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

Scaling-critical well-posedness for
continuum Calogero–Moser models

Rowan Killip Department of Mathematics, University of California, Los Angeles, CA 90095, USA [email protected] Thierry Laurens Department of Mathematics, University of Wisconsin–Madison, WI, 53706, USA [email protected]  and  Monica Vişan Department of Mathematics, University of California, Los Angeles, CA 90095, USA [email protected]
Abstract.

We prove that the focusing and defocusing continuum Calogero–Moser models are well-posed in the scaling-critical space L+2()L^{2}_{+}(\mathbb{R}). In the focusing case, this requires solutions to have mass less than that of the soliton.

1. Introduction

The goal of this paper is to address the well-posedness theory of the following two dispersive equations:

(CCM) iddtq=q′′±2iqC+(|q|2).i\tfrac{d}{dt}q=-q^{\prime\prime}\pm 2iqC_{+}\big{(}|q|^{2}\big{)}^{\prime}.

As we will explain, a plus sign in front of the nonlinearity corresponds to a focusing equation; a minus sign yields a defocusing model. Throughout the paper, the ±\pm and \mp signs will be reserved for this dichotomy: the upper sign will correspond to the focusing case and the lower sign to the defocusing case.

The unknown field q(t,x)q(t,x) appearing in (CCM) is a complex-valued function of xx\in\mathbb{R}. We further demand that q(t)q(t) belongs to the Hardy space

L+2():={fL2():f^(ξ)=0 for ξ<0}.L^{2}_{+}(\mathbb{R}):=\{f\in L^{2}(\mathbb{R}):\,\widehat{f}(\xi)=0\text{ for }\xi<0\}.

The operator C+C_{+} appearing in the nonlinearity of (CCM) denotes the Cauchy–Szegő projection from L2()L^{2}(\mathbb{R}) onto L+2()L^{2}_{+}(\mathbb{R}); see (1.17) and (1.19).

The defocusing (CCM) appears first as a special case of the ‘intermediate nonlinear Schrödinger equation’ introduced in [Pelinovsky1995]. Concretely, this equation is derived as a model for the envelope of an approximately monochromatic wave packet at the interface between otherwise quiescent fluid layers of infinite total depth, that is, it provides a modulation theory for the setting of the Benjamin–Ono equation. In addition to deriving a model for the intermediate depth case, Pelinovsky also discovered the important nonlocal structure of the nonlinearity, which had been overlooked in the prior work [Tanaka] on the infinite-depth problem.

Our interest in (CCM) was first sparked by the recent work [GerardLenzmann] centered on the focusing case. This model was formally derived in [Abanov2009] as a continuum limit of the famous Calogero–Moser particle system [Calogero1, Calogero2, Moser]:

(CM) d2xjdt2=kj1(xjxk)3.\displaystyle\frac{d^{2}x_{j}}{dt^{2}}=\sum_{k\neq j}\frac{1}{(x_{j}-x_{k})^{3}}.

This discrete completely integrable system describes the dynamics of a gas of particles interacting pairwise through an inverse square potential. It was subsequently discovered that the system remains completely integrable for interaction potentials derived from the Weierstrass \wp function. Included in this class is the cosecant-squared potential, which arises naturally in the periodic setting. This latter model was first championed by Sutherland [Sutherland1, Sutherland2] in the quantum mechanical setting. Correspondingly, the periodic analogue of (CCM) has been studied under the name Calogero–Sutherland derivative NLS; see [Badreddine2023].

Whether derived as a modulation theory for the Benjamin–Ono equation or as a continuum limit of the Calogero–Moser system, there can be no surprise that the (CCM) equations are completely integrable. This fact will ultimately play a key role in what follows.

The most physically intuitive conserved quantities are

M(q):=|q|2𝑑xandP(q):=iq¯q12|q|4dx,\displaystyle M(q):=\int|q|^{2}\,dx\quad\text{and}\quad P(q):=\int-i\overline{q}q^{\prime}\mp\tfrac{1}{2}|q|^{4}\,dx,

which represent mass and momentum, together with the Hamiltonians

(1.1) H±(q)\displaystyle H_{\pm}(q) :=12|qiqC+(|q|2)|2𝑑x.\displaystyle:=\frac{1}{2}\int_{\mathbb{R}}\bigl{|}q^{\prime}\mp iqC_{+}(|q|^{2})\bigr{|}^{2}\,dx.

We caution the reader that these identifications rely on the non-standard Poisson structure naturally associated to (CCM); see [GerardLenzmann]. This exotic structure originates from the use of a gauge transformation to simplify the underlying phase space.

The (CCM) equations are mass-critical: both the class of solutions and the conserved mass M(q)M(q) are invariant under the scaling

q(t,x)qλ(t,x):=λq(λ2t,λx)for λ>0.q(t,x)\mapsto q_{\lambda}(t,x):=\sqrt{\lambda}q\bigl{(}\lambda^{2}t,\lambda x\bigr{)}\quad\text{for $\lambda>0$}.

These considerations make L+2()L^{2}_{+}(\mathbb{R}) the most natural space in which to address the well-posedness problem. In this sense, we are able to give a complete solution to the defocusing problem:

Theorem 1.1 (Defocusing case).

The defocusing (CCM) equation is globally well-posed in the space L+2()L^{2}_{+}(\mathbb{R}). Moreover, if QL+2()Q\subset L^{2}_{+}(\mathbb{R}) is a bounded and equicontinuous set, then the set of orbits

(1.2) Q:={etJHq:qQ,t}Q^{*}:=\{e^{tJ\nabla H_{-}}q:q\in Q,\ t\in\mathbb{R}\}

is bounded and equicontinuous in L+2()L^{2}_{+}(\mathbb{R}).

The analogous theorem for periodic initial data was proved recently by Badreddine [Badreddine2023], whose influence on our work will be discussed in due course. One point in which we must diverge is in the treatment of equicontinuity.

Definition 1.2 (Equicontinuity).

Fix ss\in\mathbb{R}. A bounded set QH+s()Q\subset H^{s}_{+}(\mathbb{R}) is said to be equicontinuous in the Hs()H^{s}(\mathbb{R}) topology if

lim supδ0supqQsup|y|<δq(+y)q()Hs()=0.\limsup_{\delta\to 0}\ \sup_{q\in Q}\ \sup_{|y|<\delta}\|q(\cdot+y)-q(\cdot)\|_{H^{s}(\mathbb{R})}=0.

Replacing the Hs()H^{s}(\mathbb{R}) norm here by the supremum norm would yield the familiar notion appearing in the Arzelà–Ascoli Theorem. Indeed, boundedness, tightness, and equicontinuity constitute necessary and sufficient conditions for a set to be Hs()H^{s}(\mathbb{R})-precompact. Equicontinuity may also be understood via Plancherel’s Theorem: a bounded set QHs()Q\subset H^{s}(\mathbb{R}) is equicontinuous if and only if

(1.3) limκsupqQ|ξ|κ|q^(ξ)|2(|ξ|+1)2s𝑑ξ=0,\lim_{\kappa\to\infty}\ \sup_{q\in Q}\int_{|\xi|\geq\kappa}|\widehat{q}(\xi)|^{2}(|\xi|+1)^{2s}\,d\xi=0,

which is to say, the family of Fourier transforms is tight.

Equicontinuity plays a central role in the treatment of scaling-critical problems. It is precisely this property that prevents physical-space concentration and so blowup. Due to its scaling-critical nature, the conservation of mass is completely insensitive to any changes of scale. On the other hand, control of higher regularity explicitly ensures equicontinuity. Such a higher regularity theory has been developed in [GerardLenzmann] in a manner applicable to both the focusing and defocusing models; see, for example, Theorem 1.3 below. To properly describe this, we must first discuss some startling phenomenology discovered in [GerardLenzmann] regarding the focusing model.

The focusing (CCM) equation admits soliton solutions:

(1.4) q(t,x)=λR(λx+x0)withλ>0,x0,andR(x):=2x+i.q(t,x)=\sqrt{\lambda}R(\lambda x+x_{0})\quad\text{with}\quad\lambda>0,\quad x_{0}\in\mathbb{R},\quad\text{and}\quad R(x):=\tfrac{\sqrt{2}}{x+i}.

These (stationary) solitons all have the same mass, namely, M(R)=2πM(R)=2\pi. Moreover, the focusing (CCM) admits multisoliton solutions; these have been analyzed in detail in [GerardLenzmann].

The multisoliton solutions are rational functions of xx at every moment of time. Their poles and residues both evolve in time, with the former satisfying a complex Calogero–Moser system. Such a solution is called an NN-soliton if there are NN poles. Their mass is quantized; it is 2πN2\pi N, that is, NN times the mass of the one-soliton RR. The most striking discovery, however, is that for N2N\geq 2, these NN-solitons exhibit frequency cascades as t±t\to\pm\infty. As a consequence, for every s>0s>0, the Hs()H^{s}(\mathbb{R}) of these smooth solutions is unbounded as t±t\to\pm\infty.

By contrast, [GerardLenzmann] shows that smooth solutions with mass below that of the one-soliton do not exhibit such norm growth:

Theorem 1.3 ([GerardLenzmann]).

Let n1n\geq 1 be an integer. The focusing (CCM) is locally well-posed in H+n()H^{n}_{+}(\mathbb{R}). Furthermore, it is globally well-posed on {qH+n():M(q)<2π}\{q\in H^{n}_{+}(\mathbb{R}):M(q)<2\pi\} and these solutions satisfy

(1.5) suptq(t)H+n()<.\sup_{t\in\mathbb{R}}\|q(t)\|_{H^{n}_{+}(\mathbb{R})}<\infty.

The defocusing equation is globally well-posed in H+n()H^{n}_{+}(\mathbb{R}), without any mass constraint, and all solutions also satisfy (1.5).

Strictly speaking, the paper [GerardLenzmann] does not discuss well-posedness in the defocusing case; nevertheless, applying the arguments described there yields the results stated above. The paper also demonstrates local well-posedness for regularities s>12s>\tfrac{1}{2} and shows that smooth initial data with mass exactly equal to 2π2\pi leads to a global smooth solution.

The specific number 2π2\pi in Theorem 1.3 originates from the sharp constant in the inequality (2.1). This threshold coincides with the mass of the soliton, precisely because RR is an optimizer for the inequality (2.1).

When considering initial data from H+()H^{\infty}_{+}(\mathbb{R}), Theorem 1.3 allows us to discuss the solution to (CCM). The first question we ask about such solutions is one of equicontinuity. The explicit form of the two-soliton solutions presented in [GerardLenzmann] shows that they do not have L2L^{2}-equicontinuous orbits. Thus, the global-in-time equicontinuity property of Theorem 1.1 does not extend to the focusing case. On the other hand, the results presented in Theorem 1.3 strongly suggest that global-in-time equicontinuity may hold under a mass constraint, even though the arguments used in [GerardLenzmann] do not yield a proof of this. This is our first result on the focusing problem:

Theorem 1.4 (Equicontinuity for the focusing (CCM)).

If QH+()Q\subset H^{\infty}_{+}(\mathbb{R}) satisfies

(1.6) supqQqL22<2π\sup_{q\in Q}\left\lVert q\right\rVert_{L^{2}}^{2}<2\pi

and is equicontinuous in L2()L^{2}(\mathbb{R}), then the set of orbits

(1.7) Q:={etJH+q:qQ,t}Q^{*}:=\{e^{tJ\nabla H_{+}}q:q\in Q,\ t\in\mathbb{R}\}

is bounded and equicontinuous in L2()L^{2}(\mathbb{R}).

Recall that the two-soliton solutions demonstrate that this conclusion would fail if the number 2π2\pi in (1.6) were replaced by any number larger than 4π4\pi. At this moment, we are not ready to a make a conjecture about the true mass threshold for global-in-time equicontinuity.

As noted above, control on the frequency distribution in a scaling-critical space is an essential ingredient for proving global well-posedness because it precludes space concentration. For this purpose, it is not necessary that orbits remain equicontinuous globally in time. It would suffice to know that equicontinuous sets of smooth initial data lead to families of orbits that are equicontinuous on compact time intervals. This line of reasoning leads to our introduction of the following mass threshold:

Definition 1.5.

Let M[0,]M^{*}\in[0,\infty] denote the maximal constant so that for any L2L^{2} bounded and equicontinuous set QH+()Q\subset H^{\infty}_{+}(\mathbb{R}) satisfying

supqQM(q)<M,\sup_{q\in Q}M(q)<M^{*},

the set of partial orbits

(1.8) QT:={etJH±q:qQ,|t|<T}Q^{*}_{T}:=\{e^{tJ\nabla H_{\pm}}q:q\in Q,\ |t|<T\}

is L2L^{2}-equicontinuous for each choice of T>0T>0.

We are not implicitly assuming here that all choices of smooth initial data lead to a global solution. In the focusing case, this is an open problem. Rather, we will show that finite-time blowup must be accompanied by a loss of equicontinuity; see Lemma 2.4 for details.

In the defocusing case, Theorem 1.1 shows that M=M^{*}=\infty. Indeed, we will first prove that M=M^{*}=\infty in Section 3 and then later use this as an ingredient in proving well-posedness in Section 5.

In the focusing case, Theorem 1.4 guarantees M2πM^{*}\geq 2\pi. The analysis of multisoliton solutions in [GerardLenzmann] places no restrictions on MM^{*}; indeed, mass moves to high frequencies in a manner that is linear in time.

The main rationale for introducing the equicontinuity threshold MM^{*} is that it sets the proper generality for our well-posedness analysis of the focusing problem. It will also allow us to treat the focusing and defocusing problems in a parallel manner.

Theorem 1.6 (Global well-posedness in the focusing case).

The focusing (CCM) equation is globally well-posed in the space

BM={qL+2():qL22<M}.B_{M^{*}}=\{q\in L^{2}_{+}(\mathbb{R}):\left\lVert q\right\rVert_{L^{2}}^{2}<M^{*}\}.

The analogous result in the periodic setting was obtained in [Badreddine2023]. The relation between the arguments used here and in [Badreddine2023] will be elaborated below.

In Section 6, we will show how to deduce well-posedness at higher regularities from our scaling-critical results. This rests on a new technique for proving HsH^{s}-equicontinuity that may be of independent interest.

As mentioned earlier, the complete integrability of (CCM) will play a central role in our analysis. This integrability will manifest in two ways: through the Lax pair and an explicit formula (in the sense of Gérard).

The Lax pair we employ is the following:

(1.9) =iqC+q¯and𝒫=i2±2qC+q¯.\mathcal{L}=-i\partial\mp qC_{+}\overline{q}\quad\text{and}\quad\mathcal{P}=i\partial^{2}\pm 2q\partial C_{+}\overline{q}.

A Lax pair formulation of the full intermediate NLS was introduced in [PelinovskyG1995]; recall that one may recover the defocusing (CCM) as an infinite depth limit of this model. The Lax pair (1.9) is a small modification of the one presented in [GerardLenzmann], which uses

(1.10) 𝒫~:=𝒫+iq2=±qC+q¯qC+q¯+iqC+|q|2C+q¯.\widetilde{\mathcal{P}}:=\mathcal{P}+i\mathcal{L}_{q}^{2}=\pm qC_{+}\overline{q}^{\prime}\mp q^{\prime}C_{+}\overline{q}+iqC_{+}|q|^{2}C_{+}\overline{q}.

Evidently, the change to 𝒫\mathcal{P} has no effect on the commutator relation. In choosing to center our analysis on (1.9), we were very much informed by our prior work [Killip2023] on the Benjamin–Ono equation. Concretely, we favor 𝒫\mathcal{P} because it enjoys two advantageous properties : First, (CCM) may be written as

(1.11) ddtq=𝒫q,\tfrac{d}{dt}q=\mathcal{P}q,

and secondly, 𝒫1=0\mathcal{P}1=0. As the constant function 11 lies outside the natural Hilbert space setting, this latter relation requires some further interpretation; see (4.3) for a precise statement.

The second manifestation of complete integrability is an explicit formula of a type championed by Gérard and collaborators [MR3301889, MR4662323, Gerard2023Szego]:

Theorem 1.7.

Let q(t)q(t) denote the global solution to (CCM) with initial data q0L+2()q^{0}\in L^{2}_{+}(\mathbb{R}) satisfying M(q0)<MM(q^{0})<M^{*}. Then

(1.12) q(t,z)=12πiI+{(X+2tq0z)1q0}q(t,z)=\tfrac{1}{2\pi i}I_{+}\big{\{}(X+2t\mathcal{L}_{q^{0}}-z)^{-1}q^{0}\big{\}}

for all zz with Imz>0\operatorname{Im}z>0. Here q(t,z)q(t,z) is defined via harmonic extension; see (1.14).

For the full definitions of the linear functional I+I_{+} and of the operator XX, see Section 4. Naively speaking, I+I_{+} represents integration over the whole line, while XX generalizes the operator of multiplication by xx. Note that I+I_{+} is not a bounded linear functional on L+2()L^{2}_{+}(\mathbb{R}), nor can multiplication by xx be interpreted as a self-adjoint operator in Hardy space.

It is worth recalling that an explicit form of the solution to the Calogero–Moser system (CM) was given long ago in [MR455039]. Concretely, it was shown that one may describe particles evolving according to this system as the eigenvalues of a straight line trajectory in the space of symmetric matrices. This is evidently not a very close analogue of (1.12), which speaks directly to the subtlety of the manner in which (CCM) arises as a continuum limit of the particle system.

The analogue of (1.12) appropriate to the periodic (CCM) was introduced in [Badreddine2023] and played a central role in that analysis.

In Theorem 4.1 we show that the explicit formula (1.12) holds for solutions with smooth well-decaying initial data. This is the form that we will employ in demonstrating scaling-critical well-posedness. Once this is achieved, we may then extend this formula to all L+2()L^{2}_{+}(\mathbb{R}) solutions by using the resulting continuity of the data-to-solution map in concert with the continuity of the right-hand side that will be demonstrated using the tools developed as a part of our analysis.

Overview of the proofs

The central theme of this paper is to demonstrate how to prove well-posedness by synthesizing explicit representations of the type (1.12) with the tools and techniques developed as part of the method of commuting flows [MR4304314, harropgriffiths2022sharp, harropgriffiths2023global, Killip2023, MR4628747, MR3990604, MR3820439]. In particular, the methods we employ to prove equicontinuty draw from these earlier works. On the other hand, we will employ no regularized flows or commuting Hamiltonians in this paper.

The question central to L+2()L^{2}_{+}(\mathbb{R}) well-posedness is this: Does any sequence of smooth and well-decaying initial data qn0q^{0}_{n} that converges in L+2()L^{2}_{+}(\mathbb{R}) lead to a sequence of solutions that converges in Ct(I;L+2())C_{t}(I;L^{2}_{+}(\mathbb{R})) on any compact time interval II?

By the conservation of M(q)M(q), such sequences of solutions will always admit subsequential limits in the weak topology (pointwise in time). A first step forward is to show that there is a unique such subsequential limit and, correspondingly, one has weak convergence without passing to a subsequence. This is one of the roles played by the explicit formula in [Badreddine2023] and in this paper; see Corollary 5.2.

To upgrade weak convergence to strong convergence one must preclude a loss of mass. In the periodic setting of [Badreddine2023], this is a question of equicontinuity. On the line, one must also demonstrate tightness.

Our route to controlling the equicontinuity properties of solutions is through the spectral theory of the Lax operator \mathcal{L}. Recall that (formally at least) the Lax equation ensures that q(t)\mathcal{L}_{q(t)} and q(0)\mathcal{L}_{q(0)} remain unitarily equivalent. In the periodic setting, the Lax operator has compact resolvent and so its spectral properties are encoded through the associated sequence of eigenvalues and their eigenvectors. In [Badreddine2023], loss of mass is precluded by demonstrating continuity properties of these eigenvalues/vectors.

In the whole line setting, the spectral theory is more complicated; indeed, even the spectral type is unknown at this time. In Section 2, we construct the Lax operator as a self-adjoint operator for general qL+2()q\in L^{2}_{+}(\mathbb{R}). This involves an improvement on the earlier analysis of [GerardLenzmann]: In Lemma 2.1, we prove that q\mathcal{L}_{q} is a relatively compact perturbation of the case q0q\equiv 0, rather than an infinitesimally form bounded perturbation as in [GerardLenzmann]. The advantages of this small improvement accumulate as we progress.

In Lemmas 2.6 and  2.7, we show that the difference between the resolvents

R(κ,q(t))=(q(t)+κ)1andR0(κ):=(0+κ)1\displaystyle R(\kappa,q(t))=(\mathcal{L}_{q(t)}+\kappa)^{-1}\quad\text{and}\quad R_{0}(\kappa):=(\mathcal{L}_{0}+\kappa)^{-1}

is trace class; moreover, the trace is a conserved quantity. Consequently,

(1.13) β(κ,q):=M(q)2πκtr{R(κ)R0(κ)}\beta(\kappa,q):=M(q)\mp 2\pi\kappa\operatorname{tr}\{R(\kappa)-R_{0}(\kappa)\}

is also conserved. In Section 3, we use β(κ,q)\beta(\kappa,q) to prove our equicontinuity results.

Lemma 3.1 describes the connections between equicontinuity and the large-κ\kappa behaviour of β(κ,q)\beta(\kappa,q). We see that β(κ,q)\beta(\kappa,q) converges to zero uniformly on equicontinuous sets. However, to prove equicontinuity, we must show convergence of the quadratic part of β(κ,q)\beta(\kappa,q). By virtue of a favorable sign of the higher-order terms in β(κ,q)\beta(\kappa,q), this approach quickly yields equicontinuity in the defocusing case.

In the focusing case, the sign is unfavorable! Nevertheless, by exploiting the inequality (2.1) of [GerardLenzmann] in concert with certain operator analysis (which appears to be novel), we are able to subordinate the higher-order terms in β(κ,q)\beta(\kappa,q) to the quadratic term and so deduce equicontinuity up to the 2π2\pi threshold.

The climax of Section 4 is the verification of the explicit formula (1.12) for smooth and well-decaying initial data; see Theorem 4.1. Much of this section is devoted to developing the relevant operator theory in a manner that will allow us to consider L+2L^{2}_{+} limits of such solutions later.

In Section 5, we complete the proof of well-posedness in L+2()L^{2}_{+}(\mathbb{R}). Recall that our goal is to show that L+2()L^{2}_{+}(\mathbb{R})-Cauchy sequences of smooth well-decaying initial data produce solutions that converge in CtLx2C_{t}L^{2}_{x} on compact time intervals. At this moment in the argument, we know weak convergence at each moment of time. We upgrade this by demonstrating compactness. This requires three inputs: equicontinuity and tightness of orbits as functions of xx and thirdly, L+2()L^{2}_{+}(\mathbb{R})-equicontinuity in time.

Spatial equicontinuity is settled already in Section 3. As we will see, this equicontinuity is then used to aid in verifying the other two requirements. See Lemma 5.4 for the treatment of equicontinuity in time and Proposition 5.6 for tightness. To prove the latter, we rely again on the explicit formula.

Section 5 concludes with the proof of Theorem 1.7. This combines the well-posedness proved earlier in the section with the operator theory developed in Section 4.

Finally, in Section 6, we extend well-posedness to H+s()H^{s}_{+}(\mathbb{R}) for 0s<10\leq s<1. The cornerstone of this extension is the demonstration of equicontinuity of orbits in H+s()H^{s}_{+}(\mathbb{R}). Breaking from previous works, we introduce a method based on Loewner’s Theorem on operator monotone functions (see the monograph [MR3969971]). We believe this represents an elegant and efficient approach to this question that will prove useful in treating other completely integrable systems.

Acknowledgements

R.K. was supported by NSF grant DMS-2154022; M.V. was supported by NSF grant DMS-2054194. The work of T.L. was also supported by these grants.

1.1. Notation

Throughout this paper, we employ the standard notation ABA\lesssim B to indicate that ACBA\leq CB for some constant C>0C>0; if ABA\lesssim B and BAB\lesssim A, we write ABA\approx B. Occasionally, we adjoin subscripts to this notation to indicate dependence of the constant CC on other parameters; for instance, we write Aα,βBA\lesssim_{\alpha,\beta}B when ACBA\leq CB for some constant C>0C>0 depending on α,β\alpha,\beta.

Hardy spaces will be denoted L+p=L+p()L^{p}_{+}=L^{p}_{+}(\mathbb{R}). These are the (closed) subspaces of Lp()L^{p}(\mathbb{R}) comprised of those functions ff whose Poisson integral

(1.14) f(z):=Imzπ|xz|2f(x)𝑑xdefined forImz>0\displaystyle f(z):=\int\frac{\operatorname{Im}z}{\pi|x-z|^{2}}f(x)\,dx\quad\text{defined for}\quad\operatorname{Im}z>0

is holomorphic (in the upper half-plane). By Hölder’s inequality,

(1.15) |f(z)|(Imz)1pfLp()\displaystyle\bigl{|}f(z)\bigr{|}\lesssim(\operatorname{Im}z)^{-\frac{1}{p}}\|f\|_{L^{p}(\mathbb{R})}

for any 1p1\leq p\leq\infty.

For ss\in\mathbb{R} we define the Sobolev spaces Hs()H^{s}(\mathbb{R}) as the completion of 𝒮()\mathcal{S}(\mathbb{R}) with respect to the norm

fHs()2=(|ξ|+1)2s|f^(ξ)|2𝑑ξ.\left\lVert f\right\rVert_{H^{s}(\mathbb{R})}^{2}=\int(|\xi|+1)^{2s}|\widehat{f}(\xi)|^{2}\,d\xi.

The Hardy–Sobolev spaces H+s()H^{s}_{+}(\mathbb{R}) comprise those functions in Hs()H^{s}(\mathbb{R}) whose Fourier transform is supported on [0,)[0,\infty).

One important property of the Hardy–Sobolev spaces that is not enjoyed by the pure Sobolev spaces is that products are well defined even at negative regularity. For example,

(1.16) fgH+5()fH+2()gH+2().\left\lVert fg\right\rVert_{H^{-5}_{+}(\mathbb{R})}\lesssim\left\lVert f\right\rVert_{H^{-2}_{+}(\mathbb{R})}\left\lVert g\right\rVert_{H^{-2}_{+}(\mathbb{R})}.

The proof is elementary; see [Killip2023, Lem. 2.2] for details.

Our convention for the Fourier transform is

(1.17) f^(ξ)=12πeiξxf(x)𝑑xso that f(x)=12πeiξxf^(ξ)𝑑ξ.\displaystyle\widehat{f}(\xi)=\tfrac{1}{\sqrt{2\pi}}\int_{\mathbb{R}}e^{-i\xi x}f(x)\,dx\quad\text{so that }\quad f(x)=\tfrac{1}{\sqrt{2\pi}}\int_{\mathbb{R}}e^{i\xi x}\widehat{f}(\xi)\,d\xi.

With this definition,

(1.18) fL2()=f^L2()andfg^(ξ)=12πf^(ξη)g^(η)𝑑η.\displaystyle\|f\|_{L^{2}(\mathbb{R})}=\|\widehat{f}\|_{L^{2}(\mathbb{R})}\quad\text{and}\quad\widehat{fg}(\xi)=\tfrac{1}{\sqrt{2\pi}}\int\widehat{f}(\xi-\eta)\widehat{g}(\eta)\,d\eta.

We use the following notation for the Cauchy–Szegő projections

(1.19) C±f^(ξ):=1[0,)(±ξ)f^(ξ).\widehat{C_{\pm}f}(\xi):=1_{[0,\infty)}(\pm\xi)\widehat{f}(\xi).

We will employ Littlewood–Paley decompositions with frequency parameters N2N\in 2^{\mathbb{Z}}. For a smooth, non-negative function φ\varphi supported on |ξ|2|\xi|\leq 2 with φ(ξ)=1\varphi(\xi)=1 for |ξ|1|\xi|\leq 1, we define PNP_{\leq N} as the Fourier multiplier with symbol φ(ξ/N)\varphi(\xi/N) and then PN=PNPN/2P_{N}=P_{\leq N}-P_{\leq N/2}. Observe that

1=N2PN.1=\sum_{N\in 2^{\mathbb{Z}}}P_{N}.

We will often adopt the more compact notations fN=PNff_{N}=P_{N}f , fN=PNff_{\leq N}=P_{\leq N}f, and f>N=[1PN]ff_{>N}=[1-P_{\leq N}]f.

Throughout the paper, we write op\|\cdot\|_{\mathrm{op}} to denote the norm of an operator acting on the Hilbert space L+2()L^{2}_{+}(\mathbb{R}). Similarly, we reserve the notations 1\mathfrak{I}_{1} and 2\mathfrak{I}_{2} to denote the trace and Hilbert–Schmidt classes over this same Hilbert space. For further information on such trace ideals, we recommend the book [TraceIdeals].

2. The Lax operator

We write 0=||=i\mathcal{L}_{0}=|\partial|=-i\partial. For κ>0\kappa>0, the resolvent associated with 0\mathcal{L}_{0} is

R0(κ):=(0+κ)1.R_{0}(\kappa):=(\mathcal{L}_{0}+\kappa)^{-1}.

In the focusing case, the 2π2\pi mass threshold marks an important transition in the spectral theory of the Lax operator q\mathcal{L}_{q} presented in (1.9). This particular number originates from the sharpness of the constant in the inequality

(2.1) C+(q¯f)L212πqL2||1/2fL2,\left\lVert C_{+}(\overline{q}f)\right\rVert_{L^{2}}\leq\tfrac{1}{\sqrt{2\pi}}\left\lVert q\right\rVert_{L^{2}}\big{\lVert}|\partial|^{1/2}f\big{\rVert}_{L^{2}},

demonstrated already in [GerardLenzmann].

The inequality (2.1) was used in [GerardLenzmann] to construct the operator q\mathcal{L}_{q} as an infinitesimally form-bounded perturbation of 0\mathcal{L}_{0}. This also allowed them to identify the form domain of q\mathcal{L}_{q}. Here we will prove the stronger result that q\mathcal{L}_{q} is a relatively compact perturbation of 0\mathcal{L}_{0}; this allows us to also identify the domain and the essential spectrum of q\mathcal{L}_{q}. The key input is the following:

Lemma 2.1.

For any κ>0\kappa>0 and g,qL+2()g,q\in L^{2}_{+}(\mathbb{R}), the operator

(2.2) fgC+[q¯R0(κ)f]is compact andgC+q¯R0(κ)opgL2qL2.f\mapsto gC_{+}\bigl{[}\overline{q}R_{0}(\kappa)f\bigr{]}\quad\text{is compact and}\quad\bigl{\|}gC_{+}\overline{q}R_{0}(\kappa)\bigr{\|}_{\mathrm{op}}\lesssim\|g\|_{L^{2}}\|q\|_{L^{2}}.

Moreover, if QL+2()Q\subseteq L^{2}_{+}(\mathbb{R}) is bounded and equicontinuous, then

(2.3) supqQC+q¯R0(κ)op2supqQqC+q¯R0(κ)op0asκ.\displaystyle\sup_{q\in Q}\,\big{\lVert}C_{+}\overline{q}\sqrt{R_{0}(\kappa)}\big{\rVert}_{\mathrm{op}}^{2}\leq\sup_{q\in Q}\,\bigl{\|}qC_{+}\overline{q}R_{0}(\kappa)\bigr{\|}_{\mathrm{op}}\to 0\quad\text{as}\quad\kappa\to\infty.
Proof.

Given ξ>0\xi>0, we may use (1.18) to see that

|q¯R0f^(ξ)|12πξ|q^(ηξ)||f^(η)|dηη,\displaystyle\bigl{|}\widehat{\bar{q}R_{0}f}(\xi)\bigr{|}\leq\tfrac{1}{\sqrt{2\pi}}\int_{\xi}^{\infty}\bigl{|}\widehat{q}(\eta-\xi)\bigr{|}\bigl{|}\widehat{f}(\eta)\bigr{|}\,\frac{d\eta}{\eta},

for any κ>0\kappa>0, and consequently,

(2.4) h|q¯R0f^(ξ)|dξh1η0η|q^(ζ)|dζ|f^(η)|dηMq^L2([h,))fL2(),\displaystyle\int_{h}^{\infty}\bigl{|}\widehat{\bar{q}R_{0}f}(\xi)\bigr{|}\,d\xi\lesssim\int_{h}^{\infty}\tfrac{1}{\eta}\int_{0}^{\eta}\bigl{|}\widehat{q}(\zeta)\bigr{|}\,d\zeta\bigl{|}\widehat{f}(\eta)\bigr{|}\,d\eta\lesssim\|\textsf{M}\widehat{q}\|_{L^{2}([h,\infty))}\|f\|_{L^{2}(\mathbb{R})},

where M denotes the Hardy–Littlewood maximal function.

Choosing h=0h=0 in (2.4), we find that

E|[gC+(q¯R0f)](x)|2𝑑xgL2(E)2C+(q¯R0f)L()2gL2(E)2qL2()2fL2()2,\displaystyle\int_{E}\bigl{|}[gC_{+}(\overline{q}R_{0}f)](x)\bigr{|}^{2}dx\leq\|g\|_{L^{2}(E)}^{2}\|C_{+}(\overline{q}R_{0}f)\|_{L^{\infty}(\mathbb{R})}^{2}\lesssim\|g\|_{L^{2}(E)}^{2}\|q\|_{L^{2}(\mathbb{R})}^{2}\|f\|_{L^{2}(\mathbb{R})}^{2},

uniformly for EE\subseteq\mathbb{R}. Taking E=E=\mathbb{R}, this shows that the operator admits the norm bound stated in (2.2). By choosing EE of the form {x:|x|>r}\{x:|x|>r\}, we also see that the image of the unit ball in L+2()L^{2}_{+}(\mathbb{R}) under the operator gC+q¯R0(κ)gC_{+}\overline{q}R_{0}(\kappa) is tight in L2()L^{2}(\mathbb{R}).

As a second application of (2.4) together with considerations of the support of a convolution, we get

2h|[gC+(q¯R0f)]\displaystyle\int_{2h}^{\infty}\bigl{|}[gC_{+}(\overline{q}R_{0}f)] ^(ξ)|2dξ\displaystyle\widehat{\ }(\xi)\bigr{|}^{2}\,d\xi
g^L2([h,))2q¯R0f^L1([0,))2+g^L2()2q¯R0f^L1([h,))2\displaystyle\leq\|\widehat{g}\|_{L^{2}([h,\infty))}^{2}\|\widehat{\overline{q}R_{0}f}\|_{L^{1}([0,\infty))}^{2}+\|\widehat{g}\|_{L^{2}(\mathbb{R})}^{2}\|\widehat{\overline{q}R_{0}f}\|_{L^{1}([h,\infty))}^{2}
[g^L2([h,))2q^L2()2+g^L2()2Mq^L2([h,))2]fL2()2\displaystyle\lesssim\Bigl{[}\|\widehat{g}\|_{L^{2}([h,\infty))}^{2}\|\widehat{q}\|_{L^{2}(\mathbb{R})}^{2}+\|\widehat{g}\|_{L^{2}(\mathbb{R})}^{2}\|\textsf{M}\widehat{q}\|_{L^{2}([h,\infty))}^{2}\Bigr{]}\|f\|_{L^{2}(\mathbb{R})}^{2}

and so the image of the unit ball in L+2()L^{2}_{+}(\mathbb{R}) under gC+q¯R0(κ)gC_{+}\overline{q}R_{0}(\kappa) is also tight on the Fourier side.

In conclusion, the image of the unit ball in L+2()L^{2}_{+}(\mathbb{R}) under the operator gC+q¯R0(κ)gC_{+}\overline{q}R_{0}(\kappa) is precompact in L2()L^{2}(\mathbb{R}) and so this is a compact operator.

We now turn our attention to property (2.3) of the operator qC+q¯R0(κ)qC_{+}\overline{q}R_{0}(\kappa). Each qQq\in Q may be decomposed in frequency as q=qN+q>Nq=q_{\leq N}+q_{>N}. Accordingly,

qC+q¯R0(κ)=q>NC+q¯R0(κ)+qNC+q¯>NR0(κ)+qNC+q¯NR0(κ).qC_{+}\overline{q}R_{0}(\kappa)=q_{>N}C_{+}\overline{q}R_{0}(\kappa)+q_{\leq N}C_{+}\overline{q}_{>N}R_{0}(\kappa)+q_{\leq N}C_{+}\overline{q}_{\leq N}R_{0}(\kappa).

The first two summands are treated using (2.2):

q>NC+q¯R0(κ)+qNC+q¯>NR0(κ)opqL2q>NL2.\bigl{\|}q_{>N}C_{+}\overline{q}R_{0}(\kappa)+q_{\leq N}C_{+}\overline{q}_{>N}R_{0}(\kappa)\bigr{\|}_{\mathrm{op}}\lesssim\|q\|_{L^{2}}\|q_{>N}\|_{L^{2}}.

As QQ is equicontinuous, choosing NN large makes this small uniformly for all qQq\in Q.

Using Bernstein’s inequality, we may bound the last summand as follows:

qNC+q¯NR0(κ)opκ1qNL2κ1NqL22.\bigl{\|}q_{\leq N}C_{+}\overline{q}_{\leq N}R_{0}(\kappa)\bigr{\|}_{\mathrm{op}}\lesssim\kappa^{-1}\|q_{\leq N}\|_{L^{\infty}}^{2}\lesssim\kappa^{-1}N\|q\|_{L^{2}}^{2}.

Irrespective of how NN is chosen, this can be made small by choosing κ\kappa large.

By employing TTT^{*}T arguments and complex interpolation, we find

C+q¯R0op4=R0qC+q¯R0op2R0qC+q¯opqC+q¯R0op=qC+q¯R0op2.\displaystyle\big{\lVert}C_{+}\overline{q}\sqrt{R_{0}}\big{\rVert}_{\mathrm{op}}^{4}=\big{\lVert}\sqrt{R_{0}}qC_{+}\overline{q}\sqrt{R_{0}}\big{\rVert}_{\mathrm{op}}^{2}\leq\big{\lVert}R_{0}qC_{+}\overline{q}\big{\rVert}_{\mathrm{op}}\big{\lVert}qC_{+}\overline{q}R_{0}\big{\rVert}_{\mathrm{op}}=\big{\lVert}qC_{+}\overline{q}R_{0}\big{\rVert}_{\mathrm{op}}^{2}.

In particular, this norm also converges to zero as κ\kappa\to\infty, uniformly for qQq\in Q. ∎

Proposition 2.2 (Lax operator).

For qL+2()q\in L^{2}_{+}(\mathbb{R}), the operator

qf=ifqC+(q¯f)\mathcal{L}_{q}f=-if^{\prime}\mp qC_{+}(\overline{q}f)

with domain D(q)=H+1()D(\mathcal{L}_{q})=H^{1}_{+}(\mathbb{R}) is self-adjoint and σess(q)=[0,)\sigma_{\mathrm{ess}}(\mathcal{L}_{q})=[0,\infty). Moreover, the mapping qqq\mapsto\mathcal{L}_{q} is continuous in the norm resolvent topology.

If QL+2()Q\subset L^{2}_{+}(\mathbb{R}) is bounded and equicontinuous, then there exists κ0=κ0(Q)>0\kappa_{0}=\kappa_{0}(Q)>0 so that

(2.5) 12(0+κ)(q+κ)32(0+κ),\tfrac{1}{2}(\mathcal{L}_{0}+\kappa)\leq(\mathcal{L}_{q}+\kappa)\leq\tfrac{3}{2}(\mathcal{L}_{0}+\kappa),\\

as quadratic forms, whenever κκ0;\kappa\geq\kappa_{0}; moreover,

(2.6) (q+κ)sfL2(0+κ)sfL2\bigl{\|}(\mathcal{L}_{q}+\kappa)^{s}f\bigr{\|}_{L^{2}}\approx\bigl{\|}(\mathcal{L}_{0}+\kappa)^{s}f\bigr{\|}_{L^{2}}

uniformly for qQq\in Q, κκ0\kappa\geq\kappa_{0}, and 1s1-1\leq s\leq 1.

Proof.

Compactness of the operator (2.2) guarantees that qC+q¯qC_{+}\bar{q} is infinitesimally 0\mathcal{L}_{0}-bounded. Thus, self-adjointness and semi-boundedness follow from the Kato–Rellich Theorem [Reed1975, §X.2]. Through Weyl’s Theorem, [Reed1978, §XIII.4], compactness of the operator also shows that the essential spectrum agrees with that of 0\mathcal{L}_{0}.

By (2.3), we may choose κ0=κ0(Q)>0\kappa_{0}=\kappa_{0}(Q)>0 so that κκ0\kappa\geq\kappa_{0} implies

f,qC+(q¯f)=C+(q¯f),C+(q¯f)\displaystyle\bigl{\langle}f,qC_{+}(\overline{q}f)\bigr{\rangle}=\bigl{\langle}C_{+}(\overline{q}f),C_{+}(\overline{q}f)\bigr{\rangle} C+q¯R0L220+κfL22\displaystyle\leq\bigl{\|}C_{+}\overline{q}\sqrt{R_{0}}\bigr{\|}_{L^{2}}^{2}\bigl{\|}\sqrt{\mathcal{L}_{0}+\kappa}\,f\bigr{\|}_{L^{2}}^{2}
12f,(0+κ)f,\displaystyle\leq\tfrac{1}{2}\bigl{\langle}f,(\mathcal{L}_{0}+\kappa)f\bigr{\rangle},

and also

qC+(q¯f)L2qC+q¯R0(κ)op(0+κ)fL212(0+κ)fL2.\bigl{\|}qC_{+}(\overline{q}f)\bigr{\|}_{L^{2}}\leq\bigl{\|}qC_{+}\overline{q}R_{0}(\kappa)\bigr{\|}_{\mathrm{op}}\bigl{\|}(\mathcal{L}_{0}+\kappa)f\bigr{\|}_{L^{2}}\leq\tfrac{1}{2}\bigl{\|}(\mathcal{L}_{0}+\kappa)f\bigr{\|}_{L^{2}}.

The former inequality proves (2.5). The latter shows that

14(0+κ)fL22(q+κ)fL2294(0+κ)fL22,\tfrac{1}{4}\bigl{\|}(\mathcal{L}_{0}+\kappa)f\bigr{\|}^{2}_{L^{2}}\leq\bigl{\|}(\mathcal{L}_{q}+\kappa)f\bigr{\|}^{2}_{L^{2}}\leq\tfrac{9}{4}\bigl{\|}(\mathcal{L}_{0}+\kappa)f\bigr{\|}^{2}_{L^{2}},

which proves the s=1s=1 case of (2.6).

Noting that s=0s=0 is trivial, (2.6) follows for all s[0,1]s\in[0,1] by (real or complex) interpolation. Negative values of ss then follow by duality.

To prove that qqq\mapsto\mathcal{L}_{q} is norm-resolvent continuous, we first choose qnqq_{n}\to q in L+2L^{2}_{+} and then choose κ\kappa large enough so that (2.6) holds for qq and for all qnq_{n}. In particular, κ\kappa belongs to the resolvent set of q\mathcal{L}_{q} and all qn\mathcal{L}_{q_{n}}. By the resolvent identity, (2.6), and (2.2),

R(κ,qn)R(κ,q)op\displaystyle\|R(\kappa,q_{n})-R(\kappa,q)\|_{\mathrm{op}} R(κ,qn)[qnC+q¯nqC+q¯]R(κ,q)op\displaystyle\lesssim\|R(\kappa,q_{n})[q_{n}C_{+}\overline{q}_{n}-qC_{+}\overline{q}]R(\kappa,q)\|_{\mathrm{op}}
R0(κ)[qnC+q¯nqC+q¯]R0(κ)op\displaystyle\lesssim\|R_{0}(\kappa)[q_{n}C_{+}\overline{q}_{n}-qC_{+}\overline{q}]R_{0}(\kappa)\|_{\mathrm{op}}
qnqL2(qnL2+qL2)R0(κ)op,\displaystyle\lesssim\|q_{n}-q\|_{L^{2}}(\|q_{n}\|_{L^{2}}+\|q\|_{L^{2}})\|R_{0}(\kappa)\|_{\mathrm{op}},

which converges to zero as nn\to\infty. ∎

The operator 𝒫(q)\mathcal{P}(q) from the Lax pair (1.9) plays a less central role in our analysis than q\mathcal{L}_{q}. Our next proposition constructs a unitary transformation built from the associated differential equation (2.7). It will suffice for us to work with smooth solutions q(t)q(t) and so avoid any functional-analytic subtleties of the type addressed in the preceding proposition.

Proposition 2.3.

Let q(t)q(t) be a global H+()H^{\infty}_{+}(\mathbb{R}) solution of (CCM). For all t0t_{0}\in\mathbb{R} and all ψ0L+2()\psi_{0}\in L^{2}_{+}(\mathbb{R}), the initial value problem

(2.7) ddtψ(t)=𝒫(q(t))ψ(t)withψ(t0)=ψ0\displaystyle\tfrac{d}{dt}\psi(t)=\mathcal{P}(q(t))\psi(t)\quad\text{with}\quad\psi(t_{0})=\psi_{0}

admits a unique CtL+2Ct1H+2C_{t}L^{2}_{+}\cap C_{t}^{1}H^{-2}_{+} solution; this is global in time. Moreover, for each tt\in\mathbb{R} the mapping U(t;t0):ψ0ψ(t)U(t;t_{0}):\psi_{0}\mapsto\psi(t) is unitary on L+2L^{2}_{+},

(2.8) q(t)=U(t;t0)q(t0),and(q(t))=U(t;t0)(q(t0))U(t;t0).\displaystyle q(t)=U(t;t_{0})q(t_{0}),\quad\text{and}\quad\mathcal{L}(q(t))=U(t;t_{0})\mathcal{L}(q(t_{0}))U^{*}(t;t_{0}).

If ψ0H+()\psi_{0}\in H^{\infty}_{+}(\mathbb{R}), then so too is ψ(t)\psi(t) for all tt\in\mathbb{R}. Finally, if ψ0H+()\psi_{0}\in H^{\infty}_{+}(\mathbb{R}), xψ0L2()\langle x\rangle\psi_{0}\in L^{2}(\mathbb{R}), and xq(0)L2()\langle x\rangle q(0)\in L^{2}(\mathbb{R}), then xψ(t)L2()\langle x\rangle\psi(t)\in L^{2}(\mathbb{R}) for all tt\in\mathbb{R}.

Proof.

Given ε>0\varepsilon>0, we consider the regularized initial value problem

(2.9) ddtψε(t)=i2ψε(t)±2q(t)1ε22C+q¯(t)ψε(t)withψε(t0)=ψ0.\displaystyle\tfrac{d}{dt}\psi_{\varepsilon}(t)=i\partial^{2}\psi_{\varepsilon}(t)\pm 2q(t)\tfrac{\partial}{1-\varepsilon^{2}\partial^{2}}C_{+}\overline{q}(t)\psi_{\varepsilon}(t)\quad\text{with}\quad\psi_{\varepsilon}(t_{0})=\psi_{0}.

This is readily solved locally in time by employing Duhamel’s formula,

(2.10) ψε(t)=ei(tt0)Δψε(t0)±2t0tei(ts)Δq(s)1ε22C+q¯(s)ψε(s)𝑑s,\displaystyle\psi_{\varepsilon}(t)=e^{i(t-t_{0})\Delta}\psi_{\varepsilon}(t_{0})\pm 2\int_{t_{0}}^{t}e^{i(t-s)\Delta}q(s)\tfrac{\partial}{1-\varepsilon^{2}\partial^{2}}C_{+}\overline{q}(s)\psi_{\varepsilon}(s)\,ds,

and contraction mapping in CtL+2C_{t}L^{2}_{+}.

For any such solution, ei(tt0)Δψε(t)e^{-i(t-t_{0})\Delta}\psi_{\varepsilon}(t) belongs to Ct1L+2C^{1}_{t}L^{2}_{+} and so one may readily verify that the L2L^{2} norm is conserved. This in turn yields global well-posedness by iterated contraction mapping.

The L2L^{2} conservation law also implies that the flow maps Uε(t;t0)U_{\varepsilon}(t;t_{0}) are isometries. Moreover, the uniqueness of solutions (guaranteed by contraction mapping) shows that Uε(t;t0)U_{\varepsilon}(t;t_{0}) and Uε(t0;t)U_{\varepsilon}(t_{0};t) are inverses of one another and so both are unitary.

Proceeding by induction, taking spatial derivatives of (2.10) and applying the Gronwall inequality, we find that for each nn\in\mathbb{N} and each T>0T>0,

(2.11) ψε(t)H+nCn(q,T)ψ0H+nfor allt,t0[T,T].\displaystyle\|\psi_{\varepsilon}(t)\|_{H^{n}_{+}}\leq C_{n}(q,T)\|\psi_{0}\|_{H^{n}_{+}}\quad\text{for all}\quad t,t_{0}\in[-T,T].

Note that the constant Cn(q,T)C_{n}(q,T) does not depend on ε>0\varepsilon>0.

Let us now compare differing regularizations, beginning with the case ψ0H+2\psi_{0}\in H^{2}_{+}. Given 0<ε<η0<\varepsilon<\eta, taking the time derivative and integrating by parts to exploit the antisymmetry of the regularization, we obtain

ddtψη(t)ψε\displaystyle\tfrac{d}{dt}\|\psi_{\eta}(t)-\psi_{\varepsilon} (t)L22\displaystyle(t)\|_{L^{2}}^{2}
ψη(t)ψε(t)L2q(t)(η2ε2)3(1η22)(1ε22)C+q¯(t)[ψη(t)+ψε(t)]L2\displaystyle\lesssim\|\psi_{\eta}(t)-\psi_{\varepsilon}(t)\|_{L^{2}}\bigl{\|}q(t)\tfrac{(\eta^{2}-\varepsilon^{2})\partial^{3}}{(1-\eta^{2}\partial^{2})(1-\varepsilon^{2}\partial^{2})}C_{+}\overline{q}(t)[\psi_{\eta}(t)+\psi_{\varepsilon}(t)]\bigr{\|}_{L^{2}}
ηψη(t)ψε(t)L2q(t)H22ψη(t)+ψε(t)H2.\displaystyle\lesssim\eta\|\psi_{\eta}(t)-\psi_{\varepsilon}(t)\|_{L^{2}}\|q(t)\|_{H^{2}}^{2}\bigl{\|}\psi_{\eta}(t)+\psi_{\varepsilon}(t)\bigr{\|}_{H^{2}}.

Combining this, (2.11), and Gronwall’s inequality, we deduce that for each T>0T>0,

ψη(t)ψε(t)L22\displaystyle\|\psi_{\eta}(t)-\psi_{\varepsilon}(t)\|_{L^{2}}^{2} ηC(q,T)ψ0H22for allt,t0[T,T].\displaystyle\lesssim\eta C(q,T)\big{\lVert}\psi_{0}\big{\rVert}_{H^{2}}^{2}\quad\text{for all}\quad t,t_{0}\in[-T,T].

This proves that ψε(t)\psi_{\varepsilon}(t) converges in Ct([T,T];L+2)C_{t}([-T,T];L^{2}_{+}) as ε0\varepsilon\to 0 for any choice of T>0T>0 and any ψ0H+2\psi_{0}\in H^{2}_{+}.

Unitarity of Uε(t;t0)U_{\varepsilon}(t;t_{0}) together with the density of H2H^{2} show that this convergence carries over to every T>0T>0 and every ψ0L+2\psi_{0}\in L^{2}_{+}. Note that the limiting functions ψ(t)\psi(t) satisfy

(2.12) ψ(t)=ei(tt0)Δψ(t0)±2t0tei(ts)Δq(s)C+q¯(s)ψ(s)ds,\displaystyle\psi(t)=e^{i(t-t_{0})\Delta}\psi(t_{0})\pm 2\int_{t_{0}}^{t}e^{i(t-s)\Delta}q(s)\partial C_{+}\overline{q}(s)\psi(s)\,ds,

and so are CtL+2Ct1H+2C_{t}L^{2}_{+}\cap C^{1}_{t}H^{-2}_{+} solutions to (2.7); moreover, these solutions inherit conservation of L2L^{2} and the persistence of regularity estimate (2.11).

Earlier we saw that Uε(t;t0)Uε(t0;t)=IdU_{\varepsilon}(t;t_{0})U_{\varepsilon}(t_{0};t)=\operatorname{Id}. Sending ε0\varepsilon\to 0, we deduce that U(t;t0)U(t;t_{0}) is unitary for all t,t0t,t_{0}.

Suppose (toward a contradiction) that (2.7) admits two CtL+2C_{t}L^{2}_{+} solutions ψ(t)\psi(t) and ψ~(t)\widetilde{\psi}(t) that differ at some time t1t_{1}\in\mathbb{R}. Evidently, there is a choice of ϕ1H+\phi_{1}\in H^{\infty}_{+} so that ϕ1,ψ(t1)ψ~(t1)0\langle\phi_{1},\psi(t_{1})-\widetilde{\psi}(t_{1})\rangle\neq 0. By the analysis presented above, we can find a CtH+C_{t}H^{\infty}_{+} solution ϕ(t)=U(t;t1)ϕ1\phi(t)=U(t;t_{1})\phi_{1}. Integrating by parts (to exhibit the antisymmetry of 𝒫\mathcal{P}), it follows that ddtϕ(t),ψ(t)ψ~(t)=0\frac{d}{dt}\langle\phi(t),\,\psi(t)-\widetilde{\psi}(t)\rangle=0. This yields a contradiction:

0ϕ1,ψ(t1)ψ~(t1)=ϕ(t0),ψ(t0)ψ~(t0)=0.0\neq\langle\phi_{1},\psi(t_{1})-\widetilde{\psi}(t_{1})\rangle=\langle\phi(t_{0}),\,\psi(t_{0})-\widetilde{\psi}(t_{0})\rangle=0.

We have now completed the proof of existence, uniqueness, and persistence of regularity for solutions to (2.7), and unitarity of the associated propagator.

We now turn to (2.8). As noted in the introduction, one of the favorable attributes of our choice of 𝒫\mathcal{P} is that ddtq=𝒫(q)q\frac{d}{dt}q=\mathcal{P}(q)q. The uniqueness of such solutions demonstrated above shows that q(t)=U(t;t0)q(t0)q(t)=U(t;t_{0})q(t_{0}).

For any ϕ0H+()\phi_{0}\in H^{\infty}_{+}(\mathbb{R}), the fact that \mathcal{L} and 𝒫\mathcal{P} form a Lax pair then shows that

t[(q(t))U(t;t0)U(t;t0)(q(t0))]ϕ0\displaystyle t\mapsto\bigl{[}\mathcal{L}(q(t))U(t;t_{0})-U(t;t_{0})\mathcal{L}(q(t_{0}))\bigr{]}\phi_{0}

constitutes a solution to (2.7) with zero initial data at time t=t0t=t_{0}. Uniqueness of such solutions combined with unitarity of U(t;t0)U(t;t_{0}) then proves (2.8).

It remains to discuss weighted estimates. We begin with q(t)q(t) itself and assume that xq0(x)L2()\langle x\rangle q^{0}(x)\in L^{2}(\mathbb{R}). Integration by parts shows that

(2.13) ddtw(t)|q(t,x)|2𝑑x=w(x)[2Im(q¯q)|q|4](t,x)𝑑x\displaystyle\frac{d}{dt}\int w(t)|q(t,x)|^{2}\,dx=\int w^{\prime}(x)\bigl{[}2\operatorname{Im}(\overline{q}q^{\prime})\mp|q|^{4}\bigr{]}(t,x)\,dx

for any smooth and bounded weight ww. Observe that the rate of growth is mediated by the size of the derivative of ww, as well as the assumed H1H^{1} bounds on q(t)q(t). In this way, using a sequence of bounded approximating weights and a Grownwall argument, we obtain a quantitative bound on the L2()L^{2}(\mathbb{R}) norm of xq(t,x)\langle x\rangle q(t,x).

The weighted norms of ψ(t)\psi(t) do not follow such a simple evolution as (2.13). Nevertheless, proceeding in the same vein, we verify that

(2.14) |ddtxψL22|xψL2ψH1+xqL2xψL2qH1ψH1\displaystyle\Bigl{|}\tfrac{d}{dt}\bigl{\|}\langle x\rangle\psi\bigr{\|}_{L^{2}}^{2}\Bigr{|}\lesssim\|\langle x\rangle\psi\|_{L^{2}}\|\psi\|_{H^{1}}+\|\langle x\rangle q\|_{L^{2}}\|\langle x\rangle\psi\|_{L^{2}}\|q\|_{H^{1}}\|\psi\|_{H^{1}}

and then deduce that xψ(t,x)L2()\langle x\rangle\psi(t,x)\in L^{2}(\mathbb{R}) for all time via Gronwall. ∎

It is not difficult to see that the arguments just presented also prove an analogue of Proposition 2.3 in the case that q(t)q(t) is merely a local-in-time smooth solution. We do not need such generality because smooth solutions with mass below the equicontinuity threshold are automatically global:

Lemma 2.4.

Let q:[0,T)H+()q:[0,T)\to H^{\infty}_{+}(\mathbb{R}) be a solution to (CCM). If {q(t):t[0,T)}\{q(t):t\in[0,T)\} is L2L^{2}-equicontinuous, then q(t)q(t) extends H+()H^{\infty}_{+}(\mathbb{R})-continuously to [0,T][0,T].

Proof.

We will use conservation of the Hamiltonian to show that q(t)q(t) is H1H^{1}-bounded on [0,T)[0,T). This guarantees H1H^{1}-continuous extension by the results of [GerardLenzmann] described in Theorem 1.3. It also provides the base step of their inductive argument yielding control over all higher Sobolev norms.

Given a frequency cutoff parameter NN, we may combine the Bernstein and Gagliardo–Nirenberg inequalities to see that

qL66N2qNL26+q>NL24q>NL22N2M(q)3+q>NL24qL22\displaystyle\|q\|_{L^{6}}^{6}\lesssim N^{2}\|q_{\leq N}\|_{L^{2}}^{6}+\|q_{>N}\|_{L^{2}}^{4}\|q_{>N}^{\prime}\|_{L^{2}}^{2}\lesssim N^{2}M(q)^{3}+\|q_{>N}\|_{L^{2}}^{4}\|q^{\prime}\|_{L^{2}}^{2}

uniformly for 0t<T0\leq t<T. As q(t)q(t) is assumed to be L2L^{2}-equicontinuous, given η>0\eta>0, we may choose N(η)N(\eta) so that

(2.15) q(t)L66η2q(t)L22+N(η)2M(q)uniformly for t[0,T).\displaystyle\|q(t)\|_{L^{6}}^{6}\lesssim\eta^{2}\|q^{\prime}(t)\|_{L^{2}}^{2}+N(\eta)^{2}M(q)\quad\text{uniformly for $t\in[0,T)$.}

The role of this estimate will be to control the discrepancy between the full Hamiltonians (1.1) and their quadratic part: H±[2]:=12|q|2𝑑xH_{\pm}^{[2]}:=\int\frac{1}{2}|q^{\prime}|^{2}\,dx.

Using L3L^{3}-boundedness of C+C_{+}, we find that for any 0<η<10<\eta<1,

|H±(q(t))H±[2](q(t))|ηq(t)L22+η1q(t)L66.\displaystyle\Bigl{|}H_{\pm}(q(t))-H_{\pm}^{[2]}(q(t))\Bigr{|}\lesssim\eta\|q^{\prime}(t)\|_{L^{2}}^{2}+\eta^{-1}\|q(t)\|_{L^{6}}^{6}.

Combining this with (2.15) and choosing η\eta sufficiently small, we deduce that q(t)q(t) is bounded in H1H^{1} from the conservation of the Hamiltonian. ∎

Schatten-class properties of the resolvent of the Lax operator q\mathcal{L}_{q} will play a key role in proving equicontinuity of orbits in the next section. To demonstrate these properties, we will employ the following information about the free resolvent:

Lemma 2.5.

For qL+2()q\in L^{2}_{+}(\mathbb{R}) and κ>0\kappa>0,

(2.16) C+q¯R0(κ)22=R0(κ)q22\displaystyle\left\lVert C_{+}\overline{q}R_{0}(\kappa)\right\rVert_{\mathfrak{I}_{2}}^{2}=\left\lVert R_{0}(\kappa)q\right\rVert_{\mathfrak{I}_{2}}^{2} =12π0|q^(ξ)|2ξ+κ𝑑ξ,\displaystyle=\frac{1}{2\pi}\int_{0}^{\infty}\frac{|\widehat{q}(\xi)|^{2}}{\xi+\kappa}\,d\xi,
(2.17) C+q¯R0(κ)q2\displaystyle\left\lVert C_{+}\overline{q}R_{0}(\kappa)q\right\rVert_{\mathfrak{I}_{2}} qL22.\displaystyle\lesssim\|q\|_{L^{2}}^{2}.
Proof.

The two operators in (2.16) are adjoints of one another. Their Hilbert–Schmidt norm follows from an explicit computation in Fourier variables:

R0(κ)q22=tr{R0(κ)qC+q¯R0(κ)}\displaystyle\left\lVert R_{0}(\kappa)q\right\rVert_{\mathfrak{I}_{2}}^{2}=\operatorname{tr}\big{\{}R_{0}(\kappa)qC_{+}\overline{q}R_{0}(\kappa)\big{\}} =12π001(ξ+κ)2q^(ξη)q¯^(ηξ)𝑑η𝑑ξ\displaystyle=\frac{1}{2\pi}\int_{0}^{\infty}\!\!\int_{0}^{\infty}\!\!\frac{1}{(\xi+\kappa)^{2}}\widehat{q}(\xi-\eta)\widehat{\overline{q}}(\eta-\xi)\,d\eta\,d\xi
=12π00|q^(ξ)|2(ξ+η+κ)2𝑑η𝑑ξ\displaystyle=\frac{1}{2\pi}\!\!\int_{0}^{\infty}\!\!\int_{0}^{\infty}\frac{|\widehat{q}(\xi)|^{2}}{(\xi+\eta+\kappa)^{2}}\,d\eta\,d\xi
=12π0|q^(ξ)|2ξ+κ𝑑ξ.\displaystyle=\frac{1}{2\pi}\!\!\int_{0}^{\infty}\frac{|\widehat{q}(\xi)|^{2}}{\xi+\kappa}\,d\xi.

We turn now to (2.17). A simple frequency analysis shows that on L+2L^{2}_{+} we have the operator identity

C+q¯R0(κ)q=N1,N2M110N1N2C+q¯N1R0(κ)PMqN2,C_{+}\overline{q}R_{0}(\kappa)q=\sum_{N_{1},N_{2}}\ \sum_{M\geq\frac{1}{10}N_{1}\vee N_{2}}C_{+}\smash[b]{\overline{q}}_{N_{1}}R_{0}(\kappa)P_{M}q_{N_{2}},

where we employ the notation N1N2=max{N1,N2}N_{1}\vee N_{2}=\max\{N_{1},N_{2}\}. Thus, using the Hölder inequality for Schatten classes together with [TraceIdeals, Theorem 4.1] and the Bernstein inequality, we may bound

C+q¯R0(κ)q2\displaystyle\left\lVert C_{+}\overline{q}R_{0}(\kappa)q\right\rVert_{\mathfrak{I}_{2}} N1,N2M110N1N2q¯N1R0(κ)PM4PMR0(κ)qN24\displaystyle\lesssim\sum_{N_{1},N_{2}}\sum_{M\geq\frac{1}{10}N_{1}\vee N_{2}}\left\lVert\overline{q}_{N_{1}}\sqrt{R_{0}(\kappa)}P_{M}\right\rVert_{\mathfrak{I}_{4}}\left\lVert P_{M}\sqrt{R_{0}(\kappa)}q_{N_{2}}\right\rVert_{\mathfrak{I}_{4}}
N1,N2M110N1N2qN1L4qN2L41ξ+κL4(|ξ|M)2\displaystyle\lesssim\sum_{N_{1},N_{2}}\sum_{M\geq\frac{1}{10}N_{1}\vee N_{2}}\left\lVert q_{N_{1}}\right\rVert_{L^{4}}\left\lVert q_{N_{2}}\right\rVert_{L^{4}}\bigl{\|}\sqrt{\tfrac{1}{\xi+\kappa}}\bigr{\|}_{L^{4}(|\xi|\sim M)}^{2}
N1,N2M110N1N2(N1N2)14M12M+κqN1L2qN2L2\displaystyle\lesssim\sum_{N_{1},N_{2}}\sum_{M\geq\frac{1}{10}N_{1}\vee N_{2}}(N_{1}N_{2})^{\frac{1}{4}}\frac{M^{\frac{1}{2}}}{M+\kappa}\left\lVert q_{N_{1}}\right\rVert_{L^{2}}\left\lVert q_{N_{2}}\right\rVert_{L^{2}}
N1,N2(N1N2)14(N1N2)12qN1L2qN2L2\displaystyle\lesssim\sum_{N_{1},N_{2}}\frac{(N_{1}N_{2})^{\frac{1}{4}}}{(N_{1}\vee N_{2})^{\frac{1}{2}}}\left\lVert q_{N_{1}}\right\rVert_{L^{2}}\left\lVert q_{N_{2}}\right\rVert_{L^{2}}
qL22,\displaystyle\lesssim\|q\|_{L^{2}}^{2},

where the last line follows from applying the Schur test to sum in N1N_{1} and N2N_{2}. ∎

We adopt the following notation for the resolvent of q\mathcal{L}_{q}:

(2.18) R(κ,q)=(q+κ)1.R(\kappa,q)=(\mathcal{L}_{q}+\kappa)^{-1}.

Although R(κ)R(\kappa) does not belong to trace class, the difference R(κ)R0(κ)R(\kappa)-R_{0}(\kappa) does, at least for κ\kappa sufficiently large. This is the topic of the next lemma. More striking is the fact that this trace is conserved under the (CCM) flow; this is the subject of the subsequent Proposition 2.7.

Lemma 2.6.

Given qL+2()q\in L^{2}_{+}(\mathbb{R}) and κ>0\kappa>0 such that q+κ\mathcal{L}_{q}+\kappa is positive definite,

(2.19) R(κ)R0(κ),R0(κ)qC+q¯R0(κ),andR0(κ)qC+q¯R(κ)qC+q¯R0(κ)R(\kappa)-R_{0}(\kappa),\quad R_{0}(\kappa)qC_{+}\overline{q}R_{0}(\kappa),\quad\text{and}\quad R_{0}(\kappa)qC_{+}\overline{q}R(\kappa)qC_{+}\overline{q}R_{0}(\kappa)

all belong to the trace class 1.\mathfrak{I}_{1}. Moreover, for any bounded and equicontinuous QQ,

(2.20) limκsupqQκR0(κ)qC+q¯R(κ)qC+q¯R0(κ)1=0.\displaystyle\lim_{\kappa\to\infty}\sup_{q\in Q}\kappa\big{\|}R_{0}(\kappa)qC_{+}\overline{q}R(\kappa)qC_{+}\overline{q}R_{0}(\kappa)\big{\|}_{\mathfrak{I}_{1}}=0.
Proof.

The resolvent identity

(2.21) R(κ)R0(κ)=±R0(κ)qC+q¯R0(κ)+R0(κ)qC+q¯R(κ)qC+q¯R0(κ)R(\kappa)-R_{0}(\kappa)=\pm R_{0}(\kappa)qC_{+}\overline{q}R_{0}(\kappa)+R_{0}(\kappa)qC_{+}\overline{q}R(\kappa)qC_{+}\overline{q}R_{0}(\kappa)

reduces consideration of the first operator in (2.19) to treatment of the other two.

The second operator in (2.19) has the form TTTT^{*} with T=R0(κ)qT=R_{0}(\kappa)q and so is positive definite. Together with (2.16), this yields

(2.22) R0(κ)qC+q¯R0(κ)1=tr{R0(κ)qC+q¯R0(κ)}=12π0|q^(ξ)|2ξ+κ𝑑ξ.\big{\|}R_{0}(\kappa)qC_{+}\overline{q}R_{0}(\kappa)\big{\|}_{\mathfrak{I}_{1}}=\operatorname{tr}\bigl{\{}R_{0}(\kappa)qC_{+}\overline{q}R_{0}(\kappa)\bigr{\}}=\frac{1}{2\pi}\int_{0}^{\infty}\frac{|\widehat{q}(\xi)|^{2}}{\xi+\kappa}\,d\xi.

Thus our claims about this operator hold for all κ>0\kappa>0.

To treat the third operator in (2.19), we need some information about R(κ)R(\kappa). If we choose κ0=κ0({q})\kappa_{0}=\kappa_{0}(\{q\}) according to Proposition 2.2, then (2.5) guarantees that 0R(κ)2R0(κ)0\leq R(\kappa)\leq 2R_{0}(\kappa) as quadratic forms whenever κκ0\kappa\geq\kappa_{0}. In this way, we see that

0R0(κ)qC+q¯R(κ)qC+q¯R0(κ)2R0(κ)qC+q¯R0(κ)qC+q¯R0(κ),\displaystyle 0\leq R_{0}(\kappa)qC_{+}\overline{q}R(\kappa)qC_{+}\overline{q}R_{0}(\kappa)\leq 2\,R_{0}(\kappa)qC_{+}\overline{q}R_{0}(\kappa)qC_{+}\overline{q}R_{0}(\kappa),

still in the sense of quadratic forms. From (2.16) and (2.17), it follows that

(2.23) R0(κ)qC+q¯R(κ)qC+q¯R0(κ)1\displaystyle\big{\|}R_{0}(\kappa)qC_{+}\overline{q}R(\kappa)qC_{+}\overline{q}R_{0}(\kappa)\big{\|}_{\mathfrak{I}_{1}} 2R0(κ)q22C+q¯R0(κ)qop\displaystyle\leq 2\,\big{\|}R_{0}(\kappa)q\bigr{\|}_{\mathfrak{I}_{2}}^{2}\big{\|}C_{+}\overline{q}R_{0}(\kappa)\,q\bigr{\|}_{\mathrm{op}}
κ1qL24.\displaystyle\lesssim\kappa^{-1}\|q\|_{L^{2}}^{4}.

Claim (2.20) follows from (2.23), (2.16), and (2.3) after choosing κ0=κ0(Q)\kappa_{0}=\kappa_{0}(Q). ∎

Lemma 2.7 (A conservation law).

Let q(t)q(t) denote an H+()H^{\infty}_{+}(\mathbb{R}) solution to (CCM) and suppose κ>0\kappa>0 is such that q(0)+κ\mathcal{L}_{q(0)}+\kappa is positive definite. Then q(t)+κ\mathcal{L}_{q(t)}+\kappa is positive definite for all times tt and

(2.24) ddttr{R(κ,q(t))R0(κ)}=0.\tfrac{d}{dt}\operatorname{tr}\{R(\kappa,q(t))-R_{0}(\kappa)\}=0.
Proof.

From (2.8) we see that q(t)+ϰ\mathcal{L}_{q(t)}+\varkappa is positive definite for all tt and all ϰκ\varkappa\geq\kappa. Together with Lemma 2.6, this ensures that the trace (2.24) is well defined.

We claim that both 𝒫~(q(t))R(κ,q(t))2\widetilde{\mathcal{P}}(q(t))R(\kappa,q(t))^{2} and R(κ,q(t))2𝒫~(q(t))R(\kappa,q(t))^{2}\widetilde{\mathcal{P}}(q(t)) are trace class operators, where 𝒫~(q(t))\widetilde{\mathcal{P}}(q(t)) is the alternative form of the Lax operator presented in (1.10). As a first step to verifying this, we observe that for any f,gH()f,g\in H^{\infty}(\mathbb{R}),

C+fC+gR0(κ)2\displaystyle C_{+}fC_{+}gR_{0}(\kappa)^{2} =C+fR0(κ)(0+κ)C+gR0(κ)2\displaystyle=C_{+}fR_{0}(\kappa)\cdot(\mathcal{L}_{0}+\kappa)C_{+}gR_{0}(\kappa)^{2}
=C+fR0(κ)[C+gR0(κ)iC+gR0(κ)2]\displaystyle=C_{+}fR_{0}(\kappa)\Bigl{[}C_{+}gR_{0}(\kappa)-iC_{+}g^{\prime}R_{0}(\kappa)^{2}\Bigr{]}

and so by (2.16), this operator is trace class. Combining this with Lemma 2.6, we deduce that C+fC+gR(κ)2C_{+}fC_{+}gR(\kappa)^{2} is also trace class. This in turn proves that the operator 𝒫~(q(t))R(κ,q(t))2\widetilde{\mathcal{P}}(q(t))R(\kappa,q(t))^{2} is trace class. The fact that the adjoint of this operator is also trace class settles the analogous question for R(κ,q(t))2𝒫~(q(t))R(\kappa,q(t))^{2}\widetilde{\mathcal{P}}(q(t)).

As both products are trace class, it follows that

tr{[𝒫~(q(t)),R(κ,q(t))2]}=0.\operatorname{tr}\bigl{\{}[\widetilde{\mathcal{P}}(q(t)),R(\kappa,q(t))^{2}]\bigr{\}}=0.

By the Fundamental Theorem of Calculus, this yields

tr{R(κ,q(t))2R(κ,q(0))2}=0.\operatorname{tr}\bigl{\{}R(\kappa,q(t))^{2}-R(\kappa,q(0))^{2}\bigr{\}}=0.

Integrating the spectral parameter over the interval [κ,ϰ][\kappa,\varkappa] and rearranging, we get

tr{[R(κ,q(t))R0(κ)]\displaystyle\operatorname{tr}\bigl{\{}[R(\kappa,q(t))-R_{0}(\kappa)] [R(ϰ,q(t))R0(ϰ)]}\displaystyle-[R(\varkappa,q(t))-R_{0}(\varkappa)]\bigr{\}}
=tr{[R(κ,q(0))R0(κ)][R(ϰ,q(0))R0(ϰ)]}.\displaystyle=\operatorname{tr}\bigl{\{}[R(\kappa,q(0))-R_{0}(\kappa)]-[R(\varkappa,q(0))-R_{0}(\varkappa)]\bigr{\}}.

Using (2.21), (2.16), and (2.20), we see that uniformly in tt,

R(ϰ,q(t))R0(ϰ)1=O(ϰ1).\displaystyle\bigl{\|}R(\varkappa,q(t))-R_{0}(\varkappa)\bigr{\|}_{\mathfrak{I}_{1}}=O(\varkappa^{-1}).

Thus sending ϰ\varkappa\to\infty we obtain the desired conservation law. ∎

3. Equicontinuity in L2L^{2}

In this section we prove L2L^{2}-equicontinuity of orbits of H+H^{\infty}_{+} solutions to (CCM), both in the focusing and defocusing settings. In the focusing case, we will require that the initial data satisfies (1.6).

Our key quantity for detecting equicontinuity in L2()L^{2}(\mathbb{R}) is

(3.1) β(κ,q):=M(q)2πκtr{R(κ)R0(κ)}.\beta(\kappa,q):=M(q)\mp 2\pi\kappa\operatorname{tr}\{R(\kappa)-R_{0}(\kappa)\}.

In view of Lemma 2.7, this quantity is conserved by H+H^{\infty}_{+} solutions to (CCM). Moreover, using (2.16) we find that the quadratic term is given by

(3.2) β[2](κ,q)=M(q)2πκtr{R0(κ)qC+q¯R0(κ)}=0ξξ+κ|q^(ξ)|2𝑑ξ.\beta^{[2]}(\kappa,q)=M(q)-2\pi\kappa\operatorname{tr}\{R_{0}(\kappa)qC_{+}\overline{q}R_{0}(\kappa)\}=\int_{0}^{\infty}\frac{\xi}{\xi+\kappa}|\widehat{q}(\xi)|^{2}\,d\xi.

By the resolvent identity (2.21), the remainder can be written

(3.3) β(κ,q)β[2](κ,q)=2πκtr{R0(κ)qC+q¯R(κ)qC+q¯R0(κ)}.\beta(\kappa,q)-\beta^{[2]}(\kappa,q)=\mp 2\pi\kappa\operatorname{tr}\{R_{0}(\kappa)qC_{+}\overline{q}R(\kappa)qC_{+}\overline{q}R_{0}(\kappa)\}.

In the defocusing case, it will suffice to exploit the direct connection between β[2]\beta^{[2]} and q^\widehat{q} visible in (3.2). In the focusing case, the remainder (3.3) is more troublesome and for this purpose we introduce 𝒦\mathcal{K} norms inspired by [MR4628747]: Given an infinite subset 𝒦2\mathcal{K}\subseteq 2^{\mathbb{N}}, we define

(3.4) q𝒦2\displaystyle\|q\|_{\mathcal{K}}^{2} :=qL22+κ𝒦β[2](κ,q).\displaystyle:=\|q\|_{L^{2}}^{2}+\sum_{\kappa\in\mathcal{K}}\beta^{[2]}(\kappa,q).

The connections between β[2](κ,q)\beta^{[2]}(\kappa,q), 𝒦\mathcal{K} norms, equicontinuity, and β(κ,q)\beta(\kappa,q) can be summarized thus:

Lemma 3.1.

Let QQ be a bounded subset of L+2()L^{2}_{+}(\mathbb{R}). The following are equivalent:

  1. (1)

    QQ is equicontinuous in L2()L^{2}(\mathbb{R}).

  2. (2)

    supqQβ[2](κ,q)0\sup_{q\in Q}\beta^{[2]}(\kappa,q)\to 0 as κ\kappa\to\infty.

  3. (3)

    There exists an infinite set 𝒦2\mathcal{K}\subseteq 2^{\mathbb{N}} so that supqQq𝒦<\sup_{q\in Q}\|q\|_{\mathcal{K}}<\infty.

Moreover, if QQ is L2L^{2}-equicontinuous, then

(3.5) supqQβ(κ,q)0asκ.\displaystyle\sup_{q\in Q}\;\beta(\kappa,q)\to 0\quad\text{as}\quad\kappa\to\infty.
Proof.

To see that (1) and (2) are equivalent, we note that for any 0<η10<\eta\leq 1,

(3.6) q>κL22β[2](κ,q)ηqL22+q>ηκL22.\left\lVert q_{>\kappa}\right\rVert_{L^{2}}^{2}\lesssim\beta^{[2]}(\kappa,q)\lesssim\eta\left\lVert q\right\rVert_{L^{2}}^{2}+\left\lVert q_{>\eta\kappa}\right\rVert_{L^{2}}^{2}.

Clearly, (2) implies (3). That (3) implies (1) follows from the observation that

q𝒦2\displaystyle\|q\|_{\mathcal{K}}^{2} qL22+N2#{κ𝒦:κ<N}qNL22.\displaystyle\gtrsim\|q\|_{L^{2}}^{2}+\sum_{N\in 2^{\mathbb{N}}}\#\{\kappa\in\mathcal{K}:\kappa<N\}\;\|q_{N}\|_{L^{2}}^{2}.

In order to deduce (3.5) from property (2), we must show that (3.3) converges to zero as κ\kappa\to\infty uniformly for qQq\in Q. This follows from (2.20). ∎

Theorem 3.2 (Defocusing case).

If QH+()Q\subset H^{\infty}_{+}(\mathbb{R}) is bounded and equicontinuous in L2()L^{2}(\mathbb{R}), then the totality of states Q={etJHq:qQ,t}Q^{*}=\{e^{tJ\nabla H_{-}}q:q\in Q,\ t\in\mathbb{R}\} reached under the defocusing (CCM) flow is also bounded and equicontinuous in L2()L^{2}(\mathbb{R}).

Proof.

By Proposition 2.2, we may chose κ0=κ0(Q)\kappa_{0}=\kappa_{0}(Q) so that q+κ\mathcal{L}_{q}+\kappa is positive definite for all qQq\in Q and κκ0\kappa\geq\kappa_{0}. Moreover, as noted in Lemma 2.7, this property remains true at all later times and R(κ,q(t))R(\kappa,q(t)) is positive definite. As we are in the defocusing case, (3.3) then shows that β[2](κ,q(t))β(κ,q(t))\beta^{[2]}(\kappa,q(t))\leq\beta(\kappa,q(t)). Combining this with the conservation law (2.24), we deduce that

(3.7) supqQβ[2](κ,q)supqQβ(κ,q)=supqQβ(κ,q).\sup_{q\in Q^{*}}\beta^{[2]}(\kappa,q)\leq\sup_{q\in Q^{*}}\beta(\kappa,q)=\sup_{q\in Q}\beta(\kappa,q).

By (3.5), the right-hand side converges to zero as κ\kappa\to\infty. Therefore, we conclude that the left-hand side converges to zero as κ\kappa\to\infty, which in view of Lemma 3.1 yields the equicontinuity of the set QQ^{*}.∎

The argument just presented does not adapt to the focusing case because the remainder (3.3) has an unfavorable sign.

Theorem 3.3 (Focusing case).

If QH+()Q\subset H^{\infty}_{+}(\mathbb{R}) is bounded and equicontinuous in L2()L^{2}(\mathbb{R}) and satisfies (1.6), then the set of orbits Q={etJH+q:qQ,t}Q^{*}=\{e^{tJ\nabla H_{+}}q:q\in Q,\ t\in\mathbb{R}\} reached under the focusing (CCM) flow is bounded and equicontinuous in L2()L^{2}(\mathbb{R}).

Proof.

Let

M:=sup{qL22:qQ}.M:=\sup\{\left\lVert q\right\rVert_{L^{2}}^{2}:q\in Q\}.

Under our hypotheses, M<2πM<2\pi. As the L2L^{2} norm is conserved by solutions to (CCM), we also have that

(3.8) supqQqL22M<2π.\displaystyle\sup_{q\in Q^{*}}\left\lVert q\right\rVert_{L^{2}}^{2}\leq M<2\pi.

In view of (3.8) and (2.1),

f,qC+(q¯f)=C+(q¯f)L22M2πf,0f,\displaystyle\bigl{\langle}f,qC_{+}(\overline{q}f)\bigr{\rangle}=\|C_{+}(\overline{q}f)\|^{2}_{L^{2}}\leq\tfrac{M}{2\pi}\langle f,\mathcal{L}_{0}f\rangle,

for every qQq\in Q^{*}. This implies

(3.9) q(1M2π)0and so also R(κ,q)2π2πMR0(κ)\displaystyle\mathcal{L}_{q}\geq\bigl{(}1-\tfrac{M}{2\pi}\bigr{)}\mathcal{L}_{0}\quad\text{and so also }\quad R(\kappa,q)\leq\tfrac{2\pi}{2\pi-M}R_{0}(\kappa)

for any κ>0\kappa>0 and qQq\in Q^{*}.

As QQ is equicontinuous in L2()L^{2}(\mathbb{R}), by Lemma 3.1 we may find an infinite set 𝒦2\mathcal{K}\subseteq 2^{\mathbb{N}} so that

(3.10) supqQq𝒦<.\displaystyle\sup_{q\in Q}\|q\|_{\mathcal{K}}<\infty.

Using (3.3) and (3.9), we may bound

κ𝒦|β(κ,q)β[2](κ,q)|Mκ𝒦κtr{R0(κ)qC+q¯R0(κ)qC+q¯R0(κ)}.\displaystyle\sum_{\kappa\in\mathcal{K}}\bigl{|}\beta(\kappa,q)-\beta^{[2]}(\kappa,q)\bigr{|}\lesssim_{M}\sum_{\kappa\in\mathcal{K}}\kappa\operatorname{tr}\bigl{\{}R_{0}(\kappa)qC_{+}\overline{q}R_{0}(\kappa)qC_{+}\overline{q}R_{0}(\kappa)\bigr{\}}.

To continue, we note that for κ1\kappa\geq 1 we have R0(κ)R0(1)R_{0}(\kappa)\leq R_{0}(1) as operators on L+2L^{2}_{+} and so

R0(κ)qC+q¯R0(κ)qC+q¯R0(κ)R0(κ)qC+q¯R0(1)qC+q¯R0(κ).R_{0}(\kappa)qC_{+}\overline{q}R_{0}(\kappa)qC_{+}\overline{q}R_{0}(\kappa)\leq R_{0}(\kappa)qC_{+}\overline{q}R_{0}(1)qC_{+}\overline{q}R_{0}(\kappa).

Thus, cycling the trace we obtain

κ𝒦|β(κ,q)β[2](κ,q)|\displaystyle\sum_{\kappa\in\mathcal{K}}\bigl{|}\beta(\kappa,q)-\beta^{[2]}(\kappa,q)\bigr{|} tr{R0(1)qC+q¯κ𝒦κR0(κ)2qC+q¯R0(1)}.\displaystyle\lesssim\operatorname{tr}\bigl{\{}\sqrt{R_{0}(1)}qC_{+}\overline{q}\sum_{\kappa\in\mathcal{K}}\kappa R_{0}(\kappa)^{2}qC_{+}\overline{q}\sqrt{R_{0}(1)}\bigr{\}}.

Observing that κ𝒦κR0(κ)2R0(1)\sum_{\kappa\in\mathcal{K}}\kappa R_{0}(\kappa)^{2}\leq R_{0}(1) and invoking (2.17), we deduce that

κ𝒦|β(κ,q)β[2](κ,q)|\displaystyle\sum_{\kappa\in\mathcal{K}}\bigl{|}\beta(\kappa,q)-\beta^{[2]}(\kappa,q)\bigr{|} tr{R0(1)qC+q¯R0(1)qC+q¯R0(1)}\displaystyle\lesssim\operatorname{tr}\bigl{\{}\sqrt{R_{0}(1)}qC_{+}\overline{q}R_{0}(1)qC_{+}\overline{q}\sqrt{R_{0}(1)}\bigr{\}}
C+q¯R0(1)q22qL24M2,\displaystyle\lesssim\left\lVert C_{+}\overline{q}R_{0}(1)q\right\rVert_{\mathfrak{I}_{2}}^{2}\lesssim\|q\|_{L^{2}}^{4}\lesssim M^{2},

uniformly for qQq\in Q^{*}. Using the conservation of β\beta (cf. Lemma 2.7) and (3.10), we conclude that

supqQκ𝒦β[2](κ,q)supqQκ𝒦β[2](κ,q)+M2<,\displaystyle\sup_{q\in Q^{*}}\sum_{\kappa\in\mathcal{K}}\beta^{[2]}(\kappa,q)\lesssim\sup_{q\in Q}\sum_{\kappa\in\mathcal{K}}\beta^{[2]}(\kappa,q)+M^{2}<\infty,

which in view of Lemma 3.1 establishes the equicontinuity of QQ^{*} in the L2()L^{2}(\mathbb{R}) topology. ∎

4. Explicit formula

The climax of this section is a proof of the explicit formula (1.12) in a restricted case:

Theorem 4.1.

For any H+()H^{\infty}_{+}(\mathbb{R}) solution q(t)q(t) to (CCM) with initial data satisfying xq0L2()\langle x\rangle q^{0}\in L^{2}(\mathbb{R}),

(4.1) q(t,z)=12πiI+((X+2tq0z)1q0)q(t,z)=\tfrac{1}{2\pi i}I_{+}\big{(}(X+2t\mathcal{L}_{q^{0}}-z)^{-1}q^{0}\bigr{)}

for all zz with Imz>0\operatorname{Im}z>0.

Ultimately, we will prove Theorem 1.7, which covers all L+2()L^{2}_{+}(\mathbb{R}) solutions by taking a limit of smooth solutions. As a foundation for this, we will work to understand the basic objects of the explicit formula in the full generality of qL+2()q\in L^{2}_{+}(\mathbb{R}), rather than just the case qH+()q\in H^{\infty}_{+}(\mathbb{R}) treated in Theorem 4.1.

We begin with a few necessary preliminaries. Our first lemma represents the key outgrowth of the fact that 𝒫1=0\mathcal{P}1=0. As the constant function 11 does not belong to the natural domain of 𝒫\mathcal{P}, we use the sequence of approximations χy(x):=iyx+iy\chi_{y}(x):=\tfrac{iy}{x+iy} which converge to unity (uniformly on compact sets) as yy\to\infty.

Lemma 4.2.

Let q(t)q(t) be a global H+()H^{\infty}_{+}(\mathbb{R}) solution of (CCM) and let U(t;t0)U(t;t_{0}) denote the associate family of unitary operators constructed in Proposition 2.3. Then

(4.2) lim supyU(t;0)χyχyL2=0\displaystyle\limsup_{y\to\infty}\|U^{*}(t;0)\chi_{y}-\chi_{y}\|_{L^{2}}=0

for any time tt\in\mathbb{R}. Recall that χy(x):=iyx+iy\chi_{y}(x):=\tfrac{iy}{x+iy}.

Proof.

Given tt\in\mathbb{R}, let Q={q(s):|s||t|}Q=\{q(s):|s|\leq|t|\}, which is evidently precompact in H+1()H^{1}_{+}(\mathbb{R}). We first claim that

(4.3) limy𝒫(q)χyL2=0uniformly for qQ.\lim_{y\to\infty}\big{\lVert}\mathcal{P}(q)\chi_{y}\big{\rVert}_{L^{2}}=0\quad\text{uniformly for $q\in Q$.}

To see this, we write

𝒫χy=iχy′′±2qC+(q¯(χy1))±2qC+(q¯χy),\mathcal{P}\chi_{y}=i\chi^{\prime\prime}_{y}\pm 2qC_{+}\big{(}\overline{q}^{\prime}(\chi_{y}-1)\big{)}\pm 2qC_{+}(\overline{q}\chi^{\prime}_{y}),

and use (2.1) to estimate

𝒫χyL2χy′′L2+qLq¯(χy1)L2+qLqL2χyL.\left\lVert\mathcal{P}\chi_{y}\right\rVert_{L^{2}}\lesssim\lVert\chi^{\prime\prime}_{y}\rVert_{L^{2}}+\left\lVert q\right\rVert_{L^{\infty}}\left\lVert\overline{q}^{\prime}(\chi_{y}-1)\right\rVert_{L^{2}}+\left\lVert q\right\rVert_{L^{\infty}}\left\lVert q\right\rVert_{L^{2}}\lVert\chi^{\prime}_{y}\rVert_{L^{\infty}}.

As QH+1()Q\subset H^{1}_{+}(\mathbb{R}) is precompact, we deduce (4.3).

Turning now to (4.2), we note that χyH+()\chi_{y}\in H^{\infty}_{+}(\mathbb{R}) and so

ddtχy,U(t;0)ψ0=𝒫(q(t))χy,U(t;0)ψ0\displaystyle\tfrac{d}{dt}\langle\chi_{y},U(t;0)\psi_{0}\rangle=-\langle\mathcal{P}(q(t))\chi_{y},U(t;0)\psi_{0}\rangle

for any ψ0L+2\psi_{0}\in L^{2}_{+}. Thus, by the Fundamental Theorem of Calculus,

U(t;0)χyχyL2|t||t|𝒫(q(s))χyL2𝑑s,\displaystyle\bigl{\|}U(t;0)^{*}\chi_{y}-\chi_{y}\bigr{\|}_{L^{2}}\leq\int_{-|t|}^{|t|}\|\mathcal{P}(q(s))\chi_{y}\|_{L^{2}}\,ds,

which converges to zero as yy\to\infty by (4.3). ∎

As noted above, χy(x)1\chi_{y}(x)\to 1 as yy\to\infty, uniformly on compact sets. Correspondingly, one may view χy,q\langle\chi_{y},q\rangle as an approximation for the total integral of qq. Building on this idea, we now define the unbounded linear functional I+I_{+} appearing in the (4.1). As we will see from the equivalent representations in (4.5), this is closer to twice the integral of qq.

Definition 4.3.

The (unbounded) linear functional I+I_{+} is defined by

(4.4) I+(f):=limy2π0yeyξf^(ξ)𝑑ξI_{+}(f):=\lim_{y\to\infty}\sqrt{2\pi}\int_{0}^{\infty}ye^{-y\xi}\,\widehat{f}(\xi)\,d\xi

with domain D(I+)D(I_{+}) given by the set of those fL+2()f\in L^{2}_{+}(\mathbb{R}) for which the limit exits.

By computing the Fourier transform of χy\chi_{y} and by noting that χy+χ¯y\chi_{y}+\smash[b]{\overline{\chi}_{y}} is equal to 2πy2\pi y times the Poisson kernel, we find the following equivalent definitions of I+I_{+}:

(4.5) I+(f)=limyχy,f=limy2y2x2+y2f(x)𝑑x=limy2πyf(iy)I_{+}(f)=\lim_{y\to\infty}\bigl{\langle}\chi_{y},f\bigr{\rangle}=\lim_{y\to\infty}\int\tfrac{2y^{2}}{x^{2}+y^{2}}f(x)\,dx=\lim_{y\to\infty}2\pi yf(iy)

and that D(I+)D(I_{+}) is comprised of those fL+2()f\in L^{2}_{+}(\mathbb{R}) for which these limits exist.

Let us now describe the operator XX appearing in (4.1). If we were working in L2()L^{2}(\mathbb{R}), then f(x)eitxf(x)f(x)\mapsto e^{-itx}f(x) defines a unitary group, the generator of which is multiplication by xx. (The factor i-i appearing here is the usual convention from quantum mechanics.) However, we are working in the Hardy space L+2()L^{2}_{+}(\mathbb{R}) and for t>0t>0, multiplication by eitxe^{-itx} does not preserve this space. The natural analogue on the Hardy space L+2()L^{2}_{+}(\mathbb{R}) is

(4.6) eitXf=C+(eitxf)=12π0eiξxf^(ξ+t)𝑑ξ,e^{-itX}f=C_{+}\bigl{(}e^{-itx}f\bigr{)}=\tfrac{1}{\sqrt{2\pi}}\int_{0}^{\infty}e^{i\xi x}\widehat{f}(\xi+t)\,d\xi,

and we define the operator XX as the generator of this semigroup:

D(X)={fL+2():f^H1([0,))}andXf^(ξ)=idf^dξ(ξ)forfD(X).D(X)=\bigl{\{}f\in L^{2}_{+}(\mathbb{R}):\widehat{f}\in H^{1}\bigl{(}[0,\infty)\bigr{)}\bigr{\}}\quad\text{and}\quad\widehat{Xf}(\xi)=i\tfrac{d{\widehat{f}}}{d\xi}(\xi)\quad\text{for}\quad f\in D(X).

The semigroup (4.6) is clearly a (strongly continuous) contraction semigroup. The famous Hille–Yoshida Theorem identifies the generators of such semigroups; see [EN, Reed1975]. Unfortunately, the nomenclature for these generators is not fully settled: they may be accretive or dissipative, there is a sign ambiguity, as well as the choice of whether or not to include the imaginary unit ii. Evidently, one should select the convention that best describes the operators one is working with and for this reason, we adopt the following:

Convention 4.4.

An operator TT on a Hilbert space will be called accretive if

(4.7) Imf,Tf0for allfD(T).\operatorname{Im}\langle f,Tf\rangle\leq 0\quad\text{for all}\quad f\in D(T).

It is further maximally accretive if it admits no proper accretive extension.

In this way, TT is accretive in our sense if and only if iTiT is accretive in the sense of [Reed1975] if and only if iT-iT is dissipative in the sense of [EN]. Moreover, the Hille–Yoshida Theorem then says that an operator TT is maximally accretive if and only if eitTe^{-itT} defines a contraction semigroup. In particular, the operator XX introduced above is maximally accretive.

The spectrum of XX constitutes the closed lower half-plane. For Imz>0\operatorname{Im}z>0, the resolvent is given by

(4.8) (Xz)1f=f(x)f(z)xz(X-z)^{-1}f=\tfrac{f(x)-f(z)}{x-z}

where f(z)f(z) is defined by (1.14). Together with the functional I+I_{+}, this gives rise to the following formulation of the Cauchy integral formula:

(4.9) f(z)=12πiI+((Xz)1f)=limy12πiχy,(Xz)1ff(z)=\tfrac{1}{2\pi i}I_{+}\bigl{(}(X-z)^{-1}f\bigr{)}=\lim_{y\to\infty}\tfrac{1}{2\pi i}\bigl{\langle}\chi_{y},(X-z)^{-1}f\bigr{\rangle}

valid for all fL+2()f\in L^{2}_{+}(\mathbb{R}) and Imz>0\operatorname{Im}z>0.

The commutator of 𝒫\mathcal{P} and the operator XX will be important for deriving the explicit formula (4.1). As a stepping stone, we record the commutator of XX with a generic Toeplitz operator fC+(gf)f\mapsto C_{+}(gf). This may be obtained from straightforward computations in Fourier variables. It also arises naturally in the setting of the Benjamin–Ono equation, see [Sun2021]*Lem. 3.1.

Lemma 4.5.

If gH()g\in H^{\infty}(\mathbb{R}) and fD(X)f\in D(X), then C+(gf)D(X)C_{+}(gf)\in D(X) and

(4.10) [X,C+g]f=i2πI+(f)C+g.[X,C_{+}g]f=\tfrac{i}{2\pi}I_{+}(f)C_{+}g.

We are now ready to compute the commutator of 𝒫\mathcal{P} with XX:

Lemma 4.6.

If qH+()q\in H^{\infty}_{+}(\mathbb{R}) and xqL2()\langle x\rangle q\in L^{2}(\mathbb{R}), then

(4.11) [X,𝒫]=2q.[X,\mathcal{P}]=2\mathcal{L}_{q}.
Proof.

For fL+2()f\in L^{2}_{+}(\mathbb{R}) with xfH2()\langle x\rangle f\in H^{2}(\mathbb{R}), we use Lemma 4.5 to compute

[X,𝒫]f=2if±iπI+{C+(q¯f)}q2qC+(q¯f).[X,\mathcal{P}]f=-2if^{\prime}\pm\tfrac{i}{\pi}I_{+}\big{\{}C_{+}(\overline{q}f)^{\prime}\big{\}}q\mp 2qC_{+}(\overline{q}f).

As derivatives vanish at zero frequency, we recognize that the right-hand side above is equal to 2qf2\mathcal{L}_{q}f. ∎

We now turn our attention to the operator X+2tqX+2t\mathcal{L}_{q}, whose resolvent lies at the center of the explicit formula. We begin with the case q0q\equiv 0.

As a a preliminary notion of X+2t0X+2t\mathcal{L}_{0}, we observe that

(4.12) (Xf+2t0f)^(ξ)\displaystyle(Xf+2t\mathcal{L}_{0}f)\,\widehat{\ }\,(\xi) =iξf^(ξ)+2tξf^(ξ)\displaystyle=i\partial_{\xi}\widehat{f}(\xi)+2t\xi\widehat{f}(\xi)
=eitξ2iξ[eitξ2f^(ξ)]for allfD(X)D(0).\displaystyle=e^{it\xi^{2}}i\partial_{\xi}\bigl{[}e^{-it\xi^{2}}\widehat{f}(\xi)\bigr{]}\quad\text{for all}\quad f\in D(X)\cap D(\mathcal{L}_{0}).

While it is evident that with this domain X+2t0X+2t\mathcal{L}_{0} is accretive, it is not maximally accretive. This can be remedied by taking the closure:

Lemma 4.7.

The closure of the naive sum (4.12) is given by

(4.13) [(X+2t0)f]^(ξ)=eitξ2iξ[eitξ2f^(ξ)],[(X+2t\mathcal{L}_{0})f]\,\widehat{\ }\,(\xi)=e^{it\xi^{2}}i\partial_{\xi}\big{[}e^{-it\xi^{2}}\widehat{f}(\xi)\big{]},

with domain

(4.14) D(X+2t0)={fL+2():eitξ2f^H1([0,))}.D(X+2t\mathcal{L}_{0})=\{f\in L^{2}_{+}(\mathbb{R}):e^{-it\xi^{2}}\widehat{f}\in H^{1}([0,\infty))\}.

The operator X+2t0X+2t\mathcal{L}_{0} is maximally accretive, generating the contraction semigroup

(4.15) [eis(X+2t0)f]^(ξ)=eis2te2ist(ξ+s)f^(ξ+s),\big{[}e^{-is(X+2t\mathcal{L}_{0})}f\big{]}\,\widehat{\ }\,(\xi)=e^{is^{2}t}e^{-2ist(\xi+s)}\widehat{f}(\xi+s),

and is unitarily conjugated to XX by the free Schrödinger flow:

(4.16) X+2t0=eitΔXeitΔ.X+2t\mathcal{L}_{0}=e^{-it\Delta}Xe^{it\Delta}.
Proof.

It is perhaps easiest to work backwards. One first observes that (4.15) defines a strongly continuous semigroup of contractions and then that the generator is the operator defined by (4.13) and (4.14) and satisfies (4.16).

As the last step, we check that this generator is indeed the closure of X+2t0X+2t\mathcal{L}_{0} when defined on D(X)D(0)D(X)\cap D(\mathcal{L}_{0}). As maximally accretive operators are closed, this amounts to the elementary task of verifying that our operator is contained in the closure of the naive sum. ∎

With the proper meaning of X+2t0X+2t\mathcal{L}_{0} now set, we may turn our attention to its resolvent:

Proposition 4.8.

For tt\in\mathbb{R} and Imz>0\operatorname{Im}z>0,

(4.17) A0(t,z):=(X+2t0z)1satisfiesA0op1Imz.A_{0}(t,z):=(X+2t\mathcal{L}_{0}-z)^{-1}\quad\text{satisfies}\quad\left\lVert A_{0}\right\rVert_{\mathrm{op}}\leq\tfrac{1}{\operatorname{Im}z}.

For any t0t\neq 0, we have

(4.18) A0(t,z)L+2L++A0(t,z)L+1L+2\displaystyle\left\lVert A_{0}(t,z)\right\rVert_{L^{2}_{+}\to L^{\infty}_{+}}+\left\lVert A_{0}(t,z)\right\rVert_{L^{1}_{+}\to L^{2}_{+}} (|t|Imz)12,\displaystyle\lesssim(|t|\operatorname{Im}z)^{-\frac{1}{2}},
(4.19) A0(t,z)L+1L+\displaystyle\left\lVert A_{0}(t,z)\right\rVert_{L^{1}_{+}\to L^{\infty}_{+}} |t|1.\displaystyle\lesssim|t|^{-1}.

The range of A0A_{0} lies in the domain of I+I_{+},

(4.20) I+(A0(t,z)f)=2πi[eitΔf](z)for allt,Imz>0,I_{+}\bigl{(}A_{0}(t,z)f\bigr{)}=2\pi i\big{[}e^{it\Delta}f\big{]}(z)\quad\text{for all}\quad t\in\mathbb{R},\ \ \operatorname{Im}z>0,

and the composition I+A0I_{+}\circ A_{0} is bounded:

(4.21) |I+(A0(t,z)f)|min{|Imz|1/2fL2,|t|1/2fL1}.\displaystyle\bigl{|}I_{+}\bigl{(}A_{0}(t,z)f\bigr{)}\bigr{|}\lesssim\min\bigl{\{}|\operatorname{Im}z|^{-1/2}\|f\|_{L^{2}},\ |t|^{-1/2}\|f\|_{L^{1}}\bigr{\}}.
Proof.

The existence of the inverse (4.17) for Imz>0\operatorname{Im}z>0 and the associated norm bound are basic consequences of the fact that X+2t0X+2t\mathcal{L}_{0} is maximally accretive.

From (4.16) we see that

(4.22) A0(t,z)=eitΔ(Xz)1eitΔ.A_{0}(t,z)=e^{-it\Delta}(X-z)^{-1}e^{it\Delta}.

For t0t\neq 0, the free Schrödinger propagator is given by the explicit formula

(4.23) [eitΔf](x)=14πitei(xy)2/4tf(y)𝑑y,\big{[}e^{it\Delta}f\big{]}(x)=\tfrac{1}{\sqrt{4\pi it}}\int e^{i(x-y)^{2}/4t}f(y)\,dy,

where we use the principal branch of the square root, so that the argument of 4πit\sqrt{4\pi it} is π4sgn(t)\tfrac{\pi}{4}\operatorname{sgn}(t). From this, one easily derives the standard dispersive estimates

(4.24) eitΔLpp1Lp|t|1p12for 2p.\big{\lVert}e^{it\Delta}\big{\rVert}_{L^{\frac{p}{p-1}}\to L^{p}}\lesssim|t|^{\frac{1}{p}-\frac{1}{2}}\quad\text{for }2\leq p\leq\infty.

Combining this with (1.15), we deduce

(4.25) |[eitΔf](z)|min{|t|12fL1,|Imz|12fL2}\displaystyle\bigl{|}[e^{it\Delta}f](z)\bigr{|}\lesssim\min\bigl{\{}|t|^{-\frac{1}{2}}\|f\|_{L^{1}},|\operatorname{Im}z|^{-\frac{1}{2}}\|f\|_{L^{2}}\bigr{\}}

for all t0t\neq 0 and Imz>0\operatorname{Im}z>0.

We now turn to (4.18). Using that (1xz)=2πieiξzχξ<0\mathcal{F}(\frac{1}{x-z})=\sqrt{2\pi}ie^{-i\xi z}\chi_{\xi<0} and the Van der Corput Lemma in the form given in [bigStein, Corollary VIII.1.2], we may estimate

(4.26) eitΔ(1xz)L0eixξ+itξ2iξRezeξImz𝑑ξL|t|12.\displaystyle\bigl{\|}e^{-it\Delta}\bigl{(}\tfrac{1}{x-z}\bigr{)}\bigr{\|}_{L^{\infty}}\lesssim\Bigl{\|}\int_{-\infty}^{0}e^{ix\xi+it\xi^{2}-i\xi\operatorname{Re}z}e^{\xi\operatorname{Im}z}\,d\xi\Bigr{\|}_{L^{\infty}}\lesssim|t|^{-\frac{1}{2}}.

Employing (4.22), (4.25), and (4.8), we may thus bound

A0(t,z)fL\displaystyle\left\lVert A_{0}(t,z)f\right\rVert_{L^{\infty}} |t|12(eitΔf)(x)xzL1+eitΔ(1xz)L|(eitΔf)(z)|\displaystyle\lesssim|t|^{-\frac{1}{2}}\bigl{\|}\tfrac{(e^{it\Delta}f)(x)}{x-z}\bigr{\|}_{L^{1}}+\bigl{\|}e^{-it\Delta}\bigl{(}\tfrac{1}{x-z}\bigr{)}\bigr{\|}_{L^{\infty}}\bigl{|}(e^{it\Delta}f)(z)\bigr{|}
|t|12eitΔfL2(xz)1L2+(|t|Imz)12fL2\displaystyle\lesssim|t|^{-\frac{1}{2}}\|e^{it\Delta}f\|_{L^{2}}\|(x-z)^{-1}\|_{L^{2}}+(|t|\operatorname{Im}z)^{-\frac{1}{2}}\|f\|_{L^{2}}
(|t|Imz)12fL2.\displaystyle\lesssim(|t|\operatorname{Im}z)^{-\frac{1}{2}}\|f\|_{L^{2}}.

Proceeding similarly, we obtain

A0(t,z)fL2\displaystyle\|A_{0}(t,z)f\|_{L^{2}} (Xz)1eitΔfL2\displaystyle\lesssim\bigl{\|}(X-z)^{-1}e^{it\Delta}f\bigr{\|}_{L^{2}}
(xz)1L2[eitΔfL+|(eitΔf)(z)|]\displaystyle\lesssim\|(x-z)^{-1}\|_{L^{2}}\bigl{[}\|e^{it\Delta}f\|_{L^{\infty}}+\bigl{|}(e^{it\Delta}f)(z)\bigr{|}\bigr{]}
(|t|Imz)12fL1,\displaystyle\lesssim(|t|\operatorname{Im}z)^{-\frac{1}{2}}\|f\|_{L^{1}},

which completes the proof of (4.18).

Employing (4.22), (4.23), and (4.8), a straightforward computation yields

[A0(t,z)f](x)\displaystyle[A_{0}(t,z)f](x) =i2|t|yt<xtei(x2y2)4t+i(xy)z2tf(y)𝑑y(eitΔf)(z)[eitΔ(1z)](x).\displaystyle=\tfrac{i}{2|t|}\int_{\frac{y}{t}<\frac{x}{t}}e^{-\frac{i(x^{2}-y^{2})}{4t}+\frac{i(x-y)z}{2t}}f(y)\,dy-\bigl{(}e^{it\Delta}f\bigr{)}(z)\cdot\Bigl{[}e^{-it\Delta}\bigl{(}\tfrac{1}{\cdot-z}\bigr{)}\Bigr{]}(x).

Claim (4.19) follows from this, (4.25) and (4.26).

By definition, the resolvent A0(t,z)A_{0}(t,z) carries fL+2f\in L^{2}_{+} into the domain (4.14), which is clearly a subset of D(I+)D(I_{+}). The identity (4.20) follows immediately from (4.22) and (4.9). Together with (4.25), this identity shows that I+A0I_{+}\circ A_{0} satisfies (4.21). ∎

This completes our discussion of X+2t0X+2t\mathcal{L}_{0}. We turn now to the more difficult question of realizing X+2tqX+2t\mathcal{L}_{q} as a maximally accretive operator.

If qLL+2q\in L^{\infty}\cap L^{2}_{+}, then qC+q¯qC_{+}\overline{q} is an L2L^{2}-bounded self-adjoint operator. Under this hypothesis, X+2tqX+2t\mathcal{L}_{q} is naturally defined on (4.14) and is maximally accretive; see, for example, [EN, §III.1]. In this way, we may define

(4.27) A(t,z;q):=(X+2tqz)1for alltandImz>0,\displaystyle A(t,z;q):=(X+2t\mathcal{L}_{q}-z)^{-1}\quad\text{for all}\quad t\in\mathbb{R}\quad\text{and}\quad\operatorname{Im}z>0,

whenever qL+2()L()q\in L^{2}_{+}(\mathbb{R})\cap L^{\infty}(\mathbb{R}).

For general qL+2()q\in L^{2}_{+}(\mathbb{R}), the construction of this operator is not so simple. Indeed, we find ourselves unable to apply the textbook theorems for perturbations of accretive operators, such as the analogue of the Kato–Rellich Theorem. Rather, we will build the operator as a limit of operators with bounded potentials qnq_{n} in concert with the abstract theory developed by Trotter [Trotter] and Kato [Kato]. Our first step in this direction is the following lemma, which we will also need for the proof of Proposition 5.1:

Lemma 4.9.

Fix T>0T>0 and let QL+2()Q\subset L^{2}_{+}(\mathbb{R}) be bounded and equicontinuous. For every ε>0\varepsilon>0 there exists b=b(ε,Q)b=b(\varepsilon,Q) so that Imzb\operatorname{Im}z\geq b implies both

(4.28) 2tC+g¯A0(t,z)qopεand|t|C+q¯A0(t,z)opA0(t,z)qopεImz\big{\lVert}2t\,C_{+}\overline{g}A_{0}(t,z)q\big{\rVert}_{\mathrm{op}}\leq\varepsilon\quad\text{and}\quad|t|\big{\lVert}C_{+}\overline{q}A_{0}(t,z)\big{\rVert}_{\mathrm{op}}\big{\lVert}A_{0}(t,z)q\big{\rVert}_{\mathrm{op}}\leq\tfrac{\varepsilon}{\operatorname{Im}z}

uniformly for |t|T|t|\leq T and g,qQg,q\in Q.

Proof.

We begin with the first claim in (4.28), decomposing q=qN+q>Nq=q_{\leq N}+q_{>N} and similarly g=gN+g>Ng=g_{\leq N}+g_{>N}, for a frequency cutoff N1N\geq 1 that will be chosen shortly. In this way, (4.19) yields

2|t|C+g¯A0q>N\displaystyle 2|t|\big{\lVert}C_{+}\overline{g}A_{0}q_{>N} +C+g¯>NA0qNop\displaystyle+C_{+}\overline{g}_{>N}A_{0}q_{\leq N}\big{\rVert}_{\mathrm{op}}
(4.29) |t|A0L+1L+[q>NL2gL2+qL2g>NL2]\displaystyle\lesssim|t|\|A_{0}\|_{L^{1}_{+}\to L^{\infty}_{+}}\big{[}\left\lVert q_{>N}\right\rVert_{L^{2}}\left\lVert g\right\rVert_{L^{2}}+\left\lVert q\right\rVert_{L^{2}}\left\lVert g_{>N}\right\rVert_{L^{2}}\big{]}
q>NL2gL2+qL2g>NL2.\displaystyle\lesssim\left\lVert q_{>N}\right\rVert_{L^{2}}\left\lVert g\right\rVert_{L^{2}}+\left\lVert q\right\rVert_{L^{2}}\left\lVert g_{>N}\right\rVert_{L^{2}}.

As QQ is bounded and equicontinuous, we may choose N=N(Q)N=N(Q) large to ensure that LHS(4) is bounded by ε2\frac{\varepsilon}{2}.

For the remaining low-low frequency contribution, we use Bernstein’s inequality to estimate

2tC+g¯NA0qNop2|t|A0opgNLqNLTNImzgL2qL2.\big{\lVert}2tC_{+}\overline{g}_{\leq N}A_{0}q_{\leq N}\big{\rVert}_{\mathrm{op}}\leq 2|t|\left\lVert A_{0}\right\rVert_{\mathrm{op}}\left\lVert g_{\leq N}\right\rVert_{L^{\infty}}\left\lVert q_{\leq N}\right\rVert_{L^{\infty}}\lesssim\tfrac{TN}{\operatorname{Im}z}\left\lVert g\right\rVert_{L^{2}}\left\lVert q\right\rVert_{L^{2}}.

As QQ is bounded and N(Q)N(Q) is finite, we may choose b1b\geq 1 sufficiently large so as to render this contribution smaller that ε2\frac{\varepsilon}{2}.

The treatment of the second claim in (4.28) follows a parallel path:

C+q¯A0op\displaystyle\left\lVert C_{+}\overline{q}A_{0}\right\rVert_{\mathrm{op}} q>NL2A0L+2L++qNLA0op,\displaystyle\lesssim\|q_{>N}\|_{L^{2}}\|A_{0}\|_{L^{2}_{+}\to L^{\infty}_{+}}+\|q_{\leq N}\|_{L^{\infty}}\|A_{0}\|_{\mathrm{op}},
A0qop\displaystyle\left\lVert A_{0}q\right\rVert_{\mathrm{op}} q>NL2A0L+1L+2+qNLA0op,\displaystyle\lesssim\|q_{>N}\|_{L^{2}}\|A_{0}\|_{L^{1}_{+}\to L^{2}_{+}}+\|q_{\leq N}\|_{L^{\infty}}\|A_{0}\|_{\mathrm{op}},

and so by (4.17) and (4.18),

|t|C+q¯A0opA0qop\displaystyle|t|\cdot\left\lVert C_{+}\overline{q}A_{0}\right\rVert_{\mathrm{op}}\cdot\left\lVert A_{0}q\right\rVert_{\mathrm{op}} 1Imz[q>NL22+NTImzqL22].\displaystyle\lesssim\tfrac{1}{\operatorname{Im}z}\Bigl{[}\|q_{>N}\|_{L^{2}}^{2}+\tfrac{NT}{\operatorname{Im}z}\|q\|_{L^{2}}^{2}\Bigr{]}.

As previously, we first choose N=N(Q)N=N(Q) and then b(ε,Q)b(\varepsilon,Q) to render this contribution acceptable. ∎

We are now prepared to construct the operator X+2tqX+2t\mathcal{L}_{q} for qL+2()q\in L^{2}_{+}(\mathbb{R}), as well as its resolvent. Further mapping properties of the resolvent will be elaborated in Proposition 4.11.

Proposition 4.10.

Fix qL+2()q\in L^{2}_{+}(\mathbb{R}). For any tt\in\mathbb{R}, Imz>0\operatorname{Im}z>0, and any sequence qnL+2()L+()q_{n}\in L^{2}_{+}(\mathbb{R})\cap L^{\infty}_{+}(\mathbb{R}) converging to qq in L2L^{2}-sense, the operators

(4.30) An(t,z):=(X+2t02tqnC+q¯nz)1A_{n}(t,z):=(X+2t\mathcal{L}_{0}\mp 2tq_{n}C_{+}\overline{q}_{n}-z)^{-1}

converge in norm to an operator A(t,z;q)A(t,z;q) that satisfies

(4.31) A(t,z;q)(Xf+2tqfzf)=fwheneverfD(X)D(0).A(t,z;q)(Xf+2t\mathcal{L}_{q}f-zf)=f\quad\text{whenever}\quad f\in D(X)\cap D(\mathcal{L}_{0}).

Moreover, zA(t,z)z\mapsto A(t,z) is the resolvent of a maximally accretive operator, which we denote X+2tqX+2t\mathcal{L}_{q}. Thus

(4.32) A(t,z;q)=(X+2tqz)1satisfiesA(t,z;q)op1ImzA(t,z;q)=(X+2t\mathcal{L}_{q}-z)^{-1}\quad\text{satisfies}\quad\left\lVert A(t,z;q)\right\rVert_{\mathrm{op}}\leq\tfrac{1}{\operatorname{Im}z}

for any tt\in\mathbb{R}, Imz>0\operatorname{Im}z>0, and any qL+2()q\in L^{2}_{+}(\mathbb{R}).

Proof.

To treat the sequence qnq_{n} and its limit qq in a parallel fashion, it is convenient to set q=qq_{\infty}=q. Convergence of the sequence guarantees that Q:={qn:1n}Q:=\{q_{n}:1\leq n\leq\infty\} is L2L^{2} precompact and so bounded and equicontinuous. By Lemma 4.9, we may choose b(ε)b(\varepsilon) so that

(4.33) sup1n,m2tC+q¯nA0(t,z)qmopεwheneverImzb(ε).\displaystyle\sup_{1\leq n,m\leq\infty}\ \big{\lVert}2tC_{+}\smash[b]{\overline{q}}_{n}A_{0}(t,z)q_{m}\big{\rVert}_{\mathrm{op}}\leq\varepsilon\quad\text{whenever}\quad\operatorname{Im}z\geq b(\varepsilon).

We particularly single out the value corresponding to ε=12\varepsilon=\tfrac{1}{2}, for which we write b0b_{0}.

For Imzb0\operatorname{Im}z\geq b_{0} and any nn\in\mathbb{N}, the estimate (4.33) guarantees convergence of the resolvent series

(4.34) An(t,z)=A0(t,z)±2tA0(t,z)qn0[±2tC+q¯nA0(t,z)qn]C+q¯nA0(t,z).A_{n}(t,z)=A_{0}(t,z)\pm 2tA_{0}(t,z)q_{n}\sum_{\ell\geq 0}\big{[}\pm 2tC_{+}\smash[b]{\overline{q}}_{n}A_{0}(t,z)q_{n}\big{]}^{\ell}C_{+}\overline{q}_{n}A_{0}(t,z).

Note also that by (4.18), we have

(4.35) sup1n|t|12A0(t,z)qnop+|t|12C+q¯nA0(t,z)opQ(Imz)12.\sup_{1\leq n\leq\infty}|t|^{\frac{1}{2}}\|A_{0}(t,z)q_{n}\|_{\mathrm{op}}+|t|^{\frac{1}{2}}\|C_{+}\smash[b]{\overline{q}}_{n}A_{0}(t,z)\|_{\mathrm{op}}\lesssim_{Q}(\operatorname{Im}z)^{-\frac{1}{2}}.

It follows from the preceding discussion that for Imzb0\operatorname{Im}z\geq b_{0} we may define a bounded operator A(t,z)A_{\infty}(t,z) by setting n=n=\infty in (4.34) and, by a simple telescoping argument, we see that

(4.36) An(t,z)A(t,z)op0and soA(t,z)op1Imz\|A_{n}(t,z)-A_{\infty}(t,z)\|_{\mathrm{op}}\to 0\quad\text{and so}\quad\|A_{\infty}(t,z)\|_{\mathrm{op}}\leq\tfrac{1}{\operatorname{Im}z}

for all Imzb0\operatorname{Im}z\geq b_{0} and tt\in\mathbb{R}.

As the resolvents An(t,z)A_{n}(t,z) converge at one point z+z\in\mathbb{C}_{+}, they must converge at all points in +\mathbb{C}_{+}. However, it is not true in general that the limit of resolvents of accretive operators is the resolvent of an accretive operator. Rather it is merely a pseudo-resolvent; see [Kato]. The kernel and range of a pseudo-resolvent are independent of the spectral parameter; however, it is only for a true resolvent that the kernel is trivial and the range is dense.

To verify that A(z)A_{\infty}(z) is a true resolvent, we will show that

(4.37) limyiyA(iy)f=ffor allfL+2().\lim_{y\to\infty}iyA_{\infty}(iy)f=-f\quad\text{for all}\quad f\in L^{2}_{+}(\mathbb{R}).

This is always true for the resolvent of a maximally accretive operator (as one may see by integrating the semigroup); in particular,

(4.38) limyiyA0(iy)f=ffor allfL+2().\lim_{y\to\infty}iyA_{0}(iy)f=-f\quad\text{for all}\quad f\in L^{2}_{+}(\mathbb{R}).

By comparison, Lemma 4.9 shows that

lim supyy|t|A0(iy)qopC+q¯A0(iy)op0[|t|C+q¯A0(iy)qop]=0\displaystyle\limsup_{y\to\infty}y|t|\bigl{\|}A_{0}(iy)q\|_{\mathrm{op}}\bigl{\|}C_{+}\overline{q}A_{0}(iy)\|_{\mathrm{op}}\sum_{\ell\geq 0}\Big{[}|t|\bigl{\|}C_{+}{\overline{q}}A_{0}(iy)q\bigr{\|}_{\mathrm{op}}\Big{]}^{\ell}=0

for any fixed qL+2()q\in L^{2}_{+}(\mathbb{R}). Thus (4.37) follows from (4.38) and (4.34).

We may now invoke the abstract Trotter–Kato theory; see [EN]. The resolvents of the maximally accretive operators X+2tqnX+2t\mathcal{L}_{q_{n}} converge for all zz with Imz>0\operatorname{Im}z>0. Moreover, (4.37) ensures that the limit A(z)A_{\infty}(z) is injective with dense range. It then follows that A(z)A_{\infty}(z) is the resolvent of a densely defined operator. By (4.36), the operator is maximally accretive.

Finally, we verify (4.31), which illustrates the strong connection between the maximally accretive operator just constructed and the naive notion of the sum X+2tqX+2t\mathcal{L}_{q} defined on D(X)D(0)D(X)\cap D(\mathcal{L}_{0}). By definition of the resolvent, (4.31) holds if we replace qq by the bounded approximations qnq_{n}. To send nn\to\infty, we first note that arguing as in the proof of Proposition 2.2 we have

C+(q¯f)L[C+(q¯f)]^L1Mq^L2ηf^L2qL2fH1\displaystyle\|C_{+}(\overline{q}f)\|_{L^{\infty}}\lesssim\|[C_{+}(\overline{q}f)]\,\widehat{\ }\,\|_{L^{1}}\lesssim\|\textsf{M}\widehat{q}\|_{L^{2}}\|\langle\eta\rangle\widehat{f}\|_{L^{2}}\lesssim\|q\|_{L^{2}}\|f\|_{H^{1}}

and consequently,

[qC+q¯qnC+q¯n]fL2\displaystyle\big{\lVert}[qC_{+}\overline{q}-q_{n}C_{+}\smash[b]{\overline{q}}_{n}]f\big{\rVert}_{L^{2}} qL2C+[q¯q¯n]fL+qqnL2C+q¯nfL\displaystyle\lesssim\|q\|_{L^{2}}\|C_{+}[\overline{q}-\smash[b]{\overline{q}}_{n}]f\|_{L^{\infty}}+\|q-q_{n}\|_{L^{2}}\|C_{+}\smash[b]{\overline{q}}_{n}f\|_{L^{\infty}}
[qnL2+qL2]qqnL2fH1.\displaystyle\lesssim\bigl{[}\|q_{n}\|_{L^{2}}+\|q\|_{L^{2}}\bigr{]}\|q-q_{n}\|_{L^{2}}\|f\|_{H^{1}}.

This converges to zero as nn\to\infty, completing the proof of (4.31). ∎

Proposition 4.11.

Given a bounded and equicontinuous set QL+2()Q\subset L^{2}_{+}(\mathbb{R}), the operators

(4.39) A(t,z;q)=(X+2tqz)1A(t,z;q)=(X+2t\mathcal{L}_{q}-z)^{-1}

introduced in Proposition 4.10 satisfy

(4.40) |t|12AL+2L++|t|12AL+1L+21Imz+1Imz\displaystyle|t|^{\frac{1}{2}}\left\lVert A\right\rVert_{L^{2}_{+}\to L^{\infty}_{+}}+|t|^{\frac{1}{2}}\left\lVert A\right\rVert_{L^{1}_{+}\to L^{2}_{+}}\lesssim\tfrac{1}{\sqrt{\operatorname{Im}z}}+\tfrac{1}{\operatorname{Im}z}
(4.41) |t|AL+1L+1+1Imz\displaystyle|t|\left\lVert A\right\rVert_{L^{1}_{+}\to L^{\infty}_{+}}\lesssim 1+\tfrac{1}{\operatorname{Im}z}

uniformly for qQq\in Q, |t|T|t|\leq T, and Imz>0\operatorname{Im}z>0. Likewise, for any fL+2()f\in L^{2}_{+}(\mathbb{R}),

(4.42) |I+[A(t,z;q)f]|[1Imz+1Imz]fL2.\displaystyle\bigl{|}I_{+}\bigl{[}A(t,z;q)f\bigr{]}\bigr{|}\lesssim\Bigl{[}\tfrac{1}{\sqrt{\operatorname{Im}z}}+\tfrac{1}{\operatorname{Im}z}\Bigr{]}\|f\|_{L^{2}}.
Proof.

In view of Lemma 4.9, one may choose b0=b0(Q)b_{0}=b_{0}(Q) so that

(4.43) supqQ2tC+q¯A0(t,z)qop12whenever |t|T and Imzb0.\sup_{q\in Q}\big{\lVert}2tC_{+}\overline{q}A_{0}(t,z)q\big{\rVert}_{\mathrm{op}}\leq\tfrac{1}{2}\quad\text{whenever $|t|\leq T$ and $\operatorname{Im}z\geq b_{0}$.}\quad

In this regime, A(t,z;q)A(t,z;q) is given by the series expansion (4.34) and correspondingly,

A(t,z;q)L+2L+\displaystyle\left\lVert A(t,z;q)\right\rVert_{L^{2}_{+}\to L^{\infty}_{+}} A0(t,z)L+2L++|t|A0(t,z)qL+2L+C+q¯A0(t,z)op\displaystyle\lesssim\left\lVert A_{0}(t,z)\right\rVert_{L^{2}_{+}\to L^{\infty}_{+}}+|t|\;\!\|A_{0}(t,z)q\|_{L^{2}_{+}\to L^{\infty}_{+}}\|C_{+}\overline{q}A_{0}(t,z)\|_{\mathrm{op}}
[1+|t|qL22A0(t,z)L+1L+]A0(t,z)L+2L+.\displaystyle\lesssim\Bigl{[}1+|t|\;\!\|q\|_{L^{2}}^{2}\|A_{0}(t,z)\|_{L^{1}_{+}\to L^{\infty}_{+}}\Bigr{]}\left\lVert A_{0}(t,z)\right\rVert_{L^{2}_{+}\to L^{\infty}_{+}}.

Similarly, we obtain

A(t,z;q)L+1L+2\displaystyle\left\lVert A(t,z;q)\right\rVert_{L^{1}_{+}\to L^{2}_{+}} [1+|t|qL22A0(t,z)L+1L+]A0(t,z)L+1L+2\displaystyle\lesssim\Bigl{[}1+|t|\;\!\|q\|_{L^{2}}^{2}\|A_{0}(t,z)\|_{L^{1}_{+}\to L^{\infty}_{+}}\Bigr{]}\left\lVert A_{0}(t,z)\right\rVert_{L^{1}_{+}\to L^{2}_{+}}

and

A(t,z;q)L+1L+\displaystyle\left\lVert A(t,z;q)\right\rVert_{L^{1}_{+}\to L^{\infty}_{+}} [1+|t|qL22A0(t,z)L+1L+]A0(t,z)L+1L+.\displaystyle\lesssim\Bigl{[}1+|t|\;\!\|q\|_{L^{2}}^{2}\|A_{0}(t,z)\|_{L^{1}_{+}\to L^{\infty}_{+}}\Bigr{]}\left\lVert A_{0}(t,z)\right\rVert_{L^{1}_{+}\to L^{\infty}_{+}}.

The claims (4.40) and (4.41) then follow from (4.18) and (4.19), albeit only for Imzb0\operatorname{Im}z\geq b_{0}.

As the resolvent of a maximally accretive operator,

(4.44) A(z;q)=A(z;q)+(zz)A(z;q)A(z;q)andA(z;q)op1Imz.\displaystyle A(z;q)=A(z^{\prime};q)+(z-z^{\prime})A(z^{\prime};q)A(z;q)\quad\text{and}\quad\|A(z;q)\|_{\mathrm{op}}\lesssim\tfrac{1}{\operatorname{Im}z}.

This extends (4.40) and (4.41) to general Imz>0\operatorname{Im}z>0 by choosing z=Rez+ib0z^{\prime}=\operatorname{Re}z+ib_{0}.

Turning now to (4.42), we use the resolvent identity

(4.45) A(z;q)=A0(z)±2tA0(z)qC+q¯A(z;q)\displaystyle A(z;q)=A_{0}(z)\pm 2tA_{0}(z)qC_{+}\overline{q}A(z;q)

together with (4.21) and (4.40) to bound

|I+[A(t,z;q)f]|\displaystyle\bigl{|}I_{+}\bigl{[}A(t,z;q)f\bigr{]}\bigr{|} 1ImzfL2+|t|12qC+q¯A(t,z;q)fL1\displaystyle\lesssim\tfrac{1}{\sqrt{\operatorname{Im}z}}\|f\|_{L^{2}}+|t|^{\frac{1}{2}}\|qC_{+}\overline{q}A(t,z;q)f\|_{L^{1}}
1ImzfL2+qL22[1Imz+1Imz]fL2.\displaystyle\lesssim\tfrac{1}{\sqrt{\operatorname{Im}z}}\|f\|_{L^{2}}+\|q\|_{L^{2}}^{2}\Bigl{[}\tfrac{1}{\sqrt{\operatorname{Im}z}}+\tfrac{1}{\operatorname{Im}z}\Bigr{]}\|f\|_{L^{2}}.

This proves (4.42). ∎

We end this section by proving the explicit formula for H+()H^{\infty}_{+}(\mathbb{R}) solutions qq whose initial data satisfy xq0L2()\langle x\rangle q^{0}\in L^{2}(\mathbb{R}).

Proof of Theorem 4.1.

Let q(t)q(t) be an H+H^{\infty}_{+} solution to (CCM) with initial data q0q^{0} satisfying xq0L2()\langle x\rangle q^{0}\in L^{2}(\mathbb{R}). A simple Gronwall argument guarantees that this decay property is preserved by the (CCM) evolution; see the discussion surrounding (2.13).

Let U(t;0)U(t;0) denote the unitary flow maps discussed in Proposition 2.3. Given ψ0H+()\psi_{0}\in H^{\infty}_{+}(\mathbb{R}) satisfying xψ0L2()\langle x\rangle\psi_{0}\in L^{2}(\mathbb{R}), consider

(4.46) ψ(t):=[XU(t;0)U(t;0)(X+2tq0)]ψ0notingψ(0)=0.\displaystyle\psi(t):=\bigl{[}XU(t;0)-U(t;0)\bigl{(}X+2t\mathcal{L}_{q^{0}}\bigr{)}\bigr{]}\psi_{0}\quad\text{noting}\quad\psi(0)=0.

The smoothness and decay hypotheses on ψ0\psi_{0} and the fact that U(t;0)U(t;0) preserves these properties ensure that ψ(t)\psi(t) is well defined. In particular, one may verify that ψ(t)CtL+2Ct1H+2\psi(t)\in C_{t}L^{2}_{+}\cap C_{t}^{1}H^{-2}_{+}. In fact, using (4.11) and (2.8), the time derivative of ψ(t)\psi(t) simplifies dramatically:

ddtψ(t)\displaystyle\tfrac{d}{dt}\psi(t) =𝒫(q(t))ψ(t)+[X,𝒫(q(t))]U(t;0)ψ02U(t;0)q0ψ0=𝒫(q(t))ψ(t).\displaystyle=\mathcal{P}(q(t))\psi(t)+\bigl{[}X,\mathcal{P}(q(t))\bigr{]}U(t;0)\psi_{0}-2U(t;0)\mathcal{L}_{q^{0}}\psi_{0}=\mathcal{P}(q(t))\psi(t).

By the uniqueness of solutions proved in Proposition 2.3, we deduce ψ(t)0\psi(t)\equiv 0. By the density of allowable ψ0\psi_{0}, we then deduce that XU(t;0)=U(t;0)(X+2tq0)XU(t;0)=U(t;0)(X+2t\mathcal{L}_{q^{0}}) and from there, that

(4.47) U(t;0)(X+2tq0z)1=(Xz)1U(t;0).\displaystyle U(t;0)\bigl{(}X+2t\mathcal{L}_{q^{0}}-z\bigr{)}^{-1}=\bigl{(}X-z\bigr{)}^{-1}U(t;0).

From (4.9) and the first relation in (2.8), we know that

(4.48) q(t,z)=12πiI+[(Xz)1q(t)]=12πiI+[(Xz)1U(t;0)q0].\displaystyle q(t,z)=\tfrac{1}{2\pi i}I_{+}\bigl{[}\bigl{(}X-z\bigr{)}^{-1}q(t)\bigr{]}=\tfrac{1}{2\pi i}I_{+}\bigl{[}\bigl{(}X-z\bigr{)}^{-1}U(t;0)q^{0}\bigr{]}.

Using (4.47) and the first relation in (4.5) together with Lemma 4.2, we obtain

q(t,z)\displaystyle q(t,z) =12πiI+[U(t;0)(X+2tq0z)1q0]\displaystyle=\tfrac{1}{2\pi i}I_{+}\bigl{[}U(t;0)\bigl{(}X+2t\mathcal{L}_{q^{0}}-z\bigr{)}^{-1}q^{0}\bigr{]}
=12πilimyU(t;0)χy,(X+2tq0z)1q0\displaystyle=\tfrac{1}{2\pi i}\lim_{y\to\infty}\big{\langle}U(t;0)^{*}\chi_{y},\bigl{(}X+2t\mathcal{L}_{q^{0}}-z\bigr{)}^{-1}q^{0}\big{\rangle}
=12πiI+[(X+2tq0z)1q0],\displaystyle=\tfrac{1}{2\pi i}I_{+}\bigl{[}\bigl{(}X+2t\mathcal{L}_{q^{0}}-z\bigr{)}^{-1}q^{0}\bigr{]},

which proves (4.1). ∎

5. Well-posedness in L+2()L^{2}_{+}(\mathbb{R})

In this section we will demonstrate global well-posedness of (CCM) in L+2()L^{2}_{+}(\mathbb{R}). For each initial data q0L+2()q^{0}\in L^{2}_{+}(\mathbb{R}), we will construct the solution as a limit of solutions with smooth and well decaying initial data qn0q_{n}^{0}. Our concrete hypotheses will be these:

(5.1) q0L+2()withq0L22<M,qn0H+(),xqn0L2(),qn0q0 in L2-sense,andsupnqnL2<M.\begin{gathered}q^{0}\in L^{2}_{+}(\mathbb{R})\quad\text{with}\quad\|q^{0}\|_{L^{2}}^{2}<M^{*},\quad q^{0}_{n}\in H^{\infty}_{+}(\mathbb{R}),\quad\langle x\rangle q_{n}^{0}\in L^{2}(\mathbb{R}),\\ q_{n}^{0}\to q^{0}\text{ in $L^{2}$-sense},\quad\text{and}\quad\sup_{n}\|q_{n}\|_{L^{2}}<M^{*}.\end{gathered}

Recall that MM^{*} denotes the equicontinuity threshold introduced in Definition 1.5. In Section 3, we showed that M=M^{*}=\infty in the defocusing case and that M2πM^{*}\geq 2\pi in the focusing case.

Proposition 5.1.

Suppose q0q^{0} and {qn0}n\{q^{0}_{n}\}_{n\in\mathbb{N}} satisfy (5.1) and let qn(t)q_{n}(t) denote the solutions to (CCM) with initial data qn(0)=qn0q_{n}(0)=q_{n}^{0}. Then as nn\to\infty,

(5.2) qn(t,z)EF(t,z;q0):=12πiI+[(X+2tq0z)1q0]q_{n}(t,z)\longrightarrow\text{\sl EF}(t,z;q^{0}):=\tfrac{1}{2\pi i}I_{+}\bigl{[}(X+2t\mathcal{L}_{q^{0}}-z)^{-1}q^{0}\bigr{]}

for each tt\in\mathbb{R} and each Imz>0\operatorname{Im}z>0.

Proof.

In view of Theorem 4.1, each qn(t,z)q_{n}(t,z) can be expressed through the explicit formula. Thus, our task is so show that

(5.3) limn|I+[A(t,z;qn0)qn0A(t,z;q0)q0]|=0\lim_{n\to\infty}\bigl{|}I_{+}\bigl{[}A(t,z;q_{n}^{0})q_{n}^{0}-A(t,z;q^{0})q^{0}\bigr{]}\bigr{|}=0

for fixed tt\in\mathbb{R} and Imz>0\operatorname{Im}z>0. To this end, we set Q={q0}{qn0:n}Q=\{q^{0}\}\cup\{q_{n}^{0}:n\in\mathbb{N}\} and adopt the abbreviations An(t,z):=A(t,z;qn0)A_{n}(t,z):=A(t,z;q_{n}^{0}) and A(t,z)=A(t,z;q0)A(t,z)=A(t,z;q^{0}).

The set QQ is precompact in L2L^{2} and so bounded and equicontinuous. By Lemma 4.9, we may choose b0b_{0} sufficiently large so that for Imzb0\operatorname{Im}z\geq b_{0} we have

2tC+f¯A0(t,z)gop12for allf,gQ.\big{\lVert}2tC_{+}\overline{f}A_{0}(t,z)g\big{\rVert}_{\mathrm{op}}\leq\tfrac{1}{2}\quad\text{for all}\quad f,g\in Q.

Combining this with (4.18) and (4.19), we may then estimate

(2tqn0C+q¯n0A0(t,z))\displaystyle\big{\lVert}\big{(}2tq_{n}^{0}C_{+}\smash[b]{\overline{q}}_{n}^{0}A_{0}(t,z)\big{)}^{\ell} (2tq0C+q¯0A0(t,z))L+2L+1\displaystyle-\big{(}2tq^{0}C_{+}\overline{q}^{0}A_{0}(t,z)\big{)}^{\ell}\big{\rVert}_{L^{2}_{+}\to L^{1}_{+}}
2(12)1|t|12|Imz|12qn0q0L2supqQqL2,\displaystyle\qquad\lesssim 2\ell(\tfrac{1}{2})^{\ell-1}|t|^{\frac{1}{2}}|\operatorname{Im}z|^{-\frac{1}{2}}\left\lVert q_{n}^{0}-q^{0}\right\rVert_{L^{2}}\sup_{q\in Q}\left\lVert q\right\rVert_{L^{2}},

with the implicit constant independent of 1\ell\geq 1. The role of this estimate is to allow us to expand both resolvents AA and AnA_{n} in the manner of (4.34). Proceeding in this way and using (4.21), we find that

|I+[\displaystyle\big{|}I_{+}\big{[} An(t,z)fA(t,z)f]|\displaystyle A_{n}(t,z)f-A(t,z)f\big{]}\big{|}
(5.4) 1|I+[A0(t,z){(2tqn0C+q¯n0A0(t,z))(2tq0C+q¯0A0(t,z))}f]|\displaystyle\lesssim\sum_{\ell\geq 1}\Bigl{|}I_{+}\Bigl{[}A_{0}(t,z)\Bigl{\{}\big{(}2tq_{n}^{0}C_{+}\smash[b]{\overline{q}}_{n}^{0}A_{0}(t,z)\big{)}^{\ell}-\big{(}2tq^{0}C_{+}\overline{q}^{0}A_{0}(t,z)\big{)}^{\ell}\Big{\}}f\Bigr{]}\Bigr{|}
Q|Imz|12qn0q0L2fL2\displaystyle\lesssim_{Q}|\operatorname{Im}z|^{-\frac{1}{2}}\left\lVert q_{n}^{0}-q^{0}\right\rVert_{L^{2}}\left\lVert f\right\rVert_{L^{2}}

for any Imzb0\operatorname{Im}z\geq b_{0} and any fL+2()f\in L^{2}_{+}(\mathbb{R}).

Combining (5) and (4.42), we see that

(5.5) |I+[An(z)qn0A(z)q0]||I+[An(z)A(z)]qn0|+|I+A(z)[qn0q0]|\displaystyle\big{|}I_{+}\big{[}A_{n}(z)q_{n}^{0}-A(z)q^{0}\big{]}\big{|}\leq\big{|}I_{+}\circ\big{[}A_{n}(z)-A(z)\big{]}q_{n}^{0}\big{|}+\big{|}I_{+}\circ A(z)[q^{0}_{n}-q^{0}\big{]}\big{|}

converges to zero as nn\to\infty. This proves (5.3) provided Imzb0\operatorname{Im}z\geq b_{0}.

It remains to prove convergence for general Imz>0\operatorname{Im}z>0. The restriction Imzb0\operatorname{Im}z\geq b_{0} was needed for our treatment of the first term in RHS(5.5). To overcome this, we employ (4.44) to expand

I+[An(t,z)A(t,z)]\displaystyle I_{+}\circ[A_{n}(t,z)-A(t,z)] =I+[An(t,z)A(t,z)]\displaystyle=I_{+}\circ\bigl{[}A_{n}(t,z^{\prime})-A(t,z^{\prime})\bigr{]}
+(zz)I+[An(t,z)A(t,z)]An(t,z)\displaystyle\quad+(z-z^{\prime})I_{+}\circ\bigl{[}A_{n}(t,z^{\prime})-A(t,z^{\prime})\bigr{]}A_{n}(t,z)
+(zz)I+A(t,z)[An(t,z)A(t,z)]\displaystyle\quad+(z-z^{\prime})I_{+}\circ A(t,z^{\prime})\bigl{[}A_{n}(t,z)-A(t,z)\bigr{]}

in which we choose z=Rez+ib0z^{\prime}=\operatorname{Re}z+ib_{0}. The contribution of the first two terms is easily handled using (5) and (4.32). For the third term, we use (4.42) in concert with Proposition 4.10, which showed that An(z)A(z)A_{n}(z)\to A(z) in operator norm. ∎

By the conservation of mass and the Banach–Alaoglu Theorem, the solutions qn(t)q_{n}(t) converge subsequentially weakly in L+2L^{2}_{+} for each tt\in\mathbb{R}. The previous proposition allows us to uniquely identify this limit.

Corollary 5.2.

Suppose q0q^{0} and {qn0}n\{q^{0}_{n}\}_{n\in\mathbb{N}} satisfy (5.1) and let qn(t)q_{n}(t) denote the solutions to (CCM) with initial data qn(0)=qn0q_{n}(0)=q_{n}^{0}. Then

(5.6) qn(t)EF(t;q0)weakly in L+2().q_{n}(t)\rightharpoonup\text{\sl EF}(t;q^{0})\quad\text{weakly in $L^{2}_{+}(\mathbb{R})$}.

In particular, EF(t;q0)\text{\sl EF}(t;q^{0}) defined by (5.2) belongs to L+2()L^{2}_{+}(\mathbb{R}).

Proof.

Fix t0t\neq 0, and consider an arbitrary subsequence of {qn(t)}n\{q_{n}(t)\}_{n\in\mathbb{N}}. By conservation of mass, we know that

qn(t)L2=qn0L2q0L2.\left\lVert q_{n}(t)\right\rVert_{L^{2}}=\lVert q_{n}^{0}\rVert_{L^{2}}\lesssim\lVert q^{0}\rVert_{L^{2}}.

Therefore, by Banach–Alaoglu we may pass to a further subsequence along which

(5.7) qn(t)q~(t)weakly in L+2()q_{n}(t)\rightharpoonup\widetilde{q}(t)\quad\text{weakly in }L^{2}_{+}(\mathbb{R})

for some function q~(t)\widetilde{q}(t) in the Hardy space L+2()L^{2}_{+}(\mathbb{R}). In particular q~(t,z)\widetilde{q}(t,z) is holomorphic in the upper half-plane.

As noted in (1.15), evaluation at any point zz with Imz>0\operatorname{Im}z>0 is a bounded linear functional on L+2L^{2}_{+}. Correspondingly, qn(t,z)q~(t,z)q_{n}(t,z)\to\widetilde{q}(t,z) pointwise. In this way, Proposition 5.1 implies q~(t,z)=EF(t,z;q0)\widetilde{q}(t,z)=\text{\sl EF}(t,z;q^{0}) for all Imz>0\operatorname{Im}z>0. This shows that EF(t,z;q0)\text{\sl EF}(t,z;q^{0}) is holomorphic and lies in L+2()L^{2}_{+}(\mathbb{R}).

Moreover, q~(t,z)=EF(t,z;q0)\widetilde{q}(t,z)=\text{\sl EF}(t,z;q^{0}) shows that the subsequential limit q~L+2()\widetilde{q}\in L^{2}_{+}(\mathbb{R}) does not depend on the subsequence chosen. Thus, we conclude that qn(t)EF(t;q0)q_{n}(t)\rightharpoonup\text{\sl EF}(t;q^{0}) along the whole sequence. ∎

With little additional effort, Corollary 5.2 can be used to guarantee the existence of weak/distributional solutions to (CCM). To obtain well-posedness, however, we need to prove convergence in a stronger topology, one that guarantees that the limit depends continuously on both time and the initial data. To this end, we seek to upgrade the weak convergence of qn(t)q_{n}(t) to EF(t,z;q0)\text{\sl EF}(t,z;q^{0}) to strong convergence in CtLx2([T,T]×)C_{t}L^{2}_{x}([-T,T]\times\mathbb{R}) for all T>0T>0.

Theorem 5.3.

Suppose q0q^{0} and {qn0}n\{q^{0}_{n}\}_{n\in\mathbb{N}} satisfy (5.1) and let qn(t)q_{n}(t) denote the solutions to (CCM) with initial data qn(0)=qn0q_{n}(0)=q_{n}^{0}. Then for all T>0T>0, the solutions qn(t)q_{n}(t) converge in CtLx2([T,T]×)C_{t}L^{2}_{x}([-T,T]\times\mathbb{R}) to the function EF(t;q0)\text{\sl EF}(t;q^{0}).

By Corollary 5.2 and the Arzelà–Ascoli Theorem, to prove Theorem 5.3 it suffices to show that the subset {qn(t):n1 and |t|T}\{q_{n}(t):n\geq 1\text{ and }|t|\leq T\} of CtLx2([T,T]×)C_{t}L^{2}_{x}([-T,T]\times\mathbb{R}) is equicontinuous in both the time and space variables and tight in the space variable. Recall that equicontinuity in the spatial variable was demonstrated in Section 3.

Our next result proves equicontinuity in the time variable.

Lemma 5.4 (Equicontinuity in time).

Suppose q0q^{0} and {qn0}n\{q^{0}_{n}\}_{n\in\mathbb{N}} satisfy (5.1) and let qn(t)q_{n}(t) denote the solutions to (CCM) with initial data qn(0)=qn0q_{n}(0)=q_{n}^{0}. Then for each T>0T>0 we have

(5.8) supn1qn(t+h)qn(t)CtLx2([T,T]×)0as h0.\sup_{n\geq 1}\left\lVert q_{n}(t+h)-q_{n}(t)\right\rVert_{C_{t}L^{2}_{x}([-T,T]\times\mathbb{R})}\to 0\quad\text{as }h\to 0.
Proof.

Throughout the proof, all spacetime norms will be over the slab [T,T]×[-T,T]\times\mathbb{R}.

For N1N\geq 1 to be chosen later and t,t+h[T,T]t,t+h\in[-T,T], we may estimate

qn(t+h)qn(t)CtLx2\displaystyle\left\lVert q_{n}(t+h)-q_{n}(t)\right\rVert_{C_{t}L^{2}_{x}} PN[qn(t+h)qn(t)]CtLx2+2P>NqnCtLx2\displaystyle\leq\left\lVert P_{\leq N}[q_{n}(t+h)-q_{n}(t)]\right\rVert_{C_{t}L^{2}_{x}}+2\left\lVert P_{>N}q_{n}\right\rVert_{C_{t}L^{2}_{x}}
(5.9) |h|ddtPNqnLtLx2+2P>NqnCtLx2.\displaystyle\leq|h|\bigl{\|}\tfrac{d}{dt}P_{\leq N}q_{n}\bigr{\|}_{L^{\infty}_{t}L^{2}_{x}}+2\left\lVert P_{>N}q_{n}\right\rVert_{C_{t}L^{2}_{x}}.

As the set {qn(t):n1 and |t|T}\{q_{n}(t):n\geq 1\text{ and }|t|\leq T\} is equicontinuous in the spatial variable (see Section 3), for any ε>0\varepsilon>0 we may choose N=N(ε)1N=N(\varepsilon)\geq 1 large enough so that

(5.10) supn1P>NqnCtLx2<ε10.\displaystyle\sup_{n\geq 1}\left\lVert P_{>N}q_{n}\right\rVert_{C_{t}L^{2}_{x}}<\tfrac{\varepsilon}{10}.

Using the equation (CCM), the Bernstein inequality, and the conservation of mass, we may bound

ddtPNqnLtLx2N2qn0L2+PN[qnC+(|qn|2)]LtLx2.\displaystyle\bigl{\|}\tfrac{d}{dt}P_{\leq N}q_{n}\bigr{\|}_{L^{\infty}_{t}L^{2}_{x}}\lesssim N^{2}\|q_{n}^{0}\|_{L^{2}}+\bigl{\|}P_{\leq N}[q_{n}C_{+}\big{(}|q_{n}|^{2}\big{)}^{\prime}]\bigr{\|}_{L^{\infty}_{t}L^{2}_{x}}.

Applying (1.16) and recalling the embedding L1H1L^{1}\hookrightarrow H^{-1}, we may bound the contribution of the nonlinearity as follows:

PN[qnC+(|qn|2)]LtLx2\displaystyle\bigl{\|}P_{\leq N}[q_{n}C_{+}\big{(}|q_{n}|^{2}\big{)}^{\prime}]\bigr{\|}_{L^{\infty}_{t}L^{2}_{x}} N5qnC+(|qn|2)LtHx5\displaystyle\lesssim N^{5}\|q_{n}C_{+}\big{(}|q_{n}|^{2}\big{)}^{\prime}\|_{L^{\infty}_{t}H^{-5}_{x}}
N5qnLtHx2|qn|2LtHx1\displaystyle\lesssim N^{5}\|q_{n}\|_{L^{\infty}_{t}H^{-2}_{x}}\||q_{n}|^{2}\|_{L^{\infty}_{t}H^{-1}_{x}}
N5qn0L23.\displaystyle\lesssim N^{5}\|q_{n}^{0}\|_{L^{2}}^{3}.

Thus,

ddtPNqnLtLx2N2qn0L2+N5qn0L23,\bigl{\|}\tfrac{d}{dt}P_{\leq N}q_{n}\bigr{\|}_{L^{\infty}_{t}L^{2}_{x}}\lesssim N^{2}\|q_{n}^{0}\|_{L^{2}}+N^{5}\|q_{n}^{0}\|_{L^{2}}^{3},

uniformly for n1n\geq 1. Choosing |h||h| sufficiently small, we may ensure that

supn1|h|ddtPNqnLtLx2<ε10.\sup_{n\geq 1}|h|\bigl{\|}\tfrac{d}{dt}P_{\leq N}q_{n}\bigr{\|}_{L^{\infty}_{t}L^{2}_{x}}<\tfrac{\varepsilon}{10}.

Combining this with (5.9) and (5.10) and recalling that ε>0\varepsilon>0 was arbitrary, completes the proof of the lemma. ∎

It remains to prove tightness in the spatial variable. This will be accomplished in two distinct steps. First, we will show that for any b>0b>0, the functions xqn(t,x+ib)x\mapsto q_{n}(t,x+ib) are tight in L2()L^{2}(\mathbb{R}); this is proved in Proposition 5.5 below. In the second step, we will upgrade this statement to tightness on the real line (b=0b=0); this is realized in Proposition 5.6.

Proposition 5.5.

Suppose q0q^{0} and {qn0}n\{q^{0}_{n}\}_{n\in\mathbb{N}} satisfy (5.1) and let qn(t)q_{n}(t) denote the solutions to (CCM) with initial data qn(0)=qn0q_{n}(0)=q_{n}^{0}. For any b>0b>0 and T>0T>0, the set

{qn(t,+ib):n1,|t|T}\big{\{}q_{n}(t,\cdot+ib):n\geq 1,\ |t|\leq T\big{\}}

is tight in L2()L^{2}(\mathbb{R}).

Proof.

By Theorem 4.1, the functions qn(t,z)q_{n}(t,z) admit the representation (4.1) for any zz with Imz>0\operatorname{Im}z>0. Together with the resolvent identity, this leads to

qn(t,z)=12πiI+[A0(t,z)qn0]±12πiI+[A0(t,z)2tqn0C+q¯n0A(t,z;qn0)qn0].\displaystyle q_{n}(t,z)=\tfrac{1}{2\pi i}I_{+}\big{[}A_{0}(t,z)q_{n}^{0}\big{]}\pm\tfrac{1}{2\pi i}I_{+}\big{[}A_{0}(t,z)2tq_{n}^{0}C_{+}\overline{q}_{n}^{0}A(t,z;q_{n}^{0})q_{n}^{0}\big{]}.

We will analyze the two summands above using the identity (4.20). To begin, we write

(5.11) 12πiI+[A0(t,z)qn0]=[eitΔqn0](z).\tfrac{1}{2\pi i}I_{+}\big{[}A_{0}(t,z)q_{n}^{0}\big{]}=\big{[}e^{it\Delta}q_{n}^{0}\big{]}(z).

Let Q:={qn0:n1}Q:=\{q^{0}_{n}:n\geq 1\}. As QQ is precompact in L2()L^{2}(\mathbb{R}) and the free Schrödinger propagator is continuous on L2()L^{2}(\mathbb{R}), the set of functions

:={[eitΔq](x):qQ,|t|T}\mathscr{F}:=\big{\{}\big{[}e^{it\Delta}q\big{]}(x):\ q\in Q,\ |t|\leq T\big{\}}

is also precompact in L2()L^{2}(\mathbb{R}).

For ff\in\mathscr{F} and b>0b>0, we may use the Poisson integral formula

(5.12) f(x+ib)=[eb||f](x)=1πb(yx)2+b2f(y)𝑑yf(x+ib)=[e^{-b|\partial|}f](x)=\tfrac{1}{\pi}\int\tfrac{b}{(y-x)^{2}+b^{2}}f(y)\,dy

and Cauchy–Schwarz to estimate

|x|L|f(x+ib)|2𝑑x\displaystyle\int_{|x|\geq L}|f(x+ib)|^{2}\,dx |x|Lb(yx)2+b2|f(y)|2𝑑y𝑑x\displaystyle\lesssim\int_{|x|\geq L}\int_{\mathbb{R}}\tfrac{b}{(y-x)^{2}+b^{2}}|f(y)|^{2}\,dy\,dx
|y|L0|f(y)|2𝑑y+bLL0fL22,\displaystyle\lesssim\int_{|y|\geq L_{0}}|f(y)|^{2}\,dy+\tfrac{b}{L-L_{0}}\left\lVert f\right\rVert_{L^{2}}^{2},

for any L2L01L\geq 2L_{0}\geq 1. As \mathscr{F} is precompact in L2()L^{2}(\mathbb{R}), we may pick L0L_{0} large and then LL0L\gg L_{0} to make the right-hand side above arbitrarily small. In view of (5.11), this demonstrates that the set

{I+[A0(t,x+ib)q]:qQ,|t|T}\big{\{}I_{+}\big{[}A_{0}(t,x+ib)q\big{]}:\ q\in Q,\ |t|\leq T\big{\}}

is tight in L2()L^{2}(\mathbb{R}).

It remains to show that the functions

12πiI+[A0(t,x+ib)2tqC+q¯\displaystyle\tfrac{1}{2\pi i}I_{+}\big{[}A_{0}(t,x+ib)2tqC_{+}\overline{q} A(t,x+ib;q)q]=[eitΔeb||(2tqC+q¯A(t,x+ib;q)q)](x)\displaystyle A(t,x+ib;q)q\big{]}\!\!=\!\!\Bigl{[}e^{it\Delta}e^{-b|\partial|}\bigl{(}2tqC_{+}\overline{q}A(t,x+ib;q)q\bigr{)}\Bigr{]}(x)

with qQq\in Q and |t|T|t|\leq T form a set that is tight in L2()L^{2}(\mathbb{R}).

We first observe that the set

𝒢={2tqC+q¯A(t,x+ib;q)q:qQ,|t|T}\mathscr{G}=\big{\{}2tqC_{+}\overline{q}A(t,x+ib;q)q:\ q\in Q,\ |t|\leq T\big{\}}

is bounded and tight in L1()L^{1}(\mathbb{R}). Indeed, using (4.40) we may bound

2tqC+q¯A(t,x+ib;qn)\displaystyle\bigl{\|}2tqC_{+}\overline{q}A(t,x+ib;q_{n}) qL1(|x|L)\displaystyle q\bigr{\|}_{L^{1}(|x|\geq L)}
|t|qL2(|x|L)qL2A(t,x+ib;q)L+2L+qL2\displaystyle\lesssim|t|\|q\|_{L^{2}(|x|\geq L)}\|q\|_{L^{2}}\|A(t,x+ib;q)\|_{L^{2}_{+}\to L^{\infty}_{+}}\|q\|_{L^{2}}
T(1b+1b)qL2(|x|L)qL22\displaystyle\lesssim\sqrt{T}\bigl{(}\tfrac{1}{\sqrt{b}}+\tfrac{1}{b}\bigr{)}\|q\|_{L^{2}(|x|\geq L)}\|q\|_{L^{2}}^{2}

uniformly for L0L\geq 0, qQq\in Q, |t|T|t|\leq T, and b>0b>0.

In view of the estimate

eb||gH12=(1+ξ2)e2b|ξ||g^(ξ)|2𝑑ξbg^L2bgL12,\big{\lVert}e^{-b|\partial|}g\big{\rVert}_{H^{1}}^{2}=\int(1+\xi^{2})e^{-2b|\xi|}|\widehat{g}(\xi)|^{2}\,d\xi\lesssim_{b}\lVert\widehat{g}\rVert_{L^{\infty}}^{2}\lesssim_{b}\left\lVert g\right\rVert_{L^{1}}^{2},

we deduce that the set

={eb||g:g𝒢}\mathscr{H}=\big{\{}e^{-b|\partial|}g:g\in\mathscr{G}\big{\}}

is bounded in H1()H^{1}(\mathbb{R}), and so it is bounded and equicontinuous in L2()L^{2}(\mathbb{R}). The set \mathscr{H} is also tight in L2()L^{2}(\mathbb{R}). To see this, we may use the Poisson integral formula (5.12) and Minkowski’s inequality to estimate

eb||gL2(|x|L)\displaystyle\big{\lVert}e^{-b|\partial|}g\big{\rVert}_{L^{2}(|x|\geq L)} |y|L0bx2+b2L2|g(y)|𝑑y\displaystyle\lesssim\int_{|y|\geq L_{0}}\big{\lVert}\tfrac{b}{x^{2}+b^{2}}\big{\rVert}_{L^{2}}\,|g(y)|\,dy
+|y|L0bx2+b2L2(|x|LL0)|g(y)|𝑑y\displaystyle\quad+\int_{|y|\leq L_{0}}\big{\lVert}\tfrac{b}{x^{2}+b^{2}}\big{\rVert}_{L^{2}(|x|\geq L-L_{0})}\,|g(y)|\,dy
|y|L01b1/2|g(y)|𝑑y+b(LL0)3/2gL1,\displaystyle\lesssim\int_{|y|\geq L_{0}}\tfrac{1}{b^{1/2}}|g(y)|\,dy+\tfrac{b}{(L-L_{0})^{3/2}}\left\lVert g\right\rVert_{L^{1}},

for any L2L01L\geq 2L_{0}\geq 1. As 𝒢\mathscr{G} is bounded and tight in L1()L^{1}(\mathbb{R}), we may pick L0L_{0} large and then LL0L\gg L_{0} to make the right-hand side above arbitrarily small.

Thus \mathscr{H} is precompact in L2()L^{2}(\mathbb{R}). As the free Schrödinger propagator is continuous on L2()L^{2}(\mathbb{R}), we conclude that the set

{I+[A0(t,x+ib)2tqC+q¯A(t,x+ib;q)q]:n1,|t|T}\big{\{}I_{+}\big{[}A_{0}(t,x+ib)2tqC_{+}\overline{q}A(t,x+ib;q)q\bigr{]}:\ n\geq 1,\ |t|\leq T\big{\}}

is precompact in L2()L^{2}(\mathbb{R}) and so tight in L2()L^{2}(\mathbb{R}). This completes the proof of the proposition. ∎

Proposition 5.6 (Tightness in space).

Suppose q0q^{0} and {qn0}n\{q^{0}_{n}\}_{n\in\mathbb{N}} satisfy (5.1) and let qn(t)q_{n}(t) denote the solutions to (CCM) with initial data qn(0)=qn0q_{n}(0)=q_{n}^{0}. Given T>0T>0, the set {qn(t):n1 and |t|T}\{q_{n}(t):n\geq 1\text{ and }|t|\leq T\} is tight in L2()L^{2}(\mathbb{R}).

Proof.

For a frequency cutoff N1N\geq 1 to be chosen later, we decompose qn=PNqn+P>Nqnq_{n}=P_{\leq N}q_{n}+P_{>N}q_{n}. Let χ:[0,1]\chi:\mathbb{R}\to[0,1] be a smooth cutoff function equal to 11 on |x|2|x|\geq 2 and to 0 on |x|1|x|\leq 1. For R>0R>0, we define χ>R(x):=χ(xR)\chi_{>R}(x):=\chi(\frac{x}{R}). We may then bound

χ>Rqn(t)L2χ>RPNqn(t)L2+P>Nqn(t)L2.\left\lVert\chi_{>R}q_{n}(t)\right\rVert_{L^{2}}\leq\left\lVert\chi_{>R}P_{\leq N}q_{n}(t)\right\rVert_{L^{2}}+\left\lVert P_{>N}q_{n}(t)\right\rVert_{L^{2}}.

The hypotheses (5.1) guarantee L2L^{2}-equicontinuity of the solutions qn(t)q_{n}(t). Therefore, we can choose N1N\geq 1 sufficiently large to render supn1sup|t|TP>Nqn(t)L2\sup_{n\geq 1}\sup_{|t|\leq T}\left\lVert P_{>N}q_{n}(t)\right\rVert_{L^{2}} arbitrarily small.

To demonstrate tightness of the low frequencies, for b>0b>0 fixed we write

PNqn(t)=(PNeb||)(eb||qn(t))=K(eb||qn(t)),P_{\leq N}q_{n}(t)=\big{(}P_{\leq N}e^{b|\partial|}\big{)}\big{(}e^{-b|\partial|}q_{n}(t)\big{)}=K*\big{(}e^{-b|\partial|}q_{n}(t)\big{)},

where the convolution kernel KK is the Schwartz function

K(x)=12πeixξeb|ξ|φ(ξN)𝑑ξ=2π0cos(xξ)ebξφ(ξN)𝑑ξK(x)=\tfrac{1}{\sqrt{2\pi}}\int e^{ix\xi}e^{b|\xi|}\varphi\big{(}\tfrac{\xi}{N}\big{)}\,d\xi=\sqrt{\tfrac{2}{\pi}}\int_{0}^{\infty}\cos(x\xi)e^{b\xi}\varphi\big{(}\tfrac{\xi}{N}\big{)}\,d\xi

and φ:[0,1]\varphi:\mathbb{R}\to[0,1] denotes the cutoff function used to define the Littlewood–Paley projections. Integrating by parts twice yields the easy bound

K(x)b,Nx2.K(x)\lesssim_{b,N}\langle x\rangle^{-2}.

Using this and the observation that χ>R[(χR4K)(χR4qn)]=0\chi_{>R}[(\chi_{\leq\frac{R}{4}}K)*(\chi_{\leq\frac{R}{4}}q_{n})]=0, which follows from support considerations, we may estimate

χ>RPNqn(t)L2\displaystyle\left\lVert\chi_{>R}P_{\leq N}q_{n}(t)\right\rVert_{L^{2}} K[χ>R4eb||qn(t)]L2+[χ>R4K][χR4eb||qn(t)]L2\displaystyle\leq\big{\lVert}K*\big{[}\chi_{>\frac{R}{4}}e^{-b|\partial|}q_{n}(t)\big{]}\big{\rVert}_{L^{2}}+\big{\lVert}\big{[}\chi_{>\frac{R}{4}}K\big{]}*\big{[}\chi_{\leq\frac{R}{4}}e^{-b|\partial|}q_{n}(t)\big{]}\big{\rVert}_{L^{2}}
χ>R4eb||qn(t)L2+R1χR4eb||qn(t)L2\displaystyle\lesssim\lVert\chi_{>\frac{R}{4}}e^{-b|\partial|}q_{n}(t)\rVert_{L^{2}}+R^{-1}\lVert\chi_{\leq\frac{R}{4}}e^{-b|\partial|}q_{n}(t)\rVert_{L^{2}}
χ>R4eb||qn(t)L2+R1qn0L2.\displaystyle\lesssim\lVert\chi_{>\frac{R}{4}}e^{-b|\partial|}q_{n}(t)\rVert_{L^{2}}+R^{-1}\|q_{n}^{0}\|_{L^{2}}.

By Proposition 5.5, the functions [eb||qn](t,x)=qn(t,x+ib)[e^{-b|\partial|}q_{n}](t,x)=q_{n}(t,x+ib) form an L2L^{2}-tight set for n1n\geq 1 and |t|T|t|\leq T. Therefore, we may choose RR sufficiently large to make the right-hand side above arbitrarily small, which demonstrates that {PNqn(t):n1,|t|T}\{P_{\leq N}q_{n}(t):\ n\geq 1,\ |t|\leq T\} is tight in L2()L^{2}(\mathbb{R}), as desired. ∎

We are now ready to prove well-posedness of (CCM) in L+2()L^{2}_{+}(\mathbb{R}) for data satisfying M(q0)<MM(q^{0})<M^{*}, with MM^{*} denoting the equicontinuity threshold introduced in Definition 1.5. This will complete the proof of both Theorem 1.1 and Theorem 1.6.

Proof of well-posedness in L+2()L^{2}_{+}(\mathbb{R}).

We will show that the data-to-solution map for (CCM) extends uniquely from {qH+():xqL2() and M(q)<M}\{q\in H^{\infty}_{+}(\mathbb{R}):\,\langle x\rangle q\in L^{2}(\mathbb{R})\text{ and }M(q)<M^{*}\} to a jointly continuous map Φ:×{qL+2():M(q)<M}L+2\Phi:\mathbb{R}\times\{q\in L^{2}_{+}(\mathbb{R}):\ M(q)<M^{*}\}\to L^{2}_{+}.

Given initial data q0L+2()q^{0}\in L^{2}_{+}(\mathbb{R}) with M(q0)<MM(q^{0})<M^{*}, let {qn0}n1\{q_{n}^{0}\}_{n\geq 1} be a sequence satisfying (5.1). Applying Theorem 5.3 to the sequence {qn0}n1\{q_{n}^{0}\}_{n\geq 1}, we see that the corresponding H+()H^{\infty}_{+}(\mathbb{R}) solutions qn(t)q_{n}(t) to (CCM) converge in L2()L^{2}(\mathbb{R}) and the limit is independent of the sequence {qn0}n1\{q_{n}^{0}\}_{n\geq 1}. Consequently,

Φ(t,q0):=limnqn(t)\Phi(t,q^{0}):=\lim_{n\to\infty}q_{n}(t)

is well-defined.

We must show that Φ\Phi is jointly continuous. Fix T>0T>0 and q0L+2()q^{0}\in L^{2}_{+}(\mathbb{R}) with M(q0)<MM(q^{0})<M^{*}. Let {qn0}n1\{q_{n}^{0}\}_{n\geq 1} be an L+2L^{2}_{+} sequence that converges to q0q^{0} in L2()L^{2}(\mathbb{R}); without loss of generality, we may assume that supn1M(qn0)<M\sup_{n\geq 1}M(q_{n}^{0})<M^{*}. By the definition of Φ\Phi, we may choose another sequence q~n(t)\widetilde{q}_{n}(t) of H+H^{\infty}_{+} solutions to (CCM) so that xq~n(0)L2()\langle x\rangle\widetilde{q}_{n}(0)\in L^{2}(\mathbb{R}) and

(5.13) sup|t|TΦ(t,qn0)q~n(t)L20as n.\sup_{|t|\leq T}\left\lVert\Phi(t,q_{n}^{0})-\widetilde{q}_{n}(t)\right\rVert_{L^{2}}\to 0\quad\text{as }n\to\infty.

In particular, q~n(0)q0\widetilde{q}_{n}(0)\to q^{0} in L2()L^{2}(\mathbb{R}), and so Theorem 5.3 yields

(5.14) sup|t|Tq~n(t)Φ(t,q0)L20as n.\sup_{|t|\leq T}\left\lVert\widetilde{q}_{n}(t)-\Phi(t,q^{0})\right\rVert_{L^{2}}\to 0\quad\text{as }n\to\infty.

Given {tn}n1[T,T]\{t_{n}\}_{n\geq 1}\subset[-T,T] that converges to some t[T,T]t\in[-T,T], we may bound

Φ(tn,qn0)Φ(t,q0)L2\displaystyle\left\lVert\Phi(t_{n},q_{n}^{0})-\Phi(t,q^{0})\right\rVert_{L^{2}}
Φ(tn,qn0)q~n(tn)L2+q~n(tn)q~n(t)L2+q~n(t)Φ(t,q0)L2\displaystyle\qquad\leq\left\lVert\Phi(t_{n},q_{n}^{0})-\widetilde{q}_{n}(t_{n})\right\rVert_{L^{2}}+\left\lVert\widetilde{q}_{n}(t_{n})-\widetilde{q}_{n}(t)\right\rVert_{L^{2}}+\left\lVert\widetilde{q}_{n}(t)-\Phi(t,q^{0})\right\rVert_{L^{2}}
sup|t|TΦ(t,qn0)q~n(t)L2+q~n(tn)q~n(t)L2+sup|t|Tq~n(t)Φ(t,q0)L2.\displaystyle\qquad\leq\sup_{|t|\leq T}\left\lVert\Phi(t,q_{n}^{0})-\widetilde{q}_{n}(t)\right\rVert_{L^{2}}+\left\lVert\widetilde{q}_{n}(t_{n})-\widetilde{q}_{n}(t)\right\rVert_{L^{2}}+\sup_{|t|\leq T}\left\lVert\widetilde{q}_{n}(t)-\Phi(t,q^{0})\right\rVert_{L^{2}}.

The right-hand side above converges to zero as nn\to\infty by (5.13), (5.14), and (5.8). This demonstrates that Φ\Phi is jointly continuous. ∎

Finally, we prove the explicit formula for general L+2()L^{2}_{+}(\mathbb{R}) initial data:

Proof of Theorem 1.7.

From Theorem 4.1, we know that the explicit formula holds for initial data that is smooth and well-decaying. With this in mind, we choose a sequence qn0q^{0}_{n} of such initial data that converges to q0q^{0} in L2L^{2}-sense and satisfies M(qn0)<MM(q^{0}_{n})<M^{*}. Now we simply need to show convergence of both sides of (4.1). In the case of the left-hand side, this is immediate from the well-posedness just proved and (1.15).

The right-hand side requires further discussion. In view of (4.21) and (4.45), we merely need to show that

2tqn0C+q¯n0A(t,z;qn0)qn02tq0C+q¯0A(t,z;q0)q0in L1 sense.\displaystyle 2tq^{0}_{n}C_{+}\smash[b]{\overline{q}}^{0}_{n}A(t,z;q^{0}_{n})q^{0}_{n}\to 2tq^{0}C_{+}{\overline{q}}^{0}A(t,z;q^{0})q^{0}\quad\text{in $L^{1}$ sense.}

Employing (4.40), elementary inequalities reduce this question to showing that

limnA(t,z;qn0)A(t,z;q0)L+2L+=0.\displaystyle\lim_{n\to\infty}\ \bigl{\|}A(t,z;q^{0}_{n})-A(t,z;q^{0})\bigr{\|}_{L^{2}_{+}\to L^{\infty}_{+}}=0.

This in turn follows from (4.40), (4.41), and the resolvent identity. ∎

6. Well-posedness in H+s()H^{s}_{+}(\mathbb{R})

The goal of this section is to prove

Theorem 6.1 (Global well-posedness in H+s()H^{s}_{+}(\mathbb{R})).

Fix 0<s<10<s<1. Both (CCM) equations (focusing and defocusing) are globally well-posed in the space

BM={qH+s():M(q)<M},B_{M^{*}}=\{q\in H^{s}_{+}(\mathbb{R}):\ M(q)<M^{*}\},

where MM^{*} denotes the equicontinuity threshold introduced in Definition 1.5.

In view of the L+2()L^{2}_{+}(\mathbb{R}) global well-posedness result demonstrated in the preceding section, it will suffice to prove that (CCM) solutions satisfy a priori HsH^{s} bounds, and that the orbits of HsH^{s}-bounded and equicontinuous sets of initial data form an HsH^{s}-equicontinuous set. These two claims will be taken up in Proposition 6.5. The proof of this proposition is based on a novel argument, in which the following family of functions play a central role:

Definition 6.2.

Fix 0<s<10<s<1. For each K0K\geq 0, we define the function

(6.1) FK(E):=sin(πs)πKEλs1dλE+λfor eachE[0,).\displaystyle F_{K}(E):=\frac{\sin(\pi s)}{\pi}\int_{K}^{\infty}\frac{E\lambda^{s-1}\,d\lambda}{E+\lambda}\quad\text{for each}\quad E\in[0,\infty).

The particular structure of these functions is dictated by our use of Loewner’s Theorem on operator monotone functions. Concretely, for any positive definite operators AB0A\geq B\geq 0, it follows that FK(A)FK(B)0F_{K}(A)\geq F_{K}(B)\geq 0.

If one chooses K=0K=0, then (6.1) can be evaluated exactly: F0(E)=EsF_{0}(E)=E^{s}. It is to ensure this identity that we have included the prefactor in (6.1). Clearly FK(E)F_{K}(E) is a decreasing function of KK; thus FK(E)EsF_{K}(E)\leq E^{s} for all E0E\geq 0 and all K0K\geq 0.

For K>0K>0, we know of no explicit formula for FK(E)F_{K}(E); however, it is not difficult to describe the order of magnitude of this function:

(6.2) FK(E){Es:EKEKs1:0EK\displaystyle F_{K}(E)\approx\begin{cases}E^{s}&:E\geq K\\ EK^{s-1}&:0\leq E\leq K\end{cases}

uniformly in K0K\geq 0. The overall shape described by (6.2) also plays an important role in the selection of the function FK(E)F_{K}(E), through the following expression of equicontinuity:

Lemma 6.3.

A bounded subset QH+s()Q\subseteq H^{s}_{+}(\mathbb{R}) is HsH^{s}-equicontinuous if and only if

(6.3) supqQq,FK((0+κ)2)q0asK\displaystyle\sup_{q\in Q}\ \bigl{\langle}q,F_{K}\bigl{(}(\mathcal{L}_{0}+\kappa)^{2}\bigr{)}q\bigr{\rangle}\to 0\quad\text{as}\quad K\to\infty

for any single κ>0\kappa>0.

Proof.

In view of (6.2), we have

q,FK((0+κ)2)q(κ2+N2K)1s(0+κ)sqNL22+(0+κ)sq>NL22.\bigl{\langle}q,F_{K}\bigl{(}(\mathcal{L}_{0}+\kappa)^{2}\bigr{)}q\bigr{\rangle}\lesssim\bigl{(}\tfrac{\kappa^{2}+N^{2}}{K}\bigr{)}^{1-s}\bigl{\|}(\mathcal{L}_{0}+\kappa)^{s}q_{\leq N}\bigr{\|}_{L^{2}}^{2}+\bigl{\|}(\mathcal{L}_{0}+\kappa)^{s}q_{>N}\bigr{\|}_{L^{2}}^{2}.

Choosing NN large and then sending KK\to\infty, we see that equicontinuous sets satisfy (6.3). Conversely,

q>NHs2q,FK((0+κ)2)qwhenK=(N+κ)2.\|q_{>N}\|_{H^{s}}^{2}\lesssim\bigl{\langle}q,F_{K}\bigl{(}(\mathcal{L}_{0}+\kappa)^{2}\bigr{)}q\bigr{\rangle}\quad\text{when}\quad K=(N+\kappa)^{2}.

Thus, (6.3) implies that QQ is equicontinuous. ∎

The use of 0\mathcal{L}_{0} in this lemma, makes for an easy connection to HsH^{s}-equicontinuity; however, it is the Lax operator q(t)\mathcal{L}_{q(t)} associated to the solution itself that enjoys a close connection with the (CCM) flows. Our next result bridges this divide.

Lemma 6.4.

Fix s(0,1)s\in(0,1). Suppose QH+()Q\subseteq H^{\infty}_{+}(\mathbb{R}) is bounded and equicontinuous in L2()L^{2}(\mathbb{R}). Then there exists κ0=κ0(Q)\kappa_{0}=\kappa_{0}(Q) so that

(6.4) q,FK((q+κ)2)qq,FK((0+κ)2)q\displaystyle\bigl{\langle}q,F_{K}\bigl{(}(\mathcal{L}_{q}+\kappa)^{2}\bigr{)}q\bigr{\rangle}\approx\bigl{\langle}q,F_{K}\bigl{(}(\mathcal{L}_{0}+\kappa)^{2}\bigr{)}q\bigr{\rangle}

uniformly for qQq\in Q, κκ0\kappa\geq\kappa_{0}, and K0K\geq 0.

Proof.

Using (2.3), we may choose κ0(Q)>0\kappa_{0}(Q)>0 so that

(6.5) supqQqC+q¯R0(κ)op14uniformly for qQ and κκ0.\displaystyle\sup_{q\in Q}\bigl{\|}qC_{+}\overline{q}R_{0}(\kappa)\bigr{\|}_{\mathrm{op}}\leq\tfrac{1}{4}\quad\text{uniformly for $q\in Q$ and $\kappa\geq\kappa_{0}$}.

Using elementary manipulations and (6.5), we find that

|f,(q+κ)2ff,(0+κ)2f|\displaystyle\Bigl{|}\bigl{\langle}f,(\mathcal{L}_{q}+\kappa)^{2}f\bigr{\rangle}-\bigl{\langle}f,(\mathcal{L}_{0}+\kappa)^{2}f\bigr{\rangle}\Bigr{|} qC+q¯fL2[2(0+κ)fL2+qC+q¯fL2]\displaystyle\leq\bigl{\|}qC_{+}\overline{q}f\bigr{\|}_{L^{2}}\bigl{[}2\|(\mathcal{L}_{0}+\kappa)f\|_{L^{2}}+\|qC_{+}\overline{q}f\|_{L^{2}}\bigr{]}
916(0+κ)fL22\displaystyle\leq\tfrac{9}{16}\|(\mathcal{L}_{0}+\kappa)f\|_{L^{2}}^{2}

for all κκ0\kappa\geq\kappa_{0} and any fH+1()f\in H^{1}_{+}(\mathbb{R}). Equivalently,

716(0+κ)2(q+κ)22516(0+κ)2for all qQ and κκ0.\displaystyle\tfrac{7}{16}(\mathcal{L}_{0}+\kappa)^{2}\leq(\mathcal{L}_{q}+\kappa)^{2}\leq\tfrac{25}{16}(\mathcal{L}_{0}+\kappa)^{2}\quad\text{for all $q\in Q$ and $\kappa\geq\kappa_{0}$}.

As the functions FKF_{K} are operator monotone, we deduce that

FK(716(0+κ)2)FK((q+κ)2)FK(2516(0+κ)2).\displaystyle F_{K}\bigl{(}\tfrac{7}{16}(\mathcal{L}_{0}+\kappa)^{2}\bigr{)}\leq F_{K}\bigl{(}(\mathcal{L}_{q}+\kappa)^{2}\bigr{)}\leq F_{K}\bigl{(}\tfrac{25}{16}(\mathcal{L}_{0}+\kappa)^{2}\bigr{)}.

From the overall shape of FK(E)F_{K}(E) described in (6.2), it then follows that

(6.6) cFK((0+κ)2)FK((q+κ)2)c1FK((0+κ)2).\displaystyle cF_{K}\bigl{(}(\mathcal{L}_{0}+\kappa)^{2}\bigr{)}\leq F_{K}\bigl{(}(\mathcal{L}_{q}+\kappa)^{2}\bigr{)}\leq c^{-1}F_{K}\bigl{(}(\mathcal{L}_{0}+\kappa)^{2}\bigr{)}.

for some (small) number cc depending only on ss. The claim (6.4) is a direct consequence of this. ∎

Proposition 6.5.

Fix s(0,1)s\in(0,1). Assume Q{qH+():M(q)<M}Q\subset\{q\in H^{\infty}_{+}(\mathbb{R}):\ M(q)<M^{*}\} is bounded and equicontinuous in Hs()H^{s}(\mathbb{R}). Then

(6.7) Q:={etJH±q0:q0Q,t}Q^{*}:=\{e^{tJ\nabla H_{\pm}}q^{0}:q^{0}\in Q,\ t\in\mathbb{R}\}

is also bounded and equicontinuous in Hs()H^{s}(\mathbb{R}).

Proof.

As the set QQ is bounded in Hs()H^{s}(\mathbb{R}), it is bounded and equicontinuous in L2()L^{2}(\mathbb{R}). By the results of Section 3, the set of orbits QQ^{*} is bounded and equicontinuous in L2()L^{2}(\mathbb{R}). Thus, Lemma 6.4 may be applied to QQ^{*}; we define κ0=κ0(Q)\kappa_{0}=\kappa_{0}(Q^{*}) accordingly.

Employing both parts of (2.8) from Proposition 2.3, we see that

(6.8) q(t),FK((q(t)+κ)2)q(t)=q(0),FK((q(0)+κ)2)q(0),\displaystyle\big{\langle}q(t),F_{K}\bigl{(}(\mathcal{L}_{q(t)}+\kappa)^{2}\bigr{)}q(t)\big{\rangle}=\big{\langle}q(0),F_{K}\bigl{(}(\mathcal{L}_{q(0)}+\kappa)^{2}\bigr{)}q(0)\big{\rangle},

for any choices of κκ0(Q)\kappa\geq\kappa_{0}(Q^{*}), s(0,1)s\in(0,1), K0K\geq 0, and any H+()H^{\infty}_{+}(\mathbb{R}) solution q(t)q(t) to (CCM).

Freezing κ=κ0(Q)\kappa=\kappa_{0}(Q^{*}) and applying Lemma 6.4, the identity (6.8) yields

(6.9) q(t),FK((0+κ)2)q(t)q(0),FK((0+κ)2)q(0),\displaystyle\big{\langle}q(t),F_{K}\bigl{(}(\mathcal{L}_{0}+\kappa)^{2}\bigr{)}q(t)\big{\rangle}\approx\big{\langle}q(0),F_{K}\bigl{(}(\mathcal{L}_{0}+\kappa)^{2}\bigr{)}q(0)\big{\rangle},

uniformly for q(0)Qq(0)\in Q, tt\in\mathbb{R}, and K0K\geq 0.

Recalling that F0(E)=EsF_{0}(E)=E^{s}, the K=0K=0 case of (6.9) shows that QQ^{*} inherits HsH^{s}-boundedness from QQ. Moreover, sending KK\to\infty in (6.9) and employing Lemma 6.3, we see that QQ^{*} inherits HsH^{s}-equicontinuity from QQ. ∎

We are now ready to finish the proof of well-posedness in H+s()H^{s}_{+}(\mathbb{R}). Based on the tools we have developed, it is most convenient to construct the data-to-solution map by extension from the class of smooth solutions. As the extension will be shown to be H+s()C([T,T];H+s())H^{s}_{+}(\mathbb{R})\to C([-T,T];H^{s}_{+}(\mathbb{R})) continuous, it coincides with the restriction to H+s()H^{s}_{+}(\mathbb{R}) of the map constructed in the previous section for initial data in L+2()L^{2}_{+}(\mathbb{R}).

Proof of Theorem 6.1.

Suppose q0H+s()q^{0}\in H^{s}_{+}(\mathbb{R}) and M(q0)<MM(q^{0})<M^{*}. Let qn(t)q_{n}(t) be a sequence of H+()H^{\infty}_{+}(\mathbb{R}) solutions with qn(0)q0q_{n}(0)\to q^{0} in HsH^{s}-sense. Our sole obligation is to show that qn(t)q_{n}(t) converges in C([T,T];H+s())C([-T,T];H^{s}_{+}(\mathbb{R})) for every finite T>0T>0. If such convergence did not hold, there would be a convergent sequence tnt_{n} so that qn(tn)q_{n}(t_{n}) did not converge in H+s()H^{s}_{+}(\mathbb{R}). We will refute this.

From the results of the previous section, we know that qn(tn)q_{n}(t_{n}) converges in L+2()L^{2}_{+}(\mathbb{R}). Moreover, Proposition 6.5 shows that the sequence qn(tn)q_{n}(t_{n}) is both uniformly bounded and equicontinuous in H+s()H^{s}_{+}(\mathbb{R}). Thus, qn(tn)q_{n}(t_{n}) does indeed converge in H+s()H^{s}_{+}(\mathbb{R}). ∎

References