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

Sharp Local LpL^{p} estimates for the Hermite eigenfunctions

Xing Wang and Cheng Zhang Department of Mathematics, Hunan University, Changsha, HN 410012, China [email protected] Mathematical Sciences Center, Tsinghua University, Beijing 100084, China [email protected]
Abstract.

We investigate the concentration of eigenfunctions for the Hermite operator H=Δ+|x|2H=-\Delta+|x|^{2} in n\mathbb{R}^{n} by establishing local LpL^{p} bounds over the compact sets with arbitrary dilations and translations. These new results extend the local estimates by Thangavelu [62] and improve those derived from Koch-Tataru [45], and explain the special phenomenon that the global LpL^{p} bounds decrease in pp when 2p2n+6n+12\leq p\leq\frac{2n+6}{n+1}. The key L2L^{2}-estimates show that the local probabilities decrease away from the boundary {|x|=λ}\{|x|=\lambda\}, and then they satisfy Bohr’s correspondence principle in any dimension. The proof uses the Hermite spectral projection operator represented by Mehler’s formula for the Hermite-Schrödinger propagator eitHe^{-itH}, and the strategy developed by Thangavelu [62] and Jeong-Lee-Ryu [40]. We also exploit an explicit version of the stationary phase lemma and Hörmander’s L2L^{2} oscillatory integral theorem. Using Koch-Tataru’s strategy, we construct appropriate examples to illustrate the possible concentrations and show the optimality of our local estimates.

1. Introduction

In the seminal work of Sogge [52], he proved the LpL^{p} eigenfunction bounds for elliptic operators on compact manifolds. They are related to a variable coefficient version of Stein-Tomas restriction theorem and a number of core problems in harmonic analysis and PDEs. Sogge’s LpL^{p} bounds are sharp on the sphere SnS^{n}, because of its periodic Hamilton flow and many highly concentrated eigenfunctions, such as Gaussian beams and zonal functions. To investigate the concentration of eigenfunctions on manifolds, the LpL^{p} bounds over geodesic balls and tubes have been studied, see Bourgain [12], Blair-Sogge [6, 7, 8, 9, 10], Burq-Gérard-Tzvetkov [17], Han [29], Hezari-Rivière [32], Sogge [53], Sogge-Zelditch [56] and references therein. Specifically, Sogge [53] proved the following eigenfunction estimates over the geodesic balls B(x,r)B(x,r) with center xx and radius rr

(1.1) supxMeλL2(B(x,r))Cr12eλL2(M),λ1rInjM,\sup_{x\in M}\|e_{\lambda}\|_{L^{2}(B(x,r))}\leq Cr^{\frac{1}{2}}\|e_{\lambda}\|_{L^{2}(M)},\ \lambda^{-1}\leq r\leq\text{Inj}\ M,

where Inj MM is the injectivity radius of the compact manifold MM. These estimates are saturated on the standard spheres by zonal functions, and can be improved under some dynamical or geometric assumption, such as (M,g)(M,g) having everywhere nonpositive curvature. Sogge [53] also established the connection between local and global estimates

(1.2) eλL2(n+1)n1(M)Cλn12(n+1)(rn+14supxMeλL2(B(x,r)))2n+1,λ1rInjM.\left\|e_{\lambda}\right\|_{L^{\frac{2(n+1)}{n-1}}(M)}\leq C\lambda^{\frac{n-1}{2(n+1)}}\Big{(}r^{-\frac{n+1}{4}}\sup_{x\in M}\left\|e_{\lambda}\right\|_{L^{2}(B(x,r))}\Big{)}^{\frac{2}{n+1}},\quad\lambda^{-1}\leq r\leq\text{Inj}\ M.

One key ingredient in the proof of (1.2) is the finite propagation speed of wave equations.

The Hermite operator H=Δ+|x|2H=-\Delta+|x|^{2} in n\mathbb{R}^{n} shares some similar features with the spherical Laplacian, such as periodic Hamilton flow and many highly concentrated eigenfunctions, and the problem of obtaining LpL^{p} eigenfunction bounds has received considerable interest in the context of Bochner-Riesz means [61, 60, 62, 44, 39, 21, 20, 19, 22], as well as unique continuation problems [25, 26, 46]. To understand the nodal sets of the Hermite eigenfunctions in n\mathbb{R}^{n}, the sizes of nodal sets in small balls have been studied, see Bérard-Helffer [4, 3], Hanin-Zelditch-Zhou [30, 31], Beck-Hanin [2] and Jin [43]. In this paper, we investigate the concentration of the Hermite eigenfunctions in n\mathbb{R}^{n} by establishing sharp LpL^{p} bounds over compact sets. Similar local estimates has already been considered by Thangavelu [62] and Koch-Tataru [45], motivated by the Bochner-Riesz conjecture and its local version, and the unique continuation problems.

The Hermite functions are eigenfunctions of the one dimensional Hermite operator

hk′′+x2hk=(2k+1)hk,k.-h^{\prime\prime}_{k}+x^{2}h_{k}=(2k+1)h_{k},\ k\in\mathbb{N}.

They are

hk(x)=ex2/2(1)kdkdxkex2,k.h_{k}(x)=e^{x^{2}/2}(-1)^{k}\frac{d^{k}}{dx^{k}}e^{-x^{2}},\ k\in\mathbb{N}.

The Hermite functions form an orthonormal basis in L2()L^{2}(\mathbb{R}) after normalization.

In dimension nn, the Hermite eigenfunctions eλe_{\lambda} satisfy

(Δ+|x|2)eλ(x)=λ2eλ(x),xn,(-\Delta+|x|^{2})e_{\lambda}(x)=\lambda^{2}e_{\lambda}(x),\ \ x\in\mathbb{R}^{n},

and a complete set of eigenfunctions is given by

(1.3) hα(x)=j=1nhαj(xj)h_{\alpha}(x)=\prod_{j=1}^{n}h_{\alpha_{j}}(x_{j})

where the corresponding eigenvalue is λ2=2|α|+n\lambda^{2}=2|\alpha|+n, αn\alpha\in\mathbb{N}^{n}, |α|=αj|\alpha|=\sum\alpha_{j}. The multiplicity of the eigenvalue 2k+n2k+n is (k+n1)!k!(n1)!\frac{(k+n-1)!}{k!(n-1)!}.

1.1. Global estimates

Koch-Tataru [45, Corollary 3.2] proved the following global LpL^{p} eigenfunction bounds

(1.4) eλLp(n)λρ(p)eλL2(n),\|e_{\lambda}\|_{L^{p}(\mathbb{R}^{n})}\lesssim\lambda^{\rho(p)}\|e_{\lambda}\|_{L^{2}(\mathbb{R}^{n})},

where for n2n\geq 2,

(1.5) ρ(p)={12+1p, 2p<2n+6n+1n26n3p,2n+6n+1<p2nn2n22np,2nn2<p\displaystyle\rho(p)=\begin{cases}-\frac{1}{2}+\frac{1}{p},\ \ \ \ \ 2\leq p<\frac{2n+6}{n+1}\\ \frac{n-2}{6}-\frac{n}{3p},\ \ \ \frac{2n+6}{n+1}<p\leq\frac{2n}{n-2}\\ \frac{n-2}{2}-\frac{n}{p},\ \ \ \ \frac{2n}{n-2}<p\leq\infty\end{cases}

and for n=1n=1,

(1.6) ρ(p)={λ12+1p, 2p<4λ1613p, 4<p.\displaystyle\rho(p)=\begin{cases}\lambda^{-\frac{1}{2}+\frac{1}{p}},\ \ \ \ 2\leq p<4\\ \lambda^{-\frac{1}{6}-\frac{1}{3p}},\ \ \ 4<p\leq\infty.\end{cases}

See Figure 1. These results strengthen those of Karadzhov [44] and Thangavelu [62]. It is interesting to observe that the exponent ρ(p)\rho(p) is decreasing when p<2n+6n+1p<\frac{2n+6}{n+1}, and increasing when p>2n+6n+1p>\frac{2n+6}{n+1}. It is due to the special concentration features of the eigenfunctions, see the discussion after Theorem 3. At the kink point p=2n+6n+1p=\frac{2n+6}{n+1}, it is known that [47]

(1.7) eλL2n+6n+1(n)λ1n+3(logλ)n+12n+6eλL2(n).\|e_{\lambda}\|_{L^{\frac{2n+6}{n+1}}(\mathbb{R}^{n})}\lesssim\lambda^{-\frac{1}{n+3}}(\log\lambda)^{\frac{n+1}{2n+6}}\|e_{\lambda}\|_{L^{2}(\mathbb{R}^{n})}.

The log factor is necessary when n=1n=1. Recently, the log loss has been removed by Jeong-Lee-Ryu [41] for n3n\geq 3. Their significant improvements are due to a new phenomenon concerning the asymmetric localization near the sphere λSn1\lambda S^{n-1}, see [41, Theorem 1.2]. One may also refer to [40, 39, 42] for their related works on Bochner-Riesz means and LpLqL^{p}-L^{q} bounds for the Hermite spectral projection operator.

Refer to caption
Figure 1. Global LpL^{p} bounds of the Hermite eigenfunctions

The Hermite eigenfunctions are essentially concentrated in the ball {|x|λ}\{|x|\leq\lambda\} and have an exponential Airy type decay beyond this threshold. As was observed in [45], the behavior of eigenfunctions inside the ball {|x|λ}\{|x|\leq\lambda\} is not very different from (a rescaling of) what happens in a bounded domain. But considerable care is required near the boundary {|x|=λ}\{|x|=\lambda\}, where the concentration scales are different. Consequently, Koch-Tataru [45] split the space n\mathbb{R}^{n} into overlapping dyadic parts with respect to the distance to the boundary

Djint={λ|x|λ22j}, 12jλ23D_{j}^{int}=\{\lambda-|x|\approx\lambda 2^{-2j}\},\ 1\leq 2^{j}\leq\lambda^{\frac{2}{3}}
Dbd={||xλ|λ13}D^{bd}=\{||x-\lambda|\leq\lambda^{-\frac{1}{3}}\}
Dext={|x|>λ+12λ13}.D^{ext}=\{|x|>\lambda+\frac{1}{2}\lambda^{-\frac{1}{3}}\}.

Note that the thickness of the annulus DjintD_{j}^{int} is comparable to its distance to the boundary.

To state the main theorem in [45], we define the spaces lλqLpl_{\lambda}^{q}L^{p} of functions in n\mathbb{R}^{n} with norm

flλqLpq=fLp(Dext)q+fLp(Dbd)q+12jλ2/3fLp(Djint)q\|f\|_{l_{\lambda}^{q}L^{p}}^{q}=\|f\|_{L^{p}\left(D^{ext}\right)}^{q}+\|f\|_{L^{p}\left(D^{bd}\right)}^{q}+\sum_{1\leq 2^{j}\leq\lambda^{2/3}}\|f\|_{L^{p}\left(D_{j}^{int}\right)}^{q}

with the usual modification when q=q=\infty. For xnx\in\mathbb{R}^{n} we let

y=λ23(λ2|x|2),y=1+y,y+=1+y+y=\lambda^{-\frac{2}{3}}\left(\lambda^{2}-|x|^{2}\right),\quad\langle y\rangle_{-}=1+y_{-},\quad\langle y\rangle_{+}=1+y_{+}

Koch-Tataru [45] proved that for 2p2n+2n12\leq p\leq\frac{2n+2}{n-1},

(1.8) λ13n3(121p)y+14+n+34(121p)y1n2(121p)eλlλLpeλL2,\left\|\lambda^{\frac{1}{3}-\frac{n}{3}\left(\frac{1}{2}-\frac{1}{p}\right)}\langle y\rangle_{+}^{-\frac{1}{4}+\frac{n+3}{4}\left(\frac{1}{2}-\frac{1}{p}\right)}\langle y\rangle_{-}^{1-\frac{n}{2}\left(\frac{1}{2}-\frac{1}{p}\right)}e_{\lambda}\right\|_{l_{\lambda}^{\infty}L^{p}}\lesssim\|e_{\lambda}\|_{L^{2}},

and for 2n+2n1p\frac{2n+2}{n-1}\leq p\leq\infty,

(1.9) λ13n3(121p)y+12n2(121p)yNeλlλLpeλL2,N.\left\|\lambda^{\frac{1}{3}-\frac{n}{3}\left(\frac{1}{2}-\frac{1}{p}\right)}\langle y\rangle_{+}^{\frac{1}{2}-\frac{n}{2}\left(\frac{1}{2}-\frac{1}{p}\right)}\langle y\rangle_{-}^{N}e_{\lambda}\right\|_{l_{\lambda}^{\infty}L^{p}}\lesssim\|e_{\lambda}\|_{L^{2}},\ \forall N.

As a corollary, one can obtain LpL^{p} bounds over these dyadic annuli. For 2p2n+2n12\leq p\leq\frac{2n+2}{n-1}

(1.10) eλLp(Djint)λ1p122j(n+14n+32p)eλ2, 12jλ23,\|e_{\lambda}\|_{L^{p}(D_{j}^{int})}\lesssim\lambda^{\frac{1}{p}-\frac{1}{2}}2^{j(\frac{n+1}{4}-\frac{n+3}{2p})}\|e_{\lambda}\|_{2},\ \ 1\leq 2^{j}\leq\lambda^{\frac{2}{3}},
(1.11) eλLp(Dbd)+eλLp(Dext)λ13+n3(121p)eλ2\|e_{\lambda}\|_{L^{p}(D^{bd})}+\|e_{\lambda}\|_{L^{p}(D^{ext})}\lesssim\lambda^{-\frac{1}{3}+\frac{n}{3}(\frac{1}{2}-\frac{1}{p})}\|e_{\lambda}\|_{2}

and for 2n+2n1p\frac{2n+2}{n-1}\leq p\leq\infty

(1.12) eλLp(Djint)(λ2j)n22npeλ2, 12jλ23,\|e_{\lambda}\|_{L^{p}(D_{j}^{int})}\lesssim(\lambda 2^{-j})^{\frac{n-2}{2}-\frac{n}{p}}\|e_{\lambda}\|_{2},\ \ 1\leq 2^{j}\leq\lambda^{\frac{2}{3}},
(1.13) eλLp(Dbd)+eλLp(Dext)λ13+n3(121p)eλ2.\|e_{\lambda}\|_{L^{p}(D^{bd})}+\|e_{\lambda}\|_{L^{p}(D^{ext})}\lesssim\lambda^{-\frac{1}{3}+\frac{n}{3}(\frac{1}{2}-\frac{1}{p})}\|e_{\lambda}\|_{2}.

One may observe that the local L2L^{2} estimate over the dilated ball D0int={|x|<λ/4}D_{0}^{int}=\{|x|<\lambda/4\} has no improvement on the trivial bound. Similarly, the local L2n+6n+1L^{\frac{2n+6}{n+1}} bounds in (1.10) cannot essentially improve the global estimate (1.7) over the whole space. Moreover, the local LpL^{p} bounds (1.12) over D0intD_{0}^{int} strengthen Thangavelu’s estimates [62, Theorem 2]

(1.14) eλLp(B)Cλn22npeλL2(n),2n+2n1p.\|e_{\lambda}\|_{L^{p}(B)}\leq C\lambda^{\frac{n-2}{2}-\frac{n}{p}}\|e_{\lambda}\|_{L^{2}(\mathbb{R}^{n})},\ \ \tfrac{2n+2}{n-1}\leq p\leq\infty.

where BB is any fixed compact set in n\mathbb{R}^{n}. This means that replacing BB by a much larger dilated ball D0intD_{0}^{int} does not affect the bounds. This interesting phenomenon can be explained by the existence of the Hermite eigenfunctions with point concentration near the origin. See Section 5 and [45, Example 5.2].

1.2. Bohr’s correspondence principle

Now let’s go back to the starting point of this study. Bohr’s correspondence principle demands that classical physics and quantum physics give the same answer when the systems become large, see [11, 63]. For the classical harmonic oscillator, when a particle’s potential energy is equal to its energy level, it moves at its slowest speed and has the highest probability of being detected around those points. We call these points turning points, which actually correspond to the boundary {|x|=λ}\{|x|=\lambda\}. By the correspondence principle, we would expect that for a quantum harmonic oscillator, for eigenstates with large energy, their probability density should also peak near these turning points in some sense. For the one-dimensional case, let

s(x)=0x|t2λ2|𝑑t,s+(x)=λx|t2λ2|𝑑t.s^{-}(x)=\int_{0}^{x}\sqrt{|t^{2}-\lambda^{2}|}dt,\ \ \ \ s^{+}(x)=\int_{\lambda}^{x}\sqrt{|t^{2}-\lambda^{2}|}dt.

As in Szegö [58, p. 201] and Koch-Tataru [45, Lemma 5.1], the normalized eigenfunctions h~k=hk/hk2\tilde{h}_{k}=h_{k}/\|h_{k}\|_{2} satisfy

h~2k\displaystyle\tilde{h}_{2k} ={a2k(λ2x2)14(coss(x)+ error )|x|<λλ13O(λ16)λλ13xλ+λ13a2k+es+(x)(λ2x2)14(1+ error )|x|>λ+λ13\displaystyle=\begin{cases}a_{2k}^{-}\left(\lambda^{2}-x^{2}\right)^{-\frac{1}{4}}\left(\cos s^{-}(x)+\text{ error }\right)&|x|<\lambda-\lambda^{-\frac{1}{3}}\\ O(\lambda^{-\frac{1}{6}})&\lambda-\lambda^{-\frac{1}{3}}\leq x\leq\lambda+\lambda^{-\frac{1}{3}}\\ a_{2k}^{+}e^{-s^{+}(x)}\left(\lambda^{2}-x^{2}\right)^{-\frac{1}{4}}(1+\text{ error })&|x|>\lambda+\lambda^{-\frac{1}{3}}\end{cases}
h~2k+1\displaystyle\tilde{h}_{2k+1} ={a2k+1(λ2x2)14(sins(x)+ error )|x|<λλ13O(λ16)λλ13xλ+λ13a2k+1+es+(x)(λ2x2)14(1+ error )x>λ+λ13\displaystyle=\begin{cases}a_{2k+1}^{-}\left(\lambda^{2}-x^{2}\right)^{-\frac{1}{4}}\left(\sin s^{-}(x)+\text{ error }\right)&|x|<\lambda-\lambda^{-\frac{1}{3}}\\ O(\lambda^{-\frac{1}{6}})&\lambda-\lambda^{-\frac{1}{3}}\leq x\leq\lambda+\lambda^{-\frac{1}{3}}\\ a_{2k+1}^{+}e^{-s^{+}(x)}\left(\lambda^{2}-x^{2}\right)^{-\frac{1}{4}}(1+\text{ error })&x>\lambda+\lambda^{-\frac{1}{3}}\end{cases}

where |ak±|1|a_{k}^{\pm}|\sim 1,  error =O((|x2λ2|12||x|λ|1)\text{ error }=O((\left|x^{2}-\lambda^{2}\right|^{-\frac{1}{2}}||x|-\lambda|^{-1}). Obviously the maximum values of Hermite eigenfunctions occur near the turning points x=±λx=\pm\lambda, see Figure 2 by Wolfram Mathematica.

Refer to caption
Figure 2. Hermite function h100h_{100}

However, the eigenfunctions become more complicated in higher dimensions. Koch-Tataru [45] showed that there are eigenfunctions that attain maximal LL^{\infty} growth near the origin. This phenomenon seems to violate Bohr’s correspondence principle, and suggests that it may be not suitable to measure only by the local LL^{\infty} norm. A more reliable choice should be the local L2L^{2} norm, i.e. local probability, since the square of the amplitude is interpreted as a probability density. As we will see in Theorem 1, the local probability over a compact set, such as the unit ball with arbitrary center, decreases as it moves away from the boundary {|x|=λ}\{|x|=\lambda\}. This satisfies Bohr’s correspondence principle.

1.3. Main theorems

Let BB be any fixed compact set in n\mathbb{R}^{n}, and νn\nu\in\mathbb{R}^{n}. For r>0r>0, let

(1.15) B(ν,r)={ν+rx:xB}B(\nu,r)=\{\nu+rx:x\in B\}

be the compact set with dilation rate rr and translation vector ν\nu. We shall restrict |ν|λ|\nu|\leq\lambda and rλr\leq\lambda, since the Hermite eigenfunctions are essentially supported in the ball {|x|λ}\{|x|\leq\lambda\}. The LpL^{p} bounds (1.14) of the Hermite eigenfunctions over the fixed compact sets BB for p2n+2n1p\geq\frac{2n+2}{n-1} have been established by Thangavelu [62], motivated by the local version of the Bochner-Riesz conjecture. Koch-Tataru [45] strengthened those of Thangavelu [62] as well as Karadzhov [44] by obtaining global bounds over n\mathbb{R}^{n} and dyadic annuli (including the dilated ball D0intD_{0}^{int}), and they also observed that eigenfunctions may concentrate in some small compact subsets, e.g. balls and tubes.

To our best knowledge, compared to the Laplace eigenfunctions on compact manifolds, the picture of the Hermite eigenfunction estimates are still far from complete. See Section 6 for a list of related open problems. So we aim to investigate these problems in a series of works. In this paper, we first establish sharp estimates over the compact sets B(ν,r)B(\nu,r) with arbitrary dilations and translations, which extend the local estimates of Thangavelu [62] and improve those derived from Koch-Tataru [45]. Our main theorem (Theorem 1) gives sharp local L2L^{2} bounds (local probabilities) over the compact sets with arbitrary dilations and translations.

Theorem 1.

Let B(ν,r)B(\nu,r) be the compact set in (1.15) and μ=max{λ43,1λ1|ν|}\mu=\max\{\lambda^{-\frac{4}{3}},1-\lambda^{-1}|\nu|\}. Then for n1n\geq 1 we have

(1.16) eλL2(B(ν,r))CΛ(λ,r,ν)eλL2(n),\|e_{\lambda}\|_{L^{2}(B(\nu,r))}\leq C\Lambda(\lambda,r,\nu)\|e_{\lambda}\|_{L^{2}(\mathbb{R}^{n})},

where

Λ(λ,r,ν)={(λμ12)n22rn2,r(λμ12)1(λμ12/r)12,(λμ12)1rλμ(λ/r)14,λμrλ.\Lambda(\lambda,r,\nu)=\begin{cases}(\lambda\mu^{\frac{1}{2}})^{\frac{n-2}{2}}r^{\frac{n}{2}},\ \ r\lesssim(\lambda\mu^{\frac{1}{2}})^{-1}\\ (\lambda\mu^{\frac{1}{2}}/r)^{-\frac{1}{2}},\ \ \ (\lambda\mu^{\frac{1}{2}})^{-1}\ll r\ll\lambda\mu\\ (\lambda/r)^{-\frac{1}{4}},\ \ \ \ \ \ \ \ \lambda\mu\lesssim r\leq\lambda.\end{cases}

These bounds are sharp.

See Figures 3, 4, 5, 6.

Refer to caption
Figure 3. Local L2L^{2} bounds with respect to rr
Refer to caption
Figure 4. Local L2L^{2} bounds with respect to μ\mu when rλ13r\gtrsim\lambda^{-\frac{1}{3}}
Refer to caption
Figure 5. Local L2L^{2} bounds with respect to μ\mu when λ1rλ13\lambda^{-1}\lesssim r\lesssim\lambda^{-\frac{1}{3}}
Refer to caption
Figure 6. Local L2L^{2} bounds with respect to μ\mu when rλ1r\lesssim\lambda^{-1}

These results improve the bounds derived from Koch-Tataru’s estimates (1.8) and (1.9) when (λμ12)1rλμ(\lambda\mu^{\frac{1}{2}})^{-1}\ll r\ll\lambda\mu. The endpoints (λμ12)1(\lambda\mu^{\frac{1}{2}})^{-1} and λμ\lambda\mu correspond to the sizes of two different kinds of eigenfunction concentrations: point concentration and tube concentration respectively. These are similar to the eigenfunctions of the spherical Laplacian, e.g. zonal functions and Gaussian beams. Indeed, Koch-Tataru [45] used the eigenbasis (1.3) and linear combinations to construct two kinds of eigenfunctions concentrated in the dyadic annulus DjintD_{j}^{int}. One concentrates in the ball of radius (λ2j)1(\lambda 2^{-j})^{-1}, while the other concentrates in the tube with length λ22j\lambda 2^{-2j} and radius 2j/22^{-j/2}. We shall exploit the strategy in [45] to construct new intermediate examples between these two extreme cases, and prove the sharpness of local LpL^{p} estimates, see Section 5 for the detailed construction.

For compact sets with fixed size rr, we shall discuss the relation between their locations and the local L2L^{2} bounds. It is interesting to note that the L2L^{2} bounds (1.16) with rr larger than the threshold λ13\lambda^{-\frac{1}{3}} are decreasing in μ\mu with the decay rate μ14\sim\mu^{-\frac{1}{4}} when μr/λ\mu\gg r/\lambda. See Figure 4. This suggests that the local probabilities are decreasing away from the boundary and then satisfy Bohr’s correspondence principle. When the size rr is smaller than the threshold λ13\lambda^{-\frac{1}{3}}, there are some significant differences between one dimension and higher dimensions. See Figures 5, 6. It is because in the sense of LL^{\infty} norm, higher dimensional eigenfunctions can concentrate away from the boundary, while one dimensional eigenfunctions only concentrate near the boundary.

An important special case of Theorem 1 is the μ=r=1\mu=r=1 case, which extends Thangavelu’s local estimates (1.14) to all p2p\geq 2.

Theorem 2.

Let BB be any fixed compact set in n\mathbb{R}^{n}. Then for n1n\geq 1 we have

(1.17) eλLp(B)Cλσ(p)12eλL2(n),\|e_{\lambda}\|_{L^{p}(B)}\leq C\lambda^{\sigma(p)-\frac{1}{2}}\|e_{\lambda}\|_{L^{2}(\mathbb{R}^{n})},

where

σ(p)={n12(121p), 2p<2n+2n1n12np,2n+2n1p\sigma(p)=\begin{cases}\frac{n-1}{2}(\frac{1}{2}-\frac{1}{p}),\ \ \ 2\leq p<\frac{2n+2}{n-1}\\ \frac{n-1}{2}-\frac{n}{p},\ \ \ \ \ \ \frac{2n+2}{n-1}\leq p\leq\infty\end{cases}

is Sogge’s exponent. These bounds are sharp.

Theorem 2 is new for 2p<2n+2n12\leq p<\frac{2n+2}{n-1}. It is enlightening to compare (1.17) with Sogge’s LpL^{p} estimates [52] on compact manifolds

(1.18) eλLp(M)Cλσ(p)eλL2(M), 2p.\|e_{\lambda}\|_{L^{p}(M)}\leq C\lambda^{\sigma(p)}\|e_{\lambda}\|_{L^{2}(M)},\ 2\leq p\leq\infty.

This phenomenon suggests that the Hermite eigenfunctions locally resemble the rescaled Laplace eigenfunctions.

Furthermore, optimal local LpL^{p} estimates for all p2p\geq 2 can be obtained by the interpolation between Theorem 1 and Koch-Tataru’s LpL^{p} bounds.

Theorem 3.

Let B(ν,r)B(\nu,r) be the compact set in (1.15) and μ=max{λ43,1λ1|ν|}\mu=\max\{\lambda^{-\frac{4}{3}},1-\lambda^{-1}|\nu|\}, and μ~=max{λ43,μλ1r}\tilde{\mu}=\max\{\lambda^{-\frac{4}{3}},\mu-\lambda^{-1}r\}. Then for n1n\geq 1 we have

(1.19) eλLp(B(ν,r))CΛ(λ,r,ν,p)eλL2(n),\|e_{\lambda}\|_{L^{p}(B(\nu,r))}\leq C\Lambda(\lambda,r,\nu,p)\|e_{\lambda}\|_{L^{2}(\mathbb{R}^{n})},

where for r(λμ12)1r\lesssim(\lambda\mu^{\frac{1}{2}})^{-1}

Λ(λ,r,ν,p)=(λμ12)n22rnp, 2p,\Lambda(\lambda,r,\nu,p)=(\lambda\mu^{\frac{1}{2}})^{\frac{n-2}{2}}r^{\frac{n}{p}},\ \ 2\leq p\leq\infty,

and for (λμ12)1rλμ(\lambda\mu^{\frac{1}{2}})^{-1}\ll r\ll\lambda\mu

Λ(λ,r,ν,p)={(λμ12/r)n14n+12p(λμ12)1p12, 2p2n+2n1(λμ12)n22np2n+2n1<p,\Lambda(\lambda,r,\nu,p)=\begin{cases}(\lambda\mu^{\frac{1}{2}}/r)^{\frac{n-1}{4}-\frac{n+1}{2p}}(\lambda\mu^{\frac{1}{2}})^{\frac{1}{p}-\frac{1}{2}},\ \ \ 2\leq p\leq\frac{2n+2}{n-1}\\ (\lambda\mu^{\frac{1}{2}})^{\frac{n-2}{2}-\frac{n}{p}}\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \quad\quad\frac{2n+2}{n-1}<p\leq\infty,\end{cases}

and for λμrλ\lambda\mu\lesssim r\leq\lambda

Λ(λ,r,ν,p)={(r/λ)n+34pn+18λ1p12, 2p<2n+6n+1μ~n+34pn+18λ1p12,2n+6n+1<p2n+2n1(λμ~12)n22np,2n+2n1<p2nn2(λ12r12)n22np,2nn2<p.\Lambda(\lambda,r,\nu,p)=\begin{cases}(r/\lambda)^{\frac{n+3}{4p}-\frac{n+1}{8}}\lambda^{\frac{1}{p}-\frac{1}{2}},\ \ 2\leq p<\frac{2n+6}{n+1}\\ \tilde{\mu}^{\frac{n+3}{4p}-\frac{n+1}{8}}\lambda^{\frac{1}{p}-\frac{1}{2}},\ \ \ \ \ \ \frac{2n+6}{n+1}<p\leq\frac{2n+2}{n-1}\\ (\lambda\tilde{\mu}^{\frac{1}{2}})^{\frac{n-2}{2}-\frac{n}{p}},\ \ \ \ \ \ \ \ \ \frac{2n+2}{n-1}<p\leq\frac{2n}{n-2}\\ (\lambda^{\frac{1}{2}}r^{\frac{1}{2}})^{\frac{n-2}{2}-\frac{n}{p}},\ \ \ \ \ \ \ \ \frac{2n}{n-2}<p\leq\infty.\end{cases}

These bounds are sharp.

Here we use the convention that the endpoints 2n+2n1=\frac{2n+2}{n-1}=\infty when n=1n=1, and 2nn2=\frac{2n}{n-2}=\infty when n=1,2n=1,2. At the kink point p=2n+6n+1p=\frac{2n+6}{n+1}, the local LpL^{p} bounds for λμrλ\lambda\mu\lesssim r\leq\lambda coincide with the global estimates (1.7). The number λμ~\lambda\tilde{\mu} is essentially the distance between the compact B(ν,r)B(\nu,r) and the boundary {|x|=λ}\{|x|=\lambda\}, and obviously λμ~λμ\lambda\tilde{\mu}\leq\lambda\mu.

Theorem 3 improves the local bounds derived from Koch-Tataru’s estimates (1.8) and (1.9) when (λμ12)1rλμ(\lambda\mu^{\frac{1}{2}})^{-1}\ll r\ll\lambda\mu and 2p2n+2n12\leq p\leq\frac{2n+2}{n-1}. Moreover, for any fixed rr, we can use Theorem 3 to determine the locations of ν\nu (i.e. conditions on μ\mu) that can attain the maximal local LpL^{p} norms

supνneλLp(B(ν,r)).\sup_{\nu\in\mathbb{R}^{n}}\|e_{\lambda}\|_{L^{p}(B(\nu,r))}.

See the following Tables 1, 2, 3 with n2n\geq 2.

These tables demonstrate some new concentration features of the eigenfunctions. For 2p<2n+6n+12\leq p<\frac{2n+6}{n+1}, the maximal local LpL^{p} bounds significantly improve the global estimates λ1p12\lambda^{\frac{1}{p}-\frac{1}{2}} whenever rλr\ll\lambda. However, the maximal local LpL^{p} bounds cannotcannot improve the global estimates (1.4) for p>2n+6n+1p>\frac{2n+6}{n+1} and rλ13r\gtrsim\lambda^{-\frac{1}{3}}. This phenomenon suggests that the Hermite eigenfunctions cannotcannot highly concentrate in any “small” compact set with size rλr\ll\lambda, in terms of the LpL^{p} norm with pp less than 2n+6n+1\frac{2n+6}{n+1}. But this kind of concentration is possible for all larger pp. These features of the eigenfunctions explain why the exponent ρ(p)\rho(p) in (1.4) is decreasing in pp when p<2n+6n+1p<\frac{2n+6}{n+1}, and increasing when p>2n+6n+1p>\frac{2n+6}{n+1}. See Figure 1.

Table 1. Maximal local estimates for 2p<2n+6n+12\leq p<\frac{2n+6}{n+1}
λ13rλ\lambda^{-\frac{1}{3}}\lesssim r\leq\lambda λ1rλ13\lambda^{-1}\lesssim r\lesssim\lambda^{-\frac{1}{3}} rλ1r\lesssim\lambda^{-1}
max LpL^{p} bounds (λ/r)n+18n+34pλ1p12(\lambda/r)^{\frac{n+1}{8}-\frac{n+3}{4p}}\lambda^{\frac{1}{p}-\frac{1}{2}} rnpn22r^{\frac{n}{p}-\frac{n-2}{2}} λn22rnp\lambda^{\frac{n-2}{2}}r^{\frac{n}{p}}
conditions on μ\mu μr/λ\mu\lesssim r/\lambda μ(λr)2\mu\approx(\lambda r)^{-2} μ1\mu\approx 1
Table 2. Maximal local estimates for 2n+6n+1<p2nn2\frac{2n+6}{n+1}<p\leq\frac{2n}{n-2}
λ13rλ\lambda^{-\frac{1}{3}}\lesssim r\leq\lambda λ1rλ13\lambda^{-1}\lesssim r\lesssim\lambda^{-\frac{1}{3}} rλ1r\lesssim\lambda^{-1}
max LpL^{p} bounds λn26n3p\lambda^{\frac{n-2}{6}-\frac{n}{3p}} rnpn22r^{\frac{n}{p}-\frac{n-2}{2}} λn22rnp\lambda^{\frac{n-2}{2}}r^{\frac{n}{p}}
conditions on μ\mu λμrλ13\lambda\mu-r\lesssim\lambda^{-\frac{1}{3}} μ(λr)2\mu\approx(\lambda r)^{-2} μ1\mu\approx 1
Table 3. Maximal local estimates for 2nn2<p\frac{2n}{n-2}<p\leq\infty
λ1rλ\lambda^{-1}\lesssim r\leq\lambda rλ1r\lesssim\lambda^{-1}
max LpL^{p} bounds λn22np\lambda^{\frac{n-2}{2}-\frac{n}{p}} λn22rnp\lambda^{\frac{n-2}{2}}r^{\frac{n}{p}}
conditions on μ\mu μ1\mu\approx 1 μ1\mu\approx 1

1.4. Paper structure and proof sketch

The paper is structured as follows. In Section 2, we presents the proof of the stationary phase lemma with a precise remainder term. In Section 3, we review the representation of the kernel of the Hermite spectral projection operator. In Section 4, we prove Theorem 1. In Section 5, we show the sharpness of local LpL^{p} bounds. In Section 6, we discuss some related questions for the Hermite eigenfunctions.

Proof sketch of Theorem 1. We handle the Hermite spectral projection operator represented by Mehler’s formula for the kernel of the Hermite-Schrödinger propagator eitHe^{-itH}. By the strategy developed by Thangavelu [61, 60, 62] and Jeong-Lee-Ryu [41, 40, 39, 42], we explicitly analyze the associated oscillatory integrals by the stationary phase lemma and Hörmander’s L2L^{2} oscillatory integral theorem [35]. The main difficulties lie in the discussions concerning the critical points of the phase functions in these oscillatory integrals, which require new insights. We find that in our local problem, the Hermite spectral projection operator essentially consists of two oscillatory integral operators with different phase functions, modulo some remainder terms. One phase function behaves like the Euclidean distance function, while the other one has no obvious geometric meaning. Nevertheless, both of the phase functions satisfy the mixed Hessian condition (with rank =n1=n-1) in Hörmander’s L2L^{2} oscillatory integral theorem.

Proof sketch of Theorem 3. When rλ1μ12r\lesssim\lambda^{-1}\mu^{-\frac{1}{2}}, we may apply Koch-Tataru’s LL^{\infty} estimates (1.12), (1.13) and then the local LpL^{p} bounds follow from Hölder’s inequality. When λμrλ\lambda\mu\lesssim r\leq\lambda, we may essentially cover the set B(ν,r)B(\nu,r) by the dyadic annuli DjintD_{j}^{int} with λμ~λ22jr\lambda\tilde{\mu}\lesssim\lambda 2^{-2j}\lesssim r (and the boundary annuli, if μ~=λ43\tilde{\mu}=\lambda^{-\frac{4}{3}}), so the local LpL^{p} bounds are implied by Koch-Tataru’s LpL^{p} bounds over these annuli. When λ1μ12rλμ\lambda^{-1}\mu^{-\frac{1}{2}}\ll r\ll\lambda\mu, the set B(ν,r)B(\nu,r) is essentially covered by the annulus DjintD_{j}^{int} with 22jμ2^{-2j}\approx\mu, so the local LpL^{p} bounds follow from the interpolation between Koch-Tataru’s LpL^{p} bounds over DjintD_{j}^{int} and the local L2L^{2} bounds in Theorem 1.

1.5. Notations

Throughout this paper, XYX\lesssim Y means CXYCX\leq Y for some positive constants CC that depend only on dimension nn and the number of times we take derivatives and integrate by parts. In particular, if C>1C>1 is a large constant and CXYCX\leq Y, then we denote XYX\ll Y. If XYX\lesssim Y and YXY\lesssim X, we denote XYX\approx Y. The notation fp\|f\|_{p} means the Lebesgue norm of ff in n\mathbb{R}^{n}. Sometimes we abbreviate the phase function ψ(t,x,y)\psi(t,x,y) as ψ(t)\psi(t) when x,yx,y are fixed, and denote its partial derivative with respect to tt by ψ(t)\psi^{\prime}(t).

2. Stationary phase lemmas

In this section, we review the one dimensional stationary phase lemma, which is important to analyze the kernel of the spectral projection operator. It is classical and there are many excellent references, e.g. Hörmander [36], Stein [57] and Sogge [54]. Using the idea in Hörmander [36, Theorem 7.7.5], we prove the explicit stationary phase Lemma 4 with a more precise remainder term than the one presented in [36].

2.1. Explicit non-stationary phase lemma

Let aC0()a\in C_{0}^{\infty}(\mathbb{R}). Let ϕC()\phi\in C^{\infty}(\mathbb{R}) be real-valued. We estimate

Iλ=eiλϕ(t)a(t)𝑑t.I_{\lambda}=\int e^{i\lambda\phi(t)}a(t)dt.

Let N1N\geq 1, bC()b\in C^{\infty}(\mathbb{R}) and =ddt\partial=\frac{d}{dt}. We define

(b)(a)=(ab)=ab+ba.(\partial\circ b)(a)=\partial(ab)=a\partial b+b\partial a.

Note that

(b)N(a)=|α|=Ncαα0aα1bαNb,(\partial\circ b)^{N}(a)=\sum_{|\alpha|=N}c_{\alpha}\partial^{\alpha_{0}}a\partial^{\alpha_{1}}b\cdot\cdot\cdot\partial^{\alpha_{N}}b,

where α=(α1,,αN)N\alpha=(\alpha_{1},...,\alpha_{N})\in\mathbb{N}^{N}, and |α|=α0++αN|\alpha|=\alpha_{0}+...+\alpha_{N}. For example,

(b)2(a)=(b)(ab+ba)=b22a+3bab+abb+ab2b.(\partial\circ b)^{2}(a)=(\partial\circ b)(a\partial b+b\partial a)=b^{2}\partial^{2}a+3b\partial a\partial b+a\partial b\partial b+ab\partial^{2}b.

Moreover, for j=1,,Nj=1,...,N,

αj(f1)=fαj1|γj|=αjcγjγ1jfγαjjf,\partial^{\alpha_{j}}(f^{-1})=f^{-\alpha_{j}-1}\sum_{|\gamma^{j}|=\alpha_{j}}c_{\gamma^{j}}\partial^{\gamma^{j}_{1}}f\cdot\cdot\cdot\partial^{\gamma^{j}_{\alpha_{j}}}f,

where γj=(γ1j,,γαjj)\gamma^{j}=(\gamma_{1}^{j},...,\gamma^{j}_{\alpha_{j}}). Here we use the convention that the sum is 1 if αj=0\alpha_{j}=0.

Suppose that ϕ0\phi^{\prime}\neq 0 on supp a\text{supp }a. Integrating by parts NN times, we have

Iλ\displaystyle I_{\lambda} λN|supp a|((ϕ1))N(a)\displaystyle\lesssim\lambda^{-N}|\text{supp }a|\|(\partial\circ(\phi^{\prime-1}))^{N}(a)\|_{\infty}
λN|supp a||α|=N(α0a)(ϕ)α02Nj=1N|γj|=αj|γ1j(ϕ)γαjj(ϕ)|\displaystyle\lesssim\lambda^{-N}|\text{supp }a|\sum_{|\alpha|=N}\|(\partial^{\alpha_{0}}a)(\phi^{\prime})^{\alpha_{0}-2N}\prod_{j=1}^{N}\sum_{|\gamma^{j}|=\alpha_{j}}|\partial^{\gamma_{1}^{j}}(\phi^{\prime})\cdot\cdot\cdot\partial^{\gamma_{\alpha_{j}}^{j}}(\phi^{\prime})|\|_{\infty}
λN|supp a|α0=0N|β|=Nα0(α0a)(ϕ)α02Nβ1(ϕ)βNα0(ϕ)\displaystyle\lesssim\lambda^{-N}|\text{supp }a|\sum_{\alpha_{0}=0}^{N}\sum_{|\beta|=N-\alpha_{0}}\|(\partial^{\alpha_{0}}a)(\phi^{\prime})^{\alpha_{0}-2N}\partial^{\beta_{1}}(\phi^{\prime})\cdot\cdot\cdot\partial^{\beta_{N-\alpha_{0}}}(\phi^{\prime})\|_{\infty}
λN|supp a|α0=0Nσ(α0a)(ϕ)α02N+σ0(ϕ′′)σ1(ϕ(Nα0+1))σNα0\displaystyle\lesssim\lambda^{-N}|\text{supp }a|\sum_{\alpha_{0}=0}^{N}\sum_{\sigma}\|(\partial^{\alpha_{0}}a)(\phi^{\prime})^{\alpha_{0}-2N+\sigma_{0}}(\phi^{\prime\prime})^{\sigma_{1}}\cdot\cdot\cdot(\phi^{(N-\alpha_{0}+1)})^{\sigma_{N-\alpha_{0}}}\|_{\infty}

where β=(β1,,βNα0)\beta=(\beta_{1},...,\beta_{N-\alpha_{0}}), and σ=(σ0,,σNα0)\sigma=(\sigma_{0},...,\sigma_{N-\alpha_{0}}) satisfies kσk=Nα0=σk\sum k\sigma_{k}=N-\alpha_{0}=\sum\sigma_{k}.

2.2. Explicit stationary phase lemma

Suppose that ϕC()\phi\in C^{\infty}(\mathbb{R}) is real-valued and satisfies ϕ(0)=ϕ(0)=0,ϕ′′(0)0\phi(0)=\phi^{\prime}(0)=0,\phi^{\prime\prime}(0)\neq 0. Let

(2.1) g(t)=ϕ(t)12ϕ′′(0)t2,g(t)=\phi(t)-\frac{1}{2}\phi^{\prime\prime}(0)t^{2},
(2.2) ϕθ(t)=12ϕ′′(0)t2+θg(t), 0θ1.\phi_{\theta}(t)=\frac{1}{2}\phi^{\prime\prime}(0)t^{2}+\theta g(t),\ 0\leq\theta\leq 1.

Then ϕ1=ϕ\phi_{1}=\phi. Suppose that

|t||ϕ′′(0)|/ϕ′′′,tsupp a.|t|\leq|\phi^{\prime\prime}(0)|/\|\phi^{\prime\prime\prime}\|_{\infty},\ \ \forall t\in\text{supp }a.

Then

|g(t)/t3|16ϕ′′′,|g(t)/t^{3}|\leq\frac{1}{6}\|\phi^{\prime\prime\prime}\|_{\infty},
|g(t)/t2|12ϕ′′′,|g^{\prime}(t)/t^{2}|\leq\frac{1}{2}\|\phi^{\prime\prime\prime}\|_{\infty},
|g′′(t)/t|ϕ′′′,|g^{\prime\prime}(t)/t|\leq\|\phi^{\prime\prime\prime}\|_{\infty},
|g(k)(t)|ϕ(k),k3.|g^{(k)}(t)|\leq\|\phi^{(k)}\|_{\infty},\ k\geq 3.
|t/ϕθ(t)|=1|ϕ′′(0)+θg(t)/t|1|ϕ′′(0)|12|t|ϕ′′′2|ϕ′′(0)|,|t/\phi_{\theta}^{\prime}(t)|=\frac{1}{|\phi^{\prime\prime}(0)+\theta g^{\prime}(t)/t|}\leq\frac{1}{|\phi^{\prime\prime}(0)|-\frac{1}{2}|t|\|\phi^{\prime\prime\prime}\|_{\infty}}\leq\frac{2}{|\phi^{\prime\prime}(0)|},
|t/ϕθ(t)|=1|ϕ′′(0)+θg(t)/t|1|ϕ′′(0)|+12|t|ϕ′′′23|ϕ′′(0)|.|t/\phi_{\theta}^{\prime}(t)|=\frac{1}{|\phi^{\prime\prime}(0)+\theta g^{\prime}(t)/t|}\geq\frac{1}{|\phi^{\prime\prime}(0)|+\frac{1}{2}|t|\|\phi^{\prime\prime\prime}\|_{\infty}}\geq\frac{2}{3|\phi^{\prime\prime}(0)|}.

Thus

|ϕθ(t)||t||ϕ′′(0)|,|\phi_{\theta}^{\prime}(t)|\approx|t||\phi^{\prime\prime}(0)|,
|ϕθ′′(t)|=|ϕ′′(0)+θg′′(t)|2|ϕ′′(0)|,|\phi_{\theta}^{\prime\prime}(t)|=|\phi^{\prime\prime}(0)+\theta g^{\prime\prime}(t)|\leq 2|\phi^{\prime\prime}(0)|,
|ϕθ(k)(t)|ϕ(k),k3.|\phi_{\theta}^{(k)}(t)|\leq\|\phi^{(k)}\|_{\infty},\ k\geq 3.

Let

Iλ(θ)=eiλϕθ(t)a(t)𝑑t,aC0().I_{\lambda}(\theta)=\int e^{i\lambda\phi_{\theta}(t)}a(t)dt,\ \ a\in C_{0}^{\infty}(\mathbb{R}).

For m1m\geq 1, we need to estimate

Iλ(1)=k=02m1Iλ(k)(0)/k!+1(2m1)!01Iλ(2m)(θ)(1θ)2m1𝑑θ.I_{\lambda}(1)=\sum_{k=0}^{2m-1}I^{(k)}_{\lambda}(0)/k!+\frac{1}{(2m-1)!}\int_{0}^{1}I_{\lambda}^{(2m)}(\theta)(1-\theta)^{2m-1}d\theta.

For the last term,

|Iλ(2m)(θ)|λ2m|eiλϕθ(t)g(t)2ma(t)𝑑t|\displaystyle|I_{\lambda}^{(2m)}(\theta)|\lesssim\lambda^{2m}\Big{|}\int e^{i\lambda\phi_{\theta}(t)}g(t)^{2m}a(t)dt\Big{|}
λ2mλ3m|supp a|α0=03mσ(α0(g2ma))(ϕθ)α06m+σ0(ϕθ′′)σ1(ϕθ(3mα0+1))σ3mα0\displaystyle\lesssim\lambda^{2m}\lambda^{-3m}|\text{supp }a|\sum_{\alpha_{0}=0}^{3m}\sum_{\sigma}\|(\partial^{\alpha_{0}}(g^{2m}a))(\phi_{\theta}^{\prime})^{\alpha_{0}-6m+\sigma_{0}}(\phi_{\theta}^{\prime\prime})^{\sigma_{1}}\cdot\cdot\cdot(\phi_{\theta}^{(3m-\alpha_{0}+1)})^{\sigma_{3m-\alpha_{0}}}\|_{\infty}
λm|supp a|α0=03mγ,σ(γa)gγ0(g)γ1(g(α0))γα0(ϕθ)α06m+σ0(ϕθ′′)σ1(ϕθ(3mα0+1))σ3mα0\displaystyle\lesssim\lambda^{-m}|\text{supp }a|\sum_{\alpha_{0}=0}^{3m}\sum_{\gamma,\sigma}\|(\partial^{\gamma^{\prime}}a)g^{\gamma_{0}}(g^{\prime})^{\gamma_{1}}\cdot\cdot\cdot(g^{(\alpha_{0})})^{\gamma_{\alpha_{0}}}(\phi_{\theta}^{\prime})^{\alpha_{0}-6m+\sigma_{0}}(\phi_{\theta}^{\prime\prime})^{\sigma_{1}}\cdot\cdot\cdot(\phi_{\theta}^{(3m-\alpha_{0}+1)})^{\sigma_{3m-\alpha_{0}}}\|_{\infty}
λm|supp a|α0=03mγ,σa(γ)|ϕ′′(0)|2α012m+3γ0+2γ1+γ2+2σ0+σ1ϕ′′′6mα02γ0γ1+γ3σ0+σ2\displaystyle\lesssim\lambda^{-m}|\text{supp }a|\sum_{\alpha_{0}=0}^{3m}\sum_{\gamma,\sigma}\|a^{(\gamma^{\prime})}\|_{\infty}|\phi^{\prime\prime}(0)|^{2\alpha_{0}-12m+3\gamma_{0}+2\gamma_{1}+\gamma_{2}+2\sigma_{0}+\sigma_{1}}\|\phi^{\prime\prime\prime}\|_{\infty}^{6m-\alpha_{0}-2\gamma_{0}-\gamma_{1}+\gamma_{3}-\sigma_{0}+\sigma_{2}}
×k4ϕ(k)γk+σk1\displaystyle\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\times\prod_{k\geq 4}\|\phi^{(k)}\|_{\infty}^{\gamma_{k}+\sigma_{k-1}}
λm|supp a|(a|ϕ′′(0)|3mϕ′′′2m+)\displaystyle\lesssim\lambda^{-m}|\text{supp }a|(\|a\|_{\infty}|\phi^{\prime\prime}(0)|^{-3m}\|\phi^{\prime\prime\prime}\|_{\infty}^{2m}+......)

where γ=(γ0,,γα0)\gamma=(\gamma_{0},...,\gamma_{\alpha_{0}}) satisfies γk=2m\sum\gamma_{k}=2m and γ+kγk=α0\gamma^{\prime}+\sum k\gamma_{k}=\alpha_{0}, and σ=(σ0,,σ3mα0)\sigma=(\sigma_{0},...,\sigma_{3m-\alpha_{0}}) satisfies kσk=3mα0=σk\sum k\sigma_{k}=3m-\alpha_{0}=\sum\sigma_{k}.

In particular, if |ϕ′′(0)|B1|\phi^{\prime\prime}(0)|\approx B^{-1}, ϕ′′′B2\|\phi^{\prime\prime\prime}\|_{\infty}\approx B^{-2}, and ϕ(k)B1k\|\phi^{(k)}\|_{\infty}\lesssim B^{1-k} for k4k\geq 4, then we have an elegant estimate

(2.3) |Iλ(2m)(θ)|λm|supp a|k=03ma(k)Bkm,|I_{\lambda}^{(2m)}(\theta)|\lesssim\lambda^{-m}|\text{supp }a|\sum_{k=0}^{3m}\|a^{(k)}\|_{\infty}B^{k-m},

by observing that

(2α012m+3γ0+2γ1+γ2+2σ0+σ1)2(6mα02γ0γ1+γ3σ0+σ2)\displaystyle-(2\alpha_{0}-12m+3\gamma_{0}+2\gamma_{1}+\gamma_{2}+2\sigma_{0}+\sigma_{1})-2(6m-\alpha_{0}-2\gamma_{0}-\gamma_{1}+\gamma_{3}-\sigma_{0}+\sigma_{2})
k4(k1)(γk+σk1)\displaystyle\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad-\sum_{k\geq 4}(k-1)(\gamma_{k}+\sigma_{k-1})
=k0(k1)γkk0kσk\displaystyle=-\sum_{k\geq 0}(k-1)\gamma_{k}-\sum_{k\geq 0}k\sigma_{k}
=2mα0+γ(3mα0)\displaystyle=2m-\alpha_{0}+\gamma^{\prime}-(3m-\alpha_{0})
=γm.\displaystyle=\gamma^{\prime}-m.

Fix M3m1M\geq 3m-1. For k=0,1,,2m1k=0,1,...,2m-1, let

(2.4) ak(t)=g(t)ka(t),a_{k}(t)=g(t)^{k}a(t),

where g(t)g(t) is defined in (2.1). Then ak(j)(0)=0a_{k}^{(j)}(0)=0 for j=0,1,,3k1j=0,1,...,3k-1.

Recall the Fourier transform formula for Gaussian functions (e.g. [36, Theorem 7.6.1]) for ω,t\omega,t\in\mathbb{R}

ei12ωt2=eiπ4sgnω2π|ω|eitxeix22ω𝑑x.e^{i\frac{1}{2}\omega t^{2}}=\frac{e^{\frac{i\pi}{4}{\text{sgn}}\omega}}{\sqrt{2\pi|\omega|}}\int e^{-itx}e^{-i\frac{x^{2}}{2\omega}}dx.

Then we obtain

Iλ(k)(0)\displaystyle I_{\lambda}^{(k)}(0) =(iλ)keiλ12ϕ′′(0)t2ak(t)𝑑t\displaystyle=(i\lambda)^{k}\int e^{i\lambda\frac{1}{2}\phi^{\prime\prime}(0)t^{2}}a_{k}(t)dt
=(iλ)keiπ4sgnϕ′′(0)2π|ϕ′′(0)|λei2ϕ′′(0)λξ2a^k(ξ)𝑑ξ\displaystyle=(i\lambda)^{k}\frac{e^{\frac{i\pi}{4}{\text{sgn}}\phi^{\prime\prime}(0)}}{\sqrt{2\pi|\phi^{\prime\prime}(0)|\lambda}}\int e^{-\frac{i}{2\phi^{\prime\prime}(0)\lambda}\xi^{2}}\hat{a}_{k}(\xi)d\xi
=(iλ)keiπ4sgnϕ′′(0)2π|ϕ′′(0)|λ(j=0M1(i2ϕ′′(0)λ)jξ2jj!+(i2ϕ′′(0)λ)Mξ2MM!eiη)a^k(ξ)𝑑ξ,\displaystyle=(i\lambda)^{k}\frac{e^{\frac{i\pi}{4}{\text{sgn}}\phi^{\prime\prime}(0)}}{\sqrt{2\pi|\phi^{\prime\prime}(0)|\lambda}}\int\Big{(}\sum_{j=0}^{M-1}(\frac{-i}{2\phi^{\prime\prime}(0)\lambda})^{j}\frac{\xi^{2j}}{j!}+(\frac{-i}{2\phi^{\prime\prime}(0)\lambda})^{M}\frac{\xi^{2M}}{M!}e^{i\eta}\Big{)}\hat{a}_{k}(\xi)d\xi,
=(iλ)keiπ4sgnϕ′′(0)2π|ϕ′′(0)|λ(2π)(j=3k2M1(i2ϕ′′(0)λ)jak(2j)(0)j!)+Rk\displaystyle=(i\lambda)^{k}\frac{e^{\frac{i\pi}{4}{\text{sgn}}\phi^{\prime\prime}(0)}}{\sqrt{2\pi|\phi^{\prime\prime}(0)|\lambda}}\cdot(2\pi)\Big{(}\sum_{j=\lceil\frac{3k}{2}\rceil}^{M-1}(\frac{i}{2\phi^{\prime\prime}(0)\lambda})^{j}\frac{a^{(2j)}_{k}(0)}{j!}\Big{)}+R_{k}

where η\eta is between 0 and ξ22ϕ′′(0)λ\frac{-\xi^{2}}{2\phi^{\prime\prime}(0)\lambda}. The remainder term RkR_{k} satisfies

|Rk|\displaystyle|R_{k}| λk12M|ϕ′′(0)|12M|ξ2Ma^k(ξ)|𝑑ξ.\displaystyle\lesssim\lambda^{k-\frac{1}{2}-M}|\phi^{\prime\prime}(0)|^{-\frac{1}{2}-M}\int|\xi^{2M}\hat{a}_{k}(\xi)|d\xi.

In summary, we state the explicit stationary phase lemma.

Lemma 4.

Let ϕC()\phi\in C^{\infty}(\mathbb{R}) be real-valued and satisfy ϕ(0)=ϕ(0)=0,ϕ′′(0)0\phi(0)=\phi^{\prime}(0)=0,\ \phi^{\prime\prime}(0)\neq 0. Suppose that aC0()a\in C_{0}^{\infty}(\mathbb{R}) satisfies supp a[B,B]\text{supp }a\subset[-B,B] for some B>0B>0. If |ϕ′′(0)|B1|\phi^{\prime\prime}(0)|\approx B^{-1}, ϕ′′′B2\|\phi^{\prime\prime\prime}\|_{\infty}\approx B^{-2}, and ϕ(k)B1k\|\phi^{(k)}\|_{\infty}\lesssim B^{1-k} for k4k\geq 4, then for m1m\geq 1 and M3m+1M\geq 3m+1, we have

|eiλϕ(t)a(t)𝑑tk=02m1j=3k2M1ckjλkj12|ϕ′′(0)|j12ak(2j)(0)|\displaystyle\Big{|}\int e^{i\lambda\phi(t)}a(t)dt-\sum_{k=0}^{2m-1}\sum_{j=\lceil\frac{3k}{2}\rceil}^{M-1}c_{kj}\lambda^{k-j-\frac{1}{2}}|\phi^{\prime\prime}(0)|^{-j-\frac{1}{2}}a_{k}^{(2j)}(0)\Big{|}
λm|supp a|k=03ma(k)Bkm\displaystyle\lesssim\lambda^{-m}|\text{supp }a|\sum_{k=0}^{3m}\|a^{(k)}\|_{\infty}B^{k-m}
+(λB1)12Mk=02m1λk|ξ2Ma^k(ξ)|𝑑ξ,\displaystyle\quad\quad\quad+(\lambda B^{-1})^{-\frac{1}{2}-M}\sum_{k=0}^{2m-1}\lambda^{k}\int_{\mathbb{R}}|\xi^{2M}\hat{a}_{k}(\xi)|d\xi,

where

ak(t)=(ϕ(t)12ϕ′′(0)t2)ka(t),a_{k}(t)=(\phi(t)-\tfrac{1}{2}\phi^{\prime\prime}(0)t^{2})^{k}a(t),
ckj=ik+j2πj!k!2j(sgnϕ′′(0))jeiπ4sgnϕ′′(0).c_{kj}=\frac{i^{k+j}\sqrt{2\pi}}{j!k!2^{j}}({\text{sgn}}\phi^{\prime\prime}(0))^{j}e^{\frac{i\pi}{4}{\text{sgn}}\phi^{\prime\prime}(0)}.

Here LL^{\infty} norms are taken over supp a\text{supp }a, and a^k(ξ)\hat{a}_{k}(\xi) is the Fourier transform of ak(t)a_{k}(t) on \mathbb{R}. The implicit constants are independent of λ,ϕ,a,B\lambda,\phi,a,B, and only depend on mm and MM.

The remainder term in Lemma 4 is more precise than the one presented in [36, Theorem 7.7.5]. Indeed, the proof in [36] uses Sobolev inequalities to simplify the integrals involving a^k(ξ)\hat{a}_{k}(\xi) and estimates them by the sup norms of the derivatives of aa. It does not exploit the rapid decay of the Fourier transform a^k(ξ)\hat{a}_{k}(\xi), which is crucial in our applications.

3. The spectral projection operator

In this section, we introduce the representation formula of the Hermite spectral projection operator. It is known that the kernel of the Hermite spectral projection operator PλP_{\lambda} can be represented by Mehler’s formula for the kernel of the Hermite-Schrödinger propagator eitHe^{-itH}. One may refer to Jeong-Lee-Ryu [40, Section 2.1] for a detailed introduction. Let

Pλf=2|α|+n=λ2f,hαhαP_{\lambda}f=\sum_{2|\alpha|+n=\lambda^{2}}\langle f,h_{\alpha}\rangle h_{\alpha}

where {hα}\{h_{\alpha}\} is an orthonormal basis in L2(n)L^{2}(\mathbb{R}^{n}) and α=(α1,,αn)n\alpha=(\alpha_{1},...,\alpha_{n})\in\mathbb{N}^{n}. Each hαh_{\alpha} is an eigenfunction of HH with eigenvalue λ2=2|α|+n\lambda^{2}=2|\alpha|+n, and is a tensor product of the Hermite functions on \mathbb{R}.

Since 12πππeit2s𝑑t=0\frac{1}{2\pi}\int_{-\pi}^{\pi}e^{i\frac{t}{2}s}dt=0 for s{0}s\in\mathbb{R}\setminus\{0\}, formally we have

Pλf=12πππeit2(λ2H)f𝑑t.P_{\lambda}f=\frac{1}{2\pi}\int_{-\pi}^{\pi}e^{i\frac{t}{2}(\lambda^{2}-H)}fdt.

See [40, Section 2.1] for a detailed proof. Recall the formula (see [50], [60, p. 11])

eit2Hf=cn(sint)n2neip(t,x,y)f(y)𝑑ye^{-i\frac{t}{2}H}f=c_{n}(\sin t)^{-\frac{n}{2}}\int_{\mathbb{R}^{n}}e^{ip(t,x,y)}f(y)dy

where cn=(2πi)n2einπ/4c_{n}=(2\pi i)^{-\frac{n}{2}}e^{in\pi/4}, p(t,x,y)=a2cost2b2sintp(t,x,y)=\frac{a^{2}\cos t-2b}{2\sin t} with a2=|x|2+|y|2a^{2}=|x|^{2}+|y|^{2} and b=xyb=x\cdot y. Then the kernel of PλP_{\lambda} can be written as

(3.1) K(x,y)=cnππ(sint)n2eiϕλ(t,x,y)𝑑tK(x,y)=c_{n}\int_{-\pi}^{\pi}(\sin t)^{-\frac{n}{2}}e^{i\phi_{\lambda}(t,x,y)}dt

where

ϕλ(t,x,y)=λ22t+a2cost2b2sint.\phi_{\lambda}(t,x,y)=\frac{\lambda^{2}}{2}t+\frac{a^{2}\cos t-2b}{2\sin t}.

In applications, it is more convenient to deal with the rescaled kernel with R=λ2R=\lambda^{2}

(3.2) KR(x,y)=K(λx,λy)=2cnπ/2π/2(sin2t)n2eiRψ(t,x,y)𝑑tK_{R}(x,y)=K(\lambda x,\lambda y)=2c_{n}\int_{-\pi/2}^{\pi/2}(\sin 2t)^{-\frac{n}{2}}e^{iR\psi(t,x,y)}dt

where

ψ(t,x,y)=t+a2cos2t2b2sin2t.\psi(t,x,y)=t+\frac{a^{2}\cos 2t-2b}{2\sin 2t}.

Here the square root z12=|z|ei12arg(z),arg(z)(π,π]z^{\frac{1}{2}}=\sqrt{|z|}e^{i\frac{1}{2}\arg(z)},\ \arg(z)\in(-\pi,\pi]. By inserting smooth cutoff functions, we split the integral (3.2) into four parts

(3.3) π/2π/2=03π/8+3π/80+π/2π/8+π/8π/2=I0+I0+I1+I1.\int_{-\pi/2}^{\pi/2}=\int_{0}^{3\pi/8}+\int_{-3\pi/8}^{0}+\int_{-\pi/2}^{-\pi/8}+\int_{\pi/8}^{\pi/2}=I_{0}+I_{0}^{-}+I_{1}+I_{1}^{-}.

By changing variables, we obtain

I0=03π/8(sin2t)n2ρ0(t)eiRψ(t,x,y)𝑑t,I_{0}=\int_{0}^{3\pi/8}(\sin 2t)^{-\frac{n}{2}}\rho_{0}(t)e^{iR\psi(t,x,y)}dt,
I0=einπ203π/8(sin2t)n2ρ0(t)eiRψ(t,x,y)𝑑t,I_{0}^{-}=e^{-i\frac{n\pi}{2}}\int_{0}^{3\pi/8}(\sin 2t)^{-\frac{n}{2}}\rho_{0}(t)e^{-iR\psi(t,x,y)}dt,
I1=eiπ2(n+R)03π/8(sin2t)n2ρ0(t)eiRψ(t,x,y)𝑑t,I_{1}=e^{-i\frac{\pi}{2}(n+R)}\int_{0}^{3\pi/8}(\sin 2t)^{-\frac{n}{2}}\rho_{0}(t)e^{iR\psi(t,x,-y)}dt,
I1=eiπ2R03π/8(sin2t)n2ρ0(t)eiRψ(t,x,y)𝑑t,I_{1}^{-}=e^{i\frac{\pi}{2}R}\int_{0}^{3\pi/8}(\sin 2t)^{-\frac{n}{2}}\rho_{0}(t)e^{-iR\psi(t,x,-y)}dt,

where ρ0\rho_{0} is a smooth even function supported in [3π8,3π8][-\frac{3\pi}{8},\frac{3\pi}{8}] satisfying ρ01\rho_{0}\equiv 1 near 0 and ρ00\rho_{0}\equiv 0 near ±3π8\pm\frac{3\pi}{8}, and ρ0(t)+ρ0(π2t)1\rho_{0}(t)+\rho_{0}(\frac{\pi}{2}-t)\equiv 1 for t[0,π2]t\in[0,\frac{\pi}{2}]. In the following, we mainly deal with I0I_{0}, and the other terms can be handled similarly.

4. Proof of Theorem 1

In this section, we prove Theorem 1. The proof uses the strategy developed by Thangavelu [61, 60, 62] and Jeong-Lee-Ryu [41, 40, 39, 42]. They handle the oscillatory integrals by an explicit analysis on the critical points of the phase functions. By a standard TTTT^{*} argument, we only need to prove the operator bound (4.3) associated with the rescaled kernel KR(x,y)K_{R}(x,y). Since the kernel KR(x,y)K_{R}(x,y) is represented as an oscillatory integral in Section 3, we analyze the critical points of the phase function, and then split the integral properly into several parts with respect to the critical points. For the parts away from the critical points, we can handle them using integration by parts and Young’s inequality. The crucial parts are those around the critical points. We need to use the stationary phase lemma in Section 2 to calculate the kernel explicitly and then apply Hörmander’s L2L^{2} oscillatory integral theorem (see e.g. [35], [54, Theorem 2.1.1]).

Let n1n\geq 1, w=λ1νw=\lambda^{-1}\nu, μ=max{λ43,1|w|}\mu=\max\{\lambda^{-\frac{4}{3}},1-|w|\} and R=λ2R=\lambda^{2}.

First, we only need to prove

(4.1) eλL2(B(ν,r))Cλ12r12μ14eλ2,forλ1μ12rλμ,\|e_{\lambda}\|_{L^{2}(B(\nu,r))}\leq C\lambda^{-\frac{1}{2}}r^{\frac{1}{2}}\mu^{-\frac{1}{4}}\|e_{\lambda}\|_{2},\ \ \text{for}\ \lambda^{-1}\mu^{-\frac{1}{2}}\ll r\ll\lambda\mu,

since the remaining two cases rλ1μ12r\lesssim\lambda^{-1}\mu^{-\frac{1}{2}} and λμrλ\lambda\mu\lesssim r\leq\lambda can be handled easily. Indeed, when rλ1μ12r\lesssim\lambda^{-1}\mu^{-\frac{1}{2}}, we use the LL^{\infty} bounds in (1.12) and (1.13) to obtain the desired local L2L^{2} bound (λμ12)n22rn2(\lambda\mu^{\frac{1}{2}})^{\frac{n-2}{2}}r^{\frac{n}{2}}. When λμrλ\lambda\mu\lesssim r\leq\lambda, we can cover B(ν,r)B(\nu,r) by some dyadic annuli and use (1.10) and (1.11) to get the desired bound λ14r14\lambda^{-\frac{1}{4}}r^{\frac{1}{4}}.

Clearly, the condition λ1μ12rλμ\lambda^{-1}\mu^{-\frac{1}{2}}\ll r\ll\lambda\mu in (4.1) implies that μ=1|w|λ43\mu=1-|w|\gg\lambda^{-\frac{4}{3}}.

Note that

PλL2(n)L2(B(ν,r))2=PλPλL2(B(ν,r))L2(B(ν,r))=PλL2(B(ν,r))L2(B(ν,r)).\|P_{\lambda}\|^{2}_{L^{2}(\mathbb{R}^{n})\to L^{2}(B(\nu,r))}=\|P_{\lambda}^{*}P_{\lambda}\|_{L^{2}(B(\nu,r))\to L^{2}(B(\nu,r))}=\|P_{\lambda}\|_{L^{2}(B(\nu,r))\to L^{2}(B(\nu,r))}.

Let TT be the operator associated with the rescaled kernel KR(x,y)K_{R}(x,y)

(4.2) KR(x,y)=π/2π/2(sin2t)n2eiRψ(t,x,y)𝑑tK_{R}(x,y)=\int_{-\pi/2}^{\pi/2}(\sin 2t)^{-\frac{n}{2}}e^{iR\psi(t,x,y)}dt

Note that

PλL2(B(ν,r))L2(B(ν,r))=Rn2TL2(B(w,R12r))L2(B(w,R12r)).\|P_{\lambda}\|_{L^{2}(B(\nu,r))\to L^{2}(B(\nu,r))}=R^{\frac{n}{2}}\|T\|_{L^{2}(B(w,R^{-\frac{1}{2}}r))\to L^{2}(B(w,R^{-\frac{1}{2}}r))}.

It suffices to prove

(4.3) TL2(B(w,R12r))L2(B(w,R12r))Rn2(Rμ)12r\|T\|_{L^{2}(B(w,R^{-\frac{1}{2}}r))\to L^{2}(B(w,R^{-\frac{1}{2}}r))}\lesssim R^{-\frac{n}{2}}(R\mu)^{-\frac{1}{2}}r

under the assumption that R12μ12rR12μR^{-\frac{1}{2}}\mu^{-\frac{1}{2}}\ll r\ll R^{\frac{1}{2}}\mu.

4.1. Proof of the operator bound (4.3)

Let x,yB(w,R12r)x,y\in B(w,R^{-\frac{1}{2}}r). Let D(x,y)=b2a2+1D(x,y)=b^{2}-a^{2}+1. Since μ=1|w|R12r\mu=1-|w|\gg R^{-\frac{1}{2}}r, we have

1b=2μμ2+O(R12r)1-b=2\mu-\mu^{2}+O(R^{-\frac{1}{2}}r)
D=(1b)2|xy|2=2μμ2+O(R12r).\sqrt{D}=\sqrt{(1-b)^{2}-|x-y|^{2}}=2\mu-\mu^{2}+O(R^{-\frac{1}{2}}r).

Thus, 1|x|1|y|1bDμ1-|x|\approx 1-|y|\approx 1-b\approx\sqrt{D}\approx\mu.

By (1.12) and λ|ν|=λμr\lambda-|\nu|=\lambda\mu\gg r we have

PλL1(B(ν,r))L(B(ν,r))=PλL2(n)L(B(ν,r))2(Rμ)n22.\|P_{\lambda}\|_{L^{1}(B(\nu,r))\to L^{\infty}(B(\nu,r))}=\|P_{\lambda}\|^{2}_{L^{2}(\mathbb{R}^{n})\to L^{\infty}(B(\nu,r))}\lesssim(R\mu)^{\frac{n-2}{2}}.

After rescaling, we still have the same L1LL^{1}-L^{\infty} bound

TL1(B(w,R12r)L(B(w,R12r))(Rμ)n22.\|T\|_{L^{1}(B(w,R^{-\frac{1}{2}}r)\to L^{\infty}(B(w,R^{-\frac{1}{2}}r))}\lesssim(R\mu)^{\frac{n-2}{2}}.

This operator bound implies the uniform bound of the kernel for x,yB(w,R12r)x,y\in B(w,R^{-\frac{1}{2}}r)

(4.4) |KR(x,y)|(Rμ)n22.|K_{R}(x,y)|\lesssim(R\mu)^{\frac{n-2}{2}}.

Consider the truncated kernel

(4.5) KR(x,y)(1η(|xy|/r0))K_{R}(x,y)(1-\eta(|x-y|/r_{0}))

where ηC()\eta\in C^{\infty}(\mathbb{R}) is supported in (1,)(1,\infty) and equal to 1 in (2,)(2,\infty), and

(4.6) r0=R12r(Rμr2)n12n.r_{0}=R^{-\frac{1}{2}}r(R\mu r^{2})^{-\frac{n-1}{2n}}.

By Young’s inequality, (4.4) and μR12r\mu\gg R^{-\frac{1}{2}}r, the operator associated with the kernel (4.5) has L2L2L^{2}-L^{2} norm bounded by

(4.7) (Rμ)n22r0n=Rn2(Rμ)12r,(R\mu)^{\frac{n-2}{2}}r_{0}^{n}=R^{-\frac{n}{2}}(R\mu)^{-\frac{1}{2}}r,

which is desired. So in the following we only need to handle the truncated kernel

KR(x,y)η(|xy|/r0),K_{R}(x,y)\eta(|x-y|/r_{0}),

where |xy|r0|x-y|\geq r_{0} on the support. Recall that (3.3) gives

KR(x,y)=I0+I0+I1+I1.K_{R}(x,y)=I_{0}+I_{0}^{-}+I_{1}+I_{1}^{-}.

We only handle I0I_{0}, and other terms are similar.

To analyze the oscillatory integral I0I_{0}, we calculate the derivative of the phase function

(4.8) ψ(t)sin22t=a212bcos2t+cos22t.-\psi^{\prime}(t)\sin^{2}2t=a^{2}-1-2b\cos 2t+\cos^{2}2t.

By solving ψ(t)=0\psi^{\prime}(t)=0, when bDb\geq\sqrt{D} we get two critical points t1,t2t_{1},t_{2} satisfying

cos2t1=b+D\cos 2t_{1}=b+\sqrt{D}
cos2t2=bD.\cos 2t_{2}=b-\sqrt{D}.

A simple calculation gives t1|xy|μ12t_{1}\approx|x-y|\mu^{-\frac{1}{2}}, and t2μ12t_{2}\approx\mu^{\frac{1}{2}}, see (4.25) and (4.49) for details. In the special case b<Db<\sqrt{D}, there is exactly one critical point t1t_{1}. To see this, we first notice that b+D1b+\sqrt{D}\leq 1, since a22b=|xy|20a^{2}-2b=|x-y|^{2}\geq 0. If b<0b<0, then |x|,|y||xy|R12rμ|x|,|y|\lesssim|x-y|\leq R^{-\frac{1}{2}}r\ll\mu, so we have a2μ2a^{2}\ll\mu^{2} and |b|μ2|b|\ll\mu^{2}, which gives |b|D|b|\ll\sqrt{D}. So we always have 0<b+D10<b+\sqrt{D}\leq 1. We may assume that bDb\geq\sqrt{D} in the following, and the special case is relatively easy and can be handled with obvious modifications.

A crucial observation is that 1μ11-\mu\approx 1. Indeed, if 1μ=|w|11-\mu=|w|\ll 1, then a21a^{2}\ll 1 and |b|1|b|\ll 1. So |b|D|b|\ll\sqrt{D}, which contradicts to the assumption above. Later we will use this observation in (4.57) to handle the second critical point t2t_{2}.

We can write

ψ(t)\displaystyle\psi^{\prime}(t) =1sin22t(cos2tcos2t1)(cos2tcos2t2)\displaystyle=-\frac{1}{\sin^{2}2t}(\cos 2t-\cos 2t_{1})(\cos 2t-\cos 2t_{2})
(4.9) =4sin22tsin(t+t1)sin(tt1)sin(t+t2)sin(tt2)\displaystyle=-\frac{4}{\sin^{2}2t}\sin(t+t_{1})\sin(t-t_{1})\sin(t+t_{2})\sin(t-t_{2})

Next, we split the integral I0I_{0} into the following five parts.

  1. (1)

    0<tt10<t\ll t_{1}

  2. (2)

    tt1t\approx t_{1}

  3. (3)

    t1tt2t_{1}\ll t\ll t_{2}

  4. (4)

    tt2t\approx t_{2}

  5. (5)

    t2tt_{2}\ll t

Let βC0(12,2)\beta\in C_{0}^{\infty}(\frac{1}{2},2) be a Littlewood-Paley bump function, we will use it for smooth cutoff, and for dyadic decompositions several times later.

We first handle the three parts (1),(3),(5)(1),(3),(5) using integration by parts and Young’s inequality.

Part (1): For tt1t\ll t_{1}, by (4.1) we have

(4.10) |ψ(k)(t)|t1kt12t22t1k|xy|2,k=1,2,.|\psi^{(k)}(t)|\approx t^{-1-k}t_{1}^{2}t_{2}^{2}\approx t^{-1-k}|x-y|^{2},\ k=1,2,....

We split the integral

(4.11) η(|xy|/r0)η0(t/t1)ρ0(t)(sin2t)n2eiRψ𝑑t=2>r02jt1Kj(x,y)\eta(|x-y|/r_{0})\int\eta_{0}(t/t_{1})\rho_{0}(t)(\sin 2t)^{-\frac{n}{2}}e^{iR\psi}dt=\sum_{2^{-\ell}>r_{0}}\sum_{2^{-j}\ll t_{1}}K_{j\ell}(x,y)
Kj(x,y)=β(2|xy|)β(2jt)η0(t/t1)ρ0(t)(sin2t)n2eiRψ𝑑t,K_{j\ell}(x,y)=\beta(2^{\ell}|x-y|)\int\beta(2^{j}t)\eta_{0}(t/t_{1})\rho_{0}(t)(\sin 2t)^{-\frac{n}{2}}e^{iR\psi}dt,

where η0C()\eta_{0}\in C^{\infty}(\mathbb{R}) is supported in (,34)(-\infty,\frac{3}{4}) and equal to 1 in (,12)(-\infty,\frac{1}{2}).

By (4.10), integration by parts gives

|Kj(x,y)|2n22j(R2j22)N,N.|K_{j\ell}(x,y)|\lesssim 2^{\frac{n-2}{2}j}(R2^{j}2^{-2\ell})^{-N},\ \forall N.

Then by Young’s inequality the operator TjT_{j\ell} associated with the kernel KjK_{j\ell} has L2L2L^{2}-L^{2} norm bounded by

2n22j(R2j22)N2n.2^{\frac{n-2}{2}j}(R2^{j}2^{-2\ell})^{-N}\cdot 2^{-n\ell}.

Since 2jt12μ122^{-j}\ll t_{1}\approx 2^{-\ell}\mu^{-\frac{1}{2}}, we have

2>r02jt1Tj22\displaystyle\sum_{2^{-\ell}>r_{0}}\sum_{2^{-j}\ll t_{1}}\|T_{j\ell}\|_{2\to 2} Rn2(Rμ)n22n2(Rμ12r0)Nn22+n\displaystyle\lesssim R^{-\frac{n}{2}}(R\mu)^{\frac{n-2}{2}-\frac{n}{2}}\cdot(R\mu^{\frac{1}{2}}r_{0})^{-N-\frac{n-2}{2}+n}
=Rn2(Rμ)12r(Rμr2)N112,\displaystyle=R^{-\frac{n}{2}}(R\mu)^{-\frac{1}{2}}r\cdot(R\mu r^{2})^{-N_{1}-\frac{1}{2}},

which is better than desired if NN is large enough, since Rμr21R\mu r^{2}\gtrsim 1. Here N1=12n(N+n22n)>0.N_{1}=\frac{1}{2n}(N+\tfrac{n-2}{2}-n)>0.

Part (3): For t1tt2t_{1}\ll t\ll t_{2}, by (4.1) we get

(4.12) |ψ(k)(t)|t1kt2t22t1kμ,k=1,2,.|\psi^{(k)}(t)|\approx t^{-1-k}t^{2}t_{2}^{2}\approx t^{1-k}\mu,\ k=1,2,....

We dyadically decompose the integral as

(4.13) η(|xy|/r0)η1(t)ρ0(t)(sin2t)n2eiRψ𝑑t=t12jt2Kj(x,y)\eta(|x-y|/r_{0})\int\eta_{1}(t)\rho_{0}(t)(\sin 2t)^{-\frac{n}{2}}e^{iR\psi}dt=\sum_{t_{1}\ll 2^{-j}\ll t_{2}}K_{j}(x,y)
Kj(x,y)=η(|xy|/r0)β(2jt)η1(t)ρ0(t)(sin2t)n2eiRψ𝑑t,K_{j}(x,y)=\eta(|x-y|/r_{0})\int\beta(2^{j}t)\eta_{1}(t)\rho_{0}(t)(\sin 2t)^{-\frac{n}{2}}e^{iR\psi}dt,

where η1C0(32t1,34t2)\eta_{1}\in C_{0}^{\infty}(\frac{3}{2}t_{1},\frac{3}{4}t_{2}) satisfies η11\eta_{1}\equiv 1 in (2t1,12t2)(2t_{1},\frac{1}{2}t_{2}).

By (4.12), integration by parts gives

|Kj(x,y)|2n22j(1+Rμ2j)N,N.|K_{j}(x,y)|\lesssim 2^{\frac{n-2}{2}j}(1+R\mu 2^{-j})^{-N},\ \forall N.

Since t1|xy|μ12r0μ12R1μ1t_{1}\approx|x-y|\mu^{-\frac{1}{2}}\geq r_{0}\mu^{-\frac{1}{2}}\gg R^{-1}\mu^{-1}, we get

t12jt2|Kj(x,y)|\displaystyle\sum_{t_{1}\ll 2^{-j}\ll t_{2}}|K_{j}(x,y)| t12jt22n22j(Rμ2j)N\displaystyle\lesssim\sum_{t_{1}\ll 2^{-j}\ll t_{2}}2^{\frac{n-2}{2}j}(R\mu 2^{-j})^{-N}
t1n22(Rμt1)N\displaystyle\lesssim t_{1}^{-\frac{n-2}{2}}(R\mu t_{1})^{-N}
(|xy|μ12)n22(1+Rμ12|xy|)N.\displaystyle\lesssim(|x-y|\mu^{-\frac{1}{2}})^{-\frac{n-2}{2}}(1+R\mu^{\frac{1}{2}}|x-y|)^{-N}.

By Young’s inequality, the operator associated with the kernel given by (4.13) has L2L2L^{2}-L^{2} norm bounded by

μn24n|x|n22(1+Rμ12|x|)N𝑑x\displaystyle\mu^{\frac{n-2}{4}}\int_{\mathbb{R}^{n}}|x|^{-\frac{n-2}{2}}(1+R\mu^{\frac{1}{2}}|x|)^{-N}dx Rn2(Rμ)n22n2\displaystyle\lesssim R^{-\frac{n}{2}}(R\mu)^{\frac{n-2}{2}-\frac{n}{2}}
=Rn2(Rμ)12r(Rμr2)12,\displaystyle=R^{-\frac{n}{2}}(R\mu)^{-\frac{1}{2}}r\cdot(R\mu r^{2})^{-\frac{1}{2}},

this is better than desired, since Rμr21R\mu r^{2}\gtrsim 1.

Part (5): For tt2t\gg t_{2}, by (4.1) we have

(4.14) |ψ(k)(t)|t3k,k=1,2,.|\psi^{(k)}(t)|\approx t^{3-k},\ k=1,2,....

We need to estimate the integral

(4.15) η(|xy|/r0)η2(t/t2)ρ0(t)(sin2t)n2eiRψ𝑑t\eta(|x-y|/r_{0})\int\eta_{2}(t/t_{2})\rho_{0}(t)(\sin 2t)^{-\frac{n}{2}}e^{iR\psi}dt

where η2C()\eta_{2}\in C^{\infty}(\mathbb{R}) is supported in (32,)(\frac{3}{2},\infty) and equal to 1 in (2,)(2,\infty). By (4.14) and t2μ12t_{2}\approx\mu^{\frac{1}{2}}, integration by parts gives the bound for the integral

t2n22(Rt23)Nμn24(Rμ32)N.t_{2}^{-\frac{n-2}{2}}(Rt_{2}^{3})^{-N}\approx\mu^{-\frac{n-2}{4}}(R\mu^{\frac{3}{2}})^{-N}.

Then by Young’s inequality, the operator associated with the kernel given by (4.15) has L2L2L^{2}-L^{2} norm bounded by

μn24(Rμ32)N(R12r)n\displaystyle\mu^{-\frac{n-2}{4}}(R\mu^{\frac{3}{2}})^{-N}(R^{-\frac{1}{2}}r)^{n} =Rn2(Rμ)12r(Rμ32)N+12(μr4)n14\displaystyle=R^{-\frac{n}{2}}(R\mu)^{-\frac{1}{2}}r\cdot(R\mu^{\frac{3}{2}})^{-N+\frac{1}{2}}(\mu r^{-4})^{-\frac{n-1}{4}}
(4.16) Rn2(Rμ)12r\displaystyle\lesssim R^{-\frac{n}{2}}(R\mu)^{-\frac{1}{2}}r

if NN is large enough. The last inequality uses μR12r\mu\gg R^{-\frac{1}{2}}r when Rr61Rr^{6}\gtrsim 1, and uses μr41\mu r^{-4}\gg 1 when Rr61Rr^{6}\lesssim 1.

Next, we handle the remaining two parts (2) and (4) by the explicit stationary phase lemma and Hörmander’s L2L^{2} oscillatory integral theorem.

Part (2): Contribution of the first critical point.

For tt1t\approx t_{1}, by (4.1) we obtain

(4.17) ψ′′(t1)=4Dsin2t1μt11\psi^{\prime\prime}(t_{1})=\frac{4\sqrt{D}}{\sin 2t_{1}}\approx\mu t_{1}^{-1}
(4.18) |ψ′′′(t)|μt12|\psi^{\prime\prime\prime}(t)|\approx\mu t_{1}^{-2}
(4.19) |ψ(k)(t)|μt11k,k=2,3,.|\psi^{(k)}(t)|\lesssim\mu t_{1}^{1-k},\ k=2,3,....

We need to estimate the kernel given by the oscillatory integral

(4.20) η(|xy|/r0)β(t/t1)ρ0(t)(sin2t)n2eiRψ𝑑t.\eta(|x-y|/r_{0})\int\beta(t/t_{1})\rho_{0}(t)(\sin 2t)^{-\frac{n}{2}}e^{iR\psi}dt.

Let

a(t)=β(t/t1)ρ0(t)(sin2t)n2a(t)=\beta(t/t_{1})\rho_{0}(t)(\sin 2t)^{-\frac{n}{2}}
g(t)=ψ(t)ψ(t1)12ψ′′(t1)(tt1)2g(t)=\psi(t)-\psi(t_{1})-\tfrac{1}{2}\psi^{\prime\prime}(t_{1})(t-t_{1})^{2}
ak(t)=g(t)ka(t),k=0,1,2,.a_{k}(t)=g(t)^{k}a(t),\ k=0,1,2,....

Thus, by (4.17), (4.18), (4.19) we have

|ak(j)(t)|μkt1n2+kj,j=0,1,2,|a_{k}^{(j)}(t)|\lesssim\mu^{k}t_{1}^{-\frac{n}{2}+k-j},\ j=0,1,2,...
|a^k(ξ)|μkt1n2+k+1(1+t1|ξ|)N,N.|\hat{a}_{k}(\xi)|\lesssim\mu^{k}t_{1}^{-\frac{n}{2}+k+1}(1+t_{1}|\xi|)^{-N},\ \forall N.

Then by the stationary phase Lemma 4, we have the following expansion with a remainder term estimate

|eiRψ(t)a(t)𝑑tk=02m1j=3k2M1ckjRkj12eiRψ(t1)|ψ′′(t1)|12jak(2j)(t1)|\displaystyle\Big{|}\int e^{iR\psi(t)}a(t)dt-\sum_{k=0}^{2m-1}\sum_{j=\lceil\frac{3k}{2}\rceil}^{M-1}c_{kj}R^{k-j-\frac{1}{2}}e^{iR\psi(t_{1})}|\psi^{\prime\prime}(t_{1})|^{-\frac{1}{2}-j}a_{k}^{(2j)}(t_{1})\Big{|}
(Rμt1)mt11n2+k=02m1(Rμt1)k12Mt11n2\displaystyle\lesssim(R\mu t_{1})^{-m}t_{1}^{1-\frac{n}{2}}+\sum_{k=0}^{2m-1}(R\mu t_{1})^{k-\frac{1}{2}-M}t_{1}^{1-\frac{n}{2}}
t11n2((Rμt1)m+(Rμt1)2mM32)\displaystyle\lesssim t_{1}^{1-\frac{n}{2}}((R\mu t_{1})^{-m}+(R\mu t_{1})^{2m-M-\frac{3}{2}})
t11n2(Rμt1)m\displaystyle\lesssim t_{1}^{1-\frac{n}{2}}(R\mu t_{1})^{-m}

where ckjc_{kj} are constant coefficients.

Since t1|xy|μ12t_{1}\approx|x-y|\mu^{-\frac{1}{2}}, the kernel corresponding to the remainder term is bounded by

(4.21) η(|xy|/r0)(Rμ)n22(Rμ12|xy|)mn2+1.\eta(|x-y|/r_{0})(R\mu)^{\frac{n-2}{2}}(R\mu^{\frac{1}{2}}|x-y|)^{-m-\frac{n}{2}+1}.

Then by Young’s inequality, the operator associated with (4.21) has L2L2L^{2}-L^{2} norm bounded by

(Rμ)n22|x|>r0(Rμ12|x|)mn2+1𝑑x\displaystyle(R\mu)^{\frac{n-2}{2}}\int_{|x|>r_{0}}(R\mu^{\frac{1}{2}}|x|)^{-m-\frac{n}{2}+1}dx
Rn2(Rμ)n22n2(Rμ12r0)mn22+n\displaystyle\lesssim R^{-\frac{n}{2}}(R\mu)^{\frac{n-2}{2}-\frac{n}{2}}\cdot(R\mu^{\frac{1}{2}}r_{0})^{-m-\frac{n-2}{2}+n}
=Rn2(Rμ)12r(Rμr2)N12,\displaystyle=R^{-\frac{n}{2}}(R\mu)^{-\frac{1}{2}}r\cdot(R\mu r^{2})^{-N-\frac{1}{2}},

which is better than desired if mm is large enough, since Rμr21R\mu r^{2}\gtrsim 1. Here N=12n(m+n22n)>0.N=\frac{1}{2n}(m+\tfrac{n-2}{2}-n)>0.

Since all the terms in the expansion share the same oscillatory factor, it suffices to handle the leading term

(4.22) η(|xy|/r0)R12D14(sin2t1)n12eiRψ(t1),\eta(|x-y|/r_{0})R^{-\frac{1}{2}}D^{-\frac{1}{4}}(\sin 2t_{1})^{-\frac{n-1}{2}}e^{iR\psi(t_{1})},

where we use (4.17).

Recall that μR12r\mu\gg R^{-\frac{1}{2}}r, 1|x|1|y|1bDμ1-|x|\approx 1-|y|\approx 1-b\approx\sqrt{D}\approx\mu, and

(4.23) |x,yαD|Cαμ1α.|\partial_{x,y}^{\alpha}\sqrt{D}|\leq C_{\alpha}\mu^{1-\alpha}.

Step 1 of Part (2): Analyze the first critical point.

Recall that

ψ(t,x,y)=t+a2cos2t2b2sin2t,\psi(t,x,y)=t+\frac{a^{2}\cos 2t-2b}{2\sin 2t},
cos2t1=b+D,\cos 2t_{1}=b+\sqrt{D},
(4.24) ψt(t1(x,y),x,y)0.\psi_{t}^{\prime}(t_{1}(x,y),x,y)\equiv 0.

We have

(4.25) sin22t1=a22b22bD=|xy|2(1+2b1b+D).\sin^{2}2t_{1}=a^{2}-2b^{2}-2b\sqrt{D}=|x-y|^{2}\Big{(}1+\frac{2b}{1-b+\sqrt{D}}\Big{)}.

Let

ξ(x,y)=1+2b1b+D.\xi(x,y)=\sqrt{1+\frac{2b}{1-b+\sqrt{D}}}.

Then

(4.26) sin2t1=|xy|ξ(x,y),\sin 2t_{1}=|x-y|\xi(x,y),

where ξC\xi\in C^{\infty} and

ξ(x,y)μ12\xi(x,y)\approx\mu^{-\frac{1}{2}}
(4.27) |x,yαξ(x,y)|μ12α.|\partial_{x,y}^{\alpha}\xi(x,y)|\lesssim\mu^{-\frac{1}{2}-\alpha}.

Thus, we obtain

t1(x,y)|xy|μ12t_{1}(x,y)\approx|x-y|\mu^{-\frac{1}{2}}
(4.28) |x,yαt1(x,y)||xy|1αμ12,xy.|\partial_{x,y}^{\alpha}t_{1}(x,y)|\lesssim|x-y|^{1-\alpha}\mu^{\frac{1}{2}},\ x\neq y.

Moreover,

(4.29) ψ(t1,x,y)=|xy|((2a2)t12|xy|+a2(t1tant1)2|xy|+|xy|2sin2t1)=|xy|ζ(x,y),\psi(t_{1},x,y)=|x-y|\Big{(}\frac{(2-a^{2})t_{1}}{2|x-y|}+\frac{a^{2}(t_{1}-\tan t_{1})}{2|x-y|}+\frac{|x-y|}{2\sin 2t_{1}}\Big{)}=|x-y|\zeta(x,y),

where ζC\zeta\in C^{\infty} and

ζ(x,x)=1|x|2μ\zeta(x,x)=\sqrt{1-|x|^{2}}\approx\sqrt{\mu}
(4.30) |x,yαζ(x,y)|μ12α|\partial^{\alpha}_{x,y}\zeta(x,y)|\lesssim\mu^{\frac{1}{2}-\alpha}

by (4.26) and (4.27).

Taking partial derivatives on (4.24) we get

x[ψt(t1(x,y),x,y)]=ψtt′′xt1+ψtx′′0,\partial_{x}[\psi_{t}^{\prime}(t_{1}(x,y),x,y)]=\psi_{tt}^{\prime\prime}\partial_{x}t_{1}+\psi_{tx}^{\prime\prime}\equiv 0,
x[ψ(t1(x,y),x,y)]=ψtxt1+ψx=ψx.\partial_{x}[\psi(t_{1}(x,y),x,y)]=\psi_{t}^{\prime}\partial_{x}t_{1}+\psi^{\prime}_{x}=\psi^{\prime}_{x}.

Then

(4.31) yx[ψ(t1,x,y)]=ψtx′′yt1+ψxy′′=ψtt′′xt1yt1+ψxy′′.\partial_{y}\partial_{x}[\psi(t_{1},x,y)]=\psi_{tx}^{\prime\prime}\partial_{y}t_{1}+\psi_{xy}^{\prime\prime}=-\psi_{tt}^{\prime\prime}\partial_{x}t_{1}\partial_{y}t_{1}+\psi_{xy}^{\prime\prime}.

Let InI_{n} be the n×nn\times n identity matrix. Note that

ψxy′′(t1,x,y)=1sin2t1In,\psi_{xy}^{\prime\prime}(t_{1},x,y)=-\frac{1}{\sin 2t_{1}}I_{n},
(4.32) ψtt′′(t1,x,y)=4sin32t1(a2cos2t1b(1+cos22t1))=4Dsin2t1,\psi_{tt}^{\prime\prime}(t_{1},x,y)=\frac{4}{\sin^{3}2t_{1}}(a^{2}\cos 2t_{1}-b(1+\cos^{2}2t_{1}))=\frac{4\sqrt{D}}{\sin 2t_{1}},
xt1=12sin2t1(xbyDy)=xycos2t12sin2t1D\partial_{x}t_{1}=\frac{1}{2\sin 2t_{1}}(\frac{x-by}{\sqrt{D}}-y)=\frac{x-y\cos 2t_{1}}{2\sin 2t_{1}\sqrt{D}}
yt1=12sin2t1(ybxDx)=yxcos2t12sin2t1D\partial_{y}t_{1}=\frac{1}{2\sin 2t_{1}}(\frac{y-bx}{\sqrt{D}}-x)=\frac{y-x\cos 2t_{1}}{2\sin 2t_{1}\sqrt{D}}
xt1yt1=(xycos2t1)(yxcos2t1)T4Dsin22t1\partial_{x}t_{1}\partial_{y}t_{1}=\frac{(x-y\cos 2t_{1})(y-x\cos 2t_{1})^{T}}{4D\sin^{2}2t_{1}}

Then

(4.33) yx[ψ(t1,x,y)]=(xycos2t1)(xcos2t1y)TDsin32t11sin2t1In.\partial_{y}\partial_{x}[\psi(t_{1},x,y)]=\frac{(x-y\cos 2t_{1})(x\cos 2t_{1}-y)^{T}}{\sqrt{D}\sin^{3}2t_{1}}-\frac{1}{\sin 2t_{1}}I_{n}.

Note that

(xcos2t1y)T(xycos2t1)=Dsin22t1.(x\cos 2t_{1}-y)^{T}(x-y\cos 2t_{1})=\sqrt{D}\sin^{2}2t_{1}.

Then by (4.33) the eigenvalues of the matrix yx[ψ(t1,x,y)]\partial_{y}\partial_{x}[\psi(t_{1},x,y)] are 1sin2t1-\frac{1}{\sin 2t_{1}} (multiplicity=n1=n-1) and 0.

Step 2 of Part (2): Apply oscillatory integral theorem.

Since x,yB(w,R12r)x,y\in B(w,R^{-\frac{1}{2}}r), by (4.22) we need to estimate the L2L2L^{2}-L^{2} norm of the operator TT associated with the kernel

K(x,y)=R12D14(sin2t1)n12eiRψ(t1,x,y)χ(R12r1(xw))χ(R12r1(yw))η(|xy|/r0),\displaystyle K(x,y)=R^{-\frac{1}{2}}D^{-\frac{1}{4}}(\sin 2t_{1})^{-\frac{n-1}{2}}e^{iR\psi(t_{1},x,y)}\chi(R^{\frac{1}{2}}r^{-1}(x-w))\chi(R^{\frac{1}{2}}r^{-1}(y-w))\eta(|x-y|/r_{0}),

where χC0(n)\chi\in C_{0}^{\infty}(\mathbb{R}^{n}). Since the phase function ψ(t1,x,y)\psi(t_{1},x,y) behaves like the rescaled distance function, we may use an argument similar to Burq-Gérard-Tzvetkov [17, Section 6].

We split the kernel KK and consider the operators TjT_{j} associated with the kernel

Kj(x,y)=K(x,y)β(2j|xy|),r0<2j<R12rK_{j}(x,y)=K(x,y)\beta(2^{j}|x-y|),\ \ r_{0}<2^{-j}<R^{-\frac{1}{2}}r

where r0r_{0} is given by (4.6).

Next, we introduce a partition of unity locally finite (uniformly with respect to jj):

1=pnχ(2j(xw)p),1=\sum_{p\in\mathbb{Z}^{n}}\chi(2^{j}(x-w)-p),

where χC0(n)\chi\in C_{0}^{\infty}(\mathbb{R}^{n}). Write

Kj(x,y)=q,q~nχ(2j(xw)q)Kj(x,y)χ(2j(yw)q~):=q,q~nKj,q,q~(x,y).K_{j}(x,y)=\sum_{q,\tilde{q}\in\mathbb{Z}^{n}}\chi(2^{j}(x-w)-q)K_{j}(x,y)\chi(2^{j}(y-w)-\tilde{q}):=\sum_{q,\tilde{q}\in\mathbb{Z}^{n}}K_{j,q,\tilde{q}}(x,y).

Note that on the support of χ(2j(xw)p)\chi(2^{j}(x-w)-p) we have |2j(xw)p|1|2^{j}(x-w)-p|\lesssim 1. So

|qq~||2j(xw)q|+|2j(xw)2j(yw)|+|2j(yw)q~|1|q-\tilde{q}|\leq|2^{j}(x-w)-q|+|2^{j}(x-w)-2^{j}(y-w)|+|2^{j}(y-w)-\tilde{q}|\lesssim 1

on the support of Kj,q,q~(x,y)K_{j,q,\tilde{q}}(x,y). Let Tj,q,q~T_{j,q,\tilde{q}} be the operator with kernel Kj,q,q~(x,y)K_{j,q,\tilde{q}}(x,y). Then we need to handle

Tj=q,q~Tj,q,q~.T_{j}=\sum_{q,\tilde{q}}T_{j,q,\tilde{q}}.

Note that the kernel of Tj,p,p~Tj,q,q~T_{j,p,\tilde{p}}^{*}T_{j,q,\tilde{q}}

χ(2j(xw)p)Kj(x,z)¯χ(2j(zw)p~)χ(2j(zw)q)Kj(z,y)χ(2j(yw)q~)𝑑z\int\chi(2^{j}(x-w)-p)\overline{K_{j}(x,z)}\chi(2^{j}(z-w)-\tilde{p})\chi(2^{j}(z-w)-q)K_{j}(z,y)\chi(2^{j}(y-w)-\tilde{q})dz

vanishes when |(p,p~)(q,q~)|1|(p,\tilde{p})-(q,\tilde{q})|\gtrsim 1. Thus, Cotlar–Stein lemma ([57, Chapter VII]) implies

(4.34) TjL2L2supp,p~q,q~Tj,p,p~Tj,q,q~L2L2supq,q~Tj,q,q~L2L2.\|T_{j}\|_{L^{2}\to L^{2}}\leq\sup_{p,\tilde{p}}\sum_{q,\tilde{q}}\sqrt{\|T_{j,p,\tilde{p}}^{*}T_{j,q,\tilde{q}}\|_{L^{2}\to L^{2}}}\lesssim\sup_{q,\tilde{q}}\|T_{j,q,\tilde{q}}\|_{L^{2}\to L^{2}}.

To estimate Tj,q,q~L2L2\|T_{j,q,\tilde{q}}\|_{L^{2}\to L^{2}}, we consider the kernel of Tj,q,q~T_{j,q,\tilde{q}}

Kj,q,q~(x,y)=χ(2j(xw)q)Kj(x,y)χ(2j(yw)q~).K_{j,q,\tilde{q}}(x,y)=\chi(2^{j}(x-w)-q)K_{j}(x,y)\chi(2^{j}(y-w)-\tilde{q}).

Let X=2j(xw)X=2^{j}(x-w), Y=2j(yw)Y=2^{j}(y-w). Let T~j,q,q~\tilde{T}_{j,q,\tilde{q}} be the operator associated with the rescaled kernel

K~j,q,q~(X,Y)=χ(Xq)Kj(2jX+w,2jY+w)χ(Yq~)\displaystyle\tilde{K}_{j,q,\tilde{q}}(X,Y)=\chi(X-q)K_{j}(2^{-j}X+w,2^{-j}Y+w)\chi(Y-\tilde{q})
=χ(Xq)χ(Yq~)β(|XY|)K(2jX+w,2jY+w)\displaystyle=\chi(X-q)\chi(Y-\tilde{q})\beta(|X-Y|)K(2^{-j}X+w,2^{-j}Y+w)
=χ(Xq)χ(Yq~)χ(R12r12jX)χ(R12r12jY)β(|XY|)eiRψj(X,Y)(sin2τj)n12R12Dj14\displaystyle=\chi(X-q)\chi(Y-\tilde{q})\chi(R^{\frac{1}{2}}r^{-1}2^{-j}X)\chi(R^{\frac{1}{2}}r^{-1}2^{-j}Y)\beta(|X-Y|)e^{iR\psi_{j}(X,Y)}(\sin 2\tau_{j})^{-\frac{n-1}{2}}R^{-\frac{1}{2}}D_{j}^{-\frac{1}{4}}

where

Dj(X,Y)=D(2jX+w,2jY+w)D_{j}(X,Y)=D(2^{-j}X+w,2^{-j}Y+w)
τj(X,Y)=t1(2jX+w,2jY+w)\tau_{j}(X,Y)=t_{1}(2^{-j}X+w,2^{-j}Y+w)
ψj(X,Y)=ψ(τj,2jX+w,2jY+w).\psi_{j}(X,Y)=\psi(\tau_{j},2^{-j}X+w,2^{-j}Y+w).

Since |XY|1|X-Y|\approx 1 on the support of β(|XY|)\beta(|X-Y|), by (4.26), (4.27), (4.29) and (4.30) we have

(4.35) Dj(X,Y)1|w|2μ\sqrt{D_{j}(X,Y)}\to 1-|w|^{2}\approx\mu
(4.36) 2jsin(2τj(X,Y))|XY|1+|w|22(1|w|2)|XY|μ122^{j}\sin(2\tau_{j}(X,Y))\to|X-Y|\sqrt{1+\frac{|w|^{2}}{2(1-|w|^{2})}}\approx|X-Y|\mu^{-\frac{1}{2}}
(4.37) 2jψj(X,Y)|XY|1|w|2|XY|μ122^{j}\psi_{j}(X,Y)\to|X-Y|\sqrt{1-|w|^{2}}\approx|X-Y|\mu^{\frac{1}{2}}

in CC^{\infty} topology as jj\to\infty. So for large jj we can write the kernel as

K~j,q,q~(X,Y)=2n12jμn34R12\displaystyle\tilde{K}_{j,q,\tilde{q}}(X,Y)=2^{\frac{n-1}{2}j}\mu^{\frac{n-3}{4}}R^{-\frac{1}{2}} eiRψj(X,Y)Aj(w,X,Y)\displaystyle e^{iR\psi_{j}(X,Y)}A_{j}(w,X,Y)
χ(Xq)χ(Yq~)χ(R12r12jX)χ(R12r12jY)β(|XY|)\displaystyle\cdot\chi(X-q)\chi(Y-\tilde{q})\chi(R^{\frac{1}{2}}r^{-1}2^{-j}X)\chi(R^{\frac{1}{2}}r^{-1}2^{-j}Y)\beta(|X-Y|)

where Aj(w,X,Y)A_{j}(w,X,Y) is a smooth function on the support of the kernel and satisfies

(4.38) |X,YαAj(w,X,Y)|1.|\partial^{\alpha}_{X,Y}A_{j}(w,X,Y)|\lesssim 1.

The implicit constant is independent of j,μ,w,q,q~,R,X,Yj,\mu,w,q,\tilde{q},R,X,Y.

By Hörmander’s oscillatory integral theorem, we have

T~j,q,q~L2L22n12jR12μn34(R2jμ12)n12.\|\tilde{T}_{j,q,\tilde{q}}\|_{L^{2}\to L^{2}}\lesssim 2^{\frac{n-1}{2}j}R^{-\frac{1}{2}}\mu^{\frac{n-3}{4}}(R2^{-j}\mu^{\frac{1}{2}})^{-\frac{n-1}{2}}.

By rescaling,

Tj,q,q~L2L2=2njT~j,q,q~L2L2Rn2μ122j.\|T_{j,q,\tilde{q}}\|_{L^{2}\to L^{2}}=2^{-nj}\|\tilde{T}_{j,q,\tilde{q}}\|_{L^{2}\to L^{2}}\lesssim R^{-\frac{n}{2}}\mu^{-\frac{1}{2}}2^{-j}.

By (4.34), we get

TjL2L2Rn2μ122j.\|T_{j}\|_{L^{2}\to L^{2}}\lesssim R^{-\frac{n}{2}}\mu^{-\frac{1}{2}}2^{-j}.

Then

2j<R1/2rTj22Rn2(Rμ)12r,\sum_{2^{-j}<R^{-1/2}r}\|T_{j}\|_{2\to 2}\lesssim R^{-\frac{n}{2}}(R\mu)^{-\frac{1}{2}}r,

which is expected.

Part (4): Contribution of the second critical point.

For tt2t\approx t_{2}, by (4.1) we obtain

(4.39) |ψ′′(t2)|=4Dsin2t2μ12|\psi^{\prime\prime}(t_{2})|=\frac{4\sqrt{D}}{\sin 2t_{2}}\approx\mu^{\frac{1}{2}}
(4.40) |ψ′′′(t)|1|\psi^{\prime\prime\prime}(t)|\approx 1
(4.41) |ψ(k)(t)|μt21kμ3k2,k=2,3,.|\psi^{(k)}(t)|\lesssim\mu t_{2}^{1-k}\lesssim\mu^{\frac{3-k}{2}},\ k=2,3,....

We need to estimate the kernel given by the oscillatory integral

(4.42) η(|xy|/r0)β(t/t2)ρ0(t)(sin2t)n2eiRψ𝑑t.\eta(|x-y|/r_{0})\int\beta(t/t_{2})\rho_{0}(t)(\sin 2t)^{-\frac{n}{2}}e^{iR\psi}dt.

Let

a(t)=β(t/t2)ρ0(t)(sin2t)n2a(t)=\beta(t/t_{2})\rho_{0}(t)(\sin 2t)^{-\frac{n}{2}}
g(t)=ψ(t)ψ(t2)12ψ′′(t2)(tt2)2g(t)=\psi(t)-\psi(t_{2})-\tfrac{1}{2}\psi^{\prime\prime}(t_{2})(t-t_{2})^{2}
ak(t)=g(t)ka(t),k=0,1,2,.a_{k}(t)=g(t)^{k}a(t),\ k=0,1,2,....

Thus, by (4.39), (4.40), (4.41) we have

|ak(j)(t)|μkt2n2+kj,j=0,1,2,|a_{k}^{(j)}(t)|\lesssim\mu^{k}t_{2}^{-\frac{n}{2}+k-j},\ j=0,1,2,...
|a^k(ξ)|μkt2n2+k+1(1+t2|ξ|)N,N.|\hat{a}_{k}(\xi)|\lesssim\mu^{k}t_{2}^{-\frac{n}{2}+k+1}(1+t_{2}|\xi|)^{-N},\ \forall N.

Then by the stationary phase Lemma 4, we have the following expansion with a remainder term estimate

|eiRψ(t)a(t)𝑑tk=02m1j=3k2M1ckjRkj12eiRψ(t2)|ψ′′(t2)|12jak(2j)(t2)|\displaystyle\Big{|}\int e^{iR\psi(t)}a(t)dt-\sum_{k=0}^{2m-1}\sum_{j=\lceil\frac{3k}{2}\rceil}^{M-1}c_{kj}R^{k-j-\frac{1}{2}}e^{iR\psi(t_{2})}|\psi^{\prime\prime}(t_{2})|^{-\frac{1}{2}-j}a_{k}^{(2j)}(t_{2})\Big{|}
(Rμt2)mt21n2+k=02m1(Rμt2)k12Mt21n2\displaystyle\lesssim(R\mu t_{2})^{-m}t_{2}^{1-\frac{n}{2}}+\sum_{k=0}^{2m-1}(R\mu t_{2})^{k-\frac{1}{2}-M}t_{2}^{1-\frac{n}{2}}
t21n2((Rμt2)m+(Rμt2)2mM32)\displaystyle\lesssim t_{2}^{1-\frac{n}{2}}((R\mu t_{2})^{-m}+(R\mu t_{2})^{2m-M-\frac{3}{2}})
t21n2(Rμt2)m\displaystyle\lesssim t_{2}^{1-\frac{n}{2}}(R\mu t_{2})^{-m}

where ckjc_{kj} are constant coefficients.

Since t2μ12t_{2}\approx\mu^{\frac{1}{2}}, the kernel corresponding to the remainder term is bounded by

(4.43) μ2n4(Rμ32)m.\mu^{\frac{2-n}{4}}(R\mu^{\frac{3}{2}})^{-m}.

Then by Young’s inequality, the operator associated with (4.43) has L2L2L^{2}-L^{2} norm bounded by

μ2n4(Rμ32)m|x|R1/2r𝑑x\displaystyle\mu^{\frac{2-n}{4}}(R\mu^{\frac{3}{2}})^{-m}\int_{|x|\leq R^{-1/2}r}dx Rn2(Rμ32)mμ2n4rn\displaystyle\lesssim R^{-\frac{n}{2}}(R\mu^{\frac{3}{2}})^{-m}\mu^{\frac{2-n}{4}}r^{n}
=Rn2(Rμ)12r(Rμ32)m+12(μr4)n14\displaystyle=R^{-\frac{n}{2}}(R\mu)^{-\frac{1}{2}}r\cdot(R\mu^{\frac{3}{2}})^{-m+\frac{1}{2}}(\mu r^{-4})^{-\frac{n-1}{4}}
(4.44) Rn2(Rμ)12r\displaystyle\lesssim R^{-\frac{n}{2}}(R\mu)^{-\frac{1}{2}}r

if mm is large enough. The last inequality uses μR12r\mu\gg R^{-\frac{1}{2}}r when Rr61Rr^{6}\gtrsim 1, and uses μr41\mu r^{-4}\gg 1 when Rr61Rr^{6}\lesssim 1.

Since all the terms in the expansion share the same oscillatory factor, it suffices to handle the leading term

(4.45) η(|xy|/r0)R12D14(sin2t2)n12eiRψ(t2),\eta(|x-y|/r_{0})R^{-\frac{1}{2}}D^{-\frac{1}{4}}(\sin 2t_{2})^{-\frac{n-1}{2}}e^{iR\psi(t_{2})},

where we use (4.39). We split (4.45) into two parts

K1(x,y)=(1η(|xy|/r0))R12D14(sin2t2)n12eiRψ(t2)K_{1}(x,y)=(1-\eta(|x-y|/r_{0}))R^{-\frac{1}{2}}D^{-\frac{1}{4}}(\sin 2t_{2})^{-\frac{n-1}{2}}e^{iR\psi(t_{2})}
K(x,y)=R12D14(sin2t2)n12eiRψ(t2).K(x,y)=R^{-\frac{1}{2}}D^{-\frac{1}{4}}(\sin 2t_{2})^{-\frac{n-1}{2}}e^{iR\psi(t_{2})}.

We can handle K1K_{1} by Young’s inequality, since it is bounded and has small support. Indeed, the operator associated with K1K_{1} has L2L2L^{2}-L^{2} norm bounded by

R12μn+14r02n/p\displaystyle R^{-\frac{1}{2}}\mu^{-\frac{n+1}{4}}r_{0}^{2n/p} =(Rμ)n22(Rμ32)n12r0n\displaystyle=(R\mu)^{\frac{n-2}{2}}(R\mu^{\frac{3}{2}})^{-\frac{n-1}{2}}r_{0}^{n}
(4.46) (Rμ)n22r0n\displaystyle\lesssim(R\mu)^{\frac{n-2}{2}}r_{0}^{n}
=Rn2(Rμ)12r.\displaystyle=R^{-\frac{n}{2}}(R\mu)^{-\frac{1}{2}}r.

Here we use (4.7) and Rμ321R\mu^{\frac{3}{2}}\gtrsim 1. So it remains to handle KK.

Step 1 of Part (4): Analyze the second critical point.

The calculation is similar to Step 1 of Part (2), but the behavior of the phase function associated with the second critical point are very different. So we still provide all the details. Recall that μR12r\mu\gg R^{-\frac{1}{2}}r, 1|x|1|y|1bDμ1-|x|\approx 1-|y|\approx 1-b\approx\sqrt{D}\approx\mu and

(4.47) |x,yαD|Cαμ1α.|\partial_{x,y}^{\alpha}\sqrt{D}|\leq C_{\alpha}\mu^{1-\alpha}.

Moreover,

ψ(t,x,y)=t+a2cos2t2b2sin2t,\psi(t,x,y)=t+\frac{a^{2}\cos 2t-2b}{2\sin 2t},
cos2t2=bD,\cos 2t_{2}=b-\sqrt{D},
(4.48) ψt(t2(x,y),x,y)0.\psi_{t}^{\prime}(t_{2}(x,y),x,y)\equiv 0.

Note that

(4.49) sin22t2=a22b2+2bD=|xy|2+2b(1b+D).\sin^{2}2t_{2}=a^{2}-2b^{2}+2b\sqrt{D}=|x-y|^{2}+2b(1-b+\sqrt{D}).

Thus, we get

t2(x,y)μ12t_{2}(x,y)\approx\mu^{\frac{1}{2}}
(4.50) |x,yαt2(x,y)|μ12α.|\partial_{x,y}^{\alpha}t_{2}(x,y)|\lesssim\mu^{\frac{1}{2}-\alpha}.

We may write

(4.51) ψ(t2,x,y)=12(2a2)t2+a22(t2tant2)+|xy|22sin2t2,\psi(t_{2},x,y)=\frac{1}{2}(2-a^{2})t_{2}+\frac{a^{2}}{2}(t_{2}-\tan t_{2})+\frac{|x-y|^{2}}{2\sin 2t_{2}},

then we obtain

(4.52) |x,yαψ(t2(x,y),x,y)|μ32α.|\partial_{x,y}^{\alpha}\psi(t_{2}(x,y),x,y)|\lesssim\mu^{\frac{3}{2}-\alpha}.

Taking partial derivatives on (4.48) we get

x[ψt(t2(x,y),x,y)]=ψtt′′xt2+ψtx′′0,\partial_{x}[\psi_{t}^{\prime}(t_{2}(x,y),x,y)]=\psi_{tt}^{\prime\prime}\partial_{x}t_{2}+\psi_{tx}^{\prime\prime}\equiv 0,
x[ψ(t2(x,y),x,y)]=ψtxt2+ψx=ψx.\partial_{x}[\psi(t_{2}(x,y),x,y)]=\psi_{t}^{\prime}\partial_{x}t_{2}+\psi^{\prime}_{x}=\psi^{\prime}_{x}.

Then

(4.53) yx[ψ(t2,x,y)]=ψtx′′yt2+ψxy′′=ψtt′′xt2yt2+ψxy′′.\partial_{y}\partial_{x}[\psi(t_{2},x,y)]=\psi_{tx}^{\prime\prime}\partial_{y}t_{2}+\psi_{xy}^{\prime\prime}=-\psi_{tt}^{\prime\prime}\partial_{x}t_{2}\partial_{y}t_{2}+\psi_{xy}^{\prime\prime}.

Note that

ψxy′′(t2,x,y)=1sin2t2In,\psi_{xy}^{\prime\prime}(t_{2},x,y)=-\frac{1}{\sin 2t_{2}}I_{n},
(4.54) ψtt′′(t2,x,y)=4sin32t2(a2cos2t2b(1+cos22t2))=4Dsin2t2,\psi_{tt}^{\prime\prime}(t_{2},x,y)=\frac{4}{\sin^{3}2t_{2}}(a^{2}\cos 2t_{2}-b(1+\cos^{2}2t_{2}))=-\frac{4\sqrt{D}}{\sin 2t_{2}},
xt2=12sin2t2(byxDy)=ycos2t2x2sin2t2D\partial_{x}t_{2}=\frac{1}{2\sin 2t_{2}}(\frac{by-x}{\sqrt{D}}-y)=\frac{y\cos 2t_{2}-x}{2\sin 2t_{2}\sqrt{D}}
yt2=12sin2t2(bxyDx)=xcos2t2y2sin2t2D\partial_{y}t_{2}=\frac{1}{2\sin 2t_{2}}(\frac{bx-y}{\sqrt{D}}-x)=\frac{x\cos 2t_{2}-y}{2\sin 2t_{2}\sqrt{D}}
xt2yt2=(ycos2t2x)(xcos2t2y)T4Dsin22t2.\partial_{x}t_{2}\partial_{y}t_{2}=\frac{(y\cos 2t_{2}-x)(x\cos 2t_{2}-y)^{T}}{4D\sin^{2}2t_{2}}.

Then

(4.55) yx[ψ(t2,x,y)]=(ycos2t2x)(xcos2t2y)TDsin32t21sin2t2In.\partial_{y}\partial_{x}[\psi(t_{2},x,y)]=\frac{(y\cos 2t_{2}-x)(x\cos 2t_{2}-y)^{T}}{\sqrt{D}\sin^{3}2t_{2}}-\frac{1}{\sin 2t_{2}}I_{n}.

Note that

(xcos2t2y)T(ycos2t2x)=Dsin22t2.(x\cos 2t_{2}-y)^{T}(y\cos 2t_{2}-x)=\sqrt{D}\sin^{2}2t_{2}.

Then by (4.55) the eigenvalues of the matrix yx[ψ(t2,x,y)]\partial_{y}\partial_{x}[\psi(t_{2},x,y)] are 1sin2t2-\frac{1}{\sin 2t_{2}} (multiplicity=n1=n-1) and 0.

Step 2 of Part (4): Apply oscillatory integral theorem.

By rotational symmetry, we may assume that w=(1μ,0,,0)w=(1-\mu,0,...,0). Since x,yB(w,R12r)B(w,μ)x,y\in B(w,R^{-\frac{1}{2}}r)\subset B(w,\mu), we consider the operator TT with the kernel

(4.56) K(x,y)=R12D14(sin2t2)n12eiRψ(t2,x,y)χ((xw)/μ)χ((yw)/μ).K(x,y)=R^{-\frac{1}{2}}D^{-\frac{1}{4}}(\sin 2t_{2})^{-\frac{n-1}{2}}e^{iR\psi(t_{2},x,y)}\chi((x-w)/\mu)\chi((y-w)/\mu).

Here χC0(n)\chi\in C_{0}^{\infty}(\mathbb{R}^{n}). Unlike the first critical point, the phase function ψ(t2,x,y)\psi(t_{2},x,y) does not behave like the distance function |xy||x-y|, so we need to use a different argument to handle the associated oscillatory integral operator.

We first simplify the matrix in (4.55). Note that x=w+O(R12r)x=w+O(R^{-\frac{1}{2}}r), y=w+O(R12r)y=w+O(R^{-\frac{1}{2}}r). The big oh notations are in the sense of some matrix norm. We have

(ycos2t2x)(xcos2t2y)T=wwT(1cos2t2)2+O(R12r),(y\cos 2t_{2}-x)(x\cos 2t_{2}-y)^{T}=ww^{T}(1-\cos 2t_{2})^{2}+O(R^{-\frac{1}{2}}r),

where wwT=diag((1μ)2,0,,0)ww^{T}=diag((1-\mu)^{2},0,...,0). Moreover,

1b=2μμ2+O(R12r)1-b=2\mu-\mu^{2}+O(R^{-\frac{1}{2}}r)
D=(1b)2|xy|2=2μμ2+O(R12r)\sqrt{D}=\sqrt{(1-b)^{2}-|x-y|^{2}}=2\mu-\mu^{2}+O(R^{-\frac{1}{2}}r)
(1cos2t2)2Dsin22t2\displaystyle\frac{(1-\cos 2t_{2})^{2}}{\sqrt{D}\sin^{2}2t_{2}} =(1b+D)2D(1(bD)2)=1b+DD(2(1b)D)\displaystyle=\frac{(1-b+\sqrt{D})^{2}}{\sqrt{D}(1-(b-\sqrt{D})^{2})}=\frac{1-b+\sqrt{D}}{\sqrt{D}(2-(1-b)-\sqrt{D})}
=2(2μμ2)+O(R12r)(2μμ2+O(R12r))(22(2μμ2)+O(R12r))\displaystyle=\frac{2(2\mu-\mu^{2})+O(R^{-\frac{1}{2}}r)}{(2\mu-\mu^{2}+O(R^{-\frac{1}{2}}r))(2-2(2\mu-\mu^{2})+O(R^{-\frac{1}{2}}r))}
(4.57) =(1μ)2+O(μ1R12r).\displaystyle=(1-\mu)^{-2}+O(\mu^{-1}R^{-\frac{1}{2}}r).

Here we use the fact that 1μ11-\mu\approx 1. See the observation before (4.1).

Thus,

yx[ψ(t2(x,y),x,y)]\displaystyle\partial_{y}\partial_{x}[\psi(t_{2}(x,y),x,y)] =(ycos2t2x)(xcos2t2y)TDsin32t21sin2t2In\displaystyle=\frac{(y\cos 2t_{2}-x)(x\cos 2t_{2}-y)^{T}}{\sqrt{D}\sin^{3}2t_{2}}-\frac{1}{\sin 2t_{2}}I_{n}
=1sin2t2(wwT((1μ)2+O(μ1R12r))In)\displaystyle=\frac{1}{\sin 2t_{2}}\Big{(}ww^{T}((1-\mu)^{-2}+O(\mu^{-1}R^{-\frac{1}{2}}r))-I_{n}\Big{)}
=1sin2t2((1μ)2wwTIn)+O(μ32R12r)\displaystyle=\frac{1}{\sin 2t_{2}}((1-\mu)^{-2}ww^{T}-I_{n})+O(\mu^{-\frac{3}{2}}R^{-\frac{1}{2}}r)
=1sin2t2diag(0,1,,1)+O(μ32R12r).\displaystyle=-\frac{1}{\sin 2t_{2}}diag(0,1,...,1)+O(\mu^{-\frac{3}{2}}R^{-\frac{1}{2}}r).

Let x=(x1,x),y=(y1,y)x=(x_{1},x^{\prime}),\ y=(y_{1},y^{\prime}). Then

yx[ψ(t2(x,y),x,y)]=1sin2t2In1+O(μ32R12r).\partial_{y^{\prime}}\partial_{x^{\prime}}[\psi(t_{2}(x,y),x,y)]=-\frac{1}{\sin 2t_{2}}I_{n-1}+O(\mu^{-\frac{3}{2}}R^{-\frac{1}{2}}r).

This implies

|detyx[ψ(t2(x,y),x,y)]|(μ12)n1,|\det\partial_{y^{\prime}}\partial_{x^{\prime}}[\psi(t_{2}(x,y),x,y)]|\approx(\mu^{-\frac{1}{2}})^{n-1},

as μR12r\mu\gg R^{-\frac{1}{2}}r.

Recall the derivatives estimates in (4.23), (4.50), (4.52)

|x,yαD(x,y)|μ1α|\partial^{\alpha}_{x,y}\sqrt{D(x,y)}|\lesssim\mu^{1-\alpha}
|x,yαt2(x,y)|μ12α|\partial_{x,y}^{\alpha}t_{2}(x,y)|\lesssim\mu^{\frac{1}{2}-\alpha}
|x,yαψ(t2(x,y),x,y)|μ32α.|\partial_{x,y}^{\alpha}\psi(t_{2}(x,y),x,y)|\lesssim\mu^{\frac{3}{2}-\alpha}.

Consider the operator TμT_{\mu} with the rescaled kernel

K(μX,μY)\displaystyle K(\mu X,\mu Y) =R12D(μX,μY)14eiRψ(t2(μX,μY),μX,μY)(sin2t2(μX,μY))n12\displaystyle=R^{-\frac{1}{2}}D(\mu X,\mu Y)^{-\frac{1}{4}}e^{iR\psi(t_{2}(\mu X,\mu Y),\mu X,\mu Y)}(\sin 2t_{2}(\mu X,\mu Y))^{-\frac{n-1}{2}}
=R12μn+14eiRμ32Φ(μ,X,Y)A(μ,X,Y)χ(Xw/μ)χ(Yw/μ)\displaystyle=R^{-\frac{1}{2}}\mu^{-\frac{n+1}{4}}e^{iR\mu^{\frac{3}{2}}\Phi(\mu,X,Y)}A(\mu,X,Y)\chi(X-w/\mu)\chi(Y-w/\mu)

where

|detXYΦ(μ,X,Y)|1,|\det\partial_{X^{\prime}}\partial_{Y^{\prime}}\Phi(\mu,X,Y)|\approx 1,
|X,YαΦ(μ,X,Y)|1|\partial_{X,Y}^{\alpha}\Phi(\mu,X,Y)|\lesssim 1
|X,YαA(μ,X,Y)|1.|\partial_{X,Y}^{\alpha}A(\mu,X,Y)|\lesssim 1.

These bounds follow from rescaling the derivatives estimates above, and the implicit constants are independent of μ,w,R,X,Y\mu,w,R,X,Y.

For fixed X1,Y1X_{1},Y_{1}, we define

Tμ,X1,Y1g(X)=K(μX,μY)g(Y)𝑑Y.T_{\mu,X_{1},Y_{1}}g(X)=\int K(\mu X,\mu Y)g(Y^{\prime})dY^{\prime}.

Suppose that supp fB(w/μ,R12r/μ)B(w/μ,1)f\subset B(w/\mu,R^{-\frac{1}{2}}r/\mu)\subset B(w/\mu,1). Then by Minkovski and Hölder inequalities,

TμfL2(B(w/μ,R12r/μ)2=|X11+μ|<R12r|Y11+μ|<R12rTμ,X1,Y1(f(Y1,))𝑑Y1L2(n1)2𝑑X1\displaystyle\|T_{\mu}f\|_{L^{2}(B(w/\mu,R^{-\frac{1}{2}}r/\mu)}^{2}=\int_{|X_{1}-1+\mu|<R^{-\frac{1}{2}}r}\|\int_{|Y_{1}-1+\mu|<R^{-\frac{1}{2}}r}T_{\mu,X_{1},Y_{1}}(f(Y_{1},\cdot))dY_{1}\|_{L^{2}(\mathbb{R}^{n-1})}^{2}dX_{1}
|X11+μ|<R12r(|Y11+μ|<R12rTμ,X1,Y1(f(Y1,))L2(n1)𝑑Y1)2𝑑X1\displaystyle\leq\int_{|X_{1}-1+\mu|<R^{-\frac{1}{2}}r}\Big{(}\int_{|Y_{1}-1+\mu|<R^{-\frac{1}{2}}r}\|T_{\mu,X_{1},Y_{1}}(f(Y_{1},\cdot))\|_{L^{2}(\mathbb{R}^{n-1})}dY_{1}\Big{)}^{2}dX_{1}
R12rμ1|X11+μ|<R12r|Y11+μ|<R12rTμ,X1,Y1(f(Y1,))L2(n1)2𝑑Y1𝑑X1\displaystyle\lesssim R^{-\frac{1}{2}}r\mu^{-1}\int_{|X_{1}-1+\mu|<R^{-\frac{1}{2}}r}\int_{|Y_{1}-1+\mu|<R^{-\frac{1}{2}}r}\|T_{\mu,X_{1},Y_{1}}(f(Y_{1},\cdot))\|_{L^{2}(\mathbb{R}^{n-1})}^{2}dY_{1}dX_{1}
R12rμ1|X11+μ|<R12rTμ,X1,Y1L2(n1)L2(n1)2f(Y1,)L2(n1)2𝑑Y1𝑑X1\displaystyle\leq R^{-\frac{1}{2}}r\mu^{-1}\int_{|X_{1}-1+\mu|<R^{-\frac{1}{2}}r}\int\|T_{\mu,X_{1},Y_{1}}\|_{L^{2}(\mathbb{R}^{n-1})\to L^{2}(\mathbb{R}^{n-1})}^{2}\|f(Y_{1},\cdot)\|_{L^{2}(\mathbb{R}^{n-1})}^{2}dY_{1}dX_{1}
R1r2μ2supX1,Y1Tμ,X1,Y1L2(n1)L2(n1)2fL2(B(w/μ,R12r/μ))2.\displaystyle\lesssim R^{-1}r^{2}\mu^{-2}\sup_{X_{1},Y_{1}}\|T_{\mu,X_{1},Y_{1}}\|_{L^{2}(\mathbb{R}^{n-1})\to L^{2}(\mathbb{R}^{n-1})}^{2}\|f\|_{L^{2}(B(w/\mu,R^{-\frac{1}{2}}r/\mu))}^{2}.

Thus, by Hörmander’s L2L^{2} oscillatory integral theorem we have

TμL2(B(w/μ,R12r/μ))L2(B(w/μ,R12r/μ))\displaystyle\|T_{\mu}\|_{L^{2}(B(w/\mu,R^{-\frac{1}{2}}r/\mu))\to L^{2}(B(w/\mu,R^{-\frac{1}{2}}r/\mu))} R12rμ1supX1,Y1Tμ,X1,Y1L2(n1)L2(n1)\displaystyle\lesssim R^{-\frac{1}{2}}r\mu^{-1}\sup_{X_{1},Y_{1}}\|T_{\mu,X_{1},Y_{1}}\|_{L^{2}(\mathbb{R}^{n-1})\to L^{2}(\mathbb{R}^{n-1})}
R12rμ1(Rμ32)n12R12μn+14\displaystyle\lesssim R^{-\frac{1}{2}}r\mu^{-1}\cdot(R\mu^{\frac{3}{2}})^{-\frac{n-1}{2}}\cdot R^{-\frac{1}{2}}\mu^{-\frac{n+1}{4}}
=Rn+12μn12r.\displaystyle=R^{-\frac{n+1}{2}}\mu^{-n-\frac{1}{2}}r.

By rescaling,

TL2(B(w,R12r))L2(B(w,R12r))\displaystyle\|T\|_{L^{2}(B(w,R^{-\frac{1}{2}}r))\to L^{2}(B(w,R^{-\frac{1}{2}}r))} =μnTμL2(B(w/μ,R12/μ))L2(B(w/μ,R12/μ))\displaystyle=\mu^{n}\|T_{\mu}\|_{L^{2}(B(w/\mu,R^{-\frac{1}{2}}/\mu))\to L^{2}(B(w/\mu,R^{-\frac{1}{2}}/\mu))}
Rn2(Rμ)12r\displaystyle\lesssim R^{-\frac{n}{2}}(R\mu)^{-\frac{1}{2}}r

which is exactly the desired bound. So we complete the proof.

5. Proof of the sharpness

In this section, we construct new examples to prove the sharpness of local LpL^{p} bounds. The sharpness means that for each pair of ν,r\nu,r, there exist eigenfunctions saturating the bound (1.19). We shall use the strategy by Koch-Tataru [45].

5.1. Construction of eigenfunctions in DjintD_{j}^{int}.

For 12jλ231\leq 2^{j}\leq\lambda^{\frac{2}{3}}, λ12jδ2j/2\lambda^{-1}2^{j}\ll\delta\lesssim 2^{-j/2}, and fixed x1>0x_{1}^{*}>0 with λx1λ22j\lambda-x_{1}^{*}\approx\lambda 2^{-2j}, let

𝒯j,δ={x=(x1,x)Djint:|x1x1|λ2jδ2,|x|δ}.\mathcal{T}_{j,\delta}=\{x=(x_{1},x^{\prime})\in D_{j}^{int}:|x_{1}-x_{1}^{*}|\ll\lambda 2^{-j}\delta^{2},\ |x^{\prime}|\ll\delta\}.

Then we can construct the normalized eigenfunctions (dependent on λ,j,δ\lambda,\ j,\ \delta) so that for most x𝒯j,δx\in\mathcal{T}_{j,\delta} we have

(5.1) eλ(x)eλL2(n)λ122j/2δn12.\frac{e_{\lambda}(x)}{\|e_{\lambda}\|_{L^{2}(\mathbb{R}^{n})}}\approx\lambda^{-\frac{1}{2}}2^{j/2}\delta^{-\frac{n-1}{2}}.

To see this, we shall modify the construction of Koch-Tataru’s Example 5.4 [45]. Fix a positive integer NN and we consider the eigenfunctions with eigenvalue λ2=n+2N\lambda^{2}=n+2N. We consider the set of indices

I={αn:|α|=N,αkevenandαkδ2for 2kn}I=\{\alpha\in\mathbb{N}^{n}:|\alpha|=N,\ \alpha_{k}\ \text{even}\ \text{and}\ \alpha_{k}\approx\delta^{-2}\ \text{for}\ 2\leq k\leq n\}

which has size |I|δ2(n1)|I|\approx\delta^{-2(n-1)}. For some subset JJ of II of comparable size, let

eλ(x)=αJk=1nh~αk(xk)e_{\lambda}(x)=\sum_{\alpha\in J}\prod_{k=1}^{n}\tilde{h}_{\alpha_{k}}(x_{k})

where h~αk(xk)\tilde{h}_{\alpha_{k}}(x_{k}) are one dimensional normalized Hermite functions. By orthogonality,

(5.2) eλL2(n)|J|12δ(n1).\|e_{\lambda}\|_{L^{2}(\mathbb{R}^{n})}\approx|J|^{\frac{1}{2}}\approx\delta^{-(n-1)}.

To determine the subset JJ such that (5.1) holds, we need to avoid the cancellations in the summation. We shall use the asymptotic formulas in Section 1.2. Note that for any αI\alpha\in I, we have

(5.3) h~αk(xk)δ12,for 2kn,\tilde{h}_{\alpha_{k}}(x_{k})\approx\delta^{\frac{1}{2}},\ \text{for}\ \ 2\leq k\leq n,

whenever |xk|δ|x_{k}|\ll\delta, since

|s(xk)|=0|xk||t2(2αk+1)|𝑑t1.|s^{-}(x_{k})|=\int_{0}^{|x_{k}|}\sqrt{|t^{2}-(2\alpha_{k}+1)|}dt\ll 1.

Moreover, let u=u(α1)=2α1+1u=u(\alpha_{1})=\sqrt{2\alpha_{1}+1} and

su(x1)=0x1|t2u2|𝑑t.s_{u}^{-}(x_{1})=\int_{0}^{x_{1}}\sqrt{|t^{2}-u^{2}|}dt.

We have for each uu, the value of su(x1)s_{u}^{-}(x_{1}) varies in an interval of size 1\approx 1 as x1x_{1} varies in an interval of size λ2jδ2\approx\lambda 2^{-j}\delta^{2}, since δλ12j\delta\gg\lambda^{-1}2^{j}. So |cos(su(x1))|1|\cos(s_{u}^{-}(x_{1}))|\approx 1 for most x1x_{1}.

Moreover, we have

(5.4) |h~α1(x1)|λ122j/2,|\tilde{h}_{\alpha_{1}}(x_{1})|\approx\lambda^{-\frac{1}{2}}2^{j/2},

since

λ2x12λ222j,λ2u2δ2λ222j.\lambda^{2}-x_{1}^{2}\approx\lambda^{2}2^{-2j},\ \ \lambda^{2}-u^{2}\approx\delta^{-2}\ll\lambda^{2}2^{-2j}.

It remains to insure that the functions h~α1(x1)\tilde{h}_{\alpha_{1}}(x_{1}) have the same sign for all αJ\alpha\in J. Since

dduddx1su(x1)=u|x12u2|2j,\frac{d}{du}\frac{d}{dx_{1}}s_{u}^{-}(x_{1})=\frac{u}{\sqrt{|x_{1}^{2}-u^{2}|}}\approx 2^{j},

we may integrate this with respect to x1x_{1} (from x1x_{1}^{*} to x1x_{1}) and then with respect to uu (from u2u_{2} to u1u_{1}):

(5.5) su1(x1)su2(x1)=su1(x1)su2(x1)+O(2j(u1u2)(x1x1)).s_{u_{1}}^{-}(x_{1})-s_{u_{2}}^{-}(x_{1})=s_{u_{1}}(x_{1}^{*})-s_{u_{2}}(x_{1}^{*})+O(2^{j}(u_{1}-u_{2})(x_{1}-x_{1}^{*})).

The remainder term is 1\ll 1 since |x1x1|λ2jδ2|x_{1}-x_{1}^{*}|\ll\lambda 2^{-j}\delta^{2} and |u1u2|λ1δ2|u_{1}-u_{2}|\lesssim\lambda^{-1}\delta^{2}.

Fix a large constant M1M\gg 1. For k=1,,Mk=1,...,M, let

Jk={αI:su(x1)mod 2π[k1M2π,kM2π]}.J_{k}=\{\alpha\in I:s_{u}^{-}(x_{1}^{*})\ \text{mod}\ 2\pi\in[\tfrac{k-1}{M}2\pi,\tfrac{k}{M}2\pi]\}.

Since k=1MJk=I\cup_{k=1}^{M}J_{k}=I, by the pigeonhole principle there is some Jk0J_{k_{0}} with

|Jk0|1M|I|.|J_{k_{0}}|\geq\frac{1}{M}|I|.

So |Jk0||I|δ2(n1)|J_{k_{0}}|\approx|I|\approx\delta^{-2(n-1)}. Then we have for any two indices in Jk0J_{k_{0}}, the corresponding two phases are close modulo 2π2\pi by (5.5), namely

supx1:|x1x1|λ2jδ2|su1(x1)su2(x1)|1mod2π.\sup_{x_{1}:|x_{1}-x_{1}^{*}|\ll\lambda 2^{-j}\delta^{2}}|s_{u_{1}}^{-}(x_{1})-s_{u_{2}}^{-}(x_{1})|\ll 1\mod 2\pi.

Together with (5.2), (5.3) and (5.4), this implies (5.1).

The examples constructed above can be viewed as intermediate cases between the two kinds of extreme examples in Koch-Tataru [45], which correspond to δλ12j\delta\approx\lambda^{-1}2^{j} (point concentration) and δ2j/2\delta\approx 2^{-j/2} (tube concentration) respectively. Furthermore, the construction of eigenfunctions in DbdD^{bd} is essentially the same as the construction above with 2j=λ232^{j}=\lambda^{\frac{2}{3}}.

5.2. Proof of the sharpness

Fix νn\nu\in\mathbb{R}^{n} such that μ=max{λ43,1λ1|ν|}22j\mu=\max\{\lambda^{-\frac{4}{3}},1-\lambda^{-1}|\nu|\}\approx 2^{-2j} and r>0r>0. Let B(ν,r)B(\nu,r) be the ball {xn:|xν|<r}\{x\in\mathbb{R}^{n}:|x-\nu|<r\}.

Case 1: When rλ1μ12r\lesssim\lambda^{-1}\mu^{-\frac{1}{2}}, we choose δ=λ1μ12\delta=\lambda^{-1}\mu^{-\frac{1}{2}} and then B(ν,r)B(\nu,r) is essentially contained in the tube 𝒯j,δ\mathcal{T}_{j,\delta}. So we get

eλLp(B(ν,r))eλL2(n)\displaystyle\frac{\|e_{\lambda}\|_{L^{p}(B(\nu,r))}}{\|e_{\lambda}\|_{L^{2}(\mathbb{R}^{n})}} λ122j/2δn12rnp\displaystyle\gtrsim\lambda^{-\frac{1}{2}}2^{j/2}\delta^{-\frac{n-1}{2}}\cdot r^{\frac{n}{p}}
(λμ12)n22rnp,\displaystyle\approx(\lambda\mu^{\frac{1}{2}})^{\frac{n-2}{2}}r^{\frac{n}{p}},

which saturates (1.19).

Case 2: When λ1μ12rλμ\lambda^{-1}\mu^{-\frac{1}{2}}\ll r\ll\lambda\mu, we may choose δ=(λμ12/r)12\delta=(\lambda\mu^{\frac{1}{2}}/r)^{-\frac{1}{2}}. So we have δr=λμ12δ2\delta\ll r=\lambda\mu^{\frac{1}{2}}\delta^{2} and then |B(ν,r)𝒯j,δ|rδn1|B(\nu,r)\cap\mathcal{T}_{j,\delta}|\approx r\delta^{n-1}. Thus,

eλLp(B(ν,r))eλL2(n)\displaystyle\frac{\|e_{\lambda}\|_{L^{p}(B(\nu,r))}}{\|e_{\lambda}\|_{L^{2}(\mathbb{R}^{n})}} λ122j/2δn12|B(ν,r)𝒯j,δ|1p\displaystyle\gtrsim\lambda^{-\frac{1}{2}}2^{j/2}\delta^{-\frac{n-1}{2}}\cdot|B(\nu,r)\cap\mathcal{T}_{j,\delta}|^{\frac{1}{p}}
λ12μ14r1pδ(n1)(1p12)\displaystyle\approx\lambda^{-\frac{1}{2}}\mu^{-\frac{1}{4}}r^{\frac{1}{p}}\delta^{(n-1)(\frac{1}{p}-\frac{1}{2})}
=(λ12r12μ14)n+1pn12(λμ12)1p12.\displaystyle=(\lambda^{-\frac{1}{2}}r^{\frac{1}{2}}\mu^{-\frac{1}{4}})^{\frac{n+1}{p}-\frac{n-1}{2}}(\lambda\mu^{\frac{1}{2}})^{\frac{1}{p}-\frac{1}{2}}.

Moreover, we can choose δ=λ1μ12\delta=\lambda^{-1}\mu^{-\frac{1}{2}}, and then the tube 𝒯j,δ\mathcal{T}_{j,\delta} is essentially a ball covered by B(ν,r)B(\nu,r). Since |𝒯j,δ|δn|\mathcal{T}_{j,\delta}|\approx\delta^{n}, we get

eλLp(B(ν,r))eλL2(n)\displaystyle\frac{\|e_{\lambda}\|_{L^{p}(B(\nu,r))}}{\|e_{\lambda}\|_{L^{2}(\mathbb{R}^{n})}} λ122j/2δn12|𝒯j,δ|1p\displaystyle\gtrsim\lambda^{-\frac{1}{2}}2^{j/2}\delta^{-\frac{n-1}{2}}\cdot|\mathcal{T}_{j,\delta}|^{\frac{1}{p}}
(λμ12)n22np.\displaystyle\approx(\lambda\mu^{\frac{1}{2}})^{\frac{n-2}{2}-\frac{n}{p}}.

These saturate (1.19).

Case 3: When λμrλ\lambda\mu\lesssim r\leq\lambda, we fix some integer kk with λμ~λ22kr\lambda\tilde{\mu}\lesssim\lambda 2^{-2k}\lesssim r. The ball B(ν,r)B(\nu,r) may intersect the annuli DkintD_{k}^{int} for these kk. If we choose δ=2k/2\delta=2^{-k/2}, then the tube 𝒯k,δ\mathcal{T}_{k,\delta} is essentially contained in B(ν,r)B(\nu,r) and |𝒯k,δ|λ2kδn+1|\mathcal{T}_{k,\delta}|\approx\lambda 2^{-k}\delta^{n+1}. So we obtain

eλLp(B(ν,r))eλL2(n)\displaystyle\frac{\|e_{\lambda}\|_{L^{p}(B(\nu,r))}}{\|e_{\lambda}\|_{L^{2}(\mathbb{R}^{n})}} λ122k/2δn12|𝒯k,δ|1p\displaystyle\gtrsim\lambda^{-\frac{1}{2}}2^{k/2}\delta^{-\frac{n-1}{2}}\cdot|\mathcal{T}_{k,\delta}|^{\frac{1}{p}}
λ1p122k(n+14n+32p).\displaystyle\approx\lambda^{\frac{1}{p}-\frac{1}{2}}2^{k(\frac{n+1}{4}-\frac{n+3}{2p})}.

Moreover, we may choose δ=λ12k\delta=\lambda^{-1}2^{k}. Then the tube 𝒯k,δ\mathcal{T}_{k,\delta} is essentially a ball covered by B(ν,r)B(\nu,r). Since |𝒯j,δ|δn|\mathcal{T}_{j,\delta}|\approx\delta^{n}, we get

eλLp(B(ν,r))eλL2(n)\displaystyle\frac{\|e_{\lambda}\|_{L^{p}(B(\nu,r))}}{\|e_{\lambda}\|_{L^{2}(\mathbb{R}^{n})}} λ122k/2δn12|𝒯k,δ|1p\displaystyle\gtrsim\lambda^{-\frac{1}{2}}2^{k/2}\delta^{-\frac{n-1}{2}}\cdot|\mathcal{T}_{k,\delta}|^{\frac{1}{p}}
(λ2k)n22np.\displaystyle\approx(\lambda 2^{-k})^{\frac{n-2}{2}-\frac{n}{p}}.

These saturate (1.19) since μ~122k(r/λ)12\tilde{\mu}^{\frac{1}{2}}\lesssim 2^{-k}\lesssim(r/\lambda)^{\frac{1}{2}}. We complete the proof of the sharpness.

6. Further discussions

In this section, we discuss some open problems on the Hermite eigenfunction estimates.

  1. (i)

    Restriction estimates on submanifolds. Let Σ\Sigma be a totally geodesic submanifold (not necessarily compact) in n\mathbb{R}^{n}, e.g. straight lines, hyperplanes passing through the origin. The question is to establish sharp LpL^{p} estimates over Σ\Sigma for the Hermite eigenfunctions. See Burq-Gérard-Tzvetkov [17] and Hu [37] for the restriction estimates for the Laplace eigenfunctions. See also Blair [5], Blair-Sogge [6, 7, 8, 9, 10], Bourgain [12], Bourgain-Rudnick [13], Canzani-Galkowski-Toth [18], Chen-Sogge [24], Chen [23], Greenleaf-Seeger [28], Hassell-Tacy [34], Hezari [33], Huang-Zhang [38], Sogge-Xi-Zhang [55], Sogge-Zelditch [56], Tataru [59], Xi-Zhang [66], Wang-Zhang [64], Wyman [65], Zhang [68].

  2. (ii)

    Kakeya-Nikodym type estimates. Let BB be any fixed compact set in n\mathbb{R}^{n}. Let νn\nu\in\mathbb{R}^{n}. For real numbers r1r2rn>0r_{1}\geq r_{2}\geq...\geq r_{n}>0, let r=diag(r1,,rn)\textbf{r}=diag(r_{1},...,r_{n}). Let

    B(ν,r)={ν+rx:xB}.B(\nu,\textbf{r})=\{\nu+\textbf{r}x:x\in B\}.

    The question is to establish sharp local Lp(p2)L^{p}\ (p\geq 2) estimates over B(ν,r)B(\nu,\textbf{r}) for the Hermite eigenfunctions. The model case r1==rn=rr_{1}=...=r_{n}=r has been handled in this paper. In the case r1r2==rnr_{1}\gg r_{2}=...=r_{n}, the set B(ν,r)B(\nu,\textbf{r}) is roughly a tube, and the local L2L^{2} norms are similar to the Kakeya-Nikodym bounds considered in a series of works by Blair-Sogge [6, 7, 8, 9, 10], and closely related to the restriction estimates on straight lines.

  3. (iii)

    Bilinear and multilinear estimates. It is interesting to investigate the sharp bilinear and multilinear estimates for the Hermite eigenfunctions. See Burq-Gérard-Tzvetkov [14, 15, 16] for the bilinear and multilinear L2L^{2} estimates for the Laplace eigenfunctions, and their applications to nonlinear Schrödinger equations. See also Guo-Han-Tacy [27] for the bilinear LpL^{p} estimates of quasimodes.

Acknowledgments

The authors would like to thank Dr. Xiaoyan Su and Dr. Ying Wang for helpful discussions and comments during the research. The authors would like to thank Professor Sanghyuk Lee for helpful comments on the preprint. X.W. is partially supported by a startup grant from Hunan University. C.Z. is partially supported by NSFC Grant No.1237010173 and a startup grant from Tsinghua University.

Declarations

Data availability statement. Data sharing not applicable to this article as no datasets were generated or analyzed during the current study.

Conflict of interests. The authors have no relevant financial or non-financial interests to disclose.

References

  • [1] George E. Andrews, Richard Askey, and Ranjan Roy. Special functions, volume 71 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 1999.
  • [2] Beck, Thomas; Hanin, Boris. Level spacings and nodal sets at infinity for radial perturbations of the harmonic oscillator. Int. Math. Res. Not. IMRN(2021), no.7, 5007–5036.
  • [3] Bérard, P. and B. Helffer. On the number of nodal domains of the 2D isotropic quantum harmonic oscillator—an extension of results of A. Stern. (2014): preprint arXiv:1409.2333.
  • [4] Bérard, Pierre; Helffer, Bernard. Some nodal properties of the quantum harmonic oscillator and other Schrödinger operators in R2. Geometric and computational spectral theory, 87–116.
  • [5] Matthew D Blair. On logarithmic improvements of critical geodesic restriction bounds in the presence of nonpositive curvature. Israel Journal of Mathematics, 224(1):407–436, 2018.
  • [6] Blair, M. D. and Sogge, C. D., Refined and microlocal Kakeya-Nikodym bounds for eigenfunctions in two dimensions Anal. PDE 8 (2015), 747–764.
  • [7] M. D. Blair and C. D. Sogge. Refined and microlocal Kakeya-Nikodym bounds of eigenfunctions in higher dimensions. Comm. Math. Phys., 356(2):501–533, 2017.
  • [8] Blair, M. D. and Sogge, C. D., Kakeya-Nikodym averages, Lp-norms and lower bounds for nodal sets of eigenfunctions in higher dimensions, J. European Math. Soc., 17 (2015), 2513–2543.
  • [9] M. D. Blair and C. D. Sogge. Concerning Toponogov’s theorem and logarithmic improvement of estimates of eigenfunctions. J. Differential Geom., 109(2):189–221, 2018.
  • [10] M. D. Blair and C. D. Sogge. Logarithmic improvements in Lp bounds for eigenfunctions at the critical exponent in the presence of nonpositive curvature. Invent. Math., 217(2):703–748, 2019.
  • [11] Bohr, N. (1920), Über die Serienspektra der Elemente [About the serial spectra of the elements], Zeitschrift für Physik (in German), 2 (5): 423–478.
  • [12] J. Bourgain. Geodesic restrictions and Lp-estimates for eigenfunctions of Riemannian surfaces. In Linear and complex analysis, volume 226 of Amer. Math. Soc. Transl. Ser. 2, pages 27–35. Amer. Math. Soc., Providence, RI, 2009.
  • [13] J. Bourgain and Z. Rudnick. Restriction of toral eigenfunctions to hypersurfaces and nodal sets. Geometric and Functional Analysis, 22(4):878–937, 2012.
  • [14] Burq, N., Gérard, P., Tzvetkov, N., Multilinear estimates for the Laplace spectral projectors on compact manifolds. C. R. Math. Acad. Sci. Paris338(2004), no.5, 359–364.
  • [15] Burq, N., Gérard, P., Tzvetkov, N., Bilinear eigenfunction estimates and the nonlinear Schrödinger equation on surfaces. Invent. Math.159(2005), no.1, 187–223.
  • [16] Burq, N., Gérard, P., Tzvetkov, N., Multilinear eigenfunction estimates and global existence for the three dimensional nonlinear Schrödinger equations.Ann. Sci. École Norm. Sup. (4)38(2005), no.2, 255–301.
  • [17] Burq, N., Gérard, P. and Tzvetkov, N., Restrictions of the Laplace-Beltrami eigenfunctions to submanifolds, Duke Math. J. 138 (2007), 445–486.
  • [18] Yaiza Canzani, Jeffrey Galkowski, and John A Toth. Averages of eigenfunctions over hypersurfaces. Communications in Mathematical Physics, 360(2):619–637, 2018.
  • [19] Chen, Peng; Li, Ji; Ward, Lesley; Yan, Lixin. Weak-type endpoint bounds for Bochner-Riesz means for the Hermite operator. C. R. Math. Acad. Sci. Paris360(2022), 111–126.
  • [20] Chen, Peng; Duong, Xuan Thinh; He, Danqing; Lee, Sanghyuk; Yan, Lixin. Almost everywhere convergence of Bochner-Riesz means for the Hermite operators. Adv. Math.392(2021), Paper No. 108042, 42 pp.
  • [21] Chen, Peng; Lee, Sanghyuk; Sikora, Adam; Yan, Lixin. Bounds on the maximal Bochner-Riesz means for elliptic operators.Trans. Amer. Math. Soc.373(2020), no.6, 3793–3828.
  • [22] Chen, Peng; Lin, Xixi; Yan, Lixin. The commutators of Bochner-Riesz operators for Hermite operator. J. Geom. Anal.33(2023), no.3, Paper No. 87, 19 pp.
  • [23] X. Chen. An improvement on eigenfunction restriction estimates for compact boundaryless Riemannian manifolds with nonpositive sectional curvature. Trans. Amer. Math. Soc., 367:4019–4039, 2015.
  • [24] X. Chen and C. D. Sogge. A few endpoint geodesic restriction estimates for eigenfunctions. Comm. Math. Phys., 329(2):435–459, 2014.
  • [25] Luis Escauriaza. Carleman inequalities and the heat operator. Duke Math. J., 104(1):113–127, 2000.
  • [26] Luis Escauriaza and Luis Vega. Carleman inequalities and the heat operator. II. Indiana Univ. Math. J., 50(3):1149–1169, 2001.
  • [27] Guo, Zihua; Han, Xiaolong; Tacy, Melissa, Lp bilinear quasimode estimates. J. Geom. Anal.29(2019), no.3, 2242–2289.
  • [28] A. Greenleaf and A. Seeger. Fourier integral operators with fold singularities. J Reine Angew. Math., 455:35–56, 1994.
  • [29] Han, Xiaolong, Small scale quantum ergodicity in negatively curved manifolds. Nonlinearity 28(2015), no.9, 3263–3288.
  • [30] Hanin, Boris; Zelditch, Steve; Zhou, Peng. Nodal sets of random eigenfunctions for the isotropic harmonic oscillator. Int. Math. Res. Not. IMRN(2015), no.13, 4813–4839.
  • [31] Hanin, Boris; Zelditch, Steve; Zhou, Peng. Scaling of harmonic oscillator eigenfunctions and their nodal sets around the caustic. Comm. Math. Phys.350(2017), no.3, 1147–1183.
  • [32] Hezari, Hamid; Rivière, Gabriel, Lp norms, nodal sets, and quantum ergodicity. Adv. Math. 290(2016), 938–966.
  • [33] H. Hezari. Quantum ergodicity and l p norms of restrictions of eigenfunctions. Communications in Mathematical Physics, 357(3):1157–1177, 2018.
  • [34] Hassell, A. and Tacy, M. Improvement of eigenfunction estimates on manifolds of non-positive curvature, Forum Math. 27 (2015), 1435–1451.
  • [35] L. Hörmander. Oscillatory integrals and multipliers on F Lp, Ark. Math. 11 (1973), 1–11.
  • [36] L. Hörmander. The Analysis of Linear Partial Differential Operators I Distribution Theory and Fourier Analysis, Springer-Verlag, Berlin, 1985.
  • [37] R. Hu. Lp norm estimates of eigenfunctions restricted to submanifolds. Forum Math., 6:1021–1052, 2009.
  • [38] X. Huang and C. Zhang. Restriction of toral eigenfunctions to totally geodesic submanifolds. Anal. PDE14(2021), no.3, 861–880..
  • [39] S. Lee and J. Ryu, Bochner-Riesz means for the Hermite and special Hermite expansions. Adv. Math.400(2022), Paper No. 108260, 52 pp.
  • [40] E. Jeong, S. Lee, J. Ryu, Hermite spectral projection operator, arXiv:2006.11762
  • [41] E. Jeong, S. Lee, J. Ryu, Endpoint eigenfunction bounds for the Hermite operator, arXiv:2205.03036.
  • [42] E. Jeong, S. Lee, J. Ryu, Bounds on the Hermite spectral projection operator, arXiv:2210.03385
  • [43] Long Jin. Semiclassical Cauchy estimates and applications. Trans. Amer. Math. Soc. March 18, 2016.
  • [44] G. B. Karadzhov, Riesz summability of multiple Hermite series in Lp spaces, C. R. Acad. Bulgare Sci. 47 (1994), 5–8.
  • [45] H. Koch, D. Tataru, Lp eigenfunction bounds for the Hermite operator, Duke Math. J. 128 (2005), 369– 392.
  • [46] H. Koch, D. Tataru, Carleman estimates and unique continuation for second order parabolic equations with nonsmooth coefficients. Comm. Partial Differential Equations 34(2009), no.4-6, 305–366.
  • [47] H. Koch, D. Tataru, M. Zworski, Semiclassical Lp estimates, Ann. Henri Poincaré 8(2007), 885–916.
  • [48] A. Reznikov. Norms of geodesic restrictions for eigenfunctions on hyperbolic surfaces and representation theory. arXiv:math/0403437
  • [49] A. Reznikov. A uniform bound for geodesic periods of eigenfunctions on hyperbolic surfaces. In Forum Mathematicum, volume 27, pages 1569–1590. De Gruyter, 2015.
  • [50] P. Sjögren, J. L. Torrea, On the boundary convergence of solutions to the Hermite-Schrödinger equation, Colloq. Math. 118 (2010), 161–174.
  • [51] C. D. Sogge, Oscillatory integrals and spherical harmonics, Duke Math. J., 53 (1986), no. 1, 43‐65.
  • [52] C. D. Sogge. Concerning the Lp norm of spectral clusters for second-order elliptic operators on compact manifolds. J. Funct. Anal., 77(1):123-138, 1988.
  • [53] C.D. Sogge, Localized Lp-estimates of eigenfunctions: a note on an article of Hezari and Rivière. Adv. Math. 289(2016), 384–396.
  • [54] C. D. Sogge. Fourier integrals in classical analysis, volume 210 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, second edition, 2017.
  • [55] Christopher D Sogge, Yakun Xi, and Cheng Zhang. Geodesic period integrals of eigenfunctions on riemannian surfaces and the gauss–bonnet theorem. Cambridge Journal of Mathematics, 5(1):123–151, 2017.
  • [56] C. D. Sogge and S. Zelditch. On eigenfunction restriction estimates and L4-bounds for compact surfaces with nonpositive curvature. In Advances in analysis: the legacy of Elias M. Stein, volume 50 of Princeton Math. Ser., pages 447–461. Princeton Univ. Press, Princeton, NJ, 2014.
  • [57] E.M. Stein, Harmonic Analysis: real-variable methods, orthogonality, and osciollatory integrals, Princeton University Press, Princeton, NJ, 1993.
  • [58] Szegö, Gábor (1955), Orthogonal Polynomials, Colloquium Publications, 23 (4th ed.), American Mathematical Society
  • [59] Daniel Tataru. On the regularity of boundary traces for the wave equation. Annali della Scuola Normale Superiore di Pisa-Classe di Scienze, 26(1):185–206, 1998.
  • [60] S. Thangavelu, Multipliers for Hermite expansions, Rev. Mat. Iberoam. 3 (1987), 1–24.
  • [61] S. Thangavelu, Lectures on Hermite and Laguerre expansions, Princeton University Press, Princeton, N.J., 1993.
  • [62] S. Thangavelu. Hermite and special Hermite expansions revisited. Duke Math. J., 94(2):257–278, 1998.
  • [63] Tipler, Paul; Llewellyn, Ralph (2008). Modern Physics (5 ed.). W. H. Freeman and Company. pp. 160–161. ISBN 978-0-7167-7550-8.
  • [64] X. Wang and C. Zhang, Sharp endpoint estimates for eigenfunctions restricted to submanifolds of codimension 2, Adv. Math. 386 (2021)
  • [65] E. Wyman. Explicit bounds on integrals of eigenfunctions over curves in surfaces of nonpositive curvature. The Journal of Geometric Analysis, pages 1–29, 2019.
  • [66] Y. Xi and C. Zhang. Improved critical eigenfunction restriction estimates on Riemannian surfaces with nonpositive curvature. Comm. Math. Phys., 350(3):1299–1325, 2017.
  • [67] S. Zelditch. Kuznecov sum formulae and szeg¨o limit formulae on manifolds. Communications in partial differential equations, 17(1-2):221–260, 1992.
  • [68] C. Zhang. Improved critical eigenfunction restriction estimates on Riemannian manifolds with constant negative curvature. Journal of Functional Analysis, 272(11):4642–4670, 2017.