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

Regularity of oscillatory integral operators

Anders Israelsson Tobias Mattsson  and  Wolfgang Staubach
Anders Israelsson, Tobias Mattsson, Wolfgang Staubach
Department of Mathematics, Uppsala University,
S-751 06 Uppsala, Sweden
[email protected], [email protected], [email protected]
Abstract.

In this paper, we establish the global boundedness of oscillatory integral operators on Besov-Lipschitz and Triebel-Lizorkin spaces, with amplitudes in general Sρ,δm(n)S^{m}_{\rho,\delta}(\mathbb{R}^{n})-classes and non-degenerate phase functions in the class Fk\textart{F}^{k}. Our results hold for a wide range of parameters 0ρ10\leq\rho\leq 1, 0δ<10\leq\delta<1, 0<p0<p\leq\infty, 0<q0<q\leq\infty and k>0k>0. We also provide a sufficient condition for the boundedness of operators with amplitudes in the forbidden class S1,1m(n)S^{m}_{1,1}(\mathbb{R}^{n}) in Triebel-Lizorkin spaces.

Key words and phrases:
Oscillatory integral operators, Besov-Lipschitz spaces, Triebel-Lizorkin spaces.
2020 Mathematics Subject Classification:
Primary: 42B20, 42B35, 42B37
The second author is supported by the Knut and Alice Wallenberg Foundation.

1. Introduction

This paper is devoted to the investigation of the global regularity of oscillatory integral operators (here referred to as OIOs) of the form

Taφf(x)=1(2π)nneiφ(x,ξ)a(x,ξ)f^(ξ)dξ,T_{a}^{\varphi}f(x)=\frac{1}{(2\pi)^{n}}\int_{\mathbb{R}^{n}}e^{i\varphi(x,\xi)}\,a(x,\xi)\,\widehat{f}(\xi)\,\mathrm{d}\xi,

with amplitudes in the general Hörmander class Sρ,δm(n)S^{m}_{\rho,\delta}(\mathbb{R}^{n}) (Definition 2.1), on Besov-Lipschitz Bp,qs(n)B^{s}_{p,q}(\mathbb{R}^{n}) and Triebel-Lizorkin Fp,qs(n)F^{s}_{p,q}(\mathbb{R}^{n}) spaces of order ss\in\mathbb{R} with 0<p0<p\leq\infty and 0<q.0<q\leq\infty. Throughout the paper, we are assuming that the phase function φ(x,ξ)\varphi(x,\xi) is in the class Fk\textart{F}^{k} for some k>0k>0 (Definition 2.3) and is strongly non-degenerate (Definition 2.5). Here we note that we are not assuming any homogeneity in the ξ\xi variable of the phase function φ\varphi, as is the case for Fourier integral operators where φ(x,ξ)\varphi(x,\xi) is assumed to be positively homogeneous of degree one in ξ\xi.

The motivation for the study of the OIOs that are investigated in this paper comes from the theory of partial differential equations where phase functions φ(x,ξ)=xξ+ϕ(ξ)\varphi(x,\xi)=x\cdot\xi+\phi(\xi) frequently appear in the study of dispersive equations. Indeed ϕ(ξ)=|ξ|1/2\phi(\xi)=|\xi|^{1/2} corresponds to the water-wave equation, ϕ(ξ)=|ξ|2\phi(\xi)=|\xi|^{2} corresponds to the Schrödinger equation, while ϕ(ξ)=|ξ|3\phi(\xi)=|\xi|^{3} and ϕ(ξ)=ξ|ξ|\phi(\xi)=\xi|\xi| (both in dimension one) corresponds to Airy and Benjamin-Ono equations respectively.

The results that are obtained in this paper also accommodate for instance the case of variable coefficient Schrödinger equations through the earlier investigations of B. Helffer and D. Robert [13], Helffer [12] and the work of E. Cordero F. Nicola and L. Rodino [4], [5].

In [3], the second and the third author in collaboration with A. Castro and M. Yerlanov established an LpLqL^{p}-L^{q}, (1<pq<1<p\leq q<\infty) regularity theory for OIOs in the following two cases: (1) when the amplitude is in S1,0m(n)S^{m}_{1,0}(\mathbb{R}^{n}) and the phase function is in Fk\textart{F}^{k} with 0<k<10<k<1. (2) when the amplitude is in S0,0m(n)S^{m}_{0,0}(\mathbb{R}^{n}) and the phase function is in Fk\textart{F}^{k} with k1k\geq 1. The authors of [3] also went beyond the scope of LpL^{p}-spaces and investigated the regularity of OIOs in classical function spaces such as Besov-Lipschitz and Triebel-Lizorkin spaces.

In [18] M. Pramanik, K. Rogers and A. Seeger (see Theorem 2.23 below), proved a Calderón-Zygmund-type estimate with far-reaching applications, including the regularity of Radon transforms and Fourier integral operators. In that paper, the authors also considered local Fourier integral operators TaφT_{a}^{\varphi} where a(x,ξ)S1,0m(n)a(x,\xi)\in S^{m}_{1,0}(\mathbb{R}^{n}) is compactly supported in xx and φ(x,ξ)\varphi(x,\xi) is non-degenerate on the support of a(x,ξ).a(x,\xi). Using their Calderón-Zygmund estimate in [18], they showed that if n2, 2<p<,q>0n\geq 2,\,2<p<\infty,\,q>0, then Taφ:Fp,p0(n)Fp,q0(n)T_{a}^{\varphi}:F^{0}_{p,p}(\mathbb{R}^{n})\rightarrow F^{0}_{p,q}(\mathbb{R}^{n}), provided that m=(n1)|1p12|.m=-(n-1)\big{|}\frac{1}{p}-\frac{1}{2}\big{|}.

In this paper we prove a variation of Theorem 2.23 (see Theorem 7.3), which is on the one hand more suitable for extensions to the case of quasi-Banach scales of Besov-Lipschitz and Triebel-Lizorkin spaces, and on the other also fits well for the applications to the regularity theory of OIOs. More specifically using certain decomposition in the frequency space and rather intricate kernel estimates for oscillatory integral operators, a composition theorem for the action of parameter-dependent pseudodifferential operators on OIOs, atomic and molecular decompositions of Triebel-Lizorkin spaces in the spirit of Frazier-Jawerth [8, 9, 10], and vector-valued inequalities of the type provided in Theorem 7.3, we manage to get a significant extension of the results in [3]. These extensions are both in terms of the types of oscillatory integral operators, and also the scales of the function spaces on which the operators act. Figure 2 illustrates the extensions that are obtained here.
More specifically in Theorem 8.1 we obtain the following result. Let 0ρ10\leq\rho\leq 1 and 0δ<10\leq\delta<1, 0<μ1,0<\mu\leq 1, k>0k>0, ϰ=min(ρ,1k)\varkappa=\min(\rho,1-k) and set

m(p):=mϰ(p)+ζ.m(p):=m_{\varkappa}(p)+\zeta.

where

mϰ(p):=n(1ϰ)|1p12|m_{\varkappa}(p):=-n(1-\varkappa)\Big{|}\frac{1}{p}-\frac{1}{2}\Big{|}

and

ζ:=min(0,n2(ρδ)).\zeta:=\min\Big{(}0,\frac{n}{2}(\rho-\delta)\Big{)}.

Assume that aSρ,δm(p)(n)a\in S^{m(p)}_{\rho,\delta}(\mathbb{R}^{n}), and suppose furthermore that φFk\varphi\in\textart{F}^{k} is SND and satisfies the L2L^{2}-condition of Definition 2.6 and the LF(μ)(\mu)-condition of Definition (2.8). Let TaφT_{a}^{\varphi} be the associated OIO.

If ss\in\mathbb{R}, 0<q0<q\leq\infty, and either one of the following cases holds

(i)\displaystyle(i)\quad 2<p< when 0<qp,\displaystyle 2<p<\infty\text{ when }0<q\leq p,
(ii)\displaystyle(ii)\quad nn+μ<p<2 when pq,\displaystyle\frac{n}{n+\mu}<p<2\text{ when }p\leq q,
(iii)\displaystyle(iii)\quad p=q=2,\displaystyle p=q=2,

then the OIO TaφT_{a}^{\varphi} is bounded from Fp,qs(n)F_{p,q}^{s}(\mathbb{R}^{n}) to Fp,qs(n).F_{p,q}^{s}(\mathbb{R}^{n}). Moreover in Theorem 9.1 we also show the boundedness of TaφT_{a}^{\varphi} on Besov-Lipschitz spaces Bp,qs(n),B_{p,q}^{s}(\mathbb{R}^{n}), for any nn+μ<p<\frac{n}{n+\mu}<p<\infty and any 0<q.0<q\leq\infty. These results are currently the most general regularity results for oscillatory integral operators with amplitudes in a general Hörmander class, which include the majority of OIOs that appear in the theory of partial differential equations.

The case of operators with amplitudes in the forbidden Hörmander class S1,1kn|1/p1/2|(n)S^{-kn|1/p-1/2|}_{1,1}(\mathbb{R}^{n}) is excluded in the results above since these operators do not in general even allow L2L^{2}-boundedness. However, for s>n(1min{1,p,q}1)s>n\big{(}\frac{1}{\min\{1,p,q\}}-1\big{)} we are able to show (Theorem 8.2) the boundedness of OIOs on Fp,qs(n)F_{p,q}^{s}(\mathbb{R}^{n}) under either of the cases (i), (ii) and (iii) above.

The paper is organized as follows; in Section 2 we provide the necessary preliminaries from the theory of oscillatory integral operators and the theory of function spaces. Here the reader will also find some of the basic results that are used throughout the paper. In Section 3 we prove a basic kernel estimate which lies at the ground for the establishment of Besov-Lipschitz boundedness of OIOs. This is done by utilizing a particular frequency-space decomposition adapted to the OIOs with phase functions in the class Fk\textart{F}^{k}. In Section 4 we prove the basic global L2L^{2}-boundedness result for OIOs with strongly non-degenerate phase functions and amplitudes in general Hörmander classes. This is done by using a continuous version of the almost orthogonality method and Cotlar-Stein’s lemma. In Section 5 we prove the hpLph^{p}\to L^{p} (p<1)p<1) and LbmoL^{\infty}\to\mathrm{bmo} boundedness of OIOs. In Sections 6 and 7 the Hardy space results are transferred to Triebel-Lizorkin spaces. Section 8 contains our main result concerning the boundedness of OIOs on Triebel-Lizorkin spaces. This section includes the results for both classical and forbidden amplitudes. In Section (9) we conclude the paper by proving our main results on the regularity of OIOs in Besov-Lipschitz spaces.

Acknowldgements. The second author is supported by the Knut and Alice Wallenberg Foundation. The authors are also grateful to Andreas Strömbergsson for his support and encouragement.

2. Preliminaries

In this section, we introduce the necessary background and preliminary results that will be required for the development of our main results. These results, while not the focus of our paper, are crucial for the understanding and appreciation of our work.

We begin by introducing the key concepts and definitions relating to oscillatory integral operators. Next, we present Theorem 2.12 on the composition of pseudodifferential operators and oscillatory integral operators and then basic facts about Triebel-Lizorkin and Besov-Lipschitz spaces which will be used in the proof of our main results.

By the end of this section, the reader should have a strong foundation and be well-equipped to move on to the main results of our paper.

As is common practice, we will denote positive constants in the inequalities by CC, which can be determined by known parameters in a given situation but whose value is not crucial to the problem at hand. Such parameters in this paper would be, for example, mm, pp, ss, nn, and the constants connected to the seminorms of various amplitudes or phase functions. The value of CC may differ from line to line, but in each instance could be estimated if necessary. We also write aba\lesssim b as shorthand for aCba\leq Cb and moreover will use the notation aba\sim b if aba\lesssim b and bab\lesssim a.

2.1. Basic facts related to oscillatory integral operators

Oscillatory integral operators play a central role in a wide range of mathematical fields, including harmonic analysis, partial differential equations, and the study of singular integral operators. In this section, we provide a brief overview of some basic facts related to oscillatory integral operators that will be used throughout the rest of our paper.

We begin by introducing the definition of an oscillatory integral operator and discussing some of its basic properties. Next, we present a key lemma and theorem that provide useful information on the compositions of these operators.

Definition 2.1.

Let mm\in\mathbb{R} and ρ,δ[0,1]\rho,\delta\in[0,1]. An amplitude (symbol) a(x,ξ)a(x,\xi) in the class Sρ,δm(n)S^{m}_{\rho,\delta}(\mathbb{R}^{n}) is a function a𝒞(n×n)a\in\mathcal{C}^{\infty}(\mathbb{R}^{n}\times\mathbb{R}^{n}) that verifies the estimate

|ξαxβa(x,ξ)|ξmρ|α|+δ|β|,\left|\partial_{\xi}^{\alpha}\partial_{x}^{\beta}a(x,\xi)\right|\lesssim\langle\xi\rangle^{m-\rho|\alpha|+\delta|\beta|},

for all multi-indices α\alpha and β\beta and (x,ξ)n×n(x,\xi)\in\mathbb{R}^{n}\times\mathbb{R}^{n}, where ξ:=(1+|ξ|2)1/2.\langle\xi\rangle:=(1+|\xi|^{2})^{1/2}.

We refer to mm as the order of the amplitude and ρ,δ\rho,\,\delta as its type. We will refer to the class Sρ,δm(n)S_{\rho,\delta}^{m}(\mathbb{R}^{n}) with 0<ρ10<\rho\leq 1, 0δ<10\leq\delta<1 as classical, to the class S0,δm(n)S_{0,\delta}^{m}(\mathbb{R}^{n}) with 0δ<10\leq\delta<1 as the exotic class, and to Sρ,1m(n)S_{\rho,1}^{m}(\mathbb{R}^{n}) with 0ρ10\leq\rho\leq 1 as the forbidden class of amplitudes.

It turns out (as we will see later) that for the Besov-Lipschitz estimates we do not require any regularity in the xx-variable and therefore we introduce the following class which was first defined by Kenig and Staubach in [17].

Definition 2.2.

Let mm\in\mathbb{R} and ρ[0,1]\rho\in[0,1]. An amplitude (symbol) a(x,ξ)a(x,\xi) in the class LSρm(n)L^{\infty}S^{m}_{\rho}(\mathbb{R}^{n}) is a function a:n×nna:\mathbb{R}^{n}\times\mathbb{R}^{n}\to\mathbb{R}^{n} that verifies the estimate

ξαa(x,ξ)Lx(n)ξmρ|α|,\|\partial_{\xi}^{\alpha}a(x,\xi)\|_{L^{\infty}_{x}(\mathbb{R}^{n})}\lesssim\langle\xi\rangle^{m-\rho|\alpha|},

for all multi-indices α\alpha and β\beta and (x,ξ)n×n(x,\xi)\in\mathbb{R}^{n}\times\mathbb{R}^{n}. Thus we only assume measurability in the xx-variable.

In our treatment of oscillatory integral operators, the phase functions take center stage, and the OIOs are classified according to their phases.

Definition 2.3.

For 0<k<0<k<\infty, we say that a real-valued phase function φ(x,ξ)\varphi(x,\xi) belongs to the class Fk\textart{F}^{k}, if φ(x,ξ)𝒞(n×n{0})\varphi(x,\xi)\in\mathcal{C}^{\infty}(\mathbb{R}^{n}\times\mathbb{R}^{n}\setminus\{0\}) and satisfies the following estimates ((depending on the range of k)k):

  • for k1k\geq 1,

    |ξα(φ(x,ξ)xξ)|cα|ξ|k1,|α|1,|\partial^{\alpha}_{\xi}(\varphi(x,\xi)-x\cdot\xi)|\leq c_{\alpha}|\xi|^{k-1},\quad|\alpha|\geq 1,
  • for 0<k<10<k<1,

    |ξαxβ(φ(x,ξ)xξ)|cα,β|ξ|k|α|,|α+β|1,|\partial^{\alpha}_{\xi}\partial^{\beta}_{x}(\varphi(x,\xi)-x\cdot\xi)|\leq c_{\alpha,\beta}|\xi|^{k-|\alpha|},\quad|\alpha+\beta|\geq 1,

for all xnx\in\mathbb{R}^{n} and |ξ|1|\xi|\geq 1.

Remark 2.4.

A well known and typical example of a phase in Fk\textart{F}^{k} is the phases |ξ|k+xξ|\xi|^{k}+x\cdot\xi which are related to the operator ei(Δ)k/2e^{i(\Delta)^{k/2}}

An important condition on the phase function in the context of global regularity of OIOs, is the so-called strong non-degeneracy condition:

Definition 2.5.

One says that the phase function φ(x,ξ)\varphi(x,\xi) satisfies the strong non-degeneracy condition (or φ\varphi is SND\mathrm{SND} for short) if

(1) |det(xjξk2φ(x,ξ))|δ,for some δ>0 and all (x,ξ)n×n{0}.\big{|}\det(\partial^{2}_{x_{j}\xi_{k}}\varphi(x,\xi))\big{|}\geq\delta,\qquad\mbox{for some $\delta>0$ and all $(x,\xi)\in\mathbb{R}^{n}\times\mathbb{R}^{n}\setminus\{0\}$}.

In order the guarantee that our operators are globally L2L^{2}-bounded we should also put yet another condition of the phase which we shall henceforth simply refer to as the L2L^{2}-condition. The L2L^{2}-condition is essential for the validity of our main results. It allows us to control the L2L^{2}-behavior of the oscillatory integral operator.

Definition 2.6.

One says that the phase function φ(x,ξ)𝒞(n×n)\varphi(x,\xi)\in\mathcal{C}^{\infty}(\mathbb{R}^{n}\times\mathbb{R}^{n}) satisfies the weak L2L^{2}-condition if

(2) |xαξφ(x,ξ)|Cα,|xξβφ(x,ξ)|Cβ|\partial_{x}^{\alpha}\partial_{\xi}\varphi(x,\xi)|\leq C_{\alpha},\quad|\partial_{x}\partial_{\xi}^{\beta}\varphi(x,\xi)|\leq C_{\beta}

for |α|1|\alpha|\geq 1, |β|1,|\beta|\geq 1, all xnx\in\mathbb{R}^{n} and |ξ|1|\xi|\geq 1.
We say that φ(x,ξ)\varphi(x,\xi) satisfies the strong L2L^{2}-condition if

(3) |xαξβφ(x,ξ)|Cα,|\partial_{x}^{\alpha}\partial_{\xi}^{\beta}\varphi(x,\xi)|\leq C_{\alpha},

for |α|1|\alpha|\geq 1, |β|1,|\beta|\geq 1, all xnx\in\mathbb{R}^{n} and |ξ|1|\xi|\geq 1.

Having the definitions of the amplitudes and the phase functions at hand, one has

Definition 2.7.

An oscillatory integral operator (OIO) TaφT_{a}^{\varphi} with amplitude aSρ,δm(n)a\in S^{m}_{\rho,\delta}(\mathbb{R}^{n}) and a real valued phase function φ\varphi, is defined (once again a-priori on 𝒮(n)\mathscr{S}(\mathbb{R}^{n})) by

(4) Taφf(x):=neiφ(x,ξ)a(x,ξ)f^(ξ)đξ.T_{a}^{\varphi}f(x):=\int_{\mathbb{R}^{n}}e^{i\varphi(x,\xi)}\,a(x,\xi)\,\widehat{f}(\xi)\,\text{\rm{\mbox{\dj}}}\xi.

If φFk\varphi\in\textart{F}^{k} and is SND\mathrm{SND}, then these operators will be referred to as oscillatory integral operators of order kk.

The LF(μ)\mathrm{LF}(\mu)-condition is a natural requirement which from the point of view of the applications into PDE’s, will always be satisfied and would not cause any loss of generality. It ensures that the operator behaves in a predictable and well-behaved manner, which is necessary for the analysis the low frequency portions of the operators.

Definition 2.8.

Assume that φ(x,ξ)𝒞(n×n{0})\varphi(x,\xi)\in\mathcal{C}^{\infty}(\mathbb{R}^{n}\times\mathbb{R}^{n}\setminus\{0\}) is real-valued and 0<μ10<\mu\leq 1. We say that φ\varphi satisfies the low frequency phase condition of order μ\mu, (φ\varphi satisfies LF(μ)(\mu)-condition for short), if one has

(5) |ξαxβ(φ(x,ξ)xξ)|cα|ξ|μ|α|,|\partial^{\alpha}_{\xi}\partial_{x}^{\beta}(\varphi(x,\xi)-x\cdot\xi)|\leq c_{\alpha}|\xi|^{\mu-|\alpha|},

for all xnx\in\mathbb{R}^{n}, 0<|ξ|20<|\xi|\leq 2 and all multi-indices α,β\alpha,\beta.

Remark 2.9.

As an example, note that the phase function associated to the water-wave equation is xξ+|ξ|1/2x\cdot\xi+|\xi|^{1/2}, which satisfies the LF(μ)\mathrm{LF}(\mu) condition with μ=12\mu=\frac{1}{2}. The phase function associated to the capillary wave equation is xξ+|ξ|3/2,x\cdot\xi+|\xi|^{3/2}, which is in LF(μ)\mathrm{LF}(\mu) for any μ(0,1]\mu\in(0,1].

We shall also need the following lemma to estimate the phase in the proofs for Sobolev-boundedness of OIOs with forbidden symbols (whenever the composition Theorem 2.12 below is used), the proof of this lemma can be found in [3, Lemma 4.1].

Lemma 2.10.

Assume a(x,ξ)a(x,\xi) is an amplitude, and φ(x,ξ)\varphi(x,\xi) is a SND phase function satisfying

(6) |ξxβφ(x,ξ)|cβ,|β|1 and |ξ|1.|\partial_{\xi}\partial_{x}^{\beta}\varphi(x,\xi)|\leq c_{\beta},\quad|\beta|\geq 1\textnormal{ and }|\xi|\geq 1.

Then for all |β|1|\beta|\geq 1, the following estimates

(7) |xφ(x,ξ)||ξ|\displaystyle|\nabla_{x}\varphi(x,\xi)|\sim|\xi|
(8) |xβφ(x,ξ)|ξ\displaystyle|\partial^{\beta}_{x}\varphi(x,\xi)|\lesssim\left\langle\xi\right\rangle

hold true for the phase function φ\varphi, on the support of a(x,ξ)a(x,\xi), provided that the ξ\xi-support of a(x,ξ)a(x,\xi) lies outside the ball B(0,R)B(0,R) for some large enough R1R\gg 1 and xβφ(x,ξ)L(n×𝕊n1)\partial^{\beta}_{x}\varphi(x,\xi)\in L^{\infty}(\mathbb{R}^{n}\times\mathbb{S}^{n-1}), for |β|1.|\beta|\geq 1.

From this lemma it readily follows that

Corollary 2.11.

The phase functions in Fk\textart{F}^{k} that also satisfy the L2L^{2}-condition (3), all verify the estimates (7) and (8) of Lemma 2.10.

We also recall a composition result proved in [14]. This will be essential in the proof of the boundedness of OIOs with forbidden amplitudes on Sobolev spaces as well as in the proof of the Triebel-Lizorkin boundedness of oscillatory integral operators and in other situations.

Theorem 2.12 ([14]).

Let m,sm,s\in\mathbb{R}, ρ[0,1]\rho\in[0,1], δ[0,1)\delta\in[0,1). Suppose that a(x,ξ)Sρ,δm(n)a(x,\xi)\in S_{\rho,\delta}^{m}(\mathbb{R}^{n}), b(x,ξ)S1,0s(n)b(x,\xi)\in S^{s}_{1,0}(\mathbb{R}^{n}) and φ(x,ξ)\varphi(x,\xi) is a phase function that is smooth on suppa\operatorname{supp}a and verifies the conditions

(9) |ξ||xφ(x,ξ)||ξ|,\displaystyle|\xi|\lesssim|\nabla_{x}\varphi(x,\xi)|\lesssim|\xi|, |ξαxβφ(x,ξ)|ξ1|α|\displaystyle|\partial_{\xi}^{\alpha}\partial_{x}^{\beta}\varphi(x,\xi)|\lesssim\left\langle\xi\right\rangle^{1-|\alpha|}

for all (x,ξ)suppa(x,\xi)\in\operatorname{supp}a and for all |α|0,|\alpha|\geq 0, and all |β|1|\beta|\geq 1. For 0<t10<t\leq 1 consider the parameter-dependent pseudodifferential operator

b(x,tD)f(x):=neixξb(x,tξ)f^(ξ)đξ,b(x,tD)f(x):=\int_{\mathbb{R}^{n}}e^{ix\cdot\xi}\,b(x,t\xi)\,\widehat{f}(\xi)\,\text{\rm{\mbox{\dj}}}\xi,

and the oscillatory integral operator

Taφf(x):=neiφ(x,ξ)a(x,ξ)f^(ξ)đξ.T_{a}^{\varphi}f(x):=\int_{\mathbb{R}^{n}}e^{i\varphi(x,\xi)}\,a(x,\xi)\,\widehat{f}(\xi)\,\text{\rm{\mbox{\dj}}}\xi.

Let σt\sigma_{t} be the amplitude of the composition operator Tσtφ:=b(x,tD)TaφT_{\sigma_{t}}^{\varphi}:=b(x,tD)T_{a}^{\varphi} given by

σt(x,ξ):=n×na(y,ξ)b(x,tη)ei(xy)η+iφ(y,ξ)iφ(x,ξ)đηdy.\sigma_{t}(x,\xi):=\iint_{\mathbb{R}^{n}\times\mathbb{R}^{n}}a(y,\xi)\,b(x,t\eta)\,e^{i(x-y)\cdot\eta+i\varphi(y,\xi)-i\varphi(x,\xi)}\,\text{\rm{\mbox{\dj}}}\eta\,\mathrm{d}y.

Then for any M1M\geq 1 and all 0<ε<1max(δ,1/2)0<\varepsilon<1-\max(\delta,1/2), one can write σt\sigma_{t} as

(10) σt(x,ξ)=b(x,txφ(x,ξ))a(x,ξ)+0<|α|<Mtε|α|α!σα(t,x,ξ)+tεMr(t,x,ξ),\sigma_{t}(x,\xi)=b(x,t\nabla_{x}\varphi(x,\xi))\,a(x,\xi)+\sum_{0<|\alpha|<M}\frac{t^{\varepsilon|\alpha|}}{\alpha!}\,\sigma_{\alpha}(t,x,\xi)+t^{\varepsilon M}r(t,x,\xi),

where, for all multi-indices β,γ\beta,\gamma one has

|ξβxγσα(t,x,ξ)|\displaystyle|\partial^{\beta}_{\xi}\partial^{\gamma}_{x}\sigma_{\alpha}(t,x,\xi)| tmin(s,0)ξs+m(1max(δ,1/2)ε)|α|ρ|β|+δ|γ|,\displaystyle\lesssim t^{\min(s,0)}{\langle{\xi}\rangle}^{s+m-(1-\max(\delta,1/2)-\varepsilon)|\alpha|-\rho|\beta|+\delta|\gamma|},
|ξβxγr(t,x,ξ)|\displaystyle|\partial^{\beta}_{\xi}\partial^{\gamma}_{x}r(t,x,\xi)| tmin(s,0)ξs+m(1max(δ,1/2)ε)Mρ|β|+δ|γ|.\displaystyle\lesssim t^{\min(s,0)}{\langle{\xi}\rangle}^{s+m-(1-\max(\delta,1/2)-\varepsilon)M-\rho|\beta|+\delta|\gamma|}.
Proof.

For a proof of this result see [14]. ∎

2.2. Some facts from the theory of classical function spaces

Classical function spaces, such as Triebel-Lizorkin spaces and Besov-Lipschitz spaces, are central to the study of partial differential equations and the analysis of functions. In this section, we review some basic facts and results from the theory of classical function spaces that will be used throughout the rest of our paper.

We begin by introducing the definitions of classical Littlewood-Paley operators, Triebel-Lizorkin spaces, and Besov-Lipschitz spaces and discussing some of their key properties. Next, we present some important lemmas and theorems that provide useful estimates and bounds for functions in these spaces. We also introduce the Hardy-Littlewood maximal function, and how this operator can be leveraged in the analysis of various mathematical problems that will arise in the subsequent sections.

Definition 2.13.

Let ψ0𝒞c(n)\psi_{0}\in\mathcal{C}_{c}^{\infty}(\mathbb{R}^{n}) be equal to 11 on B(0,1)B(0,1) and have its support in B(0,2)B(0,2). Then let

ψj(ξ):=ψ0(2jξ)ψ0(2(j1)ξ),\psi_{j}(\xi):=\psi_{0}\left(2^{-j}\xi\right)-\psi_{0}\left(2^{-(j-1)}\xi\right),

where j1j\geq 1 is an integer and ψ(ξ):=ψ1(ξ)\psi(\xi):=\psi_{1}(\xi). Then ψj(ξ)=ψ(2(j1)ξ)\psi_{j}(\xi)=\psi\left(2^{-(j-1)}\xi\right) and one has the following Littlewood-Paley partition of unity

j=0ψj(ξ)=1,for all ξn.\sum_{j=0}^{\infty}\psi_{j}(\xi)=1,\quad\text{\emph{for all }}\xi\in\mathbb{R}^{n}.

It is sometimes also useful to define a sequence of smooth and compactly supported functions Ψj\Psi_{j} with Ψj=1\Psi_{j}=1 on the support of ψj\psi_{j} and Ψj=0\Psi_{j}=0 outside a slightly larger compact set. One could for instance set

Ψj:=ψj+1+ψj+ψj1,\Psi_{j}:=\psi_{j+1}+\psi_{j}+\psi_{j-1},

with ψ1:=ψ0\psi_{-1}:=\psi_{0}.

In what follows we define the Littlewood-Paley operators by

ψj(D)f(x)=nψj(ξ)f^(ξ)eixξđξ,\psi_{j}(D)\,f(x)=\int_{\mathbb{R}^{n}}\psi_{j}(\xi)\,\widehat{f}(\xi)\,e^{ix\cdot\xi}\,\text{\rm{\mbox{\dj}}}\xi,

where đξ\,\text{\rm{\mbox{\dj}}}\xi denotes the normalised Lebesgue measure dξ/(2π)n{\,\mathrm{d}\xi}/{(2\pi)^{n}} and

f^(ξ)=neixξf(x)dx,\widehat{f}(\xi)=\int_{\mathbb{R}^{n}}e^{-ix\cdot\xi}\,f(x)\,\mathrm{d}x,

is the Fourier transform of ff.

Define Lp(q)\|\cdot\|_{L^{p}(\ell^{q})} and q(Lp)\|\cdot\|_{\ell^{q}(L^{p})} to be the quasi-norms

(fk)Lp(q)\displaystyle\|(f_{k})\|_{L^{p}(\ell^{q})} :=k|fk()|qLp(n)1/q,\displaystyle:=\Big{\|}\sum_{k}|f_{k}(\cdot)|^{q}\Big{\|}_{L^{p}(\mathbb{R}^{n})}^{1/q},
(fk)q(Lp)\displaystyle\|(f_{k})\|_{\ell^{q}(L^{p})} :=(kfkLp(n)q)1/q.\displaystyle:=\Big{(}\sum_{k}\|f_{k}\|_{L^{p}(\mathbb{R}^{n})}^{q}\Big{)}^{1/q}.

Using the Littlewood-Paley decomposition of Definition 2.13 we define the Triebel-Lizorkin space Fp,qs(n)F^{s}_{p,q}(\mathbb{R}^{n}) and Besov-Lipschitz space Bp,qs(n)B^{s}_{p,q}(\mathbb{R}^{n}).

Definition 2.14.

Let ss\in{\mathbb{R}} and 0<p<0<p<\infty, 0<q0<q\leq\infty. The Triebel-Lizorkin space is defined by

Fp,qs(n):={f𝒮(n):fFp,qs(n):=(2jqsψj(D)f)Lp(q)<},F^{s}_{p,q}(\mathbb{R}^{n}):=\Big{\{}f\in{\mathscr{S}^{\prime}}(\mathbb{R}^{n})\,:\,\|f\|_{F^{s}_{p,q}(\mathbb{R}^{n})}:=\|(2^{jqs}\psi_{j}(D)f)\|_{L^{p}(\ell^{q})}<\infty\Big{\}},

where 𝒮(n)\mathscr{S}^{\prime}(\mathbb{R}^{n}) denotes the space of tempered distributions.

Definition 2.15.

Let 0<p,q0<p,q\leq\infty and ss\in{\mathbb{R}}. The Besov-Lipschitz spaces are defined by

Bp,qs(n):={fSS(n):fBp,qs(n):=(2jqsψj(D)f)q(Lp)<}.{B}^{s}_{p,q}(\mathbb{R}^{n}):=\Big{\{}f\in{{\SS}^{\prime}(\mathbb{R}^{n})}\,:\,\|f\|_{{B}^{s}_{p,q}(\mathbb{R}^{n})}:=\|(2^{jqs}\psi_{j}(D)f)\|_{\ell^{q}(L^{p})}<\infty\Big{\}}.
Remark 2.16.

Different choices of the basis {ψj}j=0\{\psi_{j}\}_{j=0}^{\infty} give equivalent norms of Fp,qs(n)F^{s}_{p,q}(\mathbb{R}^{n}) in Definition 2.14, see e.g. [20]. We will use either {ψj}j=0\{\psi_{j}\}_{j=0}^{\infty} or {Ψj}j=0\{\Psi_{j}\}_{j=0}^{\infty} to define the norm of Fp,qs(n)F^{s}_{p,q}(\mathbb{R}^{n}).

We note that for p=q=p=q=\infty and 0<s10<s\leq 1 we obtain the familiar Lipschitz space Λs(n)\Lambda^{s}(\mathbb{R}^{n}), i.e. B,s(n)=Λs(n)B^{s}_{\infty,\infty}(\mathbb{R}^{n})=\Lambda^{s}(\mathbb{R}^{n}). For <s<-\infty<s<\infty and 1p<,1\leq p<\infty, Fp,2s(n)=Hs,p(n)F^{s}_{p,2}(\mathbb{R}^{n})=H^{s,p}(\mathbb{R}^{n}) (various LpL^{p}-based Sobolev and Sobolev-Slobodeckij spaces) and for 0<p<0<p<\infty, Fp,20(n)=hp(n)F^{0}_{p,2}(\mathbb{R}^{n})=h^{p}(\mathbb{R}^{n}) (the local Hardy spaces). Moreover the dual space of F1,20(n)F^{0}_{1,2}(\mathbb{R}^{n}) is bmo\mathrm{bmo} (the local version of BMO\mathrm{BMO}).

The following estimate will be useful in the L1(n)L^{1}(\mathbb{R}^{n}) to bmo(n)bmo(\mathbb{R}^{n}) boundedness of Oscillatory integral operators. Let Xp,qs(n)X^{s}_{p,q}(\mathbb{R}^{n}) be either Bp,qs(n)B^{s}_{p,q}(\mathbb{R}^{n}) or Fp,qs(n)F^{s}_{p,q}(\mathbb{R}^{n}). Then for all <s<-\infty<s<\infty and 0<p,q0<p,q\leq\infty one has

(11) fgXp,qs(n)(|α|Msupxn|αf(x)|)gXp,qs(n).\|fg\|_{X^{s}_{p,q}(\mathbb{R}^{n})}\lesssim\Big{(}\sum_{|\alpha|\leq M}\sup_{x\in\mathbb{R}^{n}}|\partial^{\alpha}f(x)|\Big{)}\|g\|_{X^{s}_{p,q}(\mathbb{R}^{n})}.

Two other useful facts which will be useful to us is that for <s<-\infty<s<\infty and 0<p0<p\leq\infty one has

(12) Bp,ps(n)=Fp,ps(n),B^{s}_{p,p}(\mathbb{R}^{n})=F^{s}_{p,p}(\mathbb{R}^{n}),

and

(13) Fp,q0s+ε(n)Fp,q1s(n)andBp,q0s+ε(n)Bp,q1s(n)F^{s+\varepsilon}_{p,q_{0}}(\mathbb{R}^{n})\xhookrightarrow{}F^{s}_{p,q_{1}}(\mathbb{R}^{n})\quad\text{and}\quad B^{s+\varepsilon}_{p,q_{0}}(\mathbb{R}^{n})\xhookrightarrow{}B^{s}_{p,q_{1}}(\mathbb{R}^{n})

for <s<-\infty<s<\infty, 0<p<0<p<\infty, 0<q0,q10<q_{0},q_{1}\leq\infty and all ε>0\varepsilon>0.

Furthermore, for ss^{\prime}\in\mathbb{R}, the operator (1Δ)s/2(1-\Delta)^{s^{\prime}/2} maps Fp,qs(n){F}^{s}_{p,q}(\mathbb{R}^{n}) isomorphically into Fp,qss(n){F}^{s-s^{\prime}}_{p,q}(\mathbb{R}^{n}) and Bp,qs(n){B}^{s}_{p,q}(\mathbb{R}^{n}) isomorphically into Bp,qss(n),{B}^{s-s^{\prime}}_{p,q}(\mathbb{R}^{n}), see [20, p. 58].

In connection to estimates for linear operators in Triebel-Lizorkin spaces, one often encounters the well-known Hardy-Littlewood’s maximal function

f(x):=supBx1|B|B|f(y)|dy,\mathcal{M}f(x):=\sup_{B\ni x}\frac{1}{|B|}\int_{B}|f(y)|\,\mathrm{d}y,

where the supremum is taken over all balls BB containing xx. For 0<p<0<p<\infty, one also defines

(pf(x):=((|f|p))1/p.(\mathcal{M}_{p}f(x):=\left(\mathcal{M}\left(|f|^{p}\right)\right)^{1/p}.

We now present the following abstract lemma which we will make use of in relation to the boundedness of OIOs with amplitudes in the class S1,1m(n)S^{m}_{1,1}(\mathbb{R}^{n}).

Lemma 2.17.

Let Xp,qsX^{s}_{p,q} be either Fp,qs(n)F^{s}_{p,q}(\mathbb{R}^{n}) or Bp,qs(n)B^{s}_{p,q}(\mathbb{R}^{n}). Assume that {uk}k0Xp,qs\{u_{k}\}_{k\geq 0}\in X^{s}_{p,q} is an arbitrary sequence such that {2ksuk}k=0Lp(lq)fXp,qs\big{\|}\{2^{ks}u_{k}\}_{k=0}^{\infty}\big{\|}_{{}_{L^{p}(l^{q})}}\lesssim\|f\|_{X^{s}_{p,q}}. Moreover let {hk,l(x)}k,l0\{h_{k,l}(x)\}_{k,l\geq 0} be a sequence such that the spectrum of hk,l(x)h_{k,l}(x) is in B(0,2k+l+2)B(0,2^{k+l+2}) and that there is some sufficiently large NN such that hk,lh_{k,l} satisfies the pointwise estimate,

(14) |hk,l(x)|2Nlruk(x),r(0,min{p,q}).|h_{k,l}(x)|\lesssim 2^{-Nl}\,\mathcal{M}_{r}u_{k}(x),\quad\exists r\in(0,\min\{p,q\}).

Let gl=k0hk,lg_{l}=\sum_{k\geq 0}h_{k,l}. If 0<p,q<0<p,q<\infty and s>n(1min(p,q,1)1)s>n(\frac{1}{\mathrm{min}(p,q,1)}-1) then

(15) l0glXp,qsfXp,qs.\Big{\|}\sum_{l\geq 0}g_{l}\Big{\|}_{X^{s}_{p,q}}\lesssim\|f\|_{X^{s}_{p,q}}.
Proof.

For a proof see [14]. ∎

Another important and useful fact about Besov-Lipschitz and Triebel-Lizorkin spaces is the following:

Theorem 2.18.

Let η:nn\eta:\mathbb{R}^{n}\to\mathbb{R}^{n} with η(x)=(η1(x),,ηn(x))\eta(x)=(\eta_{1}(x),\dots,\eta_{n}(x)) be a diffeomorphism such that |detDη(x)|c>0\Big{|}\det D\eta(x)\Big{|}\geq c>0, xn\forall x\in\mathbb{R}^{n} (DηD\eta denotes the Jacobian matrix of η\eta), and αηj(x)L(n)1\|\partial^{\alpha}\eta_{j}(x)\|_{L^{\infty}(\mathbb{R}^{n})}\lesssim 1 for all j{1,,n}j\in\{1,\dots,n\} and |α|1.|\alpha|\geq 1. Then for ss\in\mathbb{R}, 0<p<0<p<\infty and 0<q0<q\leq\infty one has

fηFp,qs(n)fFp,qs(n).\|f\circ\eta\|_{F^{s}_{p,q}(\mathbb{R}^{n})}\lesssim\|f\|_{F^{s}_{p,q}(\mathbb{R}^{n})}.

The same invariance estimate is also true for Besov-Lipschitz spaces Bp,qs(n)B^{s}_{p,q}(\mathbb{R}^{n}) for ss\in\mathbb{R}, 0<p0<p\leq\infty and 0<q0<q\leq\infty.

For a proof see J. Johnsen, S. Munch Hansen and W. Sickel [16, Corollary 25], and H. Triebel [21, Theorem 4.3.2].

The rest of this section is dedicated to setting up some definitions which will be used in relation to transference to Triebel-Lizorkin spaces (section 6) and hpLph^{p}\to L^{p} estimates of oscillatory integral operators (section 5).

For Definition 2.19 and Definition 2.20, let [x][x] denote the integer part of xx.

Definition 2.19.

For a closed cube QQ, define

  1. (i)(i)

    cQc_{Q} as the centre of QQ,

  2. (ii)(ii)

    lQl_{Q} as the side length of QQ,

  3. (iii)(iii)

    kQ:=[1log2(lQ)]k_{Q}:=[1-\log_{2}(l_{Q})],

  4. (iv)(iv)

    χQ\chi_{Q} as the characteristic function of QQ, i.e. χQ(x):={1,xQ0,xQ\chi_{Q}(x):=\begin{cases}1,&x\in Q\\ 0,&x\notin Q\end{cases}.

  5. (v)(v)

    cQcQ as the cube with centre cQc_{Q} and length lcQ=clQl_{c\,Q}=c\,l_{Q}, where c>0c\in\mathbb{R}_{>0}.

Observe that kQk_{Q} is the unique integer such that

2kQ<lQ2(kQ1).2^{-k_{Q}}<l_{Q}\leq 2^{-(k_{Q}-1)}.
Definition 2.20.

A function 𝔞\mathfrak{a} is called a hph^{p}-atom if there exists a cube QQ such that the following three conditions are satisfied:

  1. (i)(i)

    supp𝔞Q\operatorname{supp}\mathfrak{a}\subset Q,

  2. (ii)(ii)

    supxQ|𝔞(x)|2kQn/p,\displaystyle\sup_{x\in Q}|\mathfrak{a}(x)|\leq 2^{k_{Q}n/p},

  3. (iii)(iii)

    If kQ1k_{Q}\geq 1, 𝔐𝔞=[n(1p1)]\displaystyle\mathfrak{M}_{\mathfrak{a}}=\left[n\Big{(}{\frac{1}{p}-1}\Big{)}\right], then nxα𝔞(x)dx=0,\displaystyle\int_{\mathbb{R}^{n}}x^{\alpha}\mathfrak{a}(x)\,\mathrm{d}x=0, for |α|𝔐𝔞|\alpha|\leq\mathfrak{M}_{\mathfrak{a}}. No further condition is assumed if kQ0.k_{Q}\leq 0.

It is well known ((see [20])) that a distribution fhp(n)f\in h^{p}(\mathbb{R}^{n}) has an atomic decomposition

(16) f=j=0λj𝔞j,\begin{split}f=\sum_{j=0}^{\infty}\lambda_{j}\mathfrak{a}_{j},\end{split}

where the λj\lambda_{j} are constants such that

inf{λj}j=0|λj|pfhp(n)p=fFp,20(n)p\displaystyle\inf_{\left\{{\lambda_{j}}\right\}}\sum_{j=0}^{\infty}|\lambda_{j}|^{p}\sim\|f\|_{h^{p}(\mathbb{R}^{n})}^{p}=\|f\|_{F_{p,2}^{0}(\mathbb{R}^{n})}^{p}

and the 𝔞j\mathfrak{a}_{j} are hph^{p}-atoms.

In the analysis of the boundedness of OIOs on classical function spaces one typically decomposes the operator into a low- and a high-frequency part, where the low frequency part corresponds to an amplitude that is smooth and compactly supported in the ξ\xi variable. The following theorem, which was proven in [3], addresses the boundedness of the low-frequency portion of the OIOs. The main boundedness result for the low frequency part of OIOs.

Theorem 2.21 ([3]).

For ρ,δ[0,1]\rho,\delta\in[0,1] let a(x,ξ)Sρ,δm(n)a(x,\xi)\in S^{m}_{\rho,\delta}(\mathbb{R}^{n}) be an amplitude which is compactly supported in ξ\xi. Suppose also that 0<q1,q20<q_{1},q_{2}\leq\infty, s1,s2s_{1},s_{2}\in\mathbb{R}. If φ(x,ξ)\varphi(x,\xi) verifies the LF(μ\mu)-condition, then TaφT_{a}^{\varphi} is Fp,q1s1(n)F_{p,q_{1}}^{s_{1}}(\mathbb{R}^{n}) to Fp,q2s2(n)F_{p,q_{2}}^{s_{2}}(\mathbb{R}^{n}) bounded, for nn+μ<p\frac{n}{n+\mu}<p\leq\infty Moreover, all the Triebel-Lizorkin estimates above may be replaced by the corresponding Besov-Lipschitz estimate.

Proof.

See [3, Lemma 6.3]. ∎

Here we recall a theorem that allows one to lift hpLph^{p}\to L^{p} boundedness to Besov-Lipschitz boundedness. This result can be applied to a wide class of oscillatory integral operators. Observe that this lift is only valid for the classical amplitudes when δ<1\delta<1.

Theorem 2.22 ([15]).

Let 0<p,q<0<p,q<\infty, m0m\leq 0, aSρ,δm(n)a\in S^{m}_{\rho,\delta}(\mathbb{R}^{n}) such that aa is supported in n×nB(0,1)\mathbb{R}^{n}\times\mathbb{R}^{n}\setminus B(0,1). Suppose φ\varphi is a phase function that verifies the conditions (9) of Theorem 2.12. If TaφT^{\varphi}_{a} is hpLph^{p}\to L^{p}-bounded, then TaφT^{\varphi}_{a} is bounded from Bp,qsBp,qsB^{s}_{p,q}\to B^{s}_{p,q} for all ss\in\mathbb{R}.

Proof.

For a proof see [15]. ∎

The following theorem due to M. Pramanik, K. Rogers and A. Seeger could be found in [18, Theorem 2.1].

Theorem 2.23 ([18]).

Let 1<q<p<1<q<p<\infty, and 0<b<n0<b<n. Assume that the sequence of operators SjS_{j} satisfy

(17) supj>02jb/pSjLpLpA0,\displaystyle\sup_{j>0}2^{jb/p}\|S_{j}\|_{L^{p}\to L^{p}}\leq A_{0},
(18) supj>02jb/qSjLqLqB0.\displaystyle\sup_{j>0}2^{jb/q}\|S_{j}\|_{L^{q}\to L^{q}}\leq B_{0}.

Furthermore, assume that for each cube QQ there is a measurable set Q\mathcal{E}_{Q} and a constant Γ1\Gamma\geq 1 such that

(19) |Q|Γmax{lQnb,lQn},|\mathcal{E}_{Q}|\leq\Gamma\max\{l_{Q}^{n-b},l_{Q}^{n}\},

where lQl_{Q} is the side length of QQ as in Definition 2.19, and assume further that for every jj\in\mathbb{N} and every cube QQ with 2jlQ12^{j}l_{Q}\geq 1, one has

(20) supxQnQ|Kj(x,y)|dyB1max{2jεlQε,2jε},\begin{split}\sup_{x\in Q}\int_{\mathbb{R}^{n}\setminus\mathcal{E}_{Q}}|K_{j}(x,y)|\,\mathrm{d}y\leq B_{1}\max\{2^{-j\varepsilon}l_{Q}^{-\varepsilon},2^{-j\varepsilon}\},\end{split}

for some ε>0\varepsilon>0.
Then if

(21) B2:=B0q/p(A0Γ1/p+B1)1q/p{B}_{2}:=B_{0}^{q/p}\left(A_{0}\Gamma^{1/p}+B_{1}\right)^{1-q/p}

one has

(j=02jbr/p|Ψj(D)Sjfj|r)1/rLp(n)A0[log(3+B2A0)]1/r1/p(j=0fjLp(n)p)1/p.\displaystyle\Big{\|}\Big{(}{\sum_{j=0}^{\infty}2^{jbr/p}|\Psi_{j}(D)S_{j}f_{j}|^{r}}\Big{)}^{1/r}\Big{\|}_{L^{p}(\mathbb{R}^{n})}\lesssim A_{0}\left[\log\left(3+\frac{{B}_{2}}{A_{0}}\right)\right]^{1/r-1/p}\Big{(}{\sum_{j=0}^{\infty}\|f_{j}\|_{L^{p}(\mathbb{R}^{n})}^{p}}\Big{)}^{1/p}.

where Ψ𝒮(n)\Psi\in\mathscr{S}(\mathbb{R}^{n}), Ψj(D):=Ψ(2jD)\Psi_{j}(D):=\Psi(2^{-j}D) and fjf_{j} is a sequence of functions.

3. Basic kernel estimates for oscillatory integral operators

In order to show the Besov-Lipschitz boundedness of OIOs related to forbidden amplitudes we prove some preliminary estimates on the kernel of the operators. What follows is a decomposition which, in contrast to the second angular-radii frequency decomposition of C. Fefferman (see [7]), decomposes the annuli of the standard Littlewood-Paley decomposition into balls. Furthermore, this decomposition generalizes the decomposition in [3], which used balls of constant radii to decompose the case corresponding to the OIOs of Schrödinger type phase functions in F2\textart{F}^{2}, to the more general case of balls of varying radii adapted to all OIOs with phases in the class Fk\textart{F}^{k}.

ξjν\xi_{j}^{\nu}
Figure 1. Covering of suppψj{2j1|ξ|2j+1}\operatorname{supp}\psi_{j}\subset\{2^{j-1}\leq|\xi|\leq 2^{j+1}\} with balls of radii 2j(1k/2)2^{j(1-k/2)} and centres ξjν\xi_{j}^{\nu}.
Definition 3.1.

Let k>0k>0. We make the following decomposition of the integral kernel

K(x,y)=na(x,ξ)eiφ(x,ξ)iyξđξ.K(x,y)=\int_{\mathbb{R}^{n}}a(x,\xi)\,e^{i\varphi(x,\xi)-iy\cdot\xi}\,\,\text{\rm{\mbox{\dj}}}\xi.

We introduce a standard Littlewood-Paley partition of unity j=0ψj(ξ)=1\sum_{j=0}^{\infty}\psi_{j}(\xi)=1 with suppψ0B(0,2)\operatorname{supp}\psi_{0}\subset B(0,2), and suppψj{2j1|ξ|2j+1}\operatorname{supp}\psi_{j}\subset\{2^{j-1}\leq|\xi|\leq 2^{j+1}\} for j1j\geq 1 as in Definition 2.13.

Then for every j0j\geq 0 we cover suppψj\operatorname{supp}\psi_{j} with open balls CjνC_{j}^{\nu} with radii 2j(1k)2^{j(1-k)} and centres ξjν\xi_{j}^{\nu}, where ν\nu runs from 11 to 𝒩j:=O(2njk)\mathscr{N}_{j}:=O(2^{njk}). See Figure 1 for an illustration. Observe that |Cjν|2nj(1k)|C_{j}^{\nu}|\lesssim 2^{nj(1-k)} uniformly in ν\nu. Now take u𝒞c(n)u\in\mathcal{C}_{c}^{\infty}(\mathbb{R}^{n}), with 0u10\leq u\leq 1 and supported in B(0,2)B(0,2) with u=1u=1 on B(0,1)¯.\overline{B(0,1)}.

Next set

χjν(ξ):=u(2(1k)j(ξξjν))κ=1𝒩ju(2(1k)j(ξξjκ)),\chi_{j}^{\nu}(\xi):=\frac{u(2^{-(1-k)j}(\xi-\xi_{j}^{\nu}))}{\sum_{\kappa=1}^{\mathscr{N}_{j}}u(2^{-(1-k)j}(\xi-\xi_{j}^{\kappa}))},

and note that

j=0ν=1𝒩jχjν(ξ)ψj(ξ)=1.\sum_{j=0}^{\infty}\sum_{\nu=1}^{\mathscr{N}_{j}}\chi_{j}^{\nu}(\xi)\,\psi_{j}(\xi)=1.

Now we define the second frequency localized pieces of the kernel above as

Kjν(x,y):=nψj(ξ)χjν(ξ)eiφ(x,ξ)iyξa(x,ξ)đξ.K_{j}^{\nu}(x,y):=\int_{\mathbb{R}^{n}}\psi_{j}(\xi)\,\chi_{j}^{\nu}(\xi)\,e^{i\varphi(x,\xi)-iy\cdot\xi}\,a(x,\xi)\,\text{\rm{\mbox{\dj}}}\xi.
Lemma 3.2.

Let 0ρ10\leq\rho\leq 1, 0δ<10\leq\delta<1, n1n\geq 1, and 0<k<0<k<\infty. Assume that φFk\varphi\in\textart{F}^{k} is SND, satisfies the L2L^{2}-condition (3). Then if a(x,ξ)LSρm(n)a(x,\xi)\in L^{\infty}S^{m}_{\rho}(\mathbb{R}^{n}) ((see Definition 2.2)), and Kjν(x,y)K_{j}^{\nu}(x,y) is defined as in Definition 3.1, then we have

(22) |yαKjν(x,y)|2j(m+|α|)2jn(1k)2jw(k,ρ)(ξφ(x,ξjν)y)M,|\partial_{y}^{\alpha}K_{j}^{\nu}(x,y)|\lesssim\frac{2^{j(m+|\alpha|)}2^{jn(1-k)}}{\left\langle 2^{jw(k,\rho)}(\nabla_{\xi}\varphi(x,\xi_{j}^{\nu})-y)\right\rangle^{M}},

for all multi-indices α\alpha, all j0j\geq 0 and M0M\geq 0 and where

w(k,ρ)={min{ρ,1k}0<k<1,k1k1.w(k,\rho)=\begin{cases}\min\{\rho,1-k\}&0<k<1,\\ k-1&k\geq 1.\\ \end{cases}
Proof.

Observe that we have

|ξγχjν(ξ)|2j(k1)|γ||\partial^{\gamma}_{\xi}\chi_{j}^{\nu}(\xi)|\lesssim 2^{j(k-1)|\gamma|}

Therefore, for any multi-index α\alpha and any j0j\geq 0 we have

yαKjν(x,y)\displaystyle\partial^{\alpha}_{y}K_{j}^{\nu}(x,y) =nψj(ξ)χjν(ξ)yαei(φ(x,ξ)yξ)a(x,ξ)đξ\displaystyle=\int_{\mathbb{R}^{n}}\psi_{j}(\xi)\,\chi_{j}^{\nu}(\xi)\,\partial^{\alpha}_{y}e^{i(\varphi(x,\xi)-y\cdot\xi)}\,a(x,\xi)\,\text{\rm{\mbox{\dj}}}\xi
=nei(φ(x,ξ)yξ)σjα,ν(x,ξ)đξ,\displaystyle=\int_{\mathbb{R}^{n}}e^{i(\varphi(x,\xi)-y\cdot\xi)}\,\sigma_{j}^{\alpha,\nu}(x,\xi)\,\text{\rm{\mbox{\dj}}}\xi,

where

σjα,ν(x,ξ):=ψj(ξ)χjν(ξ)(iξ)αa(x,ξ).\sigma_{j}^{\alpha,\nu}(x,\xi):=\psi_{j}(\xi)\,\chi_{j}^{\nu}(\xi)\,(-i\xi)^{\alpha}\,a(x,\xi).

Using the assumption that a(x,ξ)LSρm(n)a(x,\xi)\in L^{\infty}S^{m}_{\rho}(\mathbb{R}^{n}), we deduce that for any multi-index γ\gamma, any j0j\geq 0 and any ν\nu one has

(23) |ξγσjα,ν(x,ξ)|2j(m+|α|min{ρ,1k}|γ|),|\partial^{\gamma}_{\xi}\sigma_{j}^{\alpha,\nu}(x,\xi)|\lesssim 2^{j(m+|\alpha|-\min\{\rho,1-k\}|\gamma|)},

If we now set ϑjν(x,ξ):=φ(x,ξ)ξξφ(x,ξjν)\vartheta_{j}^{\nu}(x,\xi):=\varphi(x,\xi)-\xi\cdot\nabla_{\xi}\varphi(x,\xi_{j}^{\nu}), then we can write

yαKjν(x,y)=nei(ξφ(x,ξjν)y)ξeiϑjν(x,ξ)σjα,ν(x,ξ)đξ.\partial^{\alpha}_{y}K_{j}^{\nu}(x,y)=\int_{\mathbb{R}^{n}}e^{i(\nabla_{\xi}\varphi(x,\xi_{j}^{\nu})-y)\cdot\xi}\,e^{i\vartheta_{j}^{\nu}(x,\xi)}\,\sigma_{j}^{\alpha,\nu}(x,\xi)\,\text{\rm{\mbox{\dj}}}\xi.

Now we estimate the derivatives of ϑ\vartheta in ξ\xi on the support of σjα,ν(x,ξ)\sigma_{j}^{\alpha,\nu}(x,\xi). To this end, the mean-value theorem and Definition 2.3 yields for k1k\geq 1

|ξlϑjν(x,ξ)|\displaystyle|\partial_{\xi_{l}}\vartheta_{j}^{\nu}(x,\xi)| =|ξlφ(x,ξ)ξlφ(x,ξjν)|=|(ξξjν)01(ξξlφ)(x,tξ+(1t)ξjν)dt|\displaystyle=|\partial_{\xi_{l}}\varphi(x,\xi)-\partial_{\xi_{l}}\varphi(x,\xi_{j}^{\nu})|=\Big{|}(\xi-\xi_{j}^{\nu})\cdot\int_{0}^{1}(\nabla_{\xi}\partial_{\xi_{l}}\varphi)(x,t\xi+(1-t)\xi_{j}^{\nu})\,\mathrm{d}t\Big{|}
2j(1k)supt[0,1]|tξ+(1t)ξjν)|k1\displaystyle\lesssim 2^{j(1-k)}\sup_{t\in[0,1]}|t\xi+(1-t)\xi_{j}^{\nu})|^{k-1}
2j(1k)2j(k1)=1\displaystyle\lesssim 2^{j(1-k)}2^{j(k-1)}=1

and

|ξαϑjν(x,ξ)|=|ξαφ(x,ξ)|2j|α|max{(k1),0}, for all |α|2.\displaystyle|\partial_{\xi}^{\alpha}\vartheta_{j}^{\nu}(x,\xi)|=|\partial_{\xi}^{\alpha}\varphi(x,\xi)|\lesssim 2^{j|\alpha|\max\{(k-1),0\}},\quad\text{ for all }|\alpha|\geq 2.

Therefore by Faa di Bruno’s formula we obtain that

(24) |ξγeiϑjν(ξ,x)|γ1++γr=γ|ξγ1ϑjν||ξγrϑjν|2jmax{(k1),0}|γ|.\displaystyle|\partial_{\xi}^{\gamma}e^{i\vartheta^{\nu}_{j}(\xi,x)}|\lesssim\sum_{\gamma_{1}+\dots+\gamma_{r}=\gamma}|\partial_{\xi}^{\gamma_{1}}\vartheta^{\nu}_{j}|\dots|\partial_{\xi}^{\gamma_{r}}\vartheta^{\nu}_{j}|\lesssim 2^{j\max\{(k-1),0\}|\gamma|}.

Observe the simple estimate

(25) |yαKjν(x,y)|2j(m+|α|)2jn(1k)|\partial^{\alpha}_{y}K_{j}^{\nu}(x,y)|\lesssim 2^{j(m+|\alpha|)}2^{jn(1-k)}

together with integration by parts yield

(26) |yαKjν(x,y)|2j(m+|α|+w(k,ρ)M)2jn(1k)|ξφ(x,ξjν)y|M.|\partial^{\alpha}_{y}K_{j}^{\nu}(x,y)|\lesssim\frac{2^{j(m+|\alpha|+w(k,\rho)M)}2^{jn(1-k)}}{|\nabla_{\xi}\varphi(x,\xi_{j}^{\nu})-y|^{M}}.

where w(k,ρ)=max{min{ρ,1k},max{(k1),0}}w(k,\rho)=\max\{-\min\{\rho,1-k\},\max\{(k-1),0\}\}. These equalities, the observation that suppσjα,νCjν,\operatorname{supp}\sigma_{j}^{\alpha,\nu}\subset C^{\nu}_{j}, with |Cjν|=O(2jn(1k))|C^{\nu}_{j}|=O(2^{jn(1-k)}) uniformly in ν\nu and jj, the estimates for the derivatives of ϑ\vartheta, and (23) yield

|yαKjν(x,y)|2j(m+|α|)2jn(1k)2jw(k,ρ)(ξφ(x,ξjν)y)M2j(m+|α|)2jn(1k)2jw(k,ρ)(ξφ(x,ξjν)y)M,|\partial_{y}^{\alpha}K_{j}^{\nu}(x,y)|\lesssim\frac{2^{j(m+|\alpha|)}2^{jn(1-k)}}{\left\langle 2^{jw(k,\rho)}(\nabla_{\xi}\varphi(x,\xi_{j}^{\nu})-y)\right\rangle^{M}}\lesssim\frac{2^{j(m+|\alpha|)}2^{jn(1-k)}}{\left\langle 2^{jw(k,\rho)}(\nabla_{\xi}\varphi(x,\xi_{j}^{\nu})-y)\right\rangle^{M}},

for all multi-indices α\alpha, all j0j\geq 0 and M0M\geq 0. ∎

4. L2L^{2}-boundedness of Oscillatory integral operators

Traditionally, the way to prove L2L^{2}-boundedness results for oscillatory integral operators at the endpoint of the range of exponents has been to utilize the almost orthogonality principle, as established in the continuous version of the Cotlar-Stein lemma by Calderón and Vaillancourt, see i.e. [1] for a proof. This lemma, which was originally developed by Cotlar and Stein, is a powerful tool for establishing bounds on oscillatory integral operators.

Lemma 4.1.

Let \mathscr{H} be a Hilbert space, and A(ξ)A(\xi) a family of bounded linear endomorphisms of \mathscr{H} depending on ξn.\xi\in\mathbb{R}^{n}. Assume the following three conditions hold:

  1. (i)(i)

    the operator norm of A(ξ)A(\xi) is less than a number CC independent of ξ.\xi.

  2. (ii)(ii)

    for every uu\in\mathscr{H} the function ξA(ξ)u\xi\mapsto A(\xi)u from n\mathbb{R}^{n}\mapsto\mathscr{H} is continuous for the norm topology of .\mathscr{H}.

  3. (iii)(iii)

    for all ξ1\xi_{1} and ξ2\xi_{2} in n\mathbb{R}^{n}

    (27) A(ξ1)A(ξ2)h(ξ1,ξ2)2,andA(ξ1)A(ξ2)h(ξ1,ξ2)2,\|A^{\ast}(\xi_{1})A(\xi_{2})\|\leq h(\xi_{1},\xi_{2})^{2},\,\,\,\text{and}\,\,\,\|A(\xi_{1})A^{\ast}(\xi_{2})\|\leq h(\xi_{1},\xi_{2})^{2},

    with h(ξ1,ξ2)0h(\xi_{1},\xi_{2})\geq 0 is the kernel of a bounded linear operator on L2L^{2} with norm KK.

Then for every EnE\subset\mathbb{R}^{n}, with |E|<|E|<\infty, the operator AE=EA(ξ)đξA_{E}=\int_{E}A(\xi)\,\text{\rm{\mbox{\dj}}}\xi defined by AEu,v=EA(ξ)u,vđξ,\langle A_{E}u,v\rangle_{\mathscr{H}}=\int_{E}\langle A(\xi)u,v\rangle_{\mathscr{H}}\,\text{\rm{\mbox{\dj}}}\xi, is a bounded linear operator on \mathscr{H} with norm less than or equal to K.K.

Another useful fact that will aid us in the estimate of the oscillatory integrals is the following lemma whose proof could be found in [6].

Lemma 4.2.

Let s(x)s(x) and F(x)F(x) be real-valued smooth functions in n\mathbb{R}^{n}, and

(28) Lu(x):=D2(1is(x)xF,x)u(x),Lu(x):=D^{-2}(1-is(x)\langle\nabla_{x}F,\nabla_{x}\rangle)u(x),

with D:=(1+s(x)|xF|2)1/2.D:=(1+s(x)|\nabla_{x}F|^{2})^{1/2}. Then

  1. (i)(i)

    L(eiF(x))=eiF(x)L(e^{iF(x)})=e^{iF(x)}

  2. (ii)(ii)

    if Lt{}^{t}L denotes the formal transpose of L,L, then for any positive integer N,N, (Lt)Nu(x)({}^{t}L)^{N}u(x) is a finite linear combination of terms of the form

    (29) CDk{μ=1pxαμs(x)}{ν=1qxβνF(x)}xγu(x),CD^{-k}\Big{\{}\prod_{\mu=1}^{p}\partial^{\alpha_{\mu}}_{x}s(x)\Big{\}}\Big{\{}\prod_{\nu=1}^{q}\partial^{\beta_{\nu}}_{x}F(x)\Big{\}}\partial^{\gamma}_{x}u(x),

    with

    (30) 2Nk4N;k2NpkN;|αμ|0;μ=1p|αμ|Nk2NqkN;|βν|1;μ=1q|βν|q+N;|γ|N.2N\leq k\leq 4N;\,k-2N\leq p\leq k-N;\,|\alpha_{\mu}|\geq 0;\,\sum_{\mu=1}^{p}|\alpha_{\mu}|\leq N\\ k-2N\leq q\leq k-N;\,|\beta_{\nu}|\geq 1;\,\sum_{\mu=1}^{q}|\beta_{\nu}|\leq q+N;\,|\gamma|\leq N.
Theorem 4.3.

If m=min(0,n2(ρδ)),m=\min(0,\frac{n}{2}(\rho-\delta)), 0ρ10\leq\rho\leq 1, 0δ<1,0\leq\delta<1, aSρ,δm(n)a\in S^{m}_{\rho,\delta}(\mathbb{R}^{n}) and φ𝒞(n×n)\varphi\in\mathcal{C}^{\infty}(\mathbb{R}^{n}\times\mathbb{R}^{n}) is SND for all (x,ξ)n×n(x,\xi)\in\mathbb{R}^{n}\times\mathbb{R}^{n} and satisfies the weak L2L^{2}-condition (2) on the support of a(x,ξ)a(x,\xi), then the operator

Taφf(x)=na(x,ξ)eiφ(x,ξ)f^(ξ)đξT^{\varphi}_{a}f(x)=\int_{\mathbb{R}^{n}}a(x,\xi)\,e^{i\varphi(x,\xi)}\hat{f}(\xi)\,\text{\rm{\mbox{\dj}}}\xi

is bounded on L2(n).L^{2}(\mathbb{R}^{n}).

Proof.

Let χ(x,ξ)𝒞c(n×n)\chi(x,\xi)\in\mathcal{C}_{c}^{\infty}(\mathbb{R}^{n}\times\mathbb{R}^{n}) be such that χ(x,ξ)=1\chi(x,\xi)=1 for |x|2+|ξ|21|x|^{2}+|\xi|^{2}\leq 1 and define aε(x,ξ):=χ(εx,εξ)a(x,ξ).a_{\varepsilon}(x,\xi):=\chi(\varepsilon x,\varepsilon\xi)\,a(x,\xi). Since aεa_{\varepsilon} converges to aa in Sρ,δm(n)S^{m}_{\rho,\delta}(\mathbb{R}^{n}) so that for any f𝒮(n)f\in\mathscr{S}(\mathbb{R}^{n}) , TaεφfT_{a_{\varepsilon}}^{\varphi}f converges, as ε0\varepsilon\to 0, to TaφfT_{a}^{\varphi}f in 𝒮(n)\mathscr{S}(\mathbb{R}^{n}). Since the seminorms of aεa_{\varepsilon} are bounded by a constant (depending on χ\chi) times the seminorms of aa, we can therefore assume from now on that a(x,ξ)𝒞c(n×n).a(x,\xi)\in\mathcal{C}_{c}^{\infty}(\mathbb{R}^{n}\times\mathbb{R}^{n}). Later on, of course, the estimates that we obtain won’t depend on the support of aa.

Furthermore, we observe that since for δρ\delta\leq\rho, Sρ,δ0(n)Sρ,ρ0(n),S^{0}_{\rho,\delta}(\mathbb{R}^{n})\subset S^{0}_{\rho,\rho}(\mathbb{R}^{n}), it is enough to show the theorem for 0ρδ<10\leq\rho\leq\delta<1 and m=n2(ρδ).m=\frac{n}{2}(\rho-\delta). Using the unitarity of the Fourier transform in L2(n)L^{2}(\mathbb{R}^{n}) and a TTTT^{\ast} argument, it is enough to show that the operator

(31) Tbf(x)=b(x,y,ξ)eiφ(x,ξ)iφ(y,ξ)f(y)dyđξ,T_{b}f(x)=\iint b(x,y,\xi)\,e^{i\varphi(x,\xi)-i\varphi(y,\xi)}\,f(y)\,\mathrm{d}y\,\text{\rm{\mbox{\dj}}}\xi,

where b(x,y,ξ):=a(x,ξ)a(y,ξ)¯b(x,y,\xi):=a(x,\xi)\overline{a(y,\xi)} satisfies the estimate

(32) |ξαxβyγb(x,y,ξ)|Cαβγξm1ρ|α|+δ(|β|+|γ|),|\partial^{\alpha}_{\xi}\partial^{\beta}_{x}\partial^{\gamma}_{y}b(x,y,\xi)|\leq C_{\alpha\,\beta\,\gamma}\langle\xi\rangle^{m_{1}-\rho|\alpha|+\delta(|\beta|+|\gamma|)},

with m1=n(ρδ)m_{1}=n(\rho-\delta) and 0ρδ<1,0\leq\rho\leq\delta<1, is bounded on L2(n).L^{2}(\mathbb{R}^{n}). Moreover due to assumption of compact support of a(x,ξ)a(x,\xi) we can also assume that b𝒞c(n×n×n)b\in\mathcal{C}_{c}^{\infty}(\mathbb{R}^{n}\times\mathbb{R}^{n}\times\mathbb{R}^{n}), under the understanding that the norm estimates that we obtain will be independent of the support of bb.
Now we introduce a differential operator

L:=D2{1iξρ(ξφ(x,ξ)ξφ(y,ξ),ξ)},L:=D^{-2}\Big{\{}1-i\langle\xi\rangle^{\rho}\big{(}\langle\nabla_{\xi}\varphi(x,\xi)-\nabla_{\xi}\varphi(y,\xi),\nabla_{\xi}\rangle\big{)}\Big{\}},

with D=(1+ξρ|ξφ(x,ξ)ξφ(y,ξ)|2)12.D=(1+\langle\xi\rangle^{\rho}|\nabla_{\xi}\varphi(x,\xi)-\nabla_{\xi}\varphi(y,\xi)|^{2})^{\frac{1}{2}}. It follows from Lemma 4.2 that

L(eiφ(x,ξ)iφ(y,ξ))=eiφ(x,ξ)iφ(y,ξ)L(e^{i\varphi(x,\xi)-i\varphi(y,\xi)})=e^{i\varphi(x,\xi)-i\varphi(y,\xi)}

and that (Lt)Nb(x,y,ξ)({}^{t}L)^{N}b(x,y,\xi) is a finite sum of terms of the form

(33) Dk{μ=1pξαμξρ}{ν=1q(ξβνφ(x,ξ)ξβνφ(y,ξ))}ξγb(x,y,ξ),D^{-k}\bigg{\{}\prod_{\mu=1}^{p}\partial^{\alpha_{\mu}}_{\xi}\langle\xi\rangle^{\rho}\bigg{\}}\bigg{\{}\prod_{\nu=1}^{q}\big{(}\partial^{\beta_{\nu}}_{\xi}\varphi(x,\xi)-\partial^{\beta_{\nu}}_{\xi}\varphi(y,\xi)\big{)}\bigg{\}}\,\partial^{\gamma}_{\xi}b(x,y,\xi),

with pp, αμ\alpha_{\mu}, βν\beta_{\nu}, qq and γ\gamma quantified by (30). Furthermore since φ\varphi is SND, we can use Proposition 1.11 in [6] to show that

(34) |ξφ(x,ξ)ξφ(y,ξ)|c1|xy||\nabla_{\xi}\varphi(x,\xi)-\nabla_{\xi}\varphi(y,\xi)|\geq c_{1}|x-y|
(35) |zφ(z,ξ1)zφ(z,ξ2)|c2|ξ1ξ2|.|\nabla_{z}\varphi(z,\xi_{1})-\nabla_{z}\varphi(z,\xi_{2})|\geq c_{2}|\xi_{1}-\xi_{2}|.

Using (34), (30), (32) and (33), we have

(36) |xσ(Lt)Nb(x,y,ξ)|CΛ(ξρ(xy))ξm1+δ|σ|,|\partial^{\sigma}_{x}({}^{t}L)^{N}b(x,y,\xi)|\leq C\Lambda(\langle\xi\rangle^{\rho}(x-y))\langle\xi\rangle^{m_{1}+\delta|\sigma|},

where Λ\Lambda is a function with nΛ(ξ)đξ1\int_{\mathbb{R}^{n}}\Lambda(\xi)\,\text{\rm{\mbox{\dj}}}\xi\lesssim 1. Integration by parts using LL, NN times, in (31) one has

(37) Tbf(x)=n×nc(x,y,ξ)eiφ(x,ξ)iφ(y,ξ)f(y)dyđξ,T_{b}f(x)=\iint_{\mathbb{R}^{n}\times\mathbb{R}^{n}}c(x,y,\xi)\,e^{i\varphi(x,\xi)-i\varphi(y,\xi)}\,f(y)\,\mathrm{d}y\,\text{\rm{\mbox{\dj}}}\xi,

with c(x,y,ξ)=(Lt)Nb(x,y,ξ)c(x,y,\xi)=({}^{t}L)^{N}b(x,y,\xi) and

(38) |xσc(x,y,ξ)|CΛ(ξρ(xy))ξm1+δ|σ||\partial^{\sigma}_{x}c(x,y,\xi)|\leq C\Lambda(\langle\xi\rangle^{\rho}(x-y))\langle\xi\rangle^{m_{1}+\delta|\sigma|}

and the same estimate is valid for yσc(x,y,ξ).\partial^{\sigma}_{y}c(x,y,\xi). From this we get the representation

(39) Tb=nA(ξ)đξ,T_{b}=\int_{\mathbb{R}^{n}}A(\xi)\,\text{\rm{\mbox{\dj}}}\xi,

where

A(ξ)f(x):=nc(x,y,ξ)eiφ(x,ξ)iφ(y,ξ)f(y)dy.A(\xi)f(x):=\int_{\mathbb{R}^{n}}c(x,y,\xi)\,e^{i\varphi(x,\xi)-i\varphi(y,\xi)}\,f(y)\,\mathrm{d}y.

Noting that A(ξ)=0A(\xi)=0 for ξ\xi outside some compact set, we observe that condition (i)(i) of Lemma 4.1 follows from Young’s inequality and (38) with σ=0,\sigma=0, and condition (ii)(ii) of Lemma 4.1 follows from the assumption of the compact support of the amplitude. To verify condition (iii)(iii) we confine ourselves to the estimate of A(ξ1)A(ξ2)\|A^{\ast}(\xi_{1})A(\xi_{2})\|, since the one for A(ξ1)A(ξ2)\|A(\xi_{1})A^{\ast}(\xi_{2})\| is similar. To this end, a calculation shows that the kernel of A(ξ1)A(ξ2)A^{\ast}(\xi_{1})A(\xi_{2}) is given by

(40) K(x,y,ξ1,ξ2):=nc¯(z,x,ξ1)c(z,y,ξ2)ei[φ(z,ξ2)φ(z,ξ1)+φ(x,ξ1)φ(y,ξ2)]dz.K(x,y,\xi_{1},\xi_{2}):=\int_{\mathbb{R}^{n}}\overline{c}(z,x,\xi_{1})\,c(z,y,\xi_{2})\,e^{i[\varphi(z,\xi_{2})-\varphi(z,\xi_{1})+\varphi(x,\xi_{1})-\varphi(y,\xi_{2})]}\,\mathrm{d}z.

The estimate (38) yields

(41) |K(x,y,ξ1,ξ2)|ξ1m1ξ2m1nΛ(ξ1ρ(xz))Λ(ξ2ρ(yz))dz.|K(x,y,\xi_{1},\xi_{2})|\lesssim\langle\xi_{1}\rangle^{m_{1}}\,\langle\xi_{2}\rangle^{m_{1}}\int_{\mathbb{R}^{n}}\Lambda(\langle\xi_{1}\rangle^{\rho}(x-z))\,\Lambda(\langle\xi_{2}\rangle^{\rho}(y-z))\,\mathrm{d}z.

Therefore by choosing NN large enough, Young’s inequality and using the fact that nΛ(x)𝑑x1\int_{\mathbb{R}^{n}}\Lambda(x)\,dx\lesssim 1 yield

(42) A(ξ1)A(ξ2)ξ1m1nρξ2m1nρ.\|A^{\ast}(\xi_{1})A(\xi_{2})\|\lesssim\langle\xi_{1}\rangle^{m_{1}-n\rho}\,\langle\xi_{2}\rangle^{m_{1}-n\rho}.

At this point we introduce another first order differential operator M:=G2{1i(zφ(z,ξ2)zφ(z,ξ1),z)}M:=G^{-2}\{1-i(\langle\nabla_{z}\varphi(z,\xi_{2})-\nabla_{z}\varphi(z,\xi_{1}),\nabla_{z}\rangle)\}, with G=(1+|zφ(z,ξ2)zφ(z,ξ1)|2)12.G=(1+|\nabla_{z}\varphi(z,\xi_{2})-\nabla_{z}\varphi(z,\xi_{1})|^{2})^{\frac{1}{2}}. Using the fact that Mei(φ(z,ξ2)φ(z,ξ1))=ei(φ(z,ξ2)φ(z,ξ1)),Me^{i(\varphi(z,\xi_{2})-\varphi(z,\xi_{1}))}=e^{i(\varphi(z,\xi_{2})-\varphi(z,\xi_{1}))}, integration by parts in (40) yields

(43) n(Mt)N{c¯(z,x,ξ1)c(z,y,ξ2)}ei[φ(z,ξ2)φ(z,ξ1)+φ(x,ξ1)φ(y,ξ2)]dz.\int_{\mathbb{R}^{n}}({}^{t}M)^{N^{\prime}}\{\overline{c}(z,x,\xi_{1})\,c(z,y,\xi_{2})\}\,e^{i[\varphi(z,\xi_{2})-\varphi(z,\xi_{1})+\varphi(x,\xi_{1})-\varphi(y,\xi_{2})]}\,\mathrm{d}z.

Using the second part of Lemma 4.2, we find that (Mt)N{c¯(z,x,ξ1)c(z,y,ξ2)}({}^{t}M)^{N^{\prime}}\{\overline{c}(z,x,\xi_{1})\,c(z,y,\xi_{2})\} is a linear combination of terms of the form

(44) Gk{ν=1q(zβνφ(z,ξ2)zβνφ(z,ξ1))}zγ1c¯(z,x,ξ1)zγ2c(z,y,ξ2),G^{-k}\bigg{\{}\prod_{\nu=1}^{q}(\partial^{\beta_{\nu}}_{z}\varphi(z,\xi_{2})-\partial^{\beta_{\nu}}_{z}\varphi(z,\xi_{1}))\bigg{\}}\partial^{\gamma_{1}}_{z}\overline{c}(z,x,\xi_{1})\,\partial^{\gamma_{2}}_{z}c(z,y,\xi_{2}),

where k,k, q,q, βν\beta_{\nu} satisfy the inequalities in (30) and |γ1|+|γ2|N.|\gamma_{1}|+|\gamma_{2}|\leq N^{\prime}.

Now we observe that (35) yields that

Gk(1+|ξ1ξ2|)k.G^{-k}\lesssim(1+|\xi_{1}-\xi_{2}|)^{-k}.

Moreover using (2) we can also deduce that

|zβνφ(z,ξ2)zβνφ(z,ξ1)||ξ1ξ2|1+|ξ1ξ2|.|\partial^{\beta_{\nu}}_{z}\varphi(z,\xi_{2})-\partial^{\beta_{\nu}}_{z}\varphi(z,\xi_{1})|\lesssim|\xi_{1}-\xi_{2}|\leq 1+|\xi_{1}-\xi_{2}|.

Moreover (30) yields that qkNq-k\leq-N^{\prime} and therefore we obtain

Gk|ν=1q(zβνφ(z,ξ2)zβνφ(z,ξ1))|(1+|ξ1ξ2|)qk(1+|ξ1ξ2|)N|ξ1ξ2|N.G^{-k}\Big{|}\prod_{\nu=1}^{q}(\partial^{\beta_{\nu}}_{z}\varphi(z,\xi_{2})-\partial^{\beta_{\nu}}_{z}\varphi(z,\xi_{1}))\Big{|}\lesssim(1+|\xi_{1}-\xi_{2}|)^{q-k}\lesssim(1+|\xi_{1}-\xi_{2}|)^{-N^{\prime}}\lesssim|\xi_{1}-\xi_{2}|^{-N^{\prime}}.

Thus (38) and (44), yield the following estimate for K(x,y,ξ1,ξ2)K(x,y,\xi_{1},\xi_{2})

(45) |K(x,y,ξ1,ξ2)||ξ1ξ2|Nξ1m1ξ2m1(1+|ξ1|+|ξ2|)δN\displaystyle|K(x,y,\xi_{1},\xi_{2})|\lesssim|\xi_{1}-\xi_{2}|^{-N^{\prime}}\langle\xi_{1}\rangle^{m_{1}}\,\langle\xi_{2}\rangle^{m_{1}}(1+|\xi_{1}|+|\xi_{2}|)^{\delta N^{\prime}}
×nΛ(ξ1ρ(xz))Λ(ξ2ρ(yz))dz.\displaystyle\qquad\times\int_{\mathbb{R}^{n}}\Lambda(\langle\xi_{1}\rangle^{\rho}(x-z))\,\Lambda(\langle\xi_{2}\rangle^{\rho}(y-z))\,\mathrm{d}z.

Once again, choosing NN large enough, Young’s inequality yields

(46) A(ξ1)A(ξ2)ξ1m1nρξ2m1nρ(1+|ξ1|+|ξ2|)δN|ξ1ξ2|N.\|A^{\ast}(\xi_{1})A(\xi_{2})\|\lesssim\langle\xi_{1}\rangle^{m_{1}-n\rho}\,\langle\xi_{2}\rangle^{m_{1}-n\rho}\frac{(1+|\xi_{1}|+|\xi_{2}|)^{\delta N^{\prime}}}{|\xi_{1}-\xi_{2}|^{N^{\prime}}}.

Using the fact that for x>0,x>0, inf(1,x)(1+1x)1\inf(1,x)\sim(1+\frac{1}{x})^{-1}, one optimizes the estimates (42) and (46) by

(47) A(ξ1)A(ξ2)\displaystyle\|A^{\ast}(\xi_{1})A(\xi_{2})\| ξ1m1nρξ2m1nρ(1+|ξ1ξ2|N(1+|ξ1|+|ξ2|)δN)1\displaystyle\lesssim\langle\xi_{1}\rangle^{m_{1}-n\rho}\,\langle\xi_{2}\rangle^{m_{1}-n\rho}\bigg{(}1+\frac{|\xi_{1}-\xi_{2}|^{N^{\prime}}}{(1+|\xi_{1}|+|\xi_{2}|)^{\delta N^{\prime}}}\bigg{)}^{-1}
:=h2(ξ1,ξ2).\displaystyle:=h^{2}(\xi_{1},\xi_{2}).

Therefore recalling that m1=n(ρδ),m_{1}=n(\rho-\delta), in applying Lemma 4.1, we need to show that

(48) K(ξ1,ξ2)=(1+|ξ1|)nδ2(1+|ξ2|)nδ2(1+|ξ1ξ2|N(1+|ξ1|+|ξ2|)δN)12K(\xi_{1},\xi_{2})=(1+|\xi_{1}|)^{\frac{-n\delta}{2}}(1+|\xi_{2}|)^{-\frac{n\delta}{2}}\bigg{(}1+\frac{|\xi_{1}-\xi_{2}|^{N^{\prime}}}{(1+|\xi_{1}|+|\xi_{2}|)^{\delta N^{\prime}}}\bigg{)}^{-\frac{1}{2}}

is the kernel of a bounded operator in L2.L^{2}. At this point we use Schur’s lemma, which yields the desired conclusion provided that

supξ1nK(ξ1,ξ2)dξ2,supξ2nK(ξ1,ξ2)dξ1\sup_{\xi_{1}}\int_{\mathbb{R}^{n}}K(\xi_{1},\xi_{2})\,\mathrm{d}\xi_{2},\quad\sup_{\xi_{2}}\int_{\mathbb{R}^{n}}K(\xi_{1},\xi_{2})\,\mathrm{d}\xi_{1}

are both finite. Due to the symmetry of the kernel, we only need to show the finiteness of one of these quantities.
To this end, we fix ξ1\xi_{1} and consider the domains 𝒜={(ξ1,ξ2);|ξ2|2|ξ1|},\mathcal{A}=\{(\xi_{1},\xi_{2});\,|\xi_{2}|\geq 2|\xi_{1}|\}, ={(ξ1,ξ2);|ξ1|2|ξ2|2|ξ1|},\mathcal{B}=\{(\xi_{1},\xi_{2});\,\frac{|\xi_{1}|}{2}\leq|\xi_{2}|\leq 2|\xi_{1}|\}, and 𝒞={(ξ1,ξ2);|ξ2||ξ1|2}.\mathcal{C}=\{(\xi_{1},\xi_{2});\,|\xi_{2}|\leq\frac{|\xi_{1}|}{2}\}. Now we observe that on the set 𝒜,\mathcal{A}, K(ξ1,ξ2)K(\xi_{1},\xi_{2}) is dominated by

(1+|ξ1|)nδ2(1+|ξ2|)nδ2+N2(δ1),(1+|\xi_{1}|)^{-\frac{n\delta}{2}}(1+|\xi_{2}|)^{-\frac{n\delta}{2}+\frac{N^{\prime}}{2}(\delta-1)},

on ,\mathcal{B}, K(ξ1,ξ2)K(\xi_{1},\xi_{2}) is dominated by

(1+|ξ1|)nδ(1+|ξ1ξ2|N(1+|ξ1|)δN)12,(1+|\xi_{1}|)^{-n\delta}\bigg{(}1+\frac{|\xi_{1}-\xi_{2}|^{N^{\prime}}}{(1+|\xi_{1}|)^{\delta N^{\prime}}}\bigg{)}^{-\frac{1}{2}},

and on 𝒞,\mathcal{C}, K(ξ1,ξ2)K(\xi_{1},\xi_{2}) is dominated by

(1+|ξ2|)nδ2(1+|ξ1|)nδ2+N2(δ1).(1+|\xi_{2}|)^{-\frac{n\delta}{2}}(1+|\xi_{1}|)^{-\frac{n\delta}{2}+\frac{N^{\prime}}{2}(\delta-1)}.

Therefore, if IΩ:=ΩK(ξ1,ξ2)đξ2,I_{\Omega}:=\int_{\Omega}K(\xi_{1},\xi_{2})\,\text{\rm{\mbox{\dj}}}\xi_{2}, then choosing N2(δ1)<n,\frac{N^{\prime}}{2}(\delta-1)<-n, which is only possible if δ<1\delta<1, we have that I𝒜<I_{\mathcal{A}}<\infty uniformly in ξ1\xi_{1}. Also,

(49) I𝒞(1+|ξ1|)nnδ2+N2(δ1)C,I_{\mathcal{C}}\leq(1+|\xi_{1}|)^{n-\frac{n\delta}{2}+\frac{N^{\prime}}{2}(\delta-1)}\leq C,

which is again possible by the fact that δ<1\delta<1 and a suitable choice of N.N^{\prime}. In II_{\mathcal{B}} let us make a change of variables to set ξ2ξ1=(1+|ξ1|)δη\xi_{2}-\xi_{1}=(1+|\xi_{1}|)^{\delta}\eta, then

(50) In(1+|η|N)12đη<,I_{\mathcal{B}}\leq\int_{\mathbb{R}^{n}}(1+|\eta|^{N^{\prime}})^{-\frac{1}{2}}\,\text{\rm{\mbox{\dj}}}\eta<\infty,

by taking NN^{\prime} large enough. These estimates yield the desired result and the proof of their theorem is therefore complete. ∎

5. A hpLph^{p}\to L^{p} estimate for oscillatory integral operators

In this section, we show regularity results for oscillatory integral operators. Apart from the distinction caused by the type of the amplitudes, the values of kk also play a decisive role in their regularity theory. When 0<k<10<k<1, it turns out that the type of the amplitude i.e. ρ\rho and δ\delta can be incorporated in the analysis in such a way that the critical order of decay of the amplitude can be improved compared to the case of k1k\geq 1 where the order of the amplitude is

(51) kn|1p12|+min{0,n2(ρδ)}.-kn\Big{|}\frac{1}{p}-\frac{1}{2}\Big{|}+\min\Big{\{}0,\frac{n}{2}(\rho-\delta)\Big{\}}.

To this end, let

(52) ϰ=min(ρ,1k)\varkappa=\min(\rho,1-k)

and set

(53) m(p):=mϰ(p)+ζ.m(p):=m_{\varkappa}(p)+\zeta.

where

mϰ(p):=n(1ϰ)|1p12|m_{\varkappa}(p):=-n(1-\varkappa)\Big{|}\frac{1}{p}-\frac{1}{2}\Big{|}

and

ζ:=min{0,n2(ρδ)}.\zeta:=\min\Big{\{}0,\frac{n}{2}(\rho-\delta)\Big{\}}.

Observe that for k1k\geq 1 one recovers (51) from (53) since then ϰ=1k\varkappa=1-k.

Lemma 5.1.

Let 0ρ10\leq\rho\leq 1 and 0δ<10\leq\delta<1. Assume that φFk\varphi\in\textart{F}^{k} is SND and satisfies the L2L^{2}-condition (3) for 0<k<0<k<\infty. Then if a(x,ξ)Sρ,δm(n)a(x,\xi)\in S^{m}_{\rho,\delta}(\mathbb{R}^{n}) for some mnm\in\mathbb{R}^{n}, let

Kj(x,y)=nei(φ(x,ξ)yξ)σj(x,ξ)đξ.K_{j}(x,y)=\int_{\mathbb{R}^{n}}e^{i(\varphi(x,\xi)-y\cdot\xi)}\,\sigma_{j}(x,\xi)\,\text{\rm{\mbox{\dj}}}\xi.

where σj(x,ξ)=ψj(ξ)a(x,ξ)\sigma_{j}(x,\xi)=\psi_{j}(\xi)\,a(x,\xi). For M0M\geq 0, yny\in\mathbb{R}^{n} and j1j\geq 1 we have

(54) (1+2jϰ|xy|)MyβKj(x,y)Lx2(n)\displaystyle\|(1+2^{j\varkappa}|x-y|)^{M}\partial^{\beta}_{y}K_{j}(x,y)\|_{L^{2}_{x}(\mathbb{R}^{n})} 2j(m~+n/2+|β|),\displaystyle\lesssim 2^{j(\tilde{m}+n/2+|\beta|)},

where m~:=mζ\tilde{m}:=m-\zeta. Moreover, under the extra assumption of φ(x,0)=0\varphi(x,0)=0 then estimate (54) is also valid when Kj(x,y)K_{j}(x,y) is replaced by the kernel of the adjoint Kj(x,y)K_{j}^{*}(x,y).

Proof.

Observe that for any multi-index β\beta and any j0j\geq 0 we have

yβKj(x,y)\displaystyle\partial^{\beta}_{y}K_{j}(x,y) =nψj(ξ)yβei(φ(x,ξ)yξ)a(x,ξ)đξ\displaystyle=\int_{\mathbb{R}^{n}}\psi_{j}(\xi)\,\partial^{\beta}_{y}e^{i(\varphi(x,\xi)-y\cdot\xi)}\,a(x,\xi)\,\text{\rm{\mbox{\dj}}}\xi
=nei(φ(x,ξ)yξ)σjβ(x,ξ)đξ,\displaystyle=\int_{\mathbb{R}^{n}}e^{i(\varphi(x,\xi)-y\cdot\xi)}\,\sigma_{j}^{\beta}(x,\xi)\,\text{\rm{\mbox{\dj}}}\xi,

where

σjβ(x,ξ):=σj(x,ξ)(iξ)β.\sigma_{j}^{\beta}(x,\xi):=\sigma_{j}(x,\xi)\,(-i\xi)^{\beta}.

Therefore, since |ξαxγσjβ(x,ξ)|2j(m+|β|ρ|α|+δ|γ|)|\partial^{\alpha}_{\xi}\partial^{\gamma}_{x}\sigma_{j}^{\beta}(x,\xi)|\lesssim 2^{j(m+|\beta|-\rho|\alpha|+\delta|\gamma|)} we can reduce ourselves to the case when β=0\beta=0.

Now, using (23) and (24) and letting Ψj\Psi_{j} be a Littlewood-Paley function that is supported on a larger annulus in the sense of Definition 2.13, we have

(xy)αKj(x,y)=n(i)|α|ξαei(xy)ξσj(x,ξ)ei(φ(x,ξ)xξ)đξ\displaystyle(x-y)^{\alpha}K_{j}(x,y)=\int_{\mathbb{R}^{n}}(-i)^{|\alpha|}\partial_{\xi}^{\alpha}\,e^{i(x-y)\cdot\xi}\,\sigma_{j}(x,\xi)\,e^{i(\varphi(x,\xi)-x\cdot\xi)}\,\text{\rm{\mbox{\dj}}}\xi
=ni|α|ξα[ei(φ(x,ξ)xξ)σj(x,ξ)]ei(xy)ξΨj(ξ)đξ\displaystyle\quad=\int_{\mathbb{R}^{n}}i^{|\alpha|}\partial_{\xi}^{\alpha}\Big{[}\,e^{i(\varphi(x,\xi)-x\cdot\xi)}\,\sigma_{j}(x,\xi)\Big{]}\,e^{i(x-y)\cdot\xi}\,{\Psi}_{j}(\xi)\,\text{\rm{\mbox{\dj}}}\xi
=α1+α2=αCα1,α2nξα1σj(x,ξ)ξα2ei(φ(x,ξ)xξ)ei(xy)ξΨj(ξ)đξ\displaystyle\quad=\sum_{\alpha_{1}+\alpha_{2}=\alpha}\!\!C_{\alpha_{1},\alpha_{2}}\int_{\mathbb{R}^{n}}\partial_{\xi}^{\alpha_{1}}\sigma_{j}(x,\xi)\,\partial_{\xi}^{\alpha_{2}}e^{i(\varphi(x,\xi)-x\cdot\xi)}\,\,e^{i(x-y)\cdot\xi}{\Psi}_{j}(\xi)\,\text{\rm{\mbox{\dj}}}\xi
=α1+α2=αλ1++λr=α2Cα1,α2,λ1,,λr2j(m~ϰ|α|)nbjα1,α2,λ1,λr(x,ξ)eiφ(x,ξ)eiyξΨj(ξ)đξ\displaystyle\quad=\!\!\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}=\alpha\\ \lambda_{1}+\dots+\lambda_{r}=\alpha_{2}\end{subarray}}\!\!\!\!C_{\alpha_{1},\alpha_{2},\lambda_{1},\dots,\lambda_{r}}2^{j(\tilde{m}-\varkappa|\alpha|)}\!\!\int_{\mathbb{R}^{n}}b_{j}^{\alpha_{1},\alpha_{2},\lambda_{1},\dots\lambda_{r}}(x,\xi)\,e^{i\varphi(x,\xi)}\,e^{-iy\cdot\xi}\,{\Psi}_{j}(\xi)\,\text{\rm{\mbox{\dj}}}\xi
=:α1+α2=αλ1++λr=α2Cα1,α2,λ1,,λr2j(m~ϰ|α|)Sjα1,α2,λ1,λr(τyΨ^j)(x),\displaystyle\quad=:\!\!\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}=\alpha\\ \lambda_{1}+\dots+\lambda_{r}=\alpha_{2}\end{subarray}}\!\!\!\!C_{\alpha_{1},\alpha_{2},\lambda_{1},\dots,\lambda_{r}}2^{j(\tilde{m}-\varkappa|\alpha|)}S_{j}^{\alpha_{1},\alpha_{2},\lambda_{1},\dots\lambda_{r}}\big{(}\tau_{-y}\widehat{\Psi}_{j}\big{)}(x),

where τy\tau_{-y} is a translation by y-y, |λj|1|\lambda_{j}|\geq 1 and

bjα1,α2,λ1,,λr(x,ξ)\displaystyle b_{j}^{\alpha_{1},\alpha_{2},\lambda_{1},\dots,\lambda_{r}}(x,\xi) :=2j(m~ϰ|α|)\displaystyle:=2^{-j(\tilde{m}-\varkappa|\alpha|)}
×ξα1σj(x,ξ)ξλ1(φ(x,ξ)xξ)ξλr(φ(x,ξ)xξ),\displaystyle\qquad\times\partial_{\xi}^{\alpha_{1}}\sigma_{j}(x,\xi)\,\partial_{\xi}^{\lambda_{1}}(\varphi(x,\xi)-x\cdot\xi)\dots\partial_{\xi}^{\lambda_{r}}(\varphi(x,\xi)-x\cdot\xi),

Now we claim that bjα1,α2,λ1,,λrS0,δζ(n),b_{j}^{\alpha_{1},\alpha_{2},\lambda_{1},\dots,\lambda_{r}}\in S^{\zeta}_{0,\delta}(\mathbb{R}^{n}), uniformly in jj. Indeed, since aSρ,δm(n)a\in S^{m}_{\rho,\delta}(\mathbb{R}^{n}) and φFk\varphi\in\textart{F}^{k}, we can write

|bjα1,α2,α3,λ1,,λr(x,ξ)|\displaystyle|b_{j}^{\alpha_{1},\alpha_{2},\alpha_{3},\lambda_{1},\dots,\lambda_{r}}(x,\xi)| 2j(m~ϰ|α|) 2jm~+jζjρ|α1| 2j(k1)|α2|\displaystyle\lesssim 2^{-j(\tilde{m}-\varkappa|\alpha|)}\,2^{j\tilde{m}+j\zeta-j\rho|\alpha_{1}|}\,2^{j(k-1)|\alpha_{2}|}
2jζ2jϰ|α| 2jρ|α1| 2j(k1)|α2|\displaystyle\lesssim 2^{j\zeta}2^{j\varkappa|\alpha|}\,2^{-j\rho|\alpha_{1}|}\,2^{j(k-1)|\alpha_{2}|}
2jζ.\displaystyle\lesssim 2^{j\zeta}.

Moreover, observe also that by the Fk\textart{F}^{k}-condition for k<1k<1,

|xβjξλj1+γj(φ(x,ξ)xξ)|2j(k|λj1+γj|) when|βj|0|\partial_{x}^{\beta_{j}}\partial_{\xi}^{\lambda_{j-1}+\gamma_{j}}(\varphi(x,\xi)-x\cdot\xi)|\lesssim 2^{j(k-|\lambda_{j-1}+\gamma_{j}|)}\text{ when}\,\,\,|\beta_{j}|\geq 0

and for k1k\geq 1 the L2L^{2}-condition (3) yields that

|xβjξλj1+γj(φ(x,ξ)xξ)|1 when|βj|1.|\partial_{x}^{\beta_{j}}\partial_{\xi}^{\lambda_{j-1}+\gamma_{j}}(\varphi(x,\xi)-x\cdot\xi)|\lesssim 1\text{ when}\,\,\,|\beta_{j}|\geq 1.

Thus, using these estimates, we have for all k<0k<0 and all 1jr+11\leq j\leq r+1 that

(55) |xβjξλj1+γj(φ(x,ξ)xξ)|2j(k1)|λj1+γj|, when|βj|1|\partial_{x}^{\beta_{j}}\partial_{\xi}^{\lambda_{j-1}+\gamma_{j}}(\varphi(x,\xi)-x\cdot\xi)|\lesssim 2^{j(k-1)|\lambda_{j-1}+\gamma_{j}|},\text{ when}\,\,\,|\beta_{j}|\geq 1

Now using (55) we can also check that, for any multi-indices γ\gamma and β\beta,

(56) |ξγxβbjα1,α2,λ1,,λr(x,ξ)|\displaystyle|\partial_{\xi}^{\gamma}\partial_{x}^{\beta}\,b_{j}^{\alpha_{1},\alpha_{2},\lambda_{1},\dots,\lambda_{r}}(x,\xi)|
γ1++γr+1=γβ1++βr+1=β2j(m~ϰ|α|)|xβ1ξα1+γ1σj(x,ξ)||xβ2ξλ1+γ2(φ(x,ξ)xξ)|\displaystyle\lesssim\sum_{\begin{subarray}{c}\gamma_{1}+\dots+\gamma_{r+1}=\gamma\\ \beta_{1}+\dots+\beta_{r+1}=\beta\end{subarray}}2^{-j(\tilde{m}-\varkappa|\alpha|)}|\partial_{x}^{\beta_{1}}\partial_{\xi}^{\alpha_{1}+\gamma_{1}}\sigma_{j}(x,\xi)|\,|\partial_{x}^{\beta_{2}}\partial_{\xi}^{\lambda_{1}+\gamma_{2}}(\varphi(x,\xi)-x\cdot\xi)|\dots
×|xβr+1ξλr+γr+1(φ(x,ξ)xξ)|\displaystyle\qquad\times|\partial_{x}^{\beta_{r+1}}\partial_{\xi}^{\lambda_{r}+\gamma_{r+1}}(\varphi(x,\xi)-x\cdot\xi)|
γ1++γr+1=γβ1++βr+1=β2j(m~ϰ|α|)2j(m~+ζ+δ|β1|ρ|α1+γ1|)2j(k1)|α2|\displaystyle\lesssim\sum_{\begin{subarray}{c}\gamma_{1}+\dots+\gamma_{r+1}=\gamma\\ \beta_{1}+\dots+\beta_{r+1}=\beta\end{subarray}}2^{-j(\tilde{m}-\varkappa|\alpha|)}2^{j(\tilde{m}+\zeta+\delta|\beta_{1}|-\rho|\alpha_{1}+\gamma_{1}|)}2^{j(k-1)|\alpha_{2}|}
γ1++γr+1=γβ1++βr+1=β2j(m~ϰ|α|)2j(m~+ζ+δ|β1|)2jϰ|α|\displaystyle\lesssim\sum_{\begin{subarray}{c}\gamma_{1}+\dots+\gamma_{r+1}=\gamma\\ \beta_{1}+\dots+\beta_{r+1}=\beta\end{subarray}}2^{-j(\tilde{m}-\varkappa|\alpha|)}2^{j(\tilde{m}+\zeta+\delta|\beta_{1}|)}2^{-j\varkappa|\alpha|}
2j(ζ+δ|β|),\displaystyle\lesssim 2^{j(\zeta+\delta|\beta|)},

hence bjα1,α2,λ1,,λrS0,δζ(n).b_{j}^{\alpha_{1},\alpha_{2},\lambda_{1},\dots,\lambda_{r}}\in S^{\zeta}_{0,\delta}(\mathbb{R}^{n}).

Therefore, Theorem 4.3 yields that

(xy)αKj(x,y)Lx2(n)\displaystyle\|(x-y)^{\alpha}K_{j}(x,y)\|_{L^{2}_{x}(\mathbb{R}^{n})}
α1+α2=αλ1++λr=α22j(m~ϰ|α|)Sjα1,α2,λ1,λr(τyΨ^j)L2(n)\displaystyle\lesssim\!\!\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}=\alpha\\ \lambda_{1}+\dots+\lambda_{r}=\alpha_{2}\end{subarray}}\!\!2^{j(\tilde{m}-\varkappa|\alpha|)}\big{\|}S_{j}^{\alpha_{1},\alpha_{2},\lambda_{1},\dots\lambda_{r}}\big{(}\tau_{y}\widehat{\Psi}_{j}\big{)}\big{\|}_{L^{2}(\mathbb{R}^{n})}
2j(|α|ϰ+m~)Ψ^jL2(n)\displaystyle\lesssim 2^{j(-|\alpha|\varkappa+\tilde{m})}\big{\|}\widehat{\Psi}_{j}\big{\|}_{L^{2}(\mathbb{R}^{n})}
2j(|α|ϰ+m~+n/2).\displaystyle\lesssim 2^{j(-|\alpha|\varkappa+\tilde{m}+n/2)}.

From this and the discussion at the beginning of the proof one can deduce that

(57) (xy)αyβKj(x,y)Lx2(n)2j(|α|ϰ+|β|+m~+n/2).\|(x-y)^{\alpha}\partial_{y}^{\beta}K_{j}(x,y)\|_{L^{2}_{x}(\mathbb{R}^{n})}\\ \lesssim 2^{j(-|\alpha|\varkappa+|\beta|+\tilde{m}+n/2)}.

Thus by summing over all |α|M|\alpha|\leq M for any integer MM one obtains

(1+2jϰ|xy|)MyβKj(x,y)Lx2(n)2j(m~+n/2+|β|).\displaystyle\|(1+2^{j\varkappa}|x-y|)^{M}\partial_{y}^{\beta}K_{j}(x,y)\|_{L^{2}_{x}(\mathbb{R}^{n})}\lesssim 2^{j(\tilde{m}+n/2+|\beta|)}.

For the kernel of the adjoint, we have that

Kj(x,y)=nei(φ(y,ξ)xξ)σj(y,ξ)¯đξ,K_{j}^{*}(x,y)=\int_{\mathbb{R}^{n}}\,e^{-i(\varphi(y,\xi)-x\cdot\xi)}\,\overline{\sigma_{j}(y,\xi)}\,\text{\rm{\mbox{\dj}}}\xi,

therefore for any multi-index α\alpha we have

(xy)αyβKj(x,y)\displaystyle(x-y)^{\alpha}\partial^{\beta}_{y}K_{j}^{*}(x,y) =(xy)αnyβ(ei(φ(y,ξ)xξ)σj(y,ξ)¯)đξ\displaystyle=(x-y)^{\alpha}\int_{\mathbb{R}^{n}}\,\partial^{\beta}_{y}\Big{(}e^{-i(\varphi(y,\xi)-x\cdot\xi)}\,\overline{\sigma_{j}(y,\xi)}\Big{)}\,\text{\rm{\mbox{\dj}}}\xi
=(xy)αnei(φ(y,ξ)xξ)σjβ,(y,ξ)đξ\displaystyle=(x-y)^{\alpha}\int_{\mathbb{R}^{n}}e^{-i(\varphi(y,\xi)-x\cdot\xi)}\,\sigma_{j}^{\beta,*}(y,\xi)\,\text{\rm{\mbox{\dj}}}\xi
=nei(xy)ξi|α|ξα[ei(φ(y,ξ)yξ)σjβ,(y,ξ)]đξ,\displaystyle=\int_{\mathbb{R}^{n}}e^{i(x-y)\cdot\xi}\,i^{|\alpha|}\partial_{\xi}^{\alpha}\Big{[}e^{-i(\varphi(y,\xi)-y\cdot\xi)}\,\sigma_{j}^{\beta,*}(y,\xi)\Big{]}\,\text{\rm{\mbox{\dj}}}\xi,

where

σjβ,(y,ξ):=β1+β2=βλ1++λr=β2Cβ1,β2,λ1,λryβ1σj(y,ξ)¯yλ1φ(y,ξ)yλrφ(y,ξ),\sigma_{j}^{\beta,*}(y,\xi):=\sum_{\begin{subarray}{c}\beta_{1}+\beta_{2}=\beta\\ \lambda_{1}+\dots+\lambda_{r}=\beta_{2}\end{subarray}}C_{\beta_{1},\beta_{2},\lambda_{1},\dots\lambda_{r}}\,\partial_{y}^{\beta_{1}}\overline{\sigma_{j}(y,\xi)}\,\partial_{y}^{\lambda_{1}}\varphi(y,\xi)\cdots\partial_{y}^{\lambda_{r}}\varphi(y,\xi),

and |λj|1|\lambda_{j}|\geq 1. Now, for |λj+β|2|\lambda_{j}+\beta|\geq 2,

|yλjξβφ(y,ξ)|1,|\partial^{\lambda_{j}}_{y}\partial^{\beta}_{\xi}\varphi(y,\xi)|\lesssim 1,

and using that φ(y,0)=0\varphi(y,0)=0, the mean-value theorem and L2L^{2}-condition (3), we obtain

|yφ(y,ξ)||ξ|.|\nabla_{y}\varphi(y,\xi)|\lesssim|\xi|.

From these estimates, we deduce that for any multi-index γ\gamma one has |yαξγσjβ,(y,ξ)|2j(m+|β|ρ|γ|+δ|α|))|\partial^{\alpha}_{y}\partial^{\gamma}_{\xi}\sigma_{j}^{\beta,*}(y,\xi)|\lesssim 2^{j(m+|\beta|-\rho|\gamma|+\delta|\alpha|))}. Therefore, following the same line of reasoning as in the case of Kj(x,y)K_{j}(x,y) yields the estimate given in (54) for Kj(x,y)K_{j}^{*}(x,y). ∎

Now we are ready to show the main result of this section.

Theorem 5.2.

Let n1n\geq 1, 0<k<0<k<\infty, and 0<p<0<p<\infty. Assume that φFk\varphi\in\textart{F}^{k} is SND, satisfies the LF(μ)(\mu)-condition for some 0<μ10<\mu\leq 1, and the L2L^{2}-condition (3). Assume also that a(x,ξ)Sρ,δm(p)(n),a(x,\xi)\in S^{m(p)}_{\rho,\delta}(\mathbb{R}^{n}), for 0ρ10\leq\rho\leq 1 and 0δ<10\leq\delta<1. Then the OIO TaφT_{a}^{\varphi} is bounded from

  1. (i)(i)

    hp(n)Lp(n),h^{p}(\mathbb{R}^{n})\to L^{p}(\mathbb{R}^{n}), and

  2. (ii)(ii)

    L(n)bmo(n)L^{\infty}(\mathbb{R}^{n})\to\mathrm{bmo}(\mathbb{R}^{n}) provided that one also has |φx(x,0)|L(n)|\nabla\varphi_{x}(x,0)|\in L^{\infty}(\mathbb{R}^{n}).

Proof.

Let χCc(n)\chi\in C_{c}^{\infty}(\mathbb{R}^{n}) be supported in {ξ:|ξ|1}\{\xi:|\xi|\lesssim 1\}, and write

Taφf(x)\displaystyle T_{a}^{\varphi}f(x) :=neiφ(x,ξ)a(x,ξ)f^(ξ)(1χ(ξ))đξ+neiφ(x,ξ)a(x,ξ)f^(ξ)χ(ξ)đξ\displaystyle:=\int_{\mathbb{R}^{n}}e^{i\varphi(x,\xi)}a(x,\xi)\widehat{f}(\xi)(1-\chi(\xi))\,\text{\rm{\mbox{\dj}}}\xi+\int_{\mathbb{R}^{n}}e^{i\varphi(x,\xi)}a(x,\xi)\widehat{f}(\xi)\chi(\xi)\,\text{\rm{\mbox{\dj}}}\xi
=Thighf(x)+Tlowf(x).\displaystyle=T_{\text{high}}f(x)+T_{\text{low}}f(x).

The boundedness of TlowT_{\text{low}} follows from the low frequency result Theorem 2.21. Thus for the remainder of the proof we only consider the high frequency portion of the operator.

Let TjνT_{j}^{\nu} be the operator associated to the kernel in Definition 3.1, so that

Thigh=j=1ν=1𝒩jTjν.T_{\text{high}}=\sum_{j=1}^{\infty}\sum_{\nu=1}^{\mathscr{N}_{j}}T_{j}^{\nu}.

Case when 𝟎<𝒑<𝟏\boldsymbol{0<p<1}:

First we consider the case when 0<p<10<p<1. Let 𝔞\mathfrak{a} be a pp-atom supported in a cube QQ with side length lQl_{Q} and let 2Q2Q be the cube with the same center and twice the side length. Since 0<p<10<p<1 we have

(58) Thigh𝔞Lp(n)p\displaystyle\|T_{\text{high}}\mathfrak{a}\|^{p}_{L^{p}(\mathbb{R}^{n})} Thigh𝔞Lp(2Q)p+j=1Tj𝔞Lp(n2Q)p\displaystyle\lesssim\|T_{\text{high}}\mathfrak{a}\|^{p}_{L^{p}(2Q)}+\sum_{j=1}^{\infty}\|T_{j}\mathfrak{a}\|^{p}_{L^{p}(\mathbb{R}^{n}\setminus 2Q)}

Observe that by Hölder’s inequality and the L2L^{2}-boundedness we have,

(59) Thigh𝔞Lp(2Q)p\displaystyle\|T_{\text{high}}\mathfrak{a}\|^{p}_{L^{p}(2Q)} Thigh𝔞L2(2Q)21L2p/(2p)(2Q)p𝔞L2(2Q)2lQn(2p)/2p\displaystyle\lesssim\|T_{\text{high}}\mathfrak{a}\|^{2}_{L^{2}(2Q)}\|1\|^{p}_{L^{2p/(2-p)}(2Q)}\lesssim\|\mathfrak{a}\|^{2}_{L^{2}(2Q)}l_{Q}^{n(2-p)/2p}
lQn(2p)/2plQn(2p)/2p1\displaystyle\lesssim l_{Q}^{-n(2-p)/2p}l_{Q}^{n(2-p)/2p}\lesssim 1

By Lemma 5.1 we have for 0<k<10<k<1 that

(1+2jϰ|xy|)MKj(x,y)Lx2(n)2j(mϰ(p)+n/2),\displaystyle\|(1+2^{-j\varkappa}|x-y|)^{M}K_{j}(x,y)\|_{L^{2}_{x}(\mathbb{R}^{n})}\lesssim 2^{j(m_{\varkappa}(p)+n/2)},

Observe that for t[0,1]t\in[0,1] and xn2Qx\in\mathbb{R}^{n}\setminus 2Q, one has

(60) |xy¯||xy+t(yy¯)||x-\bar{y}|\lesssim|x-y+t(y-\bar{y})|

Now, setting

g(x)=2jϰ|xy¯|Mg(x)=\left\langle 2^{j\varkappa}|x-\bar{y}|\right\rangle^{-M}

Observe that we have for q1q\geq 1,

(61) gLτ(n)2njϰ/τ.\|g\|_{L^{\tau}(\mathbb{R}^{n})}\lesssim 2^{nj\varkappa/\tau}.

By Hölder’s inequality and Lemma 5.1 and (61) we have for lQ>1l_{Q}>1 and w1w\geq 1 that

(62) Tj𝔞Lp(n2Q)\displaystyle\|T_{j}\mathfrak{a}\|_{L^{p}(\mathbb{R}^{n}\setminus 2Q)} 1g(x)QKj(x,y)𝔞(y)dyL2(n)gLτ(n)\displaystyle\lesssim\Big{\|}\frac{1}{g(x)}\int_{Q}K_{j}(x,y)\mathfrak{a}(y)\,\mathrm{d}y\Big{\|}_{L^{2}(\mathbb{R}^{n})}\|g\|_{L^{\tau}(\mathbb{R}^{n})}
Q1g(x)Kj(x,y)L2(n)|𝔞(y)|dygLτ(n)\displaystyle\lesssim\int_{Q}\Big{\|}\frac{1}{g(x)}K_{j}(x,y)\Big{\|}_{L^{2}(\mathbb{R}^{n})}|\mathfrak{a}(y)|\,\mathrm{d}y\,\|g\|_{L^{\tau}(\mathbb{R}^{n})}
2njϰ/τQ1g(x)Kj(x,y)L2(n)|𝔞(y)|dy\displaystyle\lesssim 2^{-nj\varkappa/\tau}\int_{Q}\Big{\|}\frac{1}{g(x)}K_{j}(x,y)\Big{\|}_{L^{2}(\mathbb{R}^{n})}|\mathfrak{a}(y)|\,\mathrm{d}y
lQnn/p2j(mϰ(p)+n/2nϰ(1p12))lQnn/p2j(nn/p)\displaystyle\lesssim l_{Q}^{n-n/p}2^{j(m_{\varkappa}(p)+n/2-n\varkappa(\frac{1}{p}-\frac{1}{2}))}\lesssim l_{Q}^{n-n/p}2^{j(n-n/p)}

where 1τ=1p12\frac{1}{\tau}=\frac{1}{p}-\frac{1}{2}.

Now, if lQ<1l_{Q}<1, Taylor expansion of KK in the yy-variable around y¯\bar{y}, using the moment conditions of 𝔞\mathfrak{a} yields for all t[0,1]t\in[0,1] and N:=[n(1/p1)]N:=[n(1/p-1)] yield that

Tj𝔞Lp(n2Q)p|α|=N+1n2Q(Q|yαKj(x,t(yy¯))||yy¯|N+1|𝔞(y)|dy)pdx.\displaystyle\|T_{j}\mathfrak{a}\|_{L^{p}(\mathbb{R}^{n}\setminus 2Q)}^{p}\lesssim\sum_{|\alpha|=N+1}\int_{\mathbb{R}^{n}\setminus 2Q}\Big{(}\int_{Q}|{\partial_{y}^{\alpha}K_{j}(x,t(y-\bar{y}))}||y-\bar{y}|^{N+1}|\mathfrak{a}(y)|\,\mathrm{d}y\Big{)}^{p}\,\mathrm{d}x.

Moreover, by using that |yy¯|N+1rN+1|y-\bar{y}|^{N+1}\lesssim r^{N+1} with (60) and Lemma 5.1 we obtain the following estimate by a similar calculation as in (62),

(63) Tj𝔞Lp(n2Q)p2jmϰ(p)p+jnp/2jnpϰ(1/p1/2)2jp(N+1)lQnpn+p(N+1).\displaystyle\|T_{j}\mathfrak{a}\|_{L^{p}(\mathbb{R}^{n}\setminus 2Q)}^{p}\lesssim 2^{jm_{\varkappa}(p)p+jnp/2-jnp\varkappa(1/p-1/2)}2^{jp(N+1)}l_{Q}^{np-n+p(N+1)}.

Now, since lQ<1l_{Q}<1, take the unique integer kQ>1k_{Q}\in\mathbb{Z}_{>1} such that 2kQ<lQ2(kQ1)2^{-k_{Q}}<l_{Q}\leq 2^{-(k_{Q}-1)}. Then using (62) and (63) we have

j=1Tj𝔞Lp(n2Q)p\displaystyle\sum_{j=1}^{\infty}\|T_{j}\mathfrak{a}\|^{p}_{L^{p}(\mathbb{R}^{n}\setminus 2Q)}
j=kQ2jp(mϰ(p)+n/2+nϰ(1p12))lQnpn\displaystyle\qquad\lesssim\sum_{j=k_{Q}}^{\infty}2^{jp(m_{\varkappa}(p)+n/2+n\varkappa(\frac{1}{p}-\frac{1}{2}))}l_{Q}^{np-n}
+j=1kQ12jmϰ(p)p+np/2+jnpϰ(1/p1/2)2jp(N+1)lQnpn+p(N+1)\displaystyle\qquad\qquad+\sum_{j=1}^{k_{Q}-1}2^{jm_{\varkappa}(p)p+np/2+jnp\varkappa(1/p-1/2)}2^{jp(N+1)}l_{Q}^{np-n+p(N+1)}
j=kQ2j(npn)lQnpn+j=1kQ12j(npn+p(N+1))lQnpn+p(N+1)\displaystyle\qquad\lesssim\sum_{j=k_{Q}}^{\infty}2^{j(np-n)}l_{Q}^{np-n}+\sum_{j=1}^{k_{Q}-1}2^{j(np-n+p(N+1))}l_{Q}^{np-n+p(N+1)}
2kQ(npn)lQnpn+2kQ(npn+p(N+1))lQnpn+p(N+1)\displaystyle\qquad\lesssim 2^{k_{Q}(np-n)}l_{Q}^{np-n}+2^{k_{Q}(np-n+p(N+1))}l_{Q}^{np-n+p(N+1)}
lQ(npn)lQnpn+lQ(npn+p(N+1))lQnpn+p(N+1)\displaystyle\qquad\lesssim l_{Q}^{-(np-n)}l_{Q}^{np-n}+l_{Q}^{-(np-n+p(N+1))}l_{Q}^{np-n+p(N+1)}
1.\displaystyle\qquad\sim 1.

Now, if lQ1l_{Q}\geq 1 we do the same calculation as above but with kQ=0k_{Q}=0, and do not consider the case j<kQj<k_{Q}. Thus we conclude that

(64) Thigh𝔞Lp(n)p\displaystyle\|T_{\text{high}}\mathfrak{a}\|^{p}_{L^{p}(\mathbb{R}^{n})} Thigh𝔞Lp(2Q)p+j=1Tj𝔞Lp(n2Q)p1.\displaystyle\lesssim\|T_{\text{high}}\mathfrak{a}\|^{p}_{L^{p}(2Q)}+\sum_{j=1}^{\infty}\|T_{j}\mathfrak{a}\|^{p}_{L^{p}(\mathbb{R}^{n}\setminus 2Q)}\lesssim 1.

Interpolation and arguments for the adjoint operator:

Now, interpolating the result abovewith the L2L^{2}-boundedness result in Theorem 4.3 for 0ρ10\leq\rho\leq 1 and 0δ<10\leq\delta<1 using Riesz-Thorin’s interpolation theorem, one obtains the hpLph^{p}-L^{p}–boundedness of TaφT_{a}^{\varphi} (with the decay m(p)m(p)), in the range 0<p20<p\leq 2.

Next we observe that one may write the phase as φ(x,ξ)=φ(x,ξ)φ(x,0)+φ(x,0)=:ψ(x,ξ)+φ(x,0)\varphi(x,\xi)=\varphi(x,\xi)-\varphi(x,0)+\varphi(x,0)=:\psi(x,\xi)+\varphi(x,0). This would certainly yield that

TaφfLp(n)=TaψfLp(n),\|T^{\varphi}_{a}f\|_{L^{p}(\mathbb{R}^{n})}=\|T^{\psi}_{a}f\|_{L^{p}(\mathbb{R}^{n})},

for 2p<2\leq p<\infty. Moreover observe that ψ(x,0)=0\psi(x,0)=0, and therefore we can without loss of generality assume that φ(x,0)=0\varphi(x,0)=0 in TaφT_{a}^{\varphi}. Now in order to prove the hpLph^{p}-L^{p}–boundedness of TaφT_{a}^{\varphi} for 2p<2\leq p<\infty, using duality and interpolation, it is enough to show that the adjoint operator (Taφ){(T_{a}^{\varphi})}^{*} is bounded from hp(n)h^{p}(\mathbb{R}^{n}) to Lp(n)L^{p}(\mathbb{R}^{n}), for 0<p2.0<p\leq 2. However, this can be shown by the same argument as in the proof of the hpLph^{p}-L^{p}–boundedness of (Taφ){(T_{a}^{\varphi})}, replacing the L2L^{2}-inequality for the kernel with the corresponding inequality for the kernel of the adjoint (given in Lemma 5.1).

Now for the boundedness of TaφT_{a}^{\varphi} from L(n)L^{\infty}(\mathbb{R}^{n}) to bmo(n)\mathrm{bmo}(\mathbb{R}^{n}) one can write Taφ=eiφ(x,0)TaψT_{a}^{\varphi}=e^{i\varphi(x,0)}T_{a}^{{\psi}} with ψ(x,0)=0.{\psi}(x,0)=0. Then given the assumption that φFk\varphi\in\textart{F}^{k}, that φ\varphi satisfies the L2L^{2}-condition of Definition 2.6, and the extra assumption |xφ(x,0)|L(n)|\nabla_{x}\varphi(x,0)|\in L^{\infty}(\mathbb{R}^{n}) on the phase function, one can use (11) to reduce matters to the boundedness of TaψT_{a}^{{\psi}}. But the boundedness of TaψT_{a}^{{\psi}} from L(n)L^{\infty}(\mathbb{R}^{n}) to bmo(n)\mathrm{bmo}(\mathbb{R}^{n}) is a consequence of the boundedness of (Taψ)(T_{a}^{{\psi}})^{\ast} from h1(n)h^{1}(\mathbb{R}^{n}) to L1(n)L^{1}(\mathbb{R}^{n}) which was achieved above, and therefore the proof of the theorem is concluded.

6. Transference to Triebel-Lizorkin spaces for 0<p<10<p<1 and qpq\geq p

This section is devoted to one-half of the process of going from hpLph^{p}\to L^{p} to Triebel-Lizorkin boundedness. In particular, we state and prove the global boundedness of oscillatory integral operators with classical and exotic amplitudes on Triebel-Lizorkin spaces Fp,qsF^{s}_{p,q} for 0<p<20<p<2 and qpq\geq p. To this end, we define a molecular representation of the Triebel-Lizorkin spaces. Similar to the atomic representation of the Hardy spaces, this can be used to prove boundedness results.

Before we show the main results of this section we recall a number of useful lemmas from [14], and as the proofs are very similar or exactly the same we leave out the proofs. In particular, this is with regards to Lemma 6.3 through Lemma 6.7.

Definition 6.1 (Notation).

Let 𝒟\mathcal{D} be the set of all dyadic cubes in n\mathbb{R}^{n} and define the following sets:

  1. (i)(i)

    𝒟j:={Q𝒟:kQ=j}\mathcal{D}_{j}:=\left\{{Q\in\mathcal{D}:k_{Q}=j}\right\}

  2. (ii)(ii)

    𝒟+:={Q𝒟:kQ0}\mathcal{D}_{+}:=\left\{{Q\in\mathcal{D}:k_{Q}\geq 0}\right\}

  3. (iii)(iii)

    𝒟j(Q):={Q~𝒟j:Q~Q}\mathcal{D}_{j}(Q):=\{\tilde{Q}\in\mathcal{D}_{j}:\tilde{Q}\subseteq Q\}.

Observe that for j<kQj<k_{Q}, Dj(Q)=.D_{j}(Q)=\emptyset.

We start with defining a space of sequences that is easier to handle than the Triebel-Lizorkin spaces themselves.

Definition 6.2.

For a sequence of complex numbers b={bQ}Q𝒟l(Q)1b=\left\{{b_{Q}}\right\}_{\begin{subarray}{c}Q\in\mathcal{D}\\ l(Q)\leq 1\end{subarray}} we define

gs,q(b)(x):=(Q~𝒟+(2kQ~(s+n/2)|bQ~|χQ~(x))q)1/q.\displaystyle g^{s,q}(b)(x):=\Big{(}{\sum_{\tilde{Q}\in\mathcal{D}_{+}}\big{(}2^{k_{\tilde{Q}}(s+n/2)}|b_{\tilde{Q}}|\chi_{\tilde{Q}}(x)\big{)}^{q}}\Big{)}^{1/q}.

We say that bfp,qsb\in f^{s}_{p,q} is

bfp,qs:=gs,q(b)Lp(n)<.\displaystyle\|b\|_{f^{s}_{p,q}}:=\|g^{s,q}(b)\|_{L^{p}(\mathbb{R}^{n})}<\infty.

In what follows, take ΨˇQ~(x):=2nkQ~/2ΨˇkQ~(xcQ~)\check{\Psi}^{\tilde{Q}}(x):=2^{-nk_{\tilde{Q}}/2}\check{\Psi}_{k_{\tilde{Q}}}(x-c_{\tilde{Q}}), where ΨkQ~\Psi_{k_{\tilde{Q}}} as given in Definition 2.13 is the usual Littlewood-Paley piece. The following Lemma is a corollary of [8, Theorem II B].

Lemma 6.3.

Suppose 0<p<,0<p<\infty, 0<q0<q\leq\infty, s.s\in\mathbb{R}. For any sequence b={bQ~}Q~𝒟b=\left\{{b_{\tilde{Q}}}\right\}_{\tilde{Q}\in\mathcal{D}} of complex numbers satisfying bfp,qs<,\|b\|_{f^{s}_{p,q}}<\infty, one has

f(x):=Q~𝒟+bQ~ΨˇQ~(x)\displaystyle f(x):=\sum_{\tilde{Q}\in\mathcal{D}_{+}}b_{\tilde{Q}}\check{\Psi}^{\tilde{Q}}(x)

belongs to Fp,qs(n)F^{s}_{p,q}(\mathbb{R}^{n}) and

fFp,qs(n)bfp,qs.\displaystyle\|f\|_{F^{s}_{p,q}(\mathbb{R}^{n})}\lesssim\|b\|_{f^{s}_{p,q}}.
Proof.

This follows by combining [8, Theorem II A i] and [11, Theorem 1]. It is worth noting that this result can also be applied to inhomogeneous Triebel-Lizorkin spaces, as discussed in Chapter 12 of [9]. ∎

We now discuss the converse of Lemma 6.3, namely when a Triebel-Lizorkin function can be expressed in terms of molecules and so-called "\infty-atoms", here denoted bι,Q~b_{\iota,\tilde{Q}}.

Lemma 6.4 ([14]).

Suppose 0<p1,0<p\leq 1, pqp\leq q\leq\infty. Every fFp,q0(n)f\in F^{0}_{p,q}(\mathbb{R}^{n}) has an atomic decomposition

f(x)=ι=1λιQ~𝒟+bι,Q~ΨˇQ~(x),bι,Q~f,q0,\displaystyle f(x)=\sum_{\iota=1}^{\infty}\lambda_{\iota}\sum_{\tilde{Q}\in\mathcal{D}_{+}}b_{\iota,\tilde{Q}}\check{\Psi}^{\tilde{Q}}(x),\qquad b_{\iota,\tilde{Q}}\in f^{0}_{\infty,q},

where {bι,Q~}Q~𝒟+=:bι\left\{{b_{\iota,\tilde{Q}}}\right\}_{\tilde{Q}\in\mathcal{D}_{+}}=:b_{\iota} satisfies

bιf,q02kQ~n/p.\displaystyle\|b_{\iota}\|_{f^{0}_{\infty,q}}\leq 2^{k_{\tilde{Q}}n/p}.

Moreover

fFp,q0(n)inf{λι}(ι=1|λι|p)1/p\displaystyle\|f\|_{F^{0}_{p,q}(\mathbb{R}^{n})}\sim\inf_{\left\{{\lambda_{\iota}}\right\}}\Big{(}{\sum_{\iota=1}^{\infty}|\lambda_{\iota}|^{p}}\Big{)}^{1/p}

Next, we define the analog of the Hardy space atoms, which will be used to prove the boundedness results of our OIO’s.

Lemma 6.5 ([14]).

Let Q𝒟Q\in\mathcal{D}, 0<p10<p\leq 1, 0<q0<q\leq\infty, j1j\geq 1 and b={bQ~}Q~𝒟+f,q0b=\left\{{b_{\tilde{Q}}}\right\}_{\tilde{Q}\in\mathcal{D}_{+}}\in f^{0}_{\infty,q} with bf,q02kQn/p\|b\|_{f^{0}_{\infty,q}}\leq 2^{k_{Q}n/p}. Define

(65) RQ,j(x):=Q~𝒟j(Q)bQ~ΨˇQ~(x).\begin{split}R_{Q,j}(x):=\sum_{\tilde{Q}\in\mathcal{D}_{j}(Q)}b_{\tilde{Q}}\check{\Psi}^{\tilde{Q}}(x).\end{split}

Then RQ,j(x)=0R_{Q,j}(x)=0 for j<kQ~j<k_{\tilde{Q}}. Moreover, for 0<p<0<p<\infty,

RQ,jL2(n)2kQn(1/p1/2).\displaystyle\|R_{Q,j}\|_{L^{2}(\mathbb{R}^{n})}\lesssim 2^{k_{Q}n(1/p-1/2)}.

Now we start with the lifting results for OIO’s to Triebel-Lizorkin spaces. In order to do that, we need to introduce a partition of unity to RQ,jR_{Q,j}. We estimate the pieces separately. The first estimate is given in Lemma 6.6.

Lemma 6.6 ([14]).

Suppose that 0<p<10<p<1 and qpq\geq p. Let a(x,ξ)Sρ,δm(p)(n)a(x,\xi)\in S^{m(p)}_{\rho,\delta}(\mathbb{R}^{n}) be supported outside a neighborhood of the origin and assume that φFk\varphi\in\textart{F}^{k} is SND, satisfies the LF(μ)(\mu)-condition for some 0<μ10<\mu\leq 1, and satisfies the L2L^{2}-condition (3). Then

(66) Taφψj(D)χn2nQRQ,jLp(n)2n(kQj).\begin{split}\|T_{a}^{\varphi}\psi_{j}(D)\chi_{\mathbb{R}^{n}\setminus 2\sqrt{n}Q}R_{Q,j}\|_{L^{p}(\mathbb{R}^{n})}\lesssim 2^{n(k_{Q}-j)}.\end{split}

for jkQj\geq k_{Q}.

Lemma 6.7 ([14]).

Let a(x,ξ)Sρ,δm(n)a(x,\xi)\in S_{\rho,\delta}^{m}(\mathbb{R}^{n}) and that TaφT_{a}^{\varphi} is an oscillatory integral operator that is bounded from Fp,ps(n)F_{p,p}^{s}(\mathbb{R}^{n}) to Fp,ps(n)F_{p,p}^{s}(\mathbb{R}^{n}). Assume also that the phase function φ\varphi satisfies the conditions of Theorem 2.12. Then TσφT_{\sigma}^{\varphi} is bounded from Fp,qs(n)F_{p,q}^{s}(\mathbb{R}^{n}) to Fp,qs(n)F_{p,q}^{s}(\mathbb{R}^{n}) for σ(x,ξ)Sρ,δmε(n)\sigma(x,\xi)\in S_{\rho,\delta}^{m-\varepsilon}(\mathbb{R}^{n}) where ε>0\varepsilon>0 is arbitrary.

Proof.

This lemma follows similarly to [14, Lemma 5.8], with some minor modifications. One replaces the hpLph^{p}\to L^{p} boundedness result in that paper (Proposition 5.1) with Theorem 5.2 in this paper, beyond this abstract modification the proof remains the same, and therefore one only substitutes the hypothesis on TaφT_{a}^{\varphi} in [14, Lemma 5.8], by the hypothesis necessary for the hpLph^{p}\to L^{p} boundedness of TaφT_{a}^{\varphi} in Theorem 5.2. ∎

Now we are prepared to show the main lifting results of this section.

Lemma 6.8.

Let TaφT_{a}^{\varphi} be an OIO\mathrm{OIO} with an amplitude aSρ,δm(p)(n)a\in S^{m(p)}_{\rho,\delta}(\mathbb{R}^{n}) for ρ[0,1],δ[0,1)\rho\in[0,1],\,\delta\in[0,1). Assume φFk\varphi\in\textart{F}^{k} is SND and satisfies the L2L^{2}-condition (3). Moreover Tj:=Taφψj(D)T_{j}:=T_{a}^{\varphi}\psi_{j}(D), where ψj(D)\psi_{j}(D) is a Littlewood-Paley piece as in Definition 2.13. Furthermore, suppose that ff is supported in a cube QQ with fL1(n)2kQn(1/p1).\|f\|_{L^{1}(\mathbb{R}^{n})}\lesssim 2^{k_{Q}n(1/p-1)}.

  1. (i)(i)

    If 0<p<10<p<1, then

    (67) j=max{kQ+1,1}TjfLp(n2Q)p1.\displaystyle\sum_{j=\max\{k_{Q}+1,1\}}^{\infty}\|T_{j}f\|_{L^{p}(\mathbb{R}^{n}\setminus 2Q)}^{p}\lesssim 1.
  2. (ii)(ii)

    If 0<p10<p\leq 1 and ff is an hph^{p}-atom, see Definition 2.20, then

    (68) j=1max{kQ,0}TjfLp(n2Q)p1.\displaystyle\sum_{j=1}^{\max\{k_{Q},0\}}\|T_{j}f\|_{L^{p}(\mathbb{R}^{n}\setminus 2Q)}^{p}\lesssim 1.

Moreover, the same estimates hold true for the adjoint operator (Tj).(T_{j})^{*}.

Proof.

(i)(i) Using (62) we have

j=max{kQ+1,1}TjfLp(n2Q)p\displaystyle\sum_{j=\max\{k_{Q}+1,1\}}^{\infty}\|T_{j}f\|_{L^{p}(\mathbb{R}^{n}\setminus 2Q)}^{p} j=max{kQ+1,1}2j(npn)lQnpn\displaystyle\qquad\lesssim\sum_{j=\max\{k_{Q}+1,1\}}^{\infty}2^{j(np-n)}l_{Q}^{np-n}
2kQ(npn)lQnpn\displaystyle\qquad\lesssim 2^{k_{Q}(np-n)}l_{Q}^{np-n}
lQ(npn)lQnpn\displaystyle\qquad\lesssim l_{Q}^{-(np-n)}l_{Q}^{np-n}
1.\displaystyle\qquad\sim 1.

(ii)(ii) Observe that for lQ>1l_{Q}>1, kQ0k_{Q}\leq 0 and in this case the statement is trivially true. Hence we assume that lQ1l_{Q}\leq 1.

Now, using (63) we have

j=1max{kQ,0}TjfLp(n2Q)p\displaystyle\sum_{j=1}^{\max\{k_{Q},0\}}\|T_{j}f\|^{p}_{L^{p}(\mathbb{R}^{n}\setminus 2Q)}
j=1max{kQ,0}2jmϰ(p)p+np/2+jnpϰ(1/p1/2)2jp(N+1)lQnpn+p(N+1)\displaystyle\qquad\lesssim\sum_{j=1}^{\max\{k_{Q},0\}}2^{jm_{\varkappa}(p)p+np/2+jnp\varkappa(1/p-1/2)}2^{jp(N+1)}l_{Q}^{np-n+p(N+1)}
j=1max{kQ,0}2j(npn+p(N+1))lQnpn+p(N+1)\displaystyle\qquad\lesssim\sum_{j=1}^{\max\{k_{Q},0\}}2^{j(np-n+p(N+1))}l_{Q}^{np-n+p(N+1)}
2kQ(npn+p(N+1))lQnpn+p(N+1)\displaystyle\qquad\lesssim 2^{k_{Q}(np-n+p(N+1))}l_{Q}^{np-n+p(N+1)}
lQ(npn+p(N+1))lQnpn+p(N+1)\displaystyle\qquad\lesssim l_{Q}^{-(np-n+p(N+1))}l_{Q}^{np-n+p(N+1)}
1.\displaystyle\qquad\sim 1.

We observe that the same estimates are also valid for the adjoint TjT_{j}^{\ast} since (62) and (63) are valid for the adjoint. ∎

Proposition 6.9.

Suppose that 0<p<10<p<1 and qpq\geq p. Let a(x,ξ)Sρ,δm(p)(n)a(x,\xi)\in S^{m(p)}_{\rho,\delta}(\mathbb{R}^{n}) be supported outside a neighborhood of the origin and assume that φFk\varphi\in\textart{F}^{k} is SND, and satisfies the L2L^{2}-condition (3). Then the OIO TaφT_{a}^{\varphi} is bounded from Fp,qs(n)F_{p,q}^{s}(\mathbb{R}^{n}) to Fp,qs(n)F_{p,q}^{s}(\mathbb{R}^{n}).

Proof.

Observe that it is enough to show the result for s=0s=0 and by the inclusion Fp,p0Fp,q0F^{0}_{p,p}\xhookrightarrow{}F^{0}_{p,q} it is enough to show that TaφT_{a}^{\varphi} is bounded from Fp,q0(n)F_{p,q}^{0}(\mathbb{R}^{n}) to Fp,p0(n).F_{p,p}^{0}(\mathbb{R}^{n}).

Compose a Littlewood-Paley piece ψj(D)\psi_{j}(D) with TaφT_{a}^{\varphi} and apply Theorem 2.12. This yields

ψj(D)Taφf(x)=Taφψj(xφ(x,D))f(x)+αM2jεRf(x),\displaystyle\psi_{j}(D)T_{a}^{\varphi}f(x)=T_{a}^{\varphi}\psi_{j}(\nabla_{x}\varphi(x,D))f(x)+\sum_{\alpha\leq M}2^{-j\varepsilon}Rf(x),

where ε>0\varepsilon>0 stems from Theorem 2.12 and RR is an operator of better decay, therefore by Lemma 6.7 we have

j=02jε|Rf(x)|Lp(n)RfLp(n)fFp,2(1max(δ,1/2)ε)(n)fFp,q0(n).\displaystyle\Big{\|}\sum_{j=0}^{\infty}2^{-j\varepsilon}|Rf(x)|\Big{\|}_{L^{p}(\mathbb{R}^{n})}\sim\|Rf\|_{L^{p}(\mathbb{R}^{n})}\lesssim\|f\|_{F^{-(1-\max(\delta,1/2)-\varepsilon)}_{p,2}(\mathbb{R}^{n})}\lesssim\|f\|_{F^{0}_{p,q}(\mathbb{R}^{n})}.

So from now on, we will only consider the Fp,q0Lp(q)F^{0}_{p,q}\to L^{p}(\ell^{q})-boundedness of the first term.

Let 𝐭(ξ)=xφ(x,ξ)\mathbf{t}(\xi)=\nabla_{x}\varphi(x,\xi) and η:nn\eta:\mathbb{R}^{n}\to\mathbb{R}^{n} be diffeomorphisms such that η(y)𝐭1(ξ)=yξ\eta(y)\cdot\mathbf{t}^{-1}(\xi)=y\cdot\xi. Then we have

Taφψj(xφ(x,D))f(x)=n×neiφ(x,ξ)iyξψj(𝐭(ξ))a(x,ξ)f(y)dyđξ\displaystyle T_{a}^{\varphi}\psi_{j}(\nabla_{x}\varphi(x,D))f(x)=\iint_{\mathbb{R}^{n}\times\mathbb{R}^{n}}e^{i\varphi(x,\xi)-iy\cdot\xi}\psi_{j}(\mathbf{t}(\xi))a(x,\xi)f(y)\,\mathrm{d}y\,\text{\rm{\mbox{\dj}}}\xi
=1|det(t)|n×neiφ(x,𝐭1(ξ))iyt1(ξ)ψj(ξ)a(x,𝐭1(ξ))f(y)dyđξ\displaystyle=\frac{1}{|\det(\nabla t)|}\iint_{\mathbb{R}^{n}\times\mathbb{R}^{n}}e^{i\varphi(x,\mathbf{t}^{-1}(\xi))-iy\cdot t^{-1}(\xi)}\psi_{j}(\xi)a(x,\mathbf{t}^{-1}(\xi))f(y)\,\mathrm{d}y\,\text{\rm{\mbox{\dj}}}\xi
=|det(η)||det(t)|n×neiφ(x,𝐭1(ξ))iyξψj(ξ)a(x,𝐭1(ξ))f(η(y))dyđξ\displaystyle=\frac{|\det(\nabla\eta)|}{|\det(\nabla t)|}\iint_{\mathbb{R}^{n}\times\mathbb{R}^{n}}e^{i\varphi(x,\mathbf{t}^{-1}(\xi))-iy\cdot\xi}\psi_{j}(\xi)a(x,\mathbf{t}^{-1}(\xi))f(\eta(y))\,\mathrm{d}y\,\text{\rm{\mbox{\dj}}}\xi
=:Tj(fη)(x)\displaystyle=:T_{j}(f\circ\eta)(x)

Observe that by Theorem 2.18 it is enough to consider TjfT_{j}f from now on.

By Lemma 6.4, fFp,q0(n)f\in F^{0}_{p,q}(\mathbb{R}^{n}) has an atomic decomposition

f(x)=ι=1λιQ~𝒟+bι,Q~ΨˇQ~(x),bι,Q~fp,q0f,q0,\displaystyle f(x)=\sum_{\iota=1}^{\infty}\lambda_{\iota}\sum_{\tilde{Q}\in\mathcal{D}_{+}}b_{\iota,\tilde{Q}}\check{\Psi}^{\tilde{Q}}(x),\qquad b_{\iota,\tilde{Q}}\in f^{0}_{p,q}\cap f^{0}_{\infty,q},

such that

fFp,q0(n)inf{λι}(ι=1|λι|p)1/p.\displaystyle\|f\|_{F^{0}_{p,q}(\mathbb{R}^{n})}\approx\inf_{\left\{{\lambda_{\iota}}\right\}}\Big{(}{\sum_{\iota=1}^{\infty}|\lambda_{\iota}|^{p}}\Big{)}^{1/p}.

Since suppψjsuppΨQ~\operatorname{supp}\psi_{j}\subset\operatorname{supp}\Psi^{\tilde{Q}} it is enough to consider Q~𝒟j\tilde{Q}\in\mathcal{D}_{j} and hence

(j=1|Tjf|p)1/pLp(n)\displaystyle\Big{\|}\Big{(}{\sum_{j=1}^{\infty}|T_{j}f|^{p}}\Big{)}^{1/p}\Big{\|}_{L^{p}(\mathbb{R}^{n})} =(j=1|ι=1λιTjQ~𝒟jbι,Q~ΨˇQ~(x)|p)1/pLp(n)\displaystyle=\Big{\|}\Big{(}{\sum_{j=1}^{\infty}\Big{|}\sum_{\iota=1}^{\infty}\lambda_{\iota}T_{j}\sum_{\tilde{Q}\in\mathcal{D}_{j}}b_{\iota,\tilde{Q}}\check{\Psi}^{\tilde{Q}}(x)\Big{|}^{p}}\Big{)}^{1/p}\Big{\|}_{L^{p}(\mathbb{R}^{n})}
(ι=1|λι|pnj=1|TjQ𝒟jbι,Q~ΨˇQ~(x)|pdx)1/p\displaystyle\lesssim\Big{(}{\sum_{\iota=1}^{\infty}|\lambda_{\iota}|^{p}\int_{\mathbb{R}^{n}}\sum_{j=1}^{\infty}\Big{|}T_{j}\sum_{Q\in\mathcal{D}_{j}}b_{\iota,\tilde{Q}}\check{\Psi}^{\tilde{Q}}(x)\Big{|}^{p}\,\mathrm{d}x}\Big{)}^{1/p}
(ι=1|λι|p)1/psupι>0(j=1TjQ~𝒟jbι,Q~ΨˇQ~(x)Lp(n)p)1/p\displaystyle\lesssim\Big{(}{\sum_{\iota=1}^{\infty}|\lambda_{\iota}|^{p}}\Big{)}^{1/p}\sup_{\iota\in\mathbb{Z}_{>0}}\Big{(}{\sum_{j=1}^{\infty}\Big{\|}T_{j}\sum_{\tilde{Q}\in\mathcal{D}_{j}}b_{\iota,\tilde{Q}}\check{\Psi}^{\tilde{Q}}(x)\Big{\|}_{L^{p}(\mathbb{R}^{n})}^{p}}\Big{)}^{1/p}

Therefore it is enough to show that one has an expression of the form

(69) j=1TjRQ,jLp(n)p1\begin{split}\sum_{j=1}^{\infty}\|T_{j}R_{Q,j}\|_{L^{p}(\mathbb{R}^{n})}^{p}\lesssim 1\end{split}

uniformly in Q𝒟Q\in\mathcal{D}, where

RQ,j(x):=Q~𝒟j(Q)bQ~ΨˇQ~(x).\displaystyle R_{Q,j}(x):=\sum_{\tilde{Q}\in\mathcal{D}_{j}(Q)}b_{\tilde{Q}}\check{\Psi}^{\tilde{Q}}(x).

(Recall from Lemma 6.5 that RQ,j(x)=0R_{Q,j}(x)=0 for j<kQ~j<k_{\tilde{Q}}.)

To show (69) we need to show the following three estimates for 0<k<0<k<\infty:

(70) j=max{1,kQ}TjRQ,jLp(2Q)p1,\displaystyle\sum_{j=\max\{1,k_{Q}\}}^{\infty}\|T_{j}R_{Q,j}\|_{L^{p}(2Q)}^{p}\lesssim 1,
(71) j=max{1,kQ}Tjχ2nQRQ,jLp(n2Q)p1,\displaystyle\sum_{j=\max\{1,k_{Q}\}}^{\infty}\|T_{j}\chi_{2\sqrt{n}Q}R_{Q,j}\|_{L^{p}(\mathbb{R}^{n}\setminus 2Q)}^{p}\lesssim 1,
(72) j=max{1,kQ}Tjχn2nQRQ,jLp(n2Q)p1,\displaystyle\sum_{j=\max\{1,k_{Q}\}}^{\infty}\|T_{j}\chi_{\mathbb{R}^{n}\setminus 2\sqrt{n}Q}R_{Q,j}\|_{L^{p}(\mathbb{R}^{n}\setminus 2Q)}^{p}\lesssim 1,

Step 1 – Proof of \tagform@70

To see \tagform@70 consider

TjRQ,jLp(2Q)p\displaystyle\|T_{j}R_{Q,j}\|^{p}_{L^{p}(2Q)} lQn(2p)/2p22jmϰ(p)2jmϰ(p)TjRQ,jL22\displaystyle\leq l_{Q}^{n(2-p)/2p}2^{2jm_{\varkappa}(p)}\|2^{-jm_{\varkappa}(p)}T_{j}R_{Q,j}\|^{2}_{L^{2}}
lQn(2p)/2p22jmϰ(p)RQ,jL22\displaystyle\lesssim l_{Q}^{n(2-p)/2p}2^{2jm_{\varkappa}(p)}\|R_{Q,j}\|^{2}_{L^{2}}
lQn(2p)/2p22kQn(1/p1/2)22jmϰ(p)\displaystyle\lesssim l_{Q}^{n(2-p)/2p}2^{2k_{Q}n(1/p-1/2)}2^{2jm_{\varkappa}(p)}
2kQn(2p)/2p22kQn(1/p1/2)22jmϰ(p)\displaystyle\lesssim 2^{-k_{Q}n(2-p)/2p}2^{2k_{Q}n(1/p-1/2)}2^{2jm_{\varkappa}(p)}
=22jmϰ(p)\displaystyle=2^{2jm_{\varkappa}(p)}

for all 0<p<10<p<1, thus we obtain \tagform@70 by summing in jj, since mϰ(p)<0m_{\varkappa}(p)<0.

Step 3 – Proof of \tagform@71, \tagform@72
\tagform@71 follows directly from using Lemma 6.8 and Lemma 6.5. \tagform@72 follows immediately from Lemma 6.6. This concludes the proof. ∎

7. Transference to Triebel-Lizorkin spaces for p>2p>2

Definition 7.1.

In accordance to Theorem 2.23, let QQ be a cube and set

:=2Q.\mathcal{E}:=2Q.
Lemma 7.2.

Let n1n\geq 1, 0<p<10<p<1. Let 0ρ10\leq\rho\leq 1 and 0δ<10\leq\delta<1. Assume that φFk\varphi\in\textart{F}^{k} is SND and satisfies the L2L^{2}-condition (3) for 0<k<0<k<\infty. Let a(x,ξ)Sρ,δmϰ(p)(n)a(x,\xi)\in S^{m_{\varkappa}(p)}_{\rho,\delta}(\mathbb{R}^{n}), and

Kj(x,y)=nei(φ(x,ξ)yξ)σj(x,ξ)đξ.K_{j}(x,y)=\int_{\mathbb{R}^{n}}e^{i(\varphi(x,\xi)-y\cdot\xi)}\,\sigma_{j}(x,\xi)\,\text{\rm{\mbox{\dj}}}\xi.

be the kernel of TjT_{j}. Then for all ε>0\varepsilon>0 we have for 𝔞\mathfrak{a} a hph^{p}-atom supported in QQ that

  1. (i)(i)

    If 2jlQ12^{-j}\leq l_{Q}\leq 1, then

    (73) n|TjΨj(D)𝔞|dx2jmϰ(p)12p1lQε+lQ12p1(nn/p)212p1j(nn/p),\int_{\mathbb{R}^{n}}|T_{j}^{*}\Psi_{j}(D)\mathfrak{a}|\,\mathrm{d}x\lesssim 2^{jm_{\varkappa}(p)\frac{1}{\frac{2}{p}-1}}l_{Q}^{-\varepsilon}+l_{Q}^{\frac{1}{\frac{2}{p}-1}(n-n/p)}2^{\frac{1}{\frac{2}{p}-1}j(n-n/p)},
  2. (ii)(ii)

    If lQ1l_{Q}\geq 1, then

    (74) n|TjΨj(D)𝔞|dx(2jmϰ(p)+lQnn/p2j(nn/p))12p1.\int_{\mathbb{R}^{n}}|T_{j}^{*}\Psi_{j}(D)\mathfrak{a}|\,\mathrm{d}x\lesssim(2^{jm_{\varkappa}(p)}+l_{Q}^{n-n/p}2^{j(n-n/p)})^{\frac{1}{\frac{2}{p}-1}}.
Proof.

We begin by proving (73), to this end let ww be a real number such that 1/w=1/p1/21/w=1/p-1/2, and split n=n\mathbb{R}^{n}=\mathcal{E}\cup\mathbb{R}^{n}\setminus\mathcal{E}. Then

(|TjΨj(D)𝔞|pdx)1/p\displaystyle\Big{(}\int_{\mathcal{E}}|T_{j}^{*}\Psi_{j}(D)\mathfrak{a}|^{p}\,\mathrm{d}x\Big{)}^{1/p} 1Lw()(n|TjΨj(D)𝔞|2dx)1/2\displaystyle\lesssim\|1\|_{L^{w}(\mathcal{E})}\Big{(}\int_{\mathbb{R}^{n}}|T_{j}^{*}\Psi_{j}(D)\mathfrak{a}|^{2}\,\mathrm{d}x\Big{)}^{1/2}
|Q|1/p1/22jmϰ(p)(n|𝔞(x)|2dx)1/22jmϰ(p).\displaystyle\lesssim|Q|^{1/p-1/2}2^{jm_{\varkappa}(p)}\Big{(}\int_{\mathbb{R}^{n}}|\mathfrak{a}(x)|^{2}\,\mathrm{d}x\Big{)}^{1/2}\lesssim 2^{jm_{\varkappa}(p)}.

Now, recall that

g(x)=2jϰ|xy¯|Mg(x)=\left\langle 2^{j\varkappa}|x-\bar{y}|\right\rangle^{-M}

and that we have for w1w\geq 1,

(75) gLw(n)2njϰ/w.\|g\|_{L^{w}(\mathbb{R}^{n})}\lesssim 2^{nj\varkappa/w}.

Next Hölder’s and Minkowski’s inequalities and Lemma 5.1 yield that

(n|TjΨj(D)𝔞|pdy)1/p1g(x)BKj(y,x)¯Ψj(D)𝔞(y)dyL2(n)gLw(n)\displaystyle\Big{(}\int_{\mathbb{R}^{n}\setminus\mathcal{E}}|T_{j}^{*}\Psi_{j}(D)\mathfrak{a}|^{p}\,\mathrm{d}y\Big{)}^{1/p}\lesssim\Big{\|}\frac{1}{g(x)}\int_{B}\overline{K_{j}(y,x)}\Psi_{j}(D)\mathfrak{a}(y)\,\mathrm{d}y\Big{\|}_{L^{2}(\mathbb{R}^{n})}\|g\|_{L^{w}(\mathbb{R}^{n})}
B1g(x)Kj(y,x)¯L2(n)|Ψj(D)𝔞(y)|dygLw(n)\displaystyle\lesssim\int_{B}\Big{\|}\frac{1}{g(x)}\overline{K_{j}(y,x)}\Big{\|}_{L^{2}(\mathbb{R}^{n})}|\Psi_{j}(D)\mathfrak{a}(y)|\,\mathrm{d}y\,\|g\|_{L^{w}(\mathbb{R}^{n})}
2njϰ/wB1g(x)Kj(y,x)¯L2(n)|Ψj(D)𝔞(y)|dy\displaystyle\lesssim 2^{-nj\varkappa/w}\int_{B}\Big{\|}\frac{1}{g(x)}\overline{K_{j}(y,x)}\Big{\|}_{L^{2}(\mathbb{R}^{n})}|\Psi_{j}(D)\mathfrak{a}(y)|\,\mathrm{d}y
lQnn/p2j(mϰ(p)+n/2nϰ(1/p1/2))\displaystyle\lesssim l_{Q}^{n-n/p}2^{j(m_{\varkappa}(p)+n/2-n\varkappa(1/p-1/2))}
=lQnn/p2j(nn/p)\displaystyle=l_{Q}^{n-n/p}2^{j(n-n/p)}

where we have also use that 𝔞\mathfrak{a} is an hph^{p}-atom and that Ψj(D)\Psi_{j}(D) is L2L^{2}-bounded uniformly in jj.

Interpolation step:

Using the L2L^{2} result (Theorem 4.1) we obtain by Riesz-Thorin interpolation and 1pt=t2+1tp\frac{1}{p_{t}}=\frac{t}{2}+\frac{1-t}{p} with 0<p<10<p<1 and 0<k<10<k<1 that

(n|TjΨj(D)𝔞|ptdx)1/pt\displaystyle\Big{(}\int_{\mathbb{R}^{n}}|T_{j}^{*}\Psi_{j}(D)\mathfrak{a}|^{p_{t}}\,\mathrm{d}x\Big{)}^{1/p_{t}} (2jm+lQnn/p2j(nn/p))(1t)\displaystyle\lesssim(2^{jm}+l_{Q}^{n-n/p}2^{j(n-n/p)})^{(1-t)}
2jm(1t)+lQ(1t)(nn/p)2(1t)j(nn/p).\displaystyle\lesssim 2^{jm(1-t)}+l_{Q}^{(1-t)(n-n/p)}2^{(1-t)j(n-n/p)}.

Taking pt=1p_{t}=1 and t=11p121pt=\frac{1-\frac{1}{p}}{\frac{1}{2}-\frac{1}{p}} we have 1t=12p11-t=\frac{1}{\frac{2}{p}-1} and therefore

n|TjΨj(D)𝔞|dx\displaystyle\int_{\mathbb{R}^{n}}|T_{j}^{*}\Psi_{j}(D)\mathfrak{a}|\,\mathrm{d}x (2jm+lQnn/p2j(nn/p))12p1.\displaystyle\lesssim(2^{jm}+l_{Q}^{n-n/p}2^{j(n-n/p)})^{\frac{1}{\frac{2}{p}-1}}.

Thus for lQ>1l_{Q}>1 we have

n|TjΨj(D)𝔞|dx\displaystyle\int_{\mathbb{R}^{n}}|T_{j}^{*}\Psi_{j}(D)\mathfrak{a}|\,\mathrm{d}x 2jm12p1+212p1j(nn/p)\displaystyle\lesssim 2^{jm\frac{1}{\frac{2}{p}-1}}+2^{\frac{1}{\frac{2}{p}-1}j(n-n/p)}

and for lQ<1l_{Q}<1 and ε>0\varepsilon>0 we have

n|TjΨj(D)𝔞|dx\displaystyle\int_{\mathbb{R}^{n}}|T_{j}^{*}\Psi_{j}(D)\mathfrak{a}|\,\mathrm{d}x 2jm12p1lQε+lQ12p1(nn/p)212p1j(nn/p).\displaystyle\lesssim 2^{jm\frac{1}{\frac{2}{p}-1}}l_{Q}^{-\varepsilon}+l_{Q}^{\frac{1}{\frac{2}{p}-1}(n-n/p)}2^{\frac{1}{\frac{2}{p}-1}j(n-n/p)}.

Theorem 7.3.

Let n1n\geq 1. Assume that φFk\varphi\in\textart{F}^{k} is SND and satisfies the L2L^{2}-condition (3) for 0<k<0<k<\infty. Then if a(x,ξ)Sρ,δmϰ(p)(n)a(x,\xi)\in S^{m_{\varkappa}(p)}_{\rho,\delta}(\mathbb{R}^{n}), and TjT_{j} are as in Lemma 7.2. Finally, we let 2<p2<p, 0<q0<q\leq\infty, and b>0b>0. Assume that the operators TjT_{j} satisfy

(76) supj>02jb/pTjLpLp1,\displaystyle\sup_{j>0}2^{jb/p}\|T_{j}\|_{L^{p}\to L^{p}}\lesssim 1,
(77) supj>02jb/2TjL2L21.\displaystyle\sup_{j>0}2^{jb/2}\|T_{j}\|_{L^{2}\to L^{2}}\lesssim 1.

Then, the following inequality holds:

(j=02jbq/p|Ψj(D)Tjfj|q)1/qLp(n)(j=0fjLp(n)p)1/p.\displaystyle\Big{\|}\Big{(}{\sum_{j=0}^{\infty}2^{jbq/p}|\Psi_{j}(D)T_{j}f_{j}|^{q}}\Big{)}^{1/q}\Big{\|}_{L^{p}(\mathbb{R}^{n})}\lesssim\Big{(}{\sum_{j=0}^{\infty}\|f_{j}\|_{L^{p}(\mathbb{R}^{n})}^{p}}\Big{)}^{1/p}.

where Ψ𝒮(n)\Psi\in\mathscr{S}(\mathbb{R}^{n}), Ψj(D):=Ψ(2jD)\Psi_{j}(D):=\Psi(2^{-j}D) and fjf_{j} is a sequence of functions.

Proof.

We consider a measurable set \mathcal{E} (as defined in Definition 7.1). We have the following inequality:

||lQn,|\mathcal{E}|\lesssim l_{Q}^{n},

which means that all the hypotheses of Theorem (2.23) are satisfied, with q=2q=2. Note that the condition b<nb<n in Theorem (2.23) is not necessary here, because it stems from Lemma 2.2 in [18], which we substitute with our Lemma (7.2) (and duality). We also do not need the assumption (20) about the kernel, because we show Lemma 7.2 using a different argument. Therefore, we can obtain the result by applying the same method as in [18] to the adjoint TjT_{j}^{*}, and then interpolating with L2(n)L^{2}(\mathbb{R}^{n}). ∎

8. Triebel-Lizorkin estimates

In this section we use the lifting results from section 6 and 7 to lift Theorem 5.2 to Triebel-Lizorkin boundedness.

ppqq22112211\infty
Figure 2. Triebel-Lizorkin boundedness for OIO:s with their respective critical decay. The blue horizontal lines illustrate Theorem 7.3. The blue vertical lines illustrate the boundedness results obtained by applying a duality argument for the case p>2p>2. The red vertical lines illustrate Proposition 6.9 together with interpolation with the vertical blue area.
Theorem 8.1.

Let 0ρ10\leq\rho\leq 1, 0δ<10\leq\delta<1, k>0k>0, 0<μ10<\mu\leq 1, and 0<q0<q\leq\infty. Assume furthermore that φFk\varphi\in\textart{F}^{k} is SND, and satisfies the LF(μ)(\mu) condition (5) and the L2L^{2}-condition (3). For aSρ,δm(p)(n)a\in S^{m(p)}_{\rho,\delta}(\mathbb{R}^{n}), and let TaφT_{a}^{\varphi} be the associated OIO. If ss\in\mathbb{R} and either one of the following cases holds

(i)\displaystyle(i)\quad 2<p< when 0<qp,\displaystyle 2<p<\infty\text{ when }0<q\leq p,
(ii)\displaystyle(ii)\quad nn+μ<p<2 when pq,\displaystyle\frac{n}{n+\mu}<p<2\text{ when }p\leq q,
(iii)\displaystyle(iii)\quad p=q=2,\displaystyle p=q=2,

then the OIO TaφT_{a}^{\varphi} is bounded from Fp,qs(n)F_{p,q}^{s}(\mathbb{R}^{n}) to Fp,qs(n)F_{p,q}^{s}(\mathbb{R}^{n})

Proof.

We separate the operator into a low and a high-frequency part. The result for the low-frequency part follows from Theorem 2.21, so we only consider the high-frequency part from now on.

Observe also that the contents of (iii)(iii) is contained in Theorem 4.3. So from now on we only need to show (i)(i) and (ii)(ii).

We split the proof into different ranges of pp and qq, the two parts of the proof correspond to the blue and the red regions in Figure 2, respectively.

Part 1 – Proof when 𝒑>𝟐\boldsymbol{p>2} and 𝒑𝒒>𝟎\boldsymbol{p\geq q>0}
We use Theorem 2.12 to write

ψj(D)Taφf(x)=Taφψj(xφ(x,D))f(x)+Rf(x),\displaystyle\psi_{j}(D)T_{a}^{\varphi}f(x)=T_{a}^{\varphi}\psi_{j}(\nabla_{x}\varphi(x,D))f(x)+Rf(x),

The operator RR is bounded by Lemma 6.7. So from now on we will only consider the first term. Denote Tj:=Taφψj(xφ(x,D)).T_{j}:=T_{a}^{\varphi}\psi_{j}(\nabla_{x}\varphi(x,D)).

Hence we can use Theorem 7.3 with b:=n(1ϰ)b:=n(1-\varkappa) and Sj:=Tj(1Δ)b/2pS_{j}:=T_{j}(1-\Delta)^{-b/2p} to prove the desired result. Observe that the hpLph^{p}\to L^{p} boundedness (Proposition 5.2) and the L2L^{2} boundedness (Theorem 4.3)of TaφT_{a}^{\varphi} yield (76) and (77) respectively.

Theorem 7.3 immediately yields that

(j=02jbq/p|Ψj(D)Tjfj|q)1/qLp(n)=(j=02jbq/p|Ψj(D)Sjfj|q)1/qLp(n)\displaystyle\Big{\|}\Big{(}{\sum_{j=0}^{\infty}2^{jbq/p}|\Psi_{j}(D)T_{j}f_{j}|^{q}}\Big{)}^{1/q}\Big{\|}_{L^{p}(\mathbb{R}^{n})}=\Big{\|}\Big{(}{\sum_{j=0}^{\infty}2^{jbq/p}|\Psi_{j}(D)S_{j}f_{j}|^{q}}\Big{)}^{1/q}\Big{\|}_{L^{p}(\mathbb{R}^{n})}
(j=0fjLp(n)p)1/p.\displaystyle\lesssim\Big{(}{\sum_{j=0}^{\infty}\|f_{j}\|_{L^{p}(\mathbb{R}^{n})}^{p}}\Big{)}^{1/p}.

Thus Sj:Bp,p0Fp,qb/pS_{j}:B^{0}_{p,p}\to F^{b/p}_{p,q} which immediately implies that Tj:Bp,pb/pFp,qb/pT_{j}:B^{b/p}_{p,p}\to F^{b/p}_{p,q}. Now the assertion follows from the facts that Fp,q0Fp,p0=Bp,p0F^{0}_{p,q}\xhookrightarrow{}F^{0}_{p,p}=B^{0}_{p,p} and the calculus using Bessel potentials and Theorem 2.12.

Part 2 – Proof when 𝟏<𝒑<𝟐\boldsymbol{1<p<2} and 𝒑𝒒\boldsymbol{p\leq q}
Using that the operator is self-adjoint we can also obtain Theorem (7.3) for the adjoint, then apply the arguments from part 1.

Part 3 – Proof when 𝟎<𝒑<𝟏,𝒑𝒒\boldsymbol{0<p<1,p\leq q\leq\infty}
By Proposition 6.9 one obtains the result for 0<p<10<p<1 and qpq\geq p.

Part 3.1 – Proof when 𝟎<𝒑𝒒<𝟏\boldsymbol{0<p\leq q<1}
In this case using Riesz-Thorin interpolation with Fp,ps=Bp,ps,F^{s}_{p,p}=B^{s}_{p,p}, yields the result.

Part 3.2 – Proof when 𝟎<𝒑<𝟏𝒒\boldsymbol{0<p<1\leq q\leq\infty}
Here Riesz-Thorin interpolation with Part 2 above, yields the result.

Now notice that (1Δ)s2Taφ(1Δ)s2(1-\Delta)^{\frac{s}{2}}T^{\varphi}_{a}(1-\Delta)^{-\frac{s}{2}} is a similar operator associated to an amplitude in Smc(p)(n)S^{m_{c}(p)}(\mathbb{R}^{n}) and phase φ\varphi, and hence bounded from Fp,q0(n)F^{0}_{p,q}(\mathbb{R}^{n}) to itself. Therefore using the fact that the operator (1Δ)s2(1-\Delta)^{\frac{s}{2}} is an isomorphism from Fp,qs(n)F^{s}_{p,q}(\mathbb{R}^{n}) to Fp,q0(n)F^{0}_{p,q}(\mathbb{R}^{n}) for 0<p0<p\leq\infty, we obtain the desired result. ∎

8.1. Triebel-Lizorkin estimates related to forbidden amplitudes

It is well known that the oscillatory integral operators with amplitudes in S1,1m(n)S^{m}_{1,1}(\mathbb{R}^{n}) fail to be L2L^{2}-bounded. However, one may show that these operators are Sobolev-bounded for Hs(n)H^{s}(\mathbb{R}^{n}) for s>0s>0 (see [2]), and the pseudodifferential case goes back to E. Stein and independently by Y. Meyer. In this section, we state and prove two results about the boundedness of oscillatory integral operators in Triebel-Lizorkin spaces with amplitudes in S1,1m(n)S^{m}_{1,1}(\mathbb{R}^{n}).

Now we turn to the boundedness of OIOs with forbidden amplitudes in the class S1,1m(n)S^{m}_{1,1}(\mathbb{R}^{n}) and ask whether they are bounded on Triebel-Lizorkin spaces.

Theorem 8.2.

Let n1n\geq 1, k>0k>0, s>0s>0, 0<μ10<\mu\leq 1. Assume that a(x,ξ)S1,1kn|1/p1/2|(n)a(x,\xi)\in S^{-kn|1/p-1/2|}_{1,1}(\mathbb{R}^{n}), φFk\varphi\in\textart{F}^{k} is SND, satisfies the LF(μ)(\mu)-condition and the conditions in (9), and the L2L^{2}-condition (3). If s>n(1min{1,p,q}1)s>n\big{(}\frac{1}{\min\{1,p,q\}}-1\big{)}, and either one of the following cases hold

  1. (i)(i)

    2<p<2<p<\infty when 0<qp,0<q\leq p,

  2. (ii)(ii)

    nn+μ<p<2\frac{n}{n+\mu}<p<2 when pq,p\leq q,

  3. (iii)(iii)

    p=q=2,p=q=2,

or if s>nps>\frac{n}{p} with q=q=\infty, then the OIO TaφT_{a}^{\varphi} is bounded from Fp,qs(n)Fp,qs(n).F^{s}_{p,q}(\mathbb{R}^{n})\to F^{s}_{p,q}(\mathbb{R}^{n}).

Proof.

The proof is similar to the proof of Theorem 5.13 in [14], and therefore we shall only highlight the differences here. The only differences appear in connection to (100). At this point in the proof one instead uses Theorem 8.1 instead of Theorem 5.15. The rest of the proof remains the same. ∎

9. Besov-Lipschitz estimates

In this section we include both the Besov-Lipschitz boundedness results of OIO’s with amplitudes in Sρ,δmS^{m}_{\rho,\delta} for all 0ρ10\leq\rho\leq 1 and 0δ10\leq\delta\leq 1. We begin with the classical amplitudes were δ<1\delta<1.

Theorem 9.1.

Let 0ρ10\leq\rho\leq 1 and 0δ<10\leq\delta<1, k>0k>0, 0<μ1,0<\mu\leq 1, 0<q0<q\leq\infty. Assume that φFk\varphi\in\textart{F}^{k} is SND, satisfies the LF(μ)(\mu)-condition and the conditions in (9), and the L2L^{2}-condition (3), and aSρ,δm(p)(n)a\in S^{m(p)}_{\rho,\delta}(\mathbb{R}^{n}), and let TaφT_{a}^{\varphi} be the associated OIO. Then TaφT_{a}^{\varphi} is a bounded operator from Bp,qs(n)B^{s}_{p,q}(\mathbb{R}^{n}) to Bp,qs(n)B^{s}_{p,q}(\mathbb{R}^{n}) for all ss\in\mathbb{R} for nn+μ<p<\frac{n}{n+\mu}<p<\infty.

Proof.

Let χCc(n)\chi\in C_{c}^{\infty}(\mathbb{R}^{n}) be supported in {ξ:|ξ|1}\{\xi:|\xi|\lesssim 1\}, and write

Taφf(x)\displaystyle T_{a}^{\varphi}f(x) :=neiφ(x,ξ)a(x,ξ)f^(ξ)(1χ(ξ))đξ+neiφ(x,ξ)a(x,ξ)f^(ξ)χ(ξ)đξ\displaystyle:=\int_{\mathbb{R}^{n}}e^{i\varphi(x,\xi)}a(x,\xi)\widehat{f}(\xi)(1-\chi(\xi))\,\text{\rm{\mbox{\dj}}}\xi+\int_{\mathbb{R}^{n}}e^{i\varphi(x,\xi)}a(x,\xi)\widehat{f}(\xi)\chi(\xi)\,\text{\rm{\mbox{\dj}}}\xi
=Thighf(x)+Tlowf(x).\displaystyle=T_{\text{high}}f(x)+T_{\text{low}}f(x).

The boundedness of TlowT_{\text{low}} follows from the low frequency result Theorem 2.21. The boundedness of ThighT_{\text{high}} follows immediately by the Besov-Lipschitz lift Theorem 2.22 and Proposition 5.2. ∎

In this next result we obtain an estimate relating to OIO’s with amplitudes in LS0m(n)L^{\infty}S^{m}_{0}(\mathbb{R}^{n}) (see Definition 2.2), observe that this class of amplitudes contains the forbidden Hörmander class amplitudes S0,1m(n)S^{m}_{0,1}(\mathbb{R}^{n}) and S1,1m(n)S^{m}_{1,1}(\mathbb{R}^{n}).

Theorem 9.2.

Let n1n\geq 1, ss\in\mathbb{R}, 1<p<1<p<\infty, 0<q0<q\leq\infty. Assume furthermore that φ(x,ξ)Fk\varphi(x,\xi)\in\textart{F}^{k} is an SND phase of order k>0k>0, aLS0m(n)a\in L^{\infty}S^{m}_{0}(\mathbb{R}^{n}) with

m<n(kp1+(11p)w(k,ρ)),m<n\Big{(}kp-1+\Big{(}1-\frac{1}{p}\Big{)}w(k,\rho)\Big{)},

and the OIO TaφT_{a}^{\varphi} is bounded from Bp,qs(n)Bp,qs(n).B^{s}_{p,q}(\mathbb{R}^{n})\to B^{s}_{p,q}(\mathbb{R}^{n}). ((Where w(k,ρ)w(k,\rho) is as defined in Lemma 3.2)

Proof.

We separate the operator into a low and a high-frequency part. The result for the low-frequency part follows from Theorem 2.21, so we only consider the high-frequency part from now on.

Following the proof of Theorem 5.2 we obtain the kernels KjνK_{j}^{\nu} with the associated kernel estimate (22).

Let TjνT_{j}^{\nu} be the operators corresponding to the kernels KjνK_{j}^{\nu}. One observes that

|Tjνfj(x)|r\displaystyle|T_{j}^{\nu}f_{j}(x)|^{r} =|nKjν(x,y)fj(y)dy|r\displaystyle=\Big{|}\int_{\mathbb{R}^{n}}K_{j}^{\nu}(x,y)f_{j}(y)\,\mathrm{d}y\Big{|}^{r}
=|nKjν(x,y)σ(ξφ(x,ξjν)y)1σ(ξφ(x,ξjν)y)fj(y)dy|r,\displaystyle=\Big{|}\int_{\mathbb{R}^{n}}K_{j}^{\nu}(x,y)\sigma(\nabla_{\xi}\varphi(x,\xi_{j}^{\nu})-y)\frac{1}{\sigma(\nabla_{\xi}\varphi(x,\xi_{j}^{\nu})-y)}f_{j}(y)\,\mathrm{d}y\Big{|}^{r},

with weight functions σ(y)\sigma(y) which will be chosen momentarily. Therefore, Hölder’s inequality with 1r+1r=1\frac{1}{r}+\frac{1}{r^{\prime}}=1, r,r>1r,r^{\prime}>1 yields

(78) |Tjνfj(x)|r(n|Kjν(x,y)|r|σ(ξφ(x,ξjν)y)|rdy)rr\displaystyle|T_{j}^{\nu}f_{j}(x)|^{r}\leq\Big{(}\int_{\mathbb{R}^{n}}|K_{j}^{\nu}(x,y)|^{r^{\prime}}|\sigma(\nabla_{\xi}\varphi(x,\xi_{j}^{\nu})-y)|^{r^{\prime}}\,\mathrm{d}y\Big{)}^{\frac{r}{r^{\prime}}}
×(n|fj(y)|r|σ(ξφ(x,ξjν)y)|rdy).\displaystyle\qquad\times\Big{(}\int_{\mathbb{R}^{n}}\frac{|f_{j}(y)|^{r}}{|\sigma(\nabla_{\xi}\varphi(x,\xi_{j}^{\nu})-y)|^{r}}\,\mathrm{d}y\Big{)}.

where σ\sigma is defined by

σ(y)={1,|y|1;|y|λ,|y|>1.\sigma(y)=\begin{cases}1,&|y|\leq 1;\\ |y|^{\lambda},&|y|>1.\end{cases}

Observe further that

n|Kjν(x,y)|r|σ(ξφ(x,ξjν)y)|rdy\displaystyle\int_{\mathbb{R}^{n}}|K_{j}^{\nu}(x,y)|^{r^{\prime}}|\sigma(\nabla_{\xi}\varphi(x,\xi_{j}^{\nu})-y)|^{r^{\prime}}\,\mathrm{d}y
n2rj(m)2rjn(1k)2jw(k,ρ)(ξφ(x,ξjν)y)Mr|σ(ξφ(x,ξjν)y)|rdy\displaystyle\lesssim\int_{\mathbb{R}^{n}}\frac{2^{r^{\prime}j(m)}2^{r^{\prime}jn(1-k)}}{\left\langle 2^{jw(k,\rho)}(\nabla_{\xi}\varphi(x,\xi_{j}^{\nu})-y)\right\rangle^{Mr^{\prime}}}|\sigma(\nabla_{\xi}\varphi(x,\xi_{j}^{\nu})-y)|^{r^{\prime}}\,\mathrm{d}y
=n2rj(m)2rjn(1k)2jw(k,ρ)yMr|σ(y)|rdy\displaystyle=\int_{\mathbb{R}^{n}}\frac{2^{r^{\prime}j(m)}2^{r^{\prime}jn(1-k)}}{\left\langle 2^{jw(k,\rho)}y\right\rangle^{Mr^{\prime}}}|\sigma(y)|^{r^{\prime}}\,\mathrm{d}y
|y|>12jrm2(Mλ)rw(k,ρ)2jnr(1k)|y|(Mλ)rdy+|y|12jrm2jnr(1k)2jw(k,ρ)yMrdy\displaystyle\lesssim\int_{|y|>1}\frac{2^{jr^{\prime}m}2^{-(M-\lambda)r^{\prime}w(k,\rho)}2^{jnr^{\prime}(1-k)}}{|y|^{(M-\lambda)r^{\prime}}}\,\mathrm{d}y+\int_{|y|\leq 1}\frac{2^{jr^{\prime}m}2^{jnr^{\prime}(1-k)}}{\left\langle 2^{jw(k,\rho)}y\right\rangle^{Mr}}\,\mathrm{d}y
2jrm2(Mλ)rw(k,ρ)2jnr(1k)+2jrm2jn(r(1k)w(k,ρ)).\displaystyle\lesssim 2^{jr^{\prime}m}2^{-(M-\lambda)r^{\prime}w(k,\rho)}2^{jnr^{\prime}(1-k)}+2^{jr^{\prime}m}2^{jn(r^{\prime}(1-k)-w(k,\rho))}.

Furthermore, using (17) in [19, p. 57], we have

n|fj(y)|rdy|σ(ξφ(x,ξjν)y)|r(rfj(ξφ(x,ξjν)))r\int_{\mathbb{R}^{n}}\frac{|f_{j}(y)|^{r}\,\mathrm{d}y}{|\sigma(\nabla_{\xi}\varphi(x,\xi_{j}^{\nu})-y)|^{r}}\lesssim\big{(}\mathcal{M}_{r}f_{j}(\nabla_{\xi}\varphi(x,\xi_{j}^{\nu}))\big{)}^{r}

with a constant that only depends on the dimension nn. Thus (78) yields

|Tjfj(x)|2jrm/r2jn(r(1k)w(k,ρ))/rνrfj(ξφ(x,ξjν)).|T_{j}f_{j}(x)|\lesssim 2^{jr^{\prime}m/r}2^{jn(r^{\prime}(1-k)-w(k,\rho))/r}\sum_{\nu}\mathcal{M}_{r}f_{j}(\nabla_{\xi}\varphi(x,\xi_{j}^{\nu})).

From which it now follows for r<pr<p, the SND condition and the Hardy-Littlewood maximal theorem that

TjfjLp(n)\displaystyle\|T_{j}f_{j}\|_{L^{p}(\mathbb{R}^{n})} 2jrm/r2jn(r(1k)w(k,ρ))/rνrfj(ξφ(,ξjν))Lp(n)\displaystyle\lesssim 2^{jr^{\prime}m/r}2^{jn(r^{\prime}(1-k)-w(k,\rho))/r}\sum_{\nu}\big{\|}\mathcal{M}_{r}f_{j}(\nabla_{\xi}\varphi(\cdot,\xi_{j}^{\nu}))\big{\|}_{L^{p}(\mathbb{R}^{n})}
2jrm/r2jn(r(1k)w(k,ρ))/r2njkfjLp(n),\displaystyle\lesssim 2^{jr^{\prime}m/r}2^{jn(r^{\prime}(1-k)-w(k,\rho))/r}2^{njk}\|f_{j}\|_{L^{p}(\mathbb{R}^{n})},

and therefore, by Minkowski’s integral inequality,

TaφfBp,qs\displaystyle\|T_{a}^{\varphi}f\|_{B^{s}_{p,q}} (j=12jqsTjfjLp(n)q)1/q\displaystyle\lesssim\Big{(}\sum_{j=1}^{\infty}2^{jqs}\|T_{j}f_{j}\|^{q}_{L^{p}(\mathbb{R}^{n})}\Big{)}^{1/q}
(j=12jqs2qjrm/r2qjn(r(1k)w(k,ρ))/r2qnjkfjLp(n)q)1/q\displaystyle\lesssim\Big{(}\sum_{j=1}^{\infty}2^{jqs}2^{qjr^{\prime}m/r}2^{qjn(r^{\prime}(1-k)-w(k,\rho))/r}2^{qnjk}\|f_{j}\|^{q}_{L^{p}(\mathbb{R}^{n})}\Big{)}^{1/q}
fBp,qs(n),\displaystyle\lesssim\|f\|_{B^{s}_{p,q}(\mathbb{R}^{n})},

whenever

m<n(kp1+(11p)w(k,ρ)).m<n\Big{(}kp-1+\Big{(}1-\frac{1}{p}\Big{)}w(k,\rho)\Big{)}.

References

  • [1] A.-P. Calderón and R. Vaillancourt, A class of bounded pseudo-differential operators, Proc. Nat. Acad. Sci. U.S.A., 69 (1972), pp. 1185–1187.
  • [2] A. J. Castro, A. Israelsson, and W. Staubach, Regularity of fourier integral operators with amplitudes in general hörmander classes. (arXiv:2003.12878).
  • [3] A. J. Castro, A. Israelsson, W. Staubach, and M. Yerlanov, Regularity properties of Schrödinger integral operators and general oscillatory integrals. (arXiv:1912.08316).
  • [4] E. Cordero, F. Nicola, and L. Rodino, Sparsity of Gabor representation of Schrödinger propagators, Appl. Comput. Harmon. Anal., 26 (2009), pp. 357–370.
  • [5]  , Gabor analysis for Schrödinger equations and propagation of singularities, in Recent trends in operator theory and partial differential equations, vol. 258 of Oper. Theory Adv. Appl., Birkhäuser/Springer, Cham, 2017, pp. 257–274.
  • [6] D. Dos Santos Ferreira and W. Staubach, Global and local regularity of Fourier integral operators on weighted and unweighted spaces, Mem. Amer. Math. Soc., 229 (2014), pp. xiv+65.
  • [7] C. Fefferman, A note on spherical summation multipliers, Israel J. Math., 15 (1973), pp. 44–52.
  • [8] M. Frazier and B. Jawerth, The ϕ\phi-transform and applications to distribution spaces, in Function spaces and applications (Lund, 1986), vol. 1302 of Lecture Notes in Math., Springer, Berlin, 1988, pp. 223–246.
  • [9]  , A discrete transform and decompositions of distribution spaces, J. Funct. Anal., 93 (1990), pp. 34–170.
  • [10]  , Applications of the ϕ\phi and wavelet transforms to the theory of function spaces, in Wavelets and their applications, Jones and Bartlett, Boston, MA, 1992, pp. 377–417.
  • [11] Y.-S. Han, M. Paluszyński, and G. Weiss, A new atomic decomposition for the Triebel-Lizorkin spaces, in Harmonic analysis and operator theory (Caracas, 1994), vol. 189 of Contemp. Math., Amer. Math. Soc., Providence, RI, 1995, pp. 235–249.
  • [12] B. Helffer, Théorie spectrale pour des opérateurs globalement elliptiques, vol. 112 of Astérisque, Société Mathématique de France, Paris, 1984.
  • [13] B. Helffer and D. Robert, Comportement asymptotique précise du spectre d’opérateurs globalement elliptiques dans n\mathbb{R}^{n}. Goulaouic-Meyer-Schwartz Seminar, Exp. No. II, 23, École Polytech., Palaiseau, École Polytech., Palaiseau, 1981.
  • [14] A. Israelsson, T. Mattsson, and W. Staubach, Boundedness of fourier integral operators on classical function spaces, Manuscript submitted for publication, available on researchgate, (2022).
  • [15] A. Israelsson, S. Rodríguez-López, and W. Staubach, Local and global estimates for hyperbolic equations in Besov-Lipschitz and Triebel-Lizorkin spaces. To appear in Analysis & PDE 2020 arXiv:1802.05932v4.
  • [16] J. Johnsen, S. Munch Hansen, and W. Sickel, Anisotropic, mixed-norm Lizorkin-Triebel spaces and diffeomorphic maps, J. Funct. Spaces, (2014), pp. Art. ID 964794, 15.
  • [17] C. E. Kenig and W. Staubach, Ψ\Psi-pseudodifferential operators and estimates for maximal oscillatory integrals, Studia Math., 183 (2007), pp. 249–258.
  • [18] M. Pramanik, K. M. Rogers, and A. Seeger, A Calderón-Zygmund estimate with applications to generalized Radon transforms and Fourier integral operators, Studia Math., 202 (2011), pp. 1–15.
  • [19] E. M. Stein, Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, vol. 43 of Princeton Mathematical Series, Princeton University Press, Princeton, NJ, 1993.
  • [20] H. Triebel, Theory of function spaces, vol. 38 of Mathematik und ihre Anwendungen in Physik und Technik [Mathematics and its Applications in Physics and Technology], Akademische Verlagsgesellschaft Geest & Portig K.-G., Leipzig, 1983.
  • [21]  , Theory of function spaces II., vol. 84 of Monographs in Mathematics, Birkhäuser Verlag, Basel, 1992.