\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}_{t}})(\|\psi_{0}\|_{L^{2}}+\|W_{\lambda}\psi\|_{L^{\frac{4}{3}}}).

On the other hand from the bound (8.2) for the nonlinearity we have

WλψL43ψL4(ψrL2+ψL42).\|W_{\lambda}\psi\|_{L^{\frac{4}{3}}}\lesssim\|\psi\|_{L^{4}}(\|\frac{\psi}{r}\|_{L^{2}}+\|\psi\|_{L^{4}}^{2}).

Combining the two we arrive at

(9.18) ψrL2t,r+ψL4t,r(1+λλ2L2t)(ψ0L2+(ψrL2+ψL42)2)\|\frac{\psi}{r}\|_{L^{2}_{t,r}}+\|\psi\|_{L^{4}_{t,r}}\lesssim(1+\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}_{t}})(\|\psi_{0}\|_{L^{2}}+(\|\frac{\psi}{r}\|_{L^{2}}+\|\psi\|_{L^{4}}^{2})^{2})

c) A continuity argument. So far, we have established the bounds (9.17) and (9.18) in any interval II of existence for the solution where all the norms involved are finite. By Theorem 1.1 we know the problem is well-posed at least on some small time interval I=[0,σλ02]I=[0,\sigma\lambda_{0}^{-2}]; further, we note that the theory developed in [13] gives control on the two norms ψrL2t,r\|\frac{\psi}{r}\|_{L^{2}_{t,r}} and ψL4t,r\|\psi\|_{L^{4}_{t,r}}.

By Theorem 1.1 we know that the problem is well-posed at least on some small time interval I=[0,σλ02]I=[0,\sigma\lambda_{0}^{-2}]; further, we claim that on such a time interval the quantities ψL4t,r(I×[0,+))\|\psi\|_{L^{4}_{t,r}(I\times[0,+\infty))} and ψrL2t,r(I×[0,+))\|\frac{\psi}{r}\|_{L^{2}_{t,r}(I\times[0,+\infty))} are finite.

Indeed the theory developed in [13] gives control on ψL4t,r\|\psi\|_{L^{4}_{t,r}} on the interval II. While there is no formal result stated in [13], the arguments in that paper do give control on ψrL2t,r\|\frac{\psi}{r}\|_{L^{2}_{t,r}} on II as well. Alternatively, one could also argue based on a higher regularity argument. At the level of maps in H˙1H˙2\dot{H}^{1}\cap\dot{H}^{2}, it follows from (4.18) that ψL2H˙1e\psi\in L^{2}\cap\dot{H}^{1}_{e}. From this we obtain ψL4t\psi\in L^{4}_{t} and ψrL2r\frac{\psi}{r}\in L^{2}_{r}, which gives ψL4t,r\psi\in L^{4}_{t,r} and ψrL2t,r\frac{\psi}{r}\in L^{2}_{t,r} locally in time and within a compact interval inside [0,Tmax)[0,T_{max}) where λ(t)\lambda(t) belongs to a compact interval inside (0,+)(0,+\infty) and we have uniform bounds for ψ\psi in H˙2e\dot{H}^{2}_{e}.

Therefore the norms in (9.17) and (9.18) are finite on any compact subinterval of the maximal existence interval [0,Tmax)[0,T_{max}).

Next we combine the two bounds within a continuity argument in order to arrive at the conclusion of the theorem. For this we denote

(T)=λλ2L2t[0,T](T)=ψrL2t,r([0,T]×)+ψL4t,r([0,T]×).\mathcal{B}(T)=\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}_{t}[0,T]}\qquad\mathcal{M}(T)=\|\frac{\psi}{r}\|_{L^{2}_{t,r}([0,T]\times\mathbb{R})}+\|\psi\|_{L^{4}_{t,r}([0,T]\times\mathbb{R})}.

From (9.16) and (9.18) we obtain

(T)C(T),(T)C(1+(T))(δ+C2(T)+C3(T)).\mathcal{B}(T)\leq C\mathcal{M}(T),\qquad\mathcal{M}(T)\leq C(1+\mathcal{B}(T))(\delta+C\mathcal{M}^{2}(T)+C\mathcal{M}^{3}(T)).

From these relations we want to conclude that (T),(T)δ\mathcal{B}(T),\mathcal{M}(T)\lesssim\delta on the maximal interval of existence for the solutions [0,Tmax)[0,T_{max}).

Assuming that δ\delta is small enough depending on the universal constant CC, it is clear now that we can find a large enough constant DD such that if we know apriori that (T),(T)2Dδ\mathcal{B}(T),\mathcal{M}(T)\leq 2D\delta, then, based on the inequalities above, we can improve the result to (T),(T)Dδ\mathcal{B}(T),\mathcal{M}(T)\leq D\delta. This allows us to run a standard continuity argument to obtain the desired result.

For this we observe that both \mathcal{B} and \mathcal{M} are continuous functions of TT for T[0,Tmax)T\in[0,T_{max}), and that as T0T\to 0 both \mathcal{B} and \mathcal{M} have limit 0. Denoting by T1T_{1} the maximal time where ,2Dδ\mathcal{B},\mathcal{M}\leq 2D\delta, our analysis above implies that (T1),(T1)Dδ\mathcal{B}(T_{1}),\mathcal{M}(T_{1})\leq D\delta. But by the continuity of \mathcal{B}, \mathcal{M}, this contradicts the maximality of T1T_{1} unless T1=TmaxT_{1}=T_{max}.

The above argument provides the result of the Theorem for the weaker structure when norms in l2S(I+max×)l^{2}S(I^{+}_{max}\times\mathbb{R}) are replaced with ψrL2t,r(I+max×)+ψL4t,r(I+max×)\|\frac{\psi}{r}\|_{L^{2}_{t,r}(I^{+}_{max}\times\mathbb{R})}+\|\psi\|_{L^{4}_{t,r}(I^{+}_{max}\times\mathbb{R})}. However this weaker result is immediately upgraded to the full structure l2S(I+max)l^{2}S(I^{+}_{max}) based on the result in Theorem 7.1 and the dual Strichartz bound (8.2) for the nonlinearity.

9.4. Proof of Theorem 9.2

Here we already have the bound (9.1) from Theorem 9.1. Then we can apply the linear bounds in Theorem 7.3 to obtain

ψ1Sψ(0)LX¯+N(ψ)1N.\|\psi\|_{\ell^{1}S}\lesssim\|\psi(0)\|_{{L\bar{X}}}+\|N(\psi)\|_{\ell^{1}N}.

The nonlinear contribution is estimated using Lemma 8.3 to obtain

ψ1Sψ(0)LX¯+ψ2S.\|\psi\|_{\ell^{1}S}\lesssim\|\psi(0)\|_{{L\bar{X}}}+\|\psi\|_{\ell^{2}S}.

Finally we conclude using Theorem 9.1.

10. The modulation parameter dynamics - Part 2

From the previous section we recall the ODE (9.8) governing the dynamics of α\alpha and λ\lambda:

(10.1) 4e2iα(t)(αc1+iλλc2)=l(t)+q(t),4e^{2i\alpha(t)}\left(\alpha^{\prime}c_{1}+i\frac{\lambda^{\prime}}{\lambda}c_{2}\right)=l(t)+q(t),

where, as expanded in (9.11), l(t)=l1(t)+l2(t)l(t)=l^{1}(t)+l^{2}(t) with

(10.2) l1=2λ2h3λLλψ,ϰλ,l2=4λ2e2iα(t)h1λ(t)r(e2iα(t)h1λ(t)sLλψ)ds,ϰλ.l^{1}=-2\lambda^{2}\langle h_{3}^{\lambda}L_{\lambda}^{*}\psi,\varkappa^{\lambda}\rangle,\quad l^{2}=-4\lambda^{2}\langle e^{2i\alpha(t)}h_{1}^{\lambda(t)}\int_{r}^{\infty}\Re(\frac{e^{-2i\alpha(t)}h_{1}^{\lambda(t)}}{s}L_{\lambda}^{*}\psi)ds,\varkappa^{\lambda}\rangle.

Here qq largely plays a perturbative role. Indeed, we have established (see (9.16)) the following bound qL1tψrL2t,r2\|q\|_{L^{1}_{t}}\lesssim\|\frac{\psi}{r}\|_{L^{2}_{t,r}}^{2}, where the right hand side is estimated directly in terms of the initial data on the maximal time of existence, see Theorem 9.1.

It is clear that the the component responsible for a potential finite time or infinite time blow-up of λ\lambda, or for the more complex dynamics of the set of parameters (λ(t),α(t))(\lambda(t),\alpha(t)) (as opposed to stability which would correspond to a small perturbation of the original state (λ(0),α(0))(\lambda(0),\alpha(0))), is the linear component l(t)l(t); indeed if this was not present in the above ODE, then we would simply control α(t)\alpha(t) and logλ(t)\log\lambda(t) globally using the above bound for qq hence ruling out blow-up (both in finite and infinite time) and providing global bounds for both α\alpha and logλ\log\lambda. The information we have so far on ll, from Lemma 9.3 in the previous section, is not accurate enough in order to clarify the potential scenarios for the behaviour of solutions to the ODE (10.1), such as uniform bounds versus growth or even potential blow up in finite time.

In this section we perform a refined analysis of the linear term l(t)l(t), which will later be used to rule out the finite time blow-up scenario. This analysis has two main parts: an algebraic one that seeks to better understand the structure of the expressions l1l^{1} and l2l^{2} in (10.2), and an analytic one that provides effective estimates for the terms revealed by the algebraic part. The analysis carried in the algebraic part is inspired by the work of Gustafson, Nakanishi and Tsai [14], although in our context it is carried with respect to different variables. The analytic part of our analysis is entirely new, and this is what allows us to obtain the second main result of this paper, the global well-posedness result of Theorem 1.5, as well as its refined l1l^{1} Besov counterpart in Theorem 1.6. The main result of this section is the following:

Theorem 10.1.

The system (10.1) can be rewritten as follows on the maximal time interval of existence I=[0,Tmax)I=[0,T_{max}):

(10.3) α=12(ie2iα(t)t(Λ1+2Λ2))+14Q(t),(lnλ)=(e2iα(t)tΛ1)+12Q(t).\begin{split}\alpha^{\prime}&=-\frac{1}{2}\Re\left(ie^{-2i\alpha(t)}\partial_{t}(\Lambda^{1}+2\Lambda^{2})\right)+\frac{1}{4}\Re Q(t),\\ (\ln\lambda)^{\prime}&=-\Re(e^{-2i\alpha(t)}\partial_{t}\Lambda^{1})+\frac{1}{2}\Im Q(t).\end{split}

Here Λi:I\Lambda^{i}:I\rightarrow\mathbb{C}, i=1,2i=1,2 are continuous functions which satisfy the additional bounds

(10.4) ΛiH˙12[I]ψl2S[I],ΛiB˙122,1[I]ψl1S[I],\|\Lambda^{i}\|_{\dot{H}^{\frac{1}{2}}[I]}\lesssim\|\psi\|_{l^{2}S[I]},\qquad\|\Lambda^{i}\|_{\dot{B}^{\frac{1}{2}}_{2,1}[I]}\lesssim\|\psi\|_{l^{1}S[I]},

while the function QQ obeys the following estimate:

(10.5) QL1t[I]ψrL2[I](ψrL2[I]+ψL4[I]2).\|Q\|_{L^{1}_{t}[I]}\lesssim\|\frac{\psi}{r}\|_{L^{2}[I]}(\|\frac{\psi}{r}\|_{L^{2}[I]}+\|\psi\|_{L^{4}[I]}^{2}).

Further, assume [0,T]I[0,T]\subset I is a compact interval and let λTmax=maxt[0,T]λ(t)\lambda_{T}^{max}=max_{t\in[0,T]}\lambda(t). Then we have the following estimate:

(10.6) ΛiB˙122,1([0,T])(ln(2+(λTmax)2T))12ψl2S[I],i=1,2.\|\Lambda^{i}\|_{\dot{B}^{\frac{1}{2}}_{2,1}([0,T])}\lesssim\left(\ln(2+(\lambda_{T}^{max})^{2}T)\right)^{\frac{1}{2}}\|\psi\|_{l^{2}S[I]},\quad i=1,2.
Remark 2.

Here the norm for the B˙122,1[I]\dot{B}^{\frac{1}{2}}_{2,1}[I] quotient space of functions modulo constants is defined using extensions to \mathbb{R},

fB˙122,1[I]=inf{fextB˙122,1[];fext=fin I}.\|f\|_{\dot{B}^{\frac{1}{2}}_{2,1}[I]}=\inf\{\|f_{ext}\|_{\dot{B}^{\frac{1}{2}}_{2,1}[\mathbb{R}]};\ f_{ext}=f\ \text{in }I\}.

We briefly comment on the system (10.3), whose study is the main goal of the next section. The QQ term in both ODE’s is easily seen to play a perturbative role, so at leading order the dynamics is driven by Λi\Lambda^{i}. Assuming these have been properly estimated, what we have is a nonlinear ODE for the α\alpha variable and then an apparent linear ODE for λ\lambda; the equation for λ\lambda is not entirely linear since the term Λ1\Lambda^{1} depends on λ\lambda, but this dependence can be incorporated into the estimates on Λ1\Lambda^{1}, thus rendering an essentially linear ODE. In some sense, this indicates that it is the equation for α\alpha should be our main focus, though ultimately is the growth in the λ\lambda equation which we will want to control.

This section is organized as follows. In subsection 10.1 we carry the formal algebraic computations for l1l^{1} and l2l^{2}, the two components that appear in the linear term ll, and this leads us to the new form of the modulation system in (10.3). The L1L^{1} bound for QQ is proved in subsection 10.2. Finally in subsection 10.3 we turn our attention to the expressions Λ1\Lambda^{1} and Λ2\Lambda^{2} appearing in (10.3). These are formal objects which apriori are defined only if assume some additional spatial decay on the field ψ\psi, which apriori is more than what our framework allows (which is ψ(t)L2r\psi(t)\in L^{2}_{r}). The first step is to provide a rigorous justification of the fact that Λ1\Lambda^{1} and Λ2\Lambda^{2} are well-defined objects in our setting. Finally, we establish (10.4); this is based on transferring local energy decay bounds from ψ\psi to H˙12\dot{H}^{\frac{1}{2}} bounds on Λi,i=1,2\Lambda^{i},i=1,2.

10.1. An indepth analysis of the linear term l(t)l(t) and the derivation of (10.3)

In this section we proceed with a further analysis of the linear term l(t)l(t) appearing in (10.1). This section draws inspiration and technical details from the work [14] as follows: the analysis of the operator LλL_{\lambda}, their inverse and adjoints has been already developed in Section 3 of [14] and the identification of the terms Λ\Lambda is equivalent to the identification of the “normal form” in Section 9 in [14]. However, for pure comparison reasons, we note that all of our computations are performed at the level of the gauge elements, while those appearing in [14] are performed at the level of the actual map.

In later computations in this section we will need to work with a right inverse of the operator LλL_{\lambda}. It is convenient to use a regularized version defined as follows

Rλ,φf=h1λ0rr1h1λ(r)f(r)drλ2φλ(r)h1λ(r)rdr,R_{\lambda,\varphi}f=h_{1}^{\lambda}\int_{0}^{\infty}\int_{r^{\prime}}^{r}\frac{1}{h_{1}^{\lambda}(r^{\prime\prime})}f(r^{\prime\prime})dr^{\prime\prime}\lambda^{2}\varphi^{\lambda}(r^{\prime})h_{1}^{\lambda}(r^{\prime})r^{\prime}dr^{\prime},

where φC0((0,);)\varphi\in C_{0}^{\infty}((0,\infty);\mathbb{R}); in fact we assume that φ\varphi is supported in the interval (12,2)(\frac{1}{2},2).

Assuming also that we have the normalization φ,h1=1\langle\varphi,h_{1}\rangle=1, a direct computation shows that the following relations hold true:

LλRλ,φf=f,Rλ,φLλf=fh1λf,λ2φλ.L_{\lambda}R_{\lambda,\varphi}f=f,\qquad R_{\lambda,\varphi}L_{\lambda}f=f-h_{1}^{\lambda}\langle f,\lambda^{2}\varphi^{\lambda}\rangle.

We also note that

Rλ,φ=λ1DλR1,φDλ1,Lλ=λDλL1Dλ1,R_{\lambda,\varphi}=\lambda^{-1}D_{\lambda}R_{1,\varphi}D_{\lambda}^{-1},\quad L_{\lambda}=\lambda D_{\lambda}L_{1}D_{\lambda}^{-1},

where (Dλf)(x)=fλ(x)=f(λx)(D_{\lambda}f)(x)=f^{\lambda}(x)=f(\lambda x); the reason we use also the notation DλfD^{\lambda}f in addition to the standard fλf^{\lambda} is that it is more streamlined in longer formulas just like the one above.

We also need the adjoint of Rλ,φR_{\lambda,\varphi}. To compute it we set λ=1\lambda=1 and use the following inversion formula:

R1,φf(r)=00h1(r)h1(r)k(r,r,r)φ(r)h1(r)rf(r)drdr=K(r,r,r)f(r)drdr,R_{1,\varphi}f(r)=\int_{0}^{\infty}\int_{0}^{\infty}\frac{h_{1}(r)}{h_{1}(r^{\prime\prime})}k(r,r^{\prime},r^{\prime\prime})\varphi(r^{\prime})h_{1}(r^{\prime})r^{\prime}f(r^{\prime\prime})dr^{\prime\prime}dr^{\prime}=\int\int K(r,r^{\prime},r^{\prime\prime})f(r^{\prime\prime})dr^{\prime\prime}dr^{\prime},

where

k(r,r,r)=1{rrr}1{rrr}.k(r,r^{\prime},r^{\prime\prime})=1_{\{r^{\prime}\leq r^{\prime\prime}\leq r\}}-1_{\{r\leq r^{\prime\prime}\leq r^{\prime}\}}.

Then we have

R1,φg(r)=00K(r,r,r)g(r)rrdrdr=1rh1(r)00h1(r)k(r,r,r)φ(r)h1(r)rrg(r)drdr=1rh1(r)(0rrh1(r)φ(r)h1(r)rrg(r)drdrrr0drdr).\begin{split}R^{*}_{1,\varphi}g(r^{\prime\prime})&=\int_{0}^{\infty}\int_{0}^{\infty}K(r,r^{\prime},r^{\prime\prime})g(r)\frac{r}{r^{\prime\prime}}drdr^{\prime}\\ &=\frac{1}{r^{\prime\prime}h_{1}(r^{\prime\prime})}\int_{0}^{\infty}\int_{0}^{\infty}h_{1}(r)k(r,r^{\prime},r^{\prime\prime})\varphi(r^{\prime})h_{1}(r^{\prime})r^{\prime}rg(r)drdr^{\prime}\\ &=\frac{1}{r^{\prime\prime}h_{1}(r^{\prime\prime})}\left(\int_{0}^{r^{\prime\prime}}\int_{r^{\prime\prime}}^{\infty}h_{1}(r)\varphi(r^{\prime})h_{1}(r^{\prime})r^{\prime}rg(r)drdr^{\prime}-\int_{r^{\prime\prime}}^{\infty}\int^{r^{\prime\prime}}_{0}...drdr^{\prime}\right).\\ \end{split}

This implies that

(10.7) R1,φg(r)=φ,h1rh1(r)0rh1(r)rg(r)dr,r12R^{*}_{1,\varphi}g(r^{\prime\prime})=-\frac{\langle\varphi,h_{1}\rangle}{r^{\prime\prime}h_{1}(r^{\prime\prime})}\int_{0}^{r^{\prime\prime}}h_{1}(r)rg(r)dr,\qquad r^{\prime\prime}\leq\frac{1}{2}

respectively

(10.8) R1,φg(r)=φ,h1rh1(r)rh1(r)rg(r)dr,r>2.R^{*}_{1,\varphi}g(r^{\prime\prime})=\frac{\langle\varphi,h_{1}\rangle}{r^{\prime\prime}h_{1}(r^{\prime\prime})}\int_{r^{\prime\prime}}^{\infty}h_{1}(r)rg(r)dr,\qquad r^{\prime\prime}>2.

In the remaining case r[12,2]r^{\prime\prime}\in[\frac{1}{2},2] we may have contributions from both terms. It is easy to see that the general adjoint operator is given by

Rλ,φ=λ1DλR1,φDλ1.R_{\lambda,\varphi}^{*}=\lambda^{-1}D_{\lambda}R^{*}_{1,\varphi}D_{\lambda}^{-1}.

Finally, from the formula:

Rλ,φLλf=fh1λf,λ2φλR_{\lambda,\varphi}L^{\lambda}f=f-h_{1}^{\lambda}\langle f,\lambda^{2}\varphi^{\lambda}\rangle

it follows that

LλRλ,φg=gλ2φh1λ,g.L_{\lambda}^{*}R_{\lambda,\varphi}^{*}g=g-\lambda^{2}\varphi\langle h_{1}^{\lambda},g\rangle.

Note that this last formula implies that in order to have a clean recovery formula in the sense LλRλ,φg=gL_{\lambda}^{*}R_{\lambda,\varphi}^{*}g=g, we need to free gg of the h1λh_{1}^{\lambda} mode, that is

LλRλ,φg=giffg,h1λ=0.L_{\lambda}^{*}R_{\lambda,\varphi}^{*}g=g\ \ \mbox{iff}\ \ \langle g,h_{1}^{\lambda}\rangle=0.

With the above formalism in place, we are now ready to proceed with our refined analysis of the linear term l(t)=l1(t)+l2(t)l(t)=l^{1}(t)+l^{2}(t), see 10.2. We start with the analysis for l1l^{1}. Denoting

c=h3λϰλ,h1λh1λ,h1λ=h3ϰ,h1h1,h1,c=\frac{\langle h_{3}^{\lambda}\varkappa^{\lambda},h_{1}^{\lambda}\rangle}{\langle h_{1}^{\lambda},h_{1}^{\lambda}\rangle}=\frac{\langle h_{3}\varkappa,h_{1}\rangle}{\langle h_{1},h_{1}\rangle},

we compute

l12λ2=Lλψ,h3λϰλ=Lλψ,h3λϰλch1λ=Lλψ,LλRλ,φ(h3ϰch1)λ=LλLλψ,Rλ,φ(h3ϰch1)λ=H~λψ,Rλ,φ(h3ϰch1)λ=itψN(ψ),Rλ,φ(h3ϰch1)λ=itψ,Rλ,φ(h3ϰch1)λN(ψ),Rλ,φ(h3ϰch1)λ,\begin{split}\frac{l^{1}}{-2\lambda^{2}}=\langle L_{\lambda}^{*}\psi,h_{3}^{\lambda}\varkappa^{\lambda}\rangle&=\langle L_{\lambda}^{*}\psi,h_{3}^{\lambda}\varkappa^{\lambda}-ch_{1}^{\lambda}\rangle\\ &=\langle L_{\lambda}^{*}\psi,L_{\lambda}^{*}R_{\lambda,\varphi}^{*}(h_{3}\varkappa-ch_{1})^{\lambda}\rangle\\ &=\langle L_{\lambda}L_{\lambda}^{*}\psi,R_{\lambda,\varphi}^{*}(h_{3}\varkappa-ch_{1})^{\lambda}\rangle=\langle\tilde{H}_{\lambda}\psi,R_{\lambda,\varphi}^{*}(h_{3}\varkappa-ch_{1})^{\lambda}\rangle\\ &=\langle i\partial_{t}\psi-N(\psi),R_{\lambda,\varphi}^{*}(h_{3}\varkappa-ch_{1})^{\lambda}\rangle\\ &=i\langle\partial_{t}\psi,R_{\lambda,\varphi}^{*}(h_{3}\varkappa-ch_{1})^{\lambda}\rangle-\langle N(\psi),R_{\lambda,\varphi}^{*}(h_{3}\varkappa-ch_{1})^{\lambda}\rangle,\end{split}

where in the last two expressions we have used the Schrödinger map equation (4.59). The linear component itψ,Rλ,φ(h3ϰch1)λi\langle\partial_{t}\psi,R_{\lambda,\varphi}^{*}(h_{3}\varkappa-ch_{1})^{\lambda}\rangle above needs further work. We compute

(10.9) Rλ,φ(h3ϰch1)λ=λ1DλR1,φDλ1(h3ϰch1)λ=λ1DλR1,φ(h3ϰch1)=λ1DλG1,\begin{split}R_{\lambda,\varphi}^{*}(h_{3}\varkappa-ch_{1})^{\lambda}=&\ \lambda^{-1}D_{\lambda}R_{1,\varphi}^{*}D_{\lambda}^{-1}(h_{3}\varkappa-ch_{1})^{\lambda}\\ =&\ \lambda^{-1}D_{\lambda}R_{1,\varphi}^{*}(h_{3}\varkappa-ch_{1})=\lambda^{-1}D_{\lambda}G_{1},\end{split}

where

G1=R1,φ(h3ϰch1)=R1,φ(h3ϰ)cR1,φ(h1):=G1,weakcG.G_{1}=R_{1,\varphi}^{*}(h_{3}\varkappa-ch_{1})=R^{1,*}_{\varphi}(h_{3}\varkappa)-cR_{1,\varphi}^{*}(h_{1}):=G_{1,weak}-cG.

From (10.7) and (10.8) it follows G1,weak=R1,φ(h3ϰ)G_{1,weak}=R_{1,\varphi}^{*}(h_{3}\varkappa) is compactly supported in the region 12r2\frac{1}{2}\leq r^{\prime\prime}\leq 2, and it is also smooth,

|rαR1,φ(h3ϰ)(r)|α1,α0;|\partial_{r}^{\alpha}R_{1,\varphi}^{*}(h_{3}\varkappa)(r^{\prime\prime})|\lesssim_{\alpha}1,\quad\alpha\geq 0;

this is the reason we will refer to G1,weakG_{1,weak} as the weak contribution.

On the other hand GG is not compactly supported. In the regime 12r2\frac{1}{2}\leq r^{\prime\prime}\leq 2 it satisfies |rαG(r)|1,α0|\partial_{r}^{\alpha}G(r)|\lesssim 1,\alpha\geq 0. For r12r\leq\frac{1}{2}, from (10.7) it follows that

G(r)=1rh1(r)0rh12(s)sds.G(r)=-\frac{1}{rh_{1}(r)}\int_{0}^{r}h_{1}^{2}(s)sds.

A direct computation gives that for r12r\leq\frac{1}{2} we have

G(r)=(1+r4)2r3(r21+r4+tan1r2).G(r)=-\frac{(1+r^{4})}{2r^{3}}(-\frac{r^{2}}{1+r^{4}}+\tan^{-1}r^{2}).

From the above we obtain the pointwise bounds

|rαG(r)|αr3α,α0,r12.|\partial_{r}^{\alpha}G(r)|\lesssim_{\alpha}r^{3-\alpha},\quad\alpha\geq 0,\quad r\leq\frac{1}{2}.

From (10.7) it follows that for r2r\geq 2 we have

G(r)=1rh1(r)rh12(s)sds.G(r)=\frac{1}{rh_{1}(r)}\int_{r}^{\infty}h_{1}^{2}(s)sds.

A direct computation gives G(r)G(r) for r>2r>2 as follows

(10.10) G(r)=(r4+1)(tan1(1r2)+r2r4+1)2r3,r>2;G(r)=\frac{\left(r^{4}+1\right)\left(\tan^{-1}\left(\frac{1}{r^{2}}\right)+\frac{r^{2}}{r^{4}+1}\right)}{2r^{3}},\quad r>2;

in particular we have G(1r)r2=G(r)+1r,r2\frac{G(\frac{1}{r})}{r^{2}}=-G(r)+\frac{1}{r},\quad r\geq 2 and one could deduce either of the formulas in the regimes r12r\leq\frac{1}{2} and r>2r>2 from the other one. From the above we obtain the pointwise bounds

|rαG(r)|αr1α,α0.|\partial_{r}^{\alpha}G(r)|\lesssim_{\alpha}r^{-1-\alpha},\quad\alpha\geq 0.

A closer look at rGrG shows that the asymptotic behavior of rGrG at infinity is of the form

(10.11) rG=1+13r4+O(r8).rG=1+\frac{1}{3}r^{-4}+O(r^{-8}).

The following object g~=r(rG)\tilde{g}=r(rG)^{\prime} will occur later and providing estimates for it is helpful. From the above it follows that

g~=43r4+O(r8),\tilde{g}=-\frac{4}{3r^{4}}+O(r^{-8}),

for rr large. Given the trivial bounds |g~(r)|1|\tilde{g}(r)|\lesssim 1 for r1r\lesssim 1, it follows that g~L21\|\tilde{g}\|_{L^{2}}\lesssim 1. Due to the good properties of the extra term G1,weakG_{1,weak}, the above remains valid when GG is replaced by G1G_{1} and gg by g~1=r(rG1)\tilde{g}_{1}=r(rG_{1})^{\prime}. Thus we record

(10.12) g~1L21.\|\tilde{g}_{1}\|_{L^{2}}\lesssim 1.

With the characterization of G1G_{1} in place, we seek to better understand the term

itψ,λ1Gλ1=iλ20tψ(t,r)(rG1)λ(r)dr.\langle i\partial_{t}\psi,\lambda^{-1}G^{\lambda}_{1}\rangle=i\lambda^{-2}\int_{0}^{\infty}\partial_{t}\psi(t,r)(rG_{1})^{\lambda}(r)dr.

A formal computation yields

λ2itψ,λ1G1λ=i0tψ(t,r)(rG1)λ(r)dr=it0ψ(t,r)(rG1)λ(r)dri0ψ(t,r)t(rG1)λ(r)dr=it0ψ(t,r)(rG1)λ(r)driλλ0ψ(t,r)(r(rG1))λ(r)dr.\begin{split}\lambda^{2}\langle i\partial_{t}\psi,\lambda^{-1}G_{1}^{\lambda}\rangle&=i\int_{0}^{\infty}\partial_{t}\psi(t,r)(rG_{1})^{\lambda}(r)dr\\ &=i\partial_{t}\int_{0}^{\infty}\psi(t,r)(rG_{1})^{\lambda}(r)dr-i\int_{0}^{\infty}\psi(t,r)\partial_{t}(rG_{1})^{\lambda}(r)dr\\ &=i\partial_{t}\int_{0}^{\infty}\psi(t,r)(rG_{1})^{\lambda}(r)dr-i\frac{\lambda^{\prime}}{\lambda}\int_{0}^{\infty}\psi(t,r)(r(rG_{1})^{\prime})^{\lambda}(r)dr.\\ \end{split}

But the issue we face is that the integral

0ψ(t,r)(rG1)λ(t)(r)dr\int_{0}^{\infty}\psi(t,r)(rG_{1})^{\lambda(t)}(r)dr

is not finite for a generic ψL2r\psi\in L^{2}_{r} given the slow decay of G1G_{1} described earlier; indeed if λ(t)=1\lambda(t)=1 then rG1=1+O(r3)rG_{1}=1+O(r^{-3}) as rr\rightarrow\infty, and the expression

1ψ(r)dr=1ψ(r)1rrdr\int_{1}^{\infty}\psi(r)dr=\int_{1}^{\infty}\psi(r)\cdot\frac{1}{r}rdr

may fail to be integrable since χr1r1L2\chi_{r\geq 1}r^{-1}\notin L^{2}. Any additional decay on ψ\psi at infinity (in addition to ψL2\psi\in L^{2}) would fix this problem, but this would restrict the allowable initial data.

To rectify this, the key observation is that the divergent part of the above integral comes from the very low frequencies of ψ\psi, which essentially do not change in time on compact time intervals. Based on this heuristic argument, we choose instead the first correction term to be defined by

(10.13) Λ1(t)=0ψ(t,r)(rG1)λ(t)(r)ψ(0,r)(rG1)λ(0)(r)dr.\Lambda^{1}(t)=\int_{0}^{\infty}\psi(t,r)(rG_{1})^{\lambda(t)}(r)-\psi(0,r)(rG_{1})^{\lambda(0)}(r)dr.

Since the term we just added is time independent, the earlier computation is not affected. Justifying that this modified integral is finite will be the subject of the following subsection.

For now we take for granted the fact that Λ1\Lambda^{1} is well-defined and continue with our computations. Recalling what has been done above, we have established that

l1=2λ2Lλψ,h3λϰλ=2itΛ1+2q1,l^{1}=-2\lambda^{2}\langle L_{\lambda}^{*}\psi,h_{3}^{\lambda}\varkappa^{\lambda}\rangle=-2i\partial_{t}\Lambda^{1}+2q_{1},

where

q1=λN(ψ),G1λ+iλλ0ψ(r)(r(rG1))λ(r)dr.q_{1}=\lambda\langle N(\psi),G_{1}^{\lambda}\rangle+i\frac{\lambda^{\prime}}{\lambda}\int_{0}^{\infty}\psi(r)(r(rG_{1})^{\prime})^{\lambda}(r)dr.

We will prove at the end of this section that q1q_{1} has good estimates in L1tL^{1}_{t}, thus it can be absorbed into the better term QQ in (10.3).

We turn our attention to the second linear contribution:

l2=4λ2e2iα(e2iα(t)rh1λrLλψdr,h1λϰλ).l^{2}=-4\lambda^{2}e^{2i\alpha}\Re\left(e^{-2i\alpha(t)}\langle\int_{r}^{\infty}\frac{h_{1}^{\lambda}}{r}L_{\lambda}^{*}\psi dr,h_{1}^{\lambda}\varkappa^{\lambda}\rangle\right).

In the previous section we have introduced (see (9.12) and the computations right after)

𝔤1(r)=0rh1(s)ϰ(s)sds,𝔤2=𝔤1h1r2,\mathfrak{g}_{1}(r)=\int_{0}^{r}h_{1}(s)\varkappa(s)sds,\qquad\mathfrak{g}_{2}=\frac{\mathfrak{g}_{1}h_{1}}{r^{2}},

which allowed us to write

e2iα4l2=(e2iα(t)λ2Lλψ,𝔤2λ).-\frac{e^{-2i\alpha}}{4}l_{2}=\Re(e^{-2i\alpha(t)}\lambda^{2}\langle L_{\lambda}^{*}\psi,\mathfrak{g}_{2}^{\lambda}\rangle).

These functions have the following properties:

i) 𝔤1(r)=𝔤2(r)=0\mathfrak{g}_{1}(r)=\mathfrak{g}_{2}(r)=0 for r12r\leq\frac{1}{2};

ii) 𝔤1(r)=c1\mathfrak{g}_{1}(r)=c_{1} for r2r\geq 2, 𝔤2(r)=O(r4)\mathfrak{g}_{2}(r)=O(r^{-4}) as rr\rightarrow\infty.

We then proceed with the analysis of the term Lλψ,𝔤2λ\langle L_{\lambda}^{*}\psi,\mathfrak{g}_{2}^{\lambda}\rangle just as we did earlier, writing

λ2Lλψ,𝔤2λ=itΛ2+q2,\lambda^{2}\langle L_{\lambda}^{*}\psi,\mathfrak{g}_{2}^{\lambda}\rangle=i\partial_{t}\Lambda^{2}+q_{2},

where

(10.14) Λ2(t)=0ψ(t,r)(rG2)λ(t)(r)ψ(0,r)(rG2)λ(0)(r)dr,G2=R1,φ(𝔤2c3h1),\Lambda^{2}(t)=\int_{0}^{\infty}\psi(t,r)(rG_{2})^{\lambda(t)}(r)-\psi(0,r)(rG_{2})^{\lambda(0)}(r)dr,\quad G_{2}=R_{1,\varphi}^{*}(\mathfrak{g}_{2}-c_{3}h_{1}),

and

q2=λN(ψ),G2λiλλ0ψ(r)(r(rG2))λ(r)dr.q_{2}=-\lambda\langle N(\psi),G_{2}^{\lambda}\rangle-i\frac{\lambda^{\prime}}{\lambda}\int_{0}^{\infty}\psi(r)(r(rG_{2})^{\prime})^{\lambda}(r)dr.

Here we let

G2=R1,φ(𝔤2c3h1)=R1,φ𝔤2c3G:=G2,weakc3G,c3=𝔤2,h1h1,h1,G_{2}=R_{1,\varphi}^{*}(\mathfrak{g}_{2}-c_{3}h_{1})=R_{1,\varphi}^{*}\mathfrak{g}_{2}-c_{3}G:=G_{2,weak}-c_{3}G,\qquad c_{3}=\frac{\langle\mathfrak{g}_{2},h_{1}\rangle}{\langle h_{1},h_{1}\rangle},

and note that G2,weakG_{2,weak} is the weak term, while c3G-c_{3}G is the strong one, just as above. The only slight difference is that while G1,weakG_{1,weak} was compactly supported, G2,weakG_{2,weak} is not; instead G2,weak(r)=0G_{2,weak}(r)=0 for r12r\leq\frac{1}{2}, G2,weak=O(r4)G_{2,weak}=O(r^{-4}) as rr\rightarrow\infty and it is smooth. For all practical purposes G2,weakG_{2,weak} plays a similar role to G1,weakG_{1,weak}.

This gives the following representation for l2l_{2}:

l2=4e2iα(e2iα(t)itΛ2)4e2iα(e2iα(t)q2).l^{2}=-4e^{2i\alpha}\Re\left(e^{-2i\alpha(t)}i\partial_{t}\Lambda^{2}\right)-4e^{2i\alpha}\Re\left(e^{-2i\alpha(t)}q_{2}\right).

At this time our ODE system (10.1) takes the following form:

4e2iα(t)(c1α+ic2λλ)=2itΛ14e2iα(e2iαitΛ2)+Q,\begin{split}4e^{2i\alpha(t)}(c_{1}\alpha^{\prime}+ic_{2}\frac{\lambda^{\prime}}{\lambda})=-2i\partial_{t}\Lambda^{1}-4e^{2i\alpha}\Re\left(e^{-2i\alpha}i\partial_{t}\Lambda^{2}\right)+Q,\end{split}

where

Q=2q14e2iα(e2iαq2)+q.Q=2q_{1}-4e^{2i\alpha}\Re\left(e^{-2i\alpha}q_{2}\right)+q.

Taking the real and the imaginary part above we obtain the following system for α\alpha and λ\lambda:

c1α=12(ie2iα(t)t(Λ1+2Λ2))+14(e2iα(t)Q(t)),2c2(lnλ)=(e2iα(t)tΛ1)12(ie2iα(t)Q(t)).\begin{split}c_{1}\alpha^{\prime}&=-\frac{1}{2}\Re\left(ie^{-2i\alpha(t)}\partial_{t}(\Lambda^{1}+2\Lambda^{2})\right)+\frac{1}{4}\Re(e^{-2i\alpha(t)}Q(t)),\\ 2c_{2}(\ln\lambda)^{\prime}&=-\Re(e^{-2i\alpha(t)}\partial_{t}\Lambda^{1})-\frac{1}{2}\Re(ie^{-2i\alpha(t)}Q(t)).\end{split}

By simply relabeling e2iα(t)Q(t)e^{-2i\alpha(t)}Q(t) to be Q(t)Q(t), which has no effect on the QQ bounds, we obtain (10.3).

10.2. The L1L^{1} bounds for QQ

Our goal here is to prove the bound (10.5), which asserts that QQ is small in L1tL^{1}_{t}. We wll focus on the estimates for q1q_{1}; the ones for q2q_{2} are similar while qq has already been estimated in (9.16).

Using (10.9), the characterization of G1G_{1} and (8.4), we obtain the following estimate:

λ2N(ψ),Rλ((h3ϰch1)λ)L1N(ψ)rL1λ2rRλ,φ((h3ϰch1)λ)LψrL2(ψrL2+ψL42)(rG1)λLψrL2(ψrL2+ψL42).\begin{split}\|\lambda^{2}\langle N(\psi),R_{\lambda}^{*}\left((h_{3}\varkappa-ch_{1})^{\lambda}\right)\rangle\|_{L^{1}}&\lesssim\|\frac{N(\psi)}{r}\|_{L^{1}}\|\lambda^{2}rR_{\lambda,\varphi}^{*}\left((h_{3}\varkappa-ch_{1})^{\lambda}\right)\|_{L^{\infty}}\\ &\lesssim\|\frac{\psi}{r}\|_{L^{2}}(\|\frac{\psi}{r}\|_{L^{2}}+\|\psi\|_{L^{4}}^{2})\|(rG_{1})^{\lambda}\|_{L^{\infty}}\\ &\lesssim\|\frac{\psi}{r}\|_{L^{2}}(\|\frac{\psi}{r}\|_{L^{2}}+\|\psi\|_{L^{4}}^{2}).\end{split}

This provides the desired estimate for the first component in q1q_{1}. We continue with the estimate for the second component in q1q_{1}. Here we recall the estimate for g~1\tilde{g}_{1} from (10.12); based on this we obtain

|0ψrg~1λ(r)rdr|ψrL2t,rg~1λL2rψrL2rλ1g~1L2λ1ψrL2r.|\int_{0}^{\infty}\frac{\psi}{r}\tilde{g}_{1}^{\lambda}(r)rdr|\lesssim\|\frac{\psi}{r}\|_{L^{2}_{t,r}}\|\tilde{g}_{1}^{\lambda}\|_{L^{2}_{r}}\lesssim\|\frac{\psi}{r}\|_{L^{2}_{r}}\lambda^{-1}\|\tilde{g}_{1}\|_{L^{2}}\lesssim\lambda^{-1}\|\frac{\psi}{r}\|_{L^{2}_{r}}.

From this we conclude with

λλ0ψ(r)(r(rG1))λ(r)drL1tλλ2L2tλ0ψrg~1λ(r)rdrL2t(1+ψ0L2r)ψr2L2.\|\frac{\lambda^{\prime}}{\lambda}\int_{0}^{\infty}\psi(r)(r(rG_{1})^{\prime})^{\lambda}(r)dr\|_{L^{1}_{t}}\lesssim\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}_{t}}\|\lambda\int_{0}^{\infty}\frac{\psi}{r}\tilde{g}_{1}^{\lambda}(r)rdr\|_{L^{2}_{t}}\lesssim(1+\|\psi_{0}\|_{L^{2}_{r}})\|\frac{\psi}{r}\|^{2}_{L^{2}}.

This provides the desired bound for the second component in q1q_{1}.

10.3. The analysis of the correction terms Λi\Lambda^{i}

The goal of this subsection is to justify that Λi(t),i=1,2\Lambda^{i}(t),i=1,2 are well defined and continuous functions of tt, and satisfy the bounds (10.4), (10.6). We recall their definition from the previous section:

Λi(t)=0ψ(t,r)(rGi)λ(t)(r)ψ(0,r)(rGi)λ(0)(r)dr.\Lambda^{i}(t)=\int_{0}^{\infty}\psi(t,r)(rG_{i})^{\lambda(t)}(r)-\psi(0,r)(rG_{i})^{\lambda(0)}(r)dr.

The first observation is that these integrals are interpreted in a singular sense, as

(10.15) Λi(t)=limR0Rψ(t,r)(rGi)λ(t)(r)ψ(0,r)(rGi)λ(0)(r)dr.\Lambda^{i}(t)=\lim_{R\rightarrow\infty}\int_{0}^{R}\psi(t,r)(rG_{i})^{\lambda(t)}(r)-\psi(0,r)(rG_{i})^{\lambda(0)}(r)dr.

For each RR the above integrals are well-defined but the existence of the limit is not obvious for ψC(L2)\psi\in C(L^{2}). However, for such ψ\psi we can harmlessly replace the sharp cutoff with a regularized cutoff,

(10.16) Λi(t)=limR0χ0(r/R)[ψ(t,r)(rGi)λ(t)(r)ψ(0,r)(rGi)λ(0)(r)]dr.\Lambda^{i}(t)=\lim_{R\rightarrow\infty}\int_{0}^{\infty}\chi_{\leq 0}(r/R)\left[\psi(t,r)(rG_{i})^{\lambda(t)}(r)-\psi(0,r)(rG_{i})^{\lambda(0)}(r)\right]dr.

Here we recall (from Section 2) that χ0\chi_{\leq 0} is smooth, supported in (0,2)(0,2) and is identically equal to 11 on (0,1)(0,1) (we can use here any other function with these properties). Indeed, for ψC(L2)\psi\in C(L^{2}) the difference of the integrals in (10.15) and (10.16) is easily seen to converge to zero.

To understand this, we consider a dyadic decomposition in frequency for ψ\psi, and the corresponding decomposition for Λi\Lambda^{i}. Precisely, for every kk\in\mathbb{Z} we define

(10.17) Λki(t)=0Pkλ(t)ψ(t)(rGi)λ(t)dr.\Lambda_{k}^{i}(t)=\int_{0}^{\infty}P_{k}^{\lambda(t)}\psi(t)\cdot(rG_{i})^{\lambda(t)}dr.

also interpreted in a singular sense, first as as limR0R\lim_{R\rightarrow\infty}\int_{0}^{R} and then as

(10.18) Λki(t)=limR0Pkλ(t)ψ(t)χ0(r/R)(rGi)λ(t)dr.\Lambda_{k}^{i}(t)=\lim_{R\to\infty}\int_{0}^{\infty}P_{k}^{\lambda(t)}\psi(t)\cdot\chi_{\leq 0}(r/R)(rG_{i})^{\lambda(t)}dr.

With these notations we will show that

(10.19) Λi(t)=kΛki(t)Λki(0)\Lambda^{i}(t)=\sum_{k\in\mathbb{Z}}\Lambda_{k}^{i}(t)-\Lambda_{k}^{i}(0)

as a uniformly convergent series. We then estimate Λi\Lambda^{i} by separately estimating the functions Λki\Lambda_{k}^{i}, which will be thought of as the generalized Littlewood-Paley pieces of Λi\Lambda^{i}.

For now we make the observation that the limit in (10.18) does exist, uniformly in tt. To see that, we use the uniform L2L^{2} bound for ψ\psi and the Fourier representation of Pkλ(t)ψ(t)P_{k}^{\lambda(t)}\psi(t). Then it suffices to show that the limit exists if we replace Pkλ(t)ψ(t)P_{k}^{\lambda(t)}\psi(t) with a generalized eigenfunction ψξλ\psi_{\xi}^{\lambda}, uniformly for ξ\xi in a fixed dyadic range and for λ\lambda in a compact set. But this is a consequence of the asymptotic behavior of ψξ\psi_{\xi} in Theorem 5.2.

For fixed time tt we can insert an additional dyadic projection P~kλ(t)\tilde{P}_{k}^{\lambda(t)} on ψ\psi in (10.18). This is L2L^{2} selfadjoint so we can write

0Pkλ(t)ψ(t)χ0(r/R)(rGi)λ(t)dr=0Pkλ(t)ψ(t)P~kλ(χ0(r/R)λGiλ)rdr.\int_{0}^{\infty}P_{k}^{\lambda(t)}\psi(t)\cdot\chi_{\leq 0}(r/R)(rG_{i})^{\lambda(t)}dr=\int_{0}^{\infty}P_{k}^{\lambda(t)}\psi(t)\tilde{P}_{k}^{\lambda}(\chi_{\leq 0}(r/R)\lambda G_{i}^{\lambda})rdr.

This is not a-priori justified without the χ\chi cutoff as the function GiG_{i} (and thus λGiλ\lambda G_{i}^{\lambda}) does not belong to L2L^{2} since it decays like r1r^{-1} at \infty. However, it is justified with the cutoff χ0(r/R)\chi_{\leq 0}(r/R) inserted, and the limit as RR\to\infty is also well defined. This will allow us to define Pkλ(λGiλ)P_{k}^{\lambda}(\lambda G_{i}^{\lambda}) as an L2L^{2} function, and thus have the representation

(10.20) Λki(t)=0Pkλ(t)ψ(t)P~kλ(λGi)λ(t)rdr.\Lambda_{k}^{i}(t)=\int_{0}^{\infty}P_{k}^{\lambda(t)}\psi(t)\cdot\tilde{P}_{k}^{\lambda}(\lambda G_{i})^{\lambda(t)}rdr.

We will use this representation in order to obtain bounds for the functions Λki\Lambda_{k}^{i}. Our bounds will also hold uniformly with respect to the value RR used in the χ0(r/R)\chi_{\leq 0}(r/R) cutoff, which is needed in order to establish uniform convergence in (10.19). W will establish the bounds for Λik\Lambda_{i}^{k} in several steps:

Step 1: Uniform and L2L^{2} bounds for Λik\Lambda^{i}_{k}. The key part of this step is to establish L2L^{2} bounds for the Littlewood-Paley pieces of (λGi)λ(t)(\lambda G_{i})^{\lambda(t)}. These have the form

Lemma 10.2.

The following bounds hold:

(10.21) P~kλ(λGiλ)L22kλ14,\|\tilde{P}_{k}^{\lambda}(\lambda G_{i}^{\lambda})\|_{L^{2}}\lesssim\langle 2^{k}\lambda^{-1}\rangle^{-4},
(10.22) rP~kλ(λGiλ)L22k2kλ14.\|r\tilde{P}_{k}^{\lambda}(\lambda G_{i}^{\lambda})\|_{L^{2}}\lesssim 2^{-k}\langle 2^{k}\lambda^{-1}\rangle^{-4}.

These will also hold uniformly with respect to the value RR used in the cutoff χ0(r/R)\chi_{\leq 0}(r/R), and all that is used is the symbol type behavior at infinity of GiG_{i} with at least 1/r1/r decay.

For technical reasons we need to provide similar estimate for the function G~i=(rGi)\tilde{G}_{i}=(rG_{i})^{\prime}. From the properties of GiG_{i} detailed in (10.1) and in particular (10.11), it follows that G~i\tilde{G}_{i} is similar to GiG_{i} except that it has better decay at infinity, that is G~i=O(r3)\tilde{G}_{i}=O(r^{-3}).

Lemma 10.3.

The following holds true:

(10.23) Pkλ(λ2G~iλ)L22k2kλ14.\|P_{k}^{\lambda}(\lambda^{2}\tilde{G}_{i}^{\lambda})\|_{L^{2}}\lesssim 2^{k}\langle 2^{k}\lambda^{-1}\rangle^{-4}.

We remark that in order to get the estimate in (10.23) it suffices to have O(r2)O(r^{-2}) decay at infinity, which is less than what G~i\tilde{G}_{i} has. Essentially the extra decay factor of r1r^{-1} (over what GiG_{i} has) gives us the extra gain of a factor of 2k2^{k} over (10.21) and this is an improvement in the low frequency regime.

The bounds in Lemma 10.2 can be used in multiple ways. On one hand we can combine them with the uniform L2L^{2} bounds on ψ\psi, which leads us to a uniform bound for Λik(t)\Lambda^{i}_{k}(t),

(10.24) |Λik(t)|2kλ(t)14Pkλ(t)ψ(t)L2|\Lambda^{i}_{k}(t)|\lesssim\langle 2^{k}\lambda(t)^{-1}\rangle^{-4}\|P_{k}^{\lambda(t)}\psi(t)\|_{L^{2}}

On the other hand we can combine them with local energy bounds for ψ\psi, which yields an L2L^{2} bound for Λik(t)\Lambda^{i}_{k}(t),

(10.25) ΛikL2(I)2k2k(λmaxI)14Pkλ(t)ψLEk,λmaxI=suptIλ(t).\|\Lambda^{i}_{k}\|_{L^{2}(I)}\lesssim 2^{-k}\langle 2^{k}(\lambda^{max}_{I})^{-1}\rangle^{-4}\|P_{k}^{\lambda(t)}\psi\|_{LE_{k}},\qquad\lambda^{max}_{I}=\sup_{t\in I}\lambda(t).

We now return to the proof of the Lemmas:

Proof of Lemma 10.2.

Using the symbol type behavior of GiG_{i} at infinity, the Fourier transform H~λ(λGiλ)\mathcal{F}_{\tilde{H}_{\lambda}}(\lambda G_{i}^{\lambda}) can be easily defined as an improper integral by

(H~λλGiλ)(ξ)=λ0ψλξ(r)Giλ(r)rdr=limRλ0χ0(r/R)ψλξ(r)Giλ(r)rdr.(\mathcal{F}_{\tilde{H}_{\lambda}}\lambda G_{i}^{\lambda})(\xi)=\lambda\int_{0}^{\infty}\psi^{\lambda}_{\xi}(r)G_{i}^{\lambda}(r)rdr=\lim_{R\to\infty}\lambda\int_{0}^{\infty}\chi_{\leq 0}(r/R)\psi^{\lambda}_{\xi}(r)G_{i}^{\lambda}(r)rdr.

Omitting the RR truncation from here on, a direct rescaling yields

(H~λλGiλ)(ξ)=λ0ψλξ(r)Giλ(r)rdr=λ0λ12ψλ1ξ(λr)Gi(λr)rdr=0λ12ψλ1ξ(r)Gi(r)rdr.\begin{split}(\mathcal{F}_{\tilde{H}_{\lambda}}\lambda G_{i}^{\lambda})(\xi)&=\lambda\int_{0}^{\infty}\psi^{\lambda}_{\xi}(r)G_{i}^{\lambda}(r)rdr\\ &=\lambda\int_{0}^{\infty}\lambda^{\frac{1}{2}}\psi_{\lambda^{-1}\xi}(\lambda r)G_{i}(\lambda r)rdr\\ &=\int_{0}^{\infty}\lambda^{-\frac{1}{2}}\psi_{\lambda^{-1}\xi}(r)G_{i}(r)rdr.\end{split}

Thus we have

(H~λλGiλ)(ξ)=λ12fi(λ1ξ),fi(η)=0ψη(r)Gi(r)rdr,(\mathcal{F}_{\tilde{H}_{\lambda}}\lambda G_{i}^{\lambda})(\xi)=\lambda^{-\frac{1}{2}}f_{i}(\lambda^{-1}\xi),\qquad f_{i}(\eta)=\int_{0}^{\infty}\psi_{\eta}(r)G_{i}(r)rdr,

where the last integral defining fif_{i} is still defined as an improper integral. It remains to estimate the function fif_{i}. This is done by pointwise estimates in the non-oscillatory regime rη1r\lesssim\eta^{-1} and integration by parts in the oscillatory regime rη1r\gtrsim\eta^{-1}; using the properties of GiG_{i} we obtain the following:

|fi(η)|ηq(η)rη1rdr+O(η12)η12,η1;|fi(η)|ηq(η)rη1r3r4dr+O(η92)η92,η1.\begin{split}&|f_{i}(\eta)|\lesssim\eta q(\eta)\int_{r\lesssim\eta^{-1}}rdr+O(\eta^{-\frac{1}{2}})\lesssim\eta^{-\frac{1}{2}},\quad\eta\lesssim 1;\\ &|f_{i}(\eta)|\lesssim\eta q(\eta)\int_{r\lesssim\eta^{-1}}r^{3}\cdot r^{4}dr+O(\eta^{-\frac{9}{2}})\lesssim\eta^{-\frac{9}{2}},\quad\eta\gtrsim 1.\end{split}

A similar argument also yields bounds for the derivative of ff, namely

|ηfi(η)|η32η4.|\partial_{\eta}f_{i}(\eta)|\lesssim\ \eta^{-\frac{3}{2}}\langle\eta\rangle^{-4}.

For the actual H~λλGiλ\mathcal{F}_{\tilde{H}_{\lambda}}\lambda G_{i}^{\lambda}, the ff bounds translate into

|(H~λλGiλ)(ξ)|ξ12λ1ξ4.|(\mathcal{F}_{\tilde{H}_{\lambda}}\lambda G_{i}^{\lambda})(\xi)|\lesssim\xi^{-\frac{1}{2}}\langle\lambda^{-1}\xi\rangle^{-4}.

respectively

|ξ(H~λλGiλ)(ξ)|ξ32λ1ξ4.|\partial_{\xi}(\mathcal{F}_{\tilde{H}_{\lambda}}\lambda G_{i}^{\lambda})(\xi)|\lesssim\xi^{-\frac{3}{2}}\langle\lambda^{-1}\xi\rangle^{-4}.

The first set of bounds directly imply (10.21). On the other hand for (10.22) we use Lemma 5.10 and the above computations to obtain:

λrPkλGλL2rφk(ξ)fi(λ1ξ)ξL2+ξ(φk(ξ)fi(λ1ξ))L2\|\lambda rP_{k}^{\lambda}G^{\lambda}\|_{L^{2}_{r}}\lesssim\|\frac{\varphi_{k}(\xi)f_{i}(\lambda^{-1}\xi)}{\xi}\|_{L^{2}}+\|\partial_{\xi}(\varphi_{k}(\xi)f_{i}(\lambda^{-1}\xi))\|_{L^{2}}

It remains to show that each term on the right-hand side can be bounded, up to constants, by 2k2kλ142^{-k}\langle 2^{k}\lambda^{-1}\rangle^{-4}. The estimate for the first term follows from (10.21); the same applies to the second term when ξ\partial_{\xi} is applied to the dyadic cutoff, while for the term containing ξfi\partial_{\xi}f_{i} we use the bounds above for ξ(H~λλGiλ)(ξ)\partial_{\xi}(\mathcal{F}_{\tilde{H}_{\lambda}}\lambda G_{i}^{\lambda})(\xi). ∎

Proof of Lemma 10.3.

Just as in the previous lemma we compute

(H~λλ2G~1λ)(ξ)=λ12ψλ1ξ(r)G~1(r)rdr.\begin{split}(\mathcal{F}_{\tilde{H}_{\lambda}}\lambda^{2}\tilde{G}_{1}^{\lambda})(\xi)=\int\lambda^{\frac{1}{2}}\psi_{\lambda^{-1}\xi}(r)\tilde{G}_{1}(r)rdr.\end{split}

Thus with f~(η)=λ12ψη(r)G~1(r)rdr\tilde{f}(\eta)=\int\lambda^{\frac{1}{2}}\psi_{\eta}(r)\tilde{G}_{1}(r)rdr, we have (H~λλ2G~1λ)(ξ)=f~(λ1ξ)(\mathcal{F}_{\tilde{H}_{\lambda}}\lambda^{2}\tilde{G}_{1}^{\lambda})(\xi)=\tilde{f}(\lambda^{-1}\xi). A quick inspection of G~1\tilde{G}_{1} reveals that it has properties which are similar to those of G1G_{1} plus the improved decay G~1=O(r2)\tilde{G}_{1}=O(r^{-2}), see (10.11). By similar arguments as the ones used for estimating ff in Step 1, we obtain

|f~(η)|(λη)12,η1and|f~(η)|(λη)12η4,η1.|\tilde{f}(\eta)|\lesssim(\lambda\eta)^{\frac{1}{2}},\eta\lesssim 1\quad\mbox{and}\quad|\tilde{f}(\eta)|\lesssim(\lambda\eta)^{\frac{1}{2}}\eta^{-4},\eta\gtrsim 1.

Recalling that λη=ξ\lambda\eta=\xi, these bounds translate into

|(H~λλ2G~1λ)(ξ)|ξ12,ξλ,|(H~λλ2G~1λ)(ξ)|ξ12(λ1ξ)4,ξλ.\begin{split}&|(\mathcal{F}_{\tilde{H}_{\lambda}}\lambda^{2}\tilde{G}_{1}^{\lambda})(\xi)|\lesssim\xi^{\frac{1}{2}},\quad\xi\lesssim\lambda,\\ &|(\mathcal{F}_{\tilde{H}_{\lambda}}\lambda^{2}\tilde{G}_{1}^{\lambda})(\xi)|\lesssim\xi^{\frac{1}{2}}(\lambda^{-1}\xi)^{-4},\quad\xi\gtrsim\lambda.\end{split}

As a consequence of this we obtain the bound (10.23). ∎


Step 2: L2L^{2} bounds for tΛik\partial_{t}\Lambda^{i}_{k}. Our objective here is to prove that

(10.26) tΛikL2(I)2kj2|kj|10Pjλ(t)ψSj.\|\partial_{t}\Lambda^{i}_{k}\|_{L^{2}(I)}\lesssim 2^{k}\sum_{j}2^{-\frac{|k-j|}{10}}\|P_{j}^{\lambda(t)}\psi\|_{S_{j}}.

To prove this we begin with a simple computation,

tΛik=\displaystyle\partial_{t}\Lambda^{i}_{k}= tPkλψ,(λGi)λ+Pkλψ,t(λGi)λ\displaystyle\ \langle\partial_{t}P_{k}^{\lambda}\psi,(\lambda G_{i})^{\lambda}\rangle+\langle P_{k}^{\lambda}\psi,\partial_{t}(\lambda G_{i})^{\lambda}\rangle
=\displaystyle= iPkλψ,H~λ(λGi)λ+iPkλN(ψ),(λGi)λ+gk,(λGi)λ+Pkλψ,t(λGi)λ\displaystyle\ i\langle P_{k}^{\lambda}\psi,\tilde{H}_{\lambda}(\lambda G_{i})^{\lambda}\rangle+i\langle P_{k}^{\lambda}N(\psi),(\lambda G_{i})^{\lambda}\rangle+\langle g_{k},(\lambda G_{i})^{\lambda}\rangle+\langle P_{k}^{\lambda}\psi,\partial_{t}(\lambda G_{i})^{\lambda}\rangle
:=\displaystyle:= e1k+e2k+e3k+e4k,\displaystyle\ e^{1}_{k}+e^{2}_{k}+e^{3}_{k}+e^{4}_{k},

where we recall from (7.41) that gk=i[t,Pkλ(t)]ψg_{k}=i[\partial_{t},P_{k}^{\lambda}(t)]\psi; we now estimate separately each of the ekle_{k}^{l} terms.

The bound for e1ke^{1}_{k}. This is identical to the proof of (10.25), as the operator H~λ\tilde{H}_{\lambda} simply adds a 22k2^{2k} factor at frequency 2k2^{k}; the details are left as an exercise.

The bound for e2ke^{2}_{k}. This has the form

e2k=PkλN(ψ),λGiλe^{2}_{k}=\langle P_{k}^{\lambda}N(\psi),\lambda G_{i}^{\lambda}\rangle

For this we use the L2L1L^{2}L^{1} bound in (8.12):

e2kL2tPkλN(ψ)r,λrGiλL2tPkλN(ψ)rL2tL1rλrGiλLt,r2kj2|kj|10ψjSj.\|e^{2}_{k}\|_{L^{2}_{t}}\lesssim\|\langle\frac{P_{k}^{\lambda}N(\psi)}{r},\lambda rG_{i}^{\lambda}\rangle\|_{L^{2}_{t}}\lesssim\|\frac{P_{k}^{\lambda}N(\psi)}{r}\|_{L^{2}_{t}L^{1}_{r}}\|\lambda rG_{i}^{\lambda}\|_{L^{\infty}_{t,r}}\lesssim 2^{k}\sum_{j}2^{-\frac{|k-j|}{10}}\|\psi_{j}\|_{S_{j}}.

The bound for e3ke^{3}_{k}. Here we look at gk,λG1λ\langle g_{k},\lambda G_{1}^{\lambda}\rangle and establish the estimate (10.26) on [0,T][0,T]. From Lemma 5.13 we have

gk=iλ~λ1[𝒦~λ,mk]~λψ.g_{k}=i\lambda^{\prime}{\tilde{\mathcal{F}}}_{\lambda}^{-1}[\tilde{\mathcal{K}}_{\lambda},m_{k}]{\tilde{\mathcal{F}}}_{\lambda}\psi.

We split this into

gk=l,jglkj,glkj=iλ~λ1ml[𝒦~λ,mk]m~j~λψj,g_{k}=\sum_{l,j}g_{lkj},\qquad g_{lkj}=i\lambda^{\prime}{\tilde{\mathcal{F}}}_{\lambda}^{-1}m_{l}[\tilde{\mathcal{K}}_{\lambda},m_{k}]{\tilde{m}}_{j}{\tilde{\mathcal{F}}}_{\lambda}\psi_{j},

where the summands vanish unless either l=k+O(1)l=k+O(1) or j=k+O(1)j=k+O(1), and are estimated at fixed time by Lemma 5.14 as

(10.27) glkjL2r2k|λ|λ2χ2j=λχk=lχk=jψjLL2.\|g_{lkj}\|_{L^{2}_{r}}\lesssim 2^{k}\frac{|\lambda^{\prime}|}{\lambda^{2}}\chi_{2^{j}=\lambda}\chi_{k=l}\chi_{k=j}\|\psi_{j}\|_{L^{\infty}L^{2}}.

Next we write

glkj,λGλ=glkj,λP~lλGλ,\langle g_{lkj},\lambda G^{\lambda}\rangle=\langle g_{lkj},\lambda\tilde{P}_{l}^{\lambda}G^{\lambda}\rangle,

and, using (10.21), we estimate again at fixed time

|glkj,λGλ|glkjL2r2lλ142k|λ|λ2χ2k=λχk=lχk=j2lλ14ψjLL2.\begin{split}|\langle g_{lkj},\lambda G^{\lambda}\rangle|&\lesssim\|g_{lkj}\|_{L^{2}_{r}}\langle 2^{l}\lambda^{-1}\rangle^{-4}\\ &\lesssim 2^{k}\frac{|\lambda^{\prime}|}{\lambda^{2}}\chi_{2^{k}=\lambda}\chi_{k=l}\chi_{k=j}\langle 2^{l}\lambda^{-1}\rangle^{-4}\|\psi_{j}\|_{L^{\infty}L^{2}}.\end{split}

After jj and ll summation we arrive at

|gk,λGλ|2k|λ|λ2jχ2j=λχk=jψjLL2.\begin{split}|\langle g_{k},\lambda G^{\lambda}\rangle|&\lesssim 2^{k}\frac{|\lambda^{\prime}|}{\lambda^{2}}\sum_{j}\chi_{2^{j}=\lambda}\chi_{k=j}\|\psi_{j}\|_{L^{\infty}L^{2}}.\end{split}

Finally we take the L2tL^{2}_{t} norm to get a final bound

gk,λG1λL2t2kjMjχk=jPjλψLtL2r.\begin{split}\|\langle g_{k},\lambda G_{1}^{\lambda}\rangle\|_{L^{2}_{t}}\lesssim 2^{k}\sum_{j}M_{j}\chi_{k=j}\|P_{j}^{\lambda}\psi\|_{L^{\infty}_{t}L^{2}_{r}}.\end{split}

The bound for e4ke^{4}_{k}. Here we first compute

t[λG1λ]=λ(G1+rG1)λ=λλ2λ2G~1λ,G~1=(rG1).\partial_{t}[\lambda G_{1}^{\lambda}]=\lambda^{\prime}(G_{1}+rG^{\prime}_{1})^{\lambda}=\frac{\lambda^{\prime}}{\lambda^{2}}\lambda^{2}\tilde{G}_{1}^{\lambda},\qquad\tilde{G}_{1}=(rG_{1})^{\prime}.

Using (10.23) we obtain

λλ2Pkλψ,λ2G~1λ|L2tλλ2L2tsuptPkλψL2rP~kλ(λ2G~1)λL2λλ2L2t2k2k(λmaxI)14PkλψL2.\begin{split}\|\frac{\lambda^{\prime}}{\lambda^{2}}\langle P_{k}^{\lambda}\psi,\lambda^{2}\tilde{G}_{1}^{\lambda}\rangle|\|_{L^{2}_{t}}&\lesssim\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}_{t}}\sup_{t}\|P_{k}^{\lambda}\psi\|_{L^{2}_{r}}\|\tilde{P}_{k}^{\lambda}(\lambda^{2}\tilde{G}_{1})^{\lambda}\|_{L^{2}}\\ &\lesssim\|\frac{\lambda^{\prime}}{\lambda^{2}}\|_{L^{2}_{t}}2^{k}\langle 2^{k}(\lambda^{max}_{I})^{-1}\rangle^{-4}\|P_{k}^{\lambda}\psi\|_{L^{2}}.\end{split}

as needed.

Step 3: The dyadic summation of the Λik\Lambda^{i}_{k} bounds. Here we complete the proof of Theorem 10.1. Our starting point consists of the bounds (10.25) and (10.26), which hold in any time interval [0,T][0,T] of existence for the solution. Here we may assume without loss of generality that Tλ(0)2T\gtrsim\lambda(0)^{-2}, which is the existence time for the local well-posedness result. To use these bounds in order to estimate Λi\Lambda^{i} in (10.19), we separate the dyadic frequency indices kk as follows:

  1. (1)

    low frequencies, 2kT122^{k}\lesssim T^{-\frac{1}{2}}.

  2. (2)

    medium frequencies, T122kλmaxTT^{-\frac{1}{2}}\lesssim 2^{k}\lesssim\lambda^{max}_{T}.

  3. (3)

    high frequencies, λmaxT2k\lambda^{max}_{T}\lesssim 2^{k}.

Low frequencies: Here we directly use Hölder’s inequality and (10.26) to obtain a uniform bound

(10.28) |Λki(t)Λki(0)|(22kt)12jχj=kPjψSj,|\Lambda_{k}^{i}(t)-\Lambda_{k}^{i}(0)|\lesssim(2^{2k}t)^{\frac{1}{2}}\sum_{j}\chi_{j=k}\|P_{j}\psi\|_{S_{j}},

as well as an integrated form of this,

(10.29) Λki(t)Λki(0)L2[0,T]2kTjχj=kPjψSj.\|\Lambda_{k}^{i}(t)-\Lambda_{k}^{i}(0)\|_{L^{2}[0,T]}\lesssim 2^{k}T\sum_{j}\chi_{j=k}\|P_{j}\psi\|_{S_{j}}.

Medium and high frequencies: In this case our time interval is long enough so we can interpolate between the L2L^{2} and H˙1\dot{H}^{1} bound for Λki\Lambda_{k}^{i}, which yields

(10.30) ΛkiL2k(λmaxT)12jχj=kPjψSj.\|\Lambda_{k}^{i}\|_{L^{\infty}}\lesssim\langle 2^{k}(\lambda^{max}_{T})^{-1}\rangle^{-2}\sum_{j}\chi_{j=k}\|P_{j}\psi\|_{S_{j}}.

where the distinction is that we have extra decay for high frequencies.


Given the above bounds with decay both at low and at high frequencies, it is clear that we have uniform convergence in (10.19), with a uniform overall bound obtained by applying Hölder’s inequality for the intermediate frequencies, namely

(10.31) ΛiL(ln(T(λmaxT)2))12ul2S.\|\Lambda^{i}\|_{L^{\infty}}\lesssim(\ln(T(\lambda^{max}_{T})^{2}))^{\frac{1}{2}}\|u\|_{l^{2}S}.

If instead we assume l1l^{1} summability then this becomes a uniform bound,

(10.32) ΛiLul1S.\|\Lambda^{i}\|_{L^{\infty}}\lesssim\|u\|_{l^{1}S}.

We now consider the H˙12\dot{H}^{\frac{1}{2}} bound for Λi\Lambda^{i}. One difficulty we face is that we are in a bounded interval [0,T][0,T]. But this can be easily bypassed by a standard extension and truncation argument. Here the constants do not matter, but we have to consider them carefully when we truncate. The best strategy is to separate two cases, exactly as above:

  1. (1)

    Low frequency: here we take the function ΛkiΛki(0)\Lambda_{k}^{i}-\Lambda_{k}^{i}(0) in [0,T][0,T], which we simply extend by reflection to [0,2T][0,2T] and then by 0 outside this interval. This leaves the bounds (10.29) and (10.26) unchanged.

  2. (2)

    Medium and high frequencies: here take the function Λki\Lambda_{k}^{i} in [0,T][0,T], which we extend by reflection to [T,2T][-T,2T], and then truncate it outside [0,T][0,T]. This leaves the bounds (10.25) and (10.26) unchanged.

After this, we estimate the dyadic L2L^{2} norms by interpolating between the L2L^{2} and the H˙1\dot{H}^{1} norms. For the low frequency part the two norms balance exactly at time frequency T1T^{-1}, and we obtain (using 𝐏m\mathbf{P}_{m} for Littlewood-Paley projections in time)

𝐏m(Λki(t)Λki(0))extH˙12(2kT12)χm=log2Tjχj=kPjψSj.\|\mathbf{P}_{m}(\Lambda_{k}^{i}(t)-\Lambda_{k}^{i}(0))_{ext}\|_{\dot{H}^{\frac{1}{2}}}\lesssim(2^{k}T^{\frac{1}{2}})\chi_{m=-\log_{2}T}\sum_{j}\chi_{j=k}\|P_{j}\psi\|_{S_{j}}.

For the medium frequency part the two norms balance exactly at time frequency 22k2^{2k}, and we obtain

𝐏m(Λki)extH˙12χm=2kjχj=kPjψSj.\|\mathbf{P}_{m}(\Lambda_{k}^{i})_{ext}\|_{\dot{H}^{\frac{1}{2}}}\lesssim\chi_{m=2k}\sum_{j}\chi_{j=k}\|P_{j}\psi\|_{S_{j}}.

Finally for the high frequency part the two norms balance exactly at time frequency 2mk=22k(2k(λmaxT)1)42^{m_{k}}=2^{2k}(2^{k}(\lambda^{max}_{T})^{-1})^{4}, and we obtain

𝐏m(Λki)extH˙12χm=mk(2k(λmaxT)1)2jχj=kPjψSj.\|\mathbf{P}_{m}(\Lambda_{k}^{i})_{ext}\|_{\dot{H}^{\frac{1}{2}}}\lesssim\chi_{m=m_{k}}(2^{k}(\lambda^{max}_{T})^{-1})^{-2}\sum_{j}\chi_{j=k}\|P_{j}\psi\|_{S_{j}}.

Since all constants in front of the sums are 1\lesssim 1 and we have in all cases off-diagonal decay, the summation with respect to mm is straightforward, and we obtain immediately the bounds in (10.4). For (10.6) we observe that we have additional decay above both for the low and for the high frequencies, so it suffices to bound the contributions of intermediate kk. But there we simply convert the l2l^{2} norm to l1l^{1} using Hölder’s inequality, noting that the number of intermediate dyadic regions is about ln(T(λmaxT)2)\ln(T(\lambda^{max}_{T})^{2}).

11. An abstract ode result

In this section we consider the solvability question for nonlinear ode systems such as our modulation equations (10.3). As written there, the two components are partially uncoupled, in that it suffices to solve first the α\alpha equation, and then the λ\lambda equation is a direct integration. To keep the notations simple in this section we consider a more general vector valued model of the form

(11.1) α=(n(α)f(t))+g(t),α(0)=α0,\alpha^{\prime}=\Re(n(\alpha)f^{\prime}(t))+g(t),\quad\alpha(0)=\alpha_{0},

where nn is assumed to be a globally C3C^{3} function, nC3<+\|n\|_{C^{3}}<+\infty, while ff and gg are taken in the spaces

fB˙122,1[],gL1().f\in\dot{B}^{\frac{1}{2}}_{2,1}[\mathbb{R}],\qquad g\in L^{1}(\mathbb{R}).

For this problem we look for solutions in the space

(11.2) Z=(B˙122,1[]+W˙1,1[])LC0,Z=(\dot{B}^{\frac{1}{2}}_{2,1}[\mathbb{R}]+\dot{W}^{1,1}[\mathbb{R}])\cap L^{\infty}\subset C^{0},

where the LL^{\infty} component of the norm simply has the goal of controlling constants. We successively consider the small data case and then the large data case. For the latter we will also use the same space in an interval II, either bounded or unbounded,

(11.3) Z[I]=(B˙122,1[I]+W˙1,1[I])LC0.Z[I]=(\dot{B}^{\frac{1}{2}}_{2,1}[I]+\dot{W}^{1,1}[I])\cap L^{\infty}\subset C^{0}.

11.1. Iterating small Besov data

Our goal here is to show the following

Proposition 11.1.

There exists ϵ>0\epsilon>0 such that if

(11.4) fB˙122,1[]ϵ,gL1[]ϵ,\|f\|_{\dot{B}^{\frac{1}{2}}_{2,1}[\mathbb{R}]}\leq\epsilon,\qquad\|g\|_{L^{1}[\mathbb{R}]}\leq\epsilon,

then (11.1) has a unique global solution α\alpha with

(11.5) αα0Zϵ.\|\alpha-\alpha_{0}\|_{Z}\lesssim\epsilon.

We remark that as a direct consequence, the same result holds also in any interval II.

Proof.

We will use a fixed point argument in the space ZZ. We also denote by ZZ^{\prime} the space of derivatives of functions in ZZ, namely

Z=B˙12,1[]+L1[].Z^{\prime}=\dot{B}^{-\frac{1}{2},1}[\mathbb{R}]+L^{1}[\mathbb{R}].

For these spaces we have the following properties:

Lemma 11.2.

a) ZZ is an algebra.

b) We have the bilinear bound ZZZZ\cdot Z^{\prime}\subset Z^{\prime}.

c) For a C3C^{3} function nn with n(0)=0n(0)=0 we have the Moser inequality

n(f)ZfZ+fZ3.\|n(f)\|_{Z}\lesssim\|f\|_{Z}+\|f\|_{Z}^{3}.
Proof.

a) By the Leibniz rule, part (a) may be seen as a direct consequence of part (b).

b) Consider a product vwvw^{\prime} with v,wZv,w\in Z. For vv and ww we need to consider the two components of the ZZ norm. So we let v=v1+v2,w=w1+w2v=v_{1}+v_{2},w=w_{1}+w_{2} with v1,w1B˙122,1()v_{1},w_{1}\in\dot{B}^{\frac{1}{2}}_{2,1}(\mathbb{R}) and v2,w2W˙1,1()v_{2},w_{2}\in\dot{W}^{1,1}(\mathbb{R}), so that

v1B˙122,1+v2W˙1,1vZ,w1B˙122,1+w2W˙1,1wZ.\|v_{1}\|_{\dot{B}^{\frac{1}{2}}_{2,1}}+\|v_{2}\|_{\dot{W}^{1,1}}\lesssim\|v\|_{Z},\qquad\|w_{1}\|_{\dot{B}^{\frac{1}{2}}_{2,1}}+\|w_{2}\|_{\dot{W}^{1,1}}\lesssim\|w\|_{Z}.

The contribution of w2w_{2} to vwvw^{\prime} is easily estimated in L1L^{1}, so it remains to consider the product vw1vw^{\prime}_{1}. We expand it as follows

vw1=kP<k10vPkw1+kPkvP<k10w1+|k1k2|<10Pk1v1Pk2w1+|k1k2|<10Pk1v2Pk2w1,vw_{1}^{\prime}=\sum_{k}P_{<k-10}vP_{k}w_{1}^{\prime}+\sum_{k}P_{k}vP_{<k-10}w_{1}^{\prime}+\sum_{|k_{1}-k_{2}|<10}P_{k_{1}}v_{1}P_{k_{2}}w_{1}^{\prime}+\sum_{|k_{1}-k_{2}|<10}P_{k_{1}}v_{2}P_{k_{2}}w_{1}^{\prime},

and we consider each sum separately. The summands in the first sum are localized at frequency 2k2^{k}, so it is convenient to bound the sum in the Besov norm,

kP<k10vPkw1B˙12,1k2k2P<k10vLPkw1L2vLw1B˙122,1.\left\|\sum_{k}P_{<k-10}vP_{k}w_{1}^{\prime}\right\|_{\dot{B}^{-\frac{1}{2},1}}\lesssim\sum_{k}2^{-\frac{k}{2}}\|P_{<k-10}v\|_{L^{\infty}}\|P_{k}w_{1}^{\prime}\|_{L^{2}}\lesssim\|v\|_{L^{\infty}}\|w_{1}\|_{\dot{B}^{\frac{1}{2}}_{2,1}}.

The second sum is similar,

kPkvP<k10w1B˙12,1k2k2PkvLP<k10w1L2vLw1B˙122,1.\left\|\sum_{k}P_{k}vP_{<k-10}w_{1}^{\prime}\right\|_{\dot{B}^{-\frac{1}{2},1}}\lesssim\sum_{k}2^{-\frac{k}{2}}\|P_{k}v\|_{L^{\infty}}\|P_{<k-10}w_{1}^{\prime}\|_{L^{2}}\lesssim\|v\|_{L^{\infty}}\|w_{1}\|_{\dot{B}^{\frac{1}{2}}_{2,1}}.

But the remaining two sums are instead estimated in L1L^{1}:

|k1k2|<10Pk1v1Pk2w1L1|k1k2|<102k2Pk1v1L2Pk2w1L2v1B˙122,1w1B˙122,1.\|\sum_{|k_{1}-k_{2}|<10}P_{k_{1}}v_{1}P_{k_{2}}w_{1}^{\prime}\|_{L^{1}}\lesssim\sum_{|k_{1}-k_{2}|<10}2^{k_{2}}\|P_{k_{1}}v_{1}\|_{L^{2}}\|P_{k_{2}}w_{1}\|_{L^{2}}\lesssim\|v_{1}\|_{\dot{B}^{\frac{1}{2}}_{2,1}}\|w_{1}\|_{\dot{B}^{\frac{1}{2}}_{2,1}}.

respectively

|k1k2|<10Pk1v2Pk2w1L1|k1k2|<10Pk1v2L1Pk2w1Lv2L1w1B˙122,1.\|\sum_{|k_{1}-k_{2}|<10}P_{k_{1}}v_{2}P_{k_{2}}w_{1}^{\prime}\|_{L^{1}}\lesssim\sum_{|k_{1}-k_{2}|<10}\|P_{k_{1}}v_{2}\|_{L^{1}}\|P_{k_{2}}w_{1}^{\prime}\|_{L^{\infty}}\lesssim\|v_{2}^{\prime}\|_{L^{1}}\|w_{1}\|_{\dot{B}^{\frac{1}{2}}_{2,1}}.

c) Let f=f1+f2f=f_{1}+f_{2} where f1B˙122,1f_{1}\in\dot{B}^{\frac{1}{2}}_{2,1} and f2W˙1,1f_{2}\in\dot{W}^{1,1}. The easier case is when f1=0f_{1}=0, where the Moser bound is a simple application of chain rule. To deal with f1f_{1}, we consider a continuous Littlewood-Paley expansion

f1=Phf1dh.f_{1}=\int_{-\infty}^{\infty}P_{h}f_{1}dh.

For instance, for hh\in\mathbb{R}, we can define PhP_{\leq h} to be the zero order multiplier whose symbol is χ0(ξ2h)\chi_{\leq 0}(\frac{\xi}{2^{h}}); recall from Section 2 that χ0\chi_{\leq 0} is smooth, supported in (0,2)(0,2) and is identically equal to 11 on (0,1)(0,1) . Then we let Phf1=ddhPhf1P_{h}f_{1}=\frac{d}{dh}P_{\leq h}f_{1} and record that

f1,k:=Pkf1=kPhf1dh;f_{1,\leq k}:=P_{\leq k}f_{1}=\int_{-\infty}^{k}P_{h}f_{1}dh;

here kk\in\mathbb{R} (not to be confused with the standard choice in \mathbb{Z}). Accordingly we can define the operators P<h,Ph,P>hP_{<h},P_{\geq h},P_{>h}. We record the following basic inequalities:

(11.6) 2hPhfLpfLp,2hPhfLp+PhfLpfLp.\|2^{-h}P_{\leq h}f^{\prime}\|_{L^{p}}\leq\|f\|_{L^{p}},\quad\|2^{h}P_{\geq h}f\|_{L^{p}}+\|P_{\geq h}f^{\prime}\|_{L^{p}}\leq\|f^{\prime}\|_{L^{p}}.

which hold for any 1p1\leq p\leq\infty. The proof of these estimates is simply based on estimating kernels in L1L^{1} and it relies on the smoothness of the frequency cut-off used (that is χ0\chi_{\leq 0}); the details are left as an exercise.

Based on the calculus above, we obtain the following expansion

n(f)=\displaystyle n(f)= n(f2)+ddhn(f1,<h+f2)dh\displaystyle\ n(f_{2})+\int_{-\infty}^{\infty}\frac{d}{dh}n(f_{1,<h}+f_{2})dh
=\displaystyle= n(f2)+n(f1,<h+f2)Phf1dh\displaystyle n(f_{2})+\int_{-\infty}^{\infty}n^{\prime}(f_{1,<h}+f_{2})P_{h}f_{1}dh
=\displaystyle= n(f2)+(n(f1,<h+f2)n(f1,<h+f2,<h))Phf1+n(f1,<h+f2,<h)Phf1dh\displaystyle n(f_{2})+\int_{-\infty}^{\infty}(n^{\prime}(f_{1,<h}+f_{2})-n^{\prime}(f_{1,<h}+f_{2,<h}))P_{h}f_{1}+n^{\prime}(f_{1,<h}+f_{2,<h})P_{h}f_{1}dh
:=\displaystyle:= n(f2)+(g1h+g2h)dh:=n(f2)+g1+g2.\displaystyle n(f_{2})+\int_{-\infty}^{\infty}(g_{1}^{h}+g_{2}^{h})dh:=n(f_{2})+g_{1}+g_{2}.

Here we estimate g1hg_{1}^{h} and g2hg_{2}^{h} separately. We place g1hg_{1}^{h} in W˙1,1\dot{W}^{1,1}, for which we compute

(g1h)=\displaystyle(g_{1}^{h})^{\prime}= (n(f1,<h+f2)(f1,<h+f2)n(f1,<h+f2,<h)(f1,<h+f2,<h))Phf1\displaystyle\ (n^{\prime\prime}(f_{1,<h}+f_{2})(f^{\prime}_{1,<h}+f^{\prime}_{2})-n^{\prime\prime}(f_{1,<h}+f_{2,<h})(f^{\prime}_{1,<h}+f^{\prime}_{2,<h}))P_{h}f_{1}
+(n(f1,<h+f2)n(f1,<h+f2,<h))Phf1.\displaystyle\ +(n^{\prime}(f_{1,<h}+f_{2})-n^{\prime}(f_{1,<h}+f_{2,<h}))P_{h}f_{1}^{\prime}.

Estimating this in L1L^{1} we obtain

(g1h)L1\displaystyle\|(g_{1}^{h})^{\prime}\|_{L^{1}}\lesssim f2f2,<hL1f1,<h+f2,<hLPhf1L+f2f2,<hL1Phf1L\displaystyle\ \|f_{2}-f_{2,<h}\|_{L^{1}}\|f^{\prime}_{1,<h}+f^{\prime}_{2,<h}\|_{L^{\infty}}\|P_{h}f_{1}\|_{L^{\infty}}+\|f^{\prime}_{2}-f^{\prime}_{2,<h}\|_{L^{1}}\|P_{h}f_{1}\|_{L^{\infty}}
+f2f2,<hL1Phf1L\displaystyle+\|f_{2}-f_{2,<h}\|_{L^{1}}\|P_{h}f^{\prime}_{1}\|_{L^{\infty}}
\displaystyle\lesssim 2hf2,hL12h(f1,<h+f2,<h)LPhf1L+f2,hL1Phf1L\displaystyle\|2^{h}f_{2,\geq h}\|_{L^{1}}\|2^{-h}(f^{\prime}_{1,<h}+f^{\prime}_{2,<h})\|_{L^{\infty}}\|P_{h}f_{1}\|_{L^{\infty}}+\|f^{\prime}_{2,\geq h}\|_{L^{1}}\|P_{h}f_{1}\|_{L^{\infty}}
+2hf2,hL12hPhf1L\displaystyle+\|2^{h}f_{2,\geq h}\|_{L^{1}}\|2^{-h}P_{h}f^{\prime}_{1}\|_{L^{\infty}}
\displaystyle\lesssim f2L1(1+f1L+f2L)Phf1L\displaystyle\ \|f_{2}^{\prime}\|_{L^{1}}(1+\|f_{1}\|_{L^{\infty}}+\|f_{2}\|_{L^{\infty}})\|P_{h}f_{1}\|_{L^{\infty}}
\displaystyle\lesssim f2L1(1+fZ)Phf1B˙122,1,\displaystyle\ \|f_{2}^{\prime}\|_{L^{1}}(1+\|f\|_{Z})\|P_{h}f_{1}\|_{{\dot{B}^{\frac{1}{2}}_{2,1}}},

where we have used the basic estimates in (11.6) and the Phf1P_{h}f_{1} factor was bounded in terms of the Besov norm.

After hh integration this yields

g1L1f2L1f1B˙122,1(1+fZ).\|g^{\prime}_{1}\|_{L^{1}}\lesssim\|f_{2}^{\prime}\|_{L^{1}}\|f_{1}\|_{\dot{B}^{\frac{1}{2}}_{2,1}}(1+\|f\|_{Z}).

On the other hand g2hg_{2}^{h} may be placed in the Besov space, by estimating it in L2L^{2} and in H˙1\dot{H}^{1}, as follows:

g2hL2Phf1L22h2Phf1B˙122,1\|g_{2}^{h}\|_{L^{2}}\lesssim\|P_{h}f_{1}\|_{L^{2}}\lesssim 2^{-\frac{h}{2}}\|P_{h}f_{1}\|_{\dot{B}^{\frac{1}{2}}_{2,1}}

respectively

(g2h)L2Phf1L2+f1,<h+f2,<hLPhf1L22h2(1+fZ)Phf1B˙122,1\|(g_{2}^{h})^{\prime}\|_{L^{2}}\lesssim\|P_{h}f^{\prime}_{1}\|_{L^{2}}+\|f^{\prime}_{1,<h}+f^{\prime}_{2,<h}\|_{L^{\infty}}\|P_{h}f_{1}\|_{L^{2}}\lesssim 2^{\frac{h}{2}}(1+\|f\|_{Z})\|P_{h}f_{1}\|_{\dot{B}^{\frac{1}{2}}_{2,1}}

Combining the two we obtain

g2hB˙122,1(1+fZ)Phf1B˙122,1.\|g_{2}^{h}\|_{\dot{B}^{\frac{1}{2}}_{2,1}}\lesssim(1+\|f\|_{Z})\|P_{h}f_{1}\|_{\dot{B}^{\frac{1}{2}}_{2,1}}.

This concludes the argument for the Lemma. ∎

This lemma allows us to run a fixed point argument as follows. Given αZ\alpha\in Z, we define

N(α)=(n(αf)+g)N(\alpha)=\Re(n(\alpha f^{\prime})+g)

Then (11.1) is equivalent to the integral form

α(t)=0tN(α)(s)ds+α0.\alpha(t)=\int_{0}^{t}N(\alpha)(s)ds+\alpha_{0}.

We solve this using the contraction principle in a ball in ZZ of size CϵC\epsilon, where CC is a universal large constant. To achieve this it suffices to show that the operator uN(u)u\to N(u) maps BZ(Cϵ)B_{Z}(C\epsilon) to BZ(ϵ)B_{Z^{\prime}}(\epsilon) with a small Lipschitz constant. But this is a direct consequence of the lemma above.

11.2. Iterating large Besov data

Here we consider the same ode (11.1), but we allow ff and gg to be large in B˙122,1\dot{B}^{\frac{1}{2}}_{2,1}, respectively L1L^{1}.

Proposition 11.3.

For M1M\gtrsim 1, assume that

(11.7) fB˙122,1()M,gL1()M.\|f\|_{\dot{B}^{\frac{1}{2}}_{2,1}(\mathbb{R})}\leq M,\qquad\|g\|_{L^{1}(\mathbb{R})}\leq M.

then (11.1) has a unique global solution α\alpha with

(11.8) αα0ZM2.\|\alpha-\alpha_{0}\|_{Z}\lesssim M^{2}.

We remark that as a direct consequence, the same result holds also in any interval II.

Proof.

The main ingredient of the proof is the following divisibility lemma for the B˙122,1()\dot{B}^{\frac{1}{2}}_{2,1}(\mathbb{R}) norm:

Lemma 11.4.

a) Given an arbitrary partition of \mathbb{R} into intervals Ij=[aj,bj]I_{j}=[a_{j},b_{j}], we have the divisibility bounds

(11.9) jf2B˙122,1[Ij]f2B˙122,1[],\sum_{j}\|f\|^{2}_{\dot{B}^{\frac{1}{2}}_{2,1}[I_{j}]}\lesssim\|f\|^{2}_{\dot{B}^{\frac{1}{2}}_{2,1}[\mathbb{R}]},

respectively

(11.10) jf2W˙1,1[Ij]f2W˙1,1[].\sum_{j}\|f\|^{2}_{\dot{W}^{1,1}[I_{j}]}\lesssim\|f\|^{2}_{\dot{W}^{1,1}[\mathbb{R}]}.

b) For a converse bound, we have

(11.11) fB˙122,1()+W˙1,1()jfB˙122,1(Ij)+W˙1,1[Ij].\|f\|_{\dot{B}^{\frac{1}{2}}_{2,1}(\mathbb{R})+\dot{W}^{1,1}(\mathbb{R})}\lesssim\sum_{j}\|f\|_{\dot{B}^{\frac{1}{2}}_{2,1}(I_{j})+\dot{W}^{1,1}[I_{j}]}.
Proof.

a) The bound (11.10) is obvious, so we focus on (11.9). Since on the right of (11.9) we have an 1\ell^{1}-Besov norm, without any restriction in generality we can assume that ff is localized at a single dyadic frequency. By scaling, we can assume that this frequency is 11. Normalizing we can also assume that fL2=1\|f\|_{L^{2}}=1 therefore

xfL2+x2fL21.\|\partial_{x}f\|_{L^{2}}+\|\partial_{x}^{2}f\|_{L^{2}}\lesssim 1.

We measure the L2L^{2} norm of ff in unit size intervals [k,k+1][k,k+1] and bound it using a frequency envelope (ck)(c_{k}), so that the following properties hold:

  1. i)

    fL2[k,k+1]+fL2[k,k+1]ck,k\|f\|_{L^{2}[k,k+1]}+\|f^{\prime}\|_{L^{2}[k,k+1]}\leq c_{k},\forall k;

  2. ii)

    (ck)(c_{k}) is slowly varying in the following sense: ciij2cj,i,jc_{i}\lesssim\langle i-j\rangle^{2}c_{j},\forall i,j;

  3. iii)

    kck21.\sum_{k}c_{k}^{2}\lesssim 1.

In particular by Sobolev embeddings we will also have

fL[k,k+1]+fL[k,k+1]+fL2[k,k+1]ck.\|f\|_{L^{\infty}[k,k+1]}+\|f^{\prime}\|_{L^{\infty}[k,k+1]}+\|f^{\prime}\|_{L^{2}[k,k+1]}\lesssim c_{k}.

For an interval II we denote

cI2=I[k,k+1]ck2.c_{I}^{2}=\sum_{I\cap[k,k+1]\neq\emptyset}c_{k}^{2}.

To measure the Besov norm of the function ff in an interval II we use the interpolation inequality

fB˙122,1[I]2ffIL2[I]fL2[I],\|f\|_{\dot{B}^{\frac{1}{2}}_{2,1}[I]}^{2}\lesssim\|f-f_{I}\|_{L^{2}[I]}\|f^{\prime}\|_{L^{2}[I]},

where fI=|I|1Iff_{I}=|I|^{-1}\int_{I}f is the average of ff on II. This is easily seen in \mathbb{R}, and then it can be transferred to an interval II using a suitable extension which preserves the size of norms on the right (e.g. one even reflection followed by a constant extension).

Then from Poincare’s inequality we obtain

fB˙122,1[I]2|I|fL2[I]2.\|f\|_{\dot{B}^{\frac{1}{2}}_{2,1}[I]}^{2}\lesssim|I|\|f^{\prime}\|_{L^{2}[I]}^{2}.

To prove (11.9) we separate in intervals IjI_{j} into two classes:

i) long intervals, |Ij|1|I_{j}|\geq 1. Here we use the above interpolation inequality to write

fB122,1[Ij]2cIj2.\|f\|_{B^{\frac{1}{2}}_{2,1}[I_{j}]}^{2}\lesssim c_{I_{j}}^{2}.

Then after summation we get

IjlongfB122,1[Ij]2IjlongcIj21,\sum_{I_{j}\ long}\|f\|_{B^{\frac{1}{2}}_{2,1}[I_{j}]}^{2}\lesssim\sum_{I_{j}\ long}c_{I_{j}}^{2}\lesssim 1,

since each ckc_{k} is counted at most twice.

ii) Short intervals, |Ij|1|I_{j}|\leq 1. Then we use the second bound above to write

fB122,1[Ij]2|Ij|cIj2.\|f\|_{B^{\frac{1}{2}}_{2,1}[I_{j}]}^{2}\lesssim|I_{j}|c_{I_{j}}^{2}.

Here we may have many subintervals IjI_{j} intersecting a given unit interval [k,k+1][k,k+1], but the sum of their lengths is bounded by 33. Therefore after summation we obtain

IjshortfB122,1[Ij]2Ijshort|Ij|cIj2kck21.\sum_{I_{j}\ short}\|f\|_{B^{\frac{1}{2}}_{2,1}[I_{j}]}^{2}\lesssim\sum_{I_{j}\ short}|I_{j}|c_{I_{j}}^{2}\lesssim\sum_{k}c_{k}^{2}\lesssim 1.

This concludes the proof of part (a) of the lemma.

b) The result would be trivial for the W˙1,1\dot{W}^{1,1} norm. In order to deal with the larger space (B˙122,1+W˙1,1)[Ij](\dot{B}^{\frac{1}{2}}_{2,1}+\dot{W}^{1,1})[I_{j}], it suffices to prove the following Lemma for a single interval:

Lemma 11.5.

Let II\subset\mathbb{R} be any interval and f(B˙122,1+W˙1,1)[I]f\in(\dot{B}^{\frac{1}{2}}_{2,1}+\dot{W}^{1,1})[I]. Then there exists a decomposition f=f1+f2f=f_{1}+f_{2} in II where

f1B˙122,1()+f2W˙1,1[I]f(B˙122,1+W˙1,1)[I]\|f_{1}\|_{\dot{B}^{\frac{1}{2}}_{2,1}(\mathbb{R})}+\|f_{2}\|_{\dot{W}^{1,1}[I]}\lesssim\|f\|_{(\dot{B}^{\frac{1}{2}}_{2,1}+\dot{W}^{1,1})[I]}

with f1f_{1} supported in II.

Proof.

Rescaling we can assume that I=[0,1]I=[0,1]. It suffices to consider the case when we have fB˙122,1()f\in\dot{B}^{\frac{1}{2}}_{2,1}(\mathbb{R}). This means that there exists an extension, still denoted by ff, so that

fB˙122,1()fB˙122,1(I).\|f\|_{\dot{B}^{\frac{1}{2}}_{2,1}(\mathbb{R})}\lesssim\|f\|_{\dot{B}^{\frac{1}{2}}_{2,1}(I)}.

We consider a dyadic decomposition f=jPjff=\sum_{j\in\mathbb{Z}}P_{j}f, and construct f1f_{1} by truncating the dyadic pieces appropriately. Precisely, we set

f1=j0f1,j:=j>0χjPjf,f_{1}=\sum_{j\geq 0}f_{1,j}:=\sum_{j>0}\chi_{j}P_{j}f,

where the cutoff functions χj\chi_{j} have support in I=[0,1]I=[0,1] and equal 11 in [2j,12j]I[2^{-j},1-2^{-j}]\subset I, and satisfy |χj|2j|\chi_{j}^{\prime}|\lesssim 2^{-j}. Then we can estimate f1f_{1} as follows:

f1B˙122,1()j>02j2f1,jL2+2j2f1,jL2j>02j2PjfL2+2j2PjfL2fB˙122,1().\|f_{1}\|_{\dot{B}^{\frac{1}{2}}_{2,1}(\mathbb{R})}\lesssim\sum_{j>0}2^{\frac{j}{2}}\|f_{1,j}\|_{L^{2}}+2^{-\frac{j}{2}}\|f^{\prime}_{1,j}\|_{L^{2}}\lesssim\sum_{j>0}2^{\frac{j}{2}}\|P_{j}f\|_{L^{2}}+2^{-\frac{j}{2}}\|P_{j}f^{\prime}\|_{L^{2}}\lesssim\|f\|_{\dot{B}^{\frac{1}{2}}_{2,1}(\mathbb{R})}.

On the other hand f2f_{2} is given by

f2=f2,lo+f2,hi:=j0Pjf+j>0(1χj)Pjf,f_{2}=f_{2,lo}+f_{2,hi}:=\sum_{j\leq 0}P_{j}f+\sum_{j>0}(1-\chi_{j})P_{j}f,

and we bound its derivative in L1L^{1} by estimating separately the two sums as follows:

f2,loL1[I]j0PjfLj02jPjfLfB˙122,1(),\|f^{\prime}_{2,lo}\|_{L^{1}[I]}\lesssim\sum_{j\leq 0}\|P_{j}f^{\prime}\|_{L^{\infty}}\lesssim\sum_{j\leq 0}2^{j}\|P_{j}f\|_{L^{\infty}}\lesssim\|f\|_{\dot{B}^{\frac{1}{2}}_{2,1}(\mathbb{R})},

respectively, using Hölder’s inequality,

f2,hiL1[I]j>02j((1χj)Pjf)Lj>02j(PjfL+2jPjfL)fB˙122,1()\|f^{\prime}_{2,hi}\|_{L^{1}[I]}\lesssim\sum_{j>0}2^{-j}\|((1-\chi_{j})P_{j}f)^{\prime}\|_{L^{\infty}}\lesssim\sum_{j>0}2^{-j}(\|P_{j}f^{\prime}\|_{L^{\infty}}+2^{j}\|P_{j}f\|_{L^{\infty}})\lesssim\|f\|_{\dot{B}^{\frac{1}{2}}_{2,1}(\mathbb{R})}

which is exactly as needed. ∎

Once we have Lemma 11.5, the conclusion of part (b) of Lemma 11.4 follows directly from the triangle inequality. ∎

Now we get back to the proof of the proposition. Lemma 11.4 allows us to split \mathbb{R} into M2ϵ2\lesssim\frac{M^{2}}{\epsilon^{2}} intervals IjI_{j} with the property that fB˙122,1[Ij]+W˙1,1[Ij]ϵ\|f\|_{\dot{B}^{\frac{1}{2}}_{2,1}[I_{j}]+\dot{W}^{1,1}[I_{j}]}\leq\epsilon (to do so we use that the Besov norm fB˙122,1[Ij]+W˙1,1[I]\|f\|_{\dot{B}^{\frac{1}{2}}_{2,1}[I_{j}]+\dot{W}^{1,1}[I]} depends continuously on the interval II; since we can identify two nearby intervals via scaling and translation, this is a consequence of the fact that the scaling and translation groups are continuous in B˙122,1[]+W˙1,1[]\dot{B}^{\frac{1}{2}}_{2,1}[\mathbb{R}]+\dot{W}^{1,1}[\mathbb{R}]). Then we successively apply the small data result in Proposition 11.1 on each of these subintervals. We arrive at a solution α\alpha which is continuous in time and satisfies

(11.12) α(B˙122,1+W˙1,1)[Ij]ϵ.\|\alpha\|_{(\dot{B}^{\frac{1}{2}}_{2,1}+\dot{W}^{1,1})[I_{j}]}\lesssim\epsilon.

To obtain a global bound from here, we use the triangle type inequality (11.11), and the desired bound (11.8) follows. ∎

12. The final bootstrap

12.1. Proof of Theorem 1.5

Here we bring all the elements together to prove our second main result in this paper, that is lack of finite time blow-up and some control on the growth rate. The idea is the following: we start from the system (10.3), which we recall here:

(12.1) α=12(ie2iα(t)t(Λ1+2Λ2))+14Q(t),(lnλ)=(e2iα(t)tΛ1)+12Q(t).\begin{split}\alpha^{\prime}&=-\frac{1}{2}\Re\left(ie^{-2i\alpha(t)}\partial_{t}(\Lambda^{1}+2\Lambda^{2})\right)+\frac{1}{4}\Re Q(t),\\ (\ln\lambda)^{\prime}&=-\Re(e^{-2i\alpha(t)}\partial_{t}\Lambda^{1})+\frac{1}{2}\Im Q(t).\end{split}

The main inputs in this system are Λi\Lambda^{i} which have estimates in B˙122,1\dot{B}^{\frac{1}{2}}_{2,1} as given by (10.6); the inputs QQ are small in L1L^{1} by (5.10), and easily manageable. Thus we can employ the ode theory developed in the previous section in order to obtain information on α\alpha and λ\lambda, particularly about their rate of growth in time. There is a major conflict in this analysis, which we need to resolve: on one hand, the bounds in (10.6) depend on how large λ(t)\lambda(t) becomes on the time interval II (through the factor λTmax\lambda_{T}^{max} there), while, on the other hand, these bounds are then being used to estimate precisely the growth of λ(t)\lambda(t). Below we show how to resolve this conflict, and prove the growth bound (1.17) for λ\lambda. This in turn shows that λ\lambda cannot reach infinity nor zero in finite time, and thus that the solutions are global in time.

We now turn to the details of the strategy described above. Assume that we work on an time interval I=[0,T][0,Tmax)I=[0,T]\subset[0,T_{max}) where our solution exists; in particular λmaxT=max{λ(t):tI}\lambda^{max}_{T}=\max\{\lambda(t):t\in I\} is well defined. We recall that (10.6) provides a bound on the Besov bound of Λi\Lambda^{i} as follows

ΛiB˙122,1([0,T])ln(2+(λTmax)2T)12ψ0L2,i=1,2.\|\Lambda^{i}\|_{\dot{B}^{\frac{1}{2}}_{2,1}([0,T])}\lesssim\ln(2+(\lambda_{T}^{max})^{2}T)^{\frac{1}{2}}\|\psi_{0}\|_{L^{2}},\quad i=1,2.

while (5.10) shows that

QL1ψ0L22\|Q\|_{L^{1}}\lesssim\|\psi_{0}\|_{L^{2}}^{2}

Then applying the large data ode result in Proposition 11.3 to the system (12.1) we obtain the bound

α(t)α(0)Z+lnλ(t)lnλ(0)Zln(2+(λTmax)2T)ψ0L22.\|\alpha(t)-\alpha(0)\|_{Z}+\|\ln\lambda(t)-\ln\lambda(0)\|_{Z}\lesssim\ln(2+(\lambda_{T}^{max})^{2}T)\|\psi_{0}\|_{L^{2}}^{2}.

For simplicity we normalize λ(0)=1\lambda(0)=1. Then the above bound implies in particular that

lnλTmaxCln(2+(λTmax)2T)ψ02L2Cψ02L2ln(2+T)+Cψ02L2lnλTmax.\ln\lambda_{T}^{max}\leq C\ln(2+(\lambda_{T}^{max})^{2}T)\|\psi_{0}\|^{2}_{L^{2}}\leq C\|\psi_{0}\|^{2}_{L^{2}}\ln(2+T)+C\|\psi_{0}\|^{2}_{L^{2}}\ln\lambda_{T}^{max}.

By taking ψ0L2δ1\|\psi_{0}\|_{L^{2}}\leq\delta\ll 1 small enough, we can absorb the last term and conclude that

lnλTmaxψ02L2ln(2+T),\ln\lambda_{T}^{max}\leq\|\psi_{0}\|^{2}_{L^{2}}\ln(2+T),

which implies the bound

λTmaxeCδ2ln(2+T)=(2+T)Cδ2.\lambda_{T}^{max}\lesssim e^{C\delta^{2}\ln(2+T)}=(2+T)^{C\delta^{2}}.

The bound from below on λ(t)\lambda(t) is much simpler and follows directly from the local well-posedness result in Theorem 1.1. Precisely, if we had |λ(T)2T|1|\lambda(T)^{2}T|\ll 1 then Theorem 1.1 would imply that λ(T)λ(0)\lambda(T)\approx\lambda(0). This cannot be if T1T\gg 1, so we conclude that

λ(T)T12.\lambda(T)\gtrsim\langle T\rangle^{-\frac{1}{2}}.

12.2. Proof of Theorem 1.6

We recall that here we work with data which, at the level of ψ\psi, has smallness in L2L^{2} but also the l1l^{1} Besov bound

ψ(0)LX¯M.\|\psi(0)\|_{{L\bar{X}}}\lesssim M.

By Theorem 1.4, this in turn yields a global bound for the solution, namely

ψl1SM.\|\psi\|_{l^{1}S}\lesssim M.

This last bound allows us to apply the second part of (10.4) in Theorem 10.1, which yields an l1l^{1} Besov bound for Λ1\Lambda^{1}, Λ2\Lambda^{2} in the modulation system (12.1),

Λ1,2B122,1M.\|\Lambda^{1,2}\|_{B^{\frac{1}{2}}_{2,1}}\lesssim M.

On the other hand for QQ we still have the favourable small L1L^{1} bound as in the previous proof.

Now we can directly use Proposition 11.3 for the modulation system (12.1), which yields the global α\alpha bound

αB˙122,1+W˙1,1+lnλB˙122,1+W˙1,1M2.\|\alpha\|_{\dot{B}^{\frac{1}{2}}_{2,1}+\dot{W}^{1,1}}+\|\ln\lambda\|_{\dot{B}^{\frac{1}{2}}_{2,1}+\dot{W}^{1,1}}\lesssim M^{2}.

Finally this implies the desired bound (1.20), completing the proof of the the theorem.

12.3. Proof of Theorem 1.7

Here we establish the stability result, simply by concatenating the required building blocks.

On one hand for ψ\psi we can use Theorem 1.3, which as a corollary of (1.16) yields

ψLLX¯γ,\|\psi\|_{L^{\infty}{L\bar{X}}}\lesssim\gamma,

after which we return to uu via Proposition 6.1 to get

uQα(t),λ(t)LX¯γ.\|u-Q_{\alpha(t),\lambda(t)}\|_{L^{\infty}\bar{X}}\lesssim\gamma.

On the other hand for the modulation parameters we can use the small Besov data result in Proposition 11.1, which yields

αB˙122,1+W˙1,1+lnλB˙122,1+W˙1,1γ.\|\alpha\|_{\dot{B}^{\frac{1}{2}}_{2,1}+\dot{W}^{1,1}}+\|\ln\lambda\|_{\dot{B}^{\frac{1}{2}}_{2,1}+\dot{W}^{1,1}}\lesssim\gamma.

and the difference bound

|lnλ(t)lnλ(0)|+|α(t)α(0)|γ.|\ln\lambda(t)-\ln\lambda(0)|+|\alpha(t)-\alpha(0)|\lesssim\gamma.

Combining the two we obtain the bound (1.22) and conclude the proof of the theorem.

References

  • [1] I. Bejenaru, A. Ionescu, C. Kenig, and D. Tataru, Equivariant Schrödinger maps in two spatial dimensions, Duke Math. J. 162 (2013), no. 11, 1967–2025. MR 3090782
  • [2] I. Bejenaru, A. D. Ionescu, and C. E. Kenig, Global existence and uniqueness of Schrödinger maps in dimensions d4d\geq 4, Adv. Math. 215 (2007), no. 1, 263–291. MR 2354991
  • [3] Ioan Bejenaru, Global results for Schrödinger maps in dimensions n3n\geq 3, Comm. Partial Differential Equations 33 (2008), no. 1-3, 451–477. MR 2398238
  • [4] by same author, On Schrödinger maps, Amer. J. Math. 130 (2008), no. 4, 1033–1065. MR 2427007
  • [5] Ioan Bejenaru, Alexandru D. Ionescu, Carlos E. Kenig, and Daniel Tataru, Global Schrödinger maps in dimensions d2d\geq 2: small data in the critical Sobolev spaces, Ann. of Math. (2) 173 (2011), no. 3, 1443–1506. MR 2800718 (2012g:58048)
  • [6] Ioan Bejenaru and Daniel Tataru, Near soliton evolution for equivariant Schrödinger maps in two spatial dimensions, Mem. Amer. Math. Soc. 228 (2014), no. 1069, vi+108. MR 3157482
  • [7] Nai-Heng Chang, Jalal Shatah, and Karen Uhlenbeck, Schrödinger maps, Comm. Pure Appl. Math. 53 (2000), no. 5, 590–602. MR 1737504
  • [8] Weiyue Ding, On the Schrödinger flows, Proceedings of the International Congress of Mathematicians, Vol. II (Beijing, 2002), Higher Ed. Press, Beijing, 2002, pp. 283–291. MR 1957040
  • [9] Weiyue Ding and Youde Wang, Schrödinger flow of maps into symplectic manifolds, Sci. China Ser. A 41 (1998), no. 7, 746–755. MR 1633799
  • [10] by same author, Local Schrödinger flow into Kähler manifolds, Sci. China Ser. A 44 (2001), no. 11, 1446–1464. MR 1877231
  • [11] Fritz Gesztesy and Maxim Zinchenko, On spectral theory for Schrödinger operators with strongly singular potentials, Math. Nachr. 279 (2006), no. 9-10, 1041–1082. MR 2242965
  • [12] S. Gustafson, K. Kang, and T.-P. Tsai, Schrödinger flow near harmonic maps, Comm. Pure Appl. Math. 60 (2007), no. 4, 463–499. MR 2290708
  • [13] Stephen Gustafson, Kyungkeun Kang, and Tai-Peng Tsai, Asymptotic stability of harmonic maps under the Schrödinger flow, Duke Math. J. 145 (2008), no. 3, 537–583. MR 2462113
  • [14] Stephen Gustafson, Kenji Nakanishi, and Tai-Peng Tsai, Asymptotic stability, concentration, and oscillation in harmonic map heat-flow, Landau-Lifshitz, and Schrödinger maps on 𝐑2\mathbf{R}^{2}, Comm. Math. Phys. 300 (2010), no. 1, 205–242. MR 2725187
  • [15] Alexandru D. Ionescu and Carlos E. Kenig, Low-regularity Schrödinger maps, Differential Integral Equations 19 (2006), no. 11, 1271–1300. MR 2278007
  • [16] by same author, Low-regularity Schrödinger maps. II. Global well-posedness in dimensions d3d\geq 3, Comm. Math. Phys. 271 (2007), no. 2, 523–559. MR 2287916
  • [17] Jun Kato and Herbert Koch, Uniqueness of the modified Schrödinger map in H3/4+ϵ(𝐑2)H^{3/4+\epsilon}(\mathbf{R}^{2}), Comm. Partial Differential Equations 32 (2007), no. 1-3, 415–429. MR 2304155
  • [18] Carlos Kenig, Tobias Lamm, Daniel Pollack, Gigliola Staffilani, and Tatiana Toro, The Cauchy problem for Schrödinger flows into Kähler manifolds, Discrete Contin. Dyn. Syst. 27 (2010), no. 2, 389–439. MR 2600676
  • [19] Carlos E. Kenig and Andrea R. Nahmod, The Cauchy problem for the hyperbolic-elliptic Ishimori system and Schrödinger maps, Nonlinearity 18 (2005), no. 5, 1987–2009. MR 2164729
  • [20] Sahbi Keraani, On the defect of compactness for the Strichartz estimates of the Schrödinger equations, J. Differential Equations 175 (2001), no. 2, 353–392. MR 1855973
  • [21] J. Krieger, W. Schlag, and D. Tataru, Renormalization and blow up for charge one equivariant critical wave maps, Invent. Math. 171 (2008), no. 3, 543–615. MR 2372807
  • [22] by same author, Renormalization and blow up for the critical Yang-Mills problem, Adv. Math. 221 (2009), no. 5, 1445–1521. MR 2522426
  • [23] Andrew Lawrie, Jonas Lührmann, Sung-Jin Oh, and Sohrab Shahshahani, Asymptotic stability of harmonic maps on the hyperbolic plane under the Schrödinger maps evolution, Comm. Pure Appl. Math. 76 (2023), no. 3, 453–584. MR 4544804
  • [24] Ze Li, Global Schrödinger map flows to Kähler manifolds with small data in critical Sobolev spaces: high dimensions, J. Funct. Anal. 281 (2021), no. 6, Paper No. 109093, 76. MR 4257981
  • [25] by same author, Global Schrödinger map flows to Kähler manifolds with small data in critical Sobolev spaces: energy critical case, J. Eur. Math. Soc. (JEMS) 25 (2023), no. 12, 4879–4969. MR 4662306
  • [26] Jeremy Marzuola, Jason Metcalfe, and Daniel Tataru, Strichartz estimates and local smoothing estimates for asymptotically flat Schrödinger equations, J. Funct. Anal. 255 (2008), no. 6, 1497–1553. MR 2565717
  • [27] Helena McGahagan, An approximation scheme for Schrödinger maps, Comm. Partial Differential Equations 32 (2007), no. 1-3, 375–400. MR 2304153
  • [28] Frank Merle, Pierre Raphaël, and Igor Rodnianski, Blowup dynamics for smooth data equivariant solutions to the critical Schrödinger map problem, Invent. Math. 193 (2013), no. 2, 249–365. MR 3090180
  • [29] Andrea Nahmod, Jalal Shatah, Luis Vega, and Chongchun Zeng, Schrödinger maps and their associated frame systems, Int. Math. Res. Not. IMRN (2007), no. 21, Art. ID rnm088, 29. MR 2352219
  • [30] Andrea Nahmod, Atanas Stefanov, and Karen Uhlenbeck, On Schrödinger maps, Comm. Pure Appl. Math. 56 (2003), no. 1, 114–151. MR 1929444
  • [31] by same author, Erratum: “On Schrödinger maps” [Comm. Pure Appl. Math. 56 (2003), no. 1, 114–151; mr1929444], Comm. Pure Appl. Math. 57 (2004), no. 6, 833–839. MR 2038118
  • [32] Galina Perelman, Blow up dynamics for equivariant critical Schrödinger maps, Comm. Math. Phys. 330 (2014), no. 1, 69–105. MR 3215578
  • [33] I.N. Sneddon, Fourier transforms, Dover books on mathematics, Dover Publications, 1995.
  • [34] P.-L. Sulem, C. Sulem, and C. Bardos, On the continuous limit for a system of classical spins, Comm. Math. Phys. 107 (1986), no. 3, 431–454. MR 866199
  • [35] Daniel Tataru, Parametrices and dispersive estimates for Schrödinger operators with variable coefficients, Amer. J. Math. 130 (2008), no. 3, 571–634. MR 2418923
  • [36] G.N. Watson, A treatise on the theory of bessel functions, no. v. 1, Cambridge University Press, 1922.