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

Sobolev norms of L2L^{2}- solutions to NLS

Roman V. Bessonov,  Sergey A. Denisov Roman Bessonov: [email protected] St. Petersburg State University Universitetskaya nab. 7-9, 199034 St. Petersburg, RUSSIA St. Petersburg Department of Steklov Mathematical Institute Russian Academy of Sciences Fontanka 27, 191023 St.Petersburg, RUSSIA Sergey Denisov: [email protected] University of Wisconsin–Madison Department of Mathematics 480 Lincoln Dr., Madison, WI, 53706, USA
Abstract.

We apply inverse spectral theory to study Sobolev norms of solutions to the nonlinear Schrödinger equation. For initial datum q0L2()q_{0}\in L^{2}(\mathbb{R}) and s[1,0]s\in[-1,0], we prove that there exists a conserved quantity which is equivalent to Hs()H^{s}(\mathbb{R})-norm of the solution.

Key words and phrases:
Dirac operators, NLS, scattering, Sobolev norms
2010 Mathematics Subject Classification:
35Q55
The work of RB in Sections 2 and 3 is supported by the Russian Science Foundation grant 19-71-30002. The work of SD in the rest of the paper is supported by the grant NSF DMS-2054465 and Van Vleck Professorship Research Award. RB is a Young Russian Mathematics award winner and would like to thank its sponsors and jury.

1. Introduction

The classical defocusing nonlinear Schrödinger equation (NLS) [13, 23, 27] on the real line has the form

{iqt=2qξ2+2|q|2q,q|t=0=q0,ξ,t.\begin{cases}i\frac{\partial q}{\partial t}=-\frac{\partial^{2}q}{\partial\xi^{2}}+2|q|^{2}q,\\ q\big{\rvert}_{t=0}=q_{0},\end{cases}\qquad\xi\in\mathbb{R},\quad t\in\mathbb{R}. (1.1)

It is known that for sufficiently regular initial datum q0q_{0} the unique classical solution q=q(ξ,t)q=q(\xi,t) exists globally in time. For example, if q0q_{0} lies in the Schwartz class 𝒮()\mathcal{S}(\mathbb{R}), then q(,t)𝒮()q(\cdot,t)\in\mathcal{S}(\mathbb{R}) for all tt\in\mathbb{R}. The long-time asymptotics of qq is known [9, 26, 10]. For less regular initial datum q0q_{0}, one can define the solution by an approximation argument (see, e.g., [25]):

Theorem 1.1.

Let q0L2()q_{0}\in L^{2}(\mathbb{R}), and let q0,n𝒮()q_{0,n}\in\mathcal{S}(\mathbb{R}) converge to q0q_{0} in L2()L^{2}(\mathbb{R}). Denote by qn(ξ,t)q_{n}(\xi,t) the solution of (1.1) corresponding to q0,nq_{0,n}. We have

limn+qn(,t)q(,t)L2()=0,t,\lim_{n\to+\infty}\|q_{n}(\cdot,t)-q(\cdot,t)\|_{L^{2}(\mathbb{R})}=0,\qquad t\in\mathbb{R},

for some function q(ξ,t):2q(\xi,t):\mathbb{R}^{2}\to\mathbb{R} that does not depend on the choice of the sequence q0,nq_{0,n}.

The function qq in Theorem 1.1 is called the L2L^{2}–solution of (1.1) corresponding to the initial datum q0L2()q_{0}\in L^{2}(\mathbb{R}). It is clear that such a solution is unique. The total energy of the solution is its L2()L^{2}(\mathbb{R})-norm and it is conserved in time:

q(,t)L2()=q0L2(),t.\|q(\cdot,t)\|_{L^{2}(\mathbb{R})}=\|q_{0}\|_{L^{2}(\mathbb{R})},\qquad t\in\mathbb{R}\,.

By Plancherel’s formula, it is equal to (q)(,t)L2()\|(\mathcal{F}q)(\cdot,t)\|_{L^{2}(\mathbb{R})} where \mathcal{F} stands for the Fourier transform. In this paper, we work with Sobolev spaces Hs()H^{s}(\mathbb{R}), ss\in\mathbb{R}. The Hs()H^{s}(\mathbb{R})-norm of a function f𝒮()f\in\mathcal{S}(\mathbb{R}) is defined by

fHs()=((1+|η|2)s|(f)(η)|2𝑑η)12.\|f\|_{H^{s}(\mathbb{R})}=\left(\int_{\mathbb{R}}(1+|\eta|^{2})^{s}|(\mathcal{F}f)(\eta)|^{2}d\eta\right)^{\frac{1}{2}}\,. (1.2)

The space Hs()H^{s}(\mathbb{R}) is the completion of 𝒮()\mathcal{S}(\mathbb{R}) with respect to this norm. Equivalently, one can define it by

Hs()={f𝒮():(1+|η|2)s2fL2()},H^{s}(\mathbb{R})=\{f\in\mathcal{S}^{\prime}(\mathbb{R}):(1+|\eta|^{2})^{{\frac{s}{2}}}\mathcal{F}f\in L^{2}(\mathbb{R})\}\,,

where 𝒮()\mathcal{S}^{\prime}(\mathbb{R}) is the space of tempered distribution.

In contrast to the linear Schrödinger equation for which all Sobolev norms are conserved, the solutions of NLS can exhibit inflation of Sobolev norm Hs()H^{s}(\mathbb{R}) for s12s\leqslant-\frac{1}{2} (see, e.g., [8, 17] for details). Specifically, given an arbitrarily small positive ε\varepsilon and s12s\leqslant-\frac{1}{2}, there exists a solution qq to (1.1) that satisfies

q0𝒮(),q0Hs()ε,q(,ε)Hs()ε1,q_{0}\in\mathcal{S}(\mathbb{R}),\qquad\|q_{0}\|_{H^{s}(\mathbb{R})}\leqslant\varepsilon,\qquad\|q(\cdot,\varepsilon)\|_{H^{s}(\mathbb{R})}\geqslant\varepsilon^{-1}\,, (1.3)

see [8] for that construction. This result is related to the “high-to-low frequency cascade”. It occurs when for initial datum q0𝒮()q_{0}\in\mathcal{S}(\mathbb{R}), a part of L2()L^{2}(\mathbb{R})-norm of qq, when written on the Fourier side, moves from high to low frequencies as time increases. The Sobolev norms with negative index ss can be used to capture this phenomenon. Indeed, since q(,t)L2()\|q(\cdot,t)\|_{L^{2}(\mathbb{R})} is time-invariant and the weight (1+η2)s(1+\eta^{2})^{s} in (1.2) vanishes at infinity when s<0s<0, the transfer of L2L^{2}-norm from high to low values of frequency η\eta makes the Hs()H^{s}(\mathbb{R})-norm grow.

For NLS, the inflation of Hs()H^{s}(\mathbb{R})-norm can not happen for s>12s>-{\frac{1}{2}}. In [18], Koch and Tataru discovered the set of conserved quantities which agree with Hs()H^{s}(\mathbb{R})-norm up to a quadratic term for a small value of q0Hs()\|q_{0}\|_{H^{s}(\mathbb{R})} and s>12s>-\frac{1}{2}. As a corollary, they obtained the bounds on q(,t)Hs()\|q(\cdot,t)\|_{H^{s}(\mathbb{R})} that are uniform in time:

q(,t)Hs()C(s){+1+2s,s>0,+1+4s1+2s,s(12,0),=q0Hs().\|q(\cdot,t)\|_{H^{s}(\mathbb{R})}\leqslant C(s)\left\{\begin{matrix}\mathcal{R}+\mathcal{R}^{1+2s},&s>0,\\ \mathcal{R}+\mathcal{R}^{\frac{1+4s}{1+2s}},&s\in(-\frac{1}{2},0)\,,\end{matrix}\right.\quad\mathcal{R}=\|q_{0}\|_{H^{s}(\mathbb{R})}\,. (1.4)

In [16], Killip, Vişan, and Zhang proved a similar estimate using a different method. The estimates on the growth of Hs()H^{s}(\mathbb{R})-norms are related to questions of well-posedness and ill-posedness of NLS in Sobolev classes which have been extensively studied previously, see, e.g., [14, 18, 17, 5, 6, 7, 20].

In our paper, we use some recent results in the inverse spectral theory [2, 3, 1] to show that there are conserved quantities of NLS which agree with Hs()H^{s}(\mathbb{R})-norm provided that s[1,0]s\in[-1,0] and the value of q0L2()\|q_{0}\|_{L^{2}(\mathbb{R})} is under control. We apply our analysis to prove the following theorem.

Theorem 1.2.

Let q0L2()q_{0}\in L^{2}(\mathbb{R}) and let q=q(ξ,t)q=q(\xi,t) be the solution of (1.1) corresponding to q0q_{0}. Then,

C1(1+q0L2())2sq0Hs()q(,t)Hs()C2(1+q0L2())2sq0Hs(),C_{1}(1+\|q_{0}\|_{L^{2}(\mathbb{R})})^{2s}\|q_{0}\|_{H^{s}(\mathbb{R})}\leqslant\|q(\cdot,t)\|_{H^{s}(\mathbb{R})}\leqslant C_{2}(1+\|q_{0}\|_{L^{2}(\mathbb{R})})^{-2s}\|q_{0}\|_{H^{s}(\mathbb{R})}, (1.5)

where t,s[1,0]\,t\in\mathbb{R}\,,\,s\in[-1,0], and C1C_{1} and C2C_{2} are two positive absolute constants.

This result shows, in particular, that for a given function q0:q0L2()=1q_{0}:\|q_{0}\|_{L^{2}(\mathbb{R})}=1 whose L2()L^{2}(\mathbb{R})-norm is concentrated on high frequencies, we will never see a significant part of L2()L^{2}(\mathbb{R})-norm of the solution qq moving to the low frequencies. That limits the “high-to-low frequency cascade” we discussed above. The close inspection of construction used in [8] shows that the function q0q_{0} in (1.3) has Hs()H^{s}(\mathbb{R})-norm smaller than ε\varepsilon but its L2()L^{2}(\mathbb{R})-norm is large when ε\varepsilon is small. Hence, the bounds in Theorem 1.2 do not contradict the estimates in (1.3) when s[1,12]s\in[-1,-\frac{1}{2}]. We do not know whether Theorem 1.2 holds for s<1s<-1.

The main idea of the proof of Theorem 1.2 is based on the analysis of the conserved quantity a(z)a(z), Imz>0\operatorname{Im}z>0, which is a coefficient in the transition matrix for the Dirac equation with potential q=q(,t)q=q(\cdot,t). We take z=iz=i and show that log|a(i)|\log|a(i)| is related to a certain quantity 𝒦~Q\widetilde{\mathcal{K}}_{Q} (see the Lemma 3.3 below) that characterizes both size and oscillation of qq. Using 𝒦~Q\widetilde{\mathcal{K}}_{Q} in the context of NLS is the main novelty of our work. We study 𝒦~Q\widetilde{\mathcal{K}}_{Q} and show that it is equivalent to H1()H^{-1}(\mathbb{R}) norm of qq with constants that depend on its L2()L^{2}(\mathbb{R})-norm. That gives the estimate (1.5) for s=1s=-1 and the intermediate range of s(1,0)s\in(-1,0) is handled by interpolation. Our analysis relies heavily on the recent results [2, 3, 1] that characterize Krein – de Branges canonical systems and the Dirac operators whose spectral measures belong to the Szegő class on the real line. We also establish the framework that allows working with NLS in the context of well-studied Krein systems.


Notation

  • The symbol II stands for 2×22\times 2 identity matrix I=(1001)I=\left(\begin{smallmatrix}1&0\\ 0&1\\ \end{smallmatrix}\right) and symbol JJ stands for J=(0110)J=\left(\begin{smallmatrix}0&-1\\ 1&0\end{smallmatrix}\right). Constant matrices σ3\sigma_{3}, σ±\sigma_{\pm}, σ\sigma are defined in (2.2).

  • For a measurable set SS\subset\mathbb{R}, we say that fLloc1(S)f\in L^{1}_{\rm\rm{loc}}(S) is fL1(K)f\in L^{1}(K) for every compact KSK\subset S.

  • The Fourier transform of a function ff is defined by

    (f)(η)=12πf(x)eiηx𝑑x.(\mathcal{F}f)(\eta)=\frac{1}{\sqrt{2\pi}}\int_{\mathbb{R}}f(x)e^{-i\eta x}\,dx.
  • The symbol CC, unless we specify explicitly, denotes the absolute constant which can change the value from formula to formula. If we write, e.g., C(α)C(\alpha), this defines a positive function of parameter α\alpha.

  • For two non-negative functions f1f_{1} and f2f_{2}, we write f1f2f_{1}\lesssim f_{2} if there is an absolute constant CC such that f1Cf2f_{1}\leqslant Cf_{2} for all values of the arguments of f1f_{1} and f2f_{2}. We define \gtrsim similarly and say that f1f2f_{1}\sim f_{2} if f1f2f_{1}\lesssim f_{2} and f2f1f_{2}\lesssim f_{1} simultaneously. If |f3|f4|f_{3}|\lesssim f_{4}, we will write f3=O(f4)f_{3}=O(f_{4}).

  • Symbols {ej}\{e_{j}\} are reserved for the standard basis in 2\mathbb{C}^{2}: e1=(10)e_{1}=\left(\begin{smallmatrix}1\\ 0\\ \end{smallmatrix}\right), e2=(01)e_{2}=\left(\begin{smallmatrix}0\\ 1\\ \end{smallmatrix}\right).

  • For matrix AA, the symbol AHS\|A\|_{\rm HS} denotes its Hilbert-Schmidt norm: AHS=(tr(AA))12\|A\|_{\rm HS}=({\rm tr}(A^{*}A))^{\frac{1}{2}}.


2. Preliminaries

Our proof of Theorem 1.1 uses complete integrability of equation (1.1). In that framework, (1.1) can be solved by using the method of inverse scattering which we discuss next following [13].

2.1. The inverse scattering approach to NLS

Given a complex-valued function q𝒮()q\in\mathcal{S}(\mathbb{R}), define the differential operator

Lq=iσ3ddξ+i(qσq¯σ+),L_{q}=i\sigma_{3}\frac{d}{d\xi}+i(q\sigma_{-}-\overline{q}\sigma_{+}), (2.1)

where we borrow notation for constant matrices σ3\sigma_{3}, σ±\sigma_{\pm} from [13]:

σ3=(1001),σ+=(0100),σ=(0010),σ=σ+σ+=(0110).\sigma_{3}=\begin{pmatrix}1&0\\ 0&-1\end{pmatrix},\qquad\sigma_{+}=\begin{pmatrix}0&1\\ 0&0\end{pmatrix},\qquad\sigma_{-}=\begin{pmatrix}0&0\\ 1&0\end{pmatrix},\qquad\sigma=\sigma_{-}+\sigma_{+}=\begin{pmatrix}0&1\\ 1&0\end{pmatrix}. (2.2)

The expression LqL_{q} is one of the forms in which the Dirac operator can be written. In Section 3, we will introduce another form and will show how the two are related. Let us also define

E(ξ,λ)=eλ2iξσ3=(eλ2iξ00eλ2iξ),E(\xi,\lambda)=e^{\frac{\lambda}{2i}\xi\sigma_{3}}=\begin{pmatrix}e^{\frac{\lambda}{2i}\xi}&0\\ 0&e^{-\frac{\lambda}{2i}\xi}\end{pmatrix},

as in [13]. In the free case when q=0q=0, the matrix-function EE solves L0E=λ2E,E(0,λ)=IL_{0}E={\frac{\lambda}{2}}E,\,E(0,\lambda)=I. Since q𝒮()q\in\mathcal{S}(\mathbb{R}), it decays at infinity fast and therefore one can find two solutions T±=T±(ξ,λ)T_{\pm}=T_{\pm}(\xi,\lambda) such that

LqT±=λ2T±,T±=E(ξ,λ)+o(1),ξ±,L_{q}T_{\pm}=\frac{\lambda}{2}T_{\pm},\qquad T_{\pm}=E(\xi,\lambda)+o(1),\qquad\xi\to\pm\infty, (2.3)

for every λ\lambda\in\mathbb{R}. These solutions are called the Jost solutions for LqL_{q}. Since both T+T_{+} and TT_{-} solve the same ODE, they must satisfy

T(ξ,λ)=T+(ξ,λ)T(λ),ξ,λ,T_{-}(\xi,\lambda)=T_{+}(\xi,\lambda)T(\lambda),\qquad\xi\in\mathbb{R},\qquad\lambda\in\mathbb{R}, (2.4)

where the matrix T=T(λ)T=T(\lambda) does not depend on ξ\xi\in\mathbb{R}. One can show that it has the form

T(λ)=(a(λ)b(λ)¯b(λ)a(λ)¯),detT=|a|2|b|2=1.T(\lambda)=\begin{pmatrix}a(\lambda)&\overline{b(\lambda)}\\ b(\lambda)&\overline{a(\lambda)}\end{pmatrix},\qquad\det T=|a|^{2}-|b|^{2}=1. (2.5)

The matrix TT is called the reduced transition matrix for LqL_{q}, and the ratio 𝐫q=b/a{\bf r}_{q}=b/a is called the reflection coefficient for LqL_{q}. One can obtain TT in a different way: let Zq=Zq(ξ,λ)Z_{q}=Z_{q}(\xi,\lambda), ξ\xi\in\mathbb{R}, λ\lambda\in\mathbb{C} be the fundamental matrix for LqL_{q}, that is,

LqZq=λ2Zq,Zq(0,λ)=(1001).L_{q}Z_{q}=\frac{\lambda}{2}Z_{q},\qquad Z_{q}(0,\lambda)=\left(\begin{smallmatrix}1&0\\ 0&1\\ \end{smallmatrix}\right). (2.6)

Then, we have Zq(ξ,λ)=T±(ξ,λ)T±1(0,λ)Z_{q}(\xi,\lambda)=T_{\pm}(\xi,\lambda)T^{-1}_{\pm}(0,\lambda) and the pointwise limits

T±1(0,λ)=limξ±E1(ξ,λ)Zq(ξ,λ)T^{-1}_{\pm}(0,\lambda)=\lim_{\xi\to\pm\infty}E^{-1}(\xi,\lambda)Z_{q}(\xi,\lambda) (2.7)

exist for every λ\lambda\in\mathbb{R}. Moreover, we have T(λ)=T+1(0,λ)T(0,λ)T(\lambda)=T_{+}^{-1}(0,\lambda)T_{-}(0,\lambda) on \mathbb{R}.

The coefficients a,ba,b, and 𝐫q{\bf r}_{q} were defined for λ\lambda\in\mathbb{R} and they satisfy |a|2=1+|b|2|a|^{2}=1+|b|^{2}, 1|𝐫q|2=|a|21-|{\bf r}_{q}|^{2}=|a|^{-2} for these λ\lambda. However, one can show that a(λ)a(\lambda) is the boundary value of the outer function defined in +={z:Imz>0}{\mathbb{C}}_{+}=\{z\in\mathbb{C}:\;\operatorname{Im}z>0\} by the formula (see (6.22) in [13])

a(z)=exp(1πi1λzlog|a(λ)|dλ),z+,a(z)=\exp\left(\frac{1}{\pi i}\int_{\mathbb{R}}\frac{1}{\lambda-z}\log|a(\lambda)|\,d\lambda\right),\quad z\in\mathbb{C}_{+},

which, in view of identity 1|𝐫q|2=|a|21-|{\bf r}_{q}|^{2}=|a|^{-2} on \mathbb{R}, can be written as

a(z)=exp(12πi1λzlog(1|𝐫q(λ)|2)𝑑λ).a(z)=\exp\left(-\frac{1}{2\pi i}\int_{\mathbb{R}}\frac{1}{\lambda-z}\log(1-|{\bf r}_{q}(\lambda)|^{2})d\lambda\right)\,. (2.8)

That shows, in particular, that bb defines both aa and 𝐫q{\bf r}_{q}, and 𝐫q{\bf r}_{q} defines aa and bb.

The map q𝐫qq\mapsto{\bf r}_{q} is called the direct scattering transform and its inverse is called the inverse scattering transform. These maps are well-studied when q𝒮()q\in\mathcal{S}(\mathbb{R}). In particular, we have the following result (see [13] for the proof).

Theorem 2.1.

The map q𝐫qq\mapsto{\bf r}_{q} is a bijection from 𝒮()\mathcal{S}(\mathbb{R}) onto the set of complex-valued functions {𝐫𝒮(),𝐫L()<1}\{{\bf r}\in\mathcal{S}(\mathbb{R}),\|{\bf r}\|_{L^{\infty}(\mathbb{R})}<1\}.

The scattering transform has some symmetries:

Lemma 2.1.

If q𝒮()q\in\mathcal{S}(\mathbb{R}) and λ\lambda\in\mathbb{R}, then

(dilation): 𝐫αq(αξ)(λ)=𝐫q(ξ)(α1λ),α>0,\displaystyle\qquad{\bf r}_{\alpha q(\alpha\xi)}(\lambda)={\bf r}_{q(\xi)}(\alpha^{-1}\lambda),\quad\alpha>0\,,
(conjugation): 𝐫q¯(ξ)(λ)=𝐫q(ξ)(λ)¯,\displaystyle\qquad{\bf r}_{\overline{q}(\xi)}(\lambda)={\overline{{\bf r}_{q(\xi)}(-\lambda)}}\,,
(translation): 𝐫q(ξ)(λ)=𝐫q(ξ)(λ)eiλ,,\displaystyle\qquad{\bf r}_{q(\xi-\ell)}(\lambda)={\bf r}_{q(\xi)}(\lambda)e^{-i\lambda\ell},\quad\ell\in\mathbb{R}\,,
(modulation): 𝐫eiβξq(ξ)(λ)=𝐫q(ξ)(λ+β),β.\displaystyle\qquad{\bf r}_{e^{-i\beta\xi}q(\xi)}(\lambda)={\bf r}_{q(\xi)}(\lambda+\beta),\quad\beta\in\mathbb{R}\,.
(rotation): 𝐫μq(ξ)(λ)=μ𝐫q(ξ)(λ),μ,|μ|=1.\displaystyle\qquad{\bf r}_{\mu q(\xi)}(\lambda)={\mu}{\bf r}_{q(\xi)}(\lambda),\quad\mu\in\mathbb{C},\;|\mu|=1\,.

Proof.  Indeed, the direct substitution into (2.3) shows that if T±(ξ,λ)T_{\pm}(\xi,\lambda) are Jost solutions for q(ξ)q(\xi), then

  • (a)(a)

    T±(αξ,α1λ)T_{\pm}(\alpha\xi,\alpha^{-1}\lambda) are the Jost solutions for αq(αξ)\alpha q(\alpha\xi),

  • (b)(b)

    T¯±(ξ,λ)\overline{T}_{\pm}(\xi,{-\lambda}) are the Jost solutions for q(ξ)¯\overline{q(\xi)},

  • (c)(c)

    T±(ξ,λ)E(,λ)T_{\pm}(\xi-\ell,{\lambda})E(\ell,\lambda) are the Jost solutions for q(ξ)q(\xi-\ell),

  • (d)(d)

    E(ξ,β)T±(ξ,λ+β)E(-\xi,\beta)T_{\pm}(\xi,\lambda+\beta) are the Jost solutions for eiβξq(ξ)e^{-i\beta\xi}q(\xi),

  • (e)(e)

    (100μ)T±(ξ,λ)(100μ¯)\left(\begin{smallmatrix}1&0\\ 0&\mu\end{smallmatrix}\right)T_{\pm}(\xi,\lambda)\left(\begin{smallmatrix}1&0\\ 0&\overline{\mu}\end{smallmatrix}\right) are the Jost solutions for μq(ξ)\mu q(\xi), |μ|=1|\mu|=1.

Now, it is left to use the formula (2.4) which defines TT. A computation using (2.5) shows how aa and bb change under symmetries (a)(a)(e)(e). For example, the translation does not change aa and it multiplies bb by eiλle^{-i\lambda l}. The modulation eiβξq(ξ)e^{-i\beta\xi}q(\xi), however, gives aeiβξq(ξ)(λ)=aq(ξ)(λ+β)a_{e^{-i\beta\xi}q(\xi)}(\lambda)=a_{q(\xi)}(\lambda+\beta). Then, the claim follows from the definition of the reflection coefficient 𝐫q=b/a{\bf r}_{q}=b/a. ∎

The next result (see formula (7.5)(7.5) in [13]), along with the previous theorem, shows how the inverse scattering transform can be used to solve (1.1).

Theorem 2.2.

Let q0𝒮()q_{0}\in\mathcal{S}(\mathbb{R}) and let 𝐫q0=𝐫q0(λ){\bf r}_{q_{0}}={\bf r}_{q_{0}}(\lambda) be the reflection coefficient of Lq0L_{q_{0}}. Define the family

𝐫(λ,t)=eiλ2t𝐫q0(λ),λ,t.{\bf r}(\lambda,t)=e^{-i\lambda^{2}t}{\bf r}_{q_{0}}(\lambda),\qquad\lambda\in\mathbb{R},\quad t\in\mathbb{R}. (2.9)

For each tt\in\mathbb{R}, let q=q(ξ,t)q=q(\xi,t) be the potential in the previous theorem generated by 𝐫(λ,t){\bf r}(\lambda,t). Then, q=q(ξ,t)q=q(\xi,t) is the unique classical solution of (1.1) with the initial datum q0q_{0}. Moreover, for every tt\in\mathbb{R}, the function ξq(ξ,t)\xi\mapsto q(\xi,t) lies in 𝒮()\mathcal{S}(\mathbb{R}).

The solutions to NLS equation

iqt=2qξ2+2|q|2qi\frac{\partial q}{\partial t}=-\frac{\partial^{2}q}{\partial\xi^{2}}+2|q|^{2}q (2.10)

behave in an explicit way under some transformations. Specifically, we have

  • (a)(a)

    Dilation: if q(ξ,t)q(\xi,t) solves (2.10), then αq(αξ,α2t)\alpha q(\alpha\xi,\alpha^{2}t) solves (2.10) for every α0\alpha\neq 0.

  • (b)(b)

    Time reversal: if q(ξ,t)q(\xi,t) solves (2.10), then q¯(ξ,t)\overline{q}(\xi,-t) solves (2.10). In particular, if q0q_{0} is real-valued, then q(ξ,t)=q(ξ,t)q(\xi,t)=q(\xi,-t).

  • (c)(c)

    Translation: if q(ξ,t)q(\xi,t) solves (2.10), then q(ξ,t)q(\xi-\ell,t) solves (2.10) for every \ell\in\mathbb{R}.

  • (d)(d)

    Modulation or Galilean symmetry: if q(ξ,t)q(\xi,t) solves (2.10), then eivξiv2tq(ξ2vt,t)e^{iv\xi-iv^{2}t}q(\xi-2vt,t) solves (2.10) for every vv\in\mathbb{R}.

  • (e)(e)

    Rotation: if q(ξ,t)q(\xi,t) solves (2.10), then μq(ξ,t)\mu q(\xi,t) solves (2.10) for every μ,|μ|=1\mu\in\mathbb{C},|\mu|=1.

These properties can be checked by direct calculation (see, e.g., formula (1.19) in [14] for (d)(d)) and a simple inspection shows that the bound (1.5) is consistent with all these transformations. The statements of Theorem 2.2 and Lemma 2.1 are consistent with these symmetries as well.

Now, we can explain the idea behind the proof of the Theorem 1.2.

The idea of the proof for Theorem 1.2. One can proceed as follows. First, we assume that q0𝒮()q_{0}\in\mathcal{S}(\mathbb{R}) and notice that conservation of |𝐫(λ,t)||{\bf r}(\lambda,t)|, λ\lambda\in\mathbb{R}, guaranteed by (2.9), yields that log|a(i,t)|\log|a(i,t)| is conserved, where a(z,t)a(z,t) is defined for z+z\in\mathbb{C}_{+} by (2.8). Separately, for every Dirac operator LqL_{q} with qL2()q\in L^{2}(\mathbb{R}), we show that log|a(i)|\log|a(i)| is equivalent to some explicit quantity 𝒦~Q\widetilde{\mathcal{K}}_{Q} that involves qq. That quantity was introduced and studied in [2, 3, 1]: it resembles the matrix Muckenhoupt A2()A_{2}(\mathbb{R}) condition and it is equivalent to H1()H^{-1}(\mathbb{R}) norm of qq provided that qL2()\|q\|_{L^{2}(\mathbb{R})} is under control, e.g., qL2()<C\|q\|_{L^{2}(\mathbb{R})}<C with some fixed CC. Putting things together, we see that Sobolev H1()H^{-1}(\mathbb{R}) norm of q(,t)q(\cdot,t) does not change much in time provided that the bound q(,t)L2()<C\|q(\cdot,t)\|_{L^{2}(\mathbb{R})}<C holds. Since q(,t)L2()=q0L2()\|q(\cdot,t)\|_{L^{2}(\mathbb{R})}=\|q_{0}\|_{L^{2}(\mathbb{R})} is time-invariant, we arrive to the statement of Theorem 1.2 for q0𝒮()q_{0}\in\mathcal{S}(\mathbb{R}) and s=1s=-1. For s=0s=0, the claim of Theorem 1.2 is trivial. The intermediate range of s(1,0)s\in(-1,0) is handled by interpolation using Galilean invariance of NLS. The general case when q0L2()q_{0}\in L^{2}(\mathbb{R}) follows by a density argument if one uses the stability of L2L^{2}-solutions guaranteed by Theorem 1.1.

There are other methods that use conserved quantities that agree with negative Sobolev norms. The paper [16] uses a representation of log|a(i)|\log|a(i)| through a perturbation determinant. Then, the analysis of the perturbation series allows the authors of [16] to obtain estimates similar to (1.4). It is conceivable that this approach can provide results along the same lines as Theorem 1.2.


To focus on the Dirac operator with qL2()q\in L^{2}(\mathbb{R}), we first consider this operator on half-line +\mathbb{R}_{+} in connection to Krein systems that were introduced in [19].

2.2. Operator LqL_{q} and Krein system

Let A:+A:\mathbb{R}_{+}\to\mathbb{C} be a function on the positive half-line +=[0,+)\mathbb{R}_{+}=[0,+\infty) such that

0r|A(ξ)|𝑑ξ<,\int_{0}^{r}|A(\xi)|\,d\xi<\infty,

for every r0r\geqslant 0. Recall that we denote the set of such functions by Lloc1(+)L^{1}_{\rm{loc}}(\mathbb{R}_{+}). The Krein system (see the formula (4.52)(4.52) in [12]) with the coefficient AA has the form

{P(ξ,λ)=iλP(ξ,λ)A(ξ)¯P(ξ,λ),P(0,λ)=1,P(ξ,λ)=A(ξ)P(ξ,λ),P(0,λ)=1,\begin{cases}P^{\prime}(\xi,\lambda)=i\lambda P(\xi,\lambda)-\overline{A(\xi)}P_{*}(\xi,\lambda),&P(0,\lambda)=1,\\ P^{\prime}_{*}(\xi,\lambda)=-A(\xi)P(\xi,\lambda),&P_{*}(0,\lambda)=1,\end{cases} (2.11)

where the derivative is taken with respect to ξ+\xi\in\mathbb{R}_{+} and λ\lambda\in\mathbb{C}. Let also

{P^(ξ,λ)=iλP^(ξ,λ)+A(ξ)¯P^(ξ,λ),P^(0,λ)=1,P^(ξ,λ)=A(ξ)P^(ξ,λ),P^(0,λ)=1,\begin{cases}\widehat{P}^{\prime}(\xi,\lambda)=i\lambda\widehat{P}(\xi,\lambda)+\overline{A(\xi)}\widehat{P}_{*}(\xi,\lambda),&\widehat{P}(0,\lambda)=1,\\ \widehat{P}^{\prime}_{*}(\xi,\lambda)=A(\xi)\widehat{P}(\xi,\lambda),&\widehat{P}_{*}(0,\lambda)=1,\end{cases} (2.12)

denote the so-called dual Krein system (see Corollary 5.7 in [12]). Set

Y(ξ,λ)=eiλξ(P(2ξ,λ)iP^(2ξ,λ)P(2ξ,λ)iP^(2ξ,λ)).Y(\xi,\lambda)=e^{-i\lambda\xi}\begin{pmatrix}P(2\xi,\lambda)&i\widehat{P}(2\xi,\lambda)\\ P_{*}(2\xi,\lambda)&-i\widehat{P}_{*}(2\xi,\lambda)\end{pmatrix}. (2.13)

The matrix-function ZqZ_{q}, which was defined in (2.6) for q𝒮()q\in\mathcal{S}(\mathbb{R}), makes sense if we assume that qLloc1()q\in L^{1}_{\rm loc}(\mathbb{R}). In the next lemma, we relate YY to ZqZ_{q}.

Lemma 2.2.

Let qLloc1()q\in L^{1}_{\rm{loc}}(\mathbb{R}), A(2ξ)=q(ξ)¯/2A(2\xi)=-\overline{q(\xi)}/2 on +\mathbb{R}_{+}, and YY be the corresponding matrix-valued function defined by (2.13). Then, Zq(ξ,2λ)=σY(ξ,λ)Y1(0,λ)σZ_{q}(\xi,2\lambda)=\sigma Y(\xi,\lambda)Y^{-1}(0,\lambda)\sigma for ξ0\xi\geqslant 0 and λ\lambda\in\mathbb{C}.

Proof.  The proof is a computation. We have

Lq¯Y(ξ,λ)=\displaystyle L_{\bar{q}}Y(\xi,\lambda)= λσ3Y(ξ,λ)+iσ3eiλξddξ(P(2ξ,λ)iP^(2ξ,λ)P(2ξ,λ)iP^(2ξ,λ))+i(q¯σqσ+)Y(ξ,λ),\displaystyle\lambda\sigma_{3}Y(\xi,\lambda)+i\sigma_{3}e^{-i\lambda\xi}\frac{d}{d\xi}\begin{pmatrix}P(2\xi,\lambda)&i\widehat{P}(2\xi,\lambda)\\ P_{*}(2\xi,\lambda)&-i\widehat{P}_{*}(2\xi,\lambda)\end{pmatrix}+i(\bar{q}\sigma_{-}-q\sigma_{+})Y(\xi,\lambda),
=\displaystyle= 2iσ3eiλξ(iλP(2ξ,λ)A(2ξ)¯P(2ξ,λ)λP^(2ξ,λ)+iA(2ξ)¯P^(2ξ,λ)A(2ξ)P(2ξ,λ)iA(2ξ)P^(2ξ,λ))\displaystyle 2i\sigma_{3}e^{-i\lambda\xi}\begin{pmatrix}i\lambda P(2\xi,\lambda)-\overline{A(2\xi)}P_{*}(2\xi,\lambda)&-\lambda\widehat{P}(2\xi,\lambda)+i\overline{A(2\xi)}\widehat{P}_{*}(2\xi,\lambda)\\ -A(2\xi)P(2\xi,\lambda)&-iA(2\xi)\widehat{P}(2\xi,\lambda)\end{pmatrix}
+i(q¯σqσ+iλσ3)Y(ξ,λ).\displaystyle+i(\bar{q}\sigma_{-}-q\sigma_{+}-i\lambda\sigma_{3})Y(\xi,\lambda).

Notice, that

i(q¯σqσ+iλσ3)Y(ξ,λ)\displaystyle i(\bar{q}\sigma_{-}-q\sigma_{+}-i\lambda\sigma_{3})Y(\xi,\lambda) =ieiλξ(iλqq¯iλ)(P(2ξ,λ)iP^(2ξ,λ)P(2ξ,λ)iP^(2ξ,λ))\displaystyle=ie^{-i\lambda\xi}\begin{pmatrix}-i\lambda&-q\\ \bar{q}&i\lambda\end{pmatrix}\begin{pmatrix}P(2\xi,\lambda)&i\widehat{P}(2\xi,\lambda)\\ P_{*}(2\xi,\lambda)&-i\widehat{P}_{*}(2\xi,\lambda)\end{pmatrix}
=ieiλξ(iλP(2ξ,λ)qP(2ξ,λ)λP^(2ξ,λ)+iqP^(2ξ,λ)q¯P(2ξ,λ)+iλP(2ξ,λ)iq¯P^(2ξ,λ)+λP^(2ξ,λ)).\displaystyle=ie^{-i\lambda\xi}\begin{pmatrix}-i\lambda P(2\xi,\lambda)-qP_{*}(2\xi,\lambda)&\lambda\widehat{P}(2\xi,\lambda)+iq\widehat{P}_{*}(2\xi,\lambda)\\ \bar{q}P(2\xi,\lambda)+i\lambda P_{*}(2\xi,\lambda)&i\bar{q}\widehat{P}(2\xi,\lambda)+\lambda\widehat{P}_{*}(2\xi,\lambda)\end{pmatrix}\,.

Using relation 2A(2ξ)+q¯(ξ)=02A(2\xi)+\bar{q}(\xi)=0, we obtain

Lq¯Y(ξ,λ)=ieiλξ(iλP(2ξ,λ)λP^(2ξ,λ)iλP(2ξ,λ)λP^(2ξ,λ))=λY(ξ,λ).L_{\bar{q}}Y(\xi,\lambda)=ie^{-i\lambda\xi}\begin{pmatrix}i\lambda P(2\xi,\lambda)&-\lambda\widehat{P}(2\xi,\lambda)\\ i\lambda P_{*}(2\xi,\lambda)&\lambda\widehat{P}_{*}(2\xi,\lambda)\end{pmatrix}=-\lambda Y(\xi,\lambda).

Since σσ3σ=σ3\sigma\sigma_{3}\sigma=-\sigma_{3} and σσ±σ=σ\sigma\sigma_{\pm}\sigma=\sigma_{\mp}, one has σLq¯σ=Lq\sigma L_{\bar{q}}\sigma=-L_{q}. Therefore,

Lq(σY(ξ,λ)σ)=λ(σY(ξ,λ)σ).L_{q}(\sigma Y(\xi,\lambda)\sigma)=\lambda(\sigma Y(\xi,\lambda)\sigma).

It follows that matrix-valued functions Zq(ξ,2λ)Z_{q}(\xi,2\lambda) and σY(ξ,λ)Y1(0,λ)σ\sigma Y(\xi,\lambda)Y^{-1}(0,\lambda)\sigma solve the same Cauchy problem and thus Zq(ξ,2λ)=σY(ξ,λ)Y1(0,λ)σZ_{q}(\xi,2\lambda)=\sigma Y(\xi,\lambda)Y^{-1}(0,\lambda)\sigma, as required. ∎

Lemma 2.3.

Let qLloc1()q\in L^{1}_{\rm{loc}}(\mathbb{R}), let A(2ξ)=q(ξ)/2A(2\xi)=q(-\xi)/2 on +\mathbb{R}_{+}, and let YY be the corresponding matrix-valued function defined by (2.13). Then, Zq(ξ,2λ)=Y(ξ,λ)Y1(0,λ)Z_{q}(-\xi,2\lambda)=Y(\xi,\lambda)Y^{-1}(0,\lambda) for ξ0\xi\geqslant 0 and λ\lambda\in\mathbb{C}.

Proof.  Recall that matrices σ3\sigma_{3}, σ±\sigma_{\pm}, σ\sigma are defined in (2.2). Using relations σσ3σ=σ3\sigma\sigma_{3}\sigma=-\sigma_{3} and σσ±σ=σ\sigma\sigma_{\pm}\sigma=\sigma_{\mp}, we see that Lq~Z~q=λ2Z~qL_{\tilde{q}}\widetilde{Z}_{q}=\frac{\lambda}{2}\widetilde{Z}_{q}, where q~(ξ)=q(ξ)¯\tilde{q}(\xi)=-\overline{q(-\xi)} and Z~q(ξ,λ)=σZq(ξ,λ)σ\widetilde{Z}_{q}(\xi,\lambda)=\sigma Z_{q}(-\xi,\lambda)\sigma. Then, previous lemma applies to q~\tilde{q}, Zq~(ξ,2λ)=Z~q(ξ,2λ)Z_{\tilde{q}}(\xi,2\lambda)=\widetilde{Z}_{q}(\xi,2\lambda) and A(2ξ)=q~(ξ)¯/2=q(ξ)/2A(2\xi)=-\overline{\tilde{q}(\xi)}/2=q(-\xi)/2. It gives Z~q(ξ,2λ)=σY(ξ,λ)Y1(0,λ)σ\widetilde{Z}_{q}(\xi,2\lambda)=\sigma Y(\xi,\lambda)Y^{-1}(0,\lambda)\sigma. Returning to ZqZ_{q}, we get Zq(ξ,2λ)=Y(ξ,λ)Y1(0,λ)Z_{q}(-\xi,2\lambda)=Y(\xi,\lambda)Y^{-1}(0,\lambda). ∎

Given qL2()q\in L^{2}(\mathbb{R}), we define the continuous analogs of Wall polynomials (see [15] and Section 7 in [12]) by

𝔄±=P±+P^±2,𝔄±=P±+P^±2,𝔅±=P±P^±2,𝔅±=P±P^±2,\mathfrak{A}^{\pm}=\frac{P^{\pm}_{*}+\widehat{P}^{\pm}_{*}}{2},\qquad\mathfrak{A}^{\pm}_{*}=\frac{P^{\pm}+\widehat{P}^{\pm}}{2},\qquad\mathfrak{B}^{\pm}=\frac{P^{\pm}_{*}-\widehat{P}^{\pm}_{*}}{2},\qquad\mathfrak{B}^{\pm}_{*}=\frac{P^{\pm}-\widehat{P}^{\pm}}{2}, (2.14)

where P±P^{\pm}, P±P^{\pm}_{*}, P^±\widehat{P}^{\pm}, P^±\widehat{P}^{\pm}_{*} are the solutions of systems (2.11), (2.12) for the coefficient A+(ξ)=q(ξ/2)¯/2A^{+}(\xi)=-\overline{q(\xi/2)}/2 from Lemma 2.2 and the coefficient A(ξ)=q(ξ/2)/2A^{-}(\xi)=q(-\xi/2)/2 from Lemma 2.3, correspondingly. Functions P±P^{\pm}, P±P^{\pm}_{*}, P^±\widehat{P}^{\pm}, P^±\widehat{P}^{\pm}_{*} are continuous analogs of polynomials orthogonal on the unit circle, they depend on two parameters: ξ+\xi\in\mathbb{R}_{+} and λ\lambda\in\mathbb{C} and they satisfy identities (see formula (4.32) in [12]):

P±(ξ,λ)=eiξλP±(ξ,λ)¯,P^±(ξ,λ)=eiξλP^±(ξ,λ)¯P_{*}^{\pm}(\xi,\lambda)=e^{i\xi\lambda}\overline{P^{\pm}(\xi,\lambda)},\qquad\widehat{P}_{*}^{\pm}(\xi,\lambda)=e^{i\xi\lambda}\overline{\widehat{P}^{\pm}(\xi,\lambda)} (2.15)

for real λ\lambda.

We will use the following result (see Lemma 2 in [11] which contains a stronger statement).

Theorem 2.3.

Let AL2(+)A\in L^{2}(\mathbb{R}_{+}), and let PP, PP_{*} be the solutions of system (2.11) for the coefficient AA. Then, the limit

Π(λ)=limξ+P(ξ,λ)\Pi(\lambda)=\lim_{\xi\to+\infty}P_{*}(\xi,\lambda) (2.16)

exists for every λ+\lambda\in\mathbb{C}_{+}. That function Π\Pi is outer in +\mathbb{C}_{+}. If λ\lambda\in\mathbb{R}, the convergence in (2.16) holds in the Lebesgue measure on \mathbb{R} where Π(λ)\Pi(\lambda) now denotes the non-tangential boundary value of Π\Pi.

That theorem allows us to define

𝔞±(λ)=limξ+𝔄±(ξ,λ),𝔟±(λ)=limξ+𝔅±(ξ,λ)\mathfrak{a}^{\pm}(\lambda)=\lim_{\xi\to+\infty}\mathfrak{A}^{\pm}(\xi,\lambda),\qquad\mathfrak{b}^{\pm}(\lambda)=\lim_{\xi\to+\infty}\mathfrak{B}^{\pm}(\xi,\lambda) (2.17)

for every λ+\lambda\in\mathbb{C}_{+} and for almost every λ\lambda\in\mathbb{R}. Moreover, Corollary 12.2 in [12] gives

|𝔞±(λ)|2=1+|𝔟±(λ)|2|\mathfrak{a}^{\pm}(\lambda)|^{2}=1+|\mathfrak{b}^{\pm}(\lambda)|^{2} (2.18)

for a.e. λ\lambda\in\mathbb{R}. For every λ+\lambda\in\mathbb{C}_{+}, we define

a(λ)=𝔞+(λ)𝔞(λ)𝔟+(λ)𝔟(λ).a(\lambda)=\mathfrak{a}^{+}(\lambda)\mathfrak{a}^{-}(\lambda)-\mathfrak{b}^{+}(\lambda)\mathfrak{b}^{-}(\lambda)\,.
Proposition 2.1.

The function aa is outer in +\mathbb{C}_{+}.

Proof.  We can write

a=𝔞+𝔞(1s+s),s±=𝔟±𝔞±.a=\mathfrak{a}^{+}\mathfrak{a}^{-}(1-s^{+}s^{-}),\quad s^{\pm}=\frac{\mathfrak{b}^{\pm}}{\mathfrak{a}^{\pm}}\,.

It is known that 𝔞±\mathfrak{a}^{\pm} are outer (see the formulas (12.9) and (12.29) in [12]) and that s±s^{\pm} satisfy |s±|<1|s^{\pm}|<1 in +\mathbb{C}_{+}. The function 1s+s1-s^{+}s^{-} has a positive real part in +\mathbb{C}_{+} and so is an outer function. That shows that aa is a product of three outer functions and hence it is outer itself. ∎

Proposition 2.2.

Let qL2()q\in L^{2}(\mathbb{R}) and let ZqZ_{q} be defined by (2.6). Then, the limits in (2.7) exist in the Lebesgue measure on \mathbb{R}. The matrix T(λ)=T+1(0,λ)T(0,λ)T(\lambda)=T_{+}^{-1}(0,\lambda)T_{-}(0,\lambda) has the form (2.5) where

a=𝔞+𝔞𝔟+𝔟,b=𝔞𝔟+¯𝔟𝔞+¯,a=\mathfrak{a}^{+}\mathfrak{a}^{-}-\mathfrak{b}^{+}\mathfrak{b}^{-},\qquad b=\mathfrak{a}^{-}\overline{\mathfrak{b}^{+}}-\mathfrak{b}^{-}\overline{\mathfrak{a}^{+}}\,, (2.19)

and 𝔞±\mathfrak{a}^{\pm}, 𝔟±\mathfrak{b}^{\pm} are defined Lebesgue almost everywhere on \mathbb{R} by the convergence in (2.17) in measure.

Proof.  If qL2()q\in L^{2}(\mathbb{R}), the fundamental matrix ZqZ_{q} and the continuous Wall polynomials (2.14) are related by the formula

Zq(ξ,2λ)={eiλξ(𝔄+(2ξ,λ)𝔅+(2ξ,λ)𝔅+(2ξ,λ)𝔄+(2ξ,λ)),ξ0,eiλξ(𝔄(2ξ,λ)𝔅(2ξ,λ)𝔅(2ξ,λ)𝔄(2ξ,λ)),ξ<0.\displaystyle Z_{q}(\xi,2\lambda)=\begin{cases}e^{-i\lambda\xi}\begin{pmatrix}\mathfrak{A}^{+}(2\xi,\lambda)&\mathfrak{B}^{+}(2\xi,\lambda)\\ \mathfrak{B}^{+}_{*}(2\xi,\lambda)&\mathfrak{A}^{+}_{*}(2\xi,\lambda)\end{pmatrix},&\xi\geqslant 0,\\ e^{i\lambda\xi}\begin{pmatrix}\mathfrak{A}^{-}_{*}(-2\xi,\lambda)&\mathfrak{B}^{-}_{*}(-2\xi,\lambda)\\ \mathfrak{B}^{-}(-2\xi,\lambda)&\mathfrak{A}^{-}(-2\xi,\lambda)\end{pmatrix},&\xi<0.\end{cases} (2.20)

Indeed, it is enough to use Lemma 2.2, Lemma 2.3, and the fact that Y1(0,λ)=12(11ii)Y^{-1}(0,\lambda)=\frac{1}{2}\left(\begin{smallmatrix}1&1\\ -i&i\end{smallmatrix}\right). Our next step is to prove that the limit

T+1(0,2λ)=limξ+E1(ξ,2λ)Zq(ξ,2λ)T_{+}^{-1}(0,2\lambda)=\lim_{\xi\to+\infty}E^{-1}(\xi,2\lambda)Z_{q}(\xi,2\lambda) (2.21)

exists in Lebesgue measure when λ\lambda\in\mathbb{R}. From (2.15), we obtain

E1(ξ,2λ)Zq(ξ,2λ)=(100e2iλξ)(𝔄+(2ξ,λ)𝔅+(2ξ,λ)𝔅+(2ξ,λ)𝔄+(2ξ,λ))=(𝔄+(2ξ,λ)𝔅+(2ξ,λ)𝔅+¯(2ξ,λ)𝔄+¯(2ξ,λ)),\displaystyle E^{-1}(\xi,2\lambda)Z_{q}(\xi,2\lambda)=\begin{pmatrix}1&0\\ 0&e^{-2i\lambda\xi}\end{pmatrix}\begin{pmatrix}\mathfrak{A}^{+}(2\xi,\lambda)&\mathfrak{B}^{+}(2\xi,\lambda)\\ \mathfrak{B}^{+}_{*}(2\xi,\lambda)&\mathfrak{A}^{+}_{*}(2\xi,\lambda)\end{pmatrix}=\begin{pmatrix}\mathfrak{A}^{+}(2\xi,\lambda)&\mathfrak{B}^{+}(2\xi,\lambda)\\ \overline{\mathfrak{B}^{+}}(2\xi,\lambda)&\overline{\mathfrak{A}^{+}}(2\xi,\lambda)\end{pmatrix},

for every ξ0\xi\geqslant 0 and λ\lambda\in\mathbb{R}. Similarly,

E1(ξ,2λ)Zq(ξ,2λ)=(e2iλξ001)(𝔄(2ξ,λ)𝔅(2ξ,λ)𝔅(2ξ,λ)𝔄(2ξ,λ))=(𝔄¯(2ξ,λ)𝔅¯(2ξ,λ)𝔅(2ξ,λ)𝔄(2ξ,λ)).\displaystyle E^{-1}(-\xi,2\lambda)Z_{q}(-\xi,2\lambda)=\begin{pmatrix}e^{-2i\lambda\xi}&0\\ 0&1\end{pmatrix}\begin{pmatrix}\mathfrak{A}^{-}_{*}(2\xi,\lambda)&\mathfrak{B}^{-}_{*}(2\xi,\lambda)\\ \mathfrak{B}^{-}(2\xi,\lambda)&\mathfrak{A}^{-}(2\xi,\lambda)\end{pmatrix}=\begin{pmatrix}\overline{\mathfrak{A}^{-}}(2\xi,\lambda)&\overline{\mathfrak{B}^{-}}(2\xi,\lambda)\\ \mathfrak{B}^{-}(2\xi,\lambda)&\mathfrak{A}^{-}(2\xi,\lambda)\end{pmatrix}.

Hence, the limits

T±1(0,2λ)=limξ±E1(ξ,2λ)Zq(ξ,2λ)T^{-1}_{\pm}(0,2\lambda)=\lim_{\xi\to\pm\infty}E^{-1}(\xi,2\lambda)Z_{q}(\xi,2\lambda) (2.22)

exist in Lebesgue measure on \mathbb{R} by Theorem 2.3. Moreover,

T(2λ)\displaystyle T(2\lambda) =T+1(0,2λ)T(0,2λ)=(𝔞+(λ)𝔟+(λ)𝔟+(λ)¯𝔞+(λ)¯)(𝔞(λ)¯𝔟(λ)¯𝔟(λ)𝔞(λ))1\displaystyle=T_{+}^{-1}(0,2\lambda)T_{-}(0,2\lambda)=\begin{pmatrix}\mathfrak{a}^{+}(\lambda)&\mathfrak{b}^{+}(\lambda)\\ \overline{\mathfrak{b}^{+}(\lambda)}&\overline{\mathfrak{a}^{+}(\lambda)}\end{pmatrix}\begin{pmatrix}\overline{\mathfrak{a}^{-}(\lambda)}&\overline{\mathfrak{b}^{-}(\lambda)}\\ \mathfrak{b}^{-}(\lambda)&\mathfrak{a}^{-}(\lambda)\end{pmatrix}^{-1}
=(2.18)(𝔞+(λ)𝔟+(λ)𝔟+(λ)¯𝔞+(λ)¯)(𝔞(λ)𝔟(λ)¯𝔟(λ)𝔞(λ)¯)=(a(λ)b(λ)¯b(λ)a(λ)¯)\displaystyle\stackrel{{\scriptstyle\eqref{sa5}}}{{=}}\begin{pmatrix}\mathfrak{a}^{+}(\lambda)&\mathfrak{b}^{+}(\lambda)\\ \overline{\mathfrak{b}^{+}(\lambda)}&\overline{\mathfrak{a}^{+}(\lambda)}\end{pmatrix}\begin{pmatrix}\mathfrak{a}^{-}(\lambda)&-\overline{\mathfrak{b}^{-}(\lambda)}\\ -\mathfrak{b}^{-}(\lambda)&\overline{\mathfrak{a}^{-}(\lambda)}\end{pmatrix}=\begin{pmatrix}a(\lambda)&\overline{b(\lambda)}\\ b(\lambda)&\overline{a(\lambda)}\end{pmatrix}

and the proposition follows. ∎

We end this section with a few remarks on reflection coefficients of potentials in L2()L^{2}(\mathbb{R}). We have |a|2|b|2=1|a|^{2}-|b|^{2}=1 almost everywhere on \mathbb{R} due to the fact that detT±(0,λ)=1\det T_{\pm}(0,\lambda)=1 almost everywhere on \mathbb{R}. That can also be established directly using (2.18). Proposition 2.2 then allows to define the reflection coefficient 𝐫q=b/a{\bf r}_{q}=b/a for every qL2()q\in L^{2}(\mathbb{R}). The Lemma 2.1 holds for 𝐫q{\bf r}_{q} in that case as well. However, not all results about scattering transform can be generalized from the case q𝒮()q\in\mathcal{S}(\mathbb{R}) to qL2()q\in L^{2}(\mathbb{R}). For example, scattering transform is injective on 𝒮()\mathcal{S}(\mathbb{R}) by Theorem 2.1, but it is not longer so when extended to L2()L^{2}(\mathbb{R}) (see Example 6.1 in Appendix).

3. Another form of Dirac operator, qL2()q\in L^{2}(\mathbb{R}), and the entropy function.

Suppose qL2()q\in L^{2}(\mathbb{R}). The alternative to LqL_{q} form of writing Dirac operator on the line is given by an expression

𝒟Q:XJX+QX,Q=(ImqReqReqImq).\mathcal{D}_{Q}:X\mapsto JX^{\prime}+QX,\qquad Q=\begin{pmatrix}-\operatorname{Im}q&-\operatorname{Re}q\\ -\operatorname{Re}q&\operatorname{Im}q\end{pmatrix}\,. (3.1)

𝒟Q\mathcal{D}_{Q} is densely defined self-adjoint operator on the Hilbert space L2(,2)L^{2}(\mathbb{R},\mathbb{C}^{2}) of functions X:2X:\mathbb{R}\to\mathbb{C}^{2} such that XL2(,2)2=X(ξ)22𝑑ξ\|X\|_{L^{2}(\mathbb{R},\mathbb{C}^{2})}^{2}=\int_{\mathbb{R}}\|X(\xi)\|^{2}_{\mathbb{C}^{2}}\,d\xi is finite. 𝒟Q\mathcal{D}_{Q} and LqL_{q} defined in (2.1) are related by a simple formula:

𝒟Q=ΣLqΣ1,Σ=12(11ii),Σ1=12(1i1i).\mathcal{D}_{Q}=\Sigma L_{q}\Sigma^{-1},\quad\Sigma=\frac{1}{\sqrt{2}}\left(\begin{matrix}1&1\\ -i&i\end{matrix}\right),\quad\Sigma^{-1}=\frac{1}{\sqrt{2}}\left(\begin{matrix}1&i\\ 1&-i\end{matrix}\right)\,.

One way to study 𝒟Q\mathcal{D}_{Q} is to focus on Dirac operators on half-line +\mathbb{R}_{+} first. Given qL2(+)q\in L^{2}(\mathbb{R}_{+}), we define 𝒟Q+\mathcal{D}^{+}_{Q} on L2(+,2)L^{2}(\mathbb{R}_{+},\mathbb{C}^{2}) by

𝒟Q+:XJX+QX,Q=(ImqReqReqImq)\mathcal{D}^{+}_{Q}:X\mapsto JX^{\prime}+QX,\qquad Q=\begin{pmatrix}-\operatorname{Im}q&-\operatorname{Re}q\\ -\operatorname{Re}q&\operatorname{Im}q\end{pmatrix} (3.2)

on the dense subset of absolutely continuous functions XL2(+,2)X\in L^{2}(\mathbb{R}_{+},\mathbb{C}^{2}) such that 𝒟Q+XL2(+,2)\mathcal{D}^{+}_{Q}X\in L^{2}(\mathbb{R}_{+},\mathbb{C}^{2}), X(0)=(0)X(0)=\left(\begin{smallmatrix}*\\ 0\end{smallmatrix}\right). We will call 𝒟Q+\mathcal{D}^{+}_{Q} the Dirac operator defined on the positive half-line with boundary conditions X(0)=(0)X(0)=\left(\begin{smallmatrix}*\\ 0\end{smallmatrix}\right) or simply the half-line Dirac operator. Set A(ξ)=q(ξ/2)¯/2A(\xi)=-\overline{q(\xi/2)}/2 for ξ+\xi\in\mathbb{R}_{+}, and let P(ξ,λ)P(\xi,\lambda), P(ξ,λ)P_{*}(\xi,\lambda) be the solutions of Krein system (2.11) generated by AA. The Krein system with coefficient AA and Dirac equation (3.2) are related as follows (see the proof of Lemma 6.1 in Appendix): if NQN_{Q} solves the Cauchy problem JNQ(ξ,λ)+Q(ξ)NQ(ξ,λ)=λNQ(ξ,λ)JN^{\prime}_{Q}(\xi,\lambda)+Q(\xi)N_{Q}(\xi,\lambda)=\lambda N_{Q}(\xi,\lambda), NQ(0,λ)=(1001)N_{Q}(0,\lambda)=\left(\begin{smallmatrix}1&0\\ 0&1\\ \end{smallmatrix}\right), then

NQ(ξ,λ)=eiλξ2(11ii)(𝔄+(2ξ,λ)𝔅+(2ξ,λ)𝔅+(2ξ,λ)𝔄+(2ξ,λ))(1i1i),N_{Q}(\xi,\lambda)=\frac{e^{-i\lambda\xi}}{2}\left(\begin{matrix}1&1\\ i&-i\end{matrix}\right)\left(\begin{matrix}\mathfrak{A}_{*}^{+}(2\xi,\lambda)&\mathfrak{B}_{*}^{+}(2\xi,\lambda)\\ \mathfrak{B}^{+}(2\xi,\lambda)&\mathfrak{A}^{+}(2\xi,\lambda)\end{matrix}\right)\left(\begin{matrix}1&-i\\ 1&i\end{matrix}\right)\,,

where the continuous Wall polynomials 𝔄+,𝔅+,𝔄+,𝔅+\mathfrak{A}^{+},\mathfrak{B}^{+},\mathfrak{A}_{*}^{+},\mathfrak{B}^{+}_{*} were defined in (2.14). The Weyl function of the operator 𝒟Q+\mathcal{D}^{+}_{Q} coincides (see Lemma 6.1 in Appendix) with

mQ(z)=limξ+iP^(ξ,z)P(ξ,z),z+.m_{Q}(z)=\lim_{\xi\to+\infty}i\frac{\widehat{P}_{*}(\xi,z)}{P_{*}(\xi,z)},\qquad z\in\mathbb{C}_{+}. (3.3)

It is known (see Theorem 7.3 in [12]) that the limit above exists for every z+z\in\mathbb{C}_{+} and defines an analytic function of Herglotz-Nevanlinna class in +\mathbb{C}_{+}. The latter means that mQ(+)+m_{Q}(\mathbb{C}_{+})\subset\mathbb{C}_{+}. In the next theorem, ImmQ(λ)\operatorname{Im}m_{Q}(\lambda) denotes the nontangential boundary value on \mathbb{R} which exists Lebesgue almost everywhere. It is understood as a nonnegative function g=Immg=\operatorname{Im}m on \mathbb{R} and it satisfies g/(1+λ2)L1()g/(1+\lambda^{2})\in L^{1}(\mathbb{R}).

Theorem 3.1.

Let qL2(+)q\in L^{2}(\mathbb{R}_{+}) and let QQ, 𝒟Q+\mathcal{D}^{+}_{Q}, mQm_{Q} be defined by (3.2), (3.3). Denote by NQN_{Q} the solution of the Cauchy problem JNQ(ξ)+Q(ξ)NQ(ξ)=0JN^{\prime}_{Q}(\xi)+Q(\xi)N_{Q}(\xi)=0, NQ(0)=(1001)N_{Q}(0)=\left(\begin{smallmatrix}1&0\\ 0&1\\ \end{smallmatrix}\right), and set Q=NQNQ\mathcal{H}_{Q}=N_{Q}^{*}N_{Q}. Define also

𝒦Q+\displaystyle\mathcal{K}^{+}_{Q} =logImmQ(i)1πlogImmQ(λ)dλλ2+1,\displaystyle=\log\operatorname{Im}m_{Q}(i)-\frac{1}{\pi}\int_{\mathbb{R}}\log\operatorname{Im}m_{Q}(\lambda)\frac{d\lambda}{\lambda^{2}+1}, (3.4)
𝒦~Q+\displaystyle\widetilde{\mathcal{K}}^{+}_{Q} =k=0+(detkk+2Q(ξ)𝑑ξ4).\displaystyle=\sum_{k=0}^{+\infty}\left(\det\int_{k}^{k+2}\mathcal{H}_{Q}(\xi)\,d\xi-4\right). (3.5)

Then, we have

c1𝒦Q+𝒦~Q+c2ec2𝒦Q+c_{1}\mathcal{K}_{Q}^{+}\leqslant\widetilde{\mathcal{K}}_{Q}^{+}\leqslant c_{2}e^{c_{2}\mathcal{K}_{Q}^{+}} (3.6)

for some positive absolute constants c1c_{1}, c2c_{2}.

Proof.  Lemma 6.1 in Appendix shows that mQm_{Q} coincides with the Weyl function for the canonical system with Hamiltonian Q\mathcal{H}_{Q}. Then, the bounds in (3.6) follow from the Theorem 1.2 in [3] (see also Corollary 1.4 in [3]).∎

The quantity 𝒦Q+\mathcal{K}^{+}_{Q} will be called the entropy of the Dirac operator on +\mathbb{R}_{+}. We now turn to (3.1) to define the entropy for the Dirac operator on the whole line. Take qL2()q\in L^{2}(\mathbb{R}) and let A+(ξ)=q(ξ/2)¯/2A^{+}(\xi)=-\overline{q(\xi/2)}/2 and A(ξ)=q(ξ/2)/2A^{-}(\xi)=q(-\xi/2)/2, ξ+\xi\in\mathbb{R}_{+} be the coefficients of Krein systems associated to restrictions of qq to the half-lines +\mathbb{R}_{+} and \mathbb{R}_{-}. As in (3.3), the half-line Weyl functions m±m_{\pm} satisfy

m±(z)=limξ+iP^±(ξ,z)P±(ξ,z),z+.m_{\pm}(z)=\lim_{\xi\to+\infty}i\frac{\widehat{P}_{*}^{\pm}(\xi,z)}{P_{*}^{\pm}(\xi,z)},\qquad z\in\mathbb{C}_{+}. (3.7)

These Weyl functions m±m_{\pm} can be used to construct the spectral representation for the Dirac operator. Let

m(z)=1m+(z)+m(z)(2m+(z)m(z)m+(z)m(z)m+(z)m(z)2),z+.m(z)=-\frac{1}{m_{+}(z)+m_{-}(z)}\begin{pmatrix}-2m_{+}(z)m_{-}(z)&m_{+}(z)-m_{-}(z)\\ m_{+}(z)-m_{-}(z)&2\end{pmatrix},\qquad z\in\mathbb{C}_{+}. (3.8)

Using Imm±(z)>0\operatorname{Im}m_{\pm}(z)>0, one can show that Imm(z)\operatorname{Im}m(z) is a positive definite matrix for z+z\in\mathbb{C}_{+}. In other words, mm is the matrix-valued Herglotz function. Therefore, there exists a unique matrix-valued measure ρ\rho taking Borel subsets of \mathbb{R} into 2×22\times 2 nonnegative matrices such that

m(z)=α+βz+1π(1λzλλ2+1)𝑑ρ(λ),z+,m(z)=\alpha+\beta z+\frac{1}{\pi}\int_{\mathbb{R}}\left(\frac{1}{\lambda-z}-\frac{\lambda}{\lambda^{2}+1}\right)\,d\rho(\lambda),\qquad z\in\mathbb{C}_{+},

where α\alpha, β\beta are constant 2×22\times 2 real matrices, β0\beta\geqslant 0. The importance of ρ\rho becomes clear when we recall the spectral decomposition for 𝒟Q\mathcal{D}_{Q}. Specifically, let NQ(ξ,z)N_{Q}(\xi,z) be the solution of the Cauchy problem

JξNQ(ξ,z)+Q(ξ)NQ(ξ,z)=zNQ(ξ,z),NQ(0,z)=(1001),z,ξ.J\frac{\partial}{\partial\xi}N_{Q}(\xi,z)+Q(\xi)N_{Q}(\xi,z)=z{N}_{Q}(\xi,z),\quad\quad N_{Q}(0,z)=\left(\begin{smallmatrix}1&0\\ 0&1\\ \end{smallmatrix}\right),\quad z\in\mathbb{C},\quad\xi\in\mathbb{R}\,. (3.9)

Then, the mapping

𝒟Q:X1πNQ(ξ,λ)X(ξ)𝑑ξ,λ,\mathcal{F}_{\mathcal{D}_{Q}}:X\mapsto\frac{1}{\sqrt{\pi}}\int_{\mathbb{R}}N_{Q}^{*}(\xi,\lambda)X(\xi)\,d\xi,\qquad\lambda\in\mathbb{R}, (3.10)

initially defined on the set of compactly supported smooth functions X:2X:\mathbb{R}\to\mathbb{C}^{2}, extends (see Appendix) to the unitary operator between the Hilbert spaces L2(,2)L^{2}(\mathbb{R},\mathbb{C}^{2}) and L2(ρ)L^{2}(\rho),

L2(ρ)\displaystyle L^{2}(\rho) ={Y:2:YL2(ρ)2=Y(λ)𝑑ρ(λ)Y(λ)<}.\displaystyle=\Bigl{\{}Y:\mathbb{R}\to\mathbb{C}^{2}:\;\|Y\|^{2}_{L^{2}(\rho)}=\int_{\mathbb{R}}Y^{*}(\lambda)\,d\rho(\lambda)\,Y(\lambda)<\infty\Bigr{\}}.

Moreover, 𝒟Q\mathcal{D}_{Q} is unitary equivalent to the operator of multiplication by the independent variable in L2(ρ)L^{2}(\rho) and the unitary equivalence is given by the operator Q\mathcal{F}_{Q}. In fact, these properties of ρ\rho will not be used in the paper, we mention them only to motivate the following definition. Let us define the entropy function 𝒦Q(z)\mathcal{K}_{Q}(z) by

𝒦Q(z)=1πlog(detρac(λ))Imz|λz|2𝑑λ,z+,\mathcal{K}_{Q}(z)=-\frac{1}{\pi}\int_{\mathbb{R}}\log(\det\rho_{ac}(\lambda))\frac{\operatorname{Im}z}{|\lambda-z|^{2}}\,d\lambda,\qquad z\in\mathbb{C}_{+}, (3.11)

where ρac\rho_{ac} denotes the absolutely continuous part of the spectral measure ρ\rho and it satisfies ρac(λ)=limε0,ε>0Imm(λ+iε)\rho_{ac}(\lambda)=\lim_{\varepsilon\to 0,\varepsilon>0}\operatorname{Im}m(\lambda+i\varepsilon) for a.e. λ\lambda\in\mathbb{R}. The quantity 𝒦Q\mathcal{K}_{Q} will play a crucial role in our considerations. We first relate it to the coefficient aa of the reduced transition matrix TT which was introduced in Proposition 2.2.

Lemma 3.1.

We have detρac(λ)=|a(λ)|2\det\rho_{ac}(\lambda)=|a(\lambda)|^{-2} for almost all λ\lambda\in\mathbb{R}. In particular, 𝒦Q(z)=2log|a(z)|\mathcal{K}_{Q}(z)=2\log|a(z)| for all z+z\in\mathbb{C}_{+}.

Proof.  From the definition (or see page 59 in [21]), one has

detImm(z)=4Imm+(z)Imm(z)|m+(z)+m(z)|2,z+.\det\operatorname{Im}m(z)=4\frac{\operatorname{Im}m_{+}(z)\operatorname{Im}m_{-}(z)}{|m_{+}(z)+m_{-}(z)|^{2}},\qquad z\in\mathbb{C}_{+}. (3.12)

Substituting expressions for

m±(z)=limξ+iP^±(ξ,z)P±(ξ,z)=i𝔞±(z)𝔟±(z)𝔞±(z)+𝔟±(z),z+,m_{\pm}(z)=\lim_{\xi\to+\infty}i\frac{\widehat{P}_{*}^{\pm}(\xi,z)}{P^{\pm}(\xi,z)}=i\frac{\mathfrak{a}^{\pm}(z)-\mathfrak{b}^{\pm}(z)}{\mathfrak{a}^{\pm}(z)+\mathfrak{b}^{\pm}(z)},\qquad z\in\mathbb{C}_{+},

into (3.12), we obtain

detρac(λ)\displaystyle\det\rho_{ac}(\lambda) =limε+0detImm(λ+iε)\displaystyle=\lim_{\varepsilon\to+0}\det\operatorname{Im}m(\lambda+i\varepsilon)
=limε+0(|𝔞+(λ+iε)|2|𝔟+(λ+iε)|2)(|𝔞(λ+iε)|2|𝔟(λ+iε)|2)|𝔞+(λ+iε)𝔞(λ+iε)𝔟+(λ+iε)𝔟(λ+iε)|2\displaystyle=\lim_{\varepsilon\to+0}\frac{(|\mathfrak{a}^{+}(\lambda+i\varepsilon)|^{2}-|\mathfrak{b}^{+}(\lambda+i\varepsilon)|^{2})(|\mathfrak{a}^{-}(\lambda+i\varepsilon)|^{2}-|\mathfrak{b}^{-}(\lambda+i\varepsilon)|^{2})}{|\mathfrak{a}^{+}(\lambda+i\varepsilon)\mathfrak{a}^{-}(\lambda+i\varepsilon)-\mathfrak{b}^{+}(\lambda+i\varepsilon)\mathfrak{b}^{-}(\lambda+i\varepsilon)|^{2}}
=1|a(λ)|2,\displaystyle=\frac{1}{|a(\lambda)|^{2}},

for almost every λ\lambda\in\mathbb{R} and the first claim of the lemma follows. Then, the second claim is immediate because aa is an outer function as we showed in Proposition 2.1.∎

Consider again the half-line entropy functions

𝒦Q±(z)=logImm±(z)1πlogImm±(λ)Imz|λz|2𝑑λ,z+.\mathcal{K}_{Q}^{\pm}(z)=\log\operatorname{Im}m_{\pm}(z)-\frac{1}{\pi}\int_{\mathbb{R}}\log\operatorname{Im}m_{\pm}(\lambda)\frac{\operatorname{Im}z}{|\lambda-z|^{2}}\,d\lambda,\qquad z\in\mathbb{C}_{+}.

We see that 𝒦Q+(i)\mathcal{K}_{Q}^{+}(i) coincides with the entropy (3.4) for the restriction of QQ to +\mathbb{R}_{+} (that explains why we use the same notation for the two objects), and 𝒦Q(z)=𝒦Q+(z)\mathcal{K}^{-}_{Q}(z)=\mathcal{K}^{+}_{Q_{-}}(z) for the potential

Q(ξ)=(Imq(ξ)Req(ξ)Req(ξ)Imq(ξ)),ξ+.Q_{-}(\xi)=\begin{pmatrix}-\operatorname{Im}q(-\xi)&\operatorname{Re}q(-\xi)\\ \operatorname{Re}q(-\xi)&\operatorname{Im}q(-\xi)\end{pmatrix},\qquad\xi\in\mathbb{R}_{+}.

Our plan now is to relate 𝒦Q±(i)\mathcal{K}^{\pm}_{Q}(i) with 𝒦Q(i)\mathcal{K}_{Q}(i) and then use the fact that the full line entropy 𝒦Q(i)\mathcal{K}_{Q}(i) is conserved, see Lemma 3.1. That will eventually lead to the proof of Theorem 1.2.

Lemma 3.2.

Let qL2()q\in L^{2}(\mathbb{R}) and let q(ξ)=q(ξ)q_{\ell}(\xi)=q(\xi-\ell), where \ell\in\mathbb{R} and ξ\xi\in\mathbb{R}. Denote by QQ_{\ell} the matrix-function in (3.1) corresponding to qq_{\ell}. Then, 𝒦Q+(z)𝒦Q(z)\mathcal{K}^{+}_{Q_{\ell}}(z)\to\mathcal{K}_{Q}(z), 𝒦Q(z)0\mathcal{K}^{-}_{Q_{\ell}}(z)\to 0 as +\ell\to+\infty for every z+z\in\mathbb{C}_{+}.

Proof.  Take z+z\in\mathbb{C}_{+}. We have

𝒦Q+(z)+𝒦Q(z)=log(Imm,+(z)Imm,(z))1πlog(Imm,+(λ)Imm,(λ))Imz|λz|2𝑑λ,\mathcal{K}^{+}_{Q_{\ell}}(z)+\mathcal{K}^{-}_{Q_{\ell}}(z)=\log\Bigl{(}\operatorname{Im}m_{\ell,+}(z)\operatorname{Im}m_{\ell,-}(z)\Bigr{)}-\frac{1}{\pi}\int_{\mathbb{R}}\log\Bigl{(}\operatorname{Im}m_{\ell,+}(\lambda)\operatorname{Im}m_{\ell,-}(\lambda)\Bigr{)}\frac{\operatorname{Im}z}{|\lambda-z|^{2}}\,d\lambda,

for the corresponding Weyl functions m,±m_{\ell,\pm}. We also have

log|m,+(z)+m,(z)|2=1πlog|m,+(λ)+m,(λ)|2Imz|λz|2dλ\log|m_{\ell,+}(z)+m_{\ell,-}(z)|^{2}=\frac{1}{\pi}\int_{\mathbb{R}}\log|m_{\ell,+}(\lambda)+m_{\ell,-}(\lambda)|^{2}\frac{\operatorname{Im}z}{|\lambda-z|^{2}}\,d\lambda

by the mean value theorem for harmonic functions. From (3.12), it follows that

𝒦Q+(z)+𝒦Q(z)=log(4Imm,+(z)Imm,(z)|m,+(z)+m,(z)|2)+𝒦Q(z).\mathcal{K}^{+}_{Q_{\ell}}(z)+\mathcal{K}^{-}_{Q_{\ell}}(z)=\log\left(4\frac{\operatorname{Im}m_{\ell,+}(z)\operatorname{Im}m_{\ell,-}(z)}{|m_{\ell,+}(z)+m_{\ell,-}(z)|^{2}}\right)+\mathcal{K}_{Q_{\ell}}(z).

Notice that 𝒦Q(z)\mathcal{K}_{Q_{\ell}}(z) does not depend on \ell\in\mathbb{R} because the coefficient aa in Lemma 3.1 for the potential QQ_{\ell} does not depend on \ell. So, we only need to show that

𝒦Q(z)0andlog(4Imm,+(z)Imm,(z)|m,+(z)+m,(z)|2)0,\mathcal{K}^{-}_{Q_{\ell}}(z)\to 0\quad\,\,{\rm and}\,\,\quad\log\left(4\frac{\operatorname{Im}m_{\ell,+}(z)\operatorname{Im}m_{\ell,-}(z)}{|m_{\ell,+}(z)+m_{\ell,-}(z)|^{2}}\right)\to 0,

when +\ell\to+\infty and z+z\in\mathbb{C}_{+}. The second relation follows from m,+(z)im_{\ell,+}(z)\to i, m,(z)im_{\ell,-}(z)\to i, which hold because qq_{\ell} tends to zero weakly in L2()L^{2}(\mathbb{R}) as +\ell\to+\infty and qL2()=qL2()\|q_{\ell}\|_{L^{2}(\mathbb{R})}=\|q\|_{L^{2}(\mathbb{R})} (see Lemma 6.2 in Appendix). Moreover, relation m,(z)im_{\ell,-}(z)\to i implies that 𝒦Q(z)0\mathcal{K}^{-}_{Q_{\ell}}(z)\to 0 if and only if

1πlogImm,(λ)Imz|λz|2𝑑λ0.\frac{1}{\pi}\int_{\mathbb{R}}\log\operatorname{Im}m_{\ell,-}(\lambda)\frac{\operatorname{Im}z}{|\lambda-z|^{2}}\,d\lambda\to 0\,. (3.13)

In the rest of the proof, we will show (3.13). Let 𝔞\mathfrak{a}_{\ell}^{-}, 𝔟\mathfrak{b}_{\ell}^{-} be the limits of continuous Wall polynomials corresponding to QQ_{\ell}^{-}. Consider s=𝔟/𝔞s_{\ell}^{-}=\mathfrak{b}_{\ell}^{-}/\mathfrak{a}_{\ell}^{-}. The formula (12.57) in [12] gives

s(z)=1+im,(z)1im,(z),m,(z)=i𝔞(z)𝔟(z)𝔞(z)+𝔟(z).s_{\ell}^{-}(z)=\frac{1+im_{\ell,-}(z)}{1-im_{\ell,-}(z)}\,,\qquad{m_{\ell,-}(z)=i\frac{\mathfrak{a}_{\ell}^{-}(z)-\mathfrak{b}^{-}_{\ell}(z)}{\mathfrak{a}^{-}_{\ell}(z)+\mathfrak{b}_{\ell}^{-}(z)}}.

It implies that Imm,(λ)=|𝔞(λ)+𝔟(λ)|2\operatorname{Im}m_{\ell,-}(\lambda)=|\mathfrak{a}^{-}_{\ell}(\lambda)+\mathfrak{b}^{-}_{\ell}(\lambda)|^{-2} when λ\lambda\in\mathbb{R} and that s(z)0s^{-}_{\ell}(z)\to 0 when +\ell\to+\infty and z+z\in\mathbb{C}_{+}. Now, we can write

1πlogImm,(λ)Imz|λz|2𝑑λ\displaystyle\frac{1}{\pi}\int_{\mathbb{R}}\log\operatorname{Im}m_{\ell,-}(\lambda)\frac{\operatorname{Im}z}{|\lambda-z|^{2}}\,d\lambda =1πlog(1|𝔞(λ)+𝔟(λ)|2)Imz|λz|2𝑑λ\displaystyle=\frac{1}{\pi}\int_{\mathbb{R}}\log\left(\frac{1}{|\mathfrak{a}^{-}_{\ell}(\lambda)+\mathfrak{b}^{-}_{\ell}(\lambda)|^{2}}\right)\frac{\operatorname{Im}z}{|\lambda-z|^{2}}\,d\lambda
=log1|𝔞(z)+𝔟(z)|2=log1|𝔞(z)|2+log1|1+s(z)|2.\displaystyle=\log\frac{1}{|\mathfrak{a}^{-}_{\ell}(z)+\mathfrak{b}^{-}_{\ell}(z)|^{2}}=\log\frac{1}{|\mathfrak{a}_{\ell}^{-}(z)|^{2}}+\log\frac{1}{|1+s_{\ell}^{-}(z)|^{2}}\,.

So, it remains to show that |𝔞(z)|21|\mathfrak{a}_{\ell}^{-}(z)|^{2}\to 1 as +\ell\to+\infty. That holds because q,L2(+)0\|q_{\ell,-}\|_{L^{2}(\mathbb{R}_{+})}\to 0 as +\ell\to+\infty and

q,L2(+)2=1πlog|𝔞(λ)|2dλImzπlog|𝔞(λ)|2Imz|λz|2dλ=Imzlog|𝔞(z)|20,\|q_{\ell,-}\|_{L^{2}(\mathbb{R}_{+})}^{2}=\frac{1}{\pi}\int_{\mathbb{R}}\log|\mathfrak{a}_{\ell}^{-}(\lambda)|^{2}\,d\lambda\geqslant\frac{\operatorname{Im}z}{\pi}\int_{\mathbb{R}}\log|\mathfrak{a}_{\ell}^{-}(\lambda)|^{2}\frac{\operatorname{Im}z}{|\lambda-z|^{2}}\,d\lambda=\operatorname{Im}z\cdot\log|\mathfrak{a}_{\ell}^{-}(z)|^{2}\geqslant 0,

where the first equality follows from q,L2(+)2=2A,L2(+)2\|q_{\ell,-}\|_{L^{2}(\mathbb{R}_{+})}^{2}=2\|A_{\ell,-}\|_{L^{2}(\mathbb{R}_{+})}^{2} and the formula (12.2)(12.2) in [12]. Thus, (3.13) holds and we are done. ∎

As an immediate corollary of Theorem 3.1 and Lemma 3.2, we get the following estimate.

Lemma 3.3.

Let qL2()q\in L^{2}(\mathbb{R}). Denote by NQN_{Q} the solution of the Cauchy problem JNQ(ξ)+Q(ξ)NQ(ξ)=0JN_{Q}^{\prime}(\xi)+Q(\xi)N_{Q}(\xi)=0, NQ(0)=(1001)N_{Q}(0)=\left(\begin{smallmatrix}1&0\\ 0&1\\ \end{smallmatrix}\right), and set Q=NQNQ\mathcal{H}_{Q}=N_{Q}^{*}N_{Q}. Consider

𝒦Q=𝒦Q(i),𝒦~Q=k(detkk+2Q(ξ)𝑑ξ4).\mathcal{K}_{Q}=\mathcal{K}_{Q}(i),\qquad\widetilde{\mathcal{K}}_{Q}=\sum_{k\in\mathbb{Z}}\left(\det\int_{k}^{k+2}\mathcal{H}_{Q}(\xi)\,d\xi-4\right). (3.14)

Then, we have

c1𝒦Q𝒦~Qc2𝒦Qec2𝒦Qc_{1}\mathcal{K}_{Q}\leqslant\widetilde{\mathcal{K}}_{Q}\leqslant c_{2}\mathcal{K}_{Q}e^{c_{2}\mathcal{K}_{Q}} (3.15)

for some positive absolute constants c1c_{1}, c2c_{2}.

Proof.  By Lemma 3.2, we have 𝒦Q=lim+𝒦Q+\mathcal{K}_{Q}=\lim_{\ell\to+\infty}\mathcal{K}^{+}_{Q_{\ell}}. It remains to substitute QQ_{\ell} into the estimate (3.6) and take the limit as +\ell\to+\infty for \ell\in\mathbb{Z}. ∎

4. Proof of Theorem 1.2

The following result will play a crucial role in what follows. We postpone its proof to the next section.

Theorem 4.1.

Suppose qL2()q\in L^{2}(\mathbb{R}) and let NQN_{Q} satisfy JNQ+QNQ=0,NQ(0)=IJN_{Q}^{\prime}+QN_{Q}=0,N_{Q}(0)=I, where Q=(ImqReqReqImq).Q=\left(\begin{smallmatrix}-\operatorname{Im}q&-\operatorname{Re}q\\ -\operatorname{Re}q&\operatorname{Im}q\end{smallmatrix}\right). Then,

eC1RqH1()2𝒦~QeC2RqH1()2,\displaystyle e^{-C_{1}R}\|q\|_{H^{-1}(\mathbb{R})}^{2}\lesssim\widetilde{\mathcal{K}}_{Q}\lesssim e^{C_{2}R}\|q\|^{2}_{H^{-1}(\mathbb{R})}\,, (4.1)

where R=qL2()R=\|q\|_{L^{2}(\mathbb{R})} and C1C_{1}, C2C_{2} are two positive absolute constants.

Proof of Theorem 1.2 in the case s=1s=-1. First, assume that q0𝒮()q_{0}\in\mathcal{S}(\mathbb{R}) and let q(ξ,t)q(\xi,t) be the solution of (1.1) with the initial datum q0q_{0}. We want to prove that

C1(1+q0L2())2q0H1()q(,t)H1()C2(1+q0L2())2q0H1().C_{1}(1+\|q_{0}\|_{L^{2}(\mathbb{R})})^{-2}\|q_{0}\|_{H^{-1}(\mathbb{R})}\leqslant\|q(\cdot,t)\|_{H^{-1}(\mathbb{R})}\leqslant C_{2}(1+\|q_{0}\|_{L^{2}(\mathbb{R})})^{2}\|q_{0}\|_{H^{-1}(\mathbb{R})}\,. (4.2)

We have q(,t)L2()=q0L2()\|q(\cdot,t)\|_{L^{2}(\mathbb{R})}=\|q_{0}\|_{L^{2}(\mathbb{R})} for all tt, see formula (4.33)(4.33) in [13]. Let a(z,t)a(z,t) denote the coefficient in the matrix (2.5) given by q(ξ,t)q(\xi,t). For each tt\in\mathbb{R}, define QQ by (3.2). Let 𝒦~Q(t)\widetilde{\mathcal{K}}_{Q}(t) be as in Lemma 3.3 and 𝒦Q(z,t)\mathcal{K}_{Q}(z,t) be defined by (3.11). Formulas (2.8) and (2.9) show that a(z,t)a({z},t) is constant in tt and Lemma 3.1 says that 𝒦Q(z,t)\mathcal{K}_{Q}(z,t) is constant in tt as well. The bound (3.15) yields

c1𝒦Q(i,0)𝒦~Q(t)c2𝒦Q(i,0)ec2𝒦Q(i,0).c_{1}\mathcal{K}_{Q}(i,0)\leqslant\widetilde{\mathcal{K}}_{Q}(t)\leqslant c_{2}\mathcal{K}_{Q}(i,0)e^{c_{2}\mathcal{K}_{Q}(i,0)}\,. (4.3)

Assume first that R=q0L2()1R=\|q_{0}\|_{L^{2}(\mathbb{R})}\leqslant 1. Taking t=0t=0 in (4.3) and applying (4.1) to q0q_{0}, we get 𝒦Q(i,0)1\mathcal{K}_{Q}(i,0)\lesssim 1 since q0H1()R1\|q_{0}\|_{H^{-1}(\mathbb{R})}\leqslant R\leqslant 1. Hence, in that case (4.3) can be written as 𝒦~Q(t)𝒦Q(i,0).\widetilde{\mathcal{K}}_{Q}(t)\sim\mathcal{K}_{Q}(i,0)\,. By (4.1), q(,t)H1()2𝒦~Q(t)\|q(\cdot,t)\|_{H^{-1}(\mathbb{R})}^{2}\sim\widetilde{\mathcal{K}}_{Q}(t), and so q(,t)H1()2q(,0)H1()2\|q(\cdot,t)\|_{H^{-1}(\mathbb{R})}^{2}\sim\|q(\cdot,0)\|_{H^{-1}(\mathbb{R})}^{2}.

If R=q0L2()>1R=\|q_{0}\|_{L^{2}(\mathbb{R})}>1, we use dilation. Consider qα(ξ,t)=αq(αξ,α2t)q_{\alpha}(\xi,t)=\alpha q(\alpha\xi,\alpha^{2}t) which solves the same equation and notice that qαL2()=α12R.\|q_{\alpha}\|_{L^{2}(\mathbb{R})}=\alpha^{{\frac{1}{2}}}R\,.

Let α=αc=R2<1\alpha=\alpha_{c}=R^{-2}<1 making qαcL2()=1\|q_{\alpha_{c}}\|_{L^{2}(\mathbb{R})}=1. Then, for the Sobolev norm, we get

qα(,t)H1()=α12(11+α2η2|(q)(η,α2t)|2𝑑η)12.\|q_{\alpha}(\cdot,t)\|_{H^{-1}(\mathbb{R})}=\alpha^{{\frac{1}{2}}}\left(\int_{\mathbb{R}}\frac{1}{1+\alpha^{2}\eta^{2}}|(\mathcal{F}q)(\eta,{\alpha^{2}}t)|^{2}d\eta\right)^{{\frac{1}{2}}}\,. (4.4)

Since

11+η211+αc2η21αc2(1+η2),\frac{1}{1+\eta^{2}}\leqslant\frac{1}{1+\alpha_{c}^{2}\eta^{2}}\leqslant\frac{1}{\alpha_{c}^{2}(1+\eta^{2})}, (4.5)

one has

αc12q(,αc2t)H1()qαc(,t)H1()αc12q(,αc2t)H1().\alpha_{c}^{{\frac{1}{2}}}\|{q(\cdot,\alpha_{c}^{2}t)}\|_{H^{-1}(\mathbb{R})}{\leqslant}\|q_{\alpha_{c}}(\cdot,{t})\|_{H^{-1}(\mathbb{R})}{\leqslant}\alpha_{c}^{-{\frac{1}{2}}}\|{q(\cdot,\alpha_{c}^{2}t)}\|_{H^{-1}(\mathbb{R})}\,.

In particular, at t=0t=0 we get

αc12q(,0)H1()qαc(,0)H1()αc12q(,0)H1().\alpha_{c}^{{\frac{1}{2}}}\|q(\cdot,0)\|_{H^{-1}(\mathbb{R})}\leqslant\|q_{\alpha_{c}}(\cdot,0)\|_{H^{-1}(\mathbb{R})}\leqslant\alpha_{c}^{-{\frac{1}{2}}}\|q(\cdot,0)\|_{H^{-1}(\mathbb{R})}\,.

Since qαc(,0)L2()=1\|q_{\alpha_{c}}(\cdot,0)\|_{L^{2}(\mathbb{R})}=1, one can apply the previous bounds to obtain

qαc(,t)H1()qαc(,0)H1().\|q_{\alpha_{c}}(\cdot,t)\|_{H^{-1}(\mathbb{R})}\sim\|q_{\alpha_{c}}(\cdot,0)\|_{H^{-1}(\mathbb{R})}\,.

Then,

αc12q(,αc2t)H1()\displaystyle\alpha_{c}^{{\frac{1}{2}}}\|q(\cdot,\alpha_{c}^{2}t)\|_{H^{-1}(\mathbb{R})} qαc(,0)H1()αc12q(,0)H1(),\displaystyle\lesssim\|q_{\alpha_{c}}(\cdot,0)\|_{H^{-1}(\mathbb{R})}\lesssim\alpha_{c}^{-{\frac{1}{2}}}\|q(\cdot,0)\|_{H^{-1}(\mathbb{R})},
αc12q(,αc2t)H1()\displaystyle\alpha_{c}^{-{\frac{1}{2}}}\|q(\cdot,\alpha_{c}^{2}t)\|_{H^{-1}(\mathbb{R})} qαc(,0)H1()αc12q(,0)H1().\displaystyle\gtrsim\|q_{\alpha_{c}}(\cdot,0)\|_{H^{-1}(\mathbb{R})}\gtrsim\alpha_{c}^{{\frac{1}{2}}}\|q(\cdot,0)\|_{H^{-1}(\mathbb{R})}.

Recalling that αc=R2\alpha_{c}=R^{-2}, we obtain

R2q(,0)H1()q(,t)H1()R2q(,0)H1()R^{-2}\|q(\cdot,0)\|_{H^{-1}(\mathbb{R})}\lesssim\|q(\cdot,t)\|_{H^{-1}(\mathbb{R})}\lesssim R^{2}\|q(\cdot,0)\|_{H^{-1}(\mathbb{R})}\,

for all tt\in\mathbb{R}. Finally, having proved (4.2) for q0𝒮()q_{0}\in\mathcal{S}(\mathbb{R}), it is enough to use Theorem 1.1 to extend (4.2) to q0L2()q_{0}\in L^{2}(\mathbb{R}).∎

Our next goal is to prove the estimate

C1(1+q0L2())2sq0Hs()q(,t)Hs()C2(1+q0L2())2sq0Hs(),C_{1}(1+\|q_{0}\|_{L^{2}(\mathbb{R})})^{2s}\|q_{0}\|_{H^{s}(\mathbb{R})}\leqslant\|q(\cdot,t)\|_{H^{s}(\mathbb{R})}\leqslant C_{2}(1+\|q_{0}\|_{L^{2}(\mathbb{R})})^{-2s}\|q_{0}\|_{H^{s}(\mathbb{R})}, (4.6)

where t,s(1,0]\,t\in\mathbb{R}\,,\,s\in(-1,0], and C1C_{1} and C2C_{2} are positive absolute constants. For s=0s=0, this bound is trivial. To cover s(1,0)s\in(-1,0), we will need some auxiliary results first. One of the basic properties of NLS which we discussed in the Introduction has to do with modulation: if q(ξ,t)q(\xi,t) solves (2.10), then q~v(ξ,t)=eivξiv2tq(ξ2vt,t)\widetilde{q}_{v}(\xi,t)=e^{iv\xi-iv^{2}t}q(\xi-2vt,t) solves (2.10) for every vv\in\mathbb{R}.

Lemma 4.1.

Let q0L2()q_{0}\in L^{2}(\mathbb{R}), tt\in\mathbb{R}. Then,

q~v(,t)H1()2=|(q)(η,t)|21+(η+v)2𝑑η.\|\tilde{q}_{v}(\cdot,t)\|_{H^{-1}(\mathbb{R})}^{2}=\int_{\mathbb{R}}\frac{|(\mathcal{F}q)(\eta,t)|^{2}}{1+(\eta+v)^{2}}\,d\eta.

Proof.  It is clear that eiv2tfH1()=fH1()\|e^{-iv^{2}t}f\|_{H^{-1}(\mathbb{R})}=\|f\|_{H^{-1}(\mathbb{R})} for every fH1()f\in H^{-1}(\mathbb{R}) and tt\in\mathbb{R}, because eiv2te^{-iv^{2}t} is a unimodular constant. We have (eivξq(ξ2vt,t))(η)=(q(ξ,t))(ηv)e2ivt(ηv)\mathcal{F}(e^{iv\xi}q(\xi-2vt,t))(\eta)=(\mathcal{F}q(\xi,t))(\eta-v)e^{-2ivt(\eta-v)}, η\eta\in\mathbb{R}. Since |e2ivt(ηv)|=1|e^{-2ivt(\eta-v)}|=1, it only remains to change the variable of integration in

q~vH1()=|(q(ξ,t))(ηv)|21+η2𝑑η\|\widetilde{q}_{v}\|_{H^{-1}(\mathbb{R})}=\int_{\mathbb{R}}\frac{|(\mathcal{F}q(\xi,t))(\eta-v)|^{2}}{1+\eta^{2}}\,d\eta

to get the statement of the lemma. ∎

The next result is a standard property of convolutions.

Lemma 4.2.

Let γ(12,1]\gamma\in(-\frac{1}{2},1] and set ak=1(1+k2)γa_{k}=\frac{1}{(1+k^{2})^{\gamma}} for kk\in\mathbb{Z}. We have

kak1+(ηk)2Cγ1(1+η2)γ,η.\sum_{k\in\mathbb{Z}}\frac{a_{k}}{1+(\eta-k)^{2}}\sim C_{\gamma}\frac{1}{(1+\eta^{2})^{\gamma}},\qquad\eta\in\mathbb{R}.

Proof.  After comparing the sum to an integral, it is enough to show that

du(1+u2)γ(1+(ηu)2)Cγ1(1+η2)γ.\int_{\mathbb{R}}\frac{du}{(1+u^{2})^{\gamma}(1+(\eta-u)^{2})}\sim C_{\gamma}\frac{1}{(1+\eta^{2})^{\gamma}}\,.

The function on the left-hand side is even and continuous in η\eta and γ\gamma, so we can assume that η>1\eta>1. Then,

|ηu|<0.5ηdu(1+u2)γ(1+(ηu)2)1(1+η2)γ,|ηu|>0.5ηdu(1+u2)γ(1+(ηu)2)1+2,\int_{|\eta-u|<0.5\eta}\frac{du}{(1+u^{2})^{\gamma}(1+(\eta-u)^{2})}\sim\frac{1}{(1+\eta^{2})^{\gamma}},\quad\int_{|\eta-u|>0.5\eta}\frac{du}{(1+u^{2})^{\gamma}(1+(\eta-u)^{2})}\lesssim\mathcal{I}_{1}+\mathcal{I}_{2}\,,

where

1=u<η/2,u>3η/2du(1+u2)γ(1+(ηu)2)η/2duu2+2γ+3η/2duu2+2γCγη12γ,\mathcal{I}_{1}=\int_{u<-\eta/2,\;{u>3\eta/2}}\frac{du}{(1+u^{2})^{\gamma}(1+(\eta-u)^{2})}\lesssim\int_{-\infty}^{-\eta/2}\frac{du}{u^{2+2\gamma}}+\int_{3\eta/2}\frac{du}{u^{2+2\gamma}}\leqslant C_{\gamma}\eta^{-1-2\gamma}\,,
2=|u|<η/2du(1+u2)γ(1+(ηu)2)η2|u|<η/2du(1+u2)γη2γ.\mathcal{I}_{2}=\int_{|u|<\eta/2}\frac{du}{(1+u^{2})^{\gamma}(1+(\eta-u)^{2})}\lesssim\eta^{-2}\int_{|u|<\eta/2}\frac{du}{(1+u^{2})^{\gamma}}\lesssim\eta^{-2\gamma}\,.

Combining these bounds proves the lemma. ∎

Proof of Theorem 1.2, the case s(1,0)s\in(-1,0). We can again assume that q0𝒮()q_{0}\in\mathcal{S}(\mathbb{R}). Recall the estimate (1.5) for s=1s=-1:

C1(1+q0L2())2q0H1()q(,t)H1()C2(1+q0L2())2q0H1().C_{1}(1+\|q_{0}\|_{L^{2}(\mathbb{R})})^{-2}\|q_{0}\|_{H^{-1}(\mathbb{R})}\leqslant\|q(\cdot,t)\|_{H^{-1}(\mathbb{R})}\leqslant C_{2}(1+\|q_{0}\|_{L^{2}(\mathbb{R})})^{2}\|q_{0}\|_{H^{-1}(\mathbb{R})}\,. (4.7)

According to Lemma 4.1, we have

q~v(,t)H1()2=|(q(,t))(η)|21+(v+η)2𝑑η\|\tilde{q}_{v}(\cdot,t)\|_{H^{-1}(\mathbb{R})}^{2}=\int_{\mathbb{R}}\frac{|(\mathcal{F}q(\cdot,t))(\eta)|^{2}}{1+(v+\eta)^{2}}\,d\eta (4.8)

for q~v(ξ,t)=eivξiv2tq(ξ2vt,t)\tilde{q}_{v}(\xi,t)=e^{iv\xi-iv^{2}t}q(\xi-2vt,t). Let aka_{k}, kk\in\mathbb{Z}, be the coefficients from Lemma 4.2 with γ=s\gamma=-s. Then, (4.8) and Lemma 4.2 imply

kakq~k(,t)H1()2Csq(,t)Hs()2.\sum_{k\in\mathbb{Z}}a_{k}\|\tilde{q}_{k}(\cdot,t)\|_{H^{-1}(\mathbb{R})}^{2}\sim C_{s}\|q(\cdot,t)\|_{H^{s}(\mathbb{R})}^{2}\,. (4.9)

In particular, taking t=0t=0 gives

kakq~k(,0)H1()2Csq0Hs()2.\sum_{k\in\mathbb{Z}}a_{k}\|\tilde{q}_{k}(\cdot,0)\|_{H^{-1}(\mathbb{R})}^{2}\sim C_{s}\|q_{0}\|_{H^{s}(\mathbb{R})}^{2}\,. (4.10)

We now apply (4.7) to q~k\tilde{q}_{k} and use (4.9) and (4.10) to get

C1(s)(1+q0L2())2q0Hs()q(,t)Hs()C2(s)(1+q0L2())2q0Hs().C_{1}(s)(1+\|q_{0}\|_{L^{2}(\mathbb{R})})^{-2}\|q_{0}\|_{H^{s}(\mathbb{R})}\leqslant\|q(\cdot,t)\|_{H^{s}(\mathbb{R})}\leqslant C_{2}(s)(1+\|q_{0}\|_{L^{2}(\mathbb{R})})^{2}\|q_{0}\|_{H^{s}(\mathbb{R})}\,. (4.11)

If R=q0L2()1R=\|q_{0}\|_{L^{2}(\mathbb{R})}\leqslant 1, we have the statement of our theorem. If R=q0L2()>1R=\|q_{0}\|_{L^{2}(\mathbb{R})}>1, we use dilation transformation like in the previous proof for s=1s=-1. Consider qα(ξ,t)=αq(αξ,α2t)q_{\alpha}(\xi,t)=\alpha q(\alpha\xi,\alpha^{2}t) which solves the same equation and notice that qαL2()=α12R.\|q_{\alpha}\|_{L^{2}(\mathbb{R})}=\alpha^{\frac{1}{2}}R\,. Let α=αc=R2<1\alpha=\alpha_{c}=R^{-2}<1 making qαcL2()=1\|q_{\alpha_{c}}\|_{L^{2}(\mathbb{R})}=1. Then, for the Sobolev norm, we have

qα(,t)Hs()=α12(1(1+α2η2)|s||(q)(η,α2t)|2𝑑η)12.\|q_{\alpha}(\cdot,t)\|_{H^{s}(\mathbb{R})}=\alpha^{\frac{1}{2}}\left(\int_{\mathbb{R}}\frac{1}{(1+\alpha^{2}\eta^{2})^{|s|}}|(\mathcal{F}q)(\eta,{\alpha^{2}}t)|^{2}d\eta\right)^{\frac{1}{2}}\,.

From (4.5),

1(1+η2)|s|1(1+αc2η2)|s|1αc2|s|(1+η2)|s|.\frac{1}{(1+\eta^{2})^{|s|}}\leqslant\frac{1}{(1+\alpha_{c}^{2}\eta^{2})^{|s|}}\leqslant\frac{1}{\alpha_{c}^{2|s|}(1+\eta^{2})^{|s|}}\,.

Then, one has

αc12q(,αc2t)Hs()qαc(,t)Hs()αc12|s|q(,αc2t)Hs().\alpha_{c}^{\frac{1}{2}}\|{q(\cdot,\alpha_{c}^{2}t)}\|_{H^{s}(\mathbb{R})}{\leqslant}\|q_{\alpha_{c}}(\cdot,{t})\|_{H^{s}(\mathbb{R})}{\leqslant}\alpha_{c}^{\frac{1}{2}-|s|}\|{q(\cdot,\alpha_{c}^{2}t)}\|_{H^{s}(\mathbb{R})}\,.

In particular, taking t=0t=0 gives us

αc12q(,0)Hs()qαc(,0)Hs()αc12|s|q(,0)Hs().\alpha_{c}^{\frac{1}{2}}\|q(\cdot,0)\|_{H^{s}(\mathbb{R})}\leqslant\|q_{\alpha_{c}}(\cdot,0)\|_{H^{s}(\mathbb{R})}\leqslant\alpha_{c}^{\frac{1}{2}-|s|}\|q(\cdot,0)\|_{H^{s}(\mathbb{R})}\,.

Now qαc(,0)L2()=1\|q_{\alpha_{c}}(\cdot,0)\|_{L^{2}(\mathbb{R})}=1 and we can apply the previous bounds to get

qαc(,t)Hs()qαc(,0)Hs().\|q_{\alpha_{c}}(\cdot,t)\|_{H^{s}(\mathbb{R})}\sim\|q_{\alpha_{c}}(\cdot,0)\|_{H^{s}(\mathbb{R})}\,.

Then,

αc12q(,αc2t)Hs()\displaystyle\alpha_{c}^{\frac{1}{2}}\|q(\cdot,\alpha_{c}^{2}t)\|_{H^{s}(\mathbb{R})} qαc(,0)Hs()αc12|s|q(,0)Hs(),\displaystyle\lesssim\|q_{\alpha_{c}}(\cdot,0)\|_{H^{s}(\mathbb{R})}\lesssim\alpha_{c}^{\frac{1}{2}-|s|}\|q(\cdot,0)\|_{H^{s}(\mathbb{R})},
αc12|s|q(,αc2t)Hs()\displaystyle\alpha_{c}^{\frac{1}{2}-|s|}\|q(\cdot,\alpha_{c}^{2}t)\|_{H^{s}(\mathbb{R})} qαc(,0)Hs()αc12q(,0)Hs().\displaystyle\gtrsim\|q_{\alpha_{c}}(\cdot,0)\|_{H^{s}(\mathbb{R})}\gtrsim\alpha_{c}^{\frac{1}{2}}\|q(\cdot,0)\|_{H^{s}(\mathbb{R})}.

Recalling that αc=R2=q0L2()2\alpha_{c}=R^{-2}=\|q_{0}\|_{L^{2}(\mathbb{R})}^{-2}, we obtain

q0L2()2|s|q(,0)Hs()q(,t)Hs()q0L2()2|s|q(,0)Hs()\|q_{0}\|_{L^{2}(\mathbb{R})}^{-2|s|}\|q(\cdot,0)\|_{H^{s}(\mathbb{R})}\lesssim\|q(\cdot,t)\|_{H^{s}(\mathbb{R})}\lesssim\|q_{0}\|_{L^{2}(\mathbb{R})}^{2|s|}\|q(\cdot,0)\|_{H^{s}(\mathbb{R})}\,

for all tt\in\mathbb{R}. ∎

Our approach also provides the bounds for some positive Sobolev norms. The following proposition slightly improves (1.4) when s[0,12)s\in[0,\frac{1}{2}), q0Hs()\|q_{0}\|_{H^{s}(\mathbb{R})} is large, and q0L2()\|q_{0}\|_{L^{2}(\mathbb{R})} is much smaller than q0Hs()\|q_{0}\|_{H^{s}(\mathbb{R})}.

Proposition 4.1.

Let q0𝒮()q_{0}\in\mathcal{S}(\mathbb{R}) and let q=q(ξ,t)q=q(\xi,t) be the solution of (1.1) corresponding to q0q_{0}. Then, for each s[0,12)s\in[0,\frac{1}{2}), we get

q(,t)Hs()Csq0Hs()\|q(\cdot,t)\|_{H^{s}(\mathbb{R})}\sim C_{s}\|q_{0}\|_{H^{s}(\mathbb{R})} (4.12)

if q0L2()1\|q_{0}\|_{L^{2}(\mathbb{R})}\leqslant 1 and

q(,t)Hs()Cs(q0L2()1+2s+q0Hs())\|q(\cdot,t)\|_{H^{s}(\mathbb{R})}\lesssim C_{s}(\|q_{0}\|_{L^{2}(\mathbb{R})}^{1+2s}+\|q_{0}\|_{H^{s}(\mathbb{R})}) (4.13)

if q0L2()>1\|q_{0}\|_{L^{2}(\mathbb{R})}>1.

Proof.  In the case when qL2()1\|q\|_{L^{2}(\mathbb{R})}\leqslant 1, the proof of proposition repeats the arguments given above to get (4.11) except that the constants in the inequalities depend on ss and can blow up when s12s\to\frac{1}{2}. Suppose qL2()1\|q\|_{L^{2}(\mathbb{R})}\geqslant 1. Then, for the Sobolev norm, we have

qα(,t)Hs()=α12((1+α2η2)s|(q)(η,α2t)|2𝑑η)12.\|q_{\alpha}(\cdot,t)\|_{H^{s}(\mathbb{R})}=\alpha^{\frac{1}{2}}\left(\int_{\mathbb{R}}{(1+\alpha^{2}\eta^{2})^{s}}|(\mathcal{F}q)(\eta,{\alpha^{2}}t)|^{2}d\eta\right)^{\frac{1}{2}}\,.

Take α=αc\alpha=\alpha_{c} and write the following estimate for the integral above:

(1+η2R4)s|(q)(η,αc2t)|2𝑑ηR2R2|(q)(η,αc2t)|2𝑑η+R4s|η|>R2(1+η2)s|(q)(η,αc2t)|2𝑑η\displaystyle\int_{\mathbb{R}}{\left(1+\frac{\eta^{2}}{R^{4}}\right)^{s}}|(\mathcal{F}q)(\eta,{\alpha_{c}^{2}}t)|^{2}d\eta\sim\int_{-R^{2}}^{R^{2}}|(\mathcal{F}q)(\eta,{\alpha_{c}^{2}}t)|^{2}d\eta+R^{-4s}\int_{|\eta|>R^{2}}{\left(1+\eta^{2}\right)^{s}}|(\mathcal{F}q)(\eta,{\alpha_{c}^{2}}t)|^{2}d\eta
R2+R4s(1+η2)s|(q)(η,αc2t)|2𝑑η.\displaystyle\lesssim R^{2}+R^{-4s}\int_{\mathbb{R}}{\left(1+\eta^{2}\right)^{s}}|(\mathcal{F}q)(\eta,{\alpha_{c}^{2}}t)|^{2}d\eta\,.

We use qαc(,t)L2()=1\|q_{\alpha_{c}}(\cdot,t)\|_{L^{2}(\mathbb{R})}=1 and (4.11) to get qαc(,t)Hs()Csqαc(,0)Hs()\|q_{\alpha_{c}}(\cdot,t)\|_{H^{s}(\mathbb{R})}\sim C_{s}\|q_{\alpha_{c}}(\cdot,0)\|_{H^{s}(\mathbb{R})}. The previous estimate for t=0t=0 yields qαc(,0)Hs()1+R12sq(,0)Hs().\|q_{\alpha_{c}}(\cdot,0)\|_{H^{s}(\mathbb{R})}\lesssim 1+R^{-1-2s}\|q(\cdot,0)\|_{H^{s}(\mathbb{R})}. Hence, qαc(,t)Hs()Cs(1+R12sq(,0)Hs()).\|q_{\alpha_{c}}(\cdot,t)\|_{H^{s}(\mathbb{R})}\leqslant C_{s}(1+R^{-1-2s}\|q(\cdot,0)\|_{H^{s}(\mathbb{R})}). We can write a lower bound

qαc(,t)Hs()2=R2(1+η2R4)s|(q)(η,αc2t)|2𝑑ηR24s|η|>R2(1+η2)s|(q)(η,αc2t)|2𝑑η\|q_{\alpha_{c}}(\cdot,t)\|_{H^{s}(\mathbb{R})}^{2}=R^{-2}\int_{\mathbb{R}}{\left(1+\frac{\eta^{2}}{R^{4}}\right)^{s}}|(\mathcal{F}q)(\eta,{\alpha_{c}^{2}}t)|^{2}d\eta\gtrsim R^{-2-4s}\int_{|\eta|>R^{2}}{\left(1+\eta^{2}\right)^{s}}|(\mathcal{F}q)(\eta,{\alpha_{c}^{2}}t)|^{2}d\eta

so

|η|>R2(1+η2)s|(q)(η,αc2t)|2𝑑ηCs(R2+4s+q(,0)Hs()2).\int_{|\eta|>R^{2}}{\left(1+\eta^{2}\right)^{s}}|(\mathcal{F}q)(\eta,{\alpha_{c}^{2}}t)|^{2}d\eta\leqslant C_{s}(R^{2+4s}+\|q(\cdot,0)\|^{2}_{H^{s}(\mathbb{R})}).

Writing the integral as a sum of two:

(1+η2)s|(q)(η,αc2t)|2𝑑η=|η|>R2(1+η2)s|(q)(η,αc2t)|2𝑑η+|η|<R2(1+η2)s|(q)(η,αc2t)|2𝑑η\int_{\mathbb{R}}{\left(1+\eta^{2}\right)^{s}}|(\mathcal{F}q)(\eta,{\alpha_{c}^{2}}t)|^{2}d\eta=\int_{|\eta|>R^{2}}{\left(1+\eta^{2}\right)^{s}}|(\mathcal{F}q)(\eta,{\alpha_{c}^{2}}t)|^{2}d\eta+\int_{|\eta|<R^{2}}{\left(1+\eta^{2}\right)^{s}}|(\mathcal{F}q)(\eta,{\alpha_{c}^{2}}t)|^{2}d\eta\,

and estimating each of them, we get a bound which holds for all tt:

(1+η2)s|(q)(η,αc2t)|2𝑑ηCs(R2+4s+q(,0)Hs()2).\int_{\mathbb{R}}{\left(1+\eta^{2}\right)^{s}}|(\mathcal{F}q)(\eta,{\alpha_{c}^{2}}t)|^{2}d\eta\leqslant C_{s}(R^{2+4s}+\|q(\cdot,0)\|^{2}_{H^{s}(\mathbb{R})})\,.

That is the required upper bound (4.13). ∎


5. Oscillation and Sobolev space H1()H^{-1}(\mathbb{R}).

In this part of the paper, our goal is to prove the Theorem 4.1. Let us recall its statement.

Theorem 5.1.

Suppose that qL2()q\in L^{2}(\mathbb{R}) and let NQN_{Q} satisfy JNQ+QNQ=0,NQ(0)=IJN^{\prime}_{Q}+QN_{Q}=0,N_{Q}(0)=I, where Q=(ImqReqReqImq).Q=\left(\begin{smallmatrix}-\operatorname{Im}q&-\operatorname{Re}q\\ -\operatorname{Re}q&\operatorname{Im}q\end{smallmatrix}\right). Then,

eC1RqH1()2𝒦~QeC2RqH1()2,\displaystyle e^{-C_{1}R}\|q\|_{H^{-1}(\mathbb{R})}^{2}\lesssim\widetilde{\mathcal{K}}_{Q}\lesssim e^{C_{2}R}\|q\|^{2}_{H^{-1}(\mathbb{R})}\,, (5.1)

where R=qL2()R=\|q\|_{L^{2}(\mathbb{R})} and C1C_{1}, C2C_{2} are two positive absolute constants.

Theorem 5.1 is of independent interest in the spectral theory of Dirac operators. For example, Lemma 3.3 shows that qL2()\|q\|_{L^{2}(\mathbb{R})} and qH1()\|q\|_{H^{-1}(\mathbb{R})} control the size of 𝒦Q\mathcal{K}_{Q}.

The strategy of the proof is the following. In the next subsection, we show that H1()H^{-1}(\mathbb{R})-norm of any function can be characterized through BMO-like condition for its “antiderivative”. In Subsection 5.2, we consider solution to Cauchy problem JN+QN=0,N(0)=IJN^{\prime}+QN=0,N(0)=I on the interval [0,1][0,1] where zero-trace symmetric QQ and study the quantity det01NN𝑑x\det\int_{0}^{1}N^{*}Ndx, which represents a single term in the sum for 𝒦~Q\widetilde{\mathcal{K}}_{Q}. The results in Subsection 5.3 show that small value of 𝒦~Q\widetilde{\mathcal{K}}_{Q} guarantees that the “local” H1H^{-1} norm of QQ is also small. This rough estimate is used in the proof of Theorem 4.1 which is contained in Subsection 5.4.

5.1. One property of Sobolev space H1()H^{-1}(\mathbb{R})

Observe that a function fL2()f\in L^{2}(\mathbb{R}) belongs to the Sobolev space H1()H^{-1}(\mathbb{R}) if and only if

|f(y)χ+(xy)e(xy)𝑑y|2𝑑x<.\int_{\mathbb{R}}\left|\int_{\mathbb{R}}f(y)\chi_{\mathbb{R}_{+}}(x-y)e^{-(x-y)}\,dy\right|^{2}\,dx<\infty. (5.2)

Moreover, the last integral is equal to fH1()2.\|f\|_{H^{-1}(\mathbb{R})}^{2}. Indeed, recall that f\mathcal{F}f stands for the Fourier transform of ff:

(f)(η)=12πf(x)eiηx𝑑x.(\mathcal{F}f)(\eta)=\frac{1}{\sqrt{2\pi}}\int_{\mathbb{R}}f(x)e^{-i\eta x}\,dx\,.

Then, from Plancherel’s identity and formula

12π+exeixη𝑑x=12π11+iη,\frac{1}{\sqrt{2\pi}}\int_{\mathbb{R}_{+}}e^{-x}e^{-ix\eta}dx=\frac{1}{\sqrt{2\pi}}\frac{1}{1+i\eta},

we obtain

fH1()2=2π(f)(χ+ex)L2()2=|(f)(η)|21+η2𝑑η\|f\|_{H^{-1}(\mathbb{R})}^{2}=2\pi\|(\mathcal{F}f)\mathcal{F}(\chi_{\mathbb{R}_{+}}e^{-x})\|^{2}_{L^{2}(\mathbb{R})}=\int_{\mathbb{R}}\frac{|(\mathcal{F}f)(\eta)|^{2}}{1+\eta^{2}}\,d\eta

by properties of convolutions. We will need the following proposition.

Proposition 5.1.

Suppose that fLloc1()H1()f\in L^{1}_{\rm loc}(\mathbb{R})\cap H^{-1}(\mathbb{R}). Let gg be an absolutely continuous function on \mathbb{R} such that g=fg^{\prime}=f almost everywhere on \mathbb{R}. Then,

c1fH1()2kIk|ggIk|2𝑑xc2fH1()2,c_{1}\|f\|_{H^{-1}(\mathbb{R})}^{2}\leqslant\sum_{k\in\mathbb{Z}}\int_{I_{k}}|g-\langle g\rangle_{I_{k}}|^{2}\,dx\leqslant c_{2}\|f\|_{H^{-1}(\mathbb{R})}^{2}, (5.3)

where Ik=[k,k+2]I_{k}=[k,k+2], gI=1|I|Ig(x)𝑑x\langle g\rangle_{I}=\frac{1}{|I|}\int_{I}g(x)\,dx, and the positive constants c1c_{1} and c2c_{2} are universal.

Proof.  Take a function fLloc1()H1()f\in L^{1}_{\rm loc}(\mathbb{R})\cap H^{-1}(\mathbb{R}), and let gg be an absolutely continuous function on \mathbb{R} such that g=fg^{\prime}=f almost everywhere on \mathbb{R}. Assume first that ff has a compact support. The integral under the sum does not change if we add a constant to gg, so we can suppose without loss of generality that

g(x)=xf(s)𝑑s,x.g(x)=\int_{-\infty}^{x}f(s)\,ds,\qquad x\in\mathbb{R}.

Upper bound. Given ff, define ofo_{f} by

of(x)=exxf(y)ey𝑑yo_{f}(x)=e^{-x}\int_{-\infty}^{x}f(y)e^{y}dy

and recall (see (5.2)) that

fH1()=ofL2().\|f\|_{H^{-1}(\mathbb{R})}=\|o_{f}\|_{L^{2}(\mathbb{R})}. (5.4)

Moreover,

of+of=f.o^{\prime}_{f}+o_{f}=f\,. (5.5)

For each interval IkI_{k}, we use (5.5) for the corresponding term in the sum (5.3):

Ik|kxf𝑑x112kk+2(kx1f(x2)𝑑x2)𝑑x1|2𝑑x\displaystyle\int_{I_{k}}\left|\int_{k}^{x}fdx_{1}-\frac{1}{2}\int_{k}^{k+2}\left(\int_{k}^{x_{1}}f(x_{2})dx_{2}\right)dx_{1}\right|^{2}dx
=Ik|kxo(x1)𝑑x1+o(x)12kk+2(o(x1)+kx1o(x2)𝑑x2)𝑑x1|2𝑑x\displaystyle=\int_{I_{k}}\left|\int_{k}^{x}o(x_{1})dx_{1}+o(x)-\frac{1}{2}\int_{k}^{k+2}\left(o(x_{1})+\int_{k}^{x_{1}}o(x_{2})dx_{2}\right)dx_{1}\right|^{2}dx
Ik|o|2𝑑x\displaystyle\lesssim\int_{I_{k}}|o|^{2}dx

after the Cauchy-Schwarz inequality is applied. Summing these estimates in kk\in\mathbb{Z} and using (5.4), we get the upper bound in (5.3) for compactly supported ff.

Lower bound. Integration by parts gives

xf(y)e(xy)𝑑y\displaystyle\int_{-\infty}^{x}f(y)e^{-(x-y)}\,dy =xf(s)𝑑sx(yf(s)𝑑s)e(xy)𝑑y\displaystyle=\int_{-\infty}^{x}f(s)\,ds-\int_{-\infty}^{x}\left(\int_{-\infty}^{y}f(s)\,ds\right)\,e^{-(x-y)}\,dy
=g(x)xg(y)e(xy)𝑑y\displaystyle=g(x)-\int_{-\infty}^{x}g(y)\,e^{-(x-y)}\,dy
=x(g(x)g(y))e(xy)𝑑y.\displaystyle=\int_{-\infty}^{x}(g(x)-g(y))\,e^{-(x-y)}\,dy.

Therefore,

|xf(y)e(xy)𝑑y|2𝑑x\displaystyle\int_{\mathbb{R}}\left|\int_{-\infty}^{x}f(y)e^{-(x-y)}\,dy\right|^{2}\,dx kkk+2(x|g(x)g(y)|2e(xy)𝑑y)𝑑x\displaystyle\leqslant\sum_{k\in\mathbb{Z}}\int_{k}^{k+2}\left(\int_{-\infty}^{x}|g(x)-g(y)|^{2}e^{-(x-y)}\,dy\right)\,dx
kjke(kj)kk+2jj+2|g(x)g(y)|2𝑑x𝑑y.\displaystyle\lesssim\sum_{k\in\mathbb{Z}}\sum_{j\leqslant k}e^{-(k-j)}\int_{k}^{k+2}\int_{j}^{j+2}|g(x)-g(y)|^{2}\,dx\,dy\,.

Using the inequality (x+y+z)23(x2+y2+z2)(x+y+z)^{2}\leqslant 3(x^{2}+y^{2}+z^{2}), we continue the estimate:

kjke(kj)(Ik|ggIk|2𝑑x+|gIjgIk|2+Ij|ggIj|2𝑑x)....\lesssim\sum_{k\in\mathbb{Z}}\sum_{j\leqslant k}e^{-(k-j)}\left(\int_{I_{k}}|g-\langle g\rangle_{I_{k}}|^{2}dx+|\langle g\rangle_{I_{j}}-\langle g\rangle_{I_{k}}|^{2}+\int_{I_{j}}|g-\langle g\rangle_{I_{j}}|^{2}dx\right).

Since

kjke(kj)(Ik|ggIk|2𝑑x+Ij|ggIj|2𝑑x)kIk|ggIk|2𝑑x,\sum_{k\in\mathbb{Z}}\sum_{j\leqslant k}e^{-(k-j)}\left(\int_{I_{k}}|g-\langle g\rangle_{I_{k}}|^{2}dx+\int_{I_{j}}|g-\langle g\rangle_{I_{j}}|^{2}dx\right)\lesssim\sum_{k\in\mathbb{Z}}\int_{I_{k}}|g-\langle g\rangle_{I_{k}}|^{2}dx\,,

we are left with estimating

kjke(kj)|gIjgIk|2.\sum_{k\in\mathbb{Z}}\sum_{j\leqslant k}e^{-(k-j)}|\langle g\rangle_{I_{j}}-\langle g\rangle_{I_{k}}|^{2}.

Applying the Cauchy-Schwarz inequality for the telescoping sum

gIkgIj=s=j+1k(gIsgIs1),\langle g\rangle_{I_{k}}-\langle g\rangle_{I_{j}}=\sum_{s=j+1}^{k}\Bigl{(}\langle g\rangle_{I_{s}}-\langle g\rangle_{I_{s-1}}\Bigr{)}\,,

we get

|gIjgIk|2(kj)jsk1|gIsgIs+1|2.|\langle g\rangle_{I_{j}}-\langle g\rangle_{I_{k}}|^{2}\leqslant(k-j)\sum_{j\leqslant s\leqslant k-1}|\langle g\rangle_{I_{s}}-\langle g\rangle_{I_{s+1}}|^{2}.

Then,

kjke(kj)(kj)jsk1|gIsgIs+1|2=s|gIsgIs+1|2k,j:jsk1(kj)e(kj).\sum_{k\in\mathbb{Z}}\sum_{j\leqslant k}e^{-(k-j)}(k-j)\sum_{j\leqslant s\leqslant k-1}|\langle g\rangle_{I_{s}}-\langle g\rangle_{I_{s+1}}|^{2}=\sum_{s\in\mathbb{Z}}|\langle g\rangle_{I_{s}}-\langle g\rangle_{I_{s+1}}|^{2}\sum_{k,j:\;j\leqslant s\leqslant k-1}(k-j)e^{-(k-j)}.

We have

k,j:jsk1(kj)e(kj)\displaystyle\sum_{k,j:\;j\leqslant s\leqslant k-1}(k-j)e^{-(k-j)} =jsm1(s+mj)e(s+mj)\displaystyle=\sum_{j\leqslant s}\sum_{m\geqslant 1}(s+m-j)e^{-(s+m-j)}
=j0m1(mj)e(mj)=j0m1(m+j)e(m+j).\displaystyle=\sum_{j\leqslant 0}\sum_{m\geqslant 1}(m-j)e^{-(m-j)}=\sum_{j\geqslant 0}\sum_{m\geqslant 1}(m+j)e^{-(m+j)}.

The last sum is finite and does not depend on index ss. Now, the estimate

|gIsgIs+1|2=IsIs+1|gIsg+ggIs+1|2𝑑x2Is|ggIs|2𝑑x+2Is+1|ggIs+1|2𝑑x|\langle g\rangle_{I_{s}}-\langle g\rangle_{I_{s+1}}|^{2}=\int_{I_{s}\cap I_{s+1}}|\langle g\rangle_{I_{s}}-g+g-\langle g\rangle_{I_{s+1}}|^{2}dx\leqslant 2\int_{I_{s}}|g-\langle g\rangle_{I_{s}}|^{2}dx+2\int_{I_{s+1}}|g-\langle g\rangle_{I_{s+1}}|^{2}dx

proves that

kjke(kj)|gIjgIk|2sIs|ggIs|2𝑑x.\sum_{k\in\mathbb{Z}}\sum_{j\leqslant k}e^{-(k-j)}|\langle g\rangle_{I_{j}}-\langle g\rangle_{I_{k}}|^{2}\lesssim\sum_{s\in\mathbb{Z}}\int_{I_{s}}|g-\langle g\rangle_{I_{s}}|^{2}dx\,.

Hence, the lower bound in (5.3) holds for compactly supported ff.

Now, take any fLloc1()H1()f\in L^{1}_{\rm loc}(\mathbb{R})\cap H^{-1}(\mathbb{R}). The definition (1.2) of H1()H^{-1}(\mathbb{R}) implies that f\mathcal{F}f can be written as (1+iη)(o)(1+i\eta)(\mathcal{F}o) for some function oL2()o\in L^{2}(\mathbb{R}). Moreover, this map fof\mapsto o is a bijection between H1()H^{-1}(\mathbb{R}) and L2()L^{2}(\mathbb{R}) and fH1()=oL2()\|f\|_{H^{-1}(\mathbb{R})}=\|o\|_{L^{2}(\mathbb{R})}. Taking the inverse Fourier transform of identity f=(1+iη)(o)\mathcal{F}f=(1+i\eta)(\mathcal{F}o), one gets a formula f=o+of=o+o^{\prime} where oo^{\prime} is understood as a derivative in 𝒮()\mathcal{S}^{\prime}(\mathbb{R}). Since fLloc1()f\in L^{1}_{\rm loc}(\mathbb{R}) and oL2()o\in L^{2}(\mathbb{R}), we have oLloc1()o^{\prime}\in L^{1}_{\rm loc}(\mathbb{R}) and, therefore, oo is absolutely continuous on \mathbb{R} with the derivative equal to fof-o. Now, take on(x)=o(x)μn(x)o_{n}(x)=o(x)\mu_{n}(x) and define the corresponding fn=on+onf_{n}=o_{n}+o_{n}^{\prime}. Here, μn(x)\mu_{n}(x) is even and

μn(x)={1,0x<n,n+1x,x[n,n+1),0,xn+1.\mu_{n}(x)=\left\{\begin{array}[]{ll}1,&0\leqslant x<n,\\ n+1-x,&x\in[n,n+1),\\ 0,&x\geqslant n+1\,.\end{array}\right.

Then, {on}o\{o_{n}\}\to o in L2()L^{2}(\mathbb{R}) and so {fn}f\{f_{n}\}\to f in H1()H^{-1}(\mathbb{R}) because the mapping fof\mapsto o is unitary from H1()H^{-1}(\mathbb{R}) onto L2()L^{2}(\mathbb{R}). Also, each fnf_{n} is compactly supported and {fn}\{f_{n}\} converges to ff uniformly on every finite interval. Define gn=0xfn𝑑s,g=0xf𝑑sg_{n}=\int_{0}^{x}f_{n}ds,g=\int_{0}^{x}fds, and write (5.3) for fnf_{n}. The estimate on the right gives

|k|NIk|gngnIk|2𝑑xc2fnH1()2\sum_{|k|\leqslant N}\int_{I_{k}}|g_{n}-\langle g_{n}\rangle_{I_{k}}|^{2}\,dx\leqslant c_{2}\|f_{n}\|_{H^{-1}(\mathbb{R})}^{2}

for each NN\in\mathbb{N}. Sending nn\to\infty, the bound

|k|NIk|ggIk|2𝑑xc2fH1()2\sum_{|k|\leqslant N}\int_{I_{k}}|g-\langle g\rangle_{I_{k}}|^{2}\,dx\leqslant c_{2}\|f\|_{H^{-1}(\mathbb{R})}^{2}

appears. Taking NN\to\infty, one has the right estimate in (5.3). In particular, it shows that the sum in (5.3) converges. By construction,

kIk|gngnIk|2𝑑x=nkn2Ik|ggIk|2𝑑x+ϵn,\sum_{k\in\mathbb{Z}}\int_{I_{k}}|g_{n}-\langle g_{n}\rangle_{I_{k}}|^{2}\,dx=\sum_{-n\leqslant k\leqslant n-2}\int_{I_{k}}|g-\langle g\rangle_{I_{k}}|^{2}\,dx+\epsilon_{n}\,,

where ϵn\epsilon_{n} is a sum of integrals over In2,In1,In1,InI_{-n-2},I_{-n-1},I_{n-1},I_{n}. Since oL2()o\in L^{2}(\mathbb{R}),

limnIk|gngnIk|2𝑑x=0,k{n2,n1,n1,n}.\lim_{n\to\infty}\int_{I_{k}}|g_{n}-\langle g_{n}\rangle_{I_{k}}|^{2}\,dx=0,\quad k\in\{-n-2,-n-1,n-1,n\}\,.

Hence, limnϵn=0\lim_{n\to\infty}\epsilon_{n}=0 and, taking nn\to\infty in inequality

c1fnH1()2kIk|gngnIk|2𝑑x,c_{1}\|f_{n}\|_{H^{-1}(\mathbb{R})}^{2}\leqslant\sum_{k\in\mathbb{Z}}\int_{I_{k}}|g_{n}-\langle g_{n}\rangle_{I_{k}}|^{2}\,dx\,,

one gets the left bound in (5.3). Since all antiderivatives are different by a constant and the integral in (5.3) does not change if we add a constant to gg, the proof is finished. ∎

5.2. Auxiliary perturbative results for a single interval

Notice that for any real symmetric 2×22\times 2 matrix QQ with zero trace, we have that V=JQV=JQ is also real, symmetric and has zero trace. The converse statement is true as well. Hence, the equation JNQ+QNQ=0JN_{Q}^{\prime}+QN_{Q}=0 in Theorem 5.1, which is equivalent to NQ=JQNQN_{Q}^{\prime}=JQN_{Q}, can be written as NQ=VNQN_{Q}^{\prime}=VN_{Q} with VV having the same properties as QQ. Let U+(x,y)U_{+}(x,y) denote the solution to

ddxU+(x,y)=V(x)U+(x,y),U+(y,y)=I\frac{d}{dx}U_{+}(x,y)=V(x)U_{+}(x,y),\quad U_{+}(y,y)=I

and U(x,y)U_{-}(x,y) denote the solution to

ddxU(x,y)=V(x)U(x,y),U(y,y)=I.\frac{d}{dx}U_{-}(x,y)=-V(x)U_{-}(x,y),\quad U_{-}(y,y)=I\,.
Lemma 5.1.

Suppose N=VN,N(0)=IN^{\prime}=VN,N(0)=I, where VV is real-valued, VL1[0,1]V\in L^{1}[0,1], V=VV=V^{*}, and trV=0{\rm tr}\,V=0. Then, for =NN\mathcal{H}=N^{*}N, we have

det01(ξ)𝑑ξ=120101tr(U+(x,y)U+(x,y))𝑑x𝑑y=120101U+(x,y)HS2𝑑x𝑑y,\displaystyle\det\int_{0}^{1}\mathcal{H}(\xi)\,d\xi=\frac{1}{2}\int_{0}^{1}\int_{0}^{1}{\rm tr}\,\Bigl{(}U^{*}_{+}(x,y)U_{+}(x,y)\Bigr{)}\,dxdy=\frac{1}{2}\int_{0}^{1}\int_{0}^{1}\|U_{+}(x,y)\|^{2}_{\rm HS}\,dxdy, (5.6)
det01(ξ)𝑑ξ1=120101(U+(x,y)U(x,y))e12𝑑x𝑑y.\displaystyle\det\int_{0}^{1}\mathcal{H}(\xi)\,d\xi-1=\frac{1}{2}\int_{0}^{1}\int_{0}^{1}\left\|\Bigl{(}U_{+}(x,y)-U_{-}(x,y)\Bigr{)}e_{1}\right\|^{2}dxdy\,. (5.7)

Proof.  Notice that N,U+,USL(2,)N,U_{+},U_{-}\in{\rm SL}(2,\mathbb{R}) and that every matrix ASL(2,)A\in{\rm SL}(2,\mathbb{R}) satisfies

JA=A1J,AJ=J(A)1.JA^{*}=A^{-1}J,\qquad AJ=J(A^{*})^{-1}\,. (5.8)

Also, for any real 2×22\times 2 matrix BB, we have

detB=JBe1,Be2=JBe2,Be1.\det B=\langle JBe_{1},Be_{2}\rangle=-\langle JBe_{2},Be_{1}\rangle\,.

Hence,

:=det01(ξ)𝑑ξ=0101JN(x)N(x)e1,N(y)N(y)e2𝑑x𝑑y\displaystyle\mathcal{I}:=\det\int_{0}^{1}\mathcal{H}(\xi)d\xi=\int_{0}^{1}\int_{0}^{1}\langle JN^{*}(x)N(x)e_{1},N^{*}(y)N(y)e_{2}\rangle dxdy
=0101JN(x)N(x)e2,N(y)N(y)e1𝑑x𝑑y.\displaystyle=-\int_{0}^{1}\int_{0}^{1}\langle JN^{*}(x)N(x)e_{2},N^{*}(y)N(y)e_{1}\rangle dxdy\,.

For the second integrand, we have

JN(x)N(x)e1,N(y)N(y)e2=N(y)N(y)JN(x)N(x)e1,e2.\langle JN^{*}(x)N(x)e_{1},N^{*}(y)N(y)e_{2}\rangle=\langle N^{*}(y)N(y)JN^{*}(x)N(x)e_{1},e_{2}\rangle\,.

Then, identities (5.8) imply

N(y)N(y)JN(x)N(x)=N(y)J(N(y))1N(x)N(x)=J(N(y))1(N(y))1N(x)N(x)N^{*}(y)N(y)JN^{*}(x)N(x)=N^{*}(y)J(N^{*}(y))^{-1}N^{*}(x)N(x)=J(N(y))^{-1}(N^{*}(y))^{-1}N^{*}(x)N(x)

and, since Je1=e2Je_{1}=e_{2} and J=JJ^{*}=-J,

JN(x)N(x)e1,N(y)N(y)e2=(N(y))1(N(y))1N(x)N(x)e1,e1.\langle JN^{*}(x)N(x)e_{1},N^{*}(y)N(y)e_{2}\rangle=\langle(N(y))^{-1}(N^{*}(y))^{-1}N^{*}(x)N(x)e_{1},e_{1}\rangle\,.

Similarly, JN(x)N(x)e2,N(y)N(y)e1=(N(y))1(N(y))1N(x)N(x)e2,e2.\langle JN^{*}(x)N(x)e_{2},N^{*}(y)N(y)e_{1}\rangle=-\langle(N(y))^{-1}(N^{*}(y))^{-1}N^{*}(x)N(x)e_{2},e_{2}\rangle\,. Hence,

=120101j=12(N(y))1(N(y))1N(x)N(x)ej,ejdxdy=\displaystyle\mathcal{I}=\frac{1}{2}\int_{0}^{1}\int_{0}^{1}\sum_{j=1}^{2}\langle(N(y))^{-1}(N^{*}(y))^{-1}N^{*}(x)N(x)e_{j},e_{j}\rangle dxdy=\hskip 85.35826pt
120101tr((N(y))1(N(y))1N(x)N(x))𝑑x𝑑y=120101tr((N(y))1N(x)N(x)(N(y))1)𝑑x𝑑y.\displaystyle\frac{1}{2}\int_{0}^{1}\int_{0}^{1}{\rm tr}\Bigl{(}(N(y))^{-1}(N^{*}(y))^{-1}N^{*}(x)N(x)\Bigr{)}dxdy=\frac{1}{2}\int_{0}^{1}\int_{0}^{1}{\rm tr}\Bigl{(}(N^{*}(y))^{-1}N^{*}(x)N(x)(N(y))^{-1}\Bigr{)}dxdy\,.

Now, we use the formula N(x)(N(y))1=U+(x,y)N(x)(N(y))^{-1}=U_{+}(x,y) to rewrite the last expression as

=120101tr(U+(x,y)U+(x,y))𝑑x𝑑y.\mathcal{I}=\frac{1}{2}\int_{0}^{1}\int_{0}^{1}{\rm tr}\Bigl{(}U^{*}_{+}(x,y)U_{+}(x,y)\Bigr{)}dxdy\,.

Finally, (5.7) follows from U+(x,y)SL(2,)U_{+}(x,y)\in{\rm SL}(2,\mathbb{R}) by direct inspection after one uses the identities JU+(x,y)J=U(x,y)JU_{+}(x,y)J=-U_{-}(x,y) and tr(AA)2=(A+JAJ)e12{\rm tr}(A^{*}A)-2=\|(A+JAJ)e_{1}\|^{2}, which holds for every ASL(2,)A\in{\rm SL}(2,\mathbb{R}). ∎

Remark. The integrand in (5.6) is symmetric: tr(U+(x,y)U+(x,y))=tr(U+(y,x)U+(y,x)){\rm tr}\Bigl{(}U^{*}_{+}(x,y)U_{+}(x,y)\Bigr{)}={\rm tr}\Bigl{(}U^{*}_{+}(y,x)U_{+}(y,x)\Bigr{)} because U+(x,y)=U+1(y,x)U_{+}(x,y)=U^{-1}_{+}(y,x) and U+(x,y)SL(2,)U_{+}(x,y)\in{\rm SL}(2,\mathbb{R}). Notice also, that

tr(U+(x,y)U+(x,y))=λx,y2+λx,y22,{\rm tr}\Bigl{(}U^{*}_{+}(x,y)U_{+}(x,y)\Bigr{)}=\lambda_{x,y}^{2}+\lambda_{x,y}^{-2}\geqslant 2\,,

where λx,y\lambda_{x,y} is an eigenvalue of U+(x,y)U+(x,y)U^{*}_{+}(x,y)U_{+}(x,y) which explains why the left-hand side in (5.7) is nonnegative.

Lemma 5.2.

Suppose real-valued matrix-function V=(v1v2v2v1)V=\left(\begin{smallmatrix}v_{1}&v_{2}\\ v_{2}&-v_{1}\end{smallmatrix}\right) is defined on [0,1][0,1] and satisfies VL1[0,1]<\|V\|_{L^{1}[0,1]}<\infty. Consider =NN\mathcal{H}=N^{*}N, where NN: N=VN,N(0)=IN^{\prime}=VN,N(0)=I. Then,

det01𝑑x1VL1[0,1]2exp(CVL1[0,1]).\det\int_{0}^{1}\mathcal{H}\,dx-1\lesssim\|V\|^{2}_{L^{1}[0,1]}\exp(C\|V\|_{L^{1}[0,1]})\,. (5.9)

Proof.  The integral equation for NN is

N=I+0xVN𝑑s.N=I+\int_{0}^{x}VNds\,. (5.10)

By Gronwall’s inequality,

N(x)exp(0xV(s)𝑑s)exp(VL1[0,1]).\|N(x)\|\leqslant\exp\left(\int_{0}^{x}\|V(s)\|ds\right)\leqslant\exp(\|V\|_{L^{1}[0,1]}). (5.11)

Iteration of (5.10) gives

N=I+0xV𝑑x1+0xV(x1)(0x1V(x2)N(x2)𝑑x2)𝑑x1.N=I+\int_{0}^{x}Vdx_{1}+\int_{0}^{x}V(x_{1})\left(\int_{0}^{x_{1}}V(x_{2})N(x_{2})dx_{2}\right)dx_{1}\,.

Then,

01NN𝑑x=I+201(0xV(x1)𝑑x1)𝑑x+R,RVL1[0,1]2exp(CVL1[0,1]).\int_{0}^{1}N^{*}Ndx=I+2\int_{0}^{1}\left(\int_{0}^{x}V(x_{1})dx_{1}\right)dx+R,\;\;\|R\|\lesssim\|V\|^{2}_{L^{1}[0,1]}\exp(C\|V\|_{L^{1}[0,1]})\,.

Since trV=0\text{tr}\,V=0, the identity det(I+A)=1+trA+detA,\det(I+A)=1+\text{tr}A+\det A\,, which holds for all 2×22\times 2 matrices AA, gives

det01𝑑x1VL1[0,1]2exp(CVL1[0,1]).\det\int_{0}^{1}\mathcal{H}dx-1\lesssim\|V\|^{2}_{L^{1}[0,1]}\exp(C\|V\|_{L^{1}[0,1]})\,.

Lemma 5.3.

Suppose real-valued symmetric matrix-functions VV and OO are defined on [0,1][0,1] and satisfy

V=(v1v2v2v1)=O+O,O=O=(o1o2o2o1),\displaystyle V=\left(\begin{matrix}v_{1}&v_{2}\\ v_{2}&-v_{1}\end{matrix}\right)=O+O^{\prime},O=O^{*}=\left(\begin{matrix}o_{1}&o_{2}\\ o_{2}&-o_{1}\end{matrix}\right)\,, (5.12)
δ:=OL2[0,1]<,\displaystyle\delta:=\|O\|_{L^{2}[0,1]}<\infty\,, (5.13)
d:=OL2[0,1]<.\displaystyle d:=\|O^{\prime}\|_{L^{2}[0,1]}<\infty\,. (5.14)

Consider =NN\mathcal{H}=N^{*}N where N=VN,N(0)=IN^{\prime}=VN,N(0)=I. Then, we have

det01𝑑x1=4j=1201|gjgj|2𝑑x+r,|r|δ2.5exp(C(d+δ)),\det\int_{0}^{1}\mathcal{H}\,dx-1=4\sum_{j=1}^{2}\int_{0}^{1}|g_{j}-\langle g_{j}\rangle|^{2}\,dx+r,\,|r|\lesssim\delta^{2.5}\exp(C(d+\delta))\,, (5.15)

where

gj:=0xvj𝑑xg_{j}:=\int_{0}^{x}v_{j}\,dx (5.16)

and CC is an absolute positive constant. An analogous result holds if OO and VV are related by V=OOV=O-O^{\prime}.

Proof.  We will use the formula (5.7) for our analysis. Fix y[0,1]y\in[0,1] and take U+(x,y)U_{+}(x,y) and U(x,y)U_{-}(x,y) which solve ddxU+(x,y)=V(x)U+(x,y),U+(y,y)=I\frac{d}{dx}U_{+}(x,y)=V(x)U_{+}(x,y),U_{+}(y,y)=I and ddxU(x,y)=V(x)U(x,y),U(y,y)=I\frac{d}{dx}U_{-}(x,y)=-V(x)U_{-}(x,y),U_{-}(y,y)=I. Iterating the corresponding integral equations, one gets

U+(x,y)=I+yxV𝑑x1+yxVyx1V𝑑x2𝑑x1+yxVyx1Vyx2V𝑑x3𝑑x2𝑑x1+\displaystyle U_{+}(x,y)=I+\int_{y}^{x}Vdx_{1}+\int_{y}^{x}V\int_{y}^{x_{1}}Vdx_{2}dx_{1}+\int_{y}^{x}V\int_{y}^{x_{1}}V\int_{y}^{x_{2}}Vdx_{3}dx_{2}dx_{1}+
yxVyx1Vyx2Vyx3V𝑑x4𝑑x3𝑑x2𝑑x1+yxVyx1Vyx2Vyx3Vf+𝑑x4𝑑x3𝑑x2𝑑x1,\displaystyle\int_{y}^{x}V\int_{y}^{x_{1}}V\int_{y}^{x_{2}}V\int_{y}^{x_{3}}Vdx_{4}dx_{3}dx_{2}dx_{1}+\int_{y}^{x}V\int_{y}^{x_{1}}V\int_{y}^{x_{2}}V\int_{y}^{x_{3}}Vf_{+}dx_{4}dx_{3}dx_{2}dx_{1},
f+(x4)=yx4V(s)U+(s,y)𝑑s.\displaystyle f_{+}(x_{4})=\int_{y}^{x_{4}}V(s)U_{+}(s,y)ds\,.
U(x,y)=IyxV𝑑x1+yxVyx1V𝑑x2𝑑x1yxVyx1Vyx2V𝑑x3𝑑x2𝑑x1+\displaystyle U_{-}(x,y)=I-\int_{y}^{x}Vdx_{1}+\int_{y}^{x}V\int_{y}^{x_{1}}Vdx_{2}dx_{1}-\int_{y}^{x}V\int_{y}^{x_{1}}V\int_{y}^{x_{2}}Vdx_{3}dx_{2}dx_{1}+
yxVyx1Vyx2Vyx3V𝑑x4𝑑x3𝑑x2𝑑x1yxVyx1Vyx2Vyx3Vf𝑑x4𝑑x3𝑑x2𝑑x1,\displaystyle\int_{y}^{x}V\int_{y}^{x_{1}}V\int_{y}^{x_{2}}V\int_{y}^{x_{3}}Vdx_{4}dx_{3}dx_{2}dx_{1}-\int_{y}^{x}V\int_{y}^{x_{1}}V\int_{y}^{x_{2}}V\int_{y}^{x_{3}}Vf_{-}dx_{4}dx_{3}dx_{2}dx_{1},
f(x4)=yx4V(s)U(s,y)𝑑s.\displaystyle f_{-}(x_{4})=\int_{y}^{x_{4}}V(s)U_{-}(s,y)ds\,.

Taking U+(x,y)U(x,y)U_{+}(x,y)-U_{-}(x,y) as in (5.7) leaves us with

U+(x,y)U(x,y)2=yxV𝑑x1+1+2,\displaystyle\frac{U_{+}(x,y)-U_{-}(x,y)}{2}=\int_{y}^{x}Vdx_{1}+\mathcal{I}_{1}+\mathcal{I}_{2}\,, (5.17)
1=yxVyx1Vyx2V𝑑x3𝑑x2𝑑x1,\displaystyle\mathcal{I}_{1}=\int_{y}^{x}V\int_{y}^{x_{1}}V\int_{y}^{x_{2}}Vdx_{3}dx_{2}dx_{1}, (5.18)
2=yxVyx1Vyx2Vyx3V(f++f)𝑑x4𝑑x3𝑑x2𝑑x1.\displaystyle\mathcal{I}_{2}=\int_{y}^{x}V\int_{y}^{x_{1}}V\int_{y}^{x_{2}}V\int_{y}^{x_{3}}V(f_{+}+f_{-})dx_{4}dx_{3}dx_{2}dx_{1}\,. (5.19)

Recall that V=O+OV=O+O^{\prime} where OO satisfies (5.13) and (5.14). These assumptions are to be used in the following proposition. On +2\mathbb{R}_{+}^{2}, we define the partial order

[x1x2][y1y2]\left[\begin{array}[]{c}x_{1}\\ x_{2}\end{array}\right]\leqslant\left[\begin{array}[]{c}y_{1}\\ y_{2}\end{array}\right]

by requiring that x1y1x_{1}\leqslant y_{1} and x2y2x_{2}\leqslant y_{2}.

Proposition 5.2.

Suppose a matrix-function OO is defined on [0,1][0,1] and denote

δ=OL2[0,1],d=OL2[0,1].\displaystyle\delta=\|O\|_{L^{2}[0,1]},\,d=\|O^{\prime}\|_{L^{2}[0,1]}. (5.20)

Let an operator G(y)G_{(y)} be given by: F(G(y)F)(x)=yx(O+O)F𝑑sF\mapsto(G_{(y)}F)(x)=\int_{y}^{x}(O+O^{\prime})Fds where y[0,1]y\in[0,1] and a matrix-function FF, defined on [0,1][0,1], satisfies FL[0,1]<\|F\|_{L^{\infty}[0,1]}<\infty and FL2[0,1]<\|F^{\prime}\|_{L^{2}[0,1]}<\infty. Then,

[G(y)FL[0,1](G(y)F)L2[0,1]]C[FL[0,1]FL2[0,1]],=(δ+δdδδ+d0),\left[\begin{array}[]{l}\|G_{(y)}F\|_{L^{\infty}[0,1]}\\ \|(G_{(y)}F)^{\prime}\|_{L^{2}[0,1]}\end{array}\right]\leqslant C\mathcal{M}\left[\begin{array}[]{c}\|F\|_{L^{\infty}[0,1]}\\ \|F^{\prime}\|_{L^{2}[0,1]}\end{array}\right]\,,\quad\mathcal{M}=\left(\begin{array}[]{ll}\delta+\sqrt{\delta d}&\delta\\ \delta+d&0\end{array}\right)\,, (5.21)

where CC is an absolute positive constant, the norms and derivatives are computed with respect to xx.

Proof.  Let b=FL[0,1],c=FL2[0,1]b=\|F\|_{L^{\infty}[0,1]},c=\|F^{\prime}\|_{L^{2}[0,1]}. Write

O(x)O(x)O(y)O(y)=yx((O)O+OO)𝑑s.O^{*}(x)O(x)-O^{*}(y)O(y)=\int_{y}^{x}((O^{*})^{\prime}O+O^{*}O^{\prime})ds\,. (5.22)

Then,

O(x)2=maxξ21O(x)ξ22=maxξ21O(x)O(x)ξ,ξ(5.22)O(y)2+201O(s)O(s)𝑑s.\|O(x)\|^{2}=\max_{\|\xi\|_{\mathbb{C}^{2}}\leqslant 1}\|O(x)\xi\|^{2}_{\mathbb{C}^{2}}=\max_{\|\xi\|_{\mathbb{C}^{2}}\leqslant 1}\langle O^{*}(x)O(x)\xi,\xi\rangle\stackrel{{\scriptstyle\eqref{lak}}}{{\leqslant}}\|O(y)\|^{2}+2\int_{0}^{1}\|O^{\prime}(s)\|\cdot\|O(s)\|ds\,.

Applying Cauchy-Schwarz inequality to the integral, integrating in yy from 0 to 11 and maximizing in xx gives

OL[0,1]δ+(dδ)12.\|O\|_{L^{\infty}[0,1]}\lesssim\delta+(d\delta)^{\frac{1}{2}}\,. (5.23)

Then,

(G(y)F)(x)=yxOF𝑑s+O(x)F(x)O(y)F(y)yxOF𝑑s(G_{(y)}F)(x)=\int_{y}^{x}OFds+O(x)F(x)-O(y)F(y)-\int_{y}^{x}OF^{\prime}ds

and the estimate for the first coordinate in (5.21) follows from Cauchy-Schwarz inequality and (5.23). Since (G(y)F)=(O+O)F,(G_{(y)}F)^{\prime}=(O+O^{\prime})F\,, we get (G(y)F)L2[0,1](OL2[0,1]+OL2[0,1])FL[0,1]\|(G_{(y)}F)^{\prime}\|_{L^{2}[0,1]}\leqslant(\|O\|_{L^{2}[0,1]}+\|O^{\prime}\|_{L^{2}[0,1]})\|F\|_{L^{\infty}[0,1]} and the bound for the second coordinate in (5.21) is obtained. ∎

Continuation of the proof of Lemma 5.3. We apply the proposition to 1\mathcal{I}_{1} three times with the initial choice of FF: F=IF=I. That gives rise to taking the third power of matrix \mathcal{M}: 3\mathcal{M}^{3}, applying it to (1,0)t(1,0)^{t}, and looking at the first coordinate. As the result, one has 1L[0,1]δ32(δ+d)32\|\mathcal{I}_{1}\|_{L^{\infty}[0,1]}\lesssim\delta^{\frac{3}{2}}(\delta+d)^{\frac{3}{2}}. Therefore,

1e1L([0,1]2)δ32exp(δ+d).\|\mathcal{I}_{1}e_{1}\|_{L^{\infty}([0,1]^{2})}\lesssim\delta^{\frac{3}{2}}\exp(\delta+d). (5.24)

Similarly, we consider 2\mathcal{I}_{2} and use the previous proposition four times making the first choice of FF as F=f++fF=f_{+}+f_{-}. Applying the bound (5.11) to U+U_{+} and UU_{-}, we get f++fL[0,1](δ+d)exp(δ+d),f++fL2[0,1](δ+d)exp(δ+d)\|f_{+}+f_{-}\|_{L^{\infty}[0,1]}\lesssim(\delta+d)\exp(\delta+d),\|f^{\prime}_{+}+f_{-}^{\prime}\|_{L^{2}[0,1]}\lesssim(\delta+d)\exp(\delta+d). This time, we compute the fourth power of matrix :4\mathcal{M}:\mathcal{M}^{4}, apply it to vector (δ+d)exp(δ+d)(1,1)t(\delta+d)\exp(\delta+d)(1,1)^{t}, and look at the first coordinate. In the end, one has

2e1L([0,1]2)δ2exp(C(d+δ)).\|\mathcal{I}_{2}e_{1}\|_{L^{\infty}([0,1]^{2})}\lesssim\delta^{2}\exp(C(d+\delta))\,. (5.25)

The first term in (5.17) can be written as

yxV𝑑s=yxO𝑑s+O(x)O(y)\int_{y}^{x}Vds=\int_{y}^{x}Ods+O(x)-O(y)

and

yxO𝑑s+O(x)O(y)L2([0,1]2)δ.\left\|\int_{y}^{x}Ods+O(x)-O(y)\right\|_{L^{2}([0,1]^{2})}\lesssim\delta\,. (5.26)

For any three vectors v1,v2v_{1},v_{2} and v3v_{3} in 2\mathbb{R}^{2}, we have an estimate

|v1+v2+v3v1|v2+v3v2+v3,|\|v_{1}+v_{2}+v_{3}\|-\|v_{1}\||\leqslant\|v_{2}+v_{3}\|\leqslant\|v_{2}\|+\|v_{3}\|\,,

which follows from the triangle inequality. Multiplying with

v1+v2+v3+v12v1+v2+v3,\|v_{1}+v_{2}+v_{3}\|+\|v_{1}\|\leqslant 2\|v_{1}\|+\|v_{2}\|+\|v_{3}\|\,,

we get

|v1+v2+v32v12|2v1(v2+v3)+(v2+v3)2.|\|v_{1}+v_{2}+v_{3}\|^{2}-\|v_{1}\|^{2}|\leqslant 2\|v_{1}\|(\|v_{2}\|+\|v_{3}\|)+(\|v_{2}\|+\|v_{3}\|)^{2}\,.

Applying it to (5.17) gives

|14(U+(x,y)U(x,y))e12(yxV𝑑s)e12|\displaystyle\left|\frac{1}{4}\|(U_{+}(x,y)-U_{-}(x,y))e_{1}\|^{2}-\left\|\left(\int_{y}^{x}Vds\right)e_{1}\right\|^{2}\right|\hskip 113.81102pt
(yxV𝑑s)e1(1e1+2e1)+1e12+2e12.\displaystyle\lesssim\left\|\left(\int_{y}^{x}Vds\right)e_{1}\right\|\cdot(\|\mathcal{I}_{1}e_{1}\|+\|\mathcal{I}_{2}e_{1}\|)+\|\mathcal{I}_{1}e_{1}\|^{2}+\|\mathcal{I}_{2}e_{1}\|^{2}\,.

Taking L1([0,1]2)L^{1}([0,1]^{2}) norm in variables xx and yy of both sides and using (5.24), (5.25), (5.26) and the Cauchy-Schwartz inequality gives

140101(U+(x,y)U(x,y))e12𝑑x𝑑y=0101(yxV𝑑s)e12𝑑x𝑑y+r,|r|δ2.5exp(C(d+δ)).\frac{1}{4}\int_{0}^{1}\int_{0}^{1}\|(U_{+}(x,y)-U_{-}(x,y))e_{1}\|^{2}dxdy=\int_{0}^{1}\int_{0}^{1}\left\|\left(\int_{y}^{x}Vds\right)e_{1}\right\|^{2}dxdy+r,\,\,|r|\lesssim\delta^{2.5}\exp(C(d+\delta))\,.

Recalling the definition (5.16), we get

(yxV𝑑s)e12=j=12(gj(x)gj(y))2\left\|\left(\int_{y}^{x}Vds\right)e_{1}\right\|^{2}=\sum_{j=1}^{2}(g_{j}(x)-g_{j}(y))^{2}

so

120101(U+(x,y)U(x,y))e12𝑑x𝑑y=4j=1201|gjgj|2𝑑x+r,|r|δ2.5exp(C(d+δ)).\frac{1}{2}\int_{0}^{1}\int_{0}^{1}\|(U_{+}(x,y)-U_{-}(x,y))e_{1}\|^{2}dxdy=4\sum_{j=1}^{2}\int_{0}^{1}|g_{j}-\langle g_{j}\rangle|^{2}dx+r,\,\,|r|\lesssim\delta^{2.5}\exp(C(d+\delta))\,.

Lemma 5.3 is proved. ∎

Remark. All statements in this subsection can be easily adjusted to any interval but the constants in the inequalities will depend on the size of that interval.

5.3. Rough bound when 𝒦~Q\widetilde{\mathcal{K}}_{Q} is small

Lemma 5.4.

Suppose an absolutely continuous function ff is defined on [0,1][0,1] and satisfies

fL2[0,1],f=l1+l2,l1L1[0,1],l2L2[0,1].\displaystyle f\in L^{2}[0,1],\quad f^{\prime}=l_{1}+l_{2},\quad l_{1}\in L^{1}[0,1],\quad l_{2}\in L^{2}[0,1]\,. (5.27)

Then, fL[0,1]δ2+2(δτ+ϵ(τ+ϵ+δ)),\|f\|_{L^{\infty}[0,1]}\leqslant\sqrt{\delta^{2}+2(\delta\tau+\epsilon(\tau+\epsilon+\delta))}, where δ=fL2[0,1],ϵ=l1L1[0,1],τ=l2L2[0,1].\delta=\|f\|_{L^{2}[0,1]},\,\epsilon=\|l_{1}\|_{L^{1}[0,1]},\,\tau=\|l_{2}\|_{L^{2}[0,1]}\,.

Proof.  There is ξ[0,1]\xi\in[0,1] such that |f(ξ)|δ|f(\xi)|\leqslant\delta and

|f(x)f(ξ)||ξxf𝑑s|τ+ϵ.|f(x)-f(\xi)|\leqslant\left|\int_{\xi}^{x}f^{\prime}ds\right|\leqslant\tau+\epsilon\,.

Thus, fL[0,1]τ+ϵ+δ\|f\|_{L^{\infty}[0,1]}\leqslant\tau+\epsilon+\delta. Then, writing

f2(x)f2(y)=2yxff𝑑s,f^{2}(x)-f^{2}(y)=2\int_{y}^{x}ff^{\prime}ds\,,

integrating in yy and maximizing in xx, we get

fL[0,1]2δ2+2(δτ+ϵ(τ+ϵ+δ)).\|f\|^{2}_{L^{\infty}[0,1]}\leqslant\delta^{2}+2(\delta\tau+\epsilon(\tau+\epsilon+\delta))\,.

Suppose QQ is real-valued, symmetric matrix-function on \mathbb{R} with zero trace and QL2()<\|Q\|_{L^{2}(\mathbb{R})}<\infty. Define Q=NN\mathcal{H}_{Q}=N^{*}N, where N:N=JQN,N(0)=IN:N^{\prime}=JQN,N(0)=I. Notice that detnn+2SQS𝑑x=detnn+2Q𝑑x\det\int_{n}^{n+2}S^{*}\mathcal{H}_{Q}Sdx=\det\int_{n}^{n+2}\mathcal{H}_{Q}dx for every constant matrix SSL(2,)S\in{\rm SL}(2,\mathbb{R}). Therefore, we can apply Lemma 5.2 to each interval [n,n+2][n,n+2] by choosing S=N1(n)S=N^{-1}(n) and get an estimate which explains how QL2()\|Q\|_{L^{2}(\mathbb{R})} controls 𝒦~Q\widetilde{\mathcal{K}}_{Q}:

𝒦~Q=n(detnn+2Q𝑑x4)nQL2[n,n+2]2exp(CQ2)QL2()2exp(CQ2).\widetilde{\mathcal{K}}_{Q}=\sum_{n\in\mathbb{Z}}\left(\det\int_{n}^{n+2}\mathcal{H}_{Q}dx-4\right)\lesssim\sum_{n\in\mathbb{Z}}\|Q\|^{2}_{L^{2}[n,n+2]}\exp(C\|Q\|_{2})\lesssim\|Q\|^{2}_{L^{2}(\mathbb{R})}\exp(C\|Q\|_{2})\,.

The next lemma shows that 𝒦~Q\widetilde{\mathcal{K}}_{Q} controls the convolution of QQ with the exponential.

Lemma 5.5.

Suppose QQ is real-valued, symmetric 2×22\times 2 matrix-function on \mathbb{R} with zero trace and entries in L2()L^{2}(\mathbb{R}). Define Q=NN\mathcal{H}_{Q}=N^{*}N where N:N=JQN,N(0)=IN:N^{\prime}=JQN,N(0)=I and assume that 𝒦~Q<\widetilde{\mathcal{K}}_{Q}<\infty. If O:=exxesQ𝑑sO:=e^{x}\int_{x}^{\infty}e^{-s}Qds, then OL()exp(C(QL2()+𝒦~Q))𝒦~Q14\|O\|_{L^{\infty}(\mathbb{R})}\lesssim\exp(C(\|Q\|_{L^{2}(\mathbb{R})}+\widetilde{\mathcal{K}}_{Q}))\widetilde{\mathcal{K}}_{Q}^{{\frac{1}{4}}} where CC is a positive absolute constant.

Proof.  Let R=QL2()R=\|Q\|_{L^{2}(\mathbb{R})} and E=𝒦~QE=\widetilde{\mathcal{K}}_{Q}. We split the proof into several steps.

1. Bound for a single interval [0,1][0,1]. The definitions (3.5) and (3.14) imply that 𝒦~Q+E\widetilde{\mathcal{K}}_{Q}^{+}\leqslant E. From Theorem 1.2 and Theorem 3.2 in [3], we know that Q\mathcal{H}_{Q} admits the following factorization on +\mathbb{R}_{+}: Q=GWG\mathcal{H}_{Q}=G^{*}WG where GG and WW satisfy conditions:

G=J(v1+v2)G,v1L1(+)E,v2L2(+)E12,\displaystyle G^{\prime}=J(v_{1}+v_{2})G,\quad\|v_{1}\|_{L^{1}(\mathbb{R}_{+})}\lesssim E,\quad\|v_{2}\|_{L^{2}(\mathbb{R}_{+})}\lesssim E^{{\frac{1}{2}}}\,, (5.28)
detG=1,v1+v2=(v1+v2),\displaystyle\det G=1,\quad v_{1}+v_{2}=(v_{1}+v_{2})^{*}\,, (5.29)

and

W0,detW=1,trW2L1(+)E.\displaystyle W\geqslant 0,\quad\det W=1,\quad\|\text{tr}\,W-2\|_{L^{1}(\mathbb{R}_{+})}\lesssim E\,.

Since trW2L1[0,1]E\|\text{tr}\,W-2\|_{L^{1}[0,1]}\lesssim E, we have λ+λ12L1[0,1]E\|\lambda+\lambda^{-1}-2\|_{L^{1}[0,1]}\lesssim E, where λ\lambda is the largest eigenvalue of WW. If one denotes p=trW2=λ+λ12p=\text{tr}\,W-2=\lambda+\lambda^{-1}-2, then

λ=2+p+4p+p22,λ1=2+p4p+p22.\lambda=\frac{2+p+\sqrt{4p+p^{2}}}{2}\,,\quad\lambda^{-1}=\frac{2+p-\sqrt{4p+p^{2}}}{2}\,. (5.30)

In particular, that yields

01W𝑑x1+E.\int_{0}^{1}\|W\|\,dx\lesssim 1+E\,. (5.31)

The given conditions on QQ and (5.11) yield

N(x),N1(x)exp(CR),x[0,1],\|N(x)\|,\|N^{-1}(x)\|\lesssim\exp(CR),\quad x\in[0,1]\,,

where the second estimate follows from the first since detN=1\det N=1. The Hamiltonian Q=NN\mathcal{H}_{Q}=N^{*}N is absolutely continuous on +\mathbb{R}_{+} and

0<exp(CR)IQ(x)exp(CR)I0<\exp(-CR){I}\lesssim\mathcal{H}_{Q}(x)\lesssim\exp(CR){I} (5.32)

on [0,1][0,1]. We claim that G(0)exp(C(R+E))\|G(0)\|\lesssim\exp(C(R+E)) and that G1(0)exp(C(R+E))\|G^{-1}(0)\|\lesssim\exp(C(R+E)). Indeed, if XX satisfies X=J(v1+v2)XX^{\prime}=J(v_{1}+v_{2})X and X(0)=IX(0)=I, then G=XG(0)G=XG(0). Moreover, given conditions on v1v_{1} and v2v_{2} and detX=1\det X=1, we have

X(x)exp(CE),X1(x)exp(CE)\|X(x)\|\lesssim\exp(CE),\qquad\|X^{-1}(x)\|\lesssim\exp(CE) (5.33)

uniformly on [0,1][0,1]. Identity Q=G(0)XWXG(0)\mathcal{H}_{Q}=G^{*}(0)X^{*}WXG(0) yields

(G(0))1Q(G(0))1=XWX.(G^{*}(0))^{-1}\mathcal{H}_{Q}(G(0))^{-1}=X^{*}WX\,.

Taking an arbitrary ξ2\xi\in\mathbb{C}^{2} with ξ2=1\|\xi\|_{\mathbb{C}^{2}}=1, we get

G1(0)ξ2(5.32)exp(CR)01QG1(0)ξ,G1(0)ξ𝑑x\displaystyle\|G^{-1}(0)\xi\|^{2}\stackrel{{\scriptstyle\eqref{tot}}}{{\lesssim}}\exp(CR)\int_{0}^{1}\langle\mathcal{H}_{Q}G^{-1}(0)\xi,G^{-1}(0)\xi\rangle dx\hskip 85.35826pt
=exp(CR)01WXξ,Xξ𝑑x(5.31)+(5.33)exp(C(R+E)),\displaystyle=\exp(CR)\int_{0}^{1}\langle WX\xi,X\xi\rangle dx\stackrel{{\scriptstyle\eqref{P2p55}+\eqref{tot1}}}{{\lesssim}}\exp(C(R+E))\,,

which implies G1(0)exp(C(R+E))\|G^{-1}(0)\|\lesssim\exp(C(R+E)). We also have G(0)exp(C(R+E))\|G(0)\|\lesssim\exp(C(R+E)) since detG=1\det G=1 and the claim is proved. Finally, we have

G(x)exp(C(R+E)),G1(x)exp(C(R+E))\|G(x)\|\lesssim\exp(C(R+E)),\quad\|G^{-1}(x)\|\lesssim\exp(C(R+E))

for x[0,1]x\in[0,1] since G=XG(0)G=XG(0).

Next, let us study WW and WW^{\prime}. Since W=(G)1NNG1W=(G^{*})^{-1}N^{*}NG^{-1}, one has Wexp(C(R+E))\|W\|\lesssim\exp(C(R+E)) on x[0,1]x\in[0,1]. Recall that W0W\geqslant 0 and detW=1\det W=1, so

exp(C(R+E))IWexp(C(R+E))I,x[0,1].\exp(-C(R+E)){I}\lesssim W\lesssim\exp(C(R+E)){I},\quad x\in[0,1]\,.

Since λ\lambda is the largest eigenvalue of WW and λexp(C(R+E))\lambda\lesssim\exp(C(R+E)), then (5.30) yields λ1L2[0,1]E12exp(C(R+E))\|\lambda-1\|_{L^{2}[0,1]}\lesssim E^{{\frac{1}{2}}}\exp(C(R+E)) and λ11L2[0,1]E12exp(C(R+E))\|\lambda^{-1}-1\|_{L^{2}[0,1]}\lesssim E^{{\frac{1}{2}}}\exp(C(R+E)). Introduce Υ=WI.\Upsilon=W-I\,. The matrix Υ\Upsilon is unitarily equivalent to (λ1001/λ1)\left(\begin{smallmatrix}\lambda-1&0\\ 0&1/\lambda-1\end{smallmatrix}\right) and that gives

ΥL2[0,1]E12exp(C(R+E)).\|\Upsilon\|_{L^{2}[0,1]}\lesssim E^{{\frac{1}{2}}}\exp(C(R+E))\,. (5.34)

We need to study Υ\Upsilon^{\prime}, which is equal to WW^{\prime}. To do so, notice that

2NJQN=Q=GJ(v1+v2)WG+GWJ(v1+v2)G+GWG.\displaystyle 2N^{*}JQN=\mathcal{H}_{Q}^{\prime}=G^{*}J(v_{1}+v_{2})WG+G^{*}WJ(v_{1}+v_{2})G+G^{*}W^{\prime}G\,. (5.35)

Hence,

Υ=W=F1+F2,\Upsilon^{\prime}=W^{\prime}=F_{1}+F_{2}\,,

where

F1=J(v1+v2)WWJ(v1+v2),F2=2(G)1NJQNG1.F_{1}=-J(v_{1}+v_{2})W-WJ(v_{1}+v_{2}),\quad F_{2}=2(G^{*})^{-1}N^{*}JQNG^{-1}.

The previously obtained estimates give us

F1L1[0,1]E12exp(C(R+E)),F2L2[0,1]exp(C(R+E)).\|F_{1}\|_{L^{1}[0,1]}\lesssim E^{{\frac{1}{2}}}\exp(C(R+E)),\quad\|F_{2}\|_{L^{2}[0,1]}\lesssim\exp(C(R+E))\,. (5.36)

Now, we use (5.34), (5.36) to apply the previous lemma to each component of Υ\Upsilon to obtain

ΥL[0,1]E14exp(C(R+E)).\|\Upsilon\|_{L^{\infty}[0,1]}\lesssim E^{{\frac{1}{4}}}\exp(C(R+E))\,. (5.37)

The formula (5.35) also gives an expression for QQ:

Q=J(H1+H2),Q=-J(H_{1}+H_{2})\,,

where

H1=0.5(N)1(GJ(v1+v2)WG+GWJ(v1+v2)G)N1H_{1}=0.5(N^{*})^{-1}(G^{*}J(v_{1}+v_{2})WG+G^{*}WJ(v_{1}+v_{2})G)N^{-1}

and

H2=0.5(N)1(GΥG)N1.H_{2}=0.5(N^{*})^{-1}(G^{*}\Upsilon^{\prime}G)N^{-1}\,.

Since H1L1[0,1]E12exp(C(R+E))\|H_{1}\|_{L^{1}[0,1]}\lesssim E^{{\frac{1}{2}}}\exp(C(R+E)), we have

exx1esH1𝑑sL[0,1]E12exp(C(R+E)).\left\|e^{x}\int_{x}^{1}e^{-s}H_{1}ds\right\|_{L^{\infty}[0,1]}\lesssim E^{{\frac{1}{2}}}\exp(C(R+E))\,.

For smooth matrix-functions u1u_{1}, u2u_{2}, u3u_{3}, we have

x1u1u2u3ds=u1u2u3|1x1xu1u2u3ds1xu1u2u3ds.\int^{1}_{x}u_{1}u^{\prime}_{2}u_{3}\,ds=u_{1}u_{2}u_{3}\Bigr{\rvert}^{1}_{x}-\int^{1}_{x}u^{\prime}_{1}u_{2}u_{3}\,ds-\int^{1}_{x}u_{1}u_{2}u_{3}^{\prime}\,ds.

Then,

2exx1esH2𝑑s=\displaystyle 2e^{x}\int_{x}^{1}e^{-s}H_{2}ds=\hskip 142.26378pt
ex(e1(N(1))1G(1)Υ(1)G(1)(N(1))1ex(N(x))1G(x)Υ(x)G(x)(N(x))1)\displaystyle e^{x}\left(e^{-1}(N^{*}(1))^{-1}G^{*}(1)\Upsilon(1)G(1)(N(1))^{-1}-e^{-x}(N^{*}(x))^{-1}G^{*}(x)\Upsilon(x)G(x)(N(x))^{-1}\right)
exx1(es(N(s))1G)ΥGN1𝑑sexx1es(N(s))1GΥ(GN1)𝑑s.\displaystyle-e^{x}\int_{x}^{1}(e^{-s}(N^{*}(s))^{-1}G^{*})^{\prime}\Upsilon GN^{-1}ds-e^{x}\int_{x}^{1}e^{-s}(N^{*}(s))^{-1}G^{*}\Upsilon(GN^{-1})^{\prime}ds\,.

Since (N1)L2[0,1]exp(C(R+E))\|(N^{-1})^{\prime}\|_{L^{2}[0,1]}\lesssim\exp(C(R+E)) and GL1[0,1]exp(C(R+E))\|G^{\prime}\|_{L^{1}[0,1]}\lesssim\exp(C(R+E)), we have

exx1esH2𝑑sL[0,1]ΥL[0,1]exp(C(R+E))(5.37)E14exp(C(R+E)).\left\|e^{x}\int_{x}^{1}e^{-s}H_{2}ds\right\|_{L^{\infty}[0,1]}\lesssim\|\Upsilon\|_{L^{\infty}[0,1]}\exp(C(R+E))\stackrel{{\scriptstyle\eqref{tot9}}}{{\lesssim}}E^{{\frac{1}{4}}}\exp(C(R+E))\,.

Summing up, we get

exx1esQ𝑑sL[0,1]E14exp(C(R+E)).\left\|e^{x}\int_{x}^{1}e^{-s}Qds\right\|_{L^{\infty}[0,1]}\lesssim E^{{\frac{1}{4}}}\exp(C(R+E))\,. (5.38)

2. Handling all intervals [n,n+1],n[n,n+1],n\in\mathbb{Z}. Take any nn\in\mathbb{Z}. Our immediate goal is to show the bound

exxn+1esQ𝑑sL[n,n+1]E14exp(C(R+E))\left\|e^{x}\int_{x}^{n+1}e^{-s}Qds\right\|_{L^{\infty}[n,n+1]}\lesssim E^{{\frac{1}{4}}}\exp(C(R+E)) (5.39)

analogous to (5.38) but written for interval [n,n+1][n,n+1]. To this end, take the Hamiltonian (n)(x):=Q(x+n)\mathcal{H}^{(n)}(x):=\mathcal{H}_{Q}(x+n) defined on +\mathbb{R}_{+}. For the corresponding 𝒦~(n)+\widetilde{\mathcal{K}}_{(n)}^{+}, we get 𝒦~(n)+E\widetilde{\mathcal{K}}_{(n)}^{+}\leqslant E as follows from its definition. Since the 𝒦~\widetilde{\mathcal{K}}-characteristics of the Hamiltonians \mathcal{H} and SSS^{*}\mathcal{H}S are equal for every constant matrix SSL(2,)S\in{\rm SL}(2,\mathbb{R}), we can instead consider ^(n)=N^N^\widehat{\mathcal{H}}^{(n)}=\widehat{N}^{*}\widehat{N} where N^=JQ(x+n)N^,N^(0)=I\widehat{N}^{\prime}=JQ(x+n)\widehat{N},\widehat{N}(0)=I. Using the arguments in step 1 for ^(n)\widehat{\mathcal{H}}^{(n)}, we get (5.39).

3. Summing up. Denote On(x)=exxesQχn<s<n+1𝑑sO_{n}(x)=e^{x}\int_{x}^{\infty}e^{-s}Q\cdot\chi_{n<s<n+1}ds and notice that O=nOnO=\sum_{n\in\mathbb{Z}}O_{n}. Then, since On(x)=0O_{n}(x)=0 for x>n+1x>n+1 and On(x)exnOnL[n,n+1]\|O_{n}(x)\|\lesssim e^{x-n}\|O_{n}\|_{L^{\infty}[n,n+1]} for x<nx<n, we get

O(x)nOn(x)E14exp(C(R+E))n0enE14exp(C(R+E))\|O(x)\|\leqslant\sum_{n\in\mathbb{Z}}\|O_{n}(x)\|\lesssim E^{{\frac{1}{4}}}\exp(C(R+E))\sum_{n\geqslant 0}e^{-n}\sim E^{{\frac{1}{4}}}\exp(C(R+E))

as follows from (5.39). That finishes the proof of Lemma 5.5. ∎

5.4. Proof of Theorem 4.1

Denote E=𝒦~QE=\widetilde{\mathcal{K}}_{Q}, O=exxesQ𝑑sO=e^{x}\int_{x}^{\infty}e^{-s}Qds, and recall that OL2()QH1()QL2()\|O\|_{L^{2}(\mathbb{R})}\sim\|Q\|_{H^{-1}(\mathbb{R})}\leqslant\|Q\|_{L^{2}(\mathbb{R})}.


1. Lower bound. Define δn=OL2[n,n+1]\delta_{n}=\|O\|_{L^{2}[n,n+1]}. By Lemma 5.5, we know that supnδnE14exp(C(R+E))\sup_{n}\delta_{n}\lesssim E^{{\frac{1}{4}}}\exp(C(R+E)). Next, we apply Lemma 5.3 to each interval [n,n+2][n,n+2]. The remainder rnr_{n} in that lemma allows the estimate

rn(δn+δn+1)2.5exp(C(δn+δn+1+R)),n.r_{n}\lesssim(\delta_{n}+\delta_{n+1})^{2.5}\exp(C(\delta_{n}+\delta_{n+1}+R)),\quad n\in\mathbb{Z}\,.

For each R>0R>0 and η>0\eta>0, we can find a positive E0(R,η)E_{0}(R,\eta) such that E(0,E0(R,η))E\in(0,E_{0}(R,\eta)) implies that the remainder rnr_{n} is smaller than η(δn2+δn+12){\eta}(\delta_{n}^{2}+\delta_{n+1}^{2}) uniformly in all nn. For example, one can take

E0(R,η)eCηR,E_{0}(R,\eta)\sim e^{-C_{\eta}R}, (5.40)

where CηC_{\eta} is a sufficiently large positive number that depends on η\eta. Therefore, for such EE and some positive constant cc independent of η\eta, we have

n(cη)δn2n(detnn+2Q𝑑x4)n(c+η)δn2,\displaystyle\sum_{n\in\mathbb{Z}}(c-\eta)\delta_{n}^{2}\lesssim\sum_{n\in\mathbb{Z}}\left(\det\int_{n}^{n+2}\mathcal{H}_{Q}dx-4\right)\lesssim\sum_{n\in\mathbb{Z}}(c+\eta)\delta_{n}^{2},

where the Proposition 5.1 has been applied to the terms nn+2|gjgj|2𝑑x\int_{n}^{n+2}|g_{j}-\langle g_{j}\rangle|^{2}\,dx in the right-hand side of (5.15), adjusted to the interval [n,n+2][n,n+2], to show that they are comparable to δn2+δn+12\delta_{n}^{2}+\delta_{n+1}^{2}. Taking η=c/2\eta=c/2, we see that

E=n(detnn+2Q𝑑x4)nδn2OL2()2,E=\sum_{n\in\mathbb{Z}}\left(\det\int_{n}^{n+2}\mathcal{H}_{Q}dx-4\right)\sim\sum_{n\in\mathbb{Z}}\delta_{n}^{2}\sim\|O\|_{L^{2}(\mathbb{R})}^{2},

in the case EE0(R,c2)E\leqslant E_{0}(R,\frac{c}{2}). If E>E0(R,c2)E>E_{0}(R,\frac{c}{2}), one uses inequality OL2()R\|O\|_{L^{2}(\mathbb{R})}\lesssim R to get

eCROL2()2E0(R,c2)1+R2OL2()2E,e^{-CR}\|O\|^{2}_{L^{2}(\mathbb{R})}\lesssim\frac{E_{0}(R,\frac{c}{2})}{1+R^{2}}\|O\|^{2}_{L^{2}(\mathbb{R})}\lesssim E\,, (5.41)

which holds for some positive absolute constant CC due to (5.40). That provides the required lower bound.


2. Upper bound. Let δn=OL2[n,n+1].\delta_{n}=\|O\|_{L^{2}[n,n+1]}\,. For given value of RR, apply Lemma 5.3 and Proposition 5.1 to each interval [n,n+2][n,n+2]. That gives

Enδn2eC(R+δn)E\lesssim\sum_{n\in\mathbb{Z}}\delta^{2}_{n}e^{C(R+\delta_{n})}

with an absolute constant CC. Since nδn2qH1()2\sum_{n\in\mathbb{Z}}\delta_{n}^{2}\sim\|q\|_{H^{-1}(\mathbb{R})}^{2} and qH1()R\|q\|_{H^{-1}(\mathbb{R})}\lesssim R, one has

EqH1()2eC(R+qH1())qH1()2eC2R.E\lesssim\|q\|_{H^{-1}(\mathbb{R})}^{2}e^{C(R+\|q\|_{H^{-1}(\mathbb{R})})}\lesssim\|q\|_{H^{-1}(\mathbb{R})}^{2}e^{C_{2}R}\,.


6. Appendix

Here we collect some auxiliary results used in the main text.

1. We start with an example that shows that the scattering transform is not injective when defined on qL2()q\in L^{2}(\mathbb{R}). This is an analog of Lemma 17 in [24].

Example 6.1.

There exist potentials q1,q2L2()q_{1},q_{2}\in L^{2}(\mathbb{R}) such that q1q2q_{1}\neq q_{2} in L2()L^{2}(\mathbb{R}) but we have 𝐫q1=𝐫q2{\bf r}_{q_{1}}={\bf r}_{q_{2}} a.e. on \mathbb{R} for their reflection coefficients. In other words, the scattering transform q𝐫qq\mapsto{\bf r}_{q} is not injective on L2()L^{2}(\mathbb{R}).

Proof.  Let us consider

𝔞1+=1,𝔟1+=0,𝔞1=𝔞,𝔟1=𝔟,\mathfrak{a}_{1}^{+}=1,\quad\mathfrak{b}_{1}^{+}=0,\quad\mathfrak{a}_{1}^{-}=\mathfrak{a},\quad\mathfrak{b}_{1}^{-}=\mathfrak{b},

and

𝔞2+=𝔞,𝔟2+=𝔟,𝔞2=1,𝔟2=0,\mathfrak{a}_{2}^{+}=\mathfrak{a},\quad\mathfrak{b}_{2}^{+}=\mathfrak{b},\quad\mathfrak{a}_{2}^{-}=1,\quad\mathfrak{b}_{2}^{-}=0,

where 𝔞=1+i/x\mathfrak{a}=1+i/x, 𝔟=i/x\mathfrak{b}=i/x. Note that

log(1|sk±(x)|2)𝑑x>,sk±=𝔟±𝔞k±,k=1,2.\int_{\mathbb{R}}\log(1-|s_{k}^{\pm}(x)|^{2})\,dx>-\infty,\qquad s_{k}^{\pm}=\frac{\mathfrak{b}^{\pm}}{\mathfrak{a}_{k}^{\pm}},\qquad k=1,2.

Theorem 12.11 in [12] says that for every contractive analytic function ss on +\mathbb{C}_{+} whose boundary values on \mathbb{R} satisfy log(1|s|2)L1()\log(1-|s|^{2})\in L^{1}(\mathbb{R}) there exists a unique coefficient AL2(+)A\in L^{2}(\mathbb{R}_{+}) such that s=limξ+𝔅(ξ,λ)𝔄(ξ,λ)s=\lim_{\xi\to+\infty}\frac{\mathfrak{B}(\xi,\lambda)}{\mathfrak{A}(\xi,\lambda)}, λ+\lambda\in\mathbb{C}_{+} for the continuous Wall polynomials generated by AA. Moreover, we have

2πAL2(+)2=log(1|s|2)L1(+).2\pi\|A\|_{L^{2}(\mathbb{R}_{+})}^{2}=\|\log(1-|s|^{2})\|_{L^{1}(\mathbb{R}_{+})}. (6.1)

Applying this result, we see that there exist functions A1±,A2±L2(+)A_{1}^{\pm},A_{2}^{\pm}\in L^{2}(\mathbb{R}_{+}) such that 𝔞1,2±\mathfrak{a}^{\pm}_{1,2}, 𝔟1,2±\mathfrak{b}_{1,2}^{\pm} are the limits of their continuous Wall polynomials. Now define potentials q1,2L2()q_{1,2}\in L^{2}(\mathbb{R}) by relations

A1,2+(ξ)=q1,2(ξ/2)¯/2,A1,2(ξ)=q1,2(ξ/2)/2,ξ+.A_{1,2}^{+}(\xi)=-\overline{q_{1,2}(\xi/2)}/2,\qquad A_{1,2}^{-}(\xi)=q_{1,2}(-\xi/2)/2,\qquad\xi\in\mathbb{R}_{+}.

From Proposition 2.2, we conclude that the coefficients a1,2a_{1,2}, b1,2b_{1,2} for these potentials satisfy

a1=𝔞=a2,b1=𝔟=𝔟¯=b2,a_{1}=\mathfrak{a}=a_{2},\qquad b_{1}=-\mathfrak{b}=\overline{\mathfrak{b}}=b_{2},

on {0}\mathbb{R}\setminus\{0\}. Hence, 𝐫q1=𝐫q2{\bf r}_{q_{1}}={\bf r}_{q_{2}} on {0}\mathbb{R}\setminus\{0\}. On the other hand, we have A1+=0A_{1}^{+}=0 and A2=0A_{2}^{-}=0 by construction. It follows that suppq1(,0]\mathop{\mathrm{supp}}\nolimits q_{1}\subset(-\infty,0] and suppq2[0,+)\mathop{\mathrm{supp}}\nolimits q_{2}\subset[0,+\infty). Since q1q_{1} and q2q_{2} are nonzero (they have a nonzero L2()L^{2}(\mathbb{R})-norm as follows from (6.1)), that yields q1q2q_{1}\neq q_{2} in L2()L^{2}(\mathbb{R}). ∎

2. Next, we outline how to prove that the spectral representation for the Dirac operator 𝒟Q\mathcal{D}_{Q}, defined by relation (3.1), is given by the Weyl-Titchmarsh transform (3.10). To this end, we will use the corresponding result for canonical Hamiltonian systems proved in [21].

At first, we note that if Q=NQNQ\mathcal{H}_{Q}=N_{Q}^{*}N_{Q} is the Hamiltonian from Theorem 3.1, then detQ=1\det\mathcal{H}_{Q}=1 on \mathbb{R} and the operator V:XNQ1XV:X\mapsto N_{Q}^{-1}X is unitary from L2(,2)L^{2}(\mathbb{R},\mathbb{C}^{2}) onto the Hilbert space

L2(Q)={Y:2:YL2(Q,)2=Q(ξ)Y(ξ),Y(ξ)2𝑑ξ<}.L^{2}(\mathcal{H}_{Q})=\Bigl{\{}Y:\mathbb{R}\to\mathbb{C}^{2}:\;\|Y\|_{L^{2}(\mathcal{H}_{Q},\mathbb{R})}^{2}=\int_{\mathbb{R}}\langle\mathcal{H}_{Q}(\xi)Y(\xi),Y(\xi)\rangle_{\mathbb{C}^{2}}\,d\xi<\infty\Bigr{\}}.

Moreover, V𝒟QV1V\mathcal{D}_{Q}V^{-1} coincides with the operator 𝒟Q:Y1JY\mathcal{D}_{\mathcal{H}_{Q}}:Y\mapsto\mathcal{H}^{-1}JY^{\prime} of the canonical Hamiltonian system generated by the Hamiltonian Q\mathcal{H}_{Q}. Thus, the operator 𝒟Q\mathcal{D}_{Q} on L2(,2)L^{2}(\mathbb{R},\mathbb{C}^{2}) is unitary equivalent to the operator 𝒟Q\mathcal{D}_{\mathcal{H}_{Q}} on L2(Q)L^{2}(\mathcal{H}_{Q}). Let M~\widetilde{M} be the solution of Cauchy problem

JM~(ξ,z)=zQ(ξ)M~(ξ,z),M~(0,z)=(1001),J\widetilde{M}^{\prime}(\xi,z)=z\mathcal{H}_{Q}(\xi)\widetilde{M}(\xi,z),\qquad\widetilde{M}(0,z)=\left(\begin{smallmatrix}1&0\\ 0&1\\ \end{smallmatrix}\right), (6.2)

where zz\in\mathbb{C}, ξ\xi\in\mathbb{R}, and the differentiation is taken with respect to ξ\xi\in\mathbb{R}. The Weyl-Titchmarsh transform for 𝒟Q\mathcal{D}_{\mathcal{H}_{Q}} is defined by

𝒟Q:Y1πM~(ξ,λ)Q(ξ)Y(ξ)𝑑ξ\mathcal{F}_{\mathcal{D}_{\mathcal{H}_{Q}}}:Y\mapsto\frac{1}{\sqrt{\pi}}\int_{\mathbb{R}}\widetilde{M}(\xi,\lambda)^{*}\mathcal{H}_{Q}(\xi)Y(\xi)\,d\xi

on a dense subset of L2(Q)L^{2}(\mathcal{H}_{Q}) of smooth compactly supported functions. This operator is unitary from L2(Q)L^{2}(\mathcal{H}_{Q}) onto the space L2(ρ)L^{2}(\rho) defined in the same way as at the beginning of Section 3. Specifically, we let m±m_{\pm} be the half-line Weyl functions of Q\mathcal{H}_{Q} and define ρ\rho as the representing measure for the matrix-valued Herglotz function mm in (3.8). It was proved in Theorem 3.21 in [21] that 𝒟Q𝒟Q𝒟Q1\mathcal{F}_{\mathcal{D}_{\mathcal{H}_{Q}}}\mathcal{D}_{\mathcal{H}_{Q}}\mathcal{F}_{\mathcal{D}_{\mathcal{H}_{Q}}}^{-1} coincides with the operator of multiplication by the independent variable in L2(ρ)L^{2}(\rho). We also have

𝒟Q(VX)=𝒟QX,XL2(,2).\mathcal{F}_{\mathcal{D}_{\mathcal{H}_{Q}}}(VX)=\mathcal{F}_{\mathcal{D}_{{Q}}}X,\qquad X\in L^{2}(\mathbb{R},\mathbb{C}^{2}).

Thus, we only need to check that the Weyl functions m±m_{\pm} used in Section 3 coincide with the half-line Weyl functions of the Hamiltonian Q\mathcal{H}_{Q}. For the +\mathbb{R}_{+}-Weyl functions this follows from Lemma 6.1 below. Comparing the formulas for A+A^{+}, AA^{-} in the beginning of Section 3, we see that the Weyl function mm_{-} for 𝒟Q\mathcal{D}_{Q} corresponds to the Weyl function m+m_{+} for 𝒟Q~\mathcal{D}_{\widetilde{Q}} where Q~(ξ)=σ3Q(ξ)σ3\widetilde{Q}(\xi)=\sigma_{3}Q(-\xi)\sigma_{3}. Similarly, in the setting of canonical Hamiltonian systems, the Weyl function mm_{-} for 𝒟Q\mathcal{D}_{\mathcal{H}_{Q}} coincides with the Weyl function m+m_{+} of 𝒟~Q\mathcal{D}_{\widetilde{\mathcal{H}}_{Q}}, ~Q(ξ)=σ3(ξ)σ3\widetilde{\mathcal{H}}_{Q}(\xi)=\sigma_{3}\mathcal{H}(-\xi)\sigma_{3}. Therefore, the statement for AA^{-} follows from Lemma 6.1 below and from the relation ~Q=σ3Qσ3=(σ3NQσ3)(σ3NQσ3)=Q~\widetilde{\mathcal{H}}_{Q}=\sigma_{3}\mathcal{H}_{Q}\sigma_{3}=(\sigma_{3}N_{Q}^{*}\sigma_{3})(\sigma_{3}N_{Q}\sigma_{3})=\mathcal{H}_{\widetilde{Q}}.

Lemma 6.1.

Let qL2(+)q\in L^{2}(\mathbb{R}_{+}). Define

Q(ξ)=(Imq(ξ)Req(ξ)Req(ξ)Imq(ξ)),A(ξ)=q(ξ/2)¯/2,ξ+.Q(\xi)=\begin{pmatrix}-\operatorname{Im}q(\xi)&-\operatorname{Re}q(\xi)\\ -\operatorname{Re}q(\xi)&\operatorname{Im}q(\xi)\end{pmatrix},\qquad A(\xi)=-\overline{q(\xi/2)}/2,\qquad\xi\in\mathbb{R}_{+}.

Let NQN_{Q} be defined by JNQ(ξ,λ)+Q(ξ)NQ(ξ,λ)=λNQ(ξ,λ)JN^{\prime}_{Q}(\xi,\lambda)+Q(\xi)N_{Q}(\xi,\lambda)=\lambda N_{Q}(\xi,\lambda), NQ(0,λ)=(1001)N_{Q}(0,\lambda)=\left(\begin{smallmatrix}1&0\\ 0&1\\ \end{smallmatrix}\right). Consider the Hamiltonian Q=NQ(ξ,0)NQ(ξ,0)\mathcal{H}_{Q}=N_{Q}^{*}(\xi,0)N_{Q}(\xi,0) on +\mathbb{R}_{+} and let M~=(M~11M~12M~21M~22)\widetilde{M}=\left(\begin{smallmatrix}\widetilde{M}_{11}&\widetilde{M}_{12}\\ \widetilde{M}_{21}&\widetilde{M}_{22}\end{smallmatrix}\right) be defined by JM~(ξ,z)=zQ(ξ)M~(ξ,z)J\widetilde{M}^{\prime}(\xi,z)=z\mathcal{H}_{Q}(\xi)\widetilde{M}(\xi,z), M~(0,z)=(1001)\widetilde{M}(0,z)=\left(\begin{smallmatrix}1&0\\ 0&1\\ \end{smallmatrix}\right). Let, finally, PP, PP_{*}, P^\widehat{P}, P^\widehat{P}_{*} be the solutions to Krein systems (2.11), (2.12) for the coefficient AA on +\mathbb{R}_{+}. Then,

limξ+M~22(ξ,z)M~21(ξ,z)=limξ+(NQ)22(ξ,z)(NQ)21(ξ,z)=limξ+iP^(ξ,z)P(ξ,z),z+.\lim_{\xi\to+\infty}\frac{\widetilde{M}_{22}(\xi,z)}{\widetilde{M}_{21}(\xi,z)}=\lim_{\xi\to+\infty}\frac{(N_{Q})_{22}(\xi,z)}{(N_{Q})_{21}(\xi,z)}=\lim_{\xi\to+\infty}i\frac{\widehat{P}_{*}(\xi,z)}{P_{*}(\xi,z)},\qquad z\in\mathbb{C}_{+}. (6.3)

In other words, the function m+m_{+} in (3.7) is the half-line Weyl function for the operators 𝒟Q\mathcal{D}_{\mathcal{H}_{Q}}, 𝒟Q\mathcal{D}_{Q}.

Proof.  The formula

limξ+M~22(ξ,z)M~21(ξ,z)=limξ+(NQ)22(ξ,z)(NQ)21(ξ,z)\lim_{\xi\to+\infty}\frac{\widetilde{M}_{22}(\xi,z)}{\widetilde{M}_{21}(\xi,z)}=\lim_{\xi\to+\infty}\frac{(N_{Q})_{22}(\xi,z)}{(N_{Q})_{21}(\xi,z)}

for 𝒟Q\mathcal{D}_{Q} and 𝒟Q\mathcal{D}_{\mathcal{H}_{Q}} is well-known and can be derived from the analysis of Weyl circles by using identity NQ(ξ,λ)=NQ(ξ,0)M~(ξ,λ)N_{Q}(\xi,\lambda)=N_{Q}(\xi,0)\widetilde{M}(\xi,\lambda) and the invariance of Weyl circles under transforms generated by JJ-unitary matrices (in our setting, the JJ-unitary matrix is NQ(ξ,0)N_{Q}(\xi,0): we have NQ(ξ,0)JNQ(ξ,0)=JN_{Q}^{*}(\xi,0)JN_{Q}(\xi,0)=J on \mathbb{R}). See, e.g., [4] or Section 8 in [22] for more details on Weyl circles for canonical Hamiltonian systems. Thus, we focus on the second identity in (6.3) and define

X(ξ,z)=eiξz(P(2ξ,z)+P(2ξ,z)2P^(2ξ,z)P^(2ξ,z)2iP(2ξ,z)P(2ξ,z)2iP^(2ξ,z)+P^(2ξ,z)2),ξ,z.X(\xi,z)=e^{-i\xi z}\begin{pmatrix}\displaystyle\frac{P(2\xi,z)+P_{*}(2\xi,z)}{2}&\displaystyle\frac{\widehat{P}(2\xi,z)-\widehat{P}_{*}(2\xi,z)}{2i}\\ \displaystyle\frac{P_{*}(2\xi,z)-P(2\xi,z)}{2i}&\displaystyle\frac{\widehat{P}(2\xi,z)+\widehat{P}_{*}(2\xi,z)}{2}\end{pmatrix},\qquad\xi\in\mathbb{R},\quad z\in\mathbb{C}\,.

Differentiating, one obtains JX+QX=zXJX^{\prime}+QX=zX, X(0,z)=(1001)X(0,z)=\left(\begin{smallmatrix}1&0\\ 0&1\\ \end{smallmatrix}\right). It follows that X(ξ,z)=NQ(ξ,z)X(\xi,z)=N_{Q}(\xi,z). In particular, we have

(NQ)22=eiξzP^(2ξ,z)+P^(2ξ,z)2,(NQ)21=eiξzP(2ξ,z)P(2ξ,z)2i.(N_{Q})_{22}=e^{-i\xi z}\frac{\widehat{P}(2\xi,z)+\widehat{P}_{*}(2\xi,z)}{2},\qquad(N_{Q})_{21}=e^{-i\xi z}\frac{P_{*}(2\xi,z)-P(2\xi,z)}{2i}.

Since P(ξ,z)0P(\xi,z)\to 0, P(ξ,z)Π(z)0P_{*}(\xi,z)\to\Pi(z)\neq 0 as ξ+\xi\to+\infty (see Theorem 12.1 in [12]), and analogous relations hold for P^\widehat{P} and P^\widehat{P}_{*}, we have

limξ+(NQ)22(ξ,z)(NQ)21(ξ,z)=limξ+iP^(ξ,z)P(ξ,z),z+.\lim_{\xi\to+\infty}\frac{(N_{Q})_{22}(\xi,z)}{(N_{Q})_{21}(\xi,z)}=\lim_{\xi\to+\infty}i\frac{\widehat{P}_{*}(\xi,z)}{P_{*}(\xi,z)},\qquad z\in\mathbb{C}_{+}.

The lemma is proved. ∎


3. Lemma 6.1 and some known results for canonical systems can be used to show that weak convergence of potentials of the Dirac operator implies convergence of the corresponding Weyl functions.

Lemma 6.2.

Suppose {q}>0\{q_{\ell}\}_{\ell>0} is a bounded sequence in L2(+)L^{2}(\mathbb{R}_{+}) which converges to zero weakly. Let QQ_{\ell} be the associated matrix-functions defined as in Lemma 6.1. Then, the sequence of corresponding Weyl functions {m,+}\{m_{\ell,+}\} converges to ii locally uniformly in +\mathbb{C}_{+} when +\ell\to+\infty.

Proof.  For >0\ell>0, denote by Q\mathcal{H}_{Q_{\ell}} the Hamiltonian generated by QQ_{\ell} as in Lemma 6.1. Then, m,+m_{\ell,+} is the Weyl function for the half-line operators 𝒟Q\mathcal{D}_{\mathcal{H}_{Q_{\ell}}} and 𝒟Q\mathcal{D}_{Q_{\ell}}. Since sup>0qL2(+)<\sup_{\ell>0}\|q_{\ell}\|_{L^{2}(\mathbb{R}_{+})}<\infty and qq_{\ell} converge to zero weakly in L2(+)L^{2}(\mathbb{R}_{+}) as +\ell\to+\infty, the Hamiltonians Q\mathcal{H}_{Q_{\ell}} tend to the identity matrix 0=(1001)\mathcal{H}_{0}=\left(\begin{smallmatrix}1&0\\ 0&1\\ \end{smallmatrix}\right) uniformly on compact subsets on +\mathbb{R}_{+}. Then, their Weyl functions m+,m_{+,\ell} tend to the Weyl function m+=im_{+}=i of the Hamiltonian 0\mathcal{H}_{0} locally uniformly in +\mathbb{C}_{+} by Theorem 5.75.7 (b)(b) in [21]. ∎

References

  • [1] R. Bessonov and S. Denisov. Szegő condition, scattering, and vibration of Krein strings. preprint https://arxiv.org/abs/2203.07132.
  • [2] R. Bessonov and S. Denisov. A spectral Szegő theorem on the real line. Adv. Math., 359:106851, 41, 2020.
  • [3] R. Bessonov and S. Denisov. De Branges canonical systems with finite logarithmic integral. Anal. PDE, 14(5):1509–1556, 2021.
  • [4] R. Bessonov, M. Lukic, and P. Yuditskii. Reflectionless canonical systems, I. Arov gauge and right limits. preprint arXiv:2011.05261, 2014.
  • [5] D. G. Bhimani and R. Carles. Norm inflation for nonlinear Schrödinger equations in Fourier-Lebesgue and modulation spaces of negative regularity. J. Fourier Anal. Appl., 26(6):Paper No.78, 34, 2020.
  • [6] R. Carles and T. Kappeler. Norm-inflation with infinite loss of regularity for periodic NLS equations in negative Sobolev spaces. Bull. Soc. Math. France, 145(4):623–642, 2017.
  • [7] M. Christ, J. Colliander, and T. Tao. Asymptotics, frequency modulation, and low regularity ill-posedness for canonical defocusing equations. Amer. J. Math., 125(6):1235–1293, 2003.
  • [8] M. Christ, J. Colliander, and T. Tao. Ill-posedness for nonlinear Schrödinger and wave equations. preprint arXiv:math/0311048, 2003.
  • [9] P. Deift and X. Zhou. Long-time asymptotics for solutions of the NLS equation with initial data in a weighted Sobolev space. volume 56, pages 1029–1077. 2003. Dedicated to the memory of Jürgen K. Moser.
  • [10] P. A. Deift and X. Zhou. Long-time asymptotics for integrable systems. Higher order theory. Comm. Math. Phys., 165(1):175–191, 1994.
  • [11] S. A. Denisov. On the existence of wave operators for some Dirac operators with square summable potential. Geom. Funct. Anal., 14(3):529–534, 2004.
  • [12] S. A. Denisov. Continuous analogs of polynomials orthogonal on the unit circle and Kreĭn systems. IMRS Int. Math. Res. Surv., pages 1–148, 2006, Art. ID 54517.
  • [13] L. Faddeev and L. Takhtajan. Hamiltonian methods in the theory of solitons. Classics in Mathematics. Springer, Berlin, english edition, 2007. Translated from the 1986 Russian original by Alexey G. Reyman.
  • [14] C. Kenig, G. Ponce, and L. Vega. On the ill-posedness of some canonical dispersive equations. Duke Math. J., 106(3):617–633, 2001.
  • [15] S. Khrushchev. Schur’s algorithm, orthogonal polynomials, and convergence of Wall’s continued fractions in L2(T)L^{2}(T). Journal of Approximation Theory, 108(2):161–248, 2001.
  • [16] R. Killip, M. Vişan, and X. Zhang. Low regularity conservation laws for integrable PDE. Geom. Funct. Anal., 28(4):1062–1090, 2018.
  • [17] N. Kishimoto. A remark on norm inflation for nonlinear Schrödinger equations. Commun. Pure Appl. Anal., 18(3):1375–1402, 2019.
  • [18] H. Koch and D. Tataru. Conserved energies for the cubic nonlinear Schrödinger equation in one dimension. Duke Math. J., 167(17):3207–3313, 2018.
  • [19] M. G. Kreĭn. Continuous analogues of propositions on polynomials orthogonal on the unit circle. Dokl. Akad. Nauk SSSR (N.S.), 105:637–640, 1955.
  • [20] T. Oh and C. Sulem. On the one-dimensional cubic nonlinear Schrödinger equation below L2L^{2}. Kyoto J. Math., 52(1):99–115, 2012.
  • [21] C. Remling. Spectral Theory of Canonical Systems. De Gruyter Studies in Mathematics Series. Walter De Gruyter GmbH, 2018.
  • [22] R. Romanov. Canonical systems and de Branges spaces. preprint arXiv:1408.6022, 2014.
  • [23] T. Tao. Nonlinear dispersive equations, volume 106 of CBMS Regional Conference Series in Mathematics. Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 2006. Local and global analysis.
  • [24] T. Tao and Ch. Thiele. Nonlinear Fourier analysis. preprint arXiv:1201.5129, 2012.
  • [25] Y. Tsutsumi. L2L^{2}-solutions for nonlinear Schrödinger equations and nonlinear groups. Funkcial. Ekvac., 30(1):115–125, 1987.
  • [26] V. E. Zakharov and S. V. Manakov. Asymptotic behavior of non-linear wave systems integrated by the inverse scattering method. Z. Èksper. Teoret. Fiz., 71(1):203–215, 1976.
  • [27] V. E. Zakharov and A. B. Shabat. Exact theory of two-dimensional self-focusing and one-dimensional self-modulation of waves in nonlinear media. Ž. Èksper. Teoret. Fiz., 61(1):118–134, 1971.