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

The Steklov problem and Remainder Estimates for Krein Systems generated by a Muckenhoupt weight

Michel Alexis Michel Alexis: [email protected] McMaster University Department of Mathematics & Statistics 1280 Main St West, Hamilton, ON, L8S 4L8, Canada
Abstract.

We show that solutions to Krein systems, the continuous frequency analogue of orthogonal polynomials on the unit circle, generated by an A2()A_{2}(\mathbb{R}) weight ww satisfying w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}), are uniformly bounded in Llocp(w,)L^{p}_{\mathrm{loc}}(w,\mathbb{R}) for pp sufficiently close to 22. This provides a positive answer to the Steklov problem for Krein systems. Furthermore, we define a “remainder” which measures the difference between the solution to a Krein system and a polynomial-like approximant, and we estimate these remainders in Lwp()L^{p}_{w}(\mathbb{R}) for wA2()w\in A_{2}(\mathbb{R}) satisfying some additional conditions. Such polynomial-like approximants, and hence remainder estimates, seem unique to Krein systems, with no analogue for orthogonal polynomials on the unit circle.

I am thankful to Sergey Denisov for helpful discussions and advising on this topic.

1. Introduction

Let dσ=defw(λ)dλ2πd\sigma\stackrel{{\scriptstyle\rm def}}{{=}}w(\lambda)\frac{d\lambda}{2\pi} be a measure on \mathbb{R} with weight w0w\geq 0 satisfying w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}). Then one can define a family of “orthonormal continuous polynomials” {P(r,λ;σ)}r0\{P(r,\lambda;\sigma)\}_{r\geq 0} with the following properties.

  1. (i)

    P(r,λ;σ)P(r,\lambda;\sigma) is a “continuous polynomial” in exp(iλ)\mathrm{exp}\left(i\lambda\right) of “degree” rr, where exp(x)=defex\mathrm{exp}\left(x\right)\stackrel{{\scriptstyle\rm def}}{{=}}e^{x}: for each r0r\geq 0 there exists B(r,)L2([0,r])B(r,\cdot)\in L^{2}([0,r]) so that

    P(r,λ;σ)=exp(iλr)+0rB(r,s)exp(iλs)𝑑s.P(r,\lambda;\sigma)=\mathrm{exp}\left(i\lambda r\right)+\int\limits_{0}^{r}B(r,s)\mathrm{exp}\left(i\lambda s\right)\,ds\,.
  2. (ii)

    {P(r,λ;σ)}r0\{P(r,\lambda;\sigma)\}_{r\geq 0} is an orthonormal system with respect to σ\sigma, i.e. for all f,gL2(+)f,g\in L^{2}(\mathbb{R}^{+}) we have equality of the inner-products

    f(s),g(s)(ds,+)\displaystyle\langle f(s),g(s)\rangle_{(ds\,,\,\mathbb{R}_{+})} =0f(s)P(s,λ;σ)𝑑s,0g(s)P(s,λ;σ)𝑑s(dσ(λ),).\displaystyle=\langle\int\limits_{0}^{\infty}f(s)P(s,\lambda;\sigma)\,ds,\int\limits_{0}^{\infty}g(s)P(s,\lambda;\sigma)\,ds\rangle_{(d\sigma(\lambda),\mathbb{R})}\,. (1)

    Equivalently, the generalized Fourier transform 𝒪\mathcal{O} is an isometry from L2(+)L^{2}(\mathbb{R}^{+}) to Lσ2()L^{2}_{\sigma}(\mathbb{R}), where

    𝒪f(λ)=def0f(s)P(s,λ;σ)𝑑s.\mathcal{O}f(\lambda)\stackrel{{\scriptstyle\rm def}}{{=}}\int\limits_{0}^{\infty}f(s)P(s,\lambda;\sigma)\,ds\,.
  3. (iii)

    P(r,λ;σ)P(r,\lambda;\sigma) is the solution to the Krein differential system (20), whose coefficients A(r)A(r) uniquely determine P(r,λ;σ)P(r,\lambda;\sigma). As such P(r,λ;σ)P(r,\lambda;\sigma) is referred to as “the solution to the Krein system.”

For instance, when dσ=dλ2πd\sigma=\frac{d\lambda}{2\pi}, then P(r,λ;σ)=exp(iλr)P(r,\lambda;\sigma)=\mathrm{exp}\left(i\lambda r\right), so {P(r,λ;σ)}r0\{P(r,\lambda;\sigma)\}_{r\geq 0} is simply the Fourier orthonormal system {exp(iλr)}r0\{\mathrm{exp}\left(i\lambda r\right)\}_{r\geq 0}. See e.g. references [8, 3] for more information on Krein systems.

Notation.  When the measure σ\sigma is clear from context, we simply write P(r,λ)P(r,\lambda). And given a weight ww we write P(r,λ;w)=defP(r,λ;wdλ2π)P(r,\lambda;w)\stackrel{{\scriptstyle\rm def}}{{=}}P\left(r,\lambda;w\frac{d\lambda}{2\pi}\right).

First introduced in 1954 by Krein in [8], solutions P(r,λ)P(r,\lambda) to Krein systems are the continuous frequency analogue of Orthogonal Polynomials on the Unit Circle (OPUC): given a finite measure μ\mu on the unit circle 𝕋\mathbb{T}\subset\mathbb{C} with infinite support, the OPUC consist of an orthonormal sequence {φn}n0\{\varphi_{n}\}_{n\geq 0} in Lμ2(𝕋)L^{2}_{\mu}(\mathbb{T}), where each φn\varphi_{n} is a degree nn polynomial over \mathbb{C} with positive leading coefficient. One can obtain this sequence by, e.g., applying the Gram-Schmidt algorithm to the polynomials {1,z,z2,}\{1,z,z^{2},\ldots\}. See [12] for a robust reference on OPUC.

Consider the following problem for Krein systems.

Problem 1.1 (The Steklov problem for Krein systems).

If dσ=w(λ)2πdλd\sigma=\frac{w(\lambda)}{2\pi}\,d\lambda, what conditions on ww guarantee there exists p>2p>2 such that {P(r,λ;σ)}r0\{P(r,\lambda;\sigma)\}_{r\geq 0} is bounded in Llocp(σ,)L^{p}_{loc}(\sigma,\mathbb{R})? More precisely, when does there exist p>2p>2 such that for each compact Δ\Delta\subset\mathbb{R},

supr0P(r,;σ)Lwp(Δ)<?\sup\limits_{r\geq 0}\|P(r,\cdot;\sigma)\|_{L^{p}_{w}(\Delta)}<\infty\,? (2)

The Steklov problem and estimates like (2) are related to the following problem: for which measures σ\sigma is the maximal function

Mσ(λ)=defsupr0|P(r,λ;σ)|M_{\sigma}(\lambda)\stackrel{{\scriptstyle\rm def}}{{=}}\sup\limits_{r\geq 0}|P(r,\lambda;\sigma)|

finite at almost every λ\lambda\in\mathbb{R}? Known as nonlinear Carleson’s Theorem, this was recently proved in [11] for a conjectured class of measures.

The Steklov problem was originally posed for orthogonal polynomials: given the OPUC {φn}n0\{\varphi_{n}\}_{n\geq 0} generated by the measure dμ=wdθ2πd\mu=w\frac{d\theta}{2\pi} on the unit circle 𝕋\mathbb{T}, what conditions on ww guarantee there exists p>2p>2 for which

supn0φnLwp(𝕋)<?\sup\limits_{n\geq 0}\|\varphi_{n}\|_{L^{p}_{w}(\mathbb{T})}<\infty? (3)

Nazarov first showed that when w,w1L(𝕋)w,w^{-1}\in L^{\infty}(\mathbb{T}), then (3) holds for some p>2p>2 ([4]). Then in [5], Denisov-Rush generalized Nazarov’s result to the case when w,w1BMO(𝕋)w,w^{-1}\in\mathrm{BMO}(\mathbb{T}). And most recently in [1], Alexis-Aptekarev-Denisov further generalized these results to the case when wA2(𝕋)w\in A_{2}(\mathbb{T}). Since Krein systems are the continuous analogue of OPUC, we adapt the most recent successes in [1] to solutions of Krein systems generated by a weight wA2()w\in A_{2}(\mathbb{R}) with A2()A_{2}(\mathbb{R}) as defined below (see [13, p.194]).

Definition. Let p(1,)p\in(1,\infty). The weight wAp()w\in A_{p}(\mathbb{R}) if

[w]Ap()=defsupIwIw11pIp1<,[w]_{A_{p}(\mathbb{R})}\stackrel{{\scriptstyle\rm def}}{{=}}\sup_{I}\,\left\langle w\right\rangle_{I}\,\left\langle w^{\frac{1}{1-p}}\right\rangle_{I}^{{p-1}}<\infty,\,

where II ranges over the finite intervals in \mathbb{R}.

Note (2) will immediately follow if we can show there exists δ=δ(w)>0\delta=\delta(w)>0 such that if p[2,2+δ)p\in[2,2+\delta), then

supr0P(r,λ;w)exp(iλr)Lwp()<.\sup\limits_{r\geq 0}\|P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right)\|_{L^{p}_{w}(\mathbb{R})}<\infty\,. (4)

Whence the first main theorem of this paper below, for which we recall Lp1()+Lp2()L1()+L2()L^{p_{1}}(\mathbb{R})+L^{p_{2}}(\mathbb{R})\subset L^{1}(\mathbb{R})+L^{2}(\mathbb{R}) when p1,p2[1,2]p_{1},p_{2}\in[1,2].

Theorem 1.2.

Suppose that w1=u1+u2w-1=u_{1}+u_{2} where u1Lp1(),u2Lp2()u_{1}\in L^{p_{1}}(\mathbb{R}),u_{2}\in L^{p_{2}}(\mathbb{R}) with 1p1p221\leq p_{1}\leq p_{2}\leq 2. If [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma, then

  1. (a)

    there exists ϵ(γ)(0,12)\epsilon(\gamma)\in(0,\frac{1}{2}) with limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2} such that for any p[p2,)p\in[p_{2},\infty) satisfying |121p|<ϵ(γ)\left|\frac{1}{2}-\frac{1}{p}\right|<\epsilon(\gamma), we have

    supr0P(r,λ;w)exp(iλr)Lwp()<.\sup\limits_{r\geq 0}\|P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right)\|_{L^{p}_{w}(\mathbb{R})}<\infty\,.
  2. (b)

    for any p[p2,)p\in[p_{2},\infty) there exists τ0(p)(0,1)\tau_{0}(p)\in(0,1) such that whenever τ=def[w]A2()1τ0(p)\tau\stackrel{{\scriptstyle\rm def}}{{=}}[w]_{A_{2}(\mathbb{R})}-1\leq\tau_{0}(p), we have

    supr0P(r,λ;w)exp(iλr)Lwp()pτpp22p(τp2p12u1Lp1()p1+u2Lp2()p2)1p.\sup\limits_{r\geq 0}\|P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right)\|_{L^{p}_{w}(\mathbb{R})}\lesssim_{p}\tau^{\frac{p-p_{2}}{2p}}(\tau^{\frac{p_{2}-p_{1}}{2}}\|u_{1}\|_{L^{p_{1}}(\mathbb{R})}^{p_{1}}+\|u_{2}\|_{L^{p_{2}}(\mathbb{R})}^{p_{2}})^{\frac{1}{p}}\,. (5)

Solutions to the Krein system generate generalized eigenfunctions for a unique Dirac operator 𝒟\cal{D} with spectral measure 2dσ2\,d\sigma. Furthermore, if A(r)A(r), the coefficient of the Krein system (20), is real-valued and in L2(+)L^{2}(\mathbb{R}^{+}), then when written in its canonical form, 𝒟2\cal{D}^{2} is a diagonal operator whose entries are Schrödinger operators. Thus, Theorem 1.2 allows one to obtain Lwp()L^{p}_{w}(\mathbb{R})-information on generalized eigenfunctions for the Dirac and Schrödinger operators. See e.g. [3, Sections 14-16] for more details.

We note the following corollaries of Theorem 1.2.

Corollary 1.3.

If w1Lp1()+Lp2()w-1\in L^{p_{1}}(\mathbb{R})+L^{p_{2}}(\mathbb{R}) for 1p1p221\leq p_{1}\leq p_{2}\leq 2, then for any p(p2,)p\in(p_{2},\infty), we have

lim[w]A2()1supr0P(r,λ;w)exp(iλr)Lwp()=0\lim\limits_{[w]_{A_{2}(\mathbb{R})}\to 1}\sup\limits_{r\geq 0}\|P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right)\|_{L^{p}_{w}(\mathbb{R})}=0

where the limit is taken among weights satisfying w1Lp1()+Lp2()C\|w-1\|_{L^{p_{1}}(\mathbb{R})+L^{p_{2}}(\mathbb{R})}\leq C, with C0C\geq 0 an arbitrary absolute constant.

Proof.

Take τ0\tau\to 0 in (5). ∎

Corollary 1.3 states that as ww becomes flatter, P(r,λ)P(r,\lambda) tends to standard exponential, demonstrating some continuity of solutions to Krein systems with respect to the measure.

Corollary 1.4.

Suppose w1Lp1()+Lp2()w-1\in L^{p_{1}}(\mathbb{R})+L^{p_{2}}(\mathbb{R}) with 1p1p2<21\leq p_{1}\leq p_{2}<2. If [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma, then there exists δ(γ,p2)>0\delta(\gamma,p_{2})>0 such that

P(r,λ;w)exp(iλr)Lwp()P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right)\in L^{p}_{w}(\mathbb{R})

for all r>0r>0, p[2δ,)p\in[2-\delta,\infty).

Proof.

Apply Theorem 1.2 (a). ∎

Corollary 1.4 is non-trivial, since as we discuss in Section 3, a priori all we can say about P(r,λ;w)exp(iλr)P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right) is that it belongs in Lwp()L^{p}_{w}(\mathbb{R}) for p[2,)p\in[2,\infty).

The requirement that pp2p\geq p_{2} in Theorem 1.2 is sharp in the sense of Proposition 1.5 below, which morally says that the decay of λP(r,λ;w)exp(iλr)\lambda\mapsto P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right) cannot exceed the decay of λw(λ)1\lambda\mapsto w(\lambda)-1.

Proposition 1.5.

If p2(1,2]p_{2}\in(1,2], then for every δ>0\delta>0, there exists a weight ww such that w1Lp2()w-1\in L^{p_{2}}(\mathbb{R}), [w]A2()1+δ[w]_{A_{2}(\mathbb{R})}\leq 1+\delta and

supr0P(r,λ;w)exp(iλr)Lwp()=\sup\limits_{r\geq 0}\|P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right)\|_{L^{p}_{w}(\mathbb{R})}=\infty

for every p<p2p<p_{2}.

In the same vein as Theorem 1.2 (a), one may also estimate a “remainder” term, which we define precisely in Section 6. Let λ=def(1+λ2)1/2\langle\lambda\rangle\stackrel{{\scriptstyle\rm def}}{{=}}(1+\lambda^{2})^{1/2}. If λk(w1)L1()\langle\lambda\rangle^{k}(w-1)\in L^{1}(\mathbb{R}) then the remainder function Rk,r(λ)R_{k,r}(\lambda) exists and satisfies

Rk,r(λ)=λk(P(r,λ)exp(iλr))l=0k1λlakl,r(λ),R_{k,r}(\lambda)=\lambda^{k}(P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right))-\sum\limits_{l=0}^{k-1}\lambda^{l}a_{k-l,r}(\lambda)\,,

where al,r(λ)L()a_{l,r}(\lambda)\in L^{\infty}(\mathbb{R}), l=1,,kl=1,\dots,k. Note that Rk,rR_{k,r} measures how much λk(P(r,λ)exp(iλr))\lambda^{k}(P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)) deviates from a polynomial-like term of “degree” k1k-1. In Theorem 1.6 below, we estimate R1,rLwp()\|R_{1,r}\|_{L^{p}_{w}(\mathbb{R})}, which quantifies the decay of P(r,λ;w)exp(iλr)P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right), providing information with no analogue for OPUC. Furthermore, the assumptions of Theorem 1.6 below imply

a1,r(λ)=exp(iλr)α(r)+α2(r),a_{1,r}(\lambda)=\mathrm{exp}\left(i\lambda r\right)\alpha_{\infty}(r)+\alpha_{2}(r)\,,

where each αq(r)\alpha_{q}(r) does not depend on λ\lambda, αqLq(+)\alpha_{q}\in L^{q}(\mathbb{R}^{+}), and limrα(r)\lim\limits_{r\to\infty}\alpha_{\infty}(r) exists (see Lemma 6.2); as such, the estimate on R1,rLwp()\|R_{1,r}\|_{L^{p}_{w}(\mathbb{R})} below also qualifies the behavior of the “continuous polynomials” P(r,λ;w)P(r,\lambda;w) for rr large. We will need the following definitions for what follows.

  • Let H2(Ω)H^{2}(\Omega) denotes the Hardy space in the domain Ω\Omega\subset\mathbb{C}.

  • If BB is a Banach space with norm B\|\cdot\|_{B}, (μ,X)(\mu,X) is a measure space and p[1,]p\in[1,\infty], let

    Lμp(X;B)=def{f:XB|fLμp(X;B)=deffBLμp(X)<}.L^{p}_{\mu}(X;B)\stackrel{{\scriptstyle\rm def}}{{=}}\left\{f:X\to B~{}\biggr{|}~{}\|f\|_{L^{p}_{\mu}(X;B)}\stackrel{{\scriptstyle\rm def}}{{=}}\left\|\|f\|_{B}\right\|_{L^{p}_{\mu}(X)}<\infty\right\}\,.
  • Given two Banach spaces B1,B2B_{1},B_{2} with norms B1\|\cdot\|_{B_{1}} and B2\|\cdot\|_{B_{2}}, for each fB1+B2f\in B_{1}+B_{2} define

    fB1+B2=definff=f1+f2f1B1+f2B2.\|f\|_{B_{1}+B_{2}}\stackrel{{\scriptstyle\rm def}}{{=}}\inf\limits_{f=f_{1}+f_{2}}\|f_{1}\|_{B_{1}}+\|f_{2}\|_{B_{2}}\,.
Theorem 1.6.

Suppose that λq(w1)L1()\langle\lambda\rangle^{q}(w-1)\in L^{1}(\mathbb{R}) for some q>2q>2 and

exp(12πilog(w(t))tλ𝑑t)1=0h(x)exp(iλx)𝑑xH2(+).\mathrm{exp}\left(\frac{1}{2\pi i}\int\limits_{-\infty}^{\infty}\frac{\log(w(t))}{t-\lambda}\,dt\right)-1=\int\limits_{0}^{\infty}h(x)\mathrm{exp}\left(i\lambda x\right)\,dx\in H^{2}(\mathbb{C}^{+})\,. (6)

If [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma, then there exists ϵ(γ)>0\epsilon(\gamma)>0 such that whenever p(1,q]p\in(1,q] satisfies |1p12|<ϵ(γ)\left|\frac{1}{p}-\frac{1}{2}\right|<\epsilon(\gamma), then

R1,r(λ)Ldr2(+;Lwp())+Ldr(+;Lwp())<.\|R_{1,r}(\lambda)\|_{L^{2}_{dr}(\mathbb{R}^{+};L^{p}_{w}(\mathbb{R}))+L^{\infty}_{dr}(\mathbb{R}^{+};L^{p}_{w}(\mathbb{R}))}<\infty\,.

Compare Theorem 1.6 and Theorem 1.2 (a), the latter of which can be understood as stating that the Ldr(+;Lwp())L^{\infty}_{dr}(\mathbb{R}^{+};L^{p}_{w}(\mathbb{R}))-norm of 0th order remainder R0,rR_{0,r} is finite.

We can also estimate higher order remainders when the weight ww is very close to 11.

Theorem 1.7.

Suppose (w1)λkL1()(w-1)\langle\lambda\rangle^{k}\in L^{1}(\mathbb{R}) for some integer k0k\geq 0. If (w1)λkL()δ\|(w-1)\langle\lambda\rangle^{k}\|_{L^{\infty}(\mathbb{R})}\leq\delta, for some δ[0,1)\delta\in[0,1), then

  1. (a)

    there exists ϵ(δ)(0,12)\epsilon(\delta)\in(0,\frac{1}{2}), with limδ0ϵ(δ)=12\lim\limits_{\delta\to 0}\epsilon(\delta)=\frac{1}{2}, such that for all pp satisfying |121p|<ϵ(δ)\left|\frac{1}{2}-\frac{1}{p}\right|<\epsilon(\delta),

    supr>0Rk,r(λ)Lp()<.\sup\limits_{r>0}\|R_{k,r}(\lambda)\|_{L^{p}(\mathbb{R})}<\infty\,.
  2. (b)

    for any p(1,)p\in(1,\infty), there exists δ0(p)>0\delta_{0}(p)>0 such that for all δ(0,δ0(p))\delta\in(0,\delta_{0}(p)), we have

    supr>0Rk,r(λ)Lp()p,kδ11p(λk(w1)L1()1p+λk(w1)L1()1+1p).\sup\limits_{r>0}\|R_{k,r}(\lambda)\|_{L^{p}(\mathbb{R})}\lesssim_{p,k}\delta^{1-\frac{1}{p}}\left(\|\langle\lambda\rangle^{k}(w-1)\|_{L^{1}(\mathbb{R})}^{\frac{1}{p}}+\|\langle\lambda\rangle^{k}(w-1)\|_{L^{1}(\mathbb{R})}^{1+\frac{1}{p}}\right)\,.

In particular Theorem 1.7 (b) has the following corollary, which shows some continuity of Rk,rR_{k,r} with respect to the measure.

Corollary 1.8.

If (w1)λkL1()C\|(w-1)\langle\lambda\rangle^{k}\|_{L^{1}(\mathbb{R})}\leq C, where CC is any fixed constant and kk a nonnegative integer, then for any fixed 1<p<1<p<\infty we have

lim(w1)λkL()0supr>0Rk,r(λ)Lp()=0.\lim\limits_{\|(w-1)\langle\lambda\rangle^{k}\|_{L^{\infty}(\mathbb{R})}\to 0}\sup\limits_{r>0}\|R_{k,r}(\lambda)\|_{L^{p}(\mathbb{R})}=0\,.

While examples of weights to which we can apply Theorem 1.7 are easy to find, e.g. any weight satisfying |w1|δλk+2|w-1|\leq\frac{\delta}{\langle\lambda\rangle^{k+2}}, let’s mention some weights to which we can apply Theorems 1.2 and 1.6. Define

wβ(λ)=def{|λ|βwhen |λ|11when |λ|>1,w_{\beta}(\lambda)\stackrel{{\scriptstyle\rm def}}{{=}}\begin{cases}|\lambda|^{\beta}&\text{when }|\lambda|\leq 1\\ 1&\text{when }|\lambda|>1\end{cases}\,,

which is an A2()A_{2}(\mathbb{R}) weight when |β|<1|\beta|<1. Then finite products w(λ)=defj=1nwβj(λλj)w(\lambda)\stackrel{{\scriptstyle\rm def}}{{=}}\prod\limits_{j=1}^{n}w_{\beta_{j}}(\lambda-\lambda_{j}), where |βj|<1|\beta_{j}|<1 and λ1,,λn\lambda_{1},\ldots,\lambda_{n} are all distinct, satisfy the conditions of Theorems 1.2 and 1.6.

In practice, condition (6) is difficult to check, thus making it difficult to specify weights with infinitely many singularities to which Theorem 1.6 applies. However, if we require the singularities be well spaced out, one can generate other weights that satisfy the assumptions of Theorem 1.2. Consider for instance

w(λ)=defj=0wβj(λ2jϵj),w(\lambda)\stackrel{{\scriptstyle\rm def}}{{=}}\,\prod\limits_{j=0}^{\infty}w_{\beta_{j}}\left(\frac{\lambda-2^{j}}{\epsilon_{j}}\right)\,,

where supj|βj|<1\sup\limits_{j}|\beta_{j}|<1, and {ϵj}j0\{\epsilon_{j}\}_{j\geq 0} is a sequence of positive reals such that jϵj<\sum\limits_{j}\epsilon_{j}<\infty.

As far as methods are concerned, the proofs of Theorems 1.2 and 1.6 are based on [1]’s proof of a positive answer to the OPUC Steklov problem for A2(𝕋)A_{2}(\mathbb{T}) weights: this involves the theory of weighted operators, some basic spectral theory and complex interpolation. Meanwhile Theorem 1.7 is based on the original, elegantly simply method of Nazarov used to prove [4, Theorem 2.1], which involves just basic knowledge of the Hilbert transform.

1.1. Lingering questions: some potentially open problems

Regarding lingering questions, let us first comment on our methods. Essential to our proof of Theorem 1.2 is the orthogonality

P(r,λ;w)wRange𝒫[0,r],P(r,\lambda;w)\perp_{w}\mathrm{Range}\,\mathcal{P}_{[0,r]}\,,

where 𝒫[0,r]\mathcal{P}_{[0,r]} is Fourier projection onto the frequency band [0,r][0,r]. But using for instance Lemma 4.1, we end up showing

P(r,λ;w)wq(λ)Range𝒫[0,r],P(r,\lambda;w)\perp_{wq(\lambda)}\mathrm{Range}\,\mathcal{P}_{[0,r]}\,,

where q(λ)q(\lambda) is any nonnegative polynomial. Thus, as far as the algebra in the proof of Theorem 1.2 is concerned, one is tempted to consider the case w~=defwq(λ)A2()\widetilde{w}\stackrel{{\scriptstyle\rm def}}{{=}}wq(\lambda)\in A_{2}(\mathbb{R}) and prove another version of this result. But issues arise: for Krein systems and their solutions to be well-defined, the weight ww needs to be “centered” near 11, with e.g. w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}). This directly conflicts with w~A2()\widetilde{w}\in A_{2}(\mathbb{R}). Indeed, heuristically w~A2()\widetilde{w}\in A_{2}(\mathbb{R}) means w~\widetilde{w} is more or less constant, which would then mean ww decays to 0, which would contradict the required decay of w1w-1. A possible remedy here is to study the more general de Branges canonical systems, wherein the weight ww can deviate from 11 in more exotic ways. I have not investigated this potential framing of the problem and thus so far it remains an open question.

While Proposition 1.5 shows that the requirement that pp2p\geq p_{2} is sharp in Theorem 1.2, it is unclear to me whether the requirement that pqp\leq q is sharp in the statement of Theorem 1.6. Furthermore, while Proposition 1.5 had a nice interpretation — that P(r,λ;w)exp(iλr)P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right) cannot decay faster that w1w-1 — the requirement that pqp\leq q is more difficult to interpret. in elucidating this part of Theorem 1.6.

Finally, if wA2()w\in A_{2}(\mathbb{R}) and w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}), then Theorem 1.2 implies

supr>0P(r,λ)exp(iλr)Lwp()<\sup\limits_{r>0}\|P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)\|_{L^{p}_{w}(\mathbb{R})}<\infty\, (7)

for some p>2p>2. While this in turn implies

supr>0P(r,λ)Lw(λ)dλ1+λ2p()<\sup\limits_{r>0}\|P(r,\lambda)\|_{L^{p}_{w(\lambda)\frac{d\lambda}{1+\lambda^{2}}}(\mathbb{R})}<\infty (8)

for some p>2p>2, I feel the estimate (8) is more natural to try and prove directly. I wonder if there are other methods or assumptions which would more readily yield (8) without going through (7).

1.2. Organization of this Paper

This paper is organized as follows. First, we outline the basic theory of Krein systems in Section 2. In Section 3, we discuss a priori which Lwp()L^{p}_{w}(\mathbb{R}) spaces P(r,λ;w)exp(iλr)P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right) belongs to. In Section 4 we prove a useful orthogonality lemma. We spend Sections 5, 6 and 7 proving the main results of this paper, i.e. Theorem 1.2 and Proposition 1.5, Theorem 1.6, and Theorem 1.7. In Sections 8 and 9 we prove the technical Lemmas 5.3 and 5.4 essential in the proofs of the results in this paper. And finally, in Appendix A, we prove Proposition 5.5.

1.3. Notation

If BB is a Banach space, we denote its dual by BB^{*}, and the space of bounded operators on BB by ()\cal{L}(B).

If XdX\subset\mathbb{R}^{d} and dμ=w(x)dxd\mu=w(x)\,dx, we define Lwp(X)=defLμp(X)L^{p}_{w}(X)\stackrel{{\scriptstyle\rm def}}{{=}}L^{p}_{\mu}(X). If μ\mu is Lebesgue measure, we omit μ\mu, i.e. Lp(X)=defLdxp(X)L^{p}(X)\stackrel{{\scriptstyle\rm def}}{{=}}L^{p}_{dx}(X). If TT is a bounded linear operator between two Banach spaces Lμp(X)L^{p}_{\mu}(X), Lνq(Y)L^{q}_{\nu}(Y), we write its norm as Tp,q\|T\|_{p,q}.

If p[1,]p\in[1,\infty], the dual exponent is denoted by p=p/(p1)p^{\prime}=p/(p-1).

For fLμ1(X)f\in L^{1}_{\mu}(X), denote the average of ff over the measure space (μ,X)(\mu,X) by

fX,μ=def1μ(X)Xf𝑑μ.\langle f\rangle_{X,\mu}\stackrel{{\scriptstyle\rm def}}{{=}}\frac{1}{\mu(X)}\int_{X}fd\mu\,.

If XdX\subset\mathbb{R}^{d} and μ\mu is Lebesgue measure, we will simply write fX\langle f\rangle_{X}.

In this paper we work with spatial variable λ\lambda\in\mathbb{R} and frequency variable r+r\in\mathbb{R}^{+}. Thus for Banach spaces like Lσp()L^{p}_{\sigma}(\mathbb{R}) or Lwp()L^{p}_{w}(\mathbb{R}), the variable of integration is understood to be λ\lambda, i.e. f(r,λ)Lwp()=deff(r,λ)Lw(λ)dλp\|f(r,\lambda)\|_{L^{p}_{w}(\mathbb{R})}\stackrel{{\scriptstyle\rm def}}{{=}}\|f(r,\lambda)\|_{L^{p}_{w(\lambda)d\lambda}}. And given a measure, weight or function over \mathbb{R}, we write them with respect to variable λ\lambda, i.e. w=w(λ)w=w(\lambda).

Given a measure space (μ,X)(\mu,X), we denote the Lμ2(X)L^{2}_{\mu}(X) inner product by

f,g(μ,X)=defXf(x)g(x)¯𝑑μ(x).\langle f,g\rangle_{(\mu,X)}\stackrel{{\scriptstyle\rm def}}{{=}}\int\limits_{X}f(x)\overline{g(x)}\,d\mu(x)\,.

If it is clear from context that X=dX=\mathbb{R}^{d}, we will simply write f,gμ\langle f,g\rangle_{\mu}.

If exp(x)=defex\mathrm{exp}\left(x\right)\stackrel{{\scriptstyle\rm def}}{{=}}e^{x}, then the Fourier transform f^\hat{f}, or 𝒻\cal{F}f, of a function ff and its inverse transform fˇ\check{f}, or 1𝒻\cal{F}^{-1}f are given by

f^(ξ)=deff(x)exp(2πixξ)𝑑x,fˇ(x)=deff(ξ)exp(2πixξ)𝑑ξ.\widehat{f}(\xi)\stackrel{{\scriptstyle\rm def}}{{=}}\int\limits_{-\infty}^{\infty}f(x)\mathrm{exp}\left(-2\pi ix\xi\right)\,dx\,,\quad\widecheck{f}(x)\stackrel{{\scriptstyle\rm def}}{{=}}\int\limits_{-\infty}^{\infty}f(\xi)\mathrm{exp}\left(2\pi ix\xi\right)d\xi\,.

Given a set AXA\subseteq X where X=dX=\mathbb{R}^{d}, we will use notation AcA^{c} for its complement, i.e., Ac=X\AA^{c}=X\backslash A.

Let Cc(U)C_{c}^{\infty}(U) denote the space of smooth functions compactly supported in UU, an open subset of d\mathbb{R}^{d}.

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}. If the constant depends on a parameter α\alpha, we will write f1αf2f_{1}\lesssim_{\alpha}f_{2}. We define \gtrsim similarly and write f1f2f_{1}\sim f_{2} if f1f2f_{1}\lesssim f_{2} and f2f1f_{2}\lesssim f_{1} simultaneously.

If a constant CC depends on parameter α\alpha, we will express this by writing C(α)C(\alpha) or CαC_{\alpha}.


2. Krein Systems Basics

In this section we explain the basics of Krein systems following [3], listing most facts needed for this paper without proof; see [3] for an accessible in-depth survey on Krein systems.

Suppose σ\sigma is a Poisson-finite measure over \mathbb{R}, i.e.

dσ(λ)1+λ2<.\int\limits_{\mathbb{R}}\frac{d\sigma(\lambda)}{1+\lambda^{2}}<\infty\,.

Definition.  A function H:H:\mathbb{R}\to\mathbb{C} is Hermitian if H(x)=H(x)¯H(-x)=-\overline{H(x)}.

Definition.  A Hermitian function HLloc2()H\in L^{2}_{loc}(\mathbb{R}) is the accelerant associated to the measure σ\sigma if there exists a real constant β\beta such that for all xx\in\mathbb{R},

0x(xs)H(s)𝑑s=iβx+(1+iλx1+λ2exp(iλx))1λ2(dσ(λ)dλ2π).\int\limits_{0}^{x}(x-s)H(s)\,ds=i\beta x+\int\limits_{-\infty}^{\infty}\left(1+\frac{i\lambda x}{1+\lambda^{2}}-\mathrm{exp}\left(i\lambda x\right)\right)\frac{1}{\lambda^{2}}\,\left(d\sigma(\lambda)-\frac{d\lambda}{2\pi}\right)\,. (9)

Formally differentiating (9) twice yields “H(2πx)=(dσdλ2π)ˇ(x)H(2\pi x)=\widecheck{(d\sigma-\frac{d\lambda}{2\pi})}(x).” Thus intuitively the accelerant captures the moments of the signed measure dσdλ2πd\sigma-\frac{d\lambda}{2\pi}.

If an accelerant HH exists, then for each r>0r>0, define the operator r\mathcal{H}_{r} on L2([0,r])L^{2}([0,r]) by

rf(x)=def0rH(xy)f(y)𝑑y.\mathcal{H}_{r}f(x)\stackrel{{\scriptstyle\rm def}}{{=}}\int\limits_{0}^{r}H(x-y)f(y)\,dy\,. (10)

Of interest is when I+r>0I+\mathcal{H}_{r}>0, which occurs for a large class of measures.

Lemma 2.1.

Suppose dσ=wdλ2πd\sigma=w\frac{d\lambda}{2\pi}. If w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}), then

H(x)=def12π(w1ˇ)(x2π)=1(w(2π)1)(x)H(x)\stackrel{{\scriptstyle\rm def}}{{=}}\frac{1}{2\pi}(\widecheck{w-1})\left(\frac{x}{2\pi}\right)=\mathcal{F}^{-1}(w(2\pi\cdot)-1)(x)

is the accelerant associated to dσd\sigma, and I+r>0I+\mathcal{H}_{r}>0.

Proof.

Let aL1()+L2()a\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}) and first define H=def12πaˇ(x2π)C0()+L2()Lloc2()H\stackrel{{\scriptstyle\rm def}}{{=}}\frac{1}{2\pi}\widecheck{a}\left(\frac{x}{2\pi}\right)\in C_{0}(\mathbb{R})+L^{2}(\mathbb{R})\subset L^{2}_{loc}(\mathbb{R}) and

β=def12πλ1+λ2a(λ)𝑑λ.\beta\stackrel{{\scriptstyle\rm def}}{{=}}\frac{1}{2\pi}\int\limits_{\mathbb{R}}\frac{\lambda}{1+\lambda^{2}}a(\lambda)\,d\lambda\,. (11)

Suppose we can show that

0x(xs)H(s)𝑑s=iβx+12π(1+iλx1+λ2exp(iλx))1λ2a(λ)𝑑λ\int\limits_{0}^{x}(x-s)H(s)\,ds=i\beta x+\frac{1}{2\pi}\int\limits_{-\infty}^{\infty}\left(1+\frac{i\lambda x}{1+\lambda^{2}}-\mathrm{exp}\left(i\lambda x\right)\right)\frac{1}{\lambda^{2}}a(\lambda)\,d\lambda\, (12)

holds for all xx\in\mathbb{R}. Then setting a=w1a=w-1, we will show that H(x)=def12π(w1)ˇ(x2π)H(x)\stackrel{{\scriptstyle\rm def}}{{=}}\frac{1}{2\pi}\widecheck{(w-1)}\left(\frac{x}{2\pi}\right) is the accelerant associated to dσ=w(λ)2πdλd\sigma=\frac{w(\lambda)}{2\pi}d\lambda, i.e. we will show (12) holds.

We now note that I+r>0I+\mathcal{H}_{r}>0 : take fL2([0,r])f\in L^{2}([0,r]) and extend ff to a function over \mathbb{R} by defining f(x)=0f(x)=0 for x[0,r]x\notin[0,r]. Then Fourier inversion for distributions yields

(I+r)f,f(dx,)=f,f(dx,)+Hf,f(dx,)\displaystyle\langle(I+\mathcal{H}_{r})f,f\rangle_{(dx,\mathbb{R})}=\langle f,f\rangle_{(dx,\mathbb{R})}+\langle H\ast f,f\rangle_{(dx,\mathbb{R})} =f^,f^(dλ,)+(w(2π)1)f^,f^(dλ,)\displaystyle=\langle\hat{f},\hat{f}\rangle_{(d\lambda,\mathbb{R})}+\langle(w(2\pi\cdot)-1)\hat{f},\hat{f}\rangle_{(d\lambda,\mathbb{R})}
=f^Lw(2π)2()2\displaystyle=\|\hat{f}\|_{L^{2}_{w(2\pi\cdot)}(\mathbb{R})}^{2}
>0\displaystyle>0

whenever f0f\neq 0, i.e. I+r>0I+\mathcal{H}_{r}>0.

It remains to show (12) holds for H,β,aH,\beta,a as above. If we can show (12) holds for aL1()a\in L^{1}(\mathbb{R}) and for aL2()a\in L^{2}(\mathbb{R}), then by linearity (12) holds for all aL1()+L2()a\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}).

The case that aL1()a\in L^{1}(\mathbb{R}): First note that both sides of (12) are well-defined functions for aL1a\in L^{1}. Next note that the second derivative (with respect to variable xx) of the left side of (12) equals the second derivative of the right side of (12). Indeed, this follows from the dominated convergence theorem and that aL1()a\in L^{1}(\mathbb{R}) implies HC()H\in C(\mathbb{R}).

Thus the first derivatives of each side of (12) will be identical so long as we can verify they agree at a point, say x=0x=0. But β\beta is defined precisely to make this occur; one can check this using the dominated convergence theorem.

Thus to verify (12) holds for all xx, it suffices to verify each side of (12) agrees at a point, say x=0x=0. Direct computation shows each side of (12) is 0 at x=0x=0. This completes the proof of this case.

The case that aL2()a\in L^{2}(\mathbb{R}): First let anL1()L2()a_{n}\in L^{1}(\mathbb{R})\cap L^{2}(\mathbb{R}) approximate aa in L2()L^{2}(\mathbb{R}), e.g. define

an=defχ[n,n]a.a_{n}\stackrel{{\scriptstyle\rm def}}{{=}}\chi_{[-n,n]}a\,.

By the previous case, if Hn(x)=def12πanˇ(x2π)H_{n}(x)\stackrel{{\scriptstyle\rm def}}{{=}}\frac{1}{2\pi}\widecheck{a_{n}}\left(\frac{x}{2\pi}\right), and βn=12πλ1+λ2an(λ)𝑑λ\beta_{n}=\frac{1}{2\pi}\int\limits_{\mathbb{R}}\frac{\lambda}{1+\lambda^{2}}a_{n}(\lambda)\,d\lambda, then (12) holds for (Hn,βn,an)(H_{n},\beta_{n},a_{n}). We note that HnHH_{n}\to H in L2()L^{2}(\mathbb{R}) by Plancherel’s theorem; we also have βnβ\beta_{n}\to\beta for β\beta given by (11). Now consider (12) for (Hn,βn,an)(H_{n},\beta_{n},a_{n}) and take nn\to\infty to get it for (H,β,a)(H,\beta,a). This completes the proof. ∎

Assume HLloc2()H\in L^{2}_{loc}(\mathbb{R}) is an accelerant for which I+r>0I+\mathcal{H}_{r}>0 for each r>0r>0. Then the resolvent operator

𝒢r=defI(I+r)1\mathcal{G}_{r}\stackrel{{\scriptstyle\rm def}}{{=}}I-(I+\mathcal{H}_{r})^{-1} (13)

is an integral operator on L2([0,r])L^{2}([0,r]) with kernel Γr(s,t)Lds×dt2([0,r]2)\Gamma_{r}(s,t)\in L^{2}_{ds\times dt}([0,r]^{2}) possessing the following properties.

Properties of the Resolvent kernel Γ\Gamma

  1. (i)

    Resolvent Identity: from (13) we have

    Γr(t,s)+0rH(tu)Γr(u,s)𝑑u\displaystyle\Gamma_{r}(t,s)+\int\limits_{0}^{r}H(t-u)\Gamma_{r}(u,s)\,du =H(ts), 0s,tr.\displaystyle=H(t-s)\,,\;0\leq s,t\leq r\,. (14)
  2. (ii)

    Symmetries: Γ\Gamma has following symmetries:

    Γr(s,t)=Γr(t,s)¯,Γr(s,t)=Γr(rt,rs),0s,tr.\Gamma_{r}(s,t)=\overline{\Gamma_{r}(t,s)}\,,\quad\Gamma_{r}(s,t)=\Gamma_{r}(r-t,r-s)\,,\quad 0\leq s,t\leq r\,. (15)
  3. (iii)

    Regularity inherited from accelerant: if HCk()H\in C^{k}(\mathbb{R}) for some k0k\geq 0, then for each r>0r>0, Γr(,)Ck([0,r]2)\Gamma_{r}(\cdot,\cdot)\in C^{k}([0,r]^{2}). Furthermore, Γr(s,t)\Gamma_{r}(s,t) is continuously differentiable in rr and

    rΓr(s,t)=Γr(s,r)Γr(r,t),0s,tr.\partial_{r}\Gamma_{r}(s,t)=-\Gamma_{r}(s,r)\Gamma_{r}(r,t)\,,\quad 0\leq s,t\leq r\,. (16)
  4. (iv)

    Continuity in L2L^{2}: Define

    gr(s)=def{Γr(s,0) if 0sr0 if s>r.g_{r}(s)\stackrel{{\scriptstyle\rm def}}{{=}}\begin{cases}\Gamma_{r}(s,0)&\text{ if }0\leq s\leq r\\ 0&\text{ if }s>r\end{cases}\,. (17)

    If HLloc2()H\in L^{2}_{loc}(\mathbb{R}), then the mapping rgrr\mapsto g_{r} is an element of C([0,R],L2([0,R]))C([0,R],L^{2}([0,R])) for all R>0R>0 (see [3, Chapter 6]).

Using Γr\Gamma_{r}, one can define the “continuous polynomials” {P(r,λ;σ)}r0\{P(r,\lambda;\sigma)\}_{r\geq 0}. These are entire functions P(r,;σ)P(r,\cdot\,;\sigma), each of “degree” rr, given by

P(r,λ;σ)\displaystyle P(r,\lambda;\sigma) =defexp(iλr)0rΓr(r,t)exp(iλt)dt=(I+r)1(exp(iλ))(r).\displaystyle\stackrel{{\scriptstyle\rm def}}{{=}}\mathrm{exp}\left(i\lambda r\right)-\int\limits_{0}^{r}\Gamma_{r}(r,t)\,\mathrm{exp}\left(i\lambda t\right)\,dt=(I+\mathcal{H}_{r})^{-1}(\mathrm{exp}\left(i\lambda\cdot\right))(r)\,. (18)

This definition is motivated by analogy to the OPUC: if one considers the Toeplitz matrix

𝐓𝐧=def(c0c1cnc1c0cn1cncn+1c0),\mathbf{T_{n}}\stackrel{{\scriptstyle\rm def}}{{=}}\begin{pmatrix}c_{0}&c_{1}&\ldots&c_{n}\\ c_{-1}&c_{0}&\ldots&c_{n-1}\\ \vdots&\vdots&\ddots&\vdots\\ c_{-n}&c_{-n+1}&\ldots&c_{0}\end{pmatrix}\,,

where cj=def𝕋zj𝑑μc_{j}\stackrel{{\scriptstyle\rm def}}{{=}}\int\limits_{\mathbb{T}}z^{-j}\,d\mu are the moments of a measure μ\mu on 𝕋\mathbb{T}, then the orthogonal polynomial φn(z)\varphi_{n}(z) associated to μ\mu is given by last row of the column-vector

det𝐓𝐧det𝐓𝐧𝟏𝐓𝐧1(1zz2zn).\sqrt{\frac{\det\mathbf{T_{n}}}{\det\mathbf{T_{n-1}}}}\,\,\mathbf{T_{n}}^{-1}\begin{pmatrix}1&z&z^{2}&\ldots&z^{n}\end{pmatrix}^{\intercal}\,.

And so similar to the OPUC, {P(r,λ;σ)}r0\{P(r,\lambda;\sigma)\}_{r\geq 0} is an orthonormal system with respect to dσd\sigma, i.e. (1) holds for all f,gL2(+)f,g\in L^{2}(\mathbb{R}^{+}). Furthermore, like how each znz^{n} may be expressed as a sum of the OPUC {φk}0kn\{\varphi_{k}\}_{0\leq k\leq n}, each exponential exp(iλr)\mathrm{exp}\left(i\lambda r\right) is a “continuous sum” of the polynomials {P(s,λ;σ}0sr\{P(s,\lambda;\sigma\}_{0\leq s\leq r}, i.e. given r(0,R)r\in(0,R) we have

exp(iλr)=P(r,λ;σ)+0rLR(r,s)P(s,λ;σ)𝑑s=(I+)𝒫(,λ;σ)(𝓇),\mathrm{exp}\left(i\lambda r\right)=P(r,\lambda;\sigma)+\int\limits_{0}^{r}L_{R}(r,s)P(s,\lambda;\sigma)\,ds=(I+\cal{L}_{R})P(\cdot,\lambda;\sigma)(r)\,, (19)

for some Volterra operator \cal{L}_{R} bounded on L2([0,R])L^{2}([0,R]).

Similar to the role the Verblunsky coefficients {αn}\{\alpha_{n}\} play for OPUC (see e.g. [12, Theorem 1.5.2]), there exists A(r)Lloc2(+)A(r)\in L^{2}_{loc}(\mathbb{R}^{+}) such that (P(r,λ)P(r,λ))\begin{pmatrix}P(r,\lambda)&P_{*}(r,\lambda)\end{pmatrix}^{\intercal} is the solution to the Krein (differential) system

{P=iλPA¯P,P(0,λ)=1P=AP,P(0,λ)=1,λ,\begin{cases}P^{\prime}&=i\lambda P-\overline{A}P_{*}\,,\quad P(0,\lambda)=1\\ P_{*}^{\prime}&=-AP\,,\quad P_{*}(0,\lambda)=1\end{cases}\,,\qquad\lambda\in\mathbb{C}\,, (20)

where the derivative is taken with respect to rr and P(r,λ;σ)=defexp(iλr)P(r,λ¯;σ)¯P_{*}(r,\lambda;\sigma)\stackrel{{\scriptstyle\rm def}}{{=}}\mathrm{exp}\left(i\lambda r\right)\overline{P(r,\overline{\lambda};\sigma)}. If HH is continuous, then A(r)A(r) is given by the 0th0^{\text{th}} “coefficient” of P(r,λ)P(r,\lambda), i.e.

A(r)¯=Γr(r,0).\overline{A(r)}=\Gamma_{r}(r,0)\,. (21)

3. A priori estimates: when does P(r,λ)exp(iλr)Lp()P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)\in L^{p}(\mathbb{R})?

Suppose, just within the scope of this paragraph, that ww is very regular, e.g. w,w1L()w,w^{-1}\in L^{\infty}(\mathbb{R}). For (4) to hold for a fixed pp, it is then necessary that P(r,λ)exp(iλr)Lp()P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)\in L^{p}(\mathbb{R}) for each r>0r>0 to begin with. But this is not immediate for arbitrary weights. How do properties of ww determine which LpL^{p} spaces P(r,λ)exp(iλr)P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right) belongs to?

Proposition 3.1.

If w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}), then P(r,λ;w)exp(iλr)Lp()P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right)\in L^{p}(\mathbb{R}) for 2p2\leq p\leq\infty, for each r>0r>0. In particular, P(r,λ;w)exp(iλr)Lwp()P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right)\in L^{p}_{w}(\mathbb{R}) for 2p<2\leq p<\infty, for each r>0r>0.

Thus a priori, if w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}) all we can say is that P(r,λ)exp(iλr)Lwp()P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)\in L^{p}_{w}(\mathbb{R}) for 2p<2\leq p<\infty.

We spend the rest of this section proving Proposition 3.1.

Lemma 3.2.

If an accelerant HLloc2()H\in L^{2}_{loc}(\mathbb{R}) satisfies I+r>0I+\mathcal{H}_{r}>0 for each r>0r>0, where r\mathcal{H}_{r} is as in (10), then

  1. (a)

    P(r,λ)exp(iλr)Lp()P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)\in L^{p}(\mathbb{R}) for all 2p2\leq p\leq\infty.

  2. (b)

    for each R>0R>0, we have

    sup0rR|P(r,λ)exp(iλr)|R1/2grC([0,R],L2([0,R]))<,\sup\limits_{0\leq r\leq R}|P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)|\leq R^{1/2}\|g_{r}\|_{C([0,R],L^{2}([0,R]))}<\infty\,, (22)

    where grg_{r} is as in (17). Furthermore, P(r,λ)exp(iλr)P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right) is continuous in each variable.

Proof.

Let us first focus on part (a). By Hölder’s inequality it suffices to show P(r,λ)exp(iλr)P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right) is an element of Lp()L^{p}(\mathbb{R}) for p=2,p=2,\infty. By (15) we have

P(r,λ)exp(iλr)=exp(iλr)0rΓr(s,0)exp(iλs)𝑑s=exp(iλr)0rgr(s)exp(iλs)𝑑s,P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)=-\mathrm{exp}\left(i\lambda r\right)\int\limits_{0}^{r}\Gamma_{r}(s,0)\mathrm{exp}\left(-i\lambda s\right)\,ds=-\mathrm{exp}\left(i\lambda r\right)\int\limits_{0}^{r}g_{r}(s)\mathrm{exp}\left(-i\lambda s\right)\,ds\,, (23)

where grg_{r} is as in (17). Since grg_{r} is an element of L2([0,r])L^{2}([0,r]), then applying Plancherel to (23) yields P(r,λ)exp(iλr)L2()P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)\in L^{2}(\mathbb{R}).

Applying Cauchy-Schwarz to the right side of (23) gives the L()L^{\infty}(\mathbb{R}) estimate needed to complete the proof of (a); in fact Cauchy-Schwarz yields the estimate (22). The continuity in part 22 follows from rgrr\mapsto g_{r} being an element of C([0,R],L2([0,R]))C([0,R],L^{2}([0,R])) for each R>0R>0 (see Property (iv) of the resolvent kernel). ∎

Lemma 3.3.

If λk(w1)L1()\langle\lambda\rangle^{k}(w-1)\in L^{1}(\mathbb{R}) for k0k\geq 0 an integer, then wdλ2πw\frac{d\lambda}{2\pi} has accelerant 12π(w1ˇ)(x2π)Ck()\frac{1}{2\pi}(\widecheck{w-1})\left(\frac{x}{2\pi}\right)\in C^{k}(\mathbb{R}), the resolvent kernel Γ\Gamma exists, and Γr(,)Ck([0,r]2)\Gamma_{r}(\cdot,\cdot)\in C^{k}([0,r]^{2}) for each r>0r>0.

Proof.

By Lemma 2.1, H(x)=def12π(w1ˇ)(x2π)H(x)\stackrel{{\scriptstyle\rm def}}{{=}}\frac{1}{2\pi}(\widecheck{w-1})\left(\frac{x}{2\pi}\right) is the accelerant, and I+r>0I+\mathcal{H}_{r}>0. The latter conclusion Γ\Gamma exists. Since λk(w1)L1()\langle\lambda\rangle^{k}(w-1)\in L^{1}(\mathbb{R}), then HCk()H\in C^{k}(\mathbb{R}) and so Γr(,)Ck([0,r]2)\Gamma_{r}(\cdot,\cdot)\in C^{k}([0,r]^{2}) by Property (iii) of the Resolvent kernel Γ\Gamma. ∎

Proof of Proposition 3.1.

Lemma 2.1 implies that the measure dσ=w(λ)dλ2πd\sigma=w(\lambda)\frac{d\lambda}{2\pi} has accelerant HLloc2()H\in L^{2}_{loc}(\mathbb{R}). It follows from Lemma 3.2 (a) that

P(r,λ)exp(iλr)Lp() for p[2,].P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)\in L^{p}(\mathbb{R})\text{ for }p\in[2,\infty]\,.

To get the weighted-norm estimate, write

w|P(r,λ)exp(iλr)|p\displaystyle w|P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)|^{p} =(w1)|P(r,λ)exp(iλr)|p+|P(r,λ)exp(iλr)|p\displaystyle=(w-1)|P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)|^{p}+|P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)|^{p}
=u1|P(r,λ)exp(iλr)|p+u2|P(r,λ)exp(iλr)|p+|P(r,λ)exp(iλr)|p,\displaystyle=u_{1}|P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)|^{p}+u_{2}|P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)|^{p}+|P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)|^{p}\,,

where w1=u1+u2w-1=u_{1}+u_{2}, with u1L1()u_{1}\in L^{1}(\mathbb{R}) and u2L2()u_{2}\in L^{2}(\mathbb{R}). We are left with checking the sum is integrable.

We just showed |P(r,λ)exp(iλr)|p|P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)|^{p} is integrable for any p[2,]p\in[2,\infty]. Then u1|P(r,λ)exp(iλr)|pu_{1}|P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)|^{p} is integrable since P(r,λ)exp(iλr)L()P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)\in L^{\infty}(\mathbb{R}) and u1L1()u_{1}\in L^{1}(\mathbb{R}). To see that u2|P(r,λ)exp(iλr)|pu_{2}|P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)|^{p} is integrable, apply the Cauchy-Schwarz inequality, and use the fact that u2L2()u_{2}\in L^{2}(\mathbb{R}) and |P(r,λ)exp(iλr)|L2p()|P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)|\in L^{2p}(\mathbb{R}). ∎

4. Krein system solutions are orthogonal to lower Fourier frequencies

Definition.  Let 𝒫[a,b]\mathcal{P}_{[{a},{b}]} denote the Fourier projection onto the frequency band [a,b][a,b], i.e.

𝒫[a,b](f(ξ)exp(i()ξ)𝑑ξ)(λ)=abf(ξ)exp(iλξ)𝑑ξ\mathcal{P}_{[{a},{b}]}\left(\int\limits_{\mathbb{R}}f(\xi)\mathrm{exp}\left(i(\cdot)\xi\right)\,d\xi\right)(\lambda)=\int\limits_{a}^{b}f(\xi)\mathrm{exp}\left(i\lambda\xi\right)\,d\xi

for all fL2()f\in L^{2}(\mathbb{R}). Note 𝒫[a,b]=1χ[a2π,b2π]\mathcal{P}_{[{a},{b}]}=\mathcal{F}^{-1}\chi_{[\frac{a}{2\pi},\frac{b}{2\pi}]}\mathcal{F}.

Remark.  Note that 𝒫[0,r]g:L1()L2()\mathcal{P}_{[{0},{r}]}g:L^{1}(\mathbb{R})\to L^{2}(\mathbb{R}). Indeed, by Plancherel’s theorem, it suffices to show

χ[0,b2π]:L1()L2(),\chi_{[0,\frac{b}{2\pi}]}\mathcal{F}:L^{1}(\mathbb{R})\to L^{2}(\mathbb{R}),

which follows from the fact that

χ[0,b2π]:L()L2(),:L1()L(),\chi_{[0,\frac{b}{2\pi}]}:L^{\infty}(\mathbb{R})\to L^{2}(\mathbb{R})\,,\quad\mathcal{F}:L^{1}(\mathbb{R})\to L^{\infty}(\mathbb{R})\,,

where the last boundedness property follows from the Riemann-Lebesgue lemma.

Solutions to the Krein system satisfy the following orthogonality Lemma.

Lemma 4.1.

If w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}), then for each nonnegative integer kk, we have

P(r,λ;w),λk0rf(s)exp(iλs)𝑑sw(λ)dλ=P(r,λ;w)w,λk0rf(s)exp(iλs)𝑑sdλ=0\langle P(r,\lambda;w),\lambda^{k}\int\limits_{0}^{r}f(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{w(\lambda)d\lambda}=\langle P(r,\lambda;w)w,\lambda^{k}\int\limits_{0}^{r}f(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{d\lambda}=0\, (24)

for all fCc((0,r))f\in C_{c}^{\infty}((0,r)).

Furthermore, we also have

exp(iλr),λk0rf(s)exp(iλs)𝑑sdλ=0,\displaystyle\langle\mathrm{exp}\left(i\lambda r\right),\lambda^{k}\int\limits_{0}^{r}f(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{d\lambda}=0\,, (25)
1,λk0rf(s)exp(iλs)𝑑sdλ=0.\displaystyle\langle 1,\lambda^{k}\int\limits_{0}^{r}f(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{d\lambda}=0. (26)

Remark.  The main statement of interest in this lemma is (24). Ignoring issues of integrability, if k=0k=0, then this is the intuitive statement that

P(r,λ)wRange𝒫[0,r],P(r,\lambda)\perp_{w}\,\mathrm{Range}\,\mathcal{P}_{[0,r]}\,,

whose analogue for OPUC was used in [5, 1] in addressing the Steklov problem for OPUC.

For k1k\geq 1, this gives us the more surprising statement

P(r,λ)λkwRange𝒫[0,r],P(r,\lambda)\perp_{\lambda^{k}w}\,\mathrm{Range}\,\mathcal{P}_{[0,r]}\,,

possessing no analogue for OPUC.

Proof.

We focus first on (24). Use integration by parts to write

P(r,λ)w,λk0rf(s)exp(iλs)𝑑sdλ=ikP(r,λ)w,0rf(k)(s)exp(iλs)𝑑sdλ.\langle P(r,\lambda)w,\lambda^{k}\int\limits_{0}^{r}f(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{d\lambda}=i^{k}\langle P(r,\lambda)w,\int\limits_{0}^{r}f^{(k)}(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{d\lambda}\,.

Since w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}), then it follows from Lemma 2.1 and then Lemma 3.2 that the accelerant HH associated to ww is in Lloc2()L^{2}_{loc}(\mathbb{R}) and PP is continuous in both variables. This last property allows us to write P(r,λ)P(r,\lambda) as the limit of its averages in rr and so we may write the inner-product as

ik(limϵ01ϵrr+ϵP(s,λ)𝑑s)w,0rf(k)(s)exp(iλs)𝑑sdλ.i^{k}\langle\left(\lim\limits_{\epsilon\to 0}\frac{1}{\epsilon}\int\limits_{r}^{r+\epsilon}P(s,\lambda)\,ds\,\right)w,\int\limits_{0}^{r}f^{(k)}(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{d\lambda}\,.

Since f(k)(s)Cc(0,r)f^{(k)}(s)\in C_{c}^{\infty}(0,r) and w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}), then w(λ)0rf(k)(s)exp(iλs)𝑑sL1()w(\lambda)\int\limits_{0}^{r}f^{(k)}(s)\mathrm{exp}\left(i\lambda s\right)\,ds\in L^{1}(\mathbb{R}). Apply the dominated convergence theorem, using e.g. (22) as justification (which holds thanks to Lemma 2.1), to pull limϵ0\lim\limits_{\epsilon\to 0} outside the inner-product, yielding

iklimϵ01ϵrr+ϵP(s,λ)𝑑s,0rf(k)(s)exp(iλs)𝑑sw(λ)dλ.i^{k}\lim\limits_{\epsilon\to 0}\langle\frac{1}{\epsilon}\int\limits_{r}^{r+\epsilon}P(s,\lambda)\,ds\,,\int\limits_{0}^{r}f^{(k)}(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{w(\lambda)d\lambda}\,. (27)

Write

0rf(k)(s)exp(iλs)𝑑s=exp(iλs),f(k)(s)¯(ds,[0,r]),\int\limits_{0}^{r}f^{(k)}(s)\mathrm{exp}\left(i\lambda s\right)\,ds=\langle\mathrm{exp}\left(i\lambda s\right),\overline{f^{(k)}(s)}\rangle_{(ds,[0,r])}\,,

which, by the change of basis formula (19), equals

(I+𝓇)𝒫(,λ;σ)(𝓈),𝒻(𝓀)(𝓈)¯(𝒹𝓈,[0,𝓇])=𝒫(𝓈,λ;σ),(+𝓇)(𝒻(𝓀)¯)(𝓈)(𝒹𝓈,[0,𝓇])=0𝓇𝒫(𝓈,λ)(𝓈)𝒹𝓈,\langle(I+\cal{L}_{r})P(\cdot,\lambda;\sigma)(s),\overline{f^{(k)}(s)}\rangle_{(ds,[0,r])}=\langle P(s,\lambda;\sigma),(I+\cal{L}_{r}^{*})(\overline{f^{(k)}})(s)\rangle_{(ds,[0,r])}=\int\limits_{0}^{r}P(s,\lambda)g(s)\,ds\,,

where g(s)=def(I+𝓇)(𝒻(𝓀)¯)(𝓈)¯L2([0,r])g(s)\stackrel{{\scriptstyle\rm def}}{{=}}\overline{(I+\cal{L}_{r}^{*})(\overline{f^{(k)}})(s)}\in L^{2}([0,r]). Thus we can rewrite (27) as

iklimϵ01ϵrr+ϵP(s,λ)𝑑s,0rP(s,λ)g(s)𝑑sw(λ)dλ.i^{k}\lim\limits_{\epsilon\to 0}\langle\frac{1}{\epsilon}\int\limits_{r}^{r+\epsilon}P(s,\lambda)\,ds\,,\int\limits_{0}^{r}P(s,\lambda)g(s)\,ds\rangle_{w(\lambda)d\lambda}\,.

By the orthogonality of Krein system solutions (1) this equals 0, thereby completing the proof of (24).

By applying (24) with w=1w=1 and using P(r,λ;1)=exp(iλr)P(r,\lambda;1)=\mathrm{exp}\left(i\lambda r\right), we get (25).

As for (26), use integration by parts to write

1,λk0rf(s)exp(iλs)𝑑sdλ=ik1,0rf(k)(s)exp(iλs)𝑑sdλ.\langle 1,\lambda^{k}\int\limits_{0}^{r}f(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{d\lambda}=i^{k}\langle 1,\int\limits_{0}^{r}f^{(k)}(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{d\lambda}\,.

Use Fourier inversion to write the inner-product as

1,0rf(k)(s)exp(iλs)𝑑sdλ=0rf(k)(s)exp(iλs)𝑑s𝑑λ¯=f(k)(0)¯=0,\langle 1,\int\limits_{0}^{r}f^{(k)}(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{d\lambda}=\overline{\int\limits_{\mathbb{R}}\int\limits_{0}^{r}f^{(k)}(s)\mathrm{exp}\left(i\lambda s\right)\,ds\,d\lambda}=\overline{f^{(k)}(0)}=0\,,

where the last equality follows from suppf(0,r)\mathrm{supp}f\subseteq(0,r). ∎

The above lemma will often be paired with the one below.

Lemma 4.2.

Let r>0r>0. Then given gL1()+L2()g\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}), we have 𝒫[0,r]gL2()\mathcal{P}_{[{0},{r}]}g\in L^{2}(\mathbb{R}), which equals 0 if and only if

g(λ),0rf(s)exp(iλs)𝑑sdλ=0\langle g(\lambda),\int\limits_{0}^{r}f(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{d\lambda}=0\,

for all fCc(0,r)f\in C^{\infty}_{c}(0,r).

Proof.

Let g=g1+g2g=g_{1}+g_{2} with g1L1(),g2L2()g_{1}\in L^{1}(\mathbb{R}),\,g_{2}\in L^{2}(\mathbb{R}). By the remark at the beginning of the section and Plancherel’s theorem, we have 𝒫[0,r]gL2()\mathcal{P}_{[{0},{r}]}g\in L^{2}(\mathbb{R}).

As for the if and only if, note 𝒫[0,r]g=0\mathcal{P}_{[{0},{r}]}g=0 if and only if χ[0,r]2()\chi_{[0,r]}\cal{F}g\in L^{2}(\mathbb{R}) if and only if for all fCc(0,r)f\in C_{c}^{\infty}(0,r), we have

(𝓈),𝒻(𝓈)𝒹𝓈=0,\langle\cal{F}g(s),f(s)\rangle_{ds}=0\,,

which, by taking Fourier inverses of both entries in the inner product, holds if and only if

g(λ),0rf(s)exp(iλs)𝑑sdλ=0.\langle g(\lambda),\int\limits_{0}^{r}f(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{d\lambda}=0\,.

In the sections that follow, we will consider linear operators TT which satisfy

TLwp(d),Lwp(d)=w1/pTw1/pp,p([𝓌]𝒜𝓅(),𝓅),\|T\|_{L^{p}_{w}(\mathbb{R}^{d}),L^{p}_{w}(\mathbb{R}^{d})}=\|w^{1/p}Tw^{-1/p}\|_{p,p}\leq\cal{F}([w]_{A_{p}(\mathbb{R})},p)\,, (28)

where the function (𝓉,𝓅)\cal{F}(t,p) on [1,)×(1,)[1,\infty)\times(1,\infty) is continuous in tt for every fixed p(1,)p\in(1,\infty). In what follows, we do not need to know \cal{F} explicitly. However, \cal{F} is known in many applications. For example, the Hunt-Muckenhoupt-Wheeden theorem [13, p.205] shows that TT can be taken as a singular integral operator and recent breakthrough on domination of singular integrals by sparse operators provides the sharp dependence of \cal{F} on [w]Ap[w]_{A_{p}}. In particular, for a large class of singular integral operators, one can take (𝓉,𝓅)=𝒞(𝓅)𝓉max(1,(𝓅1)1),\cal{F}(t,p)=C(p)t^{\max(1,(p-1)^{-1})}, (see, e.g., [9, p.264]).

Lemma 4.3.

The Fourier projections 𝒫[0,r]\mathcal{P}_{[{0},{r}]} satisfy (28) for some \cal{F} independent of rr.

Proof.

By e.g. [9, p.264], (28) is satisfied by the Hilbert transform \cal{H}, which has Fourier multiplier isign(ξ)-i\,\mathrm{sign}(\xi) [13, p.26]. Now note each 𝒫[0,r]\mathcal{P}_{[{0},{r}]} is a linear combination of modulated Hilbert transforms, i.e.

𝒫[0,r]=i(exp(𝒾λ𝓇2π)exp(𝒾λ𝓇2π)2).\mathcal{P}_{[{0},{r}]}=i\left(\frac{\cal{H}-\mathrm{exp}\left(i\lambda\frac{r}{2\pi}\right)\cal{H}\mathrm{exp}\left(-i\lambda\frac{r}{2\pi}\right)}{2}\right)\,. (29)

One can check this by e.g. looking at the Fourier multipliers of all the operators involved. The triangle inequality then yields (28) for T=𝒫[0,r]T=\mathcal{P}_{[{0},{r}]} and function \cal{F} independent of rr. ∎

5. The Steklov problem for an A2()A_{2}(\mathbb{R}) weight: proof of Theorem 1.2 and Proposition 1.5

In this section we prove Theorem 1.2, and demonstrate its sharpness by Proposition 1.5. We need the following result, which follows from e.g. [15, Theorem 1, Corollary to Theorem 1].

Lemma 5.1 (Reverse Hölder inequality, open inclusion of ApA_{p} weights).

Suppose [w]Ap()γ[w]_{A_{p}(\mathbb{R})}\leq\gamma, p(1,)p\in(1,\infty). Then there exists q(γ,p)>1q(\gamma,p)>1, satisfying limγ1q(γ,p)=+\lim\limits_{\gamma\to 1}q(\gamma,p)=+\infty, such that for all t[0,q]t\in[0,q], we have

wtIγwIt\langle w^{t}\rangle_{I}\lesssim_{\gamma}\langle w\rangle_{I}^{t}\,

for all intervals II, and [wt]Ap()γ,p1[w^{t}]_{A_{p}(\mathbb{R})}\lesssim_{\gamma,p}1.

Furthermore, there exists s(γ,p)(1,p)s(\gamma,p)\in(1,p) such that for all t[s,)t\in[s,\infty), we have [w]At()η(γ,p)[w]_{A_{t}(\mathbb{R})}\leq\eta(\gamma,p), where ss and η\eta satisfy limγ1s(γ,p)=1\lim\limits_{\gamma\to 1}s(\gamma,p)=1 and limγ1η(γ,p)=1\lim\limits_{\gamma\to 1}\eta(\gamma,p)=1.

Assume [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma and define p^γ=s(γ,2)\widehat{p}_{\gamma}^{\prime}=s(\gamma,2), where ss is as in Lemma 5.1. Then limγ1p^γ=\lim\limits_{\gamma\to 1}\widehat{p}_{\gamma}=\infty, and

[w]Ap^γ()η(γ,2),p^γ>2.[w]_{A_{{\widehat{p}_{\gamma}}^{\prime}}(\mathbb{R})}\leq\eta(\gamma,2)\,,\quad\widehat{p}_{\gamma}>2\,.

Define

ϵ(γ)=def1pγ12.\epsilon(\gamma)\stackrel{{\scriptstyle\rm def}}{{=}}\frac{1}{p_{\gamma}^{\prime}}-\frac{1}{2}\,. (30)

Thus, if p[p^γ,p^γ]={p:|1p12|ϵ(γ)}p\in[\widehat{p}_{\gamma}^{\prime},\widehat{p}_{\gamma}]=\{p~{}:~{}|\frac{1}{p}-\frac{1}{2}|\leq\epsilon(\gamma)\}, then

[w]Ap(),[w]Ap()η(γ,2),[w]_{A_{p}(\mathbb{R})},\,[w]_{A_{p^{\prime}}(\mathbb{R})}\leq\eta(\gamma,2)\,,

which in particular implies that for all such pp, we have

[w]Ap(),[wp/p]Ap()γ1.[w]_{A_{p}(\mathbb{R})},[w^{-p/p^{\prime}}]_{A_{p}(\mathbb{R})}\lesssim_{\gamma}1\,. (31)

Note ϵ(γ)(0,12)\epsilon(\gamma)\in(0,\frac{1}{2}), with limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2}.

If we additionally assume w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}), then Proposition 3.1 implies P(r,λ;w)exp(iλr)L2()L()P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right)\in L^{2}(\mathbb{R})\cap L^{\infty}(\mathbb{R}). Together, these estimates yield

wP(r,λ;w)exp(iλr)=(w1)(P(r,λ;w)exp(iλr))+(w1)exp(iλr)+(P(r,λ;w)exp(iλr))L1()+L2().wP(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right)=(w-1)(P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right))\\ +(w-1)\mathrm{exp}\left(i\lambda r\right)+(P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right))\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R})\,. (32)

Proposition 3.1 and the fact that w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}) also imply

Xp=defw1/p(P(r,λ;w)exp(iλr))Lp(),2p<.X_{p}\stackrel{{\scriptstyle\rm def}}{{=}}w^{1/p}\left(P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right)\right)\in L^{p}(\mathbb{R})\,,\quad 2\leq p<\infty\,. (33)

Since

P(r,λ;w)exp(iλr)Lwp()=XpLp(),\|P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right)\|_{L^{p}_{w}(\mathbb{R})}=\|X_{p}\|_{L^{p}(\mathbb{R})}\,,

it suffices to estimate XpX_{p} in Lp()L^{p}(\mathbb{R}), which we’ll do using functional analysis methods. Note (33) means we can do functional analysis on XpX_{p} in the space Lp()L^{p}(\mathbb{R}) for any p[2,)p\in[2,\infty) when w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}).

Lemma 5.2.

If [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma and w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}), then there exists ϵ(γ)(0,12)\epsilon(\gamma)\in(0,\frac{1}{2}), with limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2}, such that for all p[2,)p\in[2,\infty) satisfying |1p12|<ϵ(γ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\gamma), we have

Xp\displaystyle X_{p} =w1/p𝒫[0,r]w1/pXp,\displaystyle=w^{1/p}\mathcal{P}_{[0,r]}w^{-1/p}X_{p}\,, (34)
0\displaystyle 0 =w1/p𝒫[0,r]w1/pXp+w1/p𝒫[0,r]w1/p(w1/pw1/p)exp(iλr)\displaystyle=w^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}}X_{p}+w^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}}(w^{1/p}-w^{-1/p^{\prime}})\mathrm{exp}\left(i\lambda r\right)\, (35)

in Lp()L^{p}(\mathbb{R}). In particular, for all such values of pp we have

(IQw,p)Xp=w1/p𝒫[0,r]w1/p(w1/pw1/p)exp(iλr),(I-Q_{w,p})X_{p}=-w^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}}(w^{1/p}-w^{-1/p^{\prime}})\mathrm{exp}\left(i\lambda r\right)\,, (36)

where

Qw,p=defw1/p𝒫[0,r]w1/pw1/p𝒫[0,r]w1/p.Q_{w,p}\stackrel{{\scriptstyle\rm def}}{{=}}w^{1/p}\mathcal{P}_{[0,r]}w^{-1/p}-w^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}}\,.

Remark.  One might wonder if (w1/pw1/p)exp(iλr)Lp()(w^{1/p}-w^{-1/p^{\prime}})\,\mathrm{exp}\left(i\lambda r\right)\in L^{p}(\mathbb{R}); it is by taking q=0q=0, and both p~\widetilde{p} and pp equal in (37) of the following lemma.

Lemma 5.3 (Integrability of ApA_{p} weights).

Let dμ(λ)=λqdλd\mu(\lambda)=\langle\lambda\rangle^{q}d\lambda for some q0q\geq 0. Suppose that

  • [w]Ap()γ[w]_{A_{p^{\prime}}(\mathbb{R})}\leq\gamma for some p(1,)p^{\prime}\in(1,\infty).

  • w1=u1+u2w-1=u_{1}+u_{2} with u1Lμp1(),u2Lμp2()u_{1}\in L^{p_{1}}_{\mu}(\mathbb{R}),u_{2}\in L^{p_{2}}_{\mu}(\mathbb{R}).

  • 1p1p2p1\leq p_{1}\leq p_{2}\leq p.

Then there exists ϵ(γ,p)>0\epsilon(\gamma,p)>0, with limγ1ϵ(γ,p)=min{1/p,1/p}\lim\limits_{\gamma\to 1}\epsilon(\gamma,p)=\min\{1/p,1/p^{\prime}\}, such that for all p~\widetilde{p} satisfying |1p1p~|<ϵ(γ,p)|\frac{1}{p}-\frac{1}{\widetilde{p}}|<\epsilon(\gamma,p), we have

w1/p~w1/p~Lμp()<.\|w^{1/\widetilde{p}}-w^{-1/\widetilde{p}^{\prime}}\|_{L^{p}_{\mu}(\mathbb{R})}<\infty\,. (37)

If q=0q=0 and p~=p\widetilde{p}=p, then there exists τ0(p)\tau_{0}(p) a small constant such that whenever

τ=def[w]A()1τ0(p),\tau\stackrel{{\scriptstyle\rm def}}{{=}}[w]_{A_{\infty}(\mathbb{R})}-1\leq\tau_{0}(p)\,,

we have the perturbative estimate

w1/pw1/pLp()pτ(pp2)/(2p)(τp2p12u1Lp1()p1+u2Lp2()p2)1/p.\|w^{1/p}-w^{-1/p^{\prime}}\|_{L^{p}(\mathbb{R})}\lesssim_{p}\tau^{(p-p_{2})/(2p)}(\tau^{\frac{p_{2}-p_{1}}{2}}\|u_{1}\|_{L^{p_{1}}(\mathbb{R})}^{p_{1}}+\|u_{2}\|_{L^{p_{2}}(\mathbb{R})}^{p_{2}})^{1/p}\,. (38)

We defer the proof of Lemma 5.3 till Section 8.

Proof of Lemma 5.2.

Take ϵ(γ)\epsilon(\gamma) as in (30). Then note all relevant quantities are well-defined as elements of, or operators on, Lp()L^{p}(\mathbb{R}): XpLp()X_{p}\in L^{p}(\mathbb{R}) by (33), w1/p𝒫[0,r]w1/pw^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}} and w1/p𝒫[0,r]w1/pw^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}} are bounded on Lp()L^{p}(\mathbb{R}) by the Hunt-Muckenhoupt-Wheeden theorem and the fact that w,wp/pAp()w,w^{-p/p^{\prime}}\in A_{p}(\mathbb{R}). And as per the previous remark, (w1/pw1/p)exp(iλr)Lp()(w^{1/p}-w^{-1/p^{\prime}})\mathrm{exp}\left(i\lambda r\right)\in L^{p}(\mathbb{R}) by Lemma 5.3.

We note (34) is equivalent to

P(r,λ)exp(iλr)=𝒫[0,r](P(r,λ)exp(iλr)).P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)=\mathcal{P}_{[{0},{r}]}(P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right))\,.

Since P(r,λ)exp(iλr)=0rΓr(r,t)exp(iλt)𝑑tP(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)=-\int\limits_{0}^{r}\Gamma_{r}(r,t)\,\mathrm{exp}\left(i\lambda t\right)\,dt with Γr(r,)L2([0,r])\Gamma_{r}(r,\cdot)\in L^{2}([0,r]) as in the discussion in Section 2, then this clearly holds.

Meanwhile (35) is equivalent to

𝒫[0,r](wP(r,λ)exp(iλr))=0.\mathcal{P}_{[0,r]}(wP(r,\lambda)-\mathrm{exp}\left(i\lambda r\right))=0\,.

From (32) it follows that wP(r,λ)exp(iλr)L1()+L2()wP(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}). Thus by Lemma 4.2, it suffices to show

wP(r,λ)exp(iλr),0rf(s)exp(iλs)𝑑sdλ=0\langle wP(r,\lambda)-\mathrm{exp}\left(i\lambda r\right),\int\limits_{0}^{r}f(s)\mathrm{exp}\left(i\lambda s\right)\,ds\,\rangle_{d\lambda}=0

for each fCc(0,r)f\in C^{\infty}_{c}(0,r). This follows from Lemma 4.1.

To get (36), subtract (35) from (34) and rearrange. ∎

As it turns out, we can invert IQw,pI-Q_{w,p} for pp sufficiently close to 22.

Lemma 5.4.

For w0w\geq 0, consider the formal operator

Qw,p=defw1/p𝒫[0,r]w1/pw1/p𝒫[0,r]w1/p,Q_{w,p}\stackrel{{\scriptstyle\rm def}}{{=}}w^{1/p}\mathcal{P}_{[{0},{r}]}w^{-1/p}-w^{-1/p^{\prime}}\mathcal{P}_{[{0},{r}]}w^{1/p^{\prime}}\,,

If [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma, then there exists ϵ(γ)(0,12)\epsilon(\gamma)\in(0,\frac{1}{2}), independent of rr, with limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2}, such that for all pp satisfying |1p12|<ϵ(γ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\gamma), IQw,pI-Q_{w,p} has bounded inverse on Lp()L^{p}(\mathbb{R}) with operator bound

(IQw,p)1p,p1.\|(I-Q_{w,p})^{-1}\|_{p,p}\lesssim 1\,. (39)

Let us briefly discuss the strategy for proving Theorem 1.2. Using Lemma 5.4 and (36), we have

Xp=(IQw,p)1w1/p𝒫[0,r]w1/p(w1/pw1/p)exp(iλr) in Lp()X_{p}=-(I-Q_{w,p})^{-1}w^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}}(w^{1/p}-w^{-1/p^{\prime}})\mathrm{exp}\left(i\lambda r\right)\,\quad\text{ in }L^{p}(\mathbb{R})\, (40)

for all p2p\geq 2 sufficiently close to 22. Then, we can estimate XpLp()\|X_{p}\|_{L^{p}(\mathbb{R})} by estimating (IQw,p)1p,p\|(I-Q_{w,p})^{-1}\|_{p,p}, w1/p𝒫[0,r]w1/pp,p\|w^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}}\|_{p,p}, and (w1/pw1/p)exp(iλr)p\|(w^{1/p}-w^{-1/p^{\prime}})\mathrm{exp}\left(i\lambda r\right)\|_{p}. However, such a process will only give us a bound for Xpp\|X_{p}\|_{p} when p2p\geq 2 since our starting point, (36), was only valid for p2p\geq 2. We address this is by noting that all elements involving pp in (40) are actually analytic in variable 1p\frac{1}{p}, and the right-side of (40) is well-defined for p<2p<2. Hence equality must hold for p<2p<2. In particular, we have the following Proposition.

Proposition 5.5.

Suppose w1L1()+L2()w-1\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R}). If [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma, then there exists ϵ(γ)(0,12)\epsilon(\gamma)\in(0,\frac{1}{2}), with limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2}, such that

Xp=(IQw,p)1w1/p𝒫[0,r]w1/p(w1/pw1/p)exp(iλr),X_{p}=-(I-Q_{w,p})^{-1}w^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}}(w^{1/p}-w^{-1/p^{\prime}})\mathrm{exp}\left(i\lambda r\right)\,, (41)

for all pp satisfying |1p12|<ϵ(γ)\left|\frac{1}{p}-\frac{1}{2}\right|<\epsilon(\gamma).

We prove Lemma 5.4 and Proposition 5.5 in Section 9 and Appendix A respectively.

Proof of Theorem 1.2.

By Proposition 5.5, we may estimate

XpLp()(IQw,p)1p,pw1/p𝒫[0,r]w1/pp,pw1/pw1/pLp()\|X_{p}\|_{L^{p}(\mathbb{R})}\leq\|(I-Q_{w,p})^{-1}\|_{p,p}\|w^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}}\|_{p,p}\|w^{1/p}-w^{-1/p^{\prime}}\|_{L^{p}(\mathbb{R})}\,

for all pp satisfying |1p12|<ϵ(γ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\gamma), where limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2}.

Since 𝒫[0,r]\mathcal{P}_{[{0},{r}]} satisfies (28) for some \cal{F} independent of rr, we may in fact write

XpLp()(IQw,p)1p,p([𝓌𝓅/𝓅]𝒜𝓅(),𝓅)𝓌1/𝓅𝓌1/𝓅𝓅().\|X_{p}\|_{L^{p}(\mathbb{R})}\leq\|(I-Q_{w,p})^{-1}\|_{p,p}\cal{F}([w^{-p/p^{\prime}}]_{A_{p}(\mathbb{R})},p)\|w^{1/p}-w^{-1/p^{\prime}}\|_{L^{p}(\mathbb{R})}\,.

Since [wp/p]Ap()γ1[w^{-p/p^{\prime}}]_{A_{p}(\mathbb{R})}\lesssim_{\gamma}1 by (31), then ([𝓌𝓅/𝓅]𝒜𝓅(),𝓅)γ,𝓅1\cal{F}([w^{-p/p^{\prime}}]_{A_{p}(\mathbb{R})},p)\lesssim_{\gamma,p}1. And by applying Lemma 5.4, we get (IQw,p)1p,p1\|(I-Q_{w,p})^{-1}\|_{p,p}\lesssim 1 for |1p12|<ϵ(γ)\left|\frac{1}{p}-\frac{1}{2}\right|<\epsilon(\gamma). Thus

XpLp()p,γw1/pw1/pLp()\|X_{p}\|_{L^{p}(\mathbb{R})}\lesssim_{p,\gamma}\|w^{1/p}-w^{-1/p^{\prime}}\|_{L^{p}(\mathbb{R})}

for all pp satisfying |1p12|<ϵ(γ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\gamma). Note all of these estimates are uniform in rr.

Part (a) now follows from noting w1/pw1/pLp()<\|w^{1/p}-w^{-1/p^{\prime}}\|_{L^{p}(\mathbb{R})}<\infty by Lemma 5.3 so long as p[p2,)p\in[p_{2},\infty).

Meanwhile part (b) follows by fixing pp, taking τ0(p)\tau_{0}(p) small enough so that

  • |1p12|<ϵ(γ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\gamma).

  • [w]A()1[w]A2()1=τ[w]_{A_{\infty}(\mathbb{R})}-1\leq[w]_{A_{2}(\mathbb{R})}-1=\tau is sufficiently small that (38) applies.

Then (38) implies (5). ∎

We now turn our attention towards Proposition 1.5. Let us first discuss its meaning: suppose w1L1()+Lp2()w-1\in L^{1}(\mathbb{R})+L^{p_{2}}(\mathbb{R}) for p2[1,2]p_{2}\in[1,2] and w1L1()w-1\notin L^{1}(\mathbb{R}); we think of p2p_{2} as measuring the decay of w1w-1, with smaller p2p_{2}’s indicating better decay. Recall Theorem 1.2 (a), which says that if [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma, then

P(r,λ;w)exp(iλr)Lwp()<\|P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right)\|_{L^{p}_{w}(\mathbb{R})}<\infty

whenever p[p2,)p\in[p_{2},\infty) satisfies |1p12|<ϵ(γ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\gamma), where limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2}. So let γ1\gamma-1 be sufficiently small that |1p212|<ϵ(γ)|\frac{1}{p_{2}}-\frac{1}{2}|<\epsilon(\gamma). Then the requirement that pp2p\geq p_{2} in Theorem 1.2 (a) is in part saying that P(r,λ;w)exp(iλr)P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right) cannot hope to decay faster than w1w-1 decays. This seemingly makes sense a priori: since the accelerant satisfies H(x)=12π(w1)ˇ(x2π)H(x)=\frac{1}{2\pi}\widecheck{(w-1)}\left(\frac{x}{2\pi}\right), then whatever decay w1w-1 possesses gets “converted” into the regularity of HH. But then the resolvent kernel Γr(s,t)\Gamma_{r}(s,t), as a rule of thumb, is at most as regular as HH. Since by (18) we have

P(r,λ)exp(iλr)=0rΓr(r,t)exp(iλt)𝑑t,P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)=-\int\limits_{0}^{r}\Gamma_{r}(r,t)\mathrm{exp}\left(i\lambda t\right)\,dt\,,

i.e. P(r,λ)exp(iλr)P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right) is essentially the inverse Fourier transform of Γr(r,)-\Gamma_{r}(r,\cdot) and so whatever regularity Γr(r,)\Gamma_{r}(r,\cdot) possesses gets “converted” into the decay of P(r,λ)exp(iλr)P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right). Thus heuristically, we expect the decay of P(r,λ)exp(iλr)P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right) to not exceed that of w1w-1, as the former “inherits” its decay from the latter.

Proof of Proposition 1.5.

By Hölder’s inequality, we can assume without loss of generality that p(1,p2)p\in(1,p_{2}).

Without loss of generality, assume δ12\delta\leq\frac{1}{2}. Let uu be an even, real-valued function on \mathbb{R} such that 0u10\leq u\leq 1 and uLq()u\in L^{q}(\mathbb{R}) if and only if qp2q\geq p_{2}. Then define w=1+δuw=1+\delta u so that [w]A2()1+δ[w]_{A_{2}(\mathbb{R})}\leq 1+\delta and w1Lp2()w-1\in L^{p_{2}}(\mathbb{R}).

Without loss of generality, assume δ\delta sufficiently small that |1p12|<ϵ(1+δ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(1+\delta), where ϵ\epsilon is as in Lemma 5.4. Then by Lemma 5.4, (IQw,p)1p,p1\|(I-Q_{w,p})^{-1}\|_{p,p}\lesssim 1. Thus by Proposition 5.5, we have

Xp=(IQw,p)1w1/p𝒫[0,r]w1/p(w1/pw1/p)exp(iλr),X_{p}=-(I-Q_{w,p})^{-1}w^{-1/p^{\prime}}\mathcal{P}_{[{0},{r}]}w^{1/p^{\prime}}(w^{1/p}-w^{-1/p^{\prime}})\mathrm{exp}\left(i\lambda r\right)\,,

or rather

(IQw,p)Xp=δexp(iλr)w1/p𝒫[r,0]u.(I-Q_{w,p})X_{p}=-\delta\,\mathrm{exp}\left(i\lambda r\right)w^{-1/p^{\prime}}\mathcal{P}_{[{-r},{0}]}u\,.

Estimating Lp()L^{p}(\mathbb{R}) norms yields

IQw,pp,pXpLp()p𝒫[r,0]uLp().\|I-Q_{w,p}\|_{p,p}\|X_{p}\|_{L^{p}(\mathbb{R})}\gtrsim_{p}\|\mathcal{P}_{[{-r},{0}]}u\|_{L^{p}(\mathbb{R})}\,.

Since w1w\sim 1, then [w]Ap(),[wp/p]Ap()p1[w]_{A_{p}(\mathbb{R})},[w^{-p/p^{\prime}}]_{A_{p}(\mathbb{R})}\sim_{p}1 and so by Lemma 4.3, we have IQw,pp,p1\|I-Q_{w,p}\|_{p,p}\lesssim 1 and so

XpLp()p𝒫[r,0]uLp().\|X_{p}\|_{L^{p}(\mathbb{R})}\gtrsim_{p}\|\mathcal{P}_{[{-r},{0}]}u\|_{L^{p}(\mathbb{R})}\,.

This then means

supr0XpLp()psupr0𝒫[r,0]uLp().\sup\limits_{r\geq 0}\|X_{p}\|_{L^{p}(\mathbb{R})}\gtrsim_{p}\sup\limits_{r\geq 0}\|\mathcal{P}_{[{-r},{0}]}u\|_{L^{p}(\mathbb{R})}\,.

It suffices to show

supr0𝒫[r,0]uLp()=.\sup\limits_{r\geq 0}\|\mathcal{P}_{[{-r},{0}]}u\|_{L^{p}(\mathbb{R})}=\infty\,.

Since uL2()u\in L^{2}(\mathbb{R}), write u(λ)=v(s)exp(iλs)𝑑su(\lambda)=\int\limits_{-\infty}^{\infty}v(s)\mathrm{exp}\left(i\lambda s\right)\,ds, where vL2()v\in L^{2}(\mathbb{R}). Since uu is real-valued and even, then so is vv. In particular this means 𝒫[r,0]u=𝒫[0,r]u¯\mathcal{P}_{[{-r},{0}]}u=\overline{\mathcal{P}_{[{0},{r}]}u}, and combined with the fact that uu is real-valued, we get supr0𝒫[r,0]uLp()=\sup\limits_{r\geq 0}\|\mathcal{P}_{[{-r},{0}]}u\|_{L^{p}(\mathbb{R})}=\infty if supr0𝒫[r,r]uLp()=\sup\limits_{r\geq 0}\|\mathcal{P}_{[{-r},{r}]}u\|_{L^{p}(\mathbb{R})}=\infty.

In fact, {𝒫[n,n]u}n0\{\mathcal{P}_{[{-n},{n}]}u\}_{n\geq 0} is unbounded in Lp()L^{p}(\mathbb{R}). Indeed, suppose to the contrary the sequence is bounded. Then there exists some increasing sequence of integers nkn_{k} and some u~Lp()\widetilde{u}\in L^{p}(\mathbb{R}) such that 𝒫[nk,nk]u\mathcal{P}_{[{-n_{k}},{n_{k}}]}u converges to u~\widetilde{u} weakly in Lp()L^{p}(\mathbb{R}). In particular, for every Schwarz function ff, we have

u~,f=limk𝒫[nk,nk]u,f=u,f,\langle\widetilde{u},f\rangle=\lim\limits_{k\to\infty}\langle\mathcal{P}_{[{-n_{k}},{n_{k}}]}u,f\rangle=\langle u,f\rangle\,,

where the last equality follows from Plancherel’s theorem. Thus u=u~Lp()u=\widetilde{u}\in L^{p}(\mathbb{R}), which contradicts our requirement that uLq()u\in L^{q}(\mathbb{R}) if and only if qp2q\geq p_{2}. This completes the proof. ∎

6. A mixed norm remainder estimate

In this section, we prove Theorem 1.6.

Definition.  If Γr(,)Ck([0,r]2)\Gamma_{r}(\cdot,\cdot)\in C^{k}([0,r]^{2}), integrate (18) by parts kk times to yield

P(r,λ)exp(iλr)=l=1kal,r(λ)λl+Rk,r(λ)λk,P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right)=\sum\limits_{l=1}^{k}\frac{a_{l,r}(\lambda)}{\lambda^{l}}+\frac{R_{k,r}(\lambda)}{\lambda^{k}}\,,

where

al,r(λ)=def{(i)l(exp(iλr)(t)l1Γr(r,t)|t=r(t)l1Γr(r,t)|t=0)l10l=0a_{l,r}(\lambda)\stackrel{{\scriptstyle\rm def}}{{=}}\begin{cases}(i)^{l}\left(\mathrm{exp}\left(i\lambda r\right)(\partial_{t})^{l-1}\Gamma_{r}(r,t)\Big{|}_{t=r}-(\partial_{t})^{l-1}\Gamma_{r}(r,t)\Big{|}_{t=0}\right)\quad&l\geq 1\\ 0&l=0\end{cases} (42)

and the remainder term Rk,r(λ)R_{k,r}(\lambda) is defined by

Rk,r(λ)=def(i)k0r((t)kΓr(r,t))exp(iλt)𝑑t.R_{k,r}(\lambda)\stackrel{{\scriptstyle\rm def}}{{=}}-(i)^{k}\int\limits_{0}^{r}\left((\partial_{t})^{k}\Gamma_{r}(r,t)\right)\mathrm{exp}\left(i\lambda t\right)\,dt\,. (43)

We can also express the remainder in terms of the solution P(r,λ)P(r,\lambda) to the Krein system and the “coefficients” al,r(λ)a_{l,r}(\lambda), i.e.

Rk,r(λ)=λk(P(r,λ)exp(iλr))l=1kλklal,r(λ).R_{k,r}(\lambda)=\lambda^{k}(P(r,\lambda)-\mathrm{exp}\left(i\lambda r\right))-\sum\limits_{l=1}^{k}\lambda^{k-l}a_{l,r}(\lambda)\,.

Note that R0,r=P(r,λ;w)exp(iλr)R_{0,r}=P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right).

The Steklov problem can be reformulated in terms of mixed norms: it asks when does one have the bound

P(r,λ)Ldr(+;Lwp(Δ))<\|P(r,\lambda)\|_{L^{\infty}_{dr}(\mathbb{R}^{+};L^{p}_{w}(\Delta))}<\infty (44)

for every compact Δ\Delta\subset\mathbb{R}? In Theorem 1.2, we estimated

R0,rLdr(+;Lwp())\|R_{0,r}\|_{L^{\infty}_{dr}(\mathbb{R}^{+};L^{p}_{w}(\mathbb{R}))}

for some pp close to 22, which implied (44). However, it also makes sense to consider other mixed norms, as we do in Theorem 1.6. To estimate R1,rR_{1,r} in Lwp()L^{p}_{w}(\mathbb{R}), it suffices to estimate Yp=defw1/pR1,r(λ)Y_{p}\stackrel{{\scriptstyle\rm def}}{{=}}w^{1/p}R_{1,r}(\lambda) in Lp()L^{p}(\mathbb{R}).

We begin with two lemmas.

Lemma 6.1.

Suppose wA2()w\in A_{2}(\mathbb{R}) and w1L1()w-1\in L^{1}(\mathbb{R}). If Condition (6) holds, then AL2(+)A\in L^{2}(\mathbb{R}^{+}), where A(r)A(r) is as given by (21).

Proof.

By [3, Theorem 12.14], A(r)L2(+)A(r)\in L^{2}(\mathbb{R}^{+}) follows if both Condition (6) holds and logwL1()\log w\in L^{1}(\mathbb{R}). Thus it suffices to check our assumptions imply logwL1()\log w\in L^{1}(\mathbb{R}).

Note that when x12x\geq\frac{1}{2}, then |log(x)||x1||\log(x)|\lesssim|x-1|. Whence

{w12}|logw|{w12}|w1|w1L1()<.\int\limits_{\{w\geq\frac{1}{2}\}}|\log w|\lesssim\int\limits_{\{w\geq\frac{1}{2}\}}|w-1|\leq\|w-1\|_{L^{1}(\mathbb{R})}<\infty\,.

Since |w1|𝑑λ<\int\limits_{\mathbb{R}}|w-1|d\lambda<\infty, then {w12}\{w\leq\frac{1}{2}\} has finite Lebesgue measure. By taking p~,p=2\widetilde{p},p^{\prime}=2 and q=0q=0 in Lemma 5.3, we get w1/2w1/2L2()w^{1/2}-w^{-1/2}\in L^{2}(\mathbb{R}) and in particular w1/2w1/2w^{1/2}-w^{-1/2} is square-integrable on {w12}\{w\leq\frac{1}{2}\}. Meanwhile ww is integrable on {w12}\{w\leq\frac{1}{2}\} since w1L1()w-1\in L^{1}(\mathbb{R}). Thus the L2()L^{2}(\mathbb{R})-triangle inequality implies w1w^{-1} is integrable on {w12}\{w\leq\frac{1}{2}\}: indeed, {w12}w1\int\limits_{\{w\leq\frac{1}{2}\}}w^{-1} equals

{w12}|w12|2{w12}|w12w12|2+{w12}|w12|2={w12}|w12w12|2+{w12}w1+{w12}1<.\int\limits_{\{w\leq\frac{1}{2}\}}|w^{-\frac{1}{2}}|^{2}\lesssim\int\limits_{\{w\leq\frac{1}{2}\}}|w^{\frac{1}{2}}-w^{-\frac{1}{2}}|^{2}+\int\limits_{\{w\leq\frac{1}{2}\}}|w^{\frac{1}{2}}|^{2}=\int\limits_{\{w\leq\frac{1}{2}\}}|w^{\frac{1}{2}}-w^{-\frac{1}{2}}|^{2}+\int\limits_{\{w\leq\frac{1}{2}\}}w-1+\int\limits_{\{w\leq\frac{1}{2}\}}1<\infty\,.

And hence logw\log w must be integrable on {w12}\{w\leq\frac{1}{2}\} as well. ∎

Lemma 6.2.

Suppose wA2()w\in A_{2}(\mathbb{R}), λ(w1)L1()\langle\lambda\rangle(w-1)\in L^{1}(\mathbb{R}) and Condition (6) holds. Then

a1,r(λ)=exp(iλr)α(r)+α2(r),a_{1,r}(\lambda)=\mathrm{exp}\left(i\lambda r\right)\alpha_{\infty}(r)+\alpha_{2}(r)\,, (45)

where α2L2(+)\alpha_{2}\in L^{2}(\mathbb{R}^{+}), αL(+)\alpha_{\infty}\in L^{\infty}(\mathbb{R}^{+}) and limrα(r)\lim\limits_{r\to\infty}\alpha_{\infty}(r) exists. In particular,

|a1,r(λ)||α(r)|+|α2(r)|Ldr(+)+Ldr2(+)|a_{1,r}(\lambda)|\leq|\alpha_{\infty}(r)|+|\alpha_{2}(r)|\in L^{\infty}_{dr}(\mathbb{R}^{+})+L^{2}_{dr}(\mathbb{R}^{+})

for all λ\lambda\in\mathbb{R}.

Proof.

Since λ(w1)L1()\langle\lambda\rangle(w-1)\in L^{1}(\mathbb{R}), then by Lemma 3.3 it follows that Γr(,)C1([0,r]2)\Gamma_{r}(\cdot,\cdot)\in C^{1}([0,r]^{2}). Hence a1,ra_{1,r} is well-defined and is given by

a1,r(λ)=i[exp(iλr)Γr(r,r)Γr(r,0)]=i[exp(iλr)Γr(0,0)A(r)¯],a_{1,r}(\lambda)=i[\mathrm{exp}\left(i\lambda r\right)\Gamma_{r}(r,r)-\Gamma_{r}(r,0)]=i[\mathrm{exp}\left(i\lambda r\right)\Gamma_{r}(0,0)-\overline{A(r)}]\,,

where we used (15) and (21) in the last equality.

Then use (16) and the fundamental theorem of Calculus to write

Γr(0,0)=Γ0(0,0)+0rxΓx(0,0)dx=Γ0(0,0)0rΓx(0,x)Γx(x,0)𝑑x.\Gamma_{r}(0,0)=\Gamma_{0}(0,0)+\int\limits_{0}^{r}\partial_{x}\Gamma_{x}(0,0)\,dx=\Gamma_{0}(0,0)-\int\limits_{0}^{r}\Gamma_{x}(0,x)\Gamma_{x}(x,0)\,dx\,.

Use (21) and the resolvent symmetries (15) again, to get

a1,r(λ)=i(exp(iλr)Γ0(0,0)+exp(iλr)0r|A(s)|2𝑑s+A(r)¯).a_{1,r}(\lambda)=-i\left(-\mathrm{exp}\left(i\lambda r\right)\Gamma_{0}(0,0)+\mathrm{exp}\left(i\lambda r\right)\int\limits_{0}^{r}|A(s)|^{2}\,ds+\overline{A(r)}\right)\,.

Define

α(r)=defi(Γ0(0,0)+0r|A(s)|2𝑑s),α2(r)=defiA(r)¯,\alpha_{\infty}(r)\stackrel{{\scriptstyle\rm def}}{{=}}-i\left(-\Gamma_{0}(0,0)+\int\limits_{0}^{r}|A(s)|^{2}\,ds\right)\,,\quad\alpha_{2}(r)\stackrel{{\scriptstyle\rm def}}{{=}}-i\overline{A(r)}\,,

so that (45) holds. Since A(r)L2(+)A(r)\in L^{2}(\mathbb{R}^{+}) by Lemma 6.1, then α2(r)L2(+)\alpha_{2}(r)\in L^{2}(\mathbb{R}^{+}) and

|α(r)||Γ0(0,0)|+0|A(s)|2𝑑s<,limrα(r)=i(Γ0(0,0)+0|A(s)|2𝑑s).|\alpha_{\infty}(r)|\leq\left|\Gamma_{0}(0,0)\right|+\int\limits_{0}^{\infty}|A(s)|^{2}\,ds<\infty\,,\quad\lim\limits_{r\to\infty}\alpha_{\infty}(r)=-i\left(-\Gamma_{0}(0,0)+\int\limits_{0}^{\infty}|A(s)|^{2}\,ds\right)\,.

Now we can do functional analysis like in the proof of Theorem 1.2.

Lemma 6.3.

Suppose λq(w1)L1()\langle\lambda\rangle^{q}(w-1)\in L^{1}(\mathbb{R}) for some q>2q>2. If [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma, then there exists ϵ(γ)(0,12)\epsilon(\gamma)\in(0,\frac{1}{2}), with limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2}, such that for all p(1,q]p\in(1,q] satisfying |1p12|<ϵ(γ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\gamma), we have

Yp\displaystyle Y_{p} =w1/p𝒫[0,r]w1/pYp,\displaystyle=w^{1/p}\mathcal{P}_{[0,r]}w^{-1/p}Y_{p}\,, (46)
w1/p𝒫[0,r]w1/pYp\displaystyle w^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}}Y_{p} =w1/p𝒫[0,r]w1/p(w1/pw1/p)(λexp(iλr)+a1,r(λ))\displaystyle=-w^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}}(w^{1/p}-w^{-1/p^{\prime}})\left(\lambda\mathrm{exp}\left(i\lambda r\right)+a_{1,r}(\lambda)\right) (47)

in Lp()L^{p}(\mathbb{R}). In particular, for all such values of pp, we have

(IQw,p)Yp=w1/p𝒫[0,r]w1/p(w1/pw1/p)(λexp(iλr)+a1,r(λ)),(I-Q_{w,p})Y_{p}=-w^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}}(w^{1/p}-w^{-1/p^{\prime}})\left(\lambda\mathrm{exp}\left(i\lambda r\right)+a_{1,r}(\lambda)\right)\,, (48)

where Qw,p=defw1/p𝒫[0,r]w1/pw1/p𝒫[0,r]w1/pQ_{w,p}\stackrel{{\scriptstyle\rm def}}{{=}}w^{1/p}\mathcal{P}_{[{0},{r}]}w^{-1/p}-w^{-1/p^{\prime}}\mathcal{P}_{[{0},{r}]}w^{1/p^{\prime}}.

Remark.  Since λ2(w1)L1()\langle\lambda\rangle^{2}(w-1)\in L^{1}(\mathbb{R}), then Γr(,)C2([0,r]2)\Gamma_{r}(\cdot,\cdot)\in C^{2}([0,r]^{2}) by Lemma 3.3. Integrating (43) by parts once yields R1,r(λ)Lp()R_{1,r}(\lambda)\in L^{p}(\mathbb{R}) for 1<p1<p\leq\infty, and so YpLp()Y_{p}\in L^{p}(\mathbb{R}) for 1<p1<p\leq\infty.

We also remark that the right-side of (47) is well-defined, i.e.

(w1/pw1/p)(λexp(iλr)+a1,r(λ))Lp()(w^{1/p}-w^{-1/p^{\prime}})(\lambda\mathrm{exp}\left(i\lambda r\right)+a_{1,r}(\lambda))\in L^{p}(\mathbb{R})

for 1<pq1<p\leq q, for each r>0r>0. Indeed this follows from Lemma 5.3, (45), and that λq(w1)L1()\langle\lambda\rangle^{q}(w-1)\in L^{1}(\mathbb{R}). We require q>2q>2 just so that we may consider p>p>, which we will do later.

Proof.

Take ϵ(γ)\epsilon(\gamma) as in (30), which in particular implies (31) holds for all pp of concern, i.e. for all pp such that |1p12|<ϵ\left|\frac{1}{p}-\frac{1}{2}\right|<\epsilon. Then note all relevant quantities are well-defined as elements of, or operators on, Lp()L^{p}(\mathbb{R}): w1/p𝒫[0,r]w1/pw^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}} and w1/p𝒫[0,r]w1/pw^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}} are bounded operators on Lp()L^{p}(\mathbb{R}) by the Hunt-Muckenhoupt-Wheeden theorem and that w,wp/pAp()w,w^{-p/p^{\prime}}\in A_{p}(\mathbb{R}). And as per the remark above, YpLp()Y_{p}\in L^{p}(\mathbb{R}), and (w1/pw1/p)(λexp(iλr)+a1,r(λ))Lp()(w^{1/p}-w^{-1/p^{\prime}})(\lambda\mathrm{exp}\left(i\lambda r\right)+a_{1,r}(\lambda))\in L^{p}(\mathbb{R}) by Lemma 5.3.

Notice (43) implies (46) is equivalent to

R1,r=𝒫[0,r]R1,r.R_{1,r}=\mathcal{P}_{[{0},{r}]}R_{1,r}\,.

Since R1,r=i0rtΓr(r,t)exp(iλt)dtR_{1,r}=-i\int\limits_{0}^{r}\partial_{t}\Gamma_{r}(r,t)\,\mathrm{exp}\left(i\lambda t\right)\,dt with tΓr(r,)C([0,r])\partial_{t}\Gamma_{r}(r,\cdot)\in C([0,r]), it clearly holds.

Concerning (47), it is equivalent to

𝒫[0,r][w(R1,r(λ)+λexp(iλr)+a1,r(λ))(λexp(iλr)+a1,r(λ))]=0.\mathcal{P}_{[0,r]}[w(R_{1,r}(\lambda)+\lambda\mathrm{exp}\left(i\lambda r\right)+a_{1,r}(\lambda))-(\lambda\mathrm{exp}\left(i\lambda r\right)+a_{1,r}(\lambda))]=0\,.

We note that 𝒫[0,r]\mathcal{P}_{[{0},{r}]} is acting on

w(R1,r(λ)+λexp(iλr)+a1,r(λ))(λexp(iλr)+a1,r(λ))L1()+L2();w(R_{1,r}(\lambda)+\lambda\mathrm{exp}\left(i\lambda r\right)+a_{1,r}(\lambda))-(\lambda\mathrm{exp}\left(i\lambda r\right)+a_{1,r}(\lambda))\in L^{1}(\mathbb{R})+L^{2}(\mathbb{R})\,;

to see this, rewrite it as

(w1)R1,r+R1,r+(w1)(λexp(iλr)+a1,r).(w-1)R_{1,r}+R_{1,r}+(w-1)(\lambda\mathrm{exp}\left(i\lambda r\right)+a_{1,r})\,.

By the previous remark, R1,rL2()R_{1,r}\in L^{2}(\mathbb{R}) and also R1,rL()R_{1,r}\in L^{\infty}(\mathbb{R}). If we combine these two estimates with the assumption that λ(w1)L1()\langle\lambda\rangle(w-1)\in L^{1}(\mathbb{R}), then we get (w1)R1,rL1()(w-1)R_{1,r}\in L^{1}(\mathbb{R}). Finally, combine assumption λ(w1)L1()\langle\lambda\rangle(w-1)\in L^{1}(\mathbb{R}) with the estimate a1,rL()a_{1,r}\in L^{\infty}(\mathbb{R}), which follows from (42), to get (w1)(λexp(iλr)+a1,r)L1()(w-1)(\lambda\mathrm{exp}\left(i\lambda r\right)+a_{1,r})\in L^{1}(\mathbb{R}).

Thus by Lemma 4.2, it suffices to show

w(R1,r+a1,r+λexp(iλr))(λexp(iλr)+a1,r),0rf(s)exp(iλs)𝑑sdλ=0,\langle w(R_{1,r}+a_{1,r}+\lambda\mathrm{exp}\left(i\lambda r\right))-(\lambda\mathrm{exp}\left(i\lambda r\right)+a_{1,r}),\int\limits_{0}^{r}f(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{d\lambda}=0\,,

or equivalently

wλP(r,λ)(λexp(iλr)+a1,r),0rf(s)exp(iλs)𝑑sdλ=0\langle w\lambda P(r,\lambda)-(\lambda\mathrm{exp}\left(i\lambda r\right)+a_{1,r}),\int\limits_{0}^{r}f(s)\mathrm{exp}\left(i\lambda s\right)\,ds\rangle_{d\lambda}=0

for each fCc(0,r)f\in C_{c}^{\infty}(0,r). This now follows from Lemma 4.1 and the identity

a1,r(λ)=i(exp(iλr)Γr(r,r)Γr(r,0)),a_{1,r}(\lambda)=i(\mathrm{exp}\left(i\lambda r\right)\Gamma_{r}(r,r)-\Gamma_{r}(r,0))\,,

as given by (42).

Add (47) and (46) and rearrange to obtain (48). ∎

Proof of Theorem 1.6.

It suffices to estimate YpLdr(;Lp())+Ldr2(;Lp())\|Y_{p}\|_{L^{\infty}_{dr}(\mathbb{R};L^{p}(\mathbb{R}))+L^{2}_{dr}(\mathbb{R};L^{p}(\mathbb{R}))}; we first estimate YpLp()\|Y_{p}\|_{L^{p}(\mathbb{R})} in terms of rr.

From (48) we have

YpLp()(IQw,p)1p,pw1/p𝒫[0,r]w1/pp,p(w1/pw1/p)(λexp(iλr)+a1,r(λ))Lp().\displaystyle\|Y_{p}\|_{L^{p}(\mathbb{R})}\leq\|(I-Q_{w,p})^{-1}\|_{p,p}\,\cdot\,\|w^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}}\|_{p,p}\,\cdot\,\|(w^{1/p}-w^{-1/p^{\prime}})(\lambda\mathrm{exp}\left(i\lambda r\right)+a_{1,r}(\lambda))\|_{L^{p}(\mathbb{R})}\,.

By Lemma 5.4, we have (IQw,p)1p,p1\|(I-Q_{w,p})^{-1}\|_{p,p}\lesssim 1 whenever |1p12|<ϵ(γ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\gamma), where limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2}. Since 𝒫[0,r]\mathcal{P}_{[{0},{r}]} satisfies (28) for some \cal{F} independent of rr, we can also estimate

w1/p𝒫[0,r]w1/pp,p([𝓌𝓅/𝓅]𝒜𝓅(),𝓅)γ,𝓅1.\|w^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}}\|_{p,p}\leq\cal{F}([w^{-p/p^{\prime}}]_{A_{p}(\mathbb{R})},p)\lesssim_{\gamma,p}1\,.

If additionally p(1,q]p\in(1,q], then

(w1/pw1/p)(λexp(iλr)+a1,r(λ))Ldλp()w1/pw1/pLλqdλp()(1+|α2(r)|+|α(r)|)\|(w^{1/p}-w^{-1/p^{\prime}})(\lambda\mathrm{exp}\left(i\lambda r\right)+a_{1,r}(\lambda))\|_{L^{p}_{d\lambda}(\mathbb{R})}\lesssim\|w^{1/p}-w^{-1/p^{\prime}}\|_{L^{p}_{\langle\lambda\rangle^{q}d\lambda}(\mathbb{R})}(1+|\alpha_{2}(r)|+|\alpha_{\infty}(r)|)

where α2,α\alpha_{2},\alpha_{\infty} arise from Lemma 6.2. By Lemma 5.3 with p~=p\widetilde{p}=p, we have

w1/pw1/pLλqdλp()p,w,q1.\|w^{1/p}-w^{-1/p^{\prime}}\|_{L^{p}_{\langle\lambda\rangle^{q}d\lambda}(\mathbb{R})}\lesssim_{p,w,q}1\,.

All together, we get

YpLwp()p,w,q(1+α2(r)+α(r)).\|Y_{p}\|_{L^{p}_{w}(\mathbb{R})}\lesssim_{p,w,q}(1+\alpha_{2}(r)+\alpha_{\infty}(r))\,.

Since 1+α2(r)+α(r)Ldr(+)+Ldr2(+)1+\alpha_{2}(r)+\alpha_{\infty}(r)\in L^{\infty}_{dr}(\mathbb{R}^{+})+L^{2}_{dr}(\mathbb{R}^{+}) by Lemma 6.2, then

YpLdr(+;Lwp())+Ldr2(+;Lwp())<.\|Y_{p}\|_{L^{\infty}_{dr}(\mathbb{R}^{+};L^{p}_{w}(\mathbb{R}))+L^{2}_{dr}(\mathbb{R}^{+};L^{p}_{w}(\mathbb{R}))}<\infty\,.

7. Higher order remainder estimates: proof of Theorem 1.7

We will spend the rest of this subsection proving Theorem 1.7. Suppose

(w1)λkL1().(w-1)\langle\lambda\rangle^{k}\in L^{1}(\mathbb{R})\,.

Then ΓCk(+)\Gamma\in C^{k}(\mathbb{R}^{+}) by Lemma 3.3, and so Rk,rR_{k,r} is well-defined. The following Lemma shows {al,r}\{a_{l,r}\} are uniformly bounded in (λ,r)(\lambda,r).

Lemma 7.1.

Suppose L=def(w1)λkL1()<L\stackrel{{\scriptstyle\rm def}}{{=}}\|(w-1)\langle\lambda\rangle^{k}\|_{L^{1}(\mathbb{R})}<\infty. If λk(w1)L()δ<1\|\langle\lambda\rangle^{k}(w-1)\|_{L^{\infty}(\mathbb{R})}\leq\delta<1, then

tjΓr(r,t)L((0,r))L1δ,j=0,,k.\|\partial_{t}^{j}\Gamma_{r}(r,t)\|_{L^{\infty}((0,r))}\lesssim\frac{L}{1-\delta}\,,\,\,j=0,\ldots,k\,.

In particular |al,r(λ)|L1δ|a_{l,r}(\lambda)|\lesssim\frac{L}{1-\delta}.

Proof.

The bound on |al,r(λ)||a_{l,r}(\lambda)| follows from its definition in (42) and the bound on tjΓr(r,t)L((0,r))\|\partial_{t}^{j}\Gamma_{r}(r,t)\|_{L^{\infty}((0,r))}. To get this latter bound, by the resolvent symmetries (15) it suffices to estimate Γr(u,0)\Gamma_{r}(u,0) and all its derivatives in [0,r][0,r]. The resolvent identity (14) is equivalent to

(I+r)Γr(,s)=H(s),s[0,r](I+\mathcal{H}_{r})\Gamma_{r}(\cdot,s)=H(\cdot-s)\,,\quad s\in[0,r]

and thus

Γr(u,0)=(I+r)1H(u).\Gamma_{r}(u,0)=(I+\mathcal{H}_{r})^{-1}H(u)\,.

By Lemma 2.1 we have H(x)=1(w(2π)1)H(x)=\mathcal{F}^{-1}(w(2\pi\cdot)-1); we in fact claim rL2([0,r]),L2([0,r])δ<1\|\mathcal{H}_{r}\|_{L^{2}([0,r]),L^{2}([0,r])}\leq\delta<1. Indeed:

rfL2()=0r1(w(2π)1)(xy)f(y)dyL2()\displaystyle\|\mathcal{H}_{r}f\|_{L^{2}(\mathbb{R})}=\|\int\limits_{0}^{r}\mathcal{F}^{-1}(w(2\pi\cdot)-1)(x-y)f(y)dy\|_{L^{2}(\mathbb{R})} =(w(2π)1)(fχ[0,r])L2()\displaystyle=\|(w(2\pi\cdot)-1)\mathcal{F}(f\chi_{[0,r]})\|_{L^{2}(\mathbb{R})}
w1L()(fχ[0,r])L2()\displaystyle\leq\|w-1\|_{L^{\infty}(\mathbb{R})}\|\mathcal{F}(f\chi_{[0,r]})\|_{L^{2}(\mathbb{R})}
δfL2([0,r]).\displaystyle\leq\delta\|f\|_{L^{2}([0,r])}\,.

Thus by geometric sum

Γr(u,0)=(I+r)1H(u)=l=0(r)lH(u).\Gamma_{r}(u,0)=(I+\mathcal{H}_{r})^{-1}H(u)=\sum\limits_{l=0}^{\infty}(-\mathcal{H}_{r})^{l}H(u)\,.

We estimate l=0(r)lH(u)\sum\limits_{l=0}^{\infty}(-\mathcal{H}_{r})^{l}H(u) and all its derivatives. Differentiate (10) to get

|(ddu)jr(f)(u)|=|0rH(j)(uv)f(v)𝑑v|H(j)L2()fL2([0,r])\left|\left(\frac{d}{du}\right)^{j}\mathcal{H}_{r}(f)(u)\right|=\left|\int\limits_{0}^{r}H^{(j)}(u-v)f(v)\,dv\right|\leq\|H^{(j)}\|_{L^{2}(\mathbb{R})}\|f\|_{L^{2}([0,r])}\,

for all 0jk0\leq j\leq k. Take f=(r)l1Hf=(-\mathcal{H}_{r})^{l-1}H to get

|(ddu)j(r)lH(u)|\displaystyle\left|\left(\frac{d}{du}\right)^{j}(-\mathcal{H}_{r})^{l}H(u)\right| H(j)L2()rl1HL2([0,r])\displaystyle\leq\|H^{(j)}\|_{L^{2}(\mathbb{R})}\|\mathcal{H}_{r}^{l-1}H\|_{L^{2}([0,r])}
=H(j)L2([0,r])rL2([0,r]),L2()l1HL2([0,r])\displaystyle=\|H^{(j)}\|_{L^{2}([0,r])}\|\mathcal{H}_{r}\|_{L^{2}([0,r]),L^{2}(\mathbb{R})}^{l-1}\|H\|_{L^{2}([0,r])}
Lδl,\displaystyle\lesssim L\delta^{l}\,,

where in the last inequality we estimated

H(j)L2()(w1)λjL2()(w1)λkL2()δ1/2L1/2,\|H^{(j)}\|_{L^{2}(\mathbb{R})}\lesssim\|(w-1)\lambda^{j}\|_{L^{2}(\mathbb{R})}\leq\|(w-1)\langle\lambda\rangle^{k}\|_{L^{2}(\mathbb{R})}\leq\delta^{1/2}L^{1/2}\,,

and similarly for HL2([0,r])\|H\|_{L^{2}([0,r])}.

Thus the sum l=0(ddu)j(r)lH(u)\sum\limits_{l=0}^{\infty}\left(\frac{d}{du}\right)^{j}(-\mathcal{H}_{r})^{l}H(u) converges absolutely and so one can use the dominated convergence theorem to show we get the desired estimate on

ujΓr(u,0)=ujl=0(r)lH(u).\partial_{u}^{j}\Gamma_{r}(u,0)=\partial_{u}^{j}\sum\limits_{l=0}^{\infty}(-\mathcal{H}_{r})^{l}H(u)\,.

Now we proceed with our usual functional-analytic approach.

Lemma 7.2.

Suppose that λk(w1)L1()\langle\lambda\rangle^{k}(w-1)\in L^{1}(\mathbb{R}). Then

Rk,r=𝒫[0,r]Rk,rR_{k,r}=\mathcal{P}_{[0,r]}R_{k,r}\, (49)

and

𝒫[0,r](w1)Rk,r=𝒫[0,r]Rk,r𝒫[0,r](w1)(λkexp(iλr)+l=1kal,rλkl)\mathcal{P}_{[0,r]}(w-1)R_{k,r}=-\mathcal{P}_{[0,r]}R_{k,r}-\mathcal{P}_{[0,r]}(w-1)(\lambda^{k}\mathrm{exp}\left(i\lambda r\right)+\sum\limits_{l=1}^{k}a_{l,r}\lambda^{k-l})\, (50)

in Lp()L^{p}(\mathbb{R}) for 2p<2\leq p<\infty. If in addition |λk(w1)|δ<1|\langle\lambda\rangle^{k}(w-1)|\leq\delta<1, then

Rk,r(λ)=(I𝒫[0,r](1w))1𝒫[0,r](w1)(λkexp(iλr)+l=1kal,rλkl),R_{k,r}(\lambda)=-(I-\mathcal{P}_{[0,r]}(1-w))^{-1}\mathcal{P}_{[0,r]}(w-1)(\lambda^{k}\mathrm{exp}\left(i\lambda r\right)+\sum\limits_{l=1}^{k}a_{l,r}\lambda^{k-l})\,, (51)

where both sides are well-defined in, e.g., L2()L^{2}(\mathbb{R}).

Proof.

First note that Rk,rLp()R_{k,r}\in L^{p}(\mathbb{R}) for 2p2\leq p\leq\infty. Indeed, p=2p=2 follows by (43) and Plancherel, and p=p=\infty from (43) and that ΓrCk([0,r])\Gamma_{r}\in C^{k}([0,r]). From this and the fact that λk(w1)L1()\langle\lambda\rangle^{k}(w-1)\in L^{1}(\mathbb{R}) we have both sides of (49) and (50) are sensical. Trivially, (49) holds. Meanwhile (50) is equivalent to

𝒫[0,r]((w1)Rk,r+Rk,r+(w1)(λkexp(iλr)+l=1kal,rλkl))=0\mathcal{P}_{[0,r]}\left((w-1)R_{k,r}+R_{k,r}+(w-1)(\lambda^{k}\mathrm{exp}\left(i\lambda r\right)+\sum\limits_{l=1}^{k}a_{l,r}\lambda^{k-l})\right)=0\,

or rather

𝒫[0,r](wλkP(r,λ)λkexp(iλr)l=1kal,rλkl)=0,\mathcal{P}_{[0,r]}(w\lambda^{k}P(r,\lambda)-\lambda^{k}\mathrm{exp}\left(i\lambda r\right)-\sum\limits_{l=1}^{k}a_{l,r}\lambda^{k-l})=0\,,

which follows by applying Lemma 4.1 and (42).

Add (50) and (49) to yield

(I𝒫[0,r](1w))Rk,r(λ)=𝒫[0,r](w1)(λkexp(iλr)+l=1kal,rλkl).(I-\mathcal{P}_{[0,r]}(1-w))R_{k,r}(\lambda)=-\mathcal{P}_{[0,r]}(w-1)(\lambda^{k}\mathrm{exp}\left(i\lambda r\right)+\sum\limits_{l=1}^{k}a_{l,r}\lambda^{k-l})\,.

Then on L2()L^{2}(\mathbb{R}), we have

𝒫[0,r](1w)2,2𝒫[0,r]2,21wL()1δ<1.\|\mathcal{P}_{[{0},{r}]}(1-w)\|_{2,2}\leq\left\|\mathcal{P}_{[{0},{r}]}\right\|_{2,2}\|1-w\|_{L^{\infty}(\mathbb{R})}\leq 1\cdot\delta<1\,.

Thus the operator (I𝒫[0,r](1w))(I-\mathcal{P}_{[0,r]}(1-w)) has bounded inverse k=0(𝒫[0,r](1w))k\sum\limits_{k=0}^{\infty}(\mathcal{P}_{[0,r]}(1-w))^{k} on L2()L^{2}(\mathbb{R}), and so (51) follows. ∎

We will also need the following Lemma to estimate the Fourier projections.

Lemma 7.3.

If δ(0,1)\delta\in(0,1), then there exists ϵ(δ)(0,12)\epsilon(\delta)\in(0,\frac{1}{2}), with limδ0ϵ(δ)=12\lim\limits_{\delta\to 0}\epsilon(\delta)=\frac{1}{2}, such that for all pp satisfying |1p12|<ϵ(δ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\delta) we have

𝒫[0,r]p,p1δ.\|\mathcal{P}_{[{0},{r}]}\|_{p,p}\leq\frac{1}{\sqrt{\delta}}\,.
Proof.

By (29) and self-adjointness of 𝒫[0,r]\mathcal{P}_{[{0},{r}]}, it follows that 𝒫[0,r]p,p=𝒫[0,r]p,p𝓅,𝓅\|\mathcal{P}_{[{0},{r}]}\|_{p^{\prime},p^{\prime}}=\|\mathcal{P}_{[{0},{r}]}\|_{p,p}\leq\|\cal{H}\|_{p,p}, where \cal{H} is the Hilbert Transform. By [10], 1𝓅,𝓅=tan(π2𝓅)1\leq\|\cal{H}\|_{p,p}=\tan(\frac{\pi}{2p}) for p(1,2]p\in(1,2]. Let p0=p0(δ)(1,2]p_{0}=p_{0}(\delta)\in(1,2] be the unique element of (1,2](1,2] that satisfies tan(π2p0)=δ1/2\tan(\frac{\pi}{2p_{0}})=\delta^{-1/2}, and let ϵ(δ)=def1p012\epsilon(\delta)\stackrel{{\scriptstyle\rm def}}{{=}}\frac{1}{p_{0}}-\frac{1}{2}.

By duality and (29), we have 𝒫[0,r]p0,p0,𝒫[0,r]p0,p0δ1/2\|\mathcal{P}_{[{0},{r}]}\|_{p_{0}^{\prime},p_{0}^{\prime}},\|\mathcal{P}_{[{0},{r}]}\|_{p_{0},p_{0}}\leq\delta^{-1/2}. Interpolate between both estimates to get

𝒫[0,r]p,p1δ\|\mathcal{P}_{[{0},{r}]}\|_{p,p}\leq\frac{1}{\sqrt{\delta}}\,

for all pp satisfying |1p12|<ϵ(δ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\delta).

To see that limδ0ϵ(δ)=12\lim\limits_{\delta\to 0}\epsilon(\delta)=\frac{1}{2}, it suffices to note that tan(π2p)\tan(\frac{\pi}{2p}) is increasing in 1p\frac{1}{p} for p(1,2]p\in(1,2] and has singularity at 1p=1\frac{1}{p}=1. Thus as δ0\delta\to 0, we have δ1/2\delta^{-1/2}\to\infty, meaning that for our definition of p0p_{0}, we have 1p01\frac{1}{p_{0}}\to 1 and so ϵ(δ)12\epsilon(\delta)\to\frac{1}{2}. ∎

We are now in a position to prove Theorem 1.7.

Proof of Theorem 1.7.

By (51) and Lemma 7.1, it follows that

Rk,rLp()k(I𝒫[0,r](1w))1p,p𝒫[0,r]p,pλk(w1)Lp()(1+λk(w1)L1()1δ).\|R_{k,r}\|_{L^{p}(\mathbb{R})}\lesssim_{k}\|(I-\mathcal{P}_{[0,r]}(1-w))^{-1}\|_{p,p}\|\mathcal{P}_{[{0},{r}]}\|_{p,p}\|\langle\lambda\rangle^{k}(w-1)\|_{L^{p}(\mathbb{R})}\left(1+\frac{\|\langle\lambda\rangle^{k}(w-1)\|_{L^{1}(\mathbb{R})}}{1-\delta}\right)\,.

First note that by (29), we have 𝒫[0,r]p,p𝓅,𝓅𝓅1\|\mathcal{P}_{[{0},{r}]}\|_{p,p}\lesssim\|\cal{H}\|_{p,p}\lesssim_{p}1.

Next, by Lemma 7.3 we can choose ϵ(δ)(0,12)\epsilon(\delta)\in(0,\frac{1}{2}), with limδ0ϵ(δ)=12\lim\limits_{\delta\to 0}\epsilon(\delta)=\frac{1}{2}, such that for all pp satisfying |1p12|<ϵ(δ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\delta), we have

𝒫[0,r](1w)p,p𝒫[0,r]p,p1wδ<1.\|\mathcal{P}_{[0,r]}(1-w)\|_{p,p}\leq\|\mathcal{P}_{[0,r]}\|_{p,p}\|1-w\|_{\infty}\leq\sqrt{\delta}<1\,.

Now by geometric sum, I𝒫[0,r](1w)I-\mathcal{P}_{[0,r]}(1-w) has inverse j=0(𝒫[0,r](1w))k\sum\limits_{j=0}^{\infty}(\mathcal{P}_{[{0},{r}]}(1-w))^{k} on Lp()L^{p}(\mathbb{R}), with the bound

(I𝒫[0,r](1w))1p,pk=0(𝒫[0,r](1w))kp,p11δ1/211δ\|(I-\mathcal{P}_{[0,r]}(1-w))^{-1}\|_{p,p}\leq\sum\limits_{k=0}^{\infty}\|(\mathcal{P}_{[0,r]}(1-w))^{k}\|_{p,p}\leq\frac{1}{1-\delta^{1/2}}\sim\frac{1}{1-\delta}

whenever |1p12|<ϵ(δ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\delta).

Combine these estimates with the fact that 11δ1\frac{1}{1-\delta}\geq 1, we get

Rk,rLp()p,k1(1δ)2λk(w1)Lp()(1+λk(w1)L1())\|R_{k,r}\|_{L^{p}(\mathbb{R})}\lesssim_{p,k}\frac{1}{(1-\delta)^{2}}\|\langle\lambda\rangle^{k}(w-1)\|_{L^{p}(\mathbb{R})}(1+\|\langle\lambda\rangle^{k}(w-1)\|_{L^{1}(\mathbb{R})})\, (52)

for all pp satisfying |1p12|<ϵ(δ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\delta), from which part (a) follows. As for part (b), given a p(1,)p\in(1,\infty) choose δ0(p)\delta_{0}(p) small enough so that |1p12|<ϵ(δ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\delta) for all δ(0,δ0)\delta\in(0,\delta_{0}). Then the estimate

λk(w1)Lp()δ1/pλk(w1)L1()1/p,\|\langle\lambda\rangle^{k}(w-1)\|_{L^{p}(\mathbb{R})}\leq\delta^{1/p^{\prime}}\|\langle\lambda\rangle^{k}(w-1)\|_{L^{1}(\mathbb{R})}^{1/p}\,,

combined with (52), yields part (b). ∎

8. Proof of Lemma 5.3

Before proving Lemma 5.3, we will need terminology and lemmas for a Calderón-Zygmund decomposition.

Definition.  An interval [a,a+r][a,a+r] has left neighbor [ar,a][a-r,a] and right neighbor [a+r,a+2r][a+r,a+2r].

Two intervals are almost disjoint if their intersection is empty or a single point.

Given an interval II, let DI=defdist(I,0)D_{I}\stackrel{{\scriptstyle\rm def}}{{=}}\mathrm{dist}(I,0) denote its distance from the origin.

An interval II in \mathbb{R} is dyadic if it is of the form {[j2n,(j+1)2n]}j,n\{[j2^{-n},(j+1)2^{-n}]\}_{j,n\in\mathbb{Z}}.

Two dyadic intervals II and JJ are siblings if their union is a dyadic interval KK of strictly larger size, which we call the parent.

So define a partial ordering \preceq on the dyadic intervals: if IKI\subset K, we write IKI\preceq K, and say KK is an ancestor of II. Note that two dyadic intervals are either almost disjoint, or comparable via \preceq.

See e.g. [13, Chapter 1, Section 3] or [14, Chapter 1, Theorem 4] for other variants of the Calderón-Zygmund decomposition below.

Lemma 8.1 (Calderón-Zygmund decomposition).

Let dμ=λqdλd\mu=\langle\lambda\rangle^{q}d\lambda for some q0q\geq 0, and suppose uLμ1()u\in L^{1}_{\mu}(\mathbb{R}). If β>0\beta>0 and Eβ=def{λ:|u|>β}E_{\beta}\stackrel{{\scriptstyle\rm def}}{{=}}\{\lambda\in\mathbb{R}~{}:~{}|u|>\beta\}, then there exists a collection of dyadic intervals {Ij}\{I_{j}\} with the following properties:

  1. (i)

    Eβ=def{λ:|u|>β}jIjE_{\beta}\stackrel{{\scriptstyle\rm def}}{{=}}\{\lambda\in\mathbb{R}~{}:~{}|u|>\beta\}\subset\bigcup\limits_{j}I_{j}.

  2. (ii)

    Each IjI_{j} is maximal with respect to \preceq among those dyadic intervals II satisfying β<1μ(I)I|u|dμ\beta<\frac{1}{\mu(I)}\int\limits_{I}|u|d\mu.

  3. (iii)

    The {Ij}\{I_{j}\} are pairwise almost disjoint.

  4. (iv)

    We have the estimate

    μ(Eβ)jμ(Ij)1βuLμ1()<.\mu(E_{\beta})\leq\sum\limits_{j}\mu(I_{j})\leq\frac{1}{\beta}\|u\|_{L^{1}_{\mu}(\mathbb{R})}<\infty\,. (53)

This is known as the Calderón-Zygmund decomposition of uu at level β\beta.

Proof.

Fix β(0,1)\beta\in(0,1), and by substituting uu with |u||u|, assume without loss of generality that u0u\geq 0. Since uLμ1()u\in L^{1}_{\mu}(\mathbb{R}), then for any interval II of size |I|1βu𝑑μ|I|\geq\frac{1}{\beta}\int\limits_{\mathbb{R}}u\,d\mu, we have uI,μβ\langle u\rangle_{I,\mu}\leq\beta, since μ(I)|I|\mu(I)\geq|I|. Thus any dyadic interval JJ has an ancestor I0JI_{0}\succeq J such that for all dyadics II0I\succeq I_{0}, we have uI,μβ\langle u\rangle_{I,\mu}\leq\beta. Meaning if we consider the set of dyadic intervals

=def{ dyadic:β<𝓊,μ},\cal{I}\stackrel{{\scriptstyle\rm def}}{{=}}\{I\text{ dyadic}~{}:~{}\beta<\langle u\rangle_{I,\mu}\}\,,

then for each II\in\cal{I}, there exists ImaxI_{\max}\in\cal{I} such that IImaxI\preceq I_{\max}, and ImaxI_{\max} is maximal in \cal{I} with respect to \preceq. Define {Ij}\{I_{j}\} to be the set of maximal dyadic intervals in (,)(\cal{I},\preceq). By definition, {Ij}\{I_{j}\} satisfies Property (ii), which then immediately implies Property (iii).

Meanwhile (53) follows from Property (i) and that β<1μ(Ij)Iju𝑑μ\beta<\frac{1}{\mu(I_{j})}\int\limits_{I_{j}}u\,d\mu:

μ(Eβ)jμ(Ij)j1βIj|u|𝑑μ1βuLμ1()<.\mu(E_{\beta})\leq\sum\limits_{j}\mu(I_{j})\leq\sum\limits_{j}\frac{1}{\beta}\int\limits_{I_{j}}|u|d\mu\leq\frac{1}{\beta}\|u\|_{L^{1}_{\mu}(\mathbb{R})}<\infty\,.

We are left with showing Property (i), which will follow if we can show uβu\leq\beta for almost every xx in (jIj)c(\bigcup\limits_{j}I_{j})^{c}. Consider the L1()L^{1}(\mathbb{R}) function v=defuλqv\stackrel{{\scriptstyle\rm def}}{{=}}u\langle\lambda\rangle^{q}. By the dyadic Lebesgue differentiation theorem, for almost every xx\in\mathbb{R}, we have limnvJn=v(x)\lim\limits_{n\to\infty}\langle v\rangle_{J_{n}}=v(x), where JnJ_{n} is any sequence of dyadic intervals shrinking to xx. Fix such a point x(jIj)cx\in(\bigcup\limits_{j}I_{j})^{c}. Then by construction of {Ij}\{I_{j}\}, for every dyadic interval JJ containing xx, it follows that

βuJ,μ=|J|μ(J)vJ.\beta\geq\langle u\rangle_{J,\mu}=\frac{|J|}{\mu(J)}\langle v\rangle_{J}\,.

Letting JJ shrink down to xx, the above yields

β1xqu(x)xq,\beta\geq\frac{1}{\langle x\rangle^{q}}u(x)\langle x\rangle^{q}\,,

i.e. βu(x)\beta\geq u(x). Since xx arbitrary in (jIj)cN(\bigcup\limits_{j}I_{j})^{c}\setminus N where NN is a set of measure 0, we get Property (i). ∎

Lemma 8.2 (Calderón-Zygmund decomposition for Lp1()+Lp2()L^{p_{1}}(\mathbb{R})+L^{p_{2}}(\mathbb{R})).

Let dμ=λqdλd\mu=\langle\lambda\rangle^{q}d\lambda for some q0q\geq 0, and suppose u=u1+u2u=u_{1}+u_{2} with u1Lμp1()u_{1}\in L^{p_{1}}_{\mu}(\mathbb{R}), u2Lμp2()u_{2}\in L^{p_{2}}_{\mu}(\mathbb{R}), where 1p1p2<1\leq p_{1}\leq p_{2}<\infty. For each β>0\beta>0, define Eβ=def{λ:|u|>β}E_{\beta}\stackrel{{\scriptstyle\rm def}}{{=}}\{\lambda\in\mathbb{R}~{}:~{}|u|>\beta\} and Eβ/2i=def{λ:|ui|>β2}E^{i}_{\beta/2}\stackrel{{\scriptstyle\rm def}}{{=}}\{\lambda~{}:~{}|u_{i}|>\frac{\beta}{2}\}, i=1,2i=1,2. Then EβEβ/21Eβ/22E_{\beta}\subset E_{\beta/2}^{1}\cup E_{\beta/2}^{2}, all of which are of finite μ\mu-measure, and there exists a collection of dyadic intervals {Ij}\{I_{j}\} with the following properties:

  1. (i)

    Eβ/21Eβ/22jIjE_{\beta/2}^{1}\cup E_{\beta/2}^{2}\subset\bigcup\limits_{j}I_{j}.

  2. (ii)

    The {Ij}\{I_{j}\} are pairwise almost disjoint.

  3. (iii)

    We have the estimate

    jμ(Ij)p1,p21βp2(βp2p1u1Lμp1()p1+u2Lμp2()p2).\sum\limits_{j}\mu(I_{j})\lesssim_{p_{1},p_{2}}\frac{1}{\beta^{p_{2}}}(\beta^{p_{2}-p_{1}}\|u_{1}\|_{L^{p_{1}}_{\mu}(\mathbb{R})}^{p_{1}}+\|u_{2}\|_{L^{p_{2}}_{\mu}(\mathbb{R})}^{p_{2}})\,. (54)
  4. (iv)

    We have

    |u|K,μβ,K any ancestor of Ij.\langle|u|\rangle_{K,\mu}\leq\beta\,,\quad K\text{ any ancestor of }I_{j}\,. (55)
Proof.

By the triangle inequality, EβEβ/21Eβ/22E_{\beta}\subset E_{\beta/2}^{1}\cup E_{\beta/2}^{2}. Thus μ(Eβ)μ(Eβ/21)+(Eβ/22)<\mu(E_{\beta})\leq\mu(E_{\beta/2}^{1})+(E_{\beta/2}^{2})<\infty, since uiLμpi()u_{i}\in L^{p_{i}}_{\mu}(\mathbb{R}).

For i=1,2i=1,2, we apply Lemma 8.1 to |ui|pi|u_{i}|^{p_{i}} at level (β2)pi\left(\frac{\beta}{2}\right)^{p_{i}} and get a collection of pairwise disjoint dyadic intervals {Iji}j0\{I_{j}^{i}\}_{j\geq 0} such that

Eβ/2ijIji.E^{i}_{\beta/2}\subset\bigcup\limits_{j}I_{j}^{i}\,.

By Lemma 8.1 (ii),

|ui|piIji,μ1/pi>β2.\langle|u_{i}|^{p_{i}}\rangle_{I_{j}^{i},\mu}^{1/p_{i}}>\frac{\beta}{2}\,.

Let {Ij˙}j={Ij1}j{Ij2}j\{\dot{I_{j}}\}_{j}=\{I_{j}^{1}\}_{j}\cup\{I_{j}^{2}\}_{j}. Note by maximality of the IjiI_{j}^{i}’s specified by Lemma 8.1 (ii), each Ij˙\dot{I_{j}} is contained in at most one other Ik˙\dot{I_{k}}. So now let {Ij}\{I_{j}\} be the maximal intervals among {Ij˙}\{\dot{I_{j}}\} with respect to \preceq; maximality ensures the {Ij}\{I_{j}\} are pairwise almost disjoint and so (ii) holds. Note maximality also yields

EβEβ/21Eβ/22(jIj1)(jIj2)jIj,E_{\beta}\subset E_{\beta/2}^{1}\cup E_{\beta/2}^{2}\subset\left(\bigcup\limits_{j}I_{j}^{1}\right)\cup\left(\bigcup\limits_{j}I_{j}^{2}\right)\subset\bigcup\limits_{j}I_{j}\,,

i.e.  (i) holds. And (53) yields (54), as

μ(Ij)jμ(Ij1)+jμ(Ij2)p1,p21βp1u1Lμp1()p1+1βp2u2Lμp2()p2=1βp2(βp2p1u1Lμp1()p1+u2Lμp2()p2).\sum\limits\mu(I_{j})\leq\sum\limits_{j}\mu(I_{j}^{1})+\sum\limits_{j}\mu(I_{j}^{2})\lesssim_{p_{1},p_{2}}\frac{1}{\beta^{p_{1}}}\|u_{1}\|_{L^{p_{1}}_{\mu}(\mathbb{R})}^{p_{1}}+\frac{1}{\beta^{p_{2}}}\|u_{2}\|_{L^{p_{2}}_{\mu}(\mathbb{R})}^{p_{2}}=\frac{1}{\beta^{p_{2}}}(\beta^{p_{2}-p_{1}}\|u_{1}\|_{L^{p_{1}}_{\mu}(\mathbb{R})}^{p_{1}}+\|u_{2}\|_{L^{p_{2}}_{\mu}(\mathbb{R})}^{p_{2}})\,.

Also note by our choice of IjI_{j} as the maximal intervals among {Ij1}j{Ij2}j\{I_{j}^{1}\}_{j}\cup\{I_{j}^{2}\}_{j}, then for any ancestor KK of IjI_{j}, it follows that

|u1|p1K,μ1/p1,|u2|p2K,μ1/p2β2.\langle|u_{1}|^{p_{1}}\rangle_{K,\mu}^{1/p_{1}}\,,\,\langle|u_{2}|^{p_{2}}\rangle_{K,\mu}^{1/p_{2}}\leq\frac{\beta}{2}\,.

The triangle inequality followed by Hölder’s inequality then imply (55), i.e.

|u|K,μ|u1|K,μ+|u2|K,μ|u1|p1K,μ1/p1+|u2|p2K,μ1/p2β.\langle|u|\rangle_{K,\mu}\leq\langle|u_{1}|\rangle_{K,\mu}+\langle|u_{2}|\rangle_{K,\mu}\leq\langle|u_{1}|^{p_{1}}\rangle_{K,\mu}^{1/p_{1}}+\langle|u_{2}|^{p_{2}}\rangle_{K,\mu}^{1/p_{2}}\leq\beta\,.

We also need the following additional lemma. Recall that for an interval II, we defined DI=dist(I,0)D_{I}=\mathrm{dist}\left(I,0\right).

Lemma 8.3 (μ\mu is flat away from 0).

Let dμ=λqdλd\mu=\langle\lambda\rangle^{q}d\lambda for some q0q\geq 0, and suppose II is an interval such that μ(I)L\mu(I)\leq L, where LL is some constant. Then there exists D=D(L,q)0D=D(L,q)\geq 0 sufficiently large so that if DIDD_{I}\geq D, then

λqqDIq,λ in I or either of its neighbors,\langle\lambda\rangle^{q}\sim_{q}\langle D_{I}\rangle^{q}\,,\quad\lambda\text{ in }I\text{ or either of its neighbors}\,, (56)

In particular,

fJ,μqfJ\langle f\rangle_{J,\mu}\sim_{q}\langle f\rangle_{J} (57)

for all f0f\geq 0, where JJ may equal II, either of its neighbors, or its parent in the case that II is dyadic.

Proof.

If q=0q=0, the lemma is trivial.

For q>0q>0, choose DD so large that Dq100L\langle D\rangle^{q}\geq 100L. Thus if DIDD_{I}\geq D, then

100L|I||I|Dq|I|DIqμ(I)L100L|I|\leq|I|\langle D\rangle^{q}\leq|I|\langle D_{I}\rangle^{q}\leq\mu(I)\leq L

and so |I|<1100|I|<\frac{1}{100}. Thus on II or either of its neighbors, we have λqqDIq\langle\lambda\rangle^{q}\sim_{q}\langle D_{I}\rangle^{q}. Hence for J=IJ=I or either of its neighbors, (57) holds as

fJ,μ=1μ(J)Jf𝑑μ=1μ(J)Jf(λ)λq𝑑λq1DIq|J|Jf(λ)DIq𝑑λ=fJ.\langle f\rangle_{J,\mu}=\frac{1}{\mu(J)}\int\limits_{J}fd\mu=\frac{1}{\mu(J)}\int\limits_{J}f(\lambda)\langle\lambda\rangle^{q}d\lambda\sim_{q}\frac{1}{\langle D_{I}\rangle^{q}|J|}\int\limits_{J}f(\lambda)\langle D_{I}\rangle^{q}d\lambda=\langle f\rangle_{J}\,.

If II is dyadic, then since the parent of II is the union of II and one of its neighbors, it follows λqqDIq\langle\lambda\rangle^{q}\sim_{q}\langle D_{I}\rangle^{q} on the parent of II. Then the same computations as done previously yield (57) when JJ is the parent of II. ∎

Proof that w1/p~w1/p~Lμp()<\|w^{1/\widetilde{p}}-w^{-1/\widetilde{p}^{\prime}}\|_{L^{p}_{\mu}(\mathbb{R})}<\infty in Lemma 5.3.

Let p~(1,)\widetilde{p}\in(1,\infty) and pp2p\geq p_{2}. Let’s apply Lemma 8.2 with u=w1u=w-1 and β=α\beta=\alpha, where α(0,12)\alpha\in(0,\frac{1}{2}), so that Eα=def{λ:|w(λ)1|>α}E_{\alpha}\stackrel{{\scriptstyle\rm def}}{{=}}\{\lambda~{}:~{}|w(\lambda)-1|>\alpha\} and Eα/2i=def{λ:|ui(λ)|>α/2}E^{i}_{\alpha/2}\stackrel{{\scriptstyle\rm def}}{{=}}\{\lambda~{}:~{}|u_{i}(\lambda)|>\alpha/2\}. Then EαEα/21Eα/22E_{\alpha}\subset E_{\alpha/2}^{1}\cup E_{\alpha/2}^{2}.

We begin by splitting

w1/p~w1/p~Lμp()p=w1/p~w1/p~Lμp((Eα/21Eα/22)c)p+w1/p~w1/p~Lμp(Eα/21Eα/22)p=defA+B.\|w^{1/\widetilde{p}}-w^{-1/\widetilde{p}^{\prime}}\|_{L^{p}_{\mu}(\mathbb{R})}^{p}=\|w^{1/\widetilde{p}}-w^{-1/\widetilde{p}^{\prime}}\|_{L^{p}_{\mu}((E_{\alpha/2}^{1}\cup E_{\alpha/2}^{2})^{c})}^{p}+\|w^{1/\widetilde{p}}-w^{-1/\widetilde{p}^{\prime}}\|_{L^{p}_{\mu}(E_{\alpha/2}^{1}\cup E_{\alpha/2}^{2})}^{p}\stackrel{{\scriptstyle\rm def}}{{=}}A+B\,. (58)

Since (Eα/21Eα/22)cEαc={|w1|α}(E_{\alpha/2}^{1}\cup E_{\alpha/2}^{2})^{c}\subset E_{\alpha}^{c}=\{|w-1|\leq\alpha\}, we may estimate

A=(Eα/21Eα/22)cwp/p~|w1|p𝑑μ(1α)p/p~(Eα/21)c(Eα/22)c|w1|p𝑑μ.A=\int\limits_{(E_{\alpha/2}^{1}\cup E_{\alpha/2}^{2})^{c}}w^{-p/\widetilde{p}^{\prime}}|w-1|^{p}\,d\mu\leq(1-\alpha)^{-p/\widetilde{p}^{\prime}}\int\limits_{(E_{\alpha/2}^{1})^{c}\cap(E_{\alpha/2}^{2})^{c}}|w-1|^{p}\,d\mu\,.

By the triangle inequality, AA is

p,p~(Eα/21)c(Eα/22)c|u1|p𝑑μ+(Eα/21)c(Eα/22)c|u2|p𝑑μp1,p2αpp1(Eα/21)c|u1|p1𝑑μ+αpp2(Eα/22)c|u2|p2𝑑μ\lesssim_{p,\widetilde{p}}\int\limits_{(E_{\alpha/2}^{1})^{c}\cap(E_{\alpha/2}^{2})^{c}}|u_{1}|^{p}\,d\mu+\int\limits_{(E_{\alpha/2}^{1})^{c}\cap(E_{\alpha/2}^{2})^{c}}|u_{2}|^{p}\,d\mu\,\lesssim_{p_{1},p_{2}}\alpha^{p-p_{1}}\int\limits_{(E_{\alpha/2}^{1})^{c}}|u_{1}|^{p_{1}}\,d\mu+\alpha^{p-p_{2}}\int\limits_{(E_{\alpha/2}^{2})^{c}}|u_{2}|^{p_{2}}\,d\mu

Thus,

Ap,p~,p1,p2αpp2(αp2p1u1Lμp1()p1+u2Lμp2()p2).A\lesssim_{p,\widetilde{p},p_{1},p_{2}}\alpha^{p-p_{2}}(\alpha^{p_{2}-p_{1}}\|u_{1}\|_{L^{p_{1}}_{\mu}(\mathbb{R})}^{p_{1}}+\|u_{2}\|_{L^{p_{2}}_{\mu}(\mathbb{R})}^{p_{2}})\,. (59)

By the triangle inequality followed by Lemma 8.2 (i),

BpEα/21Eα/22wp/p~𝑑μ+Eα/21Eα/22wp/p~𝑑μ(jIjwp/p~𝑑μ)+(jIjwp/p~𝑑μ)=defB1+B2.B\lesssim_{p}\int\limits_{E_{\alpha/2}^{1}\cup E_{\alpha/2}^{2}}w^{p/\widetilde{p}}d\mu+\int\limits_{E_{\alpha/2}^{1}\cup E_{\alpha/2}^{2}}w^{-p/\widetilde{p}^{\prime}}d\mu\leq\left(\sum\limits_{j}\int\limits_{I_{j}}w^{p/\widetilde{p}}d\mu\right)+\left(\sum\limits_{j}\int\limits_{I_{j}}w^{-p/\widetilde{p^{\prime}}}d\mu\right)\stackrel{{\scriptstyle\rm def}}{{=}}B_{1}+B_{2}\,. (60)

Recall wAp()w\in A_{p^{\prime}}(\mathbb{R}) and so wp/pAp()w^{-p/p^{\prime}}\in A_{p}(\mathbb{R}). By applying Lemma 5.1 to both of these weights, we can choose ϵ(γ,p)(0,12)\epsilon(\gamma,p)\in(0,\frac{1}{2}), with limγ1ϵ(γ,p)=min{1p,1p}\lim\limits_{\gamma\to 1}\epsilon(\gamma,p)=\min\{\frac{1}{p},\frac{1}{p^{\prime}}\}, so that whenever p~\widetilde{p} is chosen so that |1p~1p|<ϵ(γ,p)|\frac{1}{\widetilde{p}}-\frac{1}{p}|<\epsilon(\gamma,p), we have

wp/p~Iγ,pwIp/p~,wp/p~Iγ,pwp/pIp/p~,for all intervals I.\langle w^{p/\widetilde{p}}\rangle_{I}\lesssim_{\gamma,p}\langle w\rangle_{I}^{p/\widetilde{p}}\,,\quad\langle w^{-p/\widetilde{p^{\prime}}}\rangle_{I}\lesssim_{\gamma,p}\langle w^{-p/p^{\prime}}\rangle_{I}^{p^{\prime}/\widetilde{p^{\prime}}}\,,\quad\text{for all intervals }I\,. (61)

In particular, this means wp/p~w^{p/\widetilde{p}} and wp/p~w^{-p/\widetilde{p}^{\prime}} are locally integrable.

Let us focus on B2B_{2} first, as B1B_{1} will be similar. Set Dj=defdist(Ij,0)D_{j}\stackrel{{\scriptstyle\rm def}}{{=}}\mathrm{dist}(I_{j},0), then write

jIjwp/p~𝑑μ=j:Dj<DIjwp/p~𝑑μ+j:DjDIjwp/p~𝑑μ,\sum\limits_{j}\int\limits_{I_{j}}w^{-p/\widetilde{p^{\prime}}}d\mu=\sum\limits_{j\,:\,D_{j}<D}\int\limits_{I_{j}}w^{-p/\widetilde{p^{\prime}}}d\mu+\sum\limits_{j\,:\,D_{j}\geq D}\int\limits_{I_{j}}w^{-p/\widetilde{p^{\prime}}}d\mu\,,

where D>0D>0 is a constant which will be chosen later.

Since |Ij|μ(Ij)p1,p21αp2(βp2p1u1Lμp1()p1+u2Lμp2()p2)|I_{j}|\leq\mu(I_{j})\lesssim_{p_{1},p_{2}}\frac{1}{\alpha^{p_{2}}}(\beta^{p_{2}-p_{1}}\|u_{1}\|_{L^{p_{1}}_{\mu}(\mathbb{R})}^{p_{1}}+\|u_{2}\|_{L^{p_{2}}_{\mu}(\mathbb{R})}^{p_{2}}) by (54), then the length of each interval IjI_{j} is bounded. Since wp/p~w^{-p/\widetilde{p}^{\prime}} is locally integrable, then

j:Dj<DIjwp/p~𝑑μI~wp/p~λq𝑑λ<,\sum\limits_{j\,:\,D_{j}<D}\int\limits_{I_{j}}w^{-p/\widetilde{p}^{\prime}}d\mu\leq\int\limits_{\widetilde{I}}w^{-p/\widetilde{p}^{\prime}}\langle\lambda\rangle^{q}\,d\lambda<\infty\,,

where I~\widetilde{I} is some sufficiently large interval centered at 0. And so it suffices to show

j:DjDIjwp/p~𝑑μ=j:DjDμ(Ij)wp/p~Ij,μ<.\sum\limits_{j\,:\,D_{j}\geq D}\int\limits_{I_{j}}w^{-p/\widetilde{p}^{\prime}}d\mu=\sum\limits_{j\,:\,D_{j}\geq D}\mu(I_{j})\langle w^{-p/\widetilde{p}^{\prime}}\rangle_{I_{j},\mu}<\infty\,.

Since jμ(Ij)<\sum_{j}\mu(I_{j})<\infty, to show B2<B_{2}<\infty, it suffices to show wp/p~Ij,μp,γ,q1\langle w^{-p/\widetilde{p}^{\prime}}\rangle_{I_{j},\mu}\lesssim_{p,\gamma,q}1; similarly B1<B_{1}<\infty will follow from showing wp/p~Ij,μp,γ1\langle w^{p/\widetilde{p}}\rangle_{I_{j},\mu}\lesssim_{p,\gamma}1 for a suitable choice of DD.

Since there exists some constant C=C(p1,p2)C=C(p_{1},p_{2}) such that

μ(Ij)C1αp2(αp2p1u1Lμp1()p1+u2Lμp2()p2),\mu(I_{j})\leq C\frac{1}{\alpha^{p_{2}}}\left(\alpha^{p_{2}-p_{1}}\|u_{1}\|_{L^{p_{1}}_{\mu}(\mathbb{R})}^{p_{1}}+\|u_{2}\|_{L^{p_{2}}_{\mu}(\mathbb{R})}^{p_{2}}\right)\,,

we can take D=D(C1αp2(αp2p1u1Lμp1()p1+u2Lμp2()p2),q)D=D(C\frac{1}{\alpha^{p_{2}}}(\alpha^{p_{2}-p_{1}}\|u_{1}\|_{L^{p_{1}}_{\mu}(\mathbb{R})}^{p_{1}}+\|u_{2}\|_{L^{p_{2}}_{\mu}(\mathbb{R})}^{p_{2}}),q) as in Lemma 8.3 so that when DjDD_{j}\geq D, we have λqqDjq\langle\lambda\rangle^{q}\sim_{q}\langle D_{j}\rangle^{q} on IjI_{j}, its neighbors and its parent KjK_{j}.

For such an interval IjI_{j}, recall (55) implies

|w1Kj,μ||w1|Kj,μα12,|\langle w-1\rangle_{K_{j},\mu}|\leq\langle|w-1|\rangle_{K_{j},\mu}\leq\alpha\leq\frac{1}{2}\,,

where KjK_{j} is the parent of IjI_{j}. And so λqqDjq\langle\lambda\rangle^{q}\sim_{q}\langle D_{j}\rangle^{q} on KjK_{j} and (57) together imply

wKjwKj,μ=1+w1Kj,μ1.\langle w\rangle_{K_{j}}\sim\langle w\rangle_{K_{j},\mu}=1+\langle w-1\rangle_{K_{j},\mu}\sim 1\,.

In particular,

wIj,μqwIjwKj1\langle w\rangle_{I_{j},\mu}\sim_{q}\langle w\rangle_{I_{j}}\lesssim\langle w\rangle_{K_{j}}\sim 1\,

and so (61) implies

wp/p~Ij,μqwp/p~Ijγ,pwIjp/p~q,p,γwIj,μp/p~p,γ,q1.\langle w^{p/\widetilde{p}}\rangle_{I_{j},\mu}\sim_{q}\langle w^{p/\widetilde{p}}\rangle_{I_{j}}\lesssim_{\gamma,p}\langle w\rangle_{I_{j}}^{p/\widetilde{p}}\sim_{q,p,\gamma}\langle w\rangle_{I_{j},\mu}^{p/\widetilde{p}}\lesssim_{p,\gamma,q}1\,.

Similarly, since wp/pAp()w^{-p/p^{\prime}}\in A_{p}(\mathbb{R}), then

wp/pIj,μqwp/pIjwp/pKj[wp/p]Ap()wKjp/pp,γ1\langle w^{-p/p^{\prime}}\rangle_{I_{j},\mu}\sim_{q}\langle w^{-p/p^{\prime}}\rangle_{I_{j}}\lesssim\langle w^{-p/p^{\prime}}\rangle_{K_{j}}\leq[w^{-p/p^{\prime}}]_{A_{p}(\mathbb{R})}\langle w\rangle_{K_{j}}^{-p^{\prime}/p}\lesssim_{p,\gamma}1\,

and therefore (61) implies

wp/p~Ij,μ,wp/p~Ij,μp,γ1.\langle w^{-p/\widetilde{p}^{\prime}}\rangle_{I_{j},\mu}\,,\quad\langle w^{p/\widetilde{p}}\rangle_{I_{j},\mu}\lesssim_{p,\gamma}1\,.

For the proof of (38), we need to understand how [w]A()[w]_{A_{\infty}(\mathbb{R})} affects logw\log w.

Lemma 8.4.

Suppose wA()w\in A_{\infty}(\mathbb{R}) and let JJ be an interval. If |w1|J12\langle|w-1|\rangle_{J}\leq\frac{1}{2}, then

|logwJ||w1|J+log[w]A().|\langle\log w\rangle_{J}|\lesssim\langle|w-1|\rangle_{J}+\log[w]_{A_{\infty}(\mathbb{R})}\,.
Proof.

By Jensen’s inequality,

logwJlogwJ=log(w1J+1)log(|w1|J+1)|w1|J.\langle\log w\rangle_{J}\leq\log\,\langle w\rangle_{J}=\log(\langle w-1\rangle_{J}+1)\leq\log(\langle|w-1|\rangle_{J}+1)\lesssim\langle|w-1|\rangle_{J}\,.

But also since [w]A()<[w]_{A_{\infty}(\mathbb{R})}<\infty, then

exp(logwJ)[w]A()wJ[w]A()1|w1|J,\mathrm{exp}\left(-\langle\log w\rangle_{J}\right)\leq\frac{[w]_{A_{\infty}(\mathbb{R})}}{\langle w\rangle_{J}}\leq\frac{[w]_{A_{\infty}(\mathbb{R})}}{1-\langle|w-1|\rangle_{J}}\,,

which implies

logwJlog(1|w1|J)log[w]A().\langle\log w\rangle_{J}\geq\log(1-\langle|w-1|\rangle_{J})-\log[w]_{A_{\infty}(\mathbb{R})}\,.

Combine the estimates of logwJ\langle\log w\rangle_{J} from above and below to get

|logwJ||w1|J+|log(1|w1|J)|+log[w]A()|w1|J+log[w]A().|\langle\log w\rangle_{J}|\lesssim\langle|w-1|\rangle_{J}+|\log(1-\langle|w-1|\rangle_{J})|+\log[w]_{A_{\infty}(\mathbb{R})}\lesssim\langle|w-1|\rangle_{J}+\log[w]_{A_{\infty}(\mathbb{R})}\,.

Next, we recall a few facts about BMO\rm{BMO}, the space of functions with bounded mean oscillation, and how it relates to Ap()A_{p}(\mathbb{R}) weights. Recall that fBMO(d)f\in{\rm BMO}(\mathbb{R}^{d}) if

fBMO(d)=defsupB|ffB|B<,\|f\|_{{\rm BMO}(\mathbb{R}^{d})}\stackrel{{\scriptstyle\rm def}}{{=}}\sup_{B}\,\langle|f-\langle f\rangle_{B}|\rangle_{B}<\infty\,,

where BB denotes a ball in d\mathbb{R}^{d} (see, e.g., p.140 in [13]). Functions in BMO(d)\rm BMO(\mathbb{R}^{d}) all satisfy the John-Nirenberg estimates below.

Theorem 8.5 (John-Nirenberg, [13, p.144-146]).

Suppose fBMO(d)f\in\mathrm{BMO}(\mathbb{R}^{d}). Then

  1. (a)

    There exist positive absolute constants c1,c2c_{1},c_{2} such that for each α>0\alpha>0 and every ball BB,

    1|B||{xB:|f(x)fB|>α}|c1exp(c2α/fBMO).\frac{1}{|B|}|\{x\in B~{}:~{}|f(x)-\langle f\rangle_{B}|>\alpha\}|\leq c_{1}\mathrm{exp}\left(-c_{2}\alpha/\|f\|_{\mathrm{BMO}}\right)\,.
  2. (b)

    For any p<p<\infty, fLlocp(d)f\in L^{p}_{loc}(\mathbb{R}^{d}) and

    |ffB|pBpfBMO(d)p\langle|f-\langle f\rangle_{B}|^{p}\rangle_{B}\lesssim_{p}\|f\|_{\mathrm{BMO}(\mathbb{R}^{d})}^{p}\,

    for all balls BB.

  3. (c)

    If 0μfBMO(d)10\leq\mu\|f\|_{\mathrm{BMO}(\mathbb{R}^{d})}\lesssim 1, then

    exp(μ|ffB|)B1.\langle\mathrm{exp}\left(\mu|f-\langle f\rangle_{B}|\right)\rangle_{B}\lesssim 1\,.

It is well known that if wAp(d)w\in A_{p}(\mathbb{R}^{d}), then logwBMO(d)\log w\in\mathrm{BMO}(\mathbb{R}^{d}). We also have the following well-known quantification (see, e.g., [7, Corollary 6]).

Lemma 8.6.

If wAp()w\in A_{p}(\mathbb{R}) for some p(1,]p\in(1,\infty], then logwBMOlog(2[w]Ap(d))\|\log w\|_{\rm BMO}\leq\log(2[w]_{A_{p}(\mathbb{R}^{d})}). If in addition [w]Ap(d)=1+τ,τ[0,1][w]_{A_{p}(\mathbb{R}^{d})}=1+\tau,\tau\in[0,1], then

logwBMOτ.\|\log w\|_{\rm BMO}\lesssim\sqrt{\tau}\,.
Proof of (38).

By Lemma 8.6, when τ0(p)12\tau_{0}(p)\leq\frac{1}{2} we have logwBMO()τ1/2\|\log w\|_{\mathrm{BMO}(\mathbb{R})}\lesssim\tau^{1/2}. Take τ0(p)\tau_{0}(p) small enough so that additionally logwBMO()110\|\log w\|_{\mathrm{BMO}(\mathbb{R})}\leq\frac{1}{10}.

Set α=defτ1/2\alpha\stackrel{{\scriptstyle\rm def}}{{=}}\tau^{1/2} and define EαE_{\alpha}, Eα/2iE_{\alpha/2}^{i} and {Ij}\{I_{j}\} as in the proof that w1/p~w1/p~Lμp()<\|w^{1/\widetilde{p}}-w^{-1/\widetilde{p}^{\prime}}\|_{L^{p}_{\mu}(\mathbb{R})}<\infty, i.e. by applying Lemma 8.2 with u=|w1|u=|w-1| and β=α\beta=\alpha. We begin by splitting w1/pw1/pLp()p\|w^{1/p}-w^{-1/p^{\prime}}\|_{L^{p}(\mathbb{R})}^{p} as in (58), with the same definitions of AA and BB. By (59) and our choice of α\alpha, it follows that

Ap,p1,p2τ(pp2)/2(τp2p12u1Lp1()p1+u2Lp2()p2).A\lesssim_{p,p_{1},p_{2}}\tau^{(p-p_{2})/2}(\tau^{\frac{p_{2}-p_{1}}{2}}\|u_{1}\|_{L^{p_{1}}(\mathbb{R})}^{p_{1}}+\|u_{2}\|_{L^{p_{2}}(\mathbb{R})}^{p_{2}})\,.

As for estimating BB, we split slightly differently than in (60):

Bpw1/p1Lp(Eα/21Eα/22)p+w1/p1Lp(Eα/21Eα/22)p=defB1+B2.B\lesssim_{p}\|w^{1/p}-1\|_{L^{p}(E_{\alpha/2}^{1}\cup E_{\alpha/2}^{2})}^{p}+\|w^{-1/p^{\prime}}-1\|_{L^{p}(E_{\alpha/2}^{1}\cup E_{\alpha/2}^{2})}^{p}\stackrel{{\scriptstyle\rm def}}{{=}}B_{1}+B_{2}\,.

It suffices to show

B1,B2p,p1,p2τ(pp2)/2(τp2p12u1Lp1()p1+u2Lp2()p2).B_{1},B_{2}\lesssim_{p,p_{1},p_{2}}\tau^{(p-p_{2})/2}(\tau^{\frac{p_{2}-p_{1}}{2}}\|u_{1}\|_{L^{p_{1}}(\mathbb{R})}^{p_{1}}+\|u_{2}\|_{L^{p_{2}}(\mathbb{R})}^{p_{2}})\,. (62)

We prove this for B1B_{1}; the estimate for B2B_{2} will follow similarly.

By Lemma 8.2 (i),

B1jw1/p1Lp(Ij)p.B_{1}\leq\sum\limits_{j}\|w^{1/p}-1\|_{L^{p}(I_{j})}^{p}\,.

Note that (62) will follow if we can show

w1/p1Lp(I)pp|I|τp/2\|w^{1/p}-1\|_{L^{p}(I)}^{p}\lesssim_{p}|I|\tau^{p/2} (63)

for all I{Ij}I\in\{I_{j}\}: indeed, sum over I=IjI=I_{j} and use (54) to get (62). As such, fix I{Ij}I\in\{I_{j}\}; our goal is to show (63).

Now note

|logwI|τ1/2.|\langle\log w\rangle_{I}|\lesssim\tau^{1/2}\,. (64)

Indeed, if KK is the dyadic parent of II, then (55) guarantees |w1|I|w1|Kα\langle|w-1|\rangle_{I}\lesssim\langle|w-1|\rangle_{K}\leq\alpha. By Lemma 8.4, we get (64) for τ0(p)\tau_{0}(p) sufficiently small, which we will make use of repeatedly.

Split the left side of (63) into

w1/p1Lp(I)pw1/pexp(logwI/p)Lp(I)p+exp(logwI/p)1Lp(I)p=defB11+B12.\|w^{1/p}-1\|_{L^{p}(I)}^{p}\leq\|w^{1/p}-\mathrm{exp}\left(\langle\log w\rangle_{I}/p\right)\|_{L^{p}(I)}^{p}+\|\mathrm{exp}\left(\langle\log w\rangle_{I}/p\right)-1\|_{L^{p}(I)}^{p}\stackrel{{\scriptstyle\rm def}}{{=}}B_{11}+B_{12}\,.

By (64) it follows that

B12=|I||exp(logwI/p)1|pp|I|τp/2.B_{12}=|I||\mathrm{exp}\left(\langle\log w\rangle_{I}/p\right)-1|^{p}\lesssim_{p}|I|\tau^{p/2}\,.

Thus it suffices to show B11p|I|τp/2B_{11}\lesssim_{p}|I|\tau^{p/2}.

But B11B_{11} equals

exp(logwI/p)exp((logwlogwI)/p)1Lp(I)pexp(f/p)1Lp(I)p,\mathrm{exp}\left(\langle\log w\rangle_{I}/p\right)\|\mathrm{exp}\left((\log w-\langle\log w\rangle_{I})/p\right)-1\|_{L^{p}(I)}^{p}\lesssim\|\mathrm{exp}\left(f/p\right)-1\|_{L^{p}(I)}^{p}\,,

where f=deflogwlogwIf\stackrel{{\scriptstyle\rm def}}{{=}}\log w-\langle\log w\rangle_{I} and we applied (64) in the inequality. Write

exp(f/p)1=01/pexp(sf)f𝑑s\mathrm{exp}\left(f/p\right)-1=\int\limits_{0}^{1/p}\mathrm{exp}\left(sf\right)f\,ds\,

and apply Minkowski’s inequality, followed by the Cauchy-Schwarz inequality, to get

exp(f/p)1Lp(I)fexp(|f|/p)Lp(I)fL2p(I)exp(|f|/p)L2p(I).\|\mathrm{exp}\left(f/p\right)-1\|_{L^{p}(I)}\lesssim\|f\mathrm{exp}\left(|f|/p\right)\|_{L^{p}(I)}\leq\|f\|_{L^{2p}(I)}\|\mathrm{exp}\left(|f|/p\right)\|_{L^{2p}(I)}\,.

The John-Nirenberg Theorem 8.5 yields

fL2p(I)plogwBMO()|I|12p,exp(|f|/p)L2p(I)p|I|12p\|f\|_{L^{2p}(I)}\lesssim_{p}\|\log w\|_{\mathrm{BMO}(\mathbb{R})}|I|^{\frac{1}{2p}}\,,\quad\|\mathrm{exp}\left(|f|/p\right)\|_{L^{2p}(I)}\lesssim_{p}|I|^{\frac{1}{2p}}

for logwBMO()\|\log w\|_{\mathrm{BMO}(\mathbb{R})} small enough, which we can arrange by taking τ0(p)\tau_{0}(p) as small as necessary and applying Lemma 8.6. Thus

exp(f/p)1Lp(I)\displaystyle\|\mathrm{exp}\left(f/p\right)-1\|_{L^{p}(I)} p|I|1/plogwBMO()|I|1/pτ1/2.\displaystyle\lesssim_{p}|I|^{1/p}\|\log w\|_{\mathrm{BMO}(\mathbb{R})}\lesssim|I|^{1/p}\tau^{1/2}\,.

Hence B11p|I|1/pτ1/2B_{11}\lesssim_{p}|I|^{1/p}\tau^{1/2}, which completes the proof. ∎

9. Inverting IQw,pI-Q_{w,p}: proof of Lemma 5.4

Let us introduce some notation and definitions.

Notation.  Given z=z=\mathbb{C}, we will generally write z=t+iyz=t+iy, where t,yt,y are real-valued.

Given KK\subset\mathbb{C}, we let N(K)N(K) denote an open set containing KK, although we allow for the particular set to change line to line.

Definition.  Given some p[1,]p_{*}\in[1,\infty], define

1p(z)=defzp+1z2,1p(z)=def11p(z)=1+z2zp,z=t+iy:1t1.\frac{1}{p(z)}\stackrel{{\scriptstyle\rm def}}{{=}}\frac{z}{p_{*}}+\frac{1-z}{2},\quad\frac{1}{p^{\prime}(z)}\stackrel{{\scriptstyle\rm def}}{{=}}1-\frac{1}{p(z)}=\frac{1+z}{2}-\frac{z}{p_{*}},\quad z=t+iy\,:-1\leq t\leq 1\,. (65)

Given ϵ0\epsilon\geq 0, define the open strip

Ω<ϵ=def{z:|1p(t)12|<ϵ}.\Omega_{<\epsilon}\stackrel{{\scriptstyle\rm def}}{{=}}\left\{z~{}:~{}\left|\frac{1}{p(t)}-\frac{1}{2}\right|<\epsilon\right\}\,.

We define the closed strip Ωϵ\Omega_{\leq\epsilon} similarly.

Given II an interval, let

ΩI=def{z:tI}.\Omega_{I}\stackrel{{\scriptstyle\rm def}}{{=}}\{z\in\mathbb{C}~{}:~{}t\in I\}\,.
Lemma 9.1.

Let 𝚲>1{\bf\Lambda}>1 be an arbitrary absolute constant, and let p(z)p(z) is as in (65) for some p[1,]p_{*}\in[1,\infty]. For [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma, consider the operator

Qw,p(z),T=defw1/p(z)Tw1/p(z)w1/p(z)Tw1/p(z),Q_{w,p(z),T}\stackrel{{\scriptstyle\rm def}}{{=}}w^{1/p(z)}Tw^{-1/p(z)}-w^{-1/p^{\prime}(z)}Tw^{1/p^{\prime}(z)}\,,

where TT is an operator satisfying (28) for some \cal{F} and is also self-adjoint as an operator on L2()L^{2}(\mathbb{R}).

Then there exists ϵ=ϵ(γ,,𝚲)(0,12)\epsilon=\epsilon(\gamma,\cal{F},{\bf\Lambda})\in(0,\frac{1}{2}), with limγ1ϵ(γ,,𝚲)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma,\cal{F},{\bf\Lambda})=\frac{1}{2}, such that on Ω<ϵ\Omega_{<\epsilon}, IQw,p(z)I-Q_{w,p(z)} has bounded inverse on Lp(t)()L^{p(t)}(\mathbb{R}) with operator bound

(IQw,p(z),T)1p(t),p(t)2𝚲,\|(I-Q_{w,p(z),T})^{-1}\|_{p(t),p(t)}\leq 2\bf{\Lambda}\,, (66)

and (IQw,p(z),T)1(I-Q_{w,p(z),T})^{-1} is analytic as a map Ω<ϵ(2())\Omega_{<\epsilon}\to\cal{L}(L^{2}(\mathbb{R})).

Lemma 5.4 follows from Lemma 9.1 by taking T=𝒫[0,r]T=\mathcal{P}_{[{0},{r}]}, 𝚲=10\mathbf{\Lambda}=10 and applying Lemma 4.3. As such, this section is dedicated to the proof of Lemma 9.1. The proof strategy will proceed more or less as follows:

  1. (1)

    Show Qw,p(z),Tp(t),p(t)\|Q_{w,p(z),T}\|_{p(t),p(t)} uniformly bounded on the strip.

  2. (2)

    Show that if we have uniform bounds on (IκQw,p(z),T)1p(t),p(t)\|(I-\kappa Q_{w,p(z),T})^{-1}\|_{p(t),p(t)} is uniformly bounded on the strip, then (IκQw,p(z),T)1(I-\kappa Q_{w,p(z),T})^{-1} is weakly analytic.

  3. (3)

    Chop (IκQw,p(z),T)1I(I-\kappa Q_{w,p(z),T})^{-1}-I into small pieces and show we have uniform bounds for II plus the small piece; using the two previous parts of the strategy, we have our function is weakly analytic. We maintain boundedness while adding the small pieces by applying analytic interpolation.

Before we begin, we will need a few preliminary lemmas.

Proposition 9.2.

Suppose XX is an Banach space and H,VH,V are linear bounded operators from XX to XX. Then,

(I+H+V)1=(I+H)1(I+H+V)1V(I+H)1,\displaystyle(I+H+V)^{-1}=(I+H)^{-1}-(I+H+V)^{-1}V(I+H)^{-1},
(I+H+V)1=(I+H)1(I+V(I+H)1)1,\displaystyle(I+H+V)^{-1}=(I+H)^{-1}(I+V(I+H)^{-1})^{-1}\,,

provided the operators involved are well-defined and bounded in XX. Moreover, assuming V(I+H)1<1\|V\|\cdot\|(I+H)^{-1}\|<1, we get

(I+H+V)1(I+H)11V(I+H)1.\|(I+H+V)^{-1}\|\leqslant\frac{\|(I+H)^{-1}\|}{1-\|V\|\cdot\|(I+H)^{-1}\|}\,. (67)

Finally, if V<1\|V\|<1, then

(I+V)111V.\|(I+V)^{-1}\|\leqslant\frac{1}{1-\|V\|}\,. (68)

The proof of this proposition is a straightforward calculation.

We will also need continuity of weighted operators as proved in [1, Theorem 1.2].

Theorem 9.3 ([1, Theorem 1.2]).

Suppose p(1,)p\in(1,\infty), [w]Ap(d)<[w]_{A_{p}(\mathbb{R}^{d})}<\infty, fBMO<\|f\|_{\rm BMO}<\infty, and TT satisfies (28). Consider wδ=weδfw_{\delta}=we^{\delta f}. Then, there is δ0(p,[w]Ap,fBMO)>0\delta_{0}(p,[w]_{A_{p}},\|f\|_{\rm BMO})>0 such that

wδ1/pTwδ1/pw1/pTw1/pp,p<|δ|C(p,[w]Ap,fBMO,)\|w_{\delta}^{1/p}Tw_{\delta}^{-1/p}-w^{1/p}Tw^{-1/p}\|_{p,p}<|\delta|C(p,[w]_{A_{p}},\|f\|_{\rm BMO},\cal{F})

for all δ:|δ|<δ0\delta:|\delta|<\delta_{0}.

Definition.  Suppose 𝒟\cal{D} is a set of linear functionals acting on a vector space VV. If UU\subset\mathbb{C} is an open set, and {a(z)}zU\{a(z)\}_{z\in U} is contained within VV, then a(z)a(z) is weakly analytic with respect to 𝒟\cal{D} if (a(z))\ell(a(z)) is analytic for all 𝒟\ell\in\cal{D}.

We will generally be concerned with two cases:

  1. (1)

    V=p1Lp()V=\bigcup\limits_{p\geq 1}L^{p}(\mathbb{R}) for p(1,)p\in(1,\infty), and 𝒟=𝒮𝒸()\cal{D}=\cal{S}_{c}(\mathbb{R}), the space of simple functions with compact support, where g𝒮𝒸()g\in\cal{S}_{c}(\mathbb{R}) acts on fp1Lp()f\in\bigcup\limits_{p\geq 1}L^{p}(\mathbb{R}) by f,g\langle f,g\rangle.

  2. (2)

    V=p1(𝓅())V=\bigcup\limits_{p\geq 1}\cal{L}(L^{p}(\mathbb{R})) and 𝒟={𝒻,}𝒻,𝒮𝒸()\cal{D}=\{\langle\cdot f,g\rangle\}_{f,g\,\in\cal{S}_{c}(\mathbb{R})}.

Lemma 9.4.

Let TT be an operator satisfying (28) that is self-adjoint on L2()L^{2}(\mathbb{R}). If [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma, then there exists ϵ0(γ)(0,12)\epsilon_{0}(\gamma)\in(0,\frac{1}{2}), with limγ1ϵ0(γ)=12\lim\limits_{\gamma\to 1}\epsilon_{0}(\gamma)=\frac{1}{2}, such that for all z=t+iyN(Ωϵ0(γ))z=t+iy\in N(\Omega_{\leq\epsilon_{0}(\gamma)}), we have

[w]Ap(t)(),[w]Ap(t)()γ1,[w]_{A_{p(t)}(\mathbb{R})},[w]_{A_{p^{\prime}(t)}(\mathbb{R})}\lesssim_{\gamma}1,\, (69)

and

Qw,p(z),Tp,pCγ,\|Q_{w,p(z),T}\|_{p,p}\leq C_{\gamma,\cal{F}}\, (70)

and Qw,p(z),TQ_{w,p(z),T} is weakly analytic with respect to {f,gf,g𝒮𝒸()}\{\langle\cdot f,g\rangle_{f,g\in\cal{S}_{c}(\mathbb{R})}\} on N(Ωϵ0(γ))N(\Omega_{\leq\epsilon_{0}(\gamma)}).

Proof.

Initially define ϵ0(γ)=def1s(2,γ)12\epsilon_{0}(\gamma)\stackrel{{\scriptstyle\rm def}}{{=}}\frac{1}{s(2,\gamma)}-\frac{1}{2} where ss is as in Lemma 5.1. Then ϵ0(γ)(0,12)\epsilon_{0}(\gamma)\in(0,\frac{1}{2}), limγ1ϵ0(γ)=12\lim\limits_{\gamma\to 1}\epsilon_{0}(\gamma)=\frac{1}{2}, and (69) holds for zΩϵ0z\in\Omega_{\leq\epsilon_{0}}; by taking a slightly smaller ϵ0(γ)(0,12)\epsilon_{0}(\gamma)\in(0,\frac{1}{2}), still with limγ1ϵ0(γ)=12\lim\limits_{\gamma\to 1}\epsilon_{0}(\gamma)=\frac{1}{2}, then (69) holds for zN(Ωϵ0(γ))z\in N(\Omega_{\leq\epsilon_{0}(\gamma)}).

If TT satisfies (28), then

w1/pTw1/pp,pCγ,,𝓅,\|w^{1/p}Tw^{-1/p}\|_{p,p}\leq C_{\gamma,\cal{F},p}\,,

for all pp satisfying |121p|ϵ0(γ)\left|\frac{1}{2}-\frac{1}{p}\right|\leq\epsilon_{0}(\gamma). Since w1/pTw1/pp,p=TLwp(),Lwp()\|w^{1/p}Tw^{-1/p}\|_{p,p}=\|T\|_{L^{p}_{w}(\mathbb{R}),L^{p}_{w}(\mathbb{R})}, we can interpolate between the extremal values of pp given by |121p|=ϵ0(γ)\left|\frac{1}{2}-\frac{1}{p}\right|=\epsilon_{0}(\gamma) to in fact get

w1/pTw1/pp,pCγ,\|w^{1/p}Tw^{-1/p}\|_{p,p}\leq C_{\gamma,\cal{F}}\,

for all pp satisfying |121p|ϵ0(γ)\left|\frac{1}{2}-\frac{1}{p}\right|\leq\epsilon_{0}(\gamma). In particular,

w1/p(z)Tw1/p(z)p(t),p(t)=w1/p(t)Tw1/p(t)p(t),p(t)Cγ,\|w^{1/p(z)}Tw^{-1/p(z)}\|_{p(t),p(t)}=\|w^{1/p(t)}Tw^{-1/p(t)}\|_{p(t),p(t)}\leq C_{\gamma,\cal{F}}\,

for all zΩϵ0(γ)z\in\Omega_{\leq\epsilon_{0}(\gamma)}. If we swap the role of pp_{*} with pp_{*}^{\prime}, this has the net effect of swapping the role p(z)p(z) with p(z)p^{\prime}(z). Thus if TT is self-adjoint on L2()L^{2}(\mathbb{R}), then swapping as described and taking adjoints yields

w1/p(z)Tw1/p(z)p(t),p(t)=w1/p(z)Tw1/p(z)p(t),p(t)Cγ,\|w^{-1/p^{\prime}(z)}Tw^{1/p^{\prime}(z)}\|_{p(t),p(t)}=\|w^{1/p^{\prime}(z)}Tw^{-1/p^{\prime}(z)}\|_{p^{\prime}(t),p^{\prime}(t)}\leq C_{\gamma,\cal{F}}\,

for zΩϵ0(γ)z\in\Omega_{\leq\epsilon_{0}(\gamma)}. Combining all of this together yields (70) for zΩϵ0(γ)z\in\Omega_{\epsilon_{0}(\gamma)}; by slightly shrinking ϵ0(γ)(0,1/2)\epsilon_{0}(\gamma)\in(0,1/2), while still requiring limγ1ϵ0(γ)=12\lim\limits_{\gamma\to 1}\epsilon_{0}(\gamma)=\frac{1}{2}, we get (70) for zN(Ωϵ0(γ))z\in N(\Omega_{\leq\epsilon_{0}(\gamma)}).

By Proposition A.4, if we again shrinking ϵ0(γ)(0,12)\epsilon_{0}(\gamma)\in(0,\frac{1}{2}) we may assume without loss of generality that Qw,p(z),TQ_{w,p(z),T} is analytic as a map N(Ωϵ0(γ))(2())N(\Omega_{\leq\epsilon_{0}(\gamma)})\to\cal{L}(L^{2}(\mathbb{R})), and hence weakly analytic with respect to {f,gf,g𝒮𝒸()}\{\langle\cdot f,g\rangle_{f,g\,\in\cal{S}_{c}(\mathbb{R})}\} on N(Ωϵ0(γ))N(\Omega_{\leq\epsilon_{0}(\gamma)}). ∎

Lemma 9.5.

Let [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma, ϵ0(γ)\epsilon_{0}(\gamma) is as Lemma 9.4 and let p(z)p(z) as in (65) be given for some p[1,]p_{*}\in[1,\infty]. If

supzΩϵ0(γ)(IκQw,p(z),T)1p(t),p(t)<,\sup\limits_{z\in\Omega_{\leq\epsilon_{0}(\gamma)}}\|(I-\kappa Q_{w,p(z),T})^{-1}\|_{p(t),p(t)}<\infty\,,

then (IκQw,p(z),T)1(I-\kappa Q_{w,p(z),T})^{-1} is weakly analytic with respect to {f,g}f,g𝒮𝒸()\{\langle\cdot f,g\rangle\}_{f,g\in\cal{S}_{c}(\mathbb{R})} on N(Ωϵ0(γ))N(\Omega_{\leq\epsilon_{0}(\gamma)}).

Proof.

Let F(z)=def(IκQw,p(z),T)1F(z)\stackrel{{\scriptstyle\rm def}}{{=}}(I-\kappa Q_{w,p(z),T})^{-1}. We show that for each z0Ωϵ0(γ)z_{0}\in\Omega_{\leq\epsilon_{0}(\gamma)}, there exists a ball Bδ(z0)B_{\delta}(z_{0}) on which F(z)F(z) is analytic as a map Bδ(z0)(𝓅(𝓉0)())B_{\delta}(z_{0})\to\cal{L}(L^{p(t_{0})}(\mathbb{R})); this will clearly imply weak analyticity on N(Ωϵ0(γ))N(\Omega_{\leq\epsilon_{0}(\gamma)}). Fix z0z_{0}, define

V(z,z0)=defκ(Qw,p(z),TQw,p(z0),T)V(z,z_{0})\stackrel{{\scriptstyle\rm def}}{{=}}\kappa(Q_{w,p(z),T}-Q_{w,p(z_{0}),T})

and write

F(z)=F(z0)+(F(z0)1V(z,z0))1F(z0)=F(z0)+((IF(z0)V(z,z0))1I)F(z0)=(IF(z0)V(z,z0))1F(z0).F(z)=F(z_{0})+(F(z_{0})^{-1}-V(z,z_{0}))^{-1}-F(z_{0})=F(z_{0})+((I-F(z_{0})V(z,z_{0}))^{-1}-I)F(z_{0})\\ =(I-F(z_{0})V(z,z_{0}))^{-1}F(z_{0})\,. (71)

We claim that there exists a ball Bδ(z0)B_{\delta}(z_{0}) on which Qw,p(z),TQ_{w,p(z),T}, and hence V(z,z0)V(z,z_{0}), is analytic as a map Bδ(z0)(𝓅(𝓉0)())B_{\delta}(z_{0})\to\cal{L}(L^{p(t_{0})}(\mathbb{R})). Given the claim, we can then take δ\delta small enough so that V(z,z0)p(t0),p(t0)<12F(z0)p(t0),p(t0)\|V(z,z_{0})\|_{p(t_{0}),p(t_{0})}<\frac{1}{2\|F(z_{0})\|_{p(t_{0}),p(t_{0})}}. Then (71) yields

F(z)=(IF(z0)V(z,z0))1F(z0)=k=0(F(z0)V(z,z0))kF(z0),F(z)=(I-F(z_{0})V(z,z_{0}))^{-1}F(z_{0})=\sum\limits_{k=0}^{\infty}(F(z_{0})V(z,z_{0}))^{k}F(z_{0})\,,

where the sum converges uniformly in (𝓅(𝓉0)())\cal{L}(L^{p(t_{0})(\mathbb{R})}). Since each partial sum is analytic in zz and the sum converges uniformly, we have F(z)F(z) is analytic as a map Bδ(z0)(𝓅(𝓉0)())B_{\delta}(z_{0})\to\cal{L}(L^{p(t_{0})}(\mathbb{R})).

We must now show there exists a ball Bδ(z0)B_{\delta}(z_{0}) on which Qw,p(z),TQ_{w,p(z),T} is analytic as a map Bδ(z0)(𝓅(𝓉0)())B_{\delta}(z_{0})\to\cal{L}(L^{p(t_{0})}(\mathbb{R})). As per Lemma 9.4, Qw,p(z),TQ_{w,p(z),T} is weakly analytic with respect to {f,g}f,g𝒮𝒸()\{\langle\cdot f,g\rangle\}_{f,g\in\cal{S}_{c}(\mathbb{R})} on N(Ωϵ0(γ))N(\Omega_{\leq\epsilon_{0}(\gamma)}), and so by Proposition A.1 it suffices to show there exists δ>0\delta>0 such that for all zBδ(z0)z\in B_{\delta}(z_{0}), we have w1/p(z)Tw1/p(z),w1/p(z)Tw1/p(z)(𝓅(𝓉0)())w^{1/p(z)}Tw^{-1/p(z)},w^{-1/p^{\prime}(z)}Tw^{1/p(z)}\in\cal{L}(L^{p(t_{0})}(\mathbb{R})). We show this for w1/p(z)Tw1/p(z)w^{1/p(z)}Tw^{-1/p(z)}, as the proof for w1/p(z)Tw1/p(z)w^{-1/p^{\prime}(z)}Tw^{1/p^{\prime}(z)} will follow similarly. Write

w1/p(z)Tw1/p(z)p(t0),p(t0)=w1/p(t)Tw1/p(t)p(t0),p(t0)w1/p(t)Tw1/p(t)w1/p(t0)Tw1/p(t0)p(t0),p(t0)+w1/p(t0)Tw1/p(t0)p(t0),p(t0).\|w^{1/p(z)}Tw^{-1/p(z)}\|_{p(t_{0}),p(t_{0})}=\|w^{1/p(t)}Tw^{-1/p(t)}\|_{p(t_{0}),p(t_{0})}\\ \leq\|w^{1/p(t)}Tw^{-1/p(t)}-w^{1/p(t_{0})}Tw^{-1/p(t_{0})}\|_{p(t_{0}),p(t_{0})}+\|w^{1/p(t_{0})}Tw^{-1/p(t_{0})}\|_{p(t_{0}),p(t_{0})}\,.

The last term is clearly finite. As for w1/p(t)Tw1/p(t)w1/p(t0)Tw1/p(t0)p(t0),p(t0)\|w^{1/p(t)}Tw^{-1/p(t)}-w^{1/p(t_{0})}Tw^{-1/p(t_{0})}\|_{p(t_{0}),p(t_{0})}, by (69) we have wAp(t)w\in A_{p(t)}. Thus we can apply the continuity of weighted operators Theorem 9.3 with f=logwf=\log w, to argue that for |tt0||t-t_{0}| sufficiently small, we have w1/p(t)Tw1/p(t)w1/p(t0)Tw1/p(t0)p(t0),p(t0)1\|w^{1/p(t)}Tw^{-1/p(t)}-w^{1/p(t_{0})}Tw^{-1/p(t_{0})}\|_{p(t_{0}),p(t_{0})}\leq 1. This completes the proof. ∎

Proposition 9.6 ([1]).

Let [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma and let p(z)p(z) as in (65) be given for some p[1,]p_{*}\in[1,\infty] such that Ω[1,1]Ωϵ0(γ)\Omega_{[-1,1]}\subset\Omega_{\leq\epsilon_{0}(\gamma)}, where ϵ0(γ)\epsilon_{0}(\gamma) is as in Lemma 9.4. Suppose

sup1Rez1Qw,p(z),Tp(t),p(t)<,\sup_{-1\leqslant\mathop{\rm Re}z\leqslant 1}\|Q_{w,p(z),T}\|_{p(t),p(t)}<\infty\,,

where t=defRezt\stackrel{{\scriptstyle\rm def}}{{=}}\mathop{\rm Re}z. If there is a positive number 𝚲>1{\bf\Lambda}>1 such that

(IκQw,p(z),T)1p(t),p(t)2𝚲\|(I-\kappa Q_{w,p(z),T})^{-1}\|_{p(t),p(t)}\leqslant 2{\bf\Lambda}

for all zΩ[1,1]z\in\Omega_{[-1,1]}, then there is a t(𝚲)(0,1]t_{*}({\bf\Lambda})\in(0,1], so that

(IκQw,p(z),T)1p(t),p(t)𝚲\|(I-\kappa Q_{w,p(z),T})^{-1}\|_{p(t),p(t)}\leqslant{\bf\Lambda}

for all zΩ[t,t]z\in\Omega_{[-t_{*},t_{*}]}.

Proof.

We notice that Qw,p(iy),TQ_{w,p(iy),T} is bounded and antisymmetric operator in Hilbert space L2()L^{2}(\mathbb{R}). Therefore, (IκQw,p(iy),T)12,21\|(I-\kappa Q_{w,p(iy),T})^{-1}\|_{2,2}\leqslant 1. By Lemma 9.5, (IκQw,p(z),T)1(I-\kappa Q_{w,p(z),T})^{-1} is weakly analytic and continuous in the sense of Stein (p.209, [2]). Applying Stein’s interpolation theorem on the strips Ω[0,1]\Omega_{[0,1]} and Ω[1,0]\Omega_{[-1,0]}, we get

(IκQw,p(z),T)1p(t),p(t)exp(sin(π|t|)2log(2𝚲)cosh(πy)+cos(π|t|)𝑑y)=1+O(|t|),t0.\|(I-\kappa Q_{w,p(z),T})^{-1}\|_{p(t),p(t)}\leqslant\exp\left(\frac{\sin(\pi|t|)}{2}\int_{\mathbb{R}}\frac{\log(2{\bf\Lambda)}}{\cosh(\pi y)+\cos(\pi|t|)}dy\right)=1+O(|t|),\quad t\to 0\,.

Remark. We emphasize here that positive tt_{*} does not depend on ww.

Proof of Lemma 9.1.

The reader may also consult a similar, but simpler proof in [1, Proof of Lemma 3.6].

We just need to find ϵ(γ,,𝚲)\epsilon(\gamma,\cal{F},\mathbf{\Lambda}) such that on Ω<ϵ(γ,,𝚲)\Omega_{<\epsilon(\gamma,\cal{F},\mathbf{\Lambda})}, we have (IQw,p(z),T)1(I-Q_{w,p(z),T})^{-1} is bounded on Lp(t)()L^{p(t)}(\mathbb{R}) and analytic as an (2())\cal{L}(L^{2}(\mathbb{R}))-valued function.

In (65), we take parameter pp_{*} as follows: define p(1)p_{*}^{(1)} by 1p(1)=12ϵ0(γ)\frac{1}{p_{*}^{(1)}}=\frac{1}{2}-\epsilon_{0}(\gamma), where ϵ0(γ)\epsilon_{0}(\gamma) is as in Lemma 9.4, and set p1(z)=defp(z)p_{1}(z)\stackrel{{\scriptstyle\rm def}}{{=}}p(z); note then Ω[1,1]Ωϵ0(γ)\Omega_{[-1,1]}\subset\Omega_{\leq\epsilon_{0}(\gamma)}. Consider Qw,p(z),T(j)=defjQw,p(z),T/NQ^{(j)}_{w,p(z),T}\stackrel{{\scriptstyle\rm def}}{{=}}jQ_{w,p(z),T}/N, j=1,,Nj=1,\ldots,N, where NN is large and will be fixed later (it will depend on γ\gamma, \cal{F}, 𝚲\bf{\Lambda} only).

We take NN to satisfy

1Cγ,𝚲/N>1/2,1-C_{\gamma,\cal{F}}{\bf\Lambda}/N>1/2\,, (72)

where Cγ,C_{\gamma,\cal{F}} is as in (70). Next, by (68) and (70),

(IQw,p(t+iy),T(1))1p(t),p(t)11Cγ,/N11Cγ,𝚲/N22𝚲,\|(I-Q^{(1)}_{w,p(t+iy),T})^{-1}\|_{p(t),p(t)}\leqslant\frac{1}{1-C_{\gamma,\mathcal{F}}/N}\leqslant\frac{1}{1-C_{\gamma,\mathcal{F}}{\bf\Lambda}/N}\leqslant 2\leqslant 2{\bf\Lambda}\,,

since 𝚲>1{\bf\Lambda}>1.

We continue with an inductive argument in which the bound for {Qw,p(z)(j)}\{Q^{(j)}_{w,p(z)}\} provides the bound for {Qw,p(z),T(j+1)}\{Q^{(j+1)}_{w,p(z),T}\} when j=1,,N1j=1,\ldots,N-1.

\bullet Base of induction: handling Qw,p(z)(1)Q^{(1)}_{w,p(z)}. Apply Proposition 9.6 with κ=1/N\kappa=1/N to get an absolute constant tt_{*} so that

(IQw,p(t+iy),T(1))1p(t),p(t)𝚲\|(I-Q^{(1)}_{w,p(t+iy),T})^{-1}\|_{p(t),p(t)}\leqslant{\bf\Lambda}

for t[t,t]t\in[-t_{*},t_{*}] and yy\in\mathbb{R}. Next, we use (67) with H=Qw,p(t+iy)(1)H=-Q^{(1)}_{w,p(t+iy)} and V=N1Qw,p(t+iy)V=-N^{-1}Q_{w,p(t+iy)}. This gives

(IQw,p(t+iy),T(2))1p(t),p(t)𝚲1Cγ,𝚲/N2𝚲,t[t,t]\|(I-Q^{(2)}_{w,p(t+iy),T})^{-1}\|_{p(t),p(t)}\leqslant\frac{{\bf\Lambda}}{1-C_{\gamma,\mathcal{F}}{\bf\Lambda}/N}\leqslant 2{\bf\Lambda},\quad t\in[-t_{*},t_{*}] (73)

by (72).

That finishes the first step. Next, we will explain how estimates on Qw,p(z)(2)Q^{(2)}_{w,p(z)} give bounds for Qw,p(z)(3)Q^{(3)}_{w,p(z)}.

\bullet Handling Qw,p(z),T(2)Q^{(2)}_{w,p(z),T}. In Proposition 9.6, we now take κ=κ2=def2/N,p(2)=defp1(t)=p(t)\kappa=\kappa_{2}\stackrel{{\scriptstyle\rm def}}{{=}}2/N,p^{(2)}_{*}\stackrel{{\scriptstyle\rm def}}{{=}}p_{1}(t_{*})=p(t_{*}) (here p(t)p(t_{*}) is obtained at the previous step) and compute new p2(z),p2(z)p_{2}(z),p_{2}^{\prime}(z) by (65):

1p2(z)=defzp(2)+1z2=ztp+1zt2=1p1(zt)=1p(zt).\frac{1}{p_{2}(z)}\stackrel{{\scriptstyle\rm def}}{{=}}\frac{z}{p^{(2)}_{*}}+\frac{1-z}{2}=\frac{zt_{*}}{p_{*}}+\frac{1-zt_{*}}{2}=\frac{1}{p_{1}(zt_{*})}=\frac{1}{p(zt_{*})}\,. (74)

Therefore, when zz belongs to 1Rez1-1\leq\mathop{\rm Re}z\leq 1, ztzt^{*} belongs to tRezt-t_{*}\leq\mathop{\rm Re}z\leq t_{*}, p2(z)=p(zt)p_{2}(z)=p(zt_{*}) and p2(z)p_{2}(z) still satisfies Ω[1,1]Ωϵ0(γ)\Omega_{[-1,1]}\subset\Omega_{\leq\epsilon_{0}(\gamma)}. In this domain, the estimate (73) can be rewritten as

(IQw,p2(t+iy),T(2))1p2(t),p2(t)2𝚲,t[1,1],y,\|(I-Q^{(2)}_{w,p_{2}(t+iy),T})^{-1}\|_{p_{2}(t),p_{2}(t)}\leqslant 2{\bf\Lambda},\quad t\in[-1,1],\quad y\in\mathbb{R}\,,

where p2(z)p_{2}(z) is different from p1(z)=p(z)p_{1}(z)=p(z) only by the choice of parameter pp_{*} in (65) and is in fact a rescaling of the original p(z)p(z) as follows from (74). From Proposition 9.6, we have

(IQw,p2(t+iy),T(2))1p2(t),p2(t)𝚲\|(I-Q^{(2)}_{w,p_{2}(t+iy),T})^{-1}\|_{p_{2}(t),p_{2}(t)}\leqslant{\bf\Lambda}

for t[t,t],yt\in[-t_{*},t_{*}],y\in\mathbb{R}. We use the perturbative bound (67) one more time with H=Qw,p2(t+iy),T(2)H=-Q^{(2)}_{w,p_{2}(t+iy),T} and V=N1Qw,p2(t+iy),TV=-N^{-1}Q_{w,p_{2}(t+iy),T} to get

(IQw,p2(t+iy),T(3))1p2(t),p2(t)2𝚲\|(I-Q^{(3)}_{w,p_{2}(t+iy),T})^{-1}\|_{p_{2}(t),p_{2}(t)}\leqslant 2{\bf\Lambda}

for t[t,t],yt\in[-t_{*},t_{*}],y\in\mathbb{R}.

\bullet Induction in jj and the bound for Qw,p(z),T(N)Q^{(N)}_{w,p(z),T}. Next, we take p(3)=defp(2)(t)p_{*}^{(3)}\stackrel{{\scriptstyle\rm def}}{{=}}p^{(2)}(t_{*}) and repeat the process in which the bound

(IQw,pj(t+iy),T(j))1pj(t),pj(t)2𝚲,z=t+iyΩ[1,1],\|(I-Q^{(j)}_{w,p_{j}(t+iy),T})^{-1}\|_{p_{j}(t),p_{j}(t)}\leqslant 2{\bf\Lambda},\quad z=t+iy\in\Omega_{[-1,1]}\,,

implies

(IQw,pj+1(t+iy),T(j+1))1pj+1(t),pj+1(t)2𝚲,z=t+iyΩ[1,1].\|(I-Q^{(j+1)}_{w,p_{j+1}(t+iy),T})^{-1}\|_{p_{j+1}(t),p_{j+1}(t)}\leqslant 2{\bf\Lambda},\quad z=t+iy\in\Omega_{[-1,1]}\,.

Notice that each time the new pj(z)p_{j}(z) is in fact a rescaling of the original p(z)p(z) by tj1t_{*}^{j-1} as can be seen from a calculation analogous to (74). In N1N-1 steps, we get

(IQw,pN1(t+iy),T(N))1pN1(t),pN1(t)2𝚲,z=t+iyΩ[t,t].\displaystyle\|(I-Q^{(N)}_{w,p_{N-1}(t+iy),T})^{-1}\|_{p_{N-1}(t),p_{N-1}(t)}\leqslant 2{\bf\Lambda}\,,\quad z=t+iy\in\Omega_{[-t_{*},t_{*}]}\,.

Thus recalling that pN1(z)=p(tN2z)p_{N-1}(z)=p(t_{*}^{N-2}z), one has

(IQw,p(tN1z),T(N))1p(tN1t),p(tN1t)2𝚲.\|(I-Q^{(N)}_{w,p(t_{*}^{N-1}z),T})^{-1}\|_{p(t_{*}^{N-1}t),p(t_{*}^{N-1}t)}\leqslant 2{\bf\Lambda}\,.

Since Qw,p(tNz)(N)=Qw,p(tNz),TQ^{(N)}_{w,p(t_{*}^{N}z)}=Q_{w,p(t_{*}^{N}z),T}, we get (66) with

ϵ(γ,𝚲,)=def𝓉𝒩1ϵ0(γ).\epsilon(\gamma,\mathbf{\Lambda},\cal{F})\stackrel{{\scriptstyle\rm def}}{{=}}t_{*}^{N-1}\epsilon_{0}(\gamma).

The estimates (72) implies that we can take NCγ,,𝚲.N\sim C_{\gamma,\cal{F},\mathbf{\Lambda}}\,.

Thus, we showed there exists ϵ(γ,,𝚲)>0\epsilon(\gamma,\cal{F},\mathbf{\Lambda})>0 such that (IQw,p,T)1p,p2𝚲\|(I-Q_{w,p,T})^{-1}\|_{p,p}\leq 2\mathbf{\Lambda} for all |1p12|<ϵ(γ,,𝚲)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\gamma,\cal{F},\mathbf{\Lambda}).

In fact, we may assume without loss of generality that limγ1ϵ(γ,,𝚲)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma,\cal{F},\mathbf{\Lambda})=\frac{1}{2}: indeed, it suffices to show that for any ϵ~(0,12)\widetilde{\epsilon}\in(0,\frac{1}{2}), we can choose γ1\gamma-1 sufficiently small so that (IQw,p(z),T)1p(t),p(t)2𝚲\|(I-Q_{w,p(z),T})^{-1}\|_{p(t),p(t)}\leq 2\mathbf{\Lambda} for all zz satisfying |1p(t)12|<ϵ~|\frac{1}{p(t)}-\frac{1}{2}|<\widetilde{\epsilon}. Note that if Qw,p(z),Tp(t),p(t)12\|Q_{w,p(z),T}\|_{p(t),p(t)}\leq\frac{1}{2}, then by geometric sum

(IQw,p(z),T)1p(t),p(t)k=0Qw,p(z)p(t),p(t)k22𝚲.\|(I-Q_{w,p(z),T})^{-1}\|_{p(t),p(t)}\leq\sum\limits_{k=0}^{\infty}\|Q_{w,p(z)}\|_{p(t),p(t)}^{k}\leq 2\leq 2\mathbf{\Lambda}\,.

Thus it suffices to show we can take γ\gamma sufficiently small so that

Qw,p(z),Tp(t),p(t)12\|Q_{w,p(z),T}\|_{p(t),p(t)}\leq\frac{1}{2}\,

for |1p(t)12|<ϵ~|\frac{1}{p(t)}-\frac{1}{2}|<\widetilde{\epsilon}. In fact, by Stein’s analytic interpolation it suffices to check this at the vertical lines such that |1p(t)12|=ϵ~|\frac{1}{p(t)}-\frac{1}{2}|=\widetilde{\epsilon}. By duality and the triangle inequality this in turn will follow from showing

w1/p(z)Tw1/p(z)Tp(t),p(t)14\|w^{1/p(z)}Tw^{-1/p(z)}-T\|_{p(t),p(t)}\leq\frac{1}{4}\,

for z=t+iyz=t+iy such that |1p(t)12|=ϵ(γ)|\frac{1}{p(t)}-\frac{1}{2}|=\epsilon(\gamma). Fix one the values tt satisfying the last equality: if γ1\gamma-1 is sufficiently small, then by Lemma 8.6 we have logwBMOγ1\|\log w\|_{\rm BMO}\lesssim\sqrt{\gamma-1}, which can be made arbitrarily small. Apply Theorem 9.3 with f=logwf=\log w to get

w1/pTw1/pTp,pp,γ1.\|w^{1/p}Tw^{-1/p}-T\|_{p,p}\lesssim_{p,\cal{F}}\sqrt{\gamma-1}\,.

We can now choose γ\gamma sufficiently small that this is at most 14\frac{1}{4}. Thus, we can choose ϵ(γ,,𝚲)\epsilon(\gamma,\cal{F},\mathbf{\Lambda}) so that as γ1\gamma\to 1, we get ϵ(γ,,𝚲)12\epsilon(\gamma,\cal{F},\mathbf{\Lambda})\to\frac{1}{2}.

As for showing (IQw,p(z),T)1(I-Q_{w,p(z),T})^{-1} is analytic as an (2())\cal{L}(L^{2}(\mathbb{R}))-valued function, by Lemma 9.5 and Proposition A.1, it suffices to show (IQw,p(z),T)1(I-Q_{w,p(z),T})^{-1} is bounded on L2()L^{2}(\mathbb{R}) for zΩϵ(γ,,𝚲)z\in\Omega_{\leq\epsilon(\gamma,\cal{F},\mathbf{\Lambda})}. Fix one such zz; we just showed previously that

(IQw,p(z))1p(t),p(t)2𝚲.\|(I-Q_{w,p(z)})^{-1}\|_{p(t),p(t)}\leq 2\mathbf{\Lambda}\,.

However the proof just as easily applies to weight w1w^{-1}, operator T-T and Hölder index p(z)p^{\prime}(z) (as opposed to p(z)p(z)) and so we get

(IQw,p(z),T)1p(t),p(t)=(IQw1,p(z),T)1p(t),p(t)2𝚲for zΩ<ϵ(γ,,𝚲).\|(I-Q_{w,p(z),T})^{-1}\|_{p^{\prime}(t),p^{\prime}(t)}=\|(I-Q_{w^{-1},p^{\prime}(z),-T})^{-1}\|_{p^{\prime}(t),p^{\prime}(t)}\leq 2\mathbf{\Lambda}\,\quad\text{for }z\in\Omega_{<\epsilon(\gamma,\cal{F},\mathbf{\Lambda})}.

Interpolate between both estimates to get

(IQw,p(z),T)12,22𝚲.\|(I-Q_{w,p(z),T})^{-1}\|_{2,2}\leq 2\mathbf{\Lambda}\,.

This completes the proof. ∎

Appendix A A detour through complex analysis: proof of Proposition 5.5

We recall the following well-known lemma, which we prove for the sake of completeness.

Proposition A.1.

Let BB be a Banach space, let 𝒟\cal{D}\subset B^{*} be a dense subset of the dual space and UU\subset\mathbb{C} an open set. If a:UBa:U\to B is weakly analytic with respect to 𝒟\cal{D}, then a:UBa:U\to B is analytic.

Proof.

We adapt the proof of [6, Theorem 8.20].

Let 𝒟\ell\in\cal{D} so that a(z)\ell\circ a(z) is analytic.

Fix ζU\zeta\in U and let |h|<ϵ|h|<\epsilon for ϵ\epsilon sufficiently small. By the Cauchy integral formula,

a(ζ+h)=12πi|zζ|=ϵa(z)z(ζ+h)𝑑z.\ell\circ a(\zeta+h)=\frac{1}{2\pi i}\oint\limits_{|z-\zeta|=\epsilon}\frac{\ell\circ a(z)}{z-(\zeta+h)}\,dz\,.

Then

a(ζ+h)a(ζ)=h2πi|zζ|=ϵa(z)(z(ζ+h))(zζ)𝑑z.\ell\circ a(\zeta+h)-\ell\circ a(\zeta)=\frac{h}{2\pi i}\oint\limits_{|z-\zeta|=\epsilon}\frac{\ell\circ a(z)}{(z-(\zeta+h))(z-\zeta)}\,dz\,. (75)

For hhh\neq h^{\prime} with 0<|h|,|h|<ϵ0<|h|,|h^{\prime}|<\epsilon, define the second order difference quotient

x(h,h)=def1hh(a(ζ+h)a(ζ)ha(ζ+h)a(ζ)h).x(h,h^{\prime})\stackrel{{\scriptstyle\rm def}}{{=}}\frac{1}{h-h^{\prime}}\left(\frac{a(\zeta+h)-a(\zeta)}{h}-\frac{a(\zeta+h^{\prime})-a(\zeta)}{h^{\prime}}\right)\,.

It suffices to show x(h,h)x(h,h^{\prime}) is uniformly bounded in h,hh,h^{\prime}, for both sufficiently small. Indeed, uniform boundedness in BB would then yield

ha(ζ+h)a(ζ)hh\mapsto\frac{a(\zeta+h)-a(\zeta)}{h}

is Lipschitz for hh near 0, and so has a limit as h0h\to 0.

By (75),

(x(h,h))\displaystyle\ell(x(h,h^{\prime})) =12πi|zζ|=ϵa(z)(zζ)(z(ζ+h))(z(ζ+h))𝑑z.\displaystyle=\frac{1}{2\pi i}\oint\limits_{|z-\zeta|=\epsilon}\frac{\ell\circ a(z)}{(z-\zeta)(z-(\zeta+h))(z-(\zeta+h^{\prime}))}\,dz\,.

For |h|,|h|<ϵ/2|h|,|h^{\prime}|<\epsilon/2, the denominator in the integral is bounded uniformly away from 0, and the numerator is bounded above by Bsup|zζ|=ϵa(z)B\|\ell\|_{B^{*}}\sup\limits_{|z-\zeta|=\epsilon}\|a(z)\|_{B}, which is finite by the uniform boundedness principle. Whence

(x(h,h))MB\ell(x(h,h^{\prime}))\leq M\|\ell\|_{B^{*}}

where MM does not depend on \ell. By duality

x(h,h)BM.\|x(h,h^{\prime})\|_{B}\leq M\,.

Fix χ\chi a characteristic function of some finite interval II, and let Xp=defw1/p(P(r,λ;w)exp(iλr))X_{p}\stackrel{{\scriptstyle\rm def}}{{=}}w^{1/p}(P(r,\lambda;w)-\mathrm{exp}\left(i\lambda r\right)), like in Section 5.

Proposition A.2.

Suppose [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma, p[1,]p_{*}\in[1,\infty] and p(z)p(z) is as in (65). Then zχw1/p(z)1/2z\mapsto\chi w^{1/p(z)-1/2} is analytic as a map

Ω(1,1)L2().\Omega_{(-1,1)}\to L^{2}(\mathbb{R})\,.
Proof.

Without loss of generality, assume p=p_{*}=\infty; the general case follows by a rescaling argument.

We first compute

χw1/p(z)1/2Lq()q=Iwqt/2𝑑λ.\|\chi w^{1/p(z)-1/2}\|_{L^{q}(\mathbb{R})}^{q}=\int\limits_{I}w^{-qt/2}\,d\lambda\,.

Since [w]A2(),[w1]A2()γ[w]_{A_{2}(\mathbb{R})},\,[w^{-1}]_{A_{2}(\mathbb{R})}\leq\gamma, then by Lemma 5.1 there exists q(γ)>2q(\gamma)>2 such that for all |t|<1|t|<1, wqt/2A2()w^{-qt/2}\in A_{2}(\mathbb{R}). Since A2()Lloc1()A_{2}(\mathbb{R})\subset L^{1}_{loc}(\mathbb{R}), then Iwqt/2𝑑λ<\int\limits_{I}w^{-qt/2}\,d\lambda<\infty and so χw1/p(z)1/2Lq()\chi w^{1/p(z)-1/2}\in L^{q}(\mathbb{R}) for some q>2q>2 by (65).

Now let us show analyticity: write

χw1/p(z)1/2\displaystyle\chi w^{1/p(z)-1/2} =(χw1/p(z)1/p(z0))(χw1/p(z0)1/2)\displaystyle=(\chi w^{1/p(z)-1/p(z_{0})})(\chi w^{1/p(z_{0})-1/2})
=k=0(1)kk! 2k(χlogw)k(zz0)k(χw1/p(z0)1/2)\displaystyle=\sum\limits_{k=0}^{\infty}\frac{(-1)^{k}}{k!\,2^{k}}(\chi\log w)^{k}(z-z_{0})^{k}(\chi w^{1/p(z_{0})-1/2})\,

by Taylor expansion of w1/p(z)1/p(z0)=exp((zz0)2logw)w^{1/p(z)-1/p(z_{0})}=\mathrm{exp}\left(-\frac{(z-z_{0})}{2}\log w\right). We simply need to check the series converges uniformly in L2()L^{2}(\mathbb{R}) norm for |zz0||z-z_{0}| sufficiently small; by Stirling’s approximation, it suffices to show

(χlogw)k(χw1/p(z0)1/2)L2()logw,I,γC(logw,I,γ)kkk.\|(\chi\log w)^{k}(\chi w^{1/p(z_{0})-1/2})\|_{L^{2}(\mathbb{R})}\lesssim_{\log w,I,\gamma}C(\log w,I,\gamma)^{k}k^{k}\,.

Apply Hölder’s inequality to bound

(χlogw)k(χw1/p(z0)1/2)L2()χw1/p(z0)1/2Lq()(χlogw)kL2(q/2)()1/21/q,\|(\chi\log w)^{k}(\chi w^{1/p(z_{0})-1/2})\|_{L^{2}(\mathbb{R})}\lesssim\|\chi w^{1/p(z_{0})-1/2}\|_{L^{q}(\mathbb{R})}\|(\chi\log w)^{k}\|_{L^{2(q/2)^{\prime}}(\mathbb{R})}^{1/2-1/q}\,,

where (q/2)(q/2)^{\prime} is the dual exponent to q/2q/2.

Since we already showed χw1/p(z)1/2Lq()\chi w^{1/p(z)-1/2}\in L^{q}(\mathbb{R}), then we will be done once we show (χlogw)kL2(q/2)()logw,I,γC(logw,I,γ)kkk\|(\chi\log w)^{k}\|_{L^{2(q/2)^{\prime}}(\mathbb{R})}\lesssim_{\log w,I,\gamma}C(\log w,I,\gamma)^{k}k^{k}. If p=2(q/2)p=2(q/2)^{\prime}, write

(χlogw)kLp()=(logw)kLp(I)=logwLpk(I)k.\|(\chi\log w)^{k}\|_{L^{p}(\mathbb{R})}=\|(\log w)^{k}\|_{L^{p}(I)}=\|\log w\|_{L^{pk}(I)}^{k}\,.

It suffices to show logwLpk(I)logw,I,γk\|\log w\|_{L^{pk}(I)}\lesssim_{\log w,I,\gamma}k.

Write

logwLpk(I)logwlogwILpk(I)+logwI|I|1/(pk)\|\log w\|_{L^{pk}(I)}\leq\|\log w-\langle\log w\rangle_{I}\|_{L^{pk}(I)}+\langle\log w\rangle_{I}|I|^{1/(pk)}

The last term is logw,I,pk\lesssim_{\log w,I,p}k. For the first term, the John-Nirenberg Theorem 8.5 yields

logwlogwILpk(I)(c|I|pklogwBMO)1/(pk)Γ(pk)1/(pk)logw,I,γΓ(pk)1/(pk)\|\log w-\langle\log w\rangle_{I}\|_{L^{pk}(I)}\leq(c|I|pk\|\log w\|_{\rm BMO})^{1/(pk)}\Gamma(pk)^{1/(pk)}\lesssim_{\log w,I,\gamma}\Gamma(pk)^{1/(pk)}

where Γ\Gamma function is the usual analytic extension of factorial. Finally, Stirling’s formula yields

logwlogwILpk(I)logw,I,γΓ(pk)1/pkpkpk,\|\log w-\langle\log w\rangle_{I}\|_{L^{pk}(I)}\lesssim_{\log w,I,\gamma}\Gamma(pk)^{1/pk}\lesssim pk\lesssim_{p}k\,,

thereby completing the proof. ∎

Remark.  To avoid repeating similar arguments, from now onwards if we write that a function or operator is analytic (or weakly analytic) thanks to a “John-Nirenberg argument,” the reader should understand this as meaning that analyticity follows from an argument similar to the one above where we showed χw1/p(z)1/2\chi w^{1/p(z)-1/2} was analytic as a map {|Rez|<1}L2()\{|\mathop{\rm Re}z|<1\}\to L^{2}(\mathbb{R}).

Lemma A.3.

Suppose [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma, p[1,]p_{*}\in[1,\infty] and p(z)p(z) is as in (65). Then there exists ϵ(γ)(0,12)\epsilon(\gamma)\in(0,\frac{1}{2}), with limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2}, such that zw1/p(z)w1/p(z)z\mapsto w^{1/p(z)}-w^{-1/p^{\prime}(z)} is analytic as a map Ω<ϵ(γ)L2()\Omega_{<\epsilon(\gamma)}\to L^{2}(\mathbb{R}).

Proof.

We first note that w1/p(z)w1/p(z)L2()w^{1/p(z)}-w^{-1/p^{\prime}(z)}\in L^{2}(\mathbb{R}) for zΩ<ϵz\in\Omega_{<\epsilon} for some ϵ>0\epsilon>0. Indeed,

|w1/p(z)w1/p(z)|=|w1/p(z)||w1|=w1/p(t)|w1|=|w1/p(t)w1/p(t)|.\left|w^{1/p(z)}-w^{-1/p^{\prime}(z)}\right|=\left|w^{-1/p^{\prime}(z)}\right|\left|w-1\right|=w^{-1/p(t)}\left|w-1\right|=\left|w^{1/p(t)}-w^{-1/p^{\prime}(t)}\right|\,.

By Lemma 5.3, this belongs to L2()L^{2}(\mathbb{R}) so long as zΩ<ϵ(γ)z\in\Omega_{<\epsilon(\gamma)}, where ϵ(γ)(0,12)\epsilon(\gamma)\in(0,\frac{1}{2}) with limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2}.

By a John-Nirenberg argument, w1/p(z)w^{1/p(z)} and w1/p(z)w^{-1/p^{\prime}(z)} are weakly analytic with respect to 𝒮𝒸()\cal{S}_{c}(\mathbb{R}) on Ω<ϵ(γ)\Omega_{<\epsilon(\gamma)}. By Proposition A.1, this means w1/p(z)w1/p(z)w^{1/p(z)}-w^{-1/p^{\prime}(z)} is analytic as a map Ω<ϵ(γ)L2()\Omega_{<\epsilon(\gamma)}\to L^{2}(\mathbb{R}). ∎

Proposition A.4.

Suppose [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma, TT is an operator satisfying (28) for some \cal{F}, and p(z)p(z) is as in (65) for some p[1,]p_{*}\in[1,\infty]. Then there exists ϵ(γ,)(0,12]\epsilon(\gamma,\cal{F})\in(0,\frac{1}{2}], with limγ1ϵ(γ,)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma,\cal{F})=\frac{1}{2}, such that w1/p(z)Tw1/p(z)w^{1/p(z)}Tw^{-1/p(z)} and w1/p(z)Tw1/p(z)w^{-1/p^{\prime}(z)}Tw^{1/p^{\prime}(z)} are analytic as maps Ω<ϵ(γ,)(2())\Omega_{<\epsilon(\gamma,\cal{F})}\to\cal{L}(L^{2}(\mathbb{R})).

Proof.

We will prove the portion of the proposition involving w1/p(z)Tw1/p(z)w^{1/p(z)}Tw^{-1/p(z)}; the result for w1/p(z)Tw1/p(z)w^{-1/p^{\prime}(z)}Tw^{1/p^{\prime}(z)} follows by then replacing ww by w1w^{-1} and pp_{*} by pp_{*}^{\prime}.

A John-Nirenberg argument shows that w1/p(z)Tw1/p(z)w^{1/p(z)}Tw^{-1/p(z)} is weakly analytic with respect to {f,g}f,g𝒮𝒸()\{\langle\cdot f,g\rangle\}_{f,g\,\in\cal{S}_{c}(\mathbb{R})} for all z=t+iyz=t+iy where wAp(t)w\in A_{p(t)}. By Proposition A.1, w1/p(z)Tw1/p(z)w^{1/p(z)}Tw^{-1/p(z)} is then analytic as a map Ω<ϵ(2())\Omega_{<\epsilon}\to\cal{L}(L^{2}(\mathbb{R})) if we can show w1/p(z)Tw1/p(z)(2())w^{1/p(z)}Tw^{-1/p(z)}\in\cal{L}(L^{2}(\mathbb{R})) on Ω<ϵ\Omega_{<\epsilon} for some ϵ>0\epsilon>0. To see this latter part, notice that

w1/p(z)Tw1/p(z)2,2=w1/p(t)Tw1/p(t)2,2=(w2/p(t))1/2T(w2/p(t))1/22,2.\|w^{1/p(z)}Tw^{-1/p(z)}\|_{2,2}=\|w^{1/p(t)}Tw^{-1/p(t)}\|_{2,2}=\|(w^{2/p(t)})^{1/2}T(w^{2/p(t)})^{-1/2}\|_{2,2}\,.

By Lemma 5.1, there exists ϵ(γ,)\epsilon(\gamma,\cal{F}) such that when t+0iΩ<ϵ(γ,)t+0i\in\Omega_{<\epsilon(\gamma,\cal{F})}, we have w2/p(t)A2()w^{2/p(t)}\in A_{2}(\mathbb{R}). By (28), the operator is then indeed bounded on L2()L^{2}(\mathbb{R}). ∎

The above Proposition A.4, combined with Lemma 4.3, yield the following.

Corollary A.5.

Suppose [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma and p(z)p(z) is as in (65) for some p[1,]p_{*}\in[1,\infty]. Then there exists ϵ(γ)(0,12)\epsilon(\gamma)\in(0,\frac{1}{2}), with limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2}, such that w1/p(z)𝒫[0,r]w1/p(z),w1/p(z)𝒫[0,r]w1/p(z)w^{1/p(z)}\mathcal{P}_{[{0},{r}]}w^{-1/p(z)},w^{-1/p^{\prime}(z)}\mathcal{P}_{[{0},{r}]}w^{1/p^{\prime}(z)} are analytic as maps Ω<ϵ(γ)(2())\Omega_{<\epsilon(\gamma)}\to\cal{L}(L^{2}(\mathbb{R})).

Remark.  In what follows, we will need to show some ϵ(γ)(0,12)\epsilon(\gamma)\in(0,\frac{1}{2}) exists, such that limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2}. In our reasoning, we will consider various other such functions ϵ(γ)\epsilon(\gamma). We consider only finitely many such functions, so by taking the minimum of all of them we will assume without loss of generality that all the ϵ(γ)\epsilon(\gamma)’s are the same.

Corollary A.6.

Suppose [w]A2()γ[w]_{A_{2}(\mathbb{R})}\leq\gamma and p(z)p(z) is as in (65) for some p[1,]p_{*}\in[1,\infty]. Then there exists ϵ(γ)(0,12)\epsilon(\gamma)\in(0,\frac{1}{2}) with limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2} such that

χ(IQw,p(z))1w1/p(z)𝒫[0,r]w1/p(z)(w1/p(z)w1/p(z))exp(iλr)\chi(I-Q_{w,p(z)})^{-1}w^{-1/p^{\prime}(z)}\mathcal{P}_{[0,r]}w^{1/p^{\prime}(z)}(w^{1/p(z)}-w^{-1/p^{\prime}(z)})\mathrm{exp}\left(i\lambda r\right)

is analytic as a map Ω<ϵ(γ)L2()\Omega_{<\epsilon(\gamma)}\to L^{2}(\mathbb{R}).

Proof.

By Corollary A.5, Lemma A.3 and Lemma 9.1 with T=𝒫[0,r]T=\mathcal{P}_{[{0},{r}]} and e.g. 𝚲=10\mathbf{\Lambda}=10, we can choose ϵ(γ)(0,12)\epsilon(\gamma)\in(0,\frac{1}{2}), with limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2}, such that w1/p(z)𝒫[0,r]w1/p(z)w^{-1/p^{\prime}(z)}\mathcal{P}_{[0,r]}w^{1/p^{\prime}(z)}, w1/p(z)w1/p(z)w^{1/p(z)}-w^{-1/p^{\prime}(z)} and (IQw,p(z))1(I-Q_{w,p(z)})^{-1} are analytic as elements or operators on L2()L^{2}(\mathbb{R}), for Ω<ϵ(γ)\Omega_{<\epsilon(\gamma)}. A difference quotient computation then yields the Corollary statement. ∎

Proof of Proposition 5.5.

We begin with (36). By applying Lemma 4.3 and Lemma 9.1 with T=𝒫[0,r]T=\mathcal{P}_{[{0},{r}]} and e.g.𝚲=10\ \mathbf{\Lambda}=10, we can invert IQw,pI-Q_{w,p} to get (41) for all p[2,)p\in[2,\infty) satisfying |121p|<ϵ(γ)|\frac{1}{2}-\frac{1}{p}|<\epsilon(\gamma), where ϵ(γ)(0,12)\epsilon(\gamma)\in(0,\frac{1}{2}) and limγ1ϵ(γ)=12\lim\limits_{\gamma\to 1}\epsilon(\gamma)=\frac{1}{2}. So for all p2p\geq 2 satisfying |121p|<ϵ(γ)|\frac{1}{2}-\frac{1}{p}|<\epsilon(\gamma), we have

χXp,f=χ(IQw,p)1w1/p𝒫[0,r]w1/p(w1/pw1/p)exp(iλr),f\langle\chi X_{p},f\rangle=-\langle\chi(I-Q_{w,p})^{-1}w^{-1/p^{\prime}}\mathcal{P}_{[0,r]}w^{1/p^{\prime}}(w^{1/p}-w^{-1/p^{\prime}})\mathrm{exp}\left(i\lambda r\right),f\rangle (76)

for all f𝒮𝒸()f\in\cal{S}_{c}(\mathbb{R}).

Now replace pp by p(z)p(z) with p=p_{*}=\infty. Then by Corollary A.6, we have

χ(IQw,p(z))1w1/p(z)𝒫[0,r]w1/p(z)(w1/p(z)w1/p(z))exp(iλr),f\langle\chi(I-Q_{w,p(z)})^{-1}w^{-1/p^{\prime}(z)}\mathcal{P}_{[0,r]}w^{1/p^{\prime}(z)}(w^{1/p(z)}-w^{-1/p^{\prime}(z)})\mathrm{exp}\left(i\lambda r\right),f\rangle

is analytic on Ω<ϵ(γ)\Omega_{<\epsilon(\gamma)}. Recall X2L2()X_{2}\in L^{2}(\mathbb{R}) by Proposition 3.1. Thus, if we write χXp(z),f=χw1/p(z)1/2X2,f\langle\chi X_{p(z)},f\rangle=\langle\chi w^{1/p(z)-1/2}X_{2},f\rangle, then by Proposition A.2 and that X2L2()X_{2}\in L^{2}(\mathbb{R}), we may assume χXp(z),f\langle\chi X_{p(z)},f\rangle is analytic on the same region.

Thus (76) shows two analytic functions are equals when zz is in the interval Ω<ϵ(γ)\Omega_{<\epsilon(\gamma)}\cap\mathbb{R}. Thus we must conclude that they are in fact equal on the strip Ω<ϵ(γ)\Omega_{<\epsilon(\gamma)}, i.e.

χXp(z),f=χ(IQw,p(z))1w1/p(z)𝒫[0,r]w1/p(z)(w1/p(z)w1/p(z))exp(iλr),f\langle\chi X_{p(z)},f\rangle=-\langle\chi(I-Q_{w,p(z)})^{-1}w^{-1/p^{\prime}(z)}\mathcal{P}_{[0,r]}w^{1/p^{\prime}(z)}(w^{1/p(z)}-w^{-1/p^{\prime}(z)})\mathrm{exp}\left(i\lambda r\right),f\rangle

on Ω<ϵ(γ)\Omega_{<\epsilon(\gamma)} for all f𝒮𝒸()f\in\cal{S}_{c}(\mathbb{R}). Since χ\chi is the indicator of an arbitrary finite interval, then by duality and density of 𝒮𝒸()\cal{S}_{c}(\mathbb{R}) in Lp()L^{p}(\mathbb{R}),

Xp(z)=(IQw,p(z))1w1/p(z)𝒫[0,r]w1/p(z)(w1/p(z)w1/p(z))exp(iλr)X_{p(z)}=-(I-Q_{w,p(z)})^{-1}w^{-1/p^{\prime}(z)}\mathcal{P}_{[0,r]}w^{1/p^{\prime}(z)}(w^{1/p(z)}-w^{-1/p^{\prime}(z)})\mathrm{exp}\left(i\lambda r\right)\,

for zΩ<ϵ(γ)z\in\Omega_{<\epsilon(\gamma)}. Taking z=tz=t yields (41) for |1p12|<ϵ(γ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\gamma). ∎

Remark.  Both sides of (41) always makes sense as elements of L2()L^{2}(\mathbb{R}) for |1p12|<ϵ(γ)|\frac{1}{p}-\frac{1}{2}|<\epsilon(\gamma).

References

  • [1] M. Alexis, A. Aptekarev, and S. Denisov. Continuity of Weighted Operators, Muckenhoupt ApA_{p} Weights, and Steklov Problem for Orthogonal Polynomials. International Mathematics Research Notices, 2022(8):5935–5972, 10 2020.
  • [2] C. Bennett and R. Sharpley. Interpolation of operators, volume 129 of Pure and Applied Mathematics. Academic Press, Inc., Boston, MA, 1988.
  • [3] S. Denisov. Continuous analogs of polynomials orthogonal on the unit circle and Kreĭn systems. International Mathematics Research Surveys, 01 2006.
  • [4] S. Denisov. On the growth of polynomials orthogonal on the unit circle with a weight ww that satisfies w,w1L(𝕋)w,w^{-1}\in L^{\infty}(\mathbb{T}). Mat. Sb., 209(7):71–105, 2018.
  • [5] S. Denisov and K. Rush. Orthogonal polynomials on the circle for the weight ww satisfying conditions w,w1BMOw,w^{-1}\in{\rm BMO}. Constr. Approx., 46(2):285–303, 2017.
  • [6] M. Einsiedler and T. Ward. Functional Analysis, Spectral Theory, and Applications. Graduate Texts in Mathematics. Springer International Publishing, 2017.
  • [7] M. B. Korey. Ideal weights: asymptotically optimal versions of doubling, absolute continuity, and bounded mean oscillation. J. Fourier Anal. Appl., 4(4-5):491–519, 1998.
  • [8] M. G. Krein. Continuous analogues of propositions on polynomials orthogonal on the unit circle. Dokl. Akad. Nauk SSSR (N.S.), 105:637–640, 1955.
  • [9] A. K. Lerner and F. Nazarov. Intuitive dyadic calculus: the basics. Expo. Math., 37(3):225–265, 2019.
  • [10] S. Pichorides. On the best values of the constants in the theorem of M. Riesz, Zygmund and Kolmogorov. Studia Mathematica, 44(2):165–179, 1972.
  • [11] Alexei Poltoratski. Pointwise convergence of the non-linear fourier transform. arXiv:2103.13349, 2021.
  • [12] B. Simon. Orthogonal Polynomials on the Unit Circle. Part 1: Classical Theory. Colloquium Publications. American Mathematical Society, 2004.
  • [13] E. M. Stein. Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, volume 43 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1993. With the assistance of Timothy S. Murphy, Monographs in Harmonic Analysis, III.
  • [14] Elias M. Stein. Singular Integrals and Differentiability Properties of Functions (PMS-30). Princeton University Press, 2016.
  • [15] V. I. Vasyunin. The exact constant in the inverse Hölder inequality for Muckenhoupt weights. Algebra i Analiz, 15(1):73–117, 2003.