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

Boundedness for maximal operators
over hypersurfaces in 3\mathbb{R}^{3} thanks: This work is supported by the National Key R&D Program of China (No.2023YFA1010800); the Natural Science Foundation of China (No.12271435, No.12301113).

Wenjuan Li
School of Mathematics and Statistics, Northwestern Polytechnical University, Xi’an, 710129, China
Huiju Wang
School of Mathematics and Statistics, Henan University, Kaifeng, 475000, China
Corresponding author’s email address: [email protected]

Abstract: In this article, we study maximal functions related to hypersurfaces with vanishing Gaussian curvature in 3\mathbb{R}^{3}. Firstly, we characterize the LpLqL^{p}\rightarrow L^{q} boundedness of local maximal operators along homogeneous hypersurfaces. Moreover, weighted LpL^{p}-estimates are obtained for the corresponding global operators. Secondly, for a class of hypersurfaces that lack a homogeneous structure and pass through the origin, we attempt to look for other geometric properties instead of height of hypersurfaces to characterize the optimal LpL^{p}-boundedness of the corresponding global maximal operators.

Keywords: Maximal function; height of hypersurfaces; vanishing Gaussian curvature.

Mathematics Subject Classification: 42B20, 42B25.

1 Introduction

Let SS be a smooth hypersurface in n\mathbb{R}^{n} and ηC0(S)\eta\in C_{0}^{\infty}(S) be a smooth non-negative function with compact support. Then the associated averaging operator is defined by

Atf(y)=Sf(ytx)η(x)𝑑σ(x),t>0,A_{t}f(y)=\int_{S}f(y-tx)\eta(x)d\sigma(x),\quad t>0, (1.1)

where dσd\sigma denotes the surface measure on SS. A fundamental and still largely open problem is to characterize the Lp(n)L^{p}(\mathbb{R}^{n})-boundedness of the global maximal operator associated to the hypersurface where the Gaussian curvature at some points is allowed to vanish. Iosevich [8] gave a confirmed answer for the 22-dimensional case, i.e. SS is a curve of finite type in 2\mathbb{R}^{2}. For higher dimension case, one can see the works by Buschenhenke-Dendrinos-Ikromov-Müller [1], Buschenhenke-Ikromov-Müller [2], Greenleaf [3], Ikromov-Müller [4, 5], Ikromov-Kempe-Müller [6, 7], Iosevich-Sawyer [9, 10], Li [11], Zimmermann [16] etc.

However, much less is known about the LpLqL^{p}\rightarrow L^{q}-boundedness (pq)(p\leq q) for local maximal operator supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| associated with hypersurface SS whose Gaussian curvature vanishes at some points, especially when SnS\subset\mathbb{R}^{n}, n3n\geq 3. Once we obtain the LpLqL^{p}\rightarrow L^{q}-boundedness for local maximal operators supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}|, then the weighted estimates will follow for the corresponding global maximal operators by the methodology of sparse domination (see Section 1.2 in [12]). In this paper, we mainly concentrate on two kinds of maximal operators, specifically, in Section 1.1 we study maximal operators associated with homogeneous hypersurfaces in 3\mathbb{R}^{3}; in Section 1.2, we investigate other maximal operators along a class of hypersurfaces that lack a homogeneous structure and pass through the origin.

1.1 Maximal operators associated with homogeneous hypersurfaces

We first consider the maximal operators along the hypersurfaces S:={(x1,x2,Φ(x1,x2)+c):(x1,x2)U}S:=\{(x_{1},x_{2},\Phi(x_{1},x_{2})+c):(x_{1},x_{2})\in U\}, where cc\in\mathbb{R}, UU is a small neighborhood of the origin in 2\mathbb{R}^{2}. Φ\Phi is a homogeneous polynomial that satisfies Φ(λ1/m)=λΦ()\Phi(\lambda^{1/m}\cdot)=\lambda\Phi(\cdot) for all λ>0\lambda>0. We always assume that m2m\geq 2 in this paper. In order to describe the structure of Φ\Phi, we introduce the following proposition, which appears in the Proposition 2.2 in [4].

Proposition 1.1.

Let PP be a (κ1,κ2)(\kappa_{1},\kappa_{2})-homogeneous polynomial of degree one, i.e. P(λκ1,λκ2)=λP(,)P(\lambda^{\kappa_{1}\cdot},\lambda^{\kappa_{2}}\cdot)=\lambda P(\cdot,\cdot) for each λ>0\lambda>0, and assume that PP is not of the form cx1ν1x2ν2cx_{{}_{1}}^{\nu_{1}}x_{2}^{\nu_{2}}. Then κ1,κ2\kappa_{1},\kappa_{2} are uniquely determined by PP, and κ1,κ2\kappa_{1},\kappa_{2}\in\mathbb{Q}. We assume that κ1κ2\kappa_{1}\leq\kappa_{2}, and write

κ1=qm,κ2=pm,p,q,m=1,\kappa_{1}=\frac{q}{m},\kappa_{2}=\frac{p}{m},\quad\langle p,q,m\rangle=1,

so that in particular qpq\leq p. Then p,q=1\langle p,q\rangle=1, and there exist non-negative integers α1,α2\alpha_{1},\alpha_{2} and a (1,1)(1,1)-homogeneous polynomial QQ such that the polynomial PP can be written as

P(x1,x2)=x1α1x2α2Q(x1p,x2q).P(x_{1},x_{2})=x_{1}^{\alpha_{1}}x_{2}^{\alpha_{2}}Q(x_{1}^{p},x_{2}^{q}).

More precisely, PP can be written in the form

P(x1,x2)=cΦx1ν1x2ν2Πh=1N(x2qλhx1p)nhP(x_{1},x_{2})=c_{\Phi}x_{1}^{\nu_{1}}x_{2}^{\nu_{2}}\Pi_{h=1}^{N}(x_{2}^{q}-\lambda_{h}x_{1}^{p})^{n_{h}}

with N1N\geq 1, distinct λh\{0}\lambda_{h}\in\mathbb{C}\backslash\{0\} and multiplicities nh\{0}n_{h}\in\mathbb{N}\backslash\{0\} with ν1,ν2\nu_{1},\nu_{2}\in\mathbb{N} (possibly different from α1,α2\alpha_{1},\alpha_{2}).

Applying Proposition 1.1 with κ1=κ2=1/m\kappa_{1}=\kappa_{2}=1/m, then we write Φ\Phi as

Φ(x1,x2)=cΦx1ν1x2ν2Πh=1N(x2λhx1)nh.\Phi(x_{1},x_{2})=c_{\Phi}x_{1}^{\nu_{1}}x_{2}^{\nu_{2}}\Pi_{h=1}^{N}(x_{2}-\lambda_{h}x_{1})^{n_{h}}. (1.2)

It is clear that ν1+ν2+h=1Nnh=m\nu_{1}+\nu_{2}+\sum_{h=1}^{N}n_{h}=m. When one of ν1\nu_{1}, ν2\nu_{2}, nhn_{h}, 1hN1\leq h\leq N equals mm, then after some linear change of coordinates, the hypersurface SS transforms to {(y1,y2,y1m+c):(y1,y2)U}\{(y_{1},y_{2},y_{1}^{m}+c):(y_{1},y_{2})\in U\}, and the related maximal operators have been studied in [7, 16, 11, 12]. In this paper, we concentrate on the case when 0ν1<m0\leq\nu_{1}<m, 0ν2<m0\leq\nu_{2}<m and 0nh<m0\leq n_{h}<m.

We also introduce the definition of the height of Φ\Phi which is original from [6]. The height of Φ\Phi is defined by

hΦ:=max{m2,orddΦ},h_{\Phi}:=\max\{\frac{m}{2},\textsf{ordd}\hskip 2.84544pt\Phi\},

where orddΦ=supxS1orddΦ(x)\textsf{ordd}\hskip 2.84544pt\Phi=\sup_{x\in S^{1}}\textsf{ordd}\hskip 2.84544pt\Phi(x), S1S^{1} is the unit circle in 2\mathbb{R}^{2} and orddΦ(x)\textsf{ordd}\hskip 2.84544pt\Phi(x) is the smallest non-negative integer jj such that the jj-th order total derivative of Φ\Phi at xx is non-zero.

Now we are ready to state  our main results. Let ZΦZ_{\Phi} be the zero set of Φ\Phi, and ZHΦZ_{H\Phi} be the zero set of the determinant of the Hessian matrix of Φ\Phi, and Δ0\Delta_{0} be constructed by the interior of the quadrilateral with vertices P1=(0,0)P_{1}=(0,0), P2=(2/3,2/3)P_{2}=(2/3,2/3), P3=(2/3,1/3)P_{3}=(2/3,1/3), P4=(3/5,1/5)P_{4}=(3/5,1/5), and the segment P1P2P_{1}P_{2} with P1P_{1} included and P2P_{2} excluded. We first consider the local maximal operator supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| when ZHΦZΦZ_{H\Phi}\subset Z_{\Phi}.

Theorem 1.2.

Assuming that ZHΦZΦZ_{H\Phi}\subset Z_{\Phi}, then we have
(1) when c=0c=0, there exists a constant Cp,q>0C_{p,q}>0 such that supt[1,2]|At|Lp(3)Lq(3)Cp,q\|sup_{t\in[1,2]}|A_{t}|\|_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\leq C_{p,q} for

(1p,1q)Δ1~={(1p,1q)Δ0:hΦ+1phΦ+1q1<0};(\frac{1}{p},\frac{1}{q})\in\widetilde{\Delta_{1}}=\{(\frac{1}{p},\frac{1}{q})\in\Delta_{0}:\frac{h_{\Phi}+1}{p}-\frac{h_{\Phi}+1}{q}-1<0\};

(2) when c0c\neq 0, there exists a constant Cp,q>0C_{p,q}>0 such that supt[1,2]|At|Lp(3)Lq(3)Cp,q\|sup_{t\in[1,2]}|A_{t}|\|_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\leq C_{p,q} for

(1p,1q)Δ2~={(1p,1q)Δ0:hΦ+1p1q1<0}.(\frac{1}{p},\frac{1}{q})\in\widetilde{\Delta_{2}}=\{(\frac{1}{p},\frac{1}{q})\in\Delta_{0}:\frac{h_{\Phi}+1}{p}-\frac{1}{q}-1<0\}.

Theorem 1.2 is sharp up to the endpoints. For example, let Φ(x1,x2)=x1ν1x2ν2\Phi(x_{1},x_{2})=x_{1}^{\nu_{1}}x_{2}^{\nu_{2}}, where ν1\nu_{1} and ν2\nu_{2} are positive integers such that ν1>ν22\nu_{1}>\nu_{2}\geq 2. Then it is clear that ZHΦZΦZ_{H\Phi}\subset Z_{\Phi} and hΦ=ν1h_{\Phi}=\nu_{1}. By a similar argument as in the proof of Theorem 2.7 in [12], when c=0c=0, it can be obtained that supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| cannot be Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) bounded unless ν1+1pν1+1q10\frac{\nu_{1}+1}{p}-\frac{\nu_{1}+1}{q}-1\leq 0, and when c0c\neq 0, ν1+1p1q10\frac{\nu_{1}+1}{p}-\frac{1}{q}-1\leq 0 is necessary for the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})-boundedness of supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}|. Using a method similar to that of the spherical maximal operator, the requirement that (1p,1q)(\frac{1}{p},\frac{1}{q}) belongs to the closure of Δ0\Delta_{0} is necessary. Proof of the necessary conditions is provided in the appendix. The proof of the positive results in Theorem 1.2 is left to Section 4 below.

[Uncaptioned image]

Figure 1: The range of Δ1~\widetilde{\Delta_{1}} when the height hΦh_{\Phi} changes.

[Uncaptioned image]

Figure 2: The range of Δ2~\widetilde{\Delta_{2}} when the height hΦh_{\Phi} changes.

Notice when hΦ=3/2h_{\Phi}=3/2, the results in Theorem 1.2 (1) coincide with that of the local spherical maximal operators (see the article [14]), and when hΦ>3/2h_{\Phi}>3/2, the results in Theorem 1.2 depends heavily on the height of Φ\Phi, see Figure 1 and Figure 2. However, when ZHΦZΦZHΦZ_{H\Phi}\cap Z_{\Phi}\subsetneqq Z_{H\Phi}, the level set {(x1,x2):Φ(x1,x2)=1}\{(x_{1},x_{2}):\Phi(x_{1},x_{2})=1\} will also play an essential role in the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) boundedness of supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}|. In what follows, we assume that the level set {(x1,x2):Φ(x1,x2)=1}\{(x_{1},x_{2}):\Phi(x_{1},x_{2})=1\} determines a curve of finite type MM near each straight line that is contained in ZHΦ\ZΦZ_{H\Phi}\backslash Z_{\Phi}, 2Mm2\leq M\leq m. When M=2M=2, we define ΔM:=Δ0\Delta_{M}:=\Delta_{0}, and when M3M\geq 3 we define ΔM\Delta_{M} in the following cases:

(1) M6M\geq 6, ΔM\Delta_{M} is constructed by the interior of the quadrilateral with vertices P1=(0,0)P_{1}=(0,0), P2=(M+12M,M+12M)P_{2}=(\frac{M+1}{2M},\frac{M+1}{2M}), P3=(4M+2,2M+2)P_{3}=(\frac{4}{M+2},\frac{2}{M+2}), P4=(3M+2,1M+2)P_{4}=(\frac{3}{M+2},\frac{1}{M+2}), and the segment P1P2P_{1}P_{2} with P1P_{1} included and P2P_{2} excluded;

(2) 4M54\leq M\leq 5, ΔM\Delta_{M} is constructed by the interior of the quadrilateral with vertices P1=(0,0)P_{1}=(0,0), P2=(M+12M,M+12M)P_{2}=(\frac{M+1}{2M},\frac{M+1}{2M}), P3=(2M+45M+2,M+25M+2)P_{3}=(\frac{2M+4}{5M+2},\frac{M+2}{5M+2}), P4=(3M+2,1M+2)P_{4}=(\frac{3}{M+2},\frac{1}{M+2}), and the segment P1P2P_{1}P_{2} with P1P_{1} included and P2P_{2} excluded;

(3) M=3M=3, ΔM\Delta_{M} is constructed by the interior of the quadrilateral with vertices P1=(0,0)P_{1}=(0,0), P2=(M+12M,M+12M)P_{2}=(\frac{M+1}{2M},\frac{M+1}{2M}), P3=(2M+45M+2,M+25M+2)P_{3}=(\frac{2M+4}{5M+2},\frac{M+2}{5M+2}), P4=(3M+68M+4,M+28M+4)P_{4}=(\frac{3M+6}{8M+4},\frac{M+2}{8M+4}), and the segment P1P2P_{1}P_{2} with P1P_{1} included and P2P_{2} excluded.

[Uncaptioned image]

Figure 3: The range of ΔM\Delta_{M} when MM changes.

Theorem 1.3.

Suppose that ZHΦZΦZHΦZ_{H\Phi}\cap Z_{\Phi}\subsetneqq Z_{H\Phi}, we have the following results.
(1) If c=0c=0, then there exists a constant Cp,q>0C_{p,q}>0 such that supt[1,2]|At|Lp(3)Lq(3)Cp,q\|sup_{t\in[1,2]}|A_{t}|\|_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\leq C_{p,q} for

(1p,1q)Δ3~={(1p,1q)ΔM:hΦ+1phΦ+1q1<0};(\frac{1}{p},\frac{1}{q})\in\widetilde{\Delta_{3}}=\{(\frac{1}{p},\frac{1}{q})\in\Delta_{M}:\frac{h_{\Phi}+1}{p}-\frac{h_{\Phi}+1}{q}-1<0\};

(2) If c0c\neq 0, then there exists a constant Cp,q>0C_{p,q}>0 such that supt[1,2]|At|Lp(3)Lq(3)Cp,q\|sup_{t\in[1,2]}|A_{t}|\|_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\leq C_{p,q} for

(1p,1q)Δ4~={(1p,1q)ΔM:hΦ+1p1q1<0}.(\frac{1}{p},\frac{1}{q})\in\widetilde{\Delta_{4}}=\{(\frac{1}{p},\frac{1}{q})\in\Delta_{M}:\frac{h_{\Phi}+1}{p}-\frac{1}{q}-1<0\}.

Theorem 1.3 cannot be improved in some cases. For example, we choose Φ(x1,x2)=x1M+x2M\Phi(x_{1},x_{2})=x_{1}^{M}+x_{2}^{M}, and consider the local maximal operator supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| related to the hypersurfaces {(x1,x2,x1M+x2M+c):|x1|<ϵ|x2|}\{(x_{1},x_{2},x_{1}^{M}+x_{2}^{M}+c):|x_{1}|<\epsilon|x_{2}|\}, M6M\geq 6, ϵ1\epsilon\ll 1. It is clear that hΦ=M2h_{\Phi}=\frac{M}{2}, the level set {(x1,x2):x1M+x2M=1}\{(x_{1},x_{2}):x_{1}^{M}+x_{2}^{M}=1\} determines a curve which is finite type MM at x1=0x_{1}=0. When c0c\neq 0, Theorem 1.3 (2) implies that supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| is Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) bounded if (1p,1q)(\frac{1}{p},\frac{1}{q}) satisfies 1q1p<3q\frac{1}{q}\leq\frac{1}{p}<\frac{3}{q} and M/2+1p1q1<0\frac{M/2+1}{p}-\frac{1}{q}-1<0. This result is sharp up to the endpoints, and the proof of the sharpness can be found in the appendix. When c=0c=0, it follows from Theorem 1.3 (1) that supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| is Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) bounded if (1p,1q)ΔM(\frac{1}{p},\frac{1}{q})\in\Delta_{M}. Moreover, it can be proved that the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) boundedness of supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| cannot be true unless 1p2MM+1\frac{1}{p}\leq\frac{2M}{M+1}, 1p1q2M+20\frac{1}{p}-\frac{1}{q}-\frac{2}{M+2}\leq 0 and 1q1p3q\frac{1}{q}\leq\frac{1}{p}\leq\frac{3}{q}. Comparing with the results obtained by Theorem 1.3 (1), only the necessity of the segment connecting P2P_{2} and P3P_{3} is unknown. Readers may refer to the appendix for the proof of the necessary conditions.

In a small neighborhood of the straight line mentioned before that is contained in ZHΦ\ZΦZ_{H\Phi}\backslash Z_{\Phi}, we can reduce the proof of Theorem 1.3 to investigating the local maximal operator related to the hypersurfaces {(r(1+sMg(s)),rs,rm+c):r[1,2],s(0,ϵ0)},\{(r(1+s^{M}g(s)),rs,r^{m}+c):r\in[1,2],s\in(0,\epsilon_{0})\}, 2Mm2\leq M\leq m, whose Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})-boundedness will be obtained in Theorem 1.4 below. In fact, Theorem 1.4 will refer to more general maximal operators. Section 2 and Section 3 are dedicated to proving Theorem 1.4. Theorem 1.3 will be proved in Section 4.

Theorem 1.4.

Let

dμ^(ξ)=2eir(1+sMg(s))ξ1irsξ2irdξ3ψ(s)ρ(r)𝑑s𝑑r,\widehat{d\mu}(\xi)=\int_{\mathbb{R}^{2}}e^{-ir(1+s^{M}g(s))\xi_{1}-irs\xi_{2}-ir^{d}\xi_{3}}\psi(s)\rho(r)dsdr,

where gg is a smooth function with g(0)0g(0)\neq 0, ρ\rho and ψ\psi are smooth cutoff functions, supp ρ[1,2]\rho\subset[1,2] and supp ψ(0,ϵ0)\psi\subset(0,\epsilon_{0}) with ϵ0\epsilon_{0} sufficiently small, d2d\geq 2 and M2M\geq 2 are positive integers. Given the averaging operator

Atf(y)=3eiyξ+itcξ3dμ^(tξ)f^(ξ)𝑑ξA_{t}f(y)=\int_{\mathbb{R}^{3}}e^{iy\cdot\xi+itc\xi_{3}}\widehat{d\mu}(t\xi)\hat{f}(\xi)d\xi (1.3)

with some constant c0c\geq 0, then we have the following estimate

supt[1,2]|Atf|Lq(3)max{1,c1q}fLp(3),\biggl{\|}\sup_{t\in[1,2]}|A_{t}f|\biggl{\|}_{L^{q}(\mathbb{R}^{3})}\lesssim\max\{1,c^{\frac{1}{q}}\}\|f\|_{L^{p}(\mathbb{R}^{3})}, (1.4)

provided that (1p,1q)(\frac{1}{p},\frac{1}{q}) belongs to ΔM\Delta_{M}.

The following weighted estimates for the global maximal operator supt>0|At|\sup_{t>0}|A_{t}| follow from Theorem 1.2, Theorem 1.3 and the methodology of sparse domination. One can see [12] and the appendix in this paper for more details.

Theorem 1.5.

Let 0<p<r<q0<p<r<q, ωArpRH(qr)\omega\in A_{\frac{r}{p}}\cap RH_{\left(\frac{q}{r}\right)^{\prime}}, α=max(1rp,q1qr)\alpha=\max\left(\frac{1}{r-p},\frac{q-1}{q-r}\right). Then there holds the weighted estimate

supt>0|At|Lr(ω)Lr(ω)C([ω]Arp[ω]RH(qr))α\biggl{\|}\sup_{t>0}|A_{t}|\biggl{\|}_{L^{r}(\omega)\rightarrow L^{r}(\omega)}\leq C\left([\omega]_{A_{\frac{r}{p}}}[\omega]_{RH_{\left(\frac{q}{r}\right)^{\prime}}}\right)^{\alpha}

in each of the following cases:
(1) AtA_{t} is defined by (1.1) with c=0c=0, ZHΦZΦZ_{H\Phi}\subset Z_{\Phi}, and (1p,1q)Δ1~(\frac{1}{p},\frac{1}{q})\in\widetilde{\Delta_{1}};
(2) AtA_{t} is defined by (1.1) with c0c\neq 0, ZHΦZΦZ_{H\Phi}\subset Z_{\Phi}, and (1p,1q)Δ2~(\frac{1}{p},\frac{1}{q})\in\widetilde{\Delta_{2}};
(3) AtA_{t} is defined by (1.1) with c=0c=0, ZHΦZΦZHΦZ_{H\Phi}\cap Z_{\Phi}\subsetneqq Z_{H\Phi}, and (1p,1q)Δ3~(\frac{1}{p},\frac{1}{q})\in\widetilde{\Delta_{3}};
(4) AtA_{t} is defined by (1.1) with c0c\neq 0, ZHΦZΦZHΦZ_{H\Phi}\cap Z_{\Phi}\subsetneqq Z_{H\Phi}, and (1p,1q)Δ4~(\frac{1}{p},\frac{1}{q})\in\widetilde{\Delta_{4}}.

We notice that the unweighted LpL^{p}-estimates for the global maximal operators associated with mixed homogeneous hypersurfaces in 3\mathbb{R}^{3} have been studied in [6], where the hypersurface is given as a graph of Φ+c\Phi+c, where cc\in\mathbb{R} and Φ\Phi is a mixed homogeneous function. They proved that the global maximal operator is Lp(3)L^{p}(\mathbb{R}^{3})-bounded provided that p>hΦ2p>h_{\Phi}\geq 2. Their results are sharp when c0c\neq 0, but the sharp results for hΦ<2h_{\Phi}<2 still remain open. When ω=1\omega=1 and c0c\neq 0 in Theorem 1.5, our results coincide with that of [6] for the homogeneous function Φ\Phi. Indeed, Theorem 1.5 also includes some results when hΦ<2h_{\Phi}<2 or c=0c=0.

Moreover, the method we used to show Theorem 1.4 and Theorem 1.5 can also be applied to get the sharp results on the boundedness of the maximal operators along a class of hypersurfaces which do not have a homogeneous structure and pass through the origin, see Section 1.2 below.

1.2 Other maximal operators over hypersurfaces

Let \mathcal{M} be the global maximal operators defined by the averaging operator (1.1) along the hypersurfaces S={(x1,x2,Ψ(x1,x2)+c):(x1,x2)U}S=\{(x_{1},x_{2},\Psi(x_{1},x_{2})+c):(x_{1},x_{2})\in U\}, where Ψ\Psi is a smooth function on UU with Ψ(0,0)=0\Psi(0,0)=0 and Ψ(0,0)=(0,0)\nabla\Psi(0,0)=(0,0), UU is a small neighborhood of the origin (0,0)(0,0), and cc\in\mathbb{R}. When c0c\not=0, and Ψ\Psi is an analytic function, according to [7, Corollary 1.4], \mathcal{M} is Lp(3)L^{p}(\mathbb{R}^{3}) bounded if and only if p>hΨp>h_{\Psi} when hΨ2h_{\Psi}\geq 2. If hΨ<2h_{\Psi}<2, Buschenhenke, Dendrinos, Ikoromov and Müller still work on this topic, see recent articles [1, 2].

When c=0c=0, the article [6] gave a remark that one can get better result than p>hΨp>h_{\Psi}. If Ψ\Psi is an analytic function, it follows from [16, Theorem 1.3.1] that \mathcal{M} is Lp(3)L^{p}(\mathbb{R}^{3}) bounded if p>2p>2. However, how to characterize the sharp exponents of LpL^{p} boundedness for \mathcal{M} from geometric properties for the hypersurfaces passing through the origin? In this section, we make an attempt to answer this question.

Now we consider the maximal operators along the hypersurfaces S={(x1,x2,(x2x1dϕ(x1))m):(x1,x2)U}S=\{(x_{1},x_{2},(x_{2}-x_{1}^{d}\phi(x_{1}))^{m}):(x_{1},x_{2})\in U\}, d2d\geq 2 and m2m\geq 2 are positive integers, UU is a small neighborhood of the origin (0,0)(0,0), the smooth function ϕ\phi satisfies ϕ(0)0\phi(0)\neq 0, ϕ(j)(0)=0\phi^{(j)}(0)=0 for 1jn11\leq j\leq n-1 and ϕ(n)(0)0\phi^{(n)}(0)\not=0, n2n\geq 2. Let Ψ(x1,x2)=(x2x1dϕ(x1))m\Psi(x_{1},x_{2})=(x_{2}-x_{1}^{d}\phi(x_{1}))^{m}. We have the following result.

Theorem 1.6.

Let AtA_{t} be defined by equality (1.1). Then for each p>max{3/2,2d/(d+1)}p>\max\{3/2,2d/(d+1)\}, there exists Cp>0C_{p}>0 such that supt>0|At|Lp(3)Lp(3)Cp\|\sup_{t>0}|A_{t}|\|_{L^{p}(\mathbb{R}^{3})\rightarrow L^{p}(\mathbb{R}^{3})}\leq C_{p}. Moreover, supt>0|At|\sup_{t>0}|A_{t}| cannot be Lp(3)L^{p}(\mathbb{R}^{3})-bounded if p<max{3/2,2d/(d+1)}p<\max\{3/2,2d/(d+1)\}.

Theorem 1.6 is sharp up to the endpoints, it will be proved in Section 5 below. From Theorem 1.6, we observe that the bounded exponents for such maximal operators depend heavily on the type of the curve determined by the level set {(x1,x2):Ψ(x1,x2)=1}\{(x_{1},x_{2}):\Psi(x_{1},x_{2})=1\}. If we replace the term x1dϕ(x1)x_{1}^{d}\phi(x_{1}) in Ψ(x1,x2)=(x2x1dϕ(x1))m\Psi(x_{1},x_{2})=(x_{2}-x_{1}^{d}\phi(x_{1}))^{m} by γ(x1)\gamma(x_{1}), which is smooth and satisfies

γ(j)(0)=0 for each j,\gamma^{(j)}(0)=0\text{ for each }j\in\mathbb{N}, (1.5)

then the level set {(x1,x2):Ψ(x1,x2)=1}\{(x_{1},x_{2}):\Psi(x_{1},x_{2})=1\} determines a flat curve near x1=0x_{1}=0. We can get the following theorem, whose proof can also be found in Section 5.

Theorem 1.7.

There exists a real-valued smooth function γ\gamma defined on \mathbb{R} that satisfies (1.5) such that the maximal operator supt>0|At|\sup_{t>0}|A_{t}| is not Lp(3)L^{p}(\mathbb{R}^{3})-bounded if p<2p<2 for any neighborhood of the origin, where AtA_{t} is given by

Atf(y):=Uf(yt(x1,x2,(x2γ(x1))m))dx1dx2.A_{t}f(y):=\int_{U}f\bigl{(}y-t(x_{1},x_{2},(x_{2}-\gamma(x_{1}))^{m})\bigl{)}dx_{1}dx_{2}. (1.6)

Conventions: Throughout this article, we shall use the notation ABA\ll B, which means that there is a sufficiently large constant GG, which is much larger than 11 and does not depend on the relevant parameters arising in the context in which the quantities AA and BB appear, such that GABGA\leq B. By ABA\lesssim B we mean that ACBA\leq CB for some constant CC independent of the parameters related to AA and BB. Given n\mathbb{R}^{n}, we write B(0,1)B(0,1) instead of the unit ball Bn(0,1)B^{n}(0,1) in n\mathbb{R}^{n} centered at the origin for short, and the same notation is valid for B(x0,r)B(x_{0},r).

2 Proof of Theorem 1.4

It suffices to establish the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})-bounded estimate of supt[1,2]|At,jf|\sup_{t\in[1,2]}|A_{t,j}f| when j1j\geq 1, where

At,jf(y)=3eiyξ+itcξ3dμ^(tξ)β(2j|ξ|)f^(ξ)𝑑ξ,j1,A_{t,j}f(y)=\int_{\mathbb{R}^{3}}e^{iy\cdot\xi+itc\xi_{3}}\widehat{d\mu}(t\xi)\beta(2^{-j}|\xi|)\hat{f}(\xi)d\xi,\quad\quad j\geq 1, (2.1)

where the cutoff function β\beta is supported in [1,2][1,2].

We choose a smooth cutoff function χ~\tilde{\chi} which is supported in [1/2,4][1/2,4], and χ~=1\tilde{\chi}=1 on [1,2][1,2], and decompose

At,jf(y)=At,j~f(y)+Rt,jf(y),A_{t,j}f(y)=\widetilde{A_{t,j}}f(y)+R_{t,j}f(y), (2.2)

where

At,j~f(y)=3eiyξ+itcξ3dμ^(tξ)β(2j|ξ|)χ~(|ξ1|+|ξ2||ξ3|)f^(ξ)dξ.\widetilde{A_{t,j}}f(y)=\int_{\mathbb{R}^{3}}e^{iy\cdot\xi+itc\xi_{3}}\widehat{d\mu}(t\xi)\beta(2^{-j}|\xi|)\tilde{\chi}\biggl{(}\frac{|\xi_{1}|+|\xi_{2}|}{|\xi_{3}|}\biggl{)}\hat{f}(\xi)d\xi.

We first consider the remainder term Rt,jfR_{t,j}f. Denote

ϕt,ξ(r,s)=tr(1+sMg(s))ξ1+trsξ2+trdξ3.\phi_{t,\xi}(r,s)=tr(1+s^{M}g(s))\xi_{1}+trs\xi_{2}+tr^{d}\xi_{3}.

Then

r,sϕt,ξ(r,s)=(tsξ2+t(1+sMg(s))ξ1+dtrd1ξ3,rtξ2+rt(MsM1g(s)+sMg(s))ξ1).\nabla_{r,s}\phi_{t,\xi}(r,s)=(ts\xi_{2}+t(1+s^{M}g(s))\xi_{1}+dtr^{d-1}\xi_{3},rt\xi_{2}+rt(Ms^{M-1}g(s)+s^{M}g^{\prime}(s))\xi_{1}).

It is clear that |r,sϕt,ξ(r,s)||ξ||\nabla_{r,s}\phi_{t,\xi}(r,s)|\geq|\xi| provided that |ξ3||ξ1|+|ξ2||\xi_{3}|\gg|\xi_{1}|+|\xi_{2}| or |ξ3||ξ1|+|ξ2||\xi_{3}|\ll|\xi_{1}|+|\xi_{2}|, and there holds

|ξαdμ^(tξ)|CN,α(1+t|ξ|)N+|α||\nabla_{\xi}^{\alpha}\widehat{d\mu}(t\xi)|\leq\frac{C_{N,\alpha}}{(1+t|\xi|)^{N+|\alpha|}}

for any multi-index α\alpha. Then integrating by parts implies that

|Rt,jf(y)|2Nj3CN(1+2j|yz(0,0,ct)|)N|f(z)|𝑑z|R_{t,j}f(y)|\leq 2^{-Nj}\int_{\mathbb{R}^{3}}\frac{C_{N}}{(1+2^{j}|y-z-(0,0,ct)|)^{N}}|f(z)|dz

for any positive integer N1N\geq 1. Therefore, Sobolev embedding and Young’s inequality yield that

supt[1,2]|Rt,jf|Lq(3)max{1,c1q}2jNfLp(3)\biggl{\|}\sup_{t\in[1,2]}|R_{t,j}f|\biggl{\|}_{L^{q}(\mathbb{R}^{3})}\lesssim\max\{1,c^{\frac{1}{q}}\}2^{-jN}\|f\|_{L^{p}(\mathbb{R}^{3})} (2.3)

whenever qp1q\geq p\geq 1. So we are reduced to investigating the main contribution term At,j~f\widetilde{A_{t,j}}f.

A dyadic decomposition of the ss-variable and a standard scaling argument imply that

supt[1,2]|At,j~|Lp(3)Lq(3)llog1ϵ02l2(M1)l(1q1p)supt[1,2]|At,j,l~|Lp(3)Lq(3),\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,j}}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\leq\sum_{l\geq\log{\frac{1}{\epsilon_{0}}}}2^{-l}2^{(M-1)l(\frac{1}{q}-\frac{1}{p})}\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,j,l}}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}, (2.4)

where

At,j,l~f(y)=3eiyξ+itcξ3dμM^(tξ)β(|ξ1|+2(M1)l|ξ2|+|ξ3|2j)χ~(|ξ1|+2(M1)l|ξ2||ξ3|)f^(ξ)dξ,\widetilde{A_{t,j,l}}f(y)=\int_{\mathbb{R}^{3}}e^{iy\cdot\xi+itc\xi_{3}}\widehat{d\mu_{M}}(t\xi)\beta\biggl{(}\frac{|\xi_{1}|+2^{-(M-1)l}|\xi_{2}|+|\xi_{3}|}{2^{j}}\biggl{)}\tilde{\chi}\biggl{(}\frac{|\xi_{1}|+2^{-(M-1)l}|\xi_{2}|}{|\xi_{3}|}\biggl{)}\hat{f}(\xi)d\xi,

and

dμM^(tξ)=eitrdξ3itrξ1eitr2Ml[sMg(2ls)ξ1+sξ2]ψ~(s)𝑑sρ(r)𝑑r,\widehat{d\mu_{M}}(t\xi)=\int_{\mathbb{R}}e^{-itr^{d}\xi_{3}-itr\xi_{1}}\int_{\mathbb{R}}e^{-itr2^{-Ml}[s^{M}g(2^{-l}s)\xi_{1}+s\xi_{2}]}\tilde{\psi}(s)ds\rho(r)dr,

the cutoff function ψ~(s):=χ~(s)ψ(2ls)\tilde{\psi}(s):=\tilde{\chi}(s)\psi(2^{-l}s) is supported in [1,2][1,2].

When jMlj\leq Ml and |ξ2|>2Ml|\xi_{2}|>2^{Ml}, integration by parts with respect to the variable ss in the inner integral of dμM^(tξ)\widehat{d\mu_{M}}(t\xi) implies that

|ξαdμM^(tξ)|CN,α(1+2Ml|ξ2|)N+|α|,\bigl{|}\nabla_{\xi}^{\alpha}\widehat{d\mu_{M}}(t\xi)\bigl{|}\lesssim\frac{C_{N,\alpha}}{(1+2^{-Ml}|\xi_{2}|)^{N+|\alpha|}},

for each non-negative multi-index α\alpha and non-negative integer NN, then the method of handling the remainder term Rt,jR_{t,j} can be applied here to complete the proof. Next we consider the case when jMlj\leq Ml and |ξ2|2Ml|\xi_{2}|\leq 2^{Ml}. Notice that in this case, the inner integral in dμM^(tξ)\widehat{d\mu_{M}}(t\xi) can be put into the amplitude. If |ξ3||ξ1||\xi_{3}|\gg|\xi_{1}|, for any tt and rr that belong to the interval [1,2][1,2], there holds

|r(trdξ3+trξ1)|=t|ξ3||drd1+ξ1ξ3|d2|ξ3|.|\partial_{r}(tr^{d}\xi_{3}+tr\xi_{1})|=t|\xi_{3}|\biggl{|}dr^{d-1}+\frac{\xi_{1}}{\xi_{3}}\biggl{|}\geq\frac{d}{2}|\xi_{3}|.

By performing integration by parts on the variable rr, we obtain

|ξαdμM^(tξ)|CN,α(1+|ξ3|)N+|α||\nabla_{\xi}^{\alpha}\widehat{d\mu_{M}}(t\xi)|\lesssim\frac{C_{N,\alpha}}{(1+|\xi_{3}|)^{N+|\alpha|}} (2.5)

for any positive integer NN and multi-index α\alpha. By employing the method of handling the remainder term Rt,jR_{t,j}, we can complete the proof when |ξ3||ξ1||\xi_{3}|\gg|\xi_{1}|. When |ξ3||ξ1||\xi_{3}|\sim|\xi_{1}|, there exists a solution rc=(ξ1dξ3)1/(d1)r_{c}=(-\frac{\xi_{1}}{d\xi_{3}})^{1/(d-1)} for the equation r(rd+rξ1ξ3)=0\partial_{r}(r^{d}+r\frac{\xi_{1}}{\xi_{3}})=0 provided that ξ1ξ3[d,d2d1]-\frac{\xi_{1}}{\xi_{3}}\in[d,d2^{d-1}]. Then by the method of stationary phase, we only need to consider the main contribution term given by

At,j,l,1~f(y)=2j23eiyξ+itcξ3+itcdξ3(ξ1ξ3)d/(d1)aj,l,1(t,ξ)f^(ξ)β0(2Ml|ξ2|)𝑑ξ,\widetilde{A_{t,j,l,1}}f(y)=2^{-\frac{j}{2}}\int_{\mathbb{R}^{3}}e^{iy\cdot\xi+itc\xi_{3}+itc_{d}\xi_{3}(-\frac{\xi_{1}}{\xi_{3}})^{d/(d-1)}}a_{j,l,1}(t,\xi)\hat{f}(\xi)\beta_{0}(2^{-Ml}|\xi_{2}|)d\xi,

where β0\beta_{0} is a smooth cutoff function supported in [1,1][-1,1],

aj,l,1(t,ξ)=2j/2(1+t|ξ3|)1/2β(|ξ1|+2(M1)l|ξ2|+|ξ3|2j)χ~(|ξ1|+2(M1)l|ξ2||ξ3|)χ~(ξ1(d1)ξ3),a_{j,l,1}(t,\xi)=\frac{2^{j/2}}{(1+t|\xi_{3}|)^{1/2}}\beta\biggl{(}\frac{|\xi_{1}|+2^{-(M-1)l}|\xi_{2}|+|\xi_{3}|}{2^{j}}\biggl{)}\tilde{\chi}\biggl{(}\frac{|\xi_{1}|+2^{-(M-1)l}|\xi_{2}|}{|\xi_{3}|}\biggl{)}\tilde{\chi}(-\frac{\xi_{1}}{(d-1)\xi_{3}}),

which is a symbol of order zero in ξ\xi. We will deal with this main contribution term later.

Next, we will consider the case when j>Mlj>Ml. According to the definition of At,j,l~\widetilde{A_{t,j,l}}, we have |ξ1|+2(M1)l|ξ2||ξ3||\xi_{1}|+2^{-(M-1)l}|\xi_{2}|\sim|\xi_{3}|. If |ξ3||ξ1||\xi_{3}|\gg|\xi_{1}|, for any tt, rr and ss that belong to the interval [1,2][1,2], we have

|r[trdξ3+trξ1+t2Mlr(sMg(2lx)ξ1+sξ2)]|d2|ξ3|.\biggl{|}\partial_{r}[tr^{d}\xi_{3}+tr\xi_{1}+t2^{-Ml}r(s^{M}g(2^{-l}x)\xi_{1}+s\xi_{2})]\biggl{|}\geq\frac{d}{2}|\xi_{3}|.

Therefore, we still maintain the validity of inequality (2.5). Furthermore, if |ξ3||ξ1||ξ2||\xi_{3}|\sim|\xi_{1}|\gg|\xi_{2}| or |ξ3||ξ1||ξ2||\xi_{3}|\sim|\xi_{1}|\ll|\xi_{2}|, there holds the uniform estimate

|s[2Mltr(sMg(2ls)ξ1+sξ2)]|M2Ml|ξ3|\biggl{|}\partial_{s}[2^{-Ml}tr(s^{M}g(2^{-l}s)\xi_{1}+s\xi_{2})]\biggl{|}\gtrsim_{M}2^{-Ml}|\xi_{3}|

for every tt, rr and ss that belong to the interval [1,2][1,2]. By utilizing integration by parts in the inner integral of dμM^(tξ)\widehat{d\mu_{M}}(t\xi), we get

|ξαdμM^(tξ)|CN,α(1+2Ml|ξ3|)N+|α||\nabla_{\xi}^{\alpha}\widehat{d\mu_{M}}(t\xi)|\lesssim\frac{C_{N,\alpha}}{(1+2^{-Ml}|\xi_{3}|)^{N+|\alpha|}} (2.6)

for any positive integer NN and multi-index α\alpha. Since j>Mlj>Ml, we can also finish the proof when |ξ3||ξ1||ξ2||\xi_{3}|\sim|\xi_{1}|\gg|\xi_{2}| or |ξ3||ξ1||ξ2||\xi_{3}|\sim|\xi_{1}|\ll|\xi_{2}| as in the similar proof for the remainder term Rt,jR_{t,j}.

Now we turn to consider the subcase when |ξ3||ξ1||ξ2||\xi_{3}|\sim|\xi_{1}|\sim|\xi_{2}|. Firstly, we treat the inner integral of dμM^(tξ)\widehat{d\mu_{M}}(t\xi). Since ll is sufficiently large and g(0)0g(0)\neq 0, then the equation s[sMg(2ls)+sξ2ξ1]=0\partial_{s}[s^{M}g(2^{-l}s)+s\frac{\xi_{2}}{\xi_{1}}]=0 has a unique solution according to the implicit function theorem. Applying the method of stationary phase to the ss-integration, the inner integral in dμM^(tξ)\widehat{d\mu_{M}}(t\xi) can be replaced by its leading term, whose phase function can be written as i2Mltrξ1Ψ~2(ξ2ξ1)-i2^{-Ml}tr\xi_{1}\widetilde{\Psi}_{2}(\frac{\xi_{2}}{\xi_{1}}). Here,

Ψ~2(ξ2ξ1)=cM(ξ2ξ1)M/(M1)+2lR~(ξ2ξ1),\widetilde{\Psi}_{2}(\frac{\xi_{2}}{\xi_{1}})=c_{M}\biggl{(}-\frac{\xi_{2}}{\xi_{1}}\biggl{)}^{M/(M-1)}+2^{-l}\tilde{R}\biggl{(}\frac{\xi_{2}}{\xi_{1}}\biggl{)},

which is a small perturbation of cM(ξ2/ξ1)M/(M1)c_{M}(-\xi_{2}/\xi_{1})^{M/(M-1)}. Secondly, we apply the method of stationary phase to the rr-integration, the phase function of which is now given by

itrdξ3itrξ1(1+2MlΨ~2(ξ2ξ1)).-itr^{d}\xi_{3}-itr\xi_{1}(1+2^{-Ml}\widetilde{\Psi}_{2}(\frac{\xi_{2}}{\xi_{1}})).

The equation

r[rd+rξ1ξ3(1+2MlΨ~2(ξ2ξ1))]=0\partial_{r}\biggl{[}r^{d}+r\frac{\xi_{1}}{\xi_{3}}\biggl{(}1+2^{-Ml}\widetilde{\Psi}_{2}(\frac{\xi_{2}}{\xi_{1}})\biggl{)}\biggl{]}=0

has a unique solution since ll is sufficiently large, and by Taylor’s expansion, the phase function can be written as

itξ3[Ψ1(ξ1ξ3)+2MlΨ2(ξ1ξ3,ξ2ξ1)],-it\xi_{3}\biggl{[}\Psi_{1}(\frac{\xi_{1}}{\xi_{3}})+2^{-Ml}\Psi_{2}(\frac{\xi_{1}}{\xi_{3}},\frac{\xi_{2}}{\xi_{1}})\biggl{]},

where Ψ1(ξ1ξ3)=cd,M,l(ξ1ξ3)d/(d1),\Psi_{1}(\frac{\xi_{1}}{\xi_{3}})=c_{d,M,l}(-\frac{\xi_{1}}{\xi_{3}})^{d/(d-1)}, and

Ψ2(ξ1ξ3,ξ2ξ1)=cd,M,l(ξ1ξ3)d/(d1)(ξ2ξ1)M/(M1)+2lR(ξ1ξ3,ξ2ξ1).\Psi_{2}(\frac{\xi_{1}}{\xi_{3}},\frac{\xi_{2}}{\xi_{1}})=c_{d,M,l}(-\frac{\xi_{1}}{\xi_{3}})^{d/(d-1)}(-\frac{\xi_{2}}{\xi_{1}})^{M/(M-1)}+2^{-l}R(\frac{\xi_{1}}{\xi_{3}},\frac{\xi_{2}}{\xi_{1}}).

Indeed, Ψ2(ξ1ξ3,ξ2ξ1)\Psi_{2}(\frac{\xi_{1}}{\xi_{3}},\frac{\xi_{2}}{\xi_{1}}) is a small perturbation of cd,M,l(ξ1ξ3)d/(d1)(ξ2ξ1)M/(M1)c_{d,M,l}(-\frac{\xi_{1}}{\xi_{3}})^{d/(d-1)}(-\frac{\xi_{2}}{\xi_{1}})^{M/(M-1)}.

Therefore, when j>Mlj>Ml, we reduce our problem to estimating the main contribution term

At,j,l,2~f(y)=2j+Ml23eiyξ+itcξ3+itξ3Ψ1(ξ1ξ3)+it2Mlξ3Ψ2(ξ1ξ3,ξ2ξ1)aj,l,2(t,ξ)f^(ξ)𝑑ξ,\displaystyle\widetilde{A_{t,j,l,2}}f(y)=2^{-j+\frac{Ml}{2}}\int_{\mathbb{R}^{3}}e^{iy\cdot\xi+itc\xi_{3}+it\xi_{3}\Psi_{1}(\frac{\xi_{1}}{\xi_{3}})+it2^{-Ml}\xi_{3}\Psi_{2}(\frac{\xi_{1}}{\xi_{3}},\frac{\xi_{2}}{\xi_{1}})}a_{j,l,2}(t,\xi)\hat{f}(\xi)d\xi,

with

aj,l,2(t,ξ)\displaystyle a_{j,l,2}(t,\xi) =2jMl2(1+t|ξ3|)1/2(1+t2Ml|ξ2|)1/2β(|ξ1|+2(M1)l|ξ2|+|ξ3|2j)χ~(|ξ1|+2(M1)l|ξ2||ξ3|)\displaystyle=\frac{2^{j-\frac{Ml}{2}}}{(1+t|\xi_{3}|)^{1/2}(1+t2^{-Ml}|\xi_{2}|)^{1/2}}\beta\biggl{(}\frac{|\xi_{1}|+2^{-(M-1)l}|\xi_{2}|+|\xi_{3}|}{2^{j}}\biggl{)}\tilde{\chi}\biggl{(}\frac{|\xi_{1}|+2^{-(M-1)l}|\xi_{2}|}{|\xi_{3}|}\biggl{)}
×χd,l,M~(ξ1ξ3)χl,M~(ξ2ξ1).\displaystyle\times\widetilde{\chi_{d,l,M}}(-\frac{\xi_{1}}{\xi_{3}})\widetilde{\chi_{l,M}}(-\frac{\xi_{2}}{\xi_{1}}). (2.7)

Here, the support of the cutoff functions χd,l,M~\widetilde{\chi_{d,l,M}} and χl,M~\widetilde{\chi_{l,M}} are uniformly contained in [1,2][1,2] for each d,l,Md,l,M.

Based on the above stationary phase method, by Sobolev embedding, there holds

supt[1,2]|At,j,l,1~f|Lq(3)max{1,c1q}2jqj2ρ~(t)Fj,l,1(y,t)Lq(3×[12,4]),\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,j,l,1}}f|\biggl{\|}_{L^{q}(\mathbb{R}^{3})}\lesssim\max\{1,c^{\frac{1}{q}}\}2^{\frac{j}{q}-\frac{j}{2}}\biggl{\|}\tilde{\rho}(t)F_{j,l,1}(y,t)\biggl{\|}_{L^{q}(\mathbb{R}^{3}\times[\frac{1}{2},4])}, (2.8)

where the bump function ρ~\tilde{\rho} is supported in [1/2,4][1/2,4] and the Fourier integral operator

Fj,l,1(y,t)=3eiyξ+itcdξ3(ξ1ξ3)d/(d1)aj,l,1(t,ξ)f^(ξ)β0(2Ml|ξ2|)𝑑ξ.F_{j,l,1}(y,t)=\int_{\mathbb{R}^{3}}e^{iy\cdot\xi+itc_{d}\xi_{3}(-\frac{\xi_{1}}{\xi_{3}})^{d/(d-1)}}a_{j,l,1}(t,\xi)\hat{f}(\xi)\beta_{0}(2^{-Ml}|\xi_{2}|)d\xi.

For the same reason, we have

supt[1,2]|At,j,l,2~f|Lq(3)max{1,c1q}2jqj+Ml2ρ~(t)Fj,l,2f(y,t)Lq(3×[12,4])\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,j,l,2}}f|\biggl{\|}_{L^{q}(\mathbb{R}^{3})}\lesssim\max\{1,c^{\frac{1}{q}}\}2^{\frac{j}{q}-j+\frac{Ml}{2}}\biggl{\|}\tilde{\rho}(t)F_{j,l,2}f(y,t)\biggl{\|}_{L^{q}(\mathbb{R}^{3}\times[\frac{1}{2},4])} (2.9)

with the Fourier integral operator

Fj,l,2f(y,t)=3eiyξ+itξ3Ψ1(ξ1ξ3)+it2Mlξ3Ψ2(ξ1ξ3,ξ2ξ1)aj,l,2(t,ξ)f^(ξ)𝑑ξ.\displaystyle F_{j,l,2}f(y,t)=\int_{\mathbb{R}^{3}}e^{iy\cdot\xi+it\xi_{3}\Psi_{1}(\frac{\xi_{1}}{\xi_{3}})+it2^{-Ml}\xi_{3}\Psi_{2}(\frac{\xi_{1}}{\xi_{3}},\frac{\xi_{2}}{\xi_{1}})}a_{j,l,2}(t,\xi)\hat{f}(\xi)d\xi. (2.10)

Theorem 1.4 can be proved by the following Lp(3)Lq(3×[1/2,4])L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}\times[1/2,4]) estimates for Fj,l,nF_{j,l,n}, n=1,2n=1,2.

Theorem 2.1.

There holds the following estimates for Fj,l,nF_{j,l,n}, n=1,2n=1,2.
(E1) For each j1j\geq 1, n=1,2n=1,2,

ρ~(t)Fj,l,nf(y,t)L2(3×[12,4])fL2(3).\biggl{\|}\tilde{\rho}(t)F_{j,l,n}f(y,t)\biggl{\|}_{L^{2}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim\|f\|_{L^{2}(\mathbb{R}^{3})}. (2.11)

(E2) For each 1jMl1\leq j\leq Ml,

ρ~(t)Fj,l,1f(y,t)L6(3×[12,4])2j2+Ml3fL2(3),\biggl{\|}\tilde{\rho}(t)F_{j,l,1}f(y,t)\biggl{\|}_{L^{6}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim 2^{\frac{j}{2}+\frac{Ml}{3}}\|f\|_{L^{2}(\mathbb{R}^{3})}, (2.12)

and for each jMlj\geq Ml,

ρ~(t)Fj,l,2f(y,t)L6(3×[12,4])25j6fL2(3).\biggl{\|}\tilde{\rho}(t)F_{j,l,2}f(y,t)\biggl{\|}_{L^{6}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim 2^{\frac{5j}{6}}\|f\|_{L^{2}(\mathbb{R}^{3})}. (2.13)

(E3) For each 1jMl1\leq j\leq Ml,

ρ~(t)Fj,l,1(y,t)L4(3×[12,4])23j8+Ml4fL2(3),\biggl{\|}\tilde{\rho}(t)F_{j,l,1}(y,t)\biggl{\|}_{L^{4}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim 2^{\frac{3j}{8}+\frac{Ml}{4}}\|f\|_{L^{2}(\mathbb{R}^{3})}, (2.14)

and for each j>Mlj>Ml,

ρ~(t)Fj,l,2(y,t)L4(3×[12,4])2j2+Ml8fL2(3).\biggl{\|}\tilde{\rho}(t)F_{j,l,2}(y,t)\biggl{\|}_{L^{4}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim 2^{\frac{j}{2}+\frac{Ml}{8}}\|f\|_{L^{2}(\mathbb{R}^{3})}. (2.15)

(E4) For each 1jMl1\leq j\leq Ml,

ρ~(t)Fj,l,1(y,t)L(3×[12,4])2j2fL(3),\biggl{\|}\tilde{\rho}(t)F_{j,l,1}(y,t)\biggl{\|}_{L^{\infty}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim 2^{\frac{j}{2}}\|f\|_{L^{\infty}(\mathbb{R}^{3})}, (2.16)

and for each j>Mlj>Ml,

ρ~(t)Fj,l,2(y,t)L(3×[12,4])2jMl2fL(3).\biggl{\|}\tilde{\rho}(t)F_{j,l,2}(y,t)\biggl{\|}_{L^{\infty}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim 2^{j-\frac{Ml}{2}}\|f\|_{L^{\infty}(\mathbb{R}^{3})}. (2.17)

(E5) For each 1jMl1\leq j\leq Ml,

ρ~(t)Fj,l,1(y,t)L1(3×[12,4])2j2fL1(3),\biggl{\|}\tilde{\rho}(t)F_{j,l,1}(y,t)\biggl{\|}_{L^{1}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim 2^{\frac{j}{2}}\|f\|_{L^{1}(\mathbb{R}^{3})}, (2.18)

and for each j>Mlj>Ml,

ρ~(t)Fj,l,2(y,t)L1(3×[12,4])2jMl2fL1(3).\biggl{\|}\tilde{\rho}(t)F_{j,l,2}(y,t)\biggl{\|}_{L^{1}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim 2^{j-\frac{Ml}{2}}\|f\|_{L^{1}(\mathbb{R}^{3})}. (2.19)

(E6) For each 1jMl1\leq j\leq Ml,

ρ~(t)Fj,l,1(y,t)L(3×[12,4])23j2+MlfL1(3),\biggl{\|}\tilde{\rho}(t)F_{j,l,1}(y,t)\biggl{\|}_{L^{\infty}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim 2^{\frac{3j}{2}+Ml}\|f\|_{L^{1}(\mathbb{R}^{3})}, (2.20)

and for each j>Mlj>Ml,

ρ~(t)Fj,l,2(y,t)L(3×[12,4])22j+Ml2fL1(3).\biggl{\|}\tilde{\rho}(t)F_{j,l,2}(y,t)\biggl{\|}_{L^{\infty}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim 2^{2j+\frac{Ml}{2}}\|f\|_{L^{1}(\mathbb{R}^{3})}. (2.21)

Now we apply Theorem 2.1 to prove Theorem 1.4, and leave the proof of Theorem 2.1 to the next section. According to inequalities (2.4), (2.8) and (2.9), there holds

supt[1,2]|At,j~f|Lq(3)\displaystyle\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,j}}f|\biggl{\|}_{L^{q}(\mathbb{R}^{3})} max{1,c1q}(ljM2(M1)l(1q1p)l2jqj2ρ~(t)Fj,l,1(y,t)Lq(3×[12,4])\displaystyle\leq\max\{1,c^{\frac{1}{q}}\}\biggl{(}\sum_{l\geq\frac{j}{M}}2^{(M-1)l(\frac{1}{q}-\frac{1}{p})-l}2^{\frac{j}{q}-\frac{j}{2}}\biggl{\|}\tilde{\rho}(t)F_{j,l,1}(y,t)\biggl{\|}_{L^{q}(\mathbb{R}^{3}\times[\frac{1}{2},4])}
+log1ϵ0l<jM2(M1)l(1q1p)l2jqj+Ml2ρ~(t)Fj,l,2(y,t)Lq(3×[12,4])).\displaystyle\quad\quad+\sum_{\log{\frac{1}{\epsilon_{0}}}\leq l<\frac{j}{M}}2^{(M-1)l(\frac{1}{q}-\frac{1}{p})-l}2^{\frac{j}{q}-j+\frac{Ml}{2}}\biggl{\|}\tilde{\rho}(t)F_{j,l,2}(y,t)\biggl{\|}_{L^{q}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\biggl{)}. (2.22)

Firstly, we show the following diagonal results for M3M\geq 3. By the estimate (E1) in Theorem 2.1, there holds

supt[1,2]|At,j~f|L2(3)max{1,c12}2jMfL2(3).\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,j}}f|\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim\max\{1,c^{\frac{1}{2}}\}2^{-\frac{j}{M}}\|f\|_{L^{2}(\mathbb{R}^{3})}. (2.23)

By the estimates (E4) in Theorem 2.1, we have

supt[1,2]|At,j~f|L(3)fL(3).\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,j}}f|\biggl{\|}_{L^{\infty}(\mathbb{R}^{3})}\lesssim\|f\|_{L^{\infty}(\mathbb{R}^{3})}.

Interpolation between the estimates in (E1) and (E5) implies

supt[1,2]|At,j~f|L2MM+1(3)\displaystyle\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,j}}f|\biggl{\|}_{L^{\frac{2M}{M+1}}(\mathbb{R}^{3})}
max{1,cM+12M}(ljM2l2(M+1)j2Mj2+j2M+log1ϵ0l<jM2l2(M+1)j2Mj+Ml22jMl2)fL2MM+1(3)\displaystyle\leq\max\{1,c^{\frac{M+1}{2M}}\}\biggl{(}\sum_{l\geq\frac{j}{M}}2^{-l}2^{\frac{(M+1)j}{2M}-\frac{j}{2}+\frac{j}{2M}}+\sum_{\log{\frac{1}{\epsilon_{0}}}\leq l<\frac{j}{M}}2^{-l}2^{\frac{(M+1)j}{2M}-j+\frac{Ml}{2}}2^{\frac{j}{M}-\frac{l}{2}}\biggl{)}\|f\|_{L^{\frac{2M}{M+1}}(\mathbb{R}^{3})}
max{1,cM+12M}fL2MM+1(3).\displaystyle\lesssim\max\{1,c^{\frac{M+1}{2M}}\}\|f\|_{L^{\frac{2M}{M+1}}(\mathbb{R}^{3})}.

Then we give the following off-diagonal results for (1p,1q)(\frac{1}{p},\frac{1}{q}) satisfying 1p=2q\frac{1}{p}=\frac{2}{q}. When M6M\geq 6, we interpolate the estimates in (E4) with the estimates in (E3) to get

supt[1,2]|At,j~f|LM+22(3)\displaystyle\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,j}}f|\biggl{\|}_{L^{\frac{M+2}{2}}(\mathbb{R}^{3})}
max{1,c2M+2}(ljM2l22(M1)M+2l22M+2jj228M+2(3j8+Ml4)2M62(M+2)j\displaystyle\leq\max\{1,c^{\frac{2}{M+2}}\}\biggl{(}\sum_{l\geq\frac{j}{M}}2^{-l}2^{-\frac{2(M-1)}{M+2}l}2^{\frac{2}{M+2}j-\frac{j}{2}}2^{\frac{8}{M+2}(\frac{3j}{8}+\frac{Ml}{4})}2^{\frac{M-6}{2(M+2)}j}
+log1ϵ0l<jM2l22(M1)M+2l22M+2jj+Ml224M+2j2MlM+22(M6)jM+2(M6)Ml2(M+2))fLM+24(3)\displaystyle+\sum_{\log{\frac{1}{\epsilon_{0}}}\leq l<\frac{j}{M}}2^{-l}2^{-\frac{2(M-1)}{M+2}l}2^{\frac{2}{M+2}j-j+\frac{Ml}{2}}2^{\frac{4}{M+2}j}2^{\frac{Ml}{M+2}}2^{\frac{(M-6)j}{M+2}-\frac{(M-6)Ml}{2(M+2)}}\biggl{)}\|f\|_{L^{\frac{M+2}{4}}(\mathbb{R}^{3})}
max{1,c2M+2}fLM+24(3).\displaystyle\lesssim\max\{1,c^{\frac{2}{M+2}}\}\|f\|_{L^{\frac{M+2}{4}}(\mathbb{R}^{3})}.

When 3M53\leq M\leq 5, the following estimates can be easily obtained by interpolation, for each 1jMl1\leq j\leq Ml,

ρ~(t)Fj,l,1(y,t)L3(3×[12,4])2j2+Ml3fL32(3),\biggl{\|}\tilde{\rho}(t)F_{j,l,1}(y,t)\biggl{\|}_{L^{3}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim 2^{\frac{j}{2}+\frac{Ml}{3}}\|f\|_{L^{\frac{3}{2}}(\mathbb{R}^{3})}, (2.24)

and for each j>Mlj>Ml,

ρ~(t)Fj,l,2(y,t)L3(3×[12,4])22j3+Ml6fL32(3),\biggl{\|}\tilde{\rho}(t)F_{j,l,2}(y,t)\biggl{\|}_{L^{3}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim 2^{\frac{2j}{3}+\frac{Ml}{6}}\|f\|_{L^{\frac{3}{2}}(\mathbb{R}^{3})}, (2.25)

then combining the estimate in (E3) and inequality (2), there holds

supt[1,2]|At,j~f|L5M+2M+2(3)\displaystyle\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,j}}f|\biggl{\|}_{L^{\frac{5M+2}{M+2}}(\mathbb{R}^{3})}
max{1,cM+25M+2}(ljM2l2(M+2)(M1)5M+2l2M+25M+2jj223M+65M+2(j2+Ml3)\displaystyle\leq\max\{1,c^{\frac{M+2}{5M+2}}\}\biggl{(}\sum_{l\geq\frac{j}{M}}2^{-l}2^{-\frac{(M+2)(M-1)}{5M+2}l}2^{\frac{M+2}{5M+2}j-\frac{j}{2}}2^{\frac{3M+6}{5M+2}(\frac{j}{2}+\frac{Ml}{3})}
+log1ϵ0ljM2l2(M+2)(M1)5M+2l2M+25M+2jj+Ml222M+45M+2j2(M2+1)Ml5M+2)fL5M+22M+4(3)\displaystyle+\sum_{\log{\frac{1}{\epsilon_{0}}}\leq l\leq\frac{j}{M}}2^{-l}2^{-\frac{(M+2)(M-1)}{5M+2}l}2^{\frac{M+2}{5M+2}j-j+\frac{Ml}{2}}2^{\frac{2M+4}{5M+2}j}2^{\frac{(\frac{M}{2}+1)Ml}{5M+2}}\biggl{)}\|f\|_{L^{\frac{5M+2}{2M+4}}(\mathbb{R}^{3})}
max{1,cM+25M+2}fL5M+22M+4(3).\displaystyle\lesssim\max\{1,c^{\frac{M+2}{5M+2}}\}\|f\|_{L^{\frac{5M+2}{2M+4}}(\mathbb{R}^{3})}.

Furthermore, we show the following off-diagonal estimates for (1p,1q)(\frac{1}{p},\frac{1}{q}) satisfying 1p=3q\frac{1}{p}=\frac{3}{q}. When M4M\geq 4, by interpolating the estimates in (E2) with the estimates in (E4), we get

supt[1,2]|At,j~f|LM+2(3)\displaystyle\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,j}}f|\biggl{\|}_{L^{M+2}(\mathbb{R}^{3})}
max{1,c1M+2}(ljM2l22(M1)M+2l2jM+2j226M+2(j2+Ml3)2M42(M+2)j\displaystyle\leq\max\{1,c^{\frac{1}{M+2}}\}\biggl{(}\sum_{l\geq\frac{j}{M}}2^{-l}2^{-\frac{2(M-1)}{M+2}l}2^{\frac{j}{M+2}-\frac{j}{2}}2^{\frac{6}{M+2}(\frac{j}{2}+\frac{Ml}{3})}2^{\frac{M-4}{2(M+2)}j}
+log1ϵ0l<jM2l22(M1)M+2l2jM+2j+Ml225M+2j2(M4)jM+2(M4)Ml2(M+2))fLM+23(3)\displaystyle+\sum_{\log{\frac{1}{\epsilon_{0}}}\leq l<\frac{j}{M}}2^{-l}2^{-\frac{2(M-1)}{M+2}l}2^{\frac{j}{M+2}-j+\frac{Ml}{2}}2^{\frac{5}{M+2}j}2^{\frac{(M-4)j}{M+2}-\frac{(M-4)Ml}{2(M+2)}}\biggl{)}\|f\|_{L^{\frac{M+2}{3}}(\mathbb{R}^{3})}
max{1,c1M+2}jfLM+23(3).\displaystyle\lesssim\max\{1,c^{\frac{1}{M+2}}\}j\|f\|_{L^{\frac{M+2}{3}}(\mathbb{R}^{3})}.

In order to obtain the results for the case M=3M=3, we give the following easy estimates by interpolation, for each 1jMl1\leq j\leq Ml,

ρ~(t)Fj,l,1(y,t)L5(3×[12,4])23j5+2Ml5fL53(3),\biggl{\|}\tilde{\rho}(t)F_{j,l,1}(y,t)\biggl{\|}_{L^{5}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim 2^{\frac{3j}{5}+\frac{2Ml}{5}}\|f\|_{L^{\frac{5}{3}}(\mathbb{R}^{3})}, (2.26)

and for each j>Mlj>Ml,

ρ~(t)Fj,l,2(y,t)L5(3×[12,4])24j5+Ml5fL53(3),\biggl{\|}\tilde{\rho}(t)F_{j,l,2}(y,t)\biggl{\|}_{L^{5}(\mathbb{R}^{3}\times[\frac{1}{2},4])}\lesssim 2^{\frac{4j}{5}+\frac{Ml}{5}}\|f\|_{L^{\frac{5}{3}}(\mathbb{R}^{3})}, (2.27)

then combining the estimates from (E2) and inequality (2), we obtain

supt[1,2]|At,j~f|L8m+4M+2(3)\displaystyle\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,j}}f|\biggl{\|}_{L^{\frac{8m+4}{M+2}}(\mathbb{R}^{3})}
max{1,cM+28M+4}(ljM2l2(2M+4)(M1)8M+4l2M+28M+4jj225M+108M+4(3j5+2Ml5)\displaystyle\leq\max\{1,c^{\frac{M+2}{8M+4}}\}\biggl{(}\sum_{l\geq\frac{j}{M}}2^{-l}2^{-\frac{(2M+4)(M-1)}{8M+4}l}2^{\frac{M+2}{8M+4}j-\frac{j}{2}}2^{\frac{5M+10}{8M+4}(\frac{3j}{5}+\frac{2Ml}{5})}
+log1ϵ0l<jM2l2(2M+4)(M1)8M+4l2M+28M+4jj+Ml227M+28M+4j2Ml(4M)4M+2)fL8M+43M+6(3)\displaystyle+\sum_{\log{\frac{1}{\epsilon_{0}}}\leq l<\frac{j}{M}}2^{-l}2^{-\frac{(2M+4)(M-1)}{8M+4}l}2^{\frac{M+2}{8M+4}j-j+\frac{Ml}{2}}2^{\frac{7M+2}{8M+4}j}2^{\frac{Ml(4-M)}{4M+2}}\biggl{)}\|f\|_{L^{\frac{8M+4}{3M+6}}(\mathbb{R}^{3})}
max{1,cM+28M+4}jfL8M+43M+6(3).\displaystyle\lesssim\max\{1,c^{\frac{M+2}{8M+4}}\}j\|f\|_{L^{\frac{8M+4}{3M+6}}(\mathbb{R}^{3})}.

Finally, by interpolation, it is easy to see that for each (1p,1q)ΔM(\frac{1}{p},\frac{1}{q})\in\Delta_{M}, M3M\geq 3, there exists ϵ(p,q)>0\epsilon(p,q)>0 such that

supt[1,2]|At,j~f|Lq(3)max{1,c1q}2ϵ(p,q)jfLp(3).\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,j}}f|\biggl{\|}_{L^{q}(\mathbb{R}^{3})}\lesssim\max\{1,c^{\frac{1}{q}}\}2^{-\epsilon(p,q)j}\|f\|_{L^{p}(\mathbb{R}^{3})}.

This completes the proof when M3M\geq 3. For the case M=2M=2, since the curve (s,s2g(s)+1)(s,s^{2}g(s)+1) is non-degenerate near s=0s=0, after conducting stationary phase analysis, we are reduced to considering Fourier integral operators related to hypersurfaces with two non-vanishing principal curvatures. Hence, we can get the estimates using a similar argument as in [14]. Then we finish the proof of Theorem 1.4.

3 Proof of Theorem 2.1

It is clear that the L2(3)L2(3×[1/2,4])L^{2}(\mathbb{R}^{3})\rightarrow L^{2}(\mathbb{R}^{3}\times[1/2,4]) estimates in (E1) follows directly from Plancherel’s theorem. So we first consider the L2(3)L6(3×[1/2,4])L^{2}(\mathbb{R}^{3})\rightarrow L^{6}(\mathbb{R}^{3}\times[1/2,4]) estimates in (E2). Notice that for each t[12,4]t\in[\frac{1}{2},4], the hypersurface {(η1,η3,cdη3(η1η3)d/(d1)):2j(η1,η3)suppaj,l,1(,ξ2,t)}\{(\eta_{1},\eta_{3},c_{d}\eta_{3}(-\frac{\eta_{1}}{\eta_{3}})^{d/(d-1)}):2^{j}(\eta_{1},\eta_{3})\in\textmd{supp}\hskip 2.84544pta_{j,l,1}(\cdot,\xi_{2},t)\} and the hypersurface

{(η,η3Ψ1(η1η3)+2Mlη3Ψ2(η1η3,η2η1):2jηsuppaj,l,2(,t)}\{(\eta,\eta_{3}\Psi_{1}(\frac{\eta_{1}}{\eta_{3}})+2^{-Ml}\eta_{3}\Psi_{2}(\frac{\eta_{1}}{\eta_{3}},\frac{\eta_{2}}{\eta_{1}}):2^{j}\eta\in\textmd{supp}\hskip 2.84544pta_{j,l,2}(\cdot,t)\}

for ll large enough have at leas one non-vanishing principal curvature. Then according to [[13], Proposition 3.4], there holds the inequalities (2.12), (2.13) for Fj,l,1F_{j,l,1} and Fj,l,2F_{j,l,2}, respectively.

Now we explain how to establish estimates for the operator Fj,l,1F_{j,l,1} in (E3)-(E5). Interpolating between L2(3)L2(3×[1/2,4])L^{2}(\mathbb{R}^{3})\rightarrow L^{2}(\mathbb{R}^{3}\times[1/2,4]) and L2(3)L6(3×[1/2,4])L^{2}(\mathbb{R}^{3})\rightarrow L^{6}(\mathbb{R}^{3}\times[1/2,4]) estimates for Fj,l,1F_{j,l,1}, then we arrive at inequality (2.14). It is noted that

Fj,l,1f(y,t)=Kj,l,1(,t)f(y),F_{j,l,1}f(y,t)=K_{j,l,1}(\cdot,t)*f(y),

where

Kj,l,1(y,t)=3eiyξ+itcdξ3(ξ1ξ3)d/(d1)aj,l,1(t,ξ)β0(2Ml|ξ2|)𝑑ξ.K_{j,l,1}(y,t)=\int_{\mathbb{R}^{3}}e^{iy\cdot\xi+itc_{d}\xi_{3}(-\frac{\xi_{1}}{\xi_{3}})^{d/(d-1)}}a_{j,l,1}(t,\xi)\beta_{0}(2^{-Ml}|\xi_{2}|)d\xi.

We rewrite Kj,l,1(y,t)K_{j,l,1}(y,t) as

Kj,l,1(y,t)=eiy2ξ2β0(2Ml|ξ2|)2eiy1ξ1+iy3ξ3+itcdξ3(ξ1ξ3)d/(d1)aj,l,1(t,ξ)𝑑ξ1𝑑ξ3𝑑ξ2.K_{j,l,1}(y,t)=\int_{\mathbb{R}}e^{iy_{2}\xi_{2}}\beta_{0}(2^{-Ml}|\xi_{2}|)\int_{\mathbb{R}^{2}}e^{iy_{1}\xi_{1}+iy_{3}\xi_{3}+itc_{d}\xi_{3}(-\frac{\xi_{1}}{\xi_{3}})^{d/(d-1)}}a_{j,l,1}(t,\xi)d\xi_{1}d\xi_{3}d\xi_{2}.

By employing the angular decomposition of the variable ξ1/ξ3\xi_{1}/\xi_{3} as utilized in [Lemma 2.7, [11]] and [Lemma 2.6, [12]], see also the proof in Lemma 3.2 below, we can obtain that for each t[1,2]t\in[1,2] and positive integer NN, there holds

|Kj,l,1(y,t)|CN23j2+Mlκ2j2[1,2](1+2j2|y1+cdtκ1/(d1)|+2Ml|y2|+2j|y3+κy1+cdtκd/(d1)|)N.\displaystyle|K_{j,l,1}(y,t)|\lesssim C_{N}2^{\frac{3j}{2}+Ml}\sum_{\kappa\in 2^{-\frac{j}{2}}\mathbb{Z}\cap[1,2]}(1+2^{\frac{j}{2}}|y_{1}+c_{d}t\kappa^{1/(d-1)}|+2^{Ml}|y_{2}|+2^{j}|y_{3}+\kappa y_{1}+c_{d}t\kappa^{d/(d-1)}|)^{-N}.

It follows that

Kj,l,1(y,t)L1(3)2j2\|K_{j,l,1}(y,t)\|_{L^{1}(\mathbb{R}^{3})}\lesssim 2^{\frac{j}{2}}

and

Kj,l,1(y,t)L(3)23j2+Ml\|K_{j,l,1}(y,t)\|_{L^{\infty}(\mathbb{R}^{3})}\lesssim 2^{\frac{3j}{2}+Ml}

for each t[1,2]t\in[1,2]. Then inequalities (2.16), (2.18), (2.20) can be proved by Young’s inequality, we omit their proofs here.

Next, we proceed to establish the estimates for the operator Fj,l,2F_{j,l,2} in (E4)-(E5). We rewrite the operator Fj,l,2F_{j,l,2} as

Fj,l,2f(y,t)=Kj,l,2(,t)f(y),F_{j,l,2}f(y,t)=K_{j,l,2}(\cdot,t)*f(y),

where

Kj,l,2(y,t)=3eiyξ+itξ3Ψ1(ξ1ξ3)+it2Mlξ3Ψ2(ξ1ξ3,ξ2ξ1)aj,l,2(t,ξ)𝑑ξ.K_{j,l,2}(y,t)=\int_{\mathbb{R}^{3}}e^{iy\cdot\xi+it\xi_{3}\Psi_{1}(\frac{\xi_{1}}{\xi_{3}})+it2^{-Ml}\xi_{3}\Psi_{2}(\frac{\xi_{1}}{\xi_{3}},\frac{\xi_{2}}{\xi_{1}})}a_{j,l,2}(t,\xi)d\xi.

Then ineqalities (2.17) and (2.19) follow from Young’s inequality and part (2) of Theorem 3.1 below. Additionally, the estimate (2.21) follows directly from part (1) of Theorem 3.1.

Finally, we are left with building the desired L2(3)L4(3×[1/2,4])L^{2}(\mathbb{R}^{3})\rightarrow L^{4}(\mathbb{R}^{3}\times[1/2,4]) estimate for the operator Fj,l,2F_{j,l,2} in (E3). We may replace f^(ξ)\hat{f}(\xi) in Fj,l,2fF_{j,l,2}f by f(ξ)f(\xi) according to Plancherel’s theorem, then by the dual method as in the proof of [Proposition 3.4, [13]], inequality (2.15) will follow if we can obtain for arbitrary GL4/3(4)G\in L^{4/3}(\mathbb{R}^{4}),

Fj,l,2Fj,l,2G(z,t)Lz,t4(4)=Kj,l,2~(,t,t)yG(,t)(z)dtLz,t4(4)2j+Ml4GL4/3(4),\biggl{\|}F_{j,l,2}F^{*}_{j,l,2}G(z,t^{\prime})\biggl{\|}_{L_{z,t^{\prime}}^{4}(\mathbb{R}^{4})}=\biggl{\|}\int_{\mathbb{R}}\widetilde{K_{j,l,2}}(\cdot,t,t^{\prime})*_{y}G(\cdot,t)(z)dt\biggl{\|}_{L_{z,t^{\prime}}^{4}(\mathbb{R}^{4})}\lesssim 2^{j+\frac{Ml}{4}}\|G\|_{L^{4/3}(\mathbb{R}^{4})}, (3.1)

where y*_{y} denotes convolution with respect to the variable yy, and

Kj,l,2~(y,t,t)=3eiyξ+i(tt)ξ3Ψ1(ξ1ξ3)+i(tt)2Mlξ3Ψ2(ξ1ξ3,ξ2ξ1)aj,l,2¯(t,ξ)aj,l,2(t,ξ)𝑑ξ.\widetilde{K_{j,l,2}}(y,t,t^{\prime})=\int_{\mathbb{R}^{3}}e^{iy\cdot\xi+i(t^{\prime}-t)\xi_{3}\Psi_{1}(\frac{\xi_{1}}{\xi_{3}})+i(t^{\prime}-t)2^{-Ml}\xi_{3}\Psi_{2}(\frac{\xi_{1}}{\xi_{3}},\frac{\xi_{2}}{\xi_{1}})}\overline{a_{j,l,2}}(t,\xi)a_{j,l,2}(t^{\prime},\xi)d\xi.

For fixed tt, tt^{\prime}, by Plancherel’s theorem, there holds

Kj,l,2~(,t,t)yG(,t)(z)Lz2(3)G(,t)L2(3).\biggl{\|}\widetilde{K_{j,l,2}}(\cdot,t,t^{\prime})*_{y}G(\cdot,t)(z)\biggl{\|}_{L_{z}^{2}(\mathbb{R}^{3})}\lesssim\|G(\cdot,t)\|_{L^{2}(\mathbb{R}^{3})}. (3.2)

Theorem 3.1 below implies that for each ttt\neq t^{\prime}, the following estimate holds true

Kj,l,2~(,t,t)yG(,t)(z)Lz(3)22j+Ml2|tt|1G(,t)L1(3).\biggl{\|}\widetilde{K_{j,l,2}}(\cdot,t,t^{\prime})*_{y}G(\cdot,t)(z)\biggl{\|}_{L_{z}^{\infty}(\mathbb{R}^{3})}\lesssim 2^{2j+\frac{Ml}{2}}|t-t^{\prime}|^{-1}\|G(\cdot,t)\|_{L^{1}(\mathbb{R}^{3})}. (3.3)

Interpolation between (3.2) and (3.3) yields

Kj,l,2~(,t,t)yG(,t)(z)Lz4(3)2j+Ml4|tt|12G(,t)L43(3).\biggl{\|}\widetilde{K_{j,l,2}}(\cdot,t,t^{\prime})*_{y}G(\cdot,t)(z)\biggl{\|}_{L_{z}^{4}(\mathbb{R}^{3})}\lesssim 2^{j+\frac{Ml}{4}}|t-t^{\prime}|^{-\frac{1}{2}}\|G(\cdot,t)\|_{L^{\frac{4}{3}}(\mathbb{R}^{3})}. (3.4)

By combining (3.4) and the Hardy-Littlewood-Sobolev inequality, we obtain

Kj,l,2~(,t,t)yG(,t)(z)dtLz,t4(4)\displaystyle\biggl{\|}\int_{\mathbb{R}}\widetilde{K_{j,l,2}}(\cdot,t,t^{\prime})*_{y}G(\cdot,t)(z)dt\biggl{\|}_{L_{z,t^{\prime}}^{4}(\mathbb{R}^{4})} Kj,l,2~(,t,t)yG(,t)(z)Lz4(3)dtLt4()\displaystyle\leq\biggl{\|}\int_{\mathbb{R}}\biggl{\|}\widetilde{K_{j,l,2}}(\cdot,t,t^{\prime})*_{y}G(\cdot,t)(z)\biggl{\|}_{L_{z}^{4}(\mathbb{R}^{3})}dt\biggl{\|}_{L_{t^{\prime}}^{4}(\mathbb{R})}
2j+Ml4|tt|12G(,t)L43(3)Lt4()\displaystyle\lesssim 2^{j+\frac{Ml}{4}}\biggl{\|}\int_{\mathbb{R}}|t-t^{\prime}|^{-\frac{1}{2}}\|G(\cdot,t)\|_{L^{\frac{4}{3}}(\mathbb{R}^{3})}\biggl{\|}_{L_{t^{\prime}}^{4}(\mathbb{R})}
2j+Ml4G(y,t)Ly,t4/3(4).\displaystyle\lesssim 2^{j+\frac{Ml}{4}}\|G(y,t)\|_{L_{y,t}^{4/3}(\mathbb{R}^{4})}.

Then we arrive at inequality (3.1).

Therefore, in order to complete the proof of Theorem 2.1, we are left to investigate the estimate for the kernel Kj,l,2K_{j,l,2} related to the Fourier integral operator Fj,l,2F_{j,l,2}.

Theorem 3.1.

We have the following estimates for Kj,l,2K_{j,l,2}, which also hold true for Kj,l,2~\widetilde{K_{j,l,2}}.
(1) For fixed y3y\in\mathbb{R}^{3} and t(0,+)t\in(0,+\infty), there holds the pointwise estimate

|Kj,l,2(y,t)|22j+Ml2t1,|K_{j,l,2}(y,t)|\lesssim 2^{2j+\frac{Ml}{2}}t^{-1}, (3.5)

where the implied constant is independent of yy and tt.
(2) For each t[1,2]t\in[1,2], we have the uniform estimate

Kj,l,2(,t)L1(3)2jMl2.\|K_{j,l,2}(\cdot,t)\|_{L^{1}(\mathbb{R}^{3})}\lesssim 2^{j-\frac{Ml}{2}}. (3.6)

Theorem 3.1 follows from Lemma 3.2 below.

Lemma 3.2.

There holds the following pointwise estimates for Kj,l,2K_{j,l,2}.
(1) If t(0,2j]t\in(0,2^{-j}], then for any positive integer NN, there holds

|Kj,l,2(y,t)|CN22jt1(1+2j|y|)N.|K_{j,l,2}(y,t)|\lesssim C_{N}2^{2j}t^{-1}(1+2^{j}|y|)^{-N}. (3.7)

(2) If t(2j,2j+Ml]t\in(2^{-j},2^{-j+Ml}], we have

|Kj,l,2(y,t)|CN22j+Ml2t1×\displaystyle|K_{j,l,2}(y,t)|\lesssim C_{N}2^{2j+\frac{Ml}{2}}t^{-1}\times
κ𝕀1(1+2j2t12|y1+cd,M,ltκ1/(d1)|+2j|y2|+2j|y3+κy1+cd,M,ltκd/(d1)|)N,\displaystyle\sum_{\kappa\in\mathbb{I}_{1}}(1+2^{\frac{j}{2}}t^{-\frac{1}{2}}|y_{1}+c_{d,M,l}t\kappa^{1/(d-1)}|+2^{j}|y_{2}|+2^{j}|y_{3}+\kappa y_{1}+c_{d,M,l}t\kappa^{d/(d-1)}|)^{-N}, (3.8)

where 𝕀1:=(2jt)12[1,2]\mathbb{I}_{1}:=(2^{j}t)^{-\frac{1}{2}}\mathbb{Z}\cap[1,2].
(3) If t(2j+Ml,+)t\in(2^{-j+Ml},+\infty), we obtain

|Kj,l,2(y,t)|CN22j+Ml2t1×\displaystyle|K_{j,l,2}(y,t)|\lesssim C_{N}2^{2j+\frac{Ml}{2}}t^{-1}\times
κ𝕀1κ~𝕀2(1+2j2t12|y1+c1,t(κ,κ~)|+2j+Ml2t12|y2+c2,t(κ,κ~)|+2j|y3+c3,t(y1,y2,κ,κ~)|)N,\displaystyle\sum_{\kappa\in\mathbb{I}_{1}}\sum_{\tilde{\kappa}\in\mathbb{I}_{2}}(1+2^{\frac{j}{2}}t^{-\frac{1}{2}}|y_{1}+c_{1,t}(\kappa,\tilde{\kappa})|+2^{\frac{j+Ml}{2}}t^{-\frac{1}{2}}|y_{2}+c_{2,t}(\kappa,\tilde{\kappa})|+2^{j}|y_{3}+c_{3,t}(y_{1},y_{2},\kappa,\tilde{\kappa})|)^{-N}, (3.9)

where 𝕀2:=(2jMlt)12[1,2]\mathbb{I}_{2}:=(2^{j-Ml}t)^{-\frac{1}{2}}\mathbb{Z}\cap[1,2],

c1,t(κ,κ~)=cd,M,ltκ1d1+cd,M,l2Mltκdd1MM11κ~MM1+2(M+1)lt1R(κ,κ~κ)2(M+1)ltκ~κ22R(κ,κ~κ),c_{1,t}(\kappa,\tilde{\kappa})=c_{d,M,l}t\kappa^{\frac{1}{d-1}}+c_{d,M,l}2^{-Ml}t\kappa^{\frac{d}{d-1}-\frac{M}{M-1}-1}\tilde{\kappa}^{\frac{M}{M-1}}+2^{-(M+1)l}t\partial_{1}R(\kappa,\frac{\tilde{\kappa}}{\kappa})-2^{-(M+1)l}t\frac{\tilde{\kappa}}{\kappa^{2}}\partial_{2}R(\kappa,\frac{\tilde{\kappa}}{\kappa}),
c2,t(κ,κ~)=cd,M,l2Mltκdd1MM1κ~1M1+t2(M+1)l1κ2R(κ,κ~κ),c_{2,t}(\kappa,\tilde{\kappa})=c_{d,M,l}2^{-Ml}t\kappa^{\frac{d}{d-1}-\frac{M}{M-1}}\tilde{\kappa}^{\frac{1}{M-1}}+t2^{-(M+1)l}\frac{1}{\kappa}\partial_{2}R(\kappa,\frac{\tilde{\kappa}}{\kappa}),
c3,t(y1,y2,κ,κ~)=κy1+κ~y2+tΨ1(κ)+t2MlΨ2(κ,κ~κ)+t2(M+1)lR(κ,κ~κ).c_{3,t}(y_{1},y_{2},\kappa,\tilde{\kappa})=\kappa y_{1}+\tilde{\kappa}y_{2}+t\Psi_{1}(\kappa)+t2^{-Ml}\Psi_{2}(\kappa,\frac{\tilde{\kappa}}{\kappa})+t2^{-(M+1)l}R(\kappa,\frac{\tilde{\kappa}}{\kappa}).

We firstly give the proof of Lemma 3.2, and show the proof how Theorem 3.1 can be deduced by Lemma 3.2 at the end of Section 3.

Proof. We first prove the part (1). By rescaling, we have

Kj,l,2(y,t)=23j3ei2jyη+it2jη3Ψ1(η1η3)+it2jMlη3Ψ2(η1η3,η2η1)aj,l,2(t,2jη)𝑑η.K_{j,l,2}(y,t)=2^{3j}\int_{\mathbb{R}^{3}}e^{i2^{j}y\cdot\eta+it2^{j}\eta_{3}\Psi_{1}(\frac{\eta_{1}}{\eta_{3}})+it2^{j-Ml}\eta_{3}\Psi_{2}(\frac{\eta_{1}}{\eta_{3}},\frac{\eta_{2}}{\eta_{1}})}a_{j,l,2}(t,2^{j}\eta)d\eta. (3.10)

Since t(0,2j)t\in(0,2^{-j}), integrating by parts implies that

|Kj,l,2(y,t)|CN23j(1+2j|y|)NCN22jt(1+2j|y|)N.|K_{j,l,2}(y,t)|\leq\frac{C_{N}2^{3j}}{(1+2^{j}|y|)^{N}}\leq\frac{C_{N}2^{2j}}{t(1+2^{j}|y|)^{N}}.

To prove the part (2), we recall the definition of aj,l,2a_{j,l,2} in equality (2), and decompose Kj,l,2K_{j,l,2} given by equality (3.10) as Kj,l,2(y,t)=κ𝕀1Kj,l,2,κ(y,t)K_{j,l,2}(y,t)=\sum_{\kappa\in\mathbb{I}_{1}}K_{j,l,2,\kappa}(y,t), where

Kj,l,2,κ(y,t)=23j3ei2jyη+it2jη3Ψ1(η1η3)+it2jMlη3Ψ2(η1η3,η2η1)aj,l,2,κ(t,2jη)𝑑η,K_{j,l,2,\kappa}(y,t)=2^{3j}\int_{\mathbb{R}^{3}}e^{i2^{j}y\cdot\eta+it2^{j}\eta_{3}\Psi_{1}(\frac{\eta_{1}}{\eta_{3}})+it2^{j-Ml}\eta_{3}\Psi_{2}(\frac{\eta_{1}}{\eta_{3}},\frac{\eta_{2}}{\eta_{1}})}a_{j,l,2,\kappa}(t,2^{j}\eta)d\eta,

and aj,l,2,κ(t,2jη)=aj,l,2(t,2jη)χ~((2jt)12(η1η3κ))a_{j,l,2,\kappa}(t,2^{j}\eta)=a_{j,l,2}(t,2^{j}\eta)\tilde{\chi}\biggl{(}(2^{j}t)^{\frac{1}{2}}(\frac{\eta_{1}}{\eta_{3}}-\kappa)\biggl{)}. For fixed κ\kappa, changing variables η1(2jt)12η1+κη3\eta_{1}\rightarrow(2^{j}t)^{-\frac{1}{2}}\eta_{1}+\kappa\eta_{3}, its phase function becomes

2j2t12y1η1+2jy2η2+2j(y3+κy1)η3\displaystyle 2^{\frac{j}{2}}t^{-\frac{1}{2}}y_{1}\eta_{1}+2^{j}y_{2}\eta_{2}+2^{j}(y_{3}+\kappa y_{1})\eta_{3}
+t2jη3Ψ1((2jt)12η1η3+κ)+t2jMlη3Ψ2((2jt)12η1η3+κ,η2(2jt)12η1+κη3).\displaystyle+t2^{j}\eta_{3}\Psi_{1}\biggl{(}\frac{(2^{j}t)^{-\frac{1}{2}}\eta_{1}}{\eta_{3}}+\kappa\biggl{)}+t2^{j-Ml}\eta_{3}\Psi_{2}\biggl{(}\frac{(2^{j}t)^{-\frac{1}{2}}\eta_{1}}{\eta_{3}}+\kappa,\frac{\eta_{2}}{(2^{j}t)^{-\frac{1}{2}}\eta_{1}+\kappa\eta_{3}}\biggl{)}.

By Taylor’s expansion of Ψ1((2jt)12η1η3+κ)\Psi_{1}\biggl{(}\frac{(2^{j}t)^{-\frac{1}{2}}\eta_{1}}{\eta_{3}}+\kappa\biggl{)} at η1=0\eta_{1}=0, the phase function can be written as

2j2t12(y1+tΨ1(κ))η1+2jy2η2+2j(y3+κy1+tΨ1(κ))η3+η122η3Ψ1′′(κ)+Rj,l,1(η,t,κ),\displaystyle 2^{\frac{j}{2}}t^{-\frac{1}{2}}(y_{1}+t\Psi_{1}^{\prime}(\kappa))\eta_{1}+2^{j}y_{2}\eta_{2}+2^{j}(y_{3}+\kappa y_{1}+t\Psi_{1}(\kappa))\eta_{3}+\frac{\eta_{1}^{2}}{2\eta_{3}}\Psi_{1}^{\prime\prime}(\kappa)+R_{j,l,1}(\eta,t,\kappa),

where we have put all the remainder terms into Rj,l,1R_{j,l,1}. Then integration by parts yields

|Kj,l,2,κ(y,t)|CN25j2t12(1+2j2t12|y1+cd,M,ltκ1d1|+2j|y2|+2j|y3+κy1+cd,M,ltκd/(d1)|)N.|K_{j,l,2,\kappa}(y,t)|\leq\frac{C_{N}2^{\frac{5j}{2}}t^{-\frac{1}{2}}}{(1+2^{\frac{j}{2}}t^{-\frac{1}{2}}|y_{1}+c_{d,M,l}t\kappa^{\frac{1}{d-1}}|+2^{j}|y_{2}|+2^{j}|y_{3}+\kappa y_{1}+c_{d,M,l}t\kappa^{d/(d-1)}|)^{N}}.

This implies the inequality (3.10) since t2j+Mlt\leq 2^{-j+Ml}.

Similarly, we can get the part (3). We decompose Kj,l,2(y,t)=κ𝕀1κ~𝕀2Kj,l,2,κ,κ~(y,t)K_{j,l,2}(y,t)=\sum_{\kappa\in\mathbb{I}_{1}}\sum_{\tilde{\kappa}\in\mathbb{I}_{2}}K_{j,l,2,\kappa,\tilde{\kappa}}(y,t), where

Kj,l,2,κ,κ~(y,t)=23j3ei2jyη+it2jη3Ψ1(η1η3)+it2jMlη3Ψ2(η1η3,η2η1)aj,l,2,κ,κ~(t,2jη)𝑑η,K_{j,l,2,\kappa,\tilde{\kappa}}(y,t)=2^{3j}\int_{\mathbb{R}^{3}}e^{i2^{j}y\cdot\eta+it2^{j}\eta_{3}\Psi_{1}(\frac{\eta_{1}}{\eta_{3}})+it2^{j-Ml}\eta_{3}\Psi_{2}(\frac{\eta_{1}}{\eta_{3}},\frac{\eta_{2}}{\eta_{1}})}a_{j,l,2,\kappa,\tilde{\kappa}}(t,2^{j}\eta)d\eta,

and aj,l,2,κ,κ~(t,2jη)=aj,l,2(t,2jη)χ~((2jt)12(η1η3κ))χ~((2jMlt)12(η2η3κ~))a_{j,l,2,\kappa,\tilde{\kappa}}(t,2^{j}\eta)=a_{j,l,2}(t,2^{j}\eta)\tilde{\chi}\biggl{(}(2^{j}t)^{\frac{1}{2}}(\frac{\eta_{1}}{\eta_{3}}-\kappa)\biggl{)}\tilde{\chi}\biggl{(}(2^{j-Ml}t)^{\frac{1}{2}}(\frac{\eta_{2}}{\eta_{3}}-\tilde{\kappa})\biggl{)}. Then we change variables η1(2jt)12η1+κη3\eta_{1}\rightarrow(2^{j}t)^{-\frac{1}{2}}\eta_{1}+\kappa\eta_{3}, η2(2jMlt)12η2+κ~η3\eta_{2}\rightarrow(2^{j-Ml}t)^{-\frac{1}{2}}\eta_{2}+\tilde{\kappa}\eta_{3}, η3η3\eta_{3}\rightarrow\eta_{3}. We get the phase function in Kj,l,2,κ,κ~K_{j,l,2,\kappa,\tilde{\kappa}} as follows

2j2t12y1η1+2j+Ml2t12y2η2+2j(y3+κy1+κ~y2)η3\displaystyle 2^{\frac{j}{2}}t^{-\frac{1}{2}}y_{1}\eta_{1}+2^{\frac{j+Ml}{2}}t^{-\frac{1}{2}}y_{2}\eta_{2}+2^{j}(y_{3}+\kappa y_{1}+\tilde{\kappa}y_{2})\eta_{3}
+t2jη3Ψ1((2jt)12η1η3+κ)+t2jMlη3Ψ2((2jt)12η1+κη3η3,(2jMlt)12η2+κ~η3(2jt)12η1+κη3).\displaystyle+t2^{j}\eta_{3}\Psi_{1}\biggl{(}\frac{(2^{j}t)^{-\frac{1}{2}}\eta_{1}}{\eta_{3}}+\kappa\biggl{)}+t2^{j-Ml}\eta_{3}\Psi_{2}\biggl{(}\frac{(2^{j}t)^{-\frac{1}{2}}\eta_{1}+\kappa\eta_{3}}{\eta_{3}},\frac{(2^{j-Ml}t)^{-\frac{1}{2}}\eta_{2}+\tilde{\kappa}\eta_{3}}{(2^{j}t)^{-\frac{1}{2}}\eta_{1}+\kappa\eta_{3}}\biggl{)}.

Then by Taylor’s expansion, we get the phase function

2j2t12(y1+tΨ1(κ)+t2Ml1Ψ2(κ,κ~κ)t2Mlκ~κ22Ψ2(κ,κ~κ))η1\displaystyle 2^{\frac{j}{2}}t^{-\frac{1}{2}}\biggl{(}y_{1}+t\Psi_{1}^{\prime}(\kappa)+t2^{-Ml}\partial_{1}\Psi_{2}(\kappa,\frac{\tilde{\kappa}}{\kappa})-t2^{-Ml}\frac{\tilde{\kappa}}{\kappa^{2}}\partial_{2}\Psi_{2}(\kappa,\frac{\tilde{\kappa}}{\kappa})\biggl{)}\eta_{1}
+2j+Ml2t12(y2+t2Ml1κ2Ψ2(κ,κ~κ))η2+2j(y3+κy1+κ~y2+tΨ1(κ)+t2MlΨ2(κ,κ~κ))η3\displaystyle+2^{\frac{j+Ml}{2}}t^{-\frac{1}{2}}\biggl{(}y_{2}+t2^{-Ml}\frac{1}{\kappa}\partial_{2}\Psi_{2}(\kappa,\frac{\tilde{\kappa}}{\kappa})\biggl{)}\eta_{2}+2^{j}\biggl{(}y_{3}+\kappa y_{1}+\tilde{\kappa}y_{2}+t\Psi_{1}(\kappa)+t2^{-Ml}\Psi_{2}(\kappa,\frac{\tilde{\kappa}}{\kappa})\biggl{)}\eta_{3}
+η122η3Ψ1′′(κ)+η222η31κ222Ψ2(κ,κ~κ)+Rj,l,2(η,t,κ,κ~),\displaystyle+\frac{\eta_{1}^{2}}{2\eta_{3}}\Psi_{1}^{\prime\prime}(\kappa)+\frac{\eta_{2}^{2}}{2\eta_{3}}\frac{1}{\kappa^{2}}\partial_{22}\Psi_{2}(\kappa,\frac{\tilde{\kappa}}{\kappa})+R_{j,l,2}(\eta,t,\kappa,\tilde{\kappa}),

where all the remainder terms are included in Rj,l,2R_{j,l,2}. Then the estimate (3) follows by integration by parts. \Box

Proof of Theorem 3.1: (1) When t(0,2j+Ml)t\in(0,2^{-j+Ml}), according to the estimates in (1) and (2) of Lemma 3.2, it is easy to obtain inequality (3.5). Next we will show that (3.5) holds for t[2j+Ml,)t\in[2^{-j+Ml},\infty). Observing the part (3) in Lemma 3.2, for fixed tt, κ\kappa and κ~\tilde{\kappa}, the term

(1+2j2t12|y1+c1,t(κ,κ~)|+2j+Ml2t12|y2+c2,t(κ,κ~)|+2j|y3+c3,t(y1,y2,κ,κ~)|)N(1+2^{\frac{j}{2}}t^{-\frac{1}{2}}|y_{1}+c_{1,t}(\kappa,\tilde{\kappa})|+2^{\frac{j+Ml}{2}}t^{-\frac{1}{2}}|y_{2}+c_{2,t}(\kappa,\tilde{\kappa})|+2^{j}|y_{3}+c_{3,t}(y_{1},y_{2},\kappa,\tilde{\kappa})|)^{-N}

can be viewed as a characteristic function over the set whose projection on the (y1,y2)(y_{1},y_{2})-plane is a rectangle centered at (c1,t(κ,κ~),c2,t(κ,κ~))(c_{1,t}(\kappa,\tilde{\kappa}),c_{2,t}(\kappa,\tilde{\kappa})) with side length 2j/2t1/2×2j+Ml2t1/22^{-j/2}t^{1/2}\times 2^{-\frac{j+Ml}{2}}t^{1/2}. Hence, in order to obtain (3.5), it suffices to prove that if (κ,κ~)(κ,κ~)(\kappa,\tilde{\kappa})\neq(\kappa^{\prime},\widetilde{\kappa}^{\prime}), then we have either |c1,t(κ,κ~)c1,t(κ,κ~)|2j2t12|c_{1,t}(\kappa,\tilde{\kappa})-c_{1,t}(\kappa^{\prime},\widetilde{\kappa}^{\prime})|\geq 2^{-\frac{j}{2}}t^{\frac{1}{2}} or |c2,t(κ,κ~)c2,t(κ,κ~)|2j+Ml2t12|c_{2,t}(\kappa,\widetilde{\kappa})-c_{2,t}(\kappa^{\prime},\widetilde{\kappa}^{\prime})|\geq 2^{-\frac{j+Ml}{2}}t^{\frac{1}{2}}. Indeed, if |κκ|2Ml+2|κ~κ~||\kappa-\kappa^{\prime}|\geq 2^{-Ml+2}|\tilde{\kappa}-\tilde{\kappa}^{\prime}|, there holds

|c1,t(κ,κ)c1,t(κ~,κ~)|t|κκ|2j2t12.|c_{1,t}(\kappa,\kappa^{\prime})-c_{1,t}(\tilde{\kappa},\widetilde{\kappa}^{\prime})|\sim t|\kappa-\kappa^{\prime}|\geq 2^{-\frac{j}{2}}t^{\frac{1}{2}}.

Conversely, if |κκ|<2Ml+2|κ~κ~||\kappa-\kappa^{\prime}|<2^{-Ml+2}|\tilde{\kappa}-\widetilde{\kappa}^{\prime}|, we have

|c2,t(κ,κ)c2,t(κ~,κ~)|2Mlt|κ~κ~|2j+Ml2t12.|c_{2,t}(\kappa,\kappa^{\prime})-c_{2,t}(\tilde{\kappa},\widetilde{\kappa}^{\prime})|\sim 2^{-Ml}t|\tilde{\kappa}-\widetilde{\kappa}^{\prime}|\geq 2^{-\frac{j+Ml}{2}}t^{\frac{1}{2}}.

Then we obtain inequality (3.5).

(2) For each t[1,2]t\in[1,2], inequality (3.6) follows from part (3) of Lemma 3.2, we omit the proof here.

4 Proofs of Theorem 1.2 and Theorem 1.3

Firstly, we introduce some necessary notations in our proof. For Φ\Phi which is given by (1.2), we denote

E~Φ={nh:nh1,λh and {(x1,x2):x2=λhx1}ZHΦ}.\tilde{E}_{\Phi}=\biggl{\{}n_{h}:n_{h}\geq 1,\lambda_{h}\in\mathbb{R}\text{ and }\{(x_{1},x_{2}):x_{2}=\lambda_{h}x_{1}\}\subset Z_{H\Phi}\biggl{\}}.

Notice that since nh1n_{h}\geq 1, for nhE~Φn_{h}\in\tilde{E}_{\Phi}, there holds {(x1,x2):x2=λhx1}ZΦZHΦ.\{(x_{1},x_{2}):x_{2}=\lambda_{h}x_{1}\}\subset Z_{\Phi}\bigcap Z_{H\Phi}. Next we define EΦE_{\Phi} according to the following four cases:

(C1) if ZΦZHΦZ_{\Phi}\bigcap Z_{H\Phi} contains {(x1,x2):x1=0}\{(x_{1},x_{2}):x_{1}=0\} but does not contain {(x1,x2):x2=0}\{(x_{1},x_{2}):x_{2}=0\}, then we define

EΦ=E~Φ{ν1,m/2};E_{\Phi}=\tilde{E}_{\Phi}\cup\{\nu_{1},m/2\};

(C2) if ZΦZHΦZ_{\Phi}\bigcap Z_{H\Phi} contains {(x1,x2):x2=0}\{(x_{1},x_{2}):x_{2}=0\} but does not contain {(x1,x2):x1=0}\{(x_{1},x_{2}):x_{1}=0\}, then we define

EΦ=E~Φ{ν2,m/2};E_{\Phi}=\tilde{E}_{\Phi}\cup\{\nu_{2},m/2\};

(C3) if ZΦZHΦZ_{\Phi}\bigcap Z_{H\Phi} contains both {(x1,x2):x2=0}\{(x_{1},x_{2}):x_{2}=0\} and {(x1,x2):x1=0}\{(x_{1},x_{2}):x_{1}=0\}, then we define

EΦ=E~Φ{ν1,ν2,m/2};E_{\Phi}=\tilde{E}_{\Phi}\cup\{\nu_{1},\nu_{2},m/2\};

(C4) if ZΦZHΦZ_{\Phi}\bigcap Z_{H\Phi} contains neither {(x1,x2):x2=0}\{(x_{1},x_{2}):x_{2}=0\} nor {(x1,x2):x1=0}\{(x_{1},x_{2}):x_{1}=0\}, then we define

EΦ=E~Φ{m/2}.E_{\Phi}=\tilde{E}_{\Phi}\cup\{m/2\}.

The following lemma will show that the height hΦh_{\Phi} of the homogeneous polynomial Φ\Phi is the maximal element of EΦE_{\Phi}. To be convenient, let us denote μΦ\mu_{\Phi} by the maximal element of EΦE_{\Phi}.

Lemma 4.1.

For the homogeneous polynomial Φ\Phi given by (1.2), we have μΦ=hΦ\mu_{\Phi}=h_{\Phi}.

We have the following observations before the proof of Lemma 4.1. For fixed xS1x\in S^{1}, if Φ(x)0\Phi(x)\neq 0, it is clear that orddΦ(x)=0\textsf{ordd}\Phi(x)=0. Otherwise, without loss of generality, we may assume that xS1{x2=λhx1}x\in S^{1}\cap\{x_{2}=\lambda_{h}x_{1}\}, where x2=λhx1x_{2}=\lambda_{h}x_{1} corresponds to the factor (x2λhx1)nh(x_{2}-\lambda_{h}x_{1})^{n_{h}} in (1.2) with λh\lambda_{h}\in\mathbb{R}, nh1n_{h}\geq 1. A simple caculation implies that orddΦ(x)=nh\textsf{ordd}\Phi(x)=n_{h}. Moreover, for a homogeneous polynomial Φ\Phi, the supremum in the definition of orddΦ\textsf{ordd}\Phi can always be achieved at some xS1x\in S^{1}. Conversely, for fixed straight line passing through the origin which is contained in ZΦZ_{\Phi}, without loss of generality, we may again consider x2=λhx1x_{2}=\lambda_{h}x_{1} corresponding to the factor (x2λhx1)nh(x_{2}-\lambda_{h}x_{1})^{n_{h}} in (1.2), there is an xS1{x2=λhx1}x\in S^{1}\cap\{x_{2}=\lambda_{h}x_{1}\} such that orddΦ(x)=nh\textsf{ordd}\Phi(x)=n_{h}. In particular, if EΦ\{m/2}E_{\Phi}\backslash\{m/2\}\neq\emptyset, then for each κEΦ\{m/2}\kappa\in E_{\Phi}\backslash\{m/2\}, there exists an xκS1x_{\kappa}\in S^{1} such that orddΦ(xκ)=κ\textsf{ordd}\Phi(x_{\kappa})=\kappa.

Proof. We first consider the case when hΦ=m/2h_{\Phi}=m/2. If EΦ={m2}E_{\Phi}=\{\frac{m}{2}\}, then it is clear that μΦ=hΦ=m2\mu_{\Phi}=h_{\Phi}=\frac{m}{2}; if EΦ\{m2}E_{\Phi}\backslash\{\frac{m}{2}\}\neq\emptyset, there holds

m2supxS1orddΦ(x)supκEΦ\{m/2}orddΦ(xκ)=supκEΦ\{m/2}κ,\frac{m}{2}\geq\sup_{x\in S^{1}}\textsf{ordd}\Phi(x)\geq\sup_{\kappa\in E_{\Phi}\backslash\{m/2\}}\textsf{ordd}\Phi(x_{\kappa})=\sup_{\kappa\in E_{\Phi}\backslash\{m/2\}}\kappa,

then it follows that μΦ=m2=hΦ\mu_{\Phi}=\frac{m}{2}=h_{\Phi}.

In the case when hΦ>m/2h_{\Phi}>m/2, there exists an xS1ZΦx\in S^{1}\cap Z_{\Phi} such that orddΦ=orddΦ(x)>m/2\textsf{ordd}\Phi=\textsf{ordd}\Phi(x)>m/2, without loss of generality, we may assume that xS1{x2=λhx1}x\in S^{1}\cap\{x_{2}=\lambda_{h}x_{1}\}, where x2=λhx1x_{2}=\lambda_{h}x_{1} corresponds to the factor (x2λhx1)nh(x_{2}-\lambda_{h}x_{1})^{n_{h}} in (1.2). Then we write Φ\Phi as

Φ(x1,x2)=(x2λhx1)nhQ(x1,x2),\Phi(x_{1},x_{2})=(x_{2}-\lambda_{h}x_{1})^{n_{h}}Q(x_{1},x_{2}),

where QQ is a homogeneous polynomial such that Q(x1,λhx1)0Q(x_{1},\lambda_{h}x_{1})\neq 0 for x10x_{1}\neq 0. It is easy to claim that when nh3n_{h}\geq 3, HΦ(x1,λhx1)=0H\Phi(x_{1},\lambda_{h}x_{1})=0.

When nh>m2n_{h}>\frac{m}{2} and m4m\geq 4, then we must have nh3n_{h}\geq 3. Therefore, when m4m\geq 4, we have nhEΦn_{h}\in E_{\Phi} and

nh=supxS1orddΦ(x)supκEΦ\{m/2}orddΦ(xκ)=supκEΦ\{m/2}κ.n_{h}=\sup_{x\in S^{1}}\textsf{ordd}\Phi(x)\geq\sup_{\kappa\in E_{\Phi}\backslash\{m/2\}}\textsf{ordd}\Phi(x_{\kappa})=\sup_{\kappa\in E_{\Phi}\backslash\{m/2\}}\kappa.

Hence there holds μΦ=hΦ=nh\mu_{\Phi}=h_{\Phi}=n_{h}.

Now we are left with the case when m=3m=3 or m=2m=2. When m=3m=3 and nh=3n_{h}=3, the conclusion follows directly by the previous claim. When m=3m=3 and nh=2n_{h}=2, we may write

Φ(x1,x2)=(x2λhx1)2Q(x1,x2),\Phi(x_{1},x_{2})=(x_{2}-\lambda_{h}x_{1})^{2}Q(x_{1},x_{2}),

where QQ is a linear polynomial that is homogeneous of degree one. A direct calculation implies that HΦ(x1,λhx1)=0H\Phi(x_{1},\lambda_{h}x_{1})=0, then we have μΦ=hΦ=2\mu_{\Phi}=h_{\Phi}=2. When m=2m=2, we must have nh=2n_{h}=2 by the previous observation, then

Φ(x1,x2)=cΦ(x2λhx1)2.\Phi(x_{1},x_{2})=c_{\Phi}(x_{2}-\lambda_{h}x_{1})^{2}.

It is obvious that HΦ0H\Phi\equiv 0, so nhEΦn_{h}\in E_{\Phi} and μΦ=hΦ=2\mu_{\Phi}=h_{\Phi}=2. This finishes the proof of Lemma 4.1. \Box

The proof of Theorem 1.2 is based on the following Lemma 4.2.

Lemma 4.2.

If HΦH\Phi vanishes along the straight line x2=λh0x1x_{2}=\lambda_{h_{0}}x_{1} for some λh0\lambda_{h_{0}}\in\mathbb{R} and 1h0N1\leq h_{0}\leq N such that nh0E~Φn_{h_{0}}\in\tilde{E}_{\Phi}, we define the local maximal operator

supt[1,2]|At,kλh0~f(y)|\displaystyle\sup_{t\in[1,2]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t,k}}f(y)\biggl{|}
=supt[1,2]|2f(y1tx1,y2tx2,y3t(Φ(x1,x2)+c2mk))η0(x2λh0x1ϵx1)η~(x1,x2)dx1dx2|,\displaystyle=\sup_{t\in[1,2]}\biggl{|}\int_{\mathbb{R}^{2}}f(y_{1}-tx_{1},y_{2}-tx_{2},y_{3}-t(\Phi(x_{1},x_{2})+c2^{mk}))\eta_{0}(\frac{x_{2}-\lambda_{h_{0}}x_{1}}{\epsilon x_{1}})\widetilde{\eta}(x_{1},x_{2})dx_{1}dx_{2}\biggl{|}, (4.1)

where k1k\gg 1, suppη~{x2:|x|1}\tilde{\eta}\subset\{x\in\mathbb{R}^{2}:|x|\sim 1\}, suppη0B(0,1)\eta_{0}\subset B(0,1), and ϵ>0\epsilon>0 is sufficiently small, such that HΦH\Phi only vanishes along x2=λh0x1x_{2}=\lambda_{h_{0}}x_{1} in the angle |x2λh0x1|ϵ|x1||x_{2}-\lambda_{h_{0}}x_{1}|\leq\epsilon|x_{1}|. It follows that when c=0c=0,

supt[1,2]|At,kλh0~|Lp(3)Lq(3)1\biggl{\|}\sup_{t\in[1,2]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t,k}}\biggl{|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\lesssim 1 (4.2)

holds for (1p,1q)Δ0(\frac{1}{p},\frac{1}{q})\in\Delta_{0} and nh0+1pnh0+1q1<0\frac{n_{h_{0}}+1}{p}-\frac{n_{h_{0}}+1}{q}-1<0. When c0c\neq 0, we get

supt[1,2]|At,kλh0~|Lp(3)Lq(3)2mkq\biggl{\|}\sup_{t\in[1,2]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t,k}}\biggl{|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\lesssim 2^{\frac{mk}{q}} (4.3)

provided that (1p,1q)Δ0(\frac{1}{p},\frac{1}{q})\in\Delta_{0} and nh0+1p1q1<0\frac{n_{h_{0}}+1}{p}-\frac{1}{q}-1<0.

Proof. After some linear transformation (which does not change the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) norm of the maximal operator), we are reduced to estimating the maximal operator defined by

supt[1,2]|2f(y1tx1,y2tx2,y3t(Φ(x1,x2+λh0x1)+c2mk))η0(x2ϵx1)η~(x1,x2+λh0x1)dx1dx2|.\sup_{t\in[1,2]}\biggl{|}\int_{\mathbb{R}^{2}}f(y_{1}-tx_{1},y_{2}-tx_{2},y_{3}-t(\Phi(x_{1},x_{2}+\lambda_{h_{0}}x_{1})+c2^{mk}))\eta_{0}(\frac{x_{2}}{\epsilon x_{1}})\widetilde{\eta}(x_{1},x_{2}+\lambda_{h_{0}}x_{1})dx_{1}dx_{2}\biggl{|}.

Notice that

Φ(x1,x2+λh0x1)\displaystyle\Phi(x_{1},x_{2}+\lambda_{h_{0}}x_{1}) =cΦx1ν1(x2+λh0x1)ν2Πh=1N(x2+λh0x1λhx1)nh\displaystyle=c_{\Phi}x_{1}^{\nu_{1}}(x_{2}+\lambda_{h_{0}}x_{1})^{\nu_{2}}\Pi_{h=1}^{N}(x_{2}+\lambda_{h_{0}}x_{1}-\lambda_{h}x_{1})^{n_{h}}
=cΦx1ν1+ν2+hh0nhx2nh0(x2x1+λh0)ν2Πhh0(x2x1+λh0λh)nh\displaystyle=c_{\Phi}x_{1}^{\nu_{1}+\nu_{2}+\sum_{h\neq h_{0}}n_{h}}x_{2}^{n_{h_{0}}}\biggl{(}\frac{x_{2}}{x_{1}}+\lambda_{h_{0}}\biggl{)}^{\nu_{2}}\Pi_{h\neq h_{0}}(\frac{x_{2}}{x_{1}}+\lambda_{h_{0}}-\lambda_{h})^{n_{h}}
=x1ν1+ν2+hh0nhx2nh0U(x2x1),\displaystyle=x_{1}^{\nu_{1}+\nu_{2}+\sum_{h\neq h_{0}}n_{h}}x_{2}^{n_{h_{0}}}U\biggl{(}\frac{x_{2}}{x_{1}}\biggl{)}, (4.4)

where

U(x2x1):=cΦ(x2x1+λh0)ν2Πhh0(x2x1+λh0λh)nhU\biggl{(}\frac{x_{2}}{x_{1}}\biggl{)}:=c_{\Phi}\biggl{(}\frac{x_{2}}{x_{1}}+\lambda_{h_{0}}\biggl{)}^{\nu_{2}}\Pi_{h\neq h_{0}}(\frac{x_{2}}{x_{1}}+\lambda_{h_{0}}-\lambda_{h})^{n_{h}}

and |U|1|U|\sim 1 since |x2/x1|<ϵ|x_{2}/x_{1}|<\epsilon and λh00\lambda_{h_{0}}\neq 0. Moreover, 1nh0<m1\leq n_{h_{0}}<m, 1ν1+ν2+hh0nh<m1\leq\nu_{1}+\nu_{2}+\sum_{h\neq h_{0}}n_{h}<m.

By isometric transform, we obtain

supt[1,2]|At,kλh0~|Lp(3)Lq(3)llog1ϵ2(nh0+1)lp(nh0+1)lqlsupt[1,2]|At,k,lλh0~|Lp(3)Lq(3),\displaystyle\biggl{\|}\sup_{t\in[1,2]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t,k}}\biggl{|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\leq\sum_{l\geq\log\frac{1}{\epsilon}}2^{\frac{(n_{h_{0}}+1)l}{p}-\frac{(n_{h_{0}}+1)l}{q}-l}\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A^{\lambda_{h_{0}}}_{t,k,l}}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})},

where At,k,lλh0~\widetilde{A^{\lambda_{h_{0}}}_{t,k,l}} denotes the averaging operator along the hypersurface

Sk,lh0={(x1,x2,x1ν1+ν2+hh0nhx2nh0U(2lx2x1)+2mk+nh0lc):|x1||x2|1}.S^{h_{0}}_{k,l}=\biggl{\{}\biggl{(}x_{1},x_{2},x_{1}^{\nu_{1}+\nu_{2}+\sum_{h\neq h_{0}}n_{h}}x_{2}^{n_{h_{0}}}U\biggl{(}2^{-l}\frac{x_{2}}{x_{1}}\biggl{)}+2^{mk+n_{h_{0}}l}c\biggl{)}:|x_{1}|\sim|x_{2}|\sim 1\biggl{\}}. (4.5)

Now it is clear that the surface Sk,lh0S^{h_{0}}_{k,l} has non-vanishing Garssian curvature, then

supt[1,2]|At,k,lλh0~|Lp(3)Lq(3)(c2mkq2nh0lq+1)\displaystyle\biggl{\|}\sup_{t\in[1,2]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t,k,l}}\biggl{|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\lesssim(c2^{\frac{mk}{q}}2^{\frac{n_{h_{0}l}}{q}}+1) (4.6)

provided that (1p,1q)Δ0(\frac{1}{p},\frac{1}{q})\in\Delta_{0}.

The estimate (4.6) arises from a general conclusion. Let UU be a compact set in 2\mathbb{R}^{2}, Ψ\Psi be a function with the non-degenerate Heissian matrix, S={(x1,x2,Ψ(x1,x1)+c):(x1,x2)U}S=\{(x_{1},x_{2},\Psi(x_{1},x_{1})+c):(x_{1},x_{2})\in U\} be a smooth hypersurface, cc\in\mathbb{R}. Then the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) norm of the local maximal operator along SS is bounded by max{1,c1/q}\max\{1,c^{1/q}\} provided that (1p,1q)Δ0(\frac{1}{p},\frac{1}{q})\in\Delta_{0}. Here we outline the main idea behind the proof of this conclusion, which will be directly referenced in the rest of this paper. Upon Littlewood-Paley decomposition, when ξ\xi is constrained to {ξ:|ξ|2j}\{\xi:|\xi|\sim 2^{j}\}, j1j\geq 1, if |ξ1|+|ξ2||ξ3||\xi_{1}|+|\xi_{2}|\gg|\xi_{3}|, it can be proven that the Fourier transform along SS decays rapidly to complete the corresponding maximal function estimate. Next, we present the application of the method of the stationary phase in

dμS^(tξ)=2eit[x1ξ1+x2ξ2+Φ(x1,x2)ξ3]η(x1,x2)𝑑x1𝑑x2,|ξ1|+|ξ2|2j,|ξ1|+|ξ2||ξ3|,\widehat{d\mu_{S}}(t\xi)=\int_{\mathbb{R}^{2}}e^{-it[x_{1}\xi_{1}+x_{2}\xi_{2}+\Phi(x_{1},x_{2})\xi_{3}]}\eta(x_{1},x_{2})dx_{1}dx_{2},\quad|\xi_{1}|+|\xi_{2}|\sim 2^{j},|\xi_{1}|+|\xi_{2}|\lesssim|\xi_{3}|,

where t[1,2]t\in[1,2], η\eta is a smooth cutoff function supported in UU. According to Theorem 1.2.1 in [15], dμS^(tξ)\widehat{d\mu_{S}}(t\xi) can be expressed as the sum of a finite number of terms of the following form

eitΦ~(ξ)a(t,ξ)(1+t|ξ|),e^{-it\tilde{\Phi}(\xi)}\frac{a(t,\xi)}{(1+t|\xi|)},

where the function Φ~(ξ)\widetilde{\Phi}(\xi) is homogeneous of degree one and its Heissian matrix has a rank of 22, a(t,ξ)a(t,\xi) is a symbol of order zero in ξ\xi. Through Sobolev embedding, we transform the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) estimate of the maximal operator into the Lp(3)Lq(3×[1,2])L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}\times[1,2]) estimate of the following Fourier integral operator

3eiyξitΦ~(ξ)a(t,ξ)(1+t|ξ|)f^(ξ)𝑑ξ.\int_{\mathbb{R}^{3}}e^{iy\cdot\xi-it\tilde{\Phi}(\xi)}\frac{a(t,\xi)}{(1+t|\xi|)}\hat{f}(\xi)d\xi.

It is worth noting that the surface {(η,Φ~(η)):ηsuppa(t,)}\{(\eta,\tilde{\Phi}(\eta)):\eta\in suppa(t,\cdot)\} has two non-vanishing principal curvatures, which implies that the method of estimate for the corresponding Fourier integral operator is consistent with the proof described in Section 4 of reference [14].

Now let’s proceed to the proof of Lemma 4.2. Inequality (4.6) yields

supt[1,2]|At,kλh0~|Lp(3)Lq(3)llog1ϵ2(nh0+1)lp(nh0+1)lql(c2mkq2nh0lq+1).\biggl{\|}\sup_{t\in[1,2]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t,k}}\biggl{|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\leq\sum_{l\geq\log\frac{1}{\epsilon}}2^{\frac{(n_{h_{0}}+1)l}{p}-\frac{(n_{h_{0}}+1)l}{q}-l}(c2^{\frac{mk}{q}}2^{\frac{n_{h_{0}}l}{q}}+1). (4.7)

It follows that when c=0c=0, the right side of (4.7) is convergent provided that nh0+1pnh0+1q1<0\frac{n_{h_{0}}+1}{p}-\frac{n_{h_{0}}+1}{q}-1<0. Then inequality (4.2) holds true for (1p,1q)Δ0(\frac{1}{p},\frac{1}{q})\in\Delta_{0} and nh0+1pnh0+1q1<0\frac{n_{h_{0}}+1}{p}-\frac{n_{h_{0}}+1}{q}-1<0. When c0c\neq 0, the right side of (4.7) converges to c2mkqc2^{\frac{mk}{q}} if nh0+1p1q1<0\frac{n_{h_{0}}+1}{p}-\frac{1}{q}-1<0. Hence, there holds inequality (4.3) provided that (1p,1q)Δ0(\frac{1}{p},\frac{1}{q})\in\Delta_{0} and nh0+1p1q1<0\frac{n_{h_{0}}+1}{p}-\frac{1}{q}-1<0. \Box

Remark 4.3.

It is clear that the similar arguments remain valid for each {x2=λhx1}EΦZHΦ\{x_{2}=\lambda_{h}x_{1}\}\subset E_{\Phi}\cap Z_{H\Phi}, or {xi=0}ZΦZHΦ\{x_{i}=0\}\subset Z_{\Phi}\cap Z_{H\Phi}, i=1,2i=1,2, whose corresponding local maximal operator can be defined in the spirit of (4.2).

Proof of Theorem 1.2: Let AtA_{t} be the operator defined by (1.1). A dyadic decomposition and scaling argument imply

supt[1,2]|At|Lp(3)Lq(3)k12(m+2)kp(m+2)kq2ksupt[1,2]|At,k~|Lp(3)Lq(3),\biggl{\|}\sup_{t\in[1,2]}|A_{t}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\leq\sum_{k\geq 1}2^{\frac{(m+2)k}{p}-{\frac{(m+2)k}{q}}-2k}\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,k}}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})},

where

At,k~f(y)=2f(y1tx1,y2tx2,y3t(Φ(x1,x2)+c2mk))η~(x1,x2)𝑑x1𝑑x2\widetilde{A_{t,k}}f(y)=\int_{\mathbb{R}^{2}}f(y_{1}-tx_{1},y_{2}-tx_{2},y_{3}-t(\Phi(x_{1},x_{2})+c2^{mk}))\widetilde{\eta}(x_{1},x_{2})dx_{1}dx_{2} (4.8)

with suppη~{x2:|x|1}\tilde{\eta}\subset\{x\in\mathbb{R}^{2}:|x|\sim 1\}. Notice that since HΦH\Phi is a homogeneous function, ZHΦZ_{H\Phi} either contains the origin only, or contain some straight lines passing through the origin. In the first case, our results follows directly since HΦH\Phi does not vanish on the support of η~\tilde{\eta}. In the later case, by the assumption of Theorem 1.2, we are reduced to considering the maximal operators which are localized near the straight lines containing in ZΦZHΦZ_{\Phi}\cap Z_{H\Phi}, then Lemma 4.2 and Remark 4.3 can be applied to get the following conclusions.

When c=0c=0, for

(1p,1q)\displaystyle(\frac{1}{p},\frac{1}{q}) μEΦ\{m/2}{(1p,1q)Δ0:μ+1pμ+1q1<0,m/2+1pm/2+1q1<0}\displaystyle\in\bigcap_{\mu\in E_{\Phi}\backslash\{m/2\}}\biggl{\{}(\frac{1}{p},\frac{1}{q})\in\Delta_{0}:\frac{\mu+1}{p}-\frac{\mu+1}{q}-1<0,\frac{m/2+1}{p}-\frac{m/2+1}{q}-1<0\biggl{\}}
={(1p,1q)Δ0:hΦ+1phΦ+1q1<0},\displaystyle=\biggl{\{}(\frac{1}{p},\frac{1}{q})\in\Delta_{0}:\frac{h_{\Phi}+1}{p}-\frac{h_{\Phi}+1}{q}-1<0\biggl{\}}, (4.9)

there holds

k1supt[1,2]|At,k~|Lp(3)Lq(3)k12(m+2)kp(m+2)kq2k1.\sum_{k\geq 1}\biggl{\|}\sup_{t\in[1,2]}\biggl{|}\widetilde{A_{t,k}}\biggl{|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\lesssim\sum_{k\geq 1}2^{\frac{(m+2)k}{p}-{\frac{(m+2)k}{q}}-2k}\lesssim 1. (4.10)

When c0c\neq 0, for

(1p,1q)\displaystyle(\frac{1}{p},\frac{1}{q}) μEΦ\{m/2}{(1p,1q)Δ0:μ+1p1q1<0,m/2+1p1q1<0}\displaystyle\in\bigcap_{\mu\in E_{\Phi}\backslash\{m/2\}}\biggl{\{}(\frac{1}{p},\frac{1}{q})\in\Delta_{0}:\frac{\mu+1}{p}-\frac{1}{q}-1<0,\frac{m/2+1}{p}-\frac{1}{q}-1<0\biggl{\}}
={(1p,1q)Δ0:hΦ+1p1q1<0},\displaystyle=\biggl{\{}(\frac{1}{p},\frac{1}{q})\in\Delta_{0}:\frac{h_{\Phi}+1}{p}-\frac{1}{q}-1<0\biggl{\}}, (4.11)

we get

k1supt[1,2]|At,k~|Lp(3)Lq(3)k12(m+2)kp2kq2k1.\sum_{k\geq 1}\biggl{\|}\sup_{t\in[1,2]}\biggl{|}\widetilde{A_{t,k}}\biggl{|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\lesssim\sum_{k\geq 1}2^{\frac{(m+2)k}{p}-{\frac{2k}{q}}-2k}\lesssim 1. (4.12)

This finishes the proof of Theorem 1.2.

Now we are left with the proof of Theorem 1.3, which follows from Lemma 4.4 below.

Lemma 4.4.

If HΦH\Phi vanishes along x2=τx1x_{2}=\tau x_{1} for some non-zero real number τ\tau such that τλh\tau\neq\lambda_{h} for any 1hN1\leq h\leq N. We consider the maximal operator

supt[1,2]|2f(y1tx1,y2tx2,y3t(Φ(x1,x2)+c2mk))η0(x2τx1ϵx1)η~(x1,x2)dx1dx2|,\sup_{t\in[1,2]}\biggl{|}\int_{\mathbb{R}^{2}}f(y_{1}-tx_{1},y_{2}-tx_{2},y_{3}-t(\Phi(x_{1},x_{2})+c2^{mk}))\eta_{0}(\frac{x_{2}-\tau x_{1}}{\epsilon x_{1}})\widetilde{\eta}(x_{1},x_{2})dx_{1}dx_{2}\biggl{|},

where ϵ>0\epsilon>0 is sufficiently small, such that Φ\Phi does not vanish in the angle |x2τx1|ϵ|x1||x_{2}-\tau x_{1}|\leq\epsilon|x_{1}|. Then there holds

supt[1,2]|At,kτ~|Lp(3)Lq(3)c2mkq+1\biggl{\|}\sup_{t\in[1,2]}\biggl{|}\widetilde{A^{\tau}_{t,k}}\biggl{|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\lesssim c2^{\frac{mk}{q}}+1 (4.13)

for (1p,1q)ΔM(\frac{1}{p},\frac{1}{q})\in\Delta_{M}.

Proof. As in Lemma 4.2, we are reduced to estimating the maximal operator defined by

supt[1,2]|At,kτ~f(y)|=supt[1,2]|2f(y1tx1,y2tx2,y3t(Φ(x1,x2+τx1)+c2mk))η0(x2ϵx1)η~(x1,x2+τx1)dx1dx2|.\sup_{t\in[1,2]}\biggl{|}\widetilde{A^{\tau}_{t,k}}f(y)\biggl{|}=\sup_{t\in[1,2]}\biggl{|}\int_{\mathbb{R}^{2}}f(y_{1}-tx_{1},y_{2}-tx_{2},y_{3}-t(\Phi(x_{1},x_{2}+\tau x_{1})+c2^{mk}))\eta_{0}(\frac{x_{2}}{\epsilon x_{1}})\widetilde{\eta}(x_{1},x_{2}+\tau x_{1})dx_{1}dx_{2}\biggl{|}.

Here

Φ(x1,x2+τx1)\displaystyle\Phi(x_{1},x_{2}+\tau x_{1}) =cΦx1m(x2x1+τ)ν2Πh=1N(x2x1+τλh)nh=x1mV(x2x1),\displaystyle=c_{\Phi}x_{1}^{m}\biggl{(}\frac{x_{2}}{x_{1}}+\tau\biggl{)}^{\nu_{2}}\Pi_{h=1}^{N}(\frac{x_{2}}{x_{1}}+\tau-\lambda_{h})^{n_{h}}=x_{1}^{m}V\biggl{(}\frac{x_{2}}{x_{1}}\biggl{)}, (4.14)

where

V(x2x1):=cΦ(x2x1+τ)ν2Πh=1N(x2x1+τλh)nhV\biggl{(}\frac{x_{2}}{x_{1}}\biggl{)}:=c_{\Phi}\biggl{(}\frac{x_{2}}{x_{1}}+\tau\biggl{)}^{\nu_{2}}\Pi_{h=1}^{N}(\frac{x_{2}}{x_{1}}+\tau-\lambda_{h})^{n_{h}}

and |V|1|V|\sim 1 since |x2/x1|ϵ|x_{2}/x_{1}|\leq\epsilon and τ0\tau\neq 0. It is clear that when |x2|ϵ|x1||x_{2}|\leq\epsilon|x_{1}| and |x1|1|x_{1}|\sim 1,

Φ(x1,x2+τx1)0,x1Φ(x1,x2+τx1)0.\Phi(x_{1},x_{2}+\tau x_{1})\neq 0,\quad\partial_{x_{1}}\Phi(x_{1},x_{2}+\tau x_{1})\neq 0.

Hence under some transformations (which does not change the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) norm of supt[1,2]|At,kτ~|\sup_{t\in[1,2]}|\widetilde{A_{t,k}^{\tau}}| up to a constant), it is sufficient to consider the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) estimate of the maximal operator along the hypersurface

S~k={(r(1+sMg(s)),rs,rm+c2mk):r[1,2],s(0,ϵ0)},\tilde{S}_{k}=\{(r(1+s^{M}g(s)),rs,r^{m}+c2^{mk}):r\in[1,2],s\in(0,\epsilon_{0})\}, (4.15)

where 2Mm2\leq M\leq m, gg is a smooth function and g(0)0g(0)\neq 0, ϵ0\epsilon_{0} is a small constant. Here we applied the assumption that the level set {(x1,x2):Φ(x1,x2)=1}\{(x_{1},x_{2}):\Phi(x_{1},x_{2})=1\} determines a curve of finite type MM near the straight line x2=τx1x_{2}=\tau x_{1}. Finally, the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) boundedness of supt[1,2]|At,kτ~|\sup_{t\in[1,2]}|\widetilde{A_{t,k}^{\tau}}| follows from Theorem 1.4 directly. \Box

Remark 4.5.

Notice that the results in Lemma 4.4 are independent of τ\tau, and remain valid if HΦH\Phi vanishes along x1=0x_{1}=0 or x2=0x_{2}=0, but Φ\Phi dose not vanish.

Proof of Theorem 1.3: As in the beginning of the proof of Theorem 1.2, then we are reduced to estimating the maximal operator supt[1,2]|At,k~|\sup_{t\in[1,2]}|\widetilde{A_{t,k}}|, where At,k~\widetilde{A_{t,k}} is given by equality (4.8). Since ZHΦZΦZHΦZ_{H\Phi}\cap Z_{\Phi}\subsetneqq Z_{H\Phi}, there exists at least one straight line passing through the origin which is contained in ZHΦZ_{H\Phi} but not contained in ZΦZ_{\Phi}. Then Lemma 4.4 and Remark 4.5 can be applied.

When c=0c=0, for

(1p,1q){(1p,1q)ΔM:hΦ+1phΦ+1q1<0},(\frac{1}{p},\frac{1}{q})\in\biggl{\{}(\frac{1}{p},\frac{1}{q})\in\Delta_{M}:\frac{\ h_{\Phi}+1}{p}-\frac{h_{\Phi}+1}{q}-1<0\biggl{\}}, (4.16)

we obtain

k1supt[1,2]|At,k~|Lp(3)Lq(3)k12(m+2)kp(m+2)kq2k1.\sum_{k\geq 1}\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,k}}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\lesssim\sum_{k\geq 1}2^{\frac{(m+2)k}{p}-{\frac{(m+2)k}{q}}-2k}\lesssim 1. (4.17)

When c0c\neq 0, for

(1p,1q){(1p,1q)ΔM:hΦ+1p1q1<0},(\frac{1}{p},\frac{1}{q})\in\biggl{\{}(\frac{1}{p},\frac{1}{q})\in\Delta_{M}:\frac{h_{\Phi}+1}{p}-\frac{1}{q}-1<0\biggl{\}}, (4.18)

there holds

k1supt[1,2]|At,k~|Lp(3)Lq(3)k12(m+2)kp2kq2k1.\sum_{k\geq 1}\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,k}}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\lesssim\sum_{k\geq 1}2^{\frac{(m+2)k}{p}-{\frac{2k}{q}}-2k}\lesssim 1. (4.19)

Then we finish the proof of Theorem 1.3.

5 Proofs of Theorem 1.6 and Theorem 1.7

Proof of Theorem 1.6: A simple calculation implies that HΨH\Psi may vanish along the curve x2=x1dϕ(x1)x_{2}=x_{1}^{d}\phi(x_{1}) and the straight line x1=0x_{1}=0. Hence we decompose the averaging operator AtA_{t} as follows

Atf(y)=i=13Atif(y).A_{t}f(y)=\sum_{i=1}^{3}A_{t}^{i}f(y).

Here, we define the operators AtiA_{t}^{i}, i=1,2,3i=1,2,3 respectively as (without of generality, we may assume that x1>0x_{1}>0):

At1f(y):=2f(y1tx1,y2tx2,y3t(x2x1dϕ(x1))m)(1χ0(2Nx2x1dϕ(x1)))η(x1,x2)dx1dx2,A_{t}^{1}f(y):=\int_{\mathbb{R}^{2}}f(y_{1}-tx_{1},y_{2}-tx_{2},y_{3}-t(x_{2}-x_{1}^{d}\phi(x_{1}))^{m})\biggl{(}1-\chi_{0}\biggl{(}2^{-N}\frac{x_{2}}{x_{1}^{d}\phi(x_{1})}\biggl{)}\biggl{)}\eta(x_{1},x_{2})dx_{1}dx_{2},
At2f(y):=2f(y1tx1,y2tx2,y3t(x2x1dϕ(x1))m)χ0(2Nx2x1dϕ(x1)x1dϕ(x1))η(x1,x2)dx1dx2,A_{t}^{2}f(y):=\int_{\mathbb{R}^{2}}f(y_{1}-tx_{1},y_{2}-tx_{2},y_{3}-t(x_{2}-x_{1}^{d}\phi(x_{1}))^{m})\chi_{0}\biggl{(}2^{N}\frac{x_{2}-x_{1}^{d}\phi(x_{1})}{x_{1}^{d}\phi(x_{1})}\biggl{)}\eta(x_{1},x_{2})dx_{1}dx_{2},

and

At3f(y):\displaystyle A_{t}^{3}f(y): =2f(y1tx1,y2tx2,y3t(x2x1dϕ(x1))m)(χ0(2Nx2x1dϕ(x1))χ0(2Nx2x1dϕ(x1)x1dϕ(x1)))\displaystyle=\int_{\mathbb{R}^{2}}f(y_{1}-tx_{1},y_{2}-tx_{2},y_{3}-t(x_{2}-x_{1}^{d}\phi(x_{1}))^{m})\biggl{(}\chi_{0}\biggl{(}2^{-N}\frac{x_{2}}{x_{1}^{d}\phi(x_{1})}\biggl{)}-\chi_{0}\biggl{(}2^{N}\frac{x_{2}-x_{1}^{d}\phi(x_{1})}{x_{1}^{d}\phi(x_{1})}\biggl{)}\biggl{)}
×η(x1,x2)dx1dx2,\displaystyle\quad\quad\times\eta(x_{1},x_{2})dx_{1}dx_{2},

where 0χ010\leq\chi_{0}\leq 1 is a smooth cutoff function, χ0=1\chi_{0}=1 on [-1,1] and suppχ0[2,2]\chi_{0}\subset[-2,2], and the positive integer NN can be sufficiently large. Next we will demonstrate that the maximal operators supt>0|Ati|\sup_{t>0}|A_{t}^{i}| are Lp(3)L^{p}(\mathbb{R}^{3})-bounded for p>max{2dd+1,32}p>\max\{\frac{2d}{d+1},\frac{3}{2}\}, i=1,2,3i=1,2,3.

We first consider the maximal operator supt>0|At3|\sup_{t>0}|A_{t}^{3}|. By using the dyadic decomposition and isometric transformation, we obtain

supt>0|At3|Lp(3)Lp(3)k12(d+1)ksupt>0|At,k3|Lp(3)Lp(3),\biggl{\|}\sup_{t>0}|A_{t}^{3}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{p}(\mathbb{R}^{3})}\leq\sum_{k\geq 1}2^{-(d+1)k}\biggl{\|}\sup_{t>0}|A_{t,k}^{3}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{p}(\mathbb{R}^{3})}, (5.1)

where the averaging operator At,k3A_{t,k}^{3} is given by

At,k3f(y):\displaystyle A_{t,k}^{3}f(y): =2f(y1tx1,y2tx2,y3t(x2x1dϕ(2kx1))m)ρ~(x1,x2)\displaystyle=\int_{\mathbb{R}^{2}}f(y_{1}-tx_{1},y_{2}-tx_{2},y_{3}-t(x_{2}-x_{1}^{d}\phi(2^{-k}x_{1}))^{m})\tilde{\rho}(x_{1},x_{2})
×(χ0(2Nx2x1dϕ(2kx1))χ0(2Nx2x1dϕ(2kx1)x1dϕ(2kx1)))η(2kx1,2dkx2)dx1dx2,\displaystyle\quad\quad\times\biggl{(}\chi_{0}\biggl{(}2^{-N}\frac{x_{2}}{x_{1}^{d}\phi(2^{-k}x_{1})}\biggl{)}-\chi_{0}\biggl{(}2^{N}\frac{x_{2}-x_{1}^{d}\phi(2^{-k}x_{1})}{x_{1}^{d}\phi(2^{-k}x_{1})}\biggl{)}\biggl{)}\eta(2^{-k}x_{1},2^{-dk}x_{2})dx_{1}dx_{2},

and ρ~\tilde{\rho} is a smooth cutoff function supported in {(x1,x2)2:1/2<|(x1,x2)|<2}\{(x_{1},x_{2})\in\mathbb{R}^{2}:1/2<|(x_{1},x_{2})|<2\}. It is easy to see that the Heissian matrix of the function (x2x1dϕ(2kx1))m(x_{2}-x_{1}^{d}\phi(2^{-k}x_{1}))^{m} degenerates only along the curve x2=x1dϕ(2kx1)x_{2}=x_{1}^{d}\phi(2^{-k}x_{1}) and the straight line x1=0x_{1}=0. Therefore, the hypersurface

{(x1,x2,(x2x1dϕ(2kx1))m):1/2<|(x1,x2)|<2}\{(x_{1},x_{2},(x_{2}-x_{1}^{d}\phi(2^{-k}x_{1}))^{m}):1/2<|(x_{1},x_{2})|<2\}

possesses a non-vanishing Gaussian curvature everywhere when (x1,x2)(x_{1},x_{2}) is restricted to the support of the function

χ0(2Nx2x1dϕ(2kx1))χ0(2Nx2x1dϕ(2kx1)x1dϕ(2kx1)).\chi_{0}\biggl{(}2^{-N}\frac{x_{2}}{x_{1}^{d}\phi(2^{-k}x_{1})}\biggl{)}-\chi_{0}\biggl{(}2^{N}\frac{x_{2}-x_{1}^{d}\phi(2^{-k}x_{1})}{x_{1}^{d}\phi(2^{-k}x_{1})}\biggl{)}.

It follows that the maximal operator supt>0|At,k3|\sup_{t>0}|A_{t,k}^{3}| is Lp(3)L^{p}(\mathbb{R}^{3})-bounded if p>3/2p>3/2. According to (5.1), the result also holds for the maximal operator supt>0|At3|\sup_{t>0}|A_{t}^{3}|.

So it is sufficient to estimate the maximal operators supt>0|At1|\sup_{t>0}|A_{t}^{1}| and supt>0|At2|\sup_{t>0}|A_{t}^{2}|. Theorem 1.6 follows from the two lemmas below.

Lemma 5.1.

For each p>max{3/2,2d/(d+1)}p>\max\{3/2,2d/(d+1)\} and i=1,2i=1,2, there exists a constant Cp>0C_{p}>0 such that

supt[1,2]|Ati|Lp(3)Lp(3)Cp.\biggl{\|}\sup_{t\in[1,2]}|A_{t}^{i}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{p}(\mathbb{R}^{3})}\leq C_{p}. (5.2)
Lemma 5.2.

For each i=1,2i=1,2, there exists q0>p01q_{0}>p_{0}\geq 1 and a constant Cp0,q0>0C_{p_{0},q_{0}}>0 such that

supt[1,2]|Ati|Lp0(3)Lq0(3)Cp0,q0.\biggl{\|}\sup_{t\in[1,2]}|A_{t}^{i}|\biggl{\|}_{L^{p_{0}}(\mathbb{R}^{3})\rightarrow L^{q_{0}}(\mathbb{R}^{3})}\leq C_{p_{0},q_{0}}. (5.3)

Combining the above lemmas with the similar arguments as in [12], we can get the weighted results for the global maximal operators supt>0|Ait|\sup_{t>0}|A^{i}_{t}|. More details can be found in the appendix of this paper. In what follows, we first prove Lemma 5.1 and Lemma 5.2 for i=1i=1. By dyadic decomposition, isometric transform and a simple changing coordinates, for each qp1q\geq p\geq 1, there holds

supt[1,2]|At1|Lp(3)Lq(3)l11kdlN2[(m+1)k+l](1p1q)2lksupt[1,2]|At,k,l1|Lp(3)Lq(3),\biggl{\|}\sup_{t\in[1,2]}|A_{t}^{1}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\leq\sum_{l\geq 1}\sum_{1\leq k\leq dl-N}2^{[(m+1)k+l](\frac{1}{p}-\frac{1}{q})}2^{-l-k}\biggl{\|}\sup_{t\in[1,2]}|A_{t,k,l}^{1}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}, (5.4)

where

At,k,l1f(y):=2f(y1tx1,y2t(x2+2kdlx1dϕ(2lx1)),y3tx2m)ηk,l1(x1,x2)dx1dx2,A_{t,k,l}^{1}f(y):=\int_{\mathbb{R}^{2}}f(y_{1}-tx_{1},y_{2}-t(x_{2}+2^{k-dl}x_{1}^{d}\phi(2^{-l}x_{1})),y_{3}-tx_{2}^{m})\eta_{k,l}^{1}(x_{1},x_{2})dx_{1}dx_{2},

and

ηk,l1(x1,x2)\displaystyle\eta_{k,l}^{1}(x_{1},x_{2}) =ρ~(x1)ρ~(x2+2kdlx1dϕ(2lx1))(1χ0(x2+2kdlx1dϕ(2lx1)2kdl+Nx1dϕ(2lx1)))\displaystyle=\tilde{\rho}(x_{1})\tilde{\rho}(x_{2}+2^{k-dl}x_{1}^{d}\phi(2^{-l}x_{1}))\biggl{(}1-\chi_{0}\biggl{(}\frac{x_{2}+2^{k-dl}x_{1}^{d}\phi(2^{-l}x_{1})}{2^{k-dl+N}x_{1}^{d}\phi(2^{-l}x_{1})}\biggl{)}\biggl{)}
×η(2lx1,2kx2+2dlx1dϕ(2lx1)),\displaystyle\quad\quad\times\eta(2^{-l}x_{1},2^{-k}x_{2}+2^{-dl}x_{1}^{d}\phi(2^{-l}x_{1})),

and ρ~\tilde{\rho} is a smooth function supported in [2,1/2][1/2,2][-2,-1/2]\cup[1/2,2].

By Littlewood-Paley decomposition, we define At,k,l,j1A_{t,k,l,j}^{1} as in equality (2.1), and we are reduced to considering the maximal operator supt>0|At,k,l,j1|\sup_{t>0}|A_{t,k,l,j}^{1}|, j1j\geq 1. Let

dμk,l^(ξ)=2eix1ξ1i[x2+2kdlx1dϕ(2lx1)]ξ2ix2mξ3ηk,l1(x1,x2)dx1dx2.\widehat{d\mu_{k,l}}(\xi)=\int_{\mathbb{R}^{2}}e^{-ix_{1}\xi_{1}-i[x_{2}+2^{k-dl}x_{1}^{d}\phi(2^{-l}x_{1})]\xi_{2}-ix_{2}^{m}\xi_{3}}\eta_{k,l}^{1}(x_{1},x_{2})dx_{1}dx_{2}.

When |ξ1|+|ξ2||ξ3||\xi_{1}|+|\xi_{2}|\gg|\xi_{3}|, we can obtain

|x1,x2[x1ξ1+[x2+2kdlx1dϕ(2lx1)]ξ2+x2mξ3]|||ξ1|2kdl|ξ2||+||ξ2||ξ3|||ξ|,\biggl{|}\nabla_{x_{1},x_{2}}[x_{1}\xi_{1}+[x_{2}+2^{k-dl}x_{1}^{d}\phi(2^{-l}x_{1})]\xi_{2}+x_{2}^{m}\xi_{3}]\biggl{|}\geq\biggl{|}|\xi_{1}|-2^{k-dl}|\xi_{2}|\biggl{|}+\biggl{|}|\xi_{2}|-|\xi_{3}|\biggl{|}\gtrsim|\xi|,

and by performing integration by parts, we achieve rapid decay of dμk,l^(ξ)\widehat{d\mu_{k,l}}(\xi). Therefore, we only need to consider the main contribution term where |ξ1|+|ξ2||ξ3||\xi_{1}|+|\xi_{2}|\lesssim|\xi_{3}|. By an abuse of notation, we define the main contribution term by

At,k,l,j1f(y)=3eiyξdμk,l^(tξ)χ0(|ξ1|+|ξ2||ξ3|)β(2j|ξ|)f^(ξ)dξ.A_{t,k,l,j}^{1}f(y)=\int_{\mathbb{R}^{3}}e^{iy\cdot\xi}\widehat{d\mu_{k,l}}(t\xi)\chi_{0}\biggl{(}\frac{|\xi_{1}|+|\xi_{2}|}{|\xi_{3}|}\biggl{)}\beta(2^{-j}|\xi|)\hat{f}(\xi)d\xi.

By scaling, there holds

supt[1,2]|At,k,l,j1|Lp(3)Lq(3)2(dlk)(1q1p)supt[1,2]|At,k,l,j1|~Lp(3)Lq(3).\biggl{\|}\sup_{t\in[1,2]}|A_{t,k,l,j}^{1}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\lesssim 2^{(dl-k)(\frac{1}{q}-\frac{1}{p})}\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,k,l,j}^{1}|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}. (5.5)

Here,

At,k,l,j1~f(y)=3eiyξdμk,l^(tδ2kdlξ)χ0(2kdl|ξ1|+|ξ2||ξ3|)β(2j|δ2kdlξ|)f^(ξ)dξ,\widetilde{A_{t,k,l,j}^{1}}f(y)=\int_{\mathbb{R}^{3}}e^{iy\cdot\xi}\widehat{d\mu_{k,l}}(t\delta_{2^{k-dl}}\xi)\chi_{0}\biggl{(}\frac{2^{k-dl}|\xi_{1}|+|\xi_{2}|}{|\xi_{3}|}\biggl{)}\beta(2^{-j}|\delta_{2^{k-dl}}\xi|)\hat{f}(\xi)d\xi,

and δ2kdlξ=(2kdlξ1,ξ2,ξ3)\delta_{2^{k-dl}}\xi=(2^{k-dl}\xi_{1},\xi_{2},\xi_{3}). By the stationary phase method and Sobolev embedding, it suffices to estimate the corresponding Fourier integral operators. More concretely, when 1jdlk1\leq j\leq dl-k,

supt[1,2]|At,k,l,j1|~Lp(3)Lq(3)2jqj2ρ(t)Fj,k,l1,1Lp(3)Lq(3×[1/2,4]),\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,k,l,j}^{1}|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\leq 2^{\frac{j}{q}-\frac{j}{2}}\biggl{\|}\rho(t)F_{j,k,l}^{1,1}\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}\times[1/2,4])}, (5.6)

where the bump function ρ\rho is supported in [1/2,4][1/2,4] and the Fourier integral operator Fk,l,j1,1F_{k,l,j}^{1,1} is defined by

Fk,l,j1,1f(y,t)\displaystyle F_{k,l,j}^{1,1}f(y,t) =3eiyξ+itξ3Ψ2(ξ2/ξ3)ak,l,j,1(tξ)χ~(ξ2ξ3)χ0(2kdl|ξ1|+|ξ2||ξ3|)\displaystyle=\int_{\mathbb{R}^{3}}e^{iy\cdot\xi+it\xi_{3}\Psi_{2}(\xi_{2}/\xi_{3})}a_{k,l,j,1}(t\xi)\tilde{\chi}\biggl{(}\frac{\xi_{2}}{\xi_{3}}\biggl{)}\chi_{0}\biggl{(}\frac{2^{k-dl}|\xi_{1}|+|\xi_{2}|}{|\xi_{3}|}\biggl{)}
×β(2j|δ2kdlξ|)β0(2kdl|ξ1|)f^(ξ)dξ,\displaystyle\quad\times\beta(2^{-j}|\delta_{2^{k-dl}}\xi|)\beta_{0}(2^{k-dl}|\xi_{1}|)\hat{f}(\xi)d\xi,

Ψ2(ξ2/ξ3)=cm(ξ2/ξ3)m/(m1)\Psi_{2}(\xi_{2}/\xi_{3})=c_{m}(-\xi_{2}/\xi_{3})^{m/(m-1)} and ak,l,j,1a_{k,l,j,1} is a symbol of order zero. When j>dlkj>dl-k,

supt[1,2]|At,k,l,j1|~Lp(3)Lq(3)2jqj+dlk2ρ(t)Fj,k,l1,2Lp(3)Lq(3×[1/2,4]),\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,k,l,j}^{1}|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\leq 2^{\frac{j}{q}-j+\frac{dl-k}{2}}\biggl{\|}\rho(t)F_{j,k,l}^{1,2}\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}\times[1/2,4])}, (5.7)

where

Fk,l,j1,2f(y,t)\displaystyle F_{k,l,j}^{1,2}f(y,t) =3eiyξ+it2kdlξ2Ψ1(ξ1/ξ2)+itξ3Ψ2(ξ2/ξ3)ak,l,j,2(tξ)χ~(ξ1ξ2)χ~(ξ2ξ3)\displaystyle=\int_{\mathbb{R}^{3}}e^{iy\cdot\xi+it2^{k-dl}\xi_{2}\Psi_{1}(\xi_{1}/\xi_{2})+it\xi_{3}\Psi_{2}(\xi_{2}/\xi_{3})}a_{k,l,j,2}(t\xi)\tilde{\chi}\biggl{(}\frac{\xi_{1}}{\xi_{2}}\biggl{)}\tilde{\chi}\biggl{(}\frac{\xi_{2}}{\xi_{3}}\biggl{)}
×χ0(2kdl|ξ1|+|ξ2||ξ3|)β(2j|δ2kdlξ|)f^(ξ)dξ,\displaystyle\quad\quad\times\chi_{0}\biggl{(}\frac{2^{k-dl}|\xi_{1}|+|\xi_{2}|}{|\xi_{3}|}\biggl{)}\beta(2^{-j}|\delta_{2^{k-dl}}\xi|)\hat{f}(\xi)d\xi,

Ψ1(ξ1/ξ2)\Psi_{1}(\xi_{1}/\xi_{2}) is a small perturbation of cd(ξ1/ξ2)d/(d1)c_{d}(-\xi_{1}/\xi_{2})^{d/(d-1)}, and ak,l,j,2a_{k,l,j,2} is a symbol of order zero.

By replacing 2Ml2^{-Ml} with 2kdl2^{k-dl}, a similar proof implies that the results in Theorem 2.1 remain valid for Fourier integral operators Fk,l,j1,1F_{k,l,j}^{1,1} and Fk,l,j1,2F_{k,l,j}^{1,2}.

By the estimate (E1) in Theorem 2.1, for 1jdlk1\leq j\leq dl-k, we have

supt[1,2]|At,k,l,j1|~L2(3)L2(3)1,\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,k,l,j}^{1}|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})\rightarrow L^{2}(\mathbb{R}^{3})}\lesssim 1, (5.8)

and for j>dlkj>dl-k,

supt[1,2]|At,k,l,j1|~L2(3)L2(3)2j22dlk2.\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,k,l,j}^{1}|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})\rightarrow L^{2}(\mathbb{R}^{3})}\lesssim 2^{-\frac{j}{2}}2^{\frac{dl-k}{2}}. (5.9)

The estimates (E5) and (E4) in Theorem 2.1 imply that for each j1j\geq 1, there hold

supt[1,2]|At,k,l,j1|~L1(3)L1(3)2j,\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,k,l,j}^{1}|}\biggl{\|}_{L^{1}(\mathbb{R}^{3})\rightarrow L^{1}(\mathbb{R}^{3})}\lesssim 2^{j}, (5.10)

and

supt[1,2]|At,k,l,j1|~L(3)L(3)1.\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,k,l,j}^{1}|}\biggl{\|}_{L^{\infty}(\mathbb{R}^{3})\rightarrow L^{\infty}(\mathbb{R}^{3})}\lesssim 1. (5.11)

The interpolation argument implies that inequality (5.2) holds for p>max{3/2,2d/(d+1)}p>\max\{3/2,2d/(d+1)\} and i=1i=1.

By the estimate (E2) in Theorem 2.1, for each j1j\geq 1, there holds

supt[1,2]|At,k,l,j1|~L2(3)L6(3)2dlk2.\biggl{\|}\sup_{t\in[1,2]}|\widetilde{A_{t,k,l,j}^{1}|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})\rightarrow L^{6}(\mathbb{R}^{3})}\lesssim 2^{\frac{dl-k}{2}}. (5.12)

Interpolation with the L2L^{2}-estimates (5.8) and (5.9) implies that when 121dm+d+1<1q12\frac{1}{2}-\frac{1}{dm+d+1}<\frac{1}{q}\leq\frac{1}{2}, inequality (5.3) for i=1i=1 holds. This completes the proof of Lemma 5.2 for i=1i=1.

Now let’s turn to consider the maximal operator supt[1,2]|At2|\sup_{t\in[1,2]}|A_{t}^{2}|, there holds

supt[1,2]|At2|Lp(3)Lq(3)k11lkNd2[(d+1)l+mk](1p1q)2lksupt[1,2]|At,k,l2|Lp(3)Lq(3).\biggl{\|}\sup_{t\in[1,2]}|A_{t}^{2}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}\leq\sum_{k\geq 1}\sum_{1\leq l\leq\frac{k-N}{d}}2^{[(d+1)l+mk](\frac{1}{p}-\frac{1}{q})}2^{-l-k}\biggl{\|}\sup_{t\in[1,2]}|A_{t,k,l}^{2}|\biggl{\|}_{L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})}. (5.13)

Here,

At,k,l2f(y):=2f(y1tx1,y2t(2dlkx2+x1dϕ(2lx1)),y3tx2m)ηk,l2(x1,x2)dx1dx2,A_{t,k,l}^{2}f(y):=\int_{\mathbb{R}^{2}}f(y_{1}-tx_{1},y_{2}-t(2^{dl-k}x_{2}+x_{1}^{d}\phi(2^{-l}x_{1})),y_{3}-tx_{2}^{m})\eta_{k,l}^{2}(x_{1},x_{2})dx_{1}dx_{2},

and

ηk,l2(x1,x2)\displaystyle\eta_{k,l}^{2}(x_{1},x_{2}) =ρ~(x1)ρ~(x2)χ0(2N2dlkx2x1dϕ(2lx1))η(2lx1,2kx2+2dlx1dϕ(2lx1)).\displaystyle=\tilde{\rho}(x_{1})\tilde{\rho}(x_{2})\chi_{0}\biggl{(}2^{N}2^{dl-k}\frac{x_{2}}{x_{1}^{d}\phi(2^{-l}x_{1})}\biggl{)}\eta(2^{-l}x_{1},2^{-k}x_{2}+2^{-dl}x_{1}^{d}\phi(2^{-l}x_{1})).

By an argument similar to that of supt[1,2]|At1|\sup_{t\in[1,2]}|A_{t}^{1}|, we can get Lp(3)L^{p}(\mathbb{R}^{3})-boundedness of supt[1,2]|At2|\sup_{t\in[1,2]}|A_{t}^{2}| provided that p>3/2p>3/2, and L2(3)Lq(3)L^{2}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3})-boundedness of supt[1,2]|At2|\sup_{t\in[1,2]}|A_{t}^{2}| for 12dmd+d+1<1q12\frac{1}{2}-\frac{d}{md+d+1}<\frac{1}{q}\leq\frac{1}{2}.

Then we completed the proof of the positive results in Theorem 1.6.

Finally, we will show that supt>0|At|\sup_{t>0}|A_{t}| cannot be Lp(3)L^{p}(\mathbb{R}^{3})-bounded for p<2d/(d+1)p<2d/(d+1) when d3d\geq 3. By a change of coordinates, it is equivalent to consider the maximal operator supt>0|A~t|\sup_{t>0}|\widetilde{A}_{t}|, where At~\widetilde{A_{t}} is the averaging operator along the hypersurface {(x1,x2+x1dϕ(x1),x2m):(x1,x2)U}\{(x_{1},x_{2}+x_{1}^{d}\phi(x_{1}),x_{2}^{m}):(x_{1},x_{2})\in U\}, UU is a small neighborhood of the origin. Let U~=(0,δ)×(ϵ/2,ϵ)U\tilde{U}=(0,\delta)\times(\epsilon/2,\epsilon)\subset U, 0<δϵ10<\delta\ll\epsilon\ll 1. Let

T={t(x1,x2,x2m):(x1,x2)U~,t[1,2]}.T=\{t(x_{1},x_{2},x_{2}^{m}):(x_{1},x_{2})\in\tilde{U},t\in[1,2]\}.

It is obvious that |T|ϵδ|T|\sim_{\epsilon}\delta, where the implied constants depend on ϵ\epsilon. We define

D=[10δ,10δ]×[10δd,10δd]×[10mδd,10mδd],D=[-10\delta,10\delta]\times[-10\delta^{d},10\delta^{d}]\times[-10m\delta^{d},10m\delta^{d}],

and f=χDf=\chi_{D} which is the characteristic function over DD. Then

fLp(3)δ2d+1p.\|f\|_{L^{p}(\mathbb{R}^{3})}\sim\delta^{\frac{2d+1}{p}}. (5.14)

Notice that for each yTy\in T, there exists a sy[1,2]s_{y}\in[1,2] and (z1,z2)U~(z_{1},z_{2})\in\tilde{U} (which depend on yy) such that y=sy(z1,z2,z2m)y=s_{y}(z_{1},z_{2},z_{2}^{m}). Then there holds

supt>0|At~f(y)|\displaystyle\sup_{t>0}|\widetilde{A_{t}}f(y)| U~χD(sy(z1x1,z2x2x1dϕ(x1),z2mx2m))dx1dx2\displaystyle\geq\int_{\tilde{U}}\chi_{D}(s_{y}(z_{1}-x_{1},z_{2}-x_{2}-x_{1}^{d}\phi(x_{1}),z_{2}^{m}-x_{2}^{m}))dx_{1}dx_{2}
U~(z1,z2)χD(sy(z1x1,z2x2x1dϕ(x1),z2mx2m))dx1dx2,\displaystyle\geq\int_{\tilde{U}(z_{1},z_{2})}\chi_{D}(s_{y}(z_{1}-x_{1},z_{2}-x_{2}-x_{1}^{d}\phi(x_{1}),z_{2}^{m}-x_{2}^{m}))dx_{1}dx_{2}, (5.15)

where U~(z1,z2)={(x1,x2)U~:|x1z1|δ/2,|x2z2|δd/2}\tilde{U}(z_{1},z_{2})=\{(x_{1},x_{2})\in\tilde{U}:|x_{1}-z_{1}|\leq\delta/2,|x_{2}-z_{2}|\leq\delta^{d}/2\}. It is clear that |U~(z1,z2)|δd+1|\tilde{U}(z_{1},z_{2})|\gtrsim\delta^{d+1}. Moreover, for each (x1,x2)U~(z1,z2)(x_{1},x_{2})\in\tilde{U}(z_{1},z_{2}), there holds

|sy(z1x1)|10δ,|sy(z2mx2m)|10mδd,|s_{y}(z_{1}-x_{1})|\leq 10\delta,\quad|s_{y}(z_{2}^{m}-x_{2}^{m})|\leq 10m\delta^{d},
|sy(z2x2x1dϕ(x1))|10δd,|s_{y}(z_{2}-x_{2}-x_{1}^{d}\phi(x_{1}))|\leq 10\delta^{d},

which implies that sy(z1x1,z2x2x1dϕ(x1),z2mx2m)Ds_{y}(z_{1}-x_{1},z_{2}-x_{2}-x_{1}^{d}\phi(x_{1}),z_{2}^{m}-x_{2}^{m})\in D. Hence for each yTy\in T, there holds

supt>0|At~f(y)||U~(z1,z2)|δd+1.\sup_{t>0}|\widetilde{A_{t}}f(y)|\geq|\tilde{U}(z_{1},z_{2})|\gtrsim\delta^{d+1}.

Furthermore,

supt>0|At~f|Lp(3)supt>0|At~f|Lp(T)ϵδd+1+1p.\biggl{\|}\sup_{t>0}|\widetilde{A_{t}}f|\biggl{\|}_{L^{p}(\mathbb{R}^{3})}\geq\biggl{\|}\sup_{t>0}|\widetilde{A_{t}}f|\biggl{\|}_{L^{p}(T)}\gtrsim_{\epsilon}\delta^{d+1+\frac{1}{p}}. (5.16)

If supt>0|At~|\sup_{t>0}|\widetilde{A_{t}}| is Lp(3)L^{p}(\mathbb{R}^{3}) bounded, then the inequalities (5.14) and (5.16) imply that

δd+1+1pϵδ2d+1p.\delta^{d+1+\frac{1}{p}}\lesssim_{\epsilon}\delta^{\frac{2d+1}{p}}.

Then it follows that 1pd+12d\frac{1}{p}\leq\frac{d+1}{2d} since δ\delta can be sufficiently small.

When d=2d=2, we can get the necessity of p>3/2p>3/2 by modifying the above arguments, i.e., just choose U~=(0,ϵ)×(ϵ,2ϵ)\tilde{U}=(0,\epsilon^{\prime})\times(\epsilon,2\epsilon), T={t(x1,x2,(x2x12ϕ(x1))m):(x1,x2)U~,t[1,2]}T=\{t(x_{1},x_{2},(x_{2}-x_{1}^{2}\phi(x_{1}))^{m}):(x_{1},x_{2})\in\tilde{U},t\in[1,2]\}, U~(z1,z2)={(x1,x2)U~:|zixi|<δ2,i=1,2}\tilde{U}(z_{1},z_{2})=\{(x_{1},x_{2})\in\tilde{U}:|z_{i}-x_{i}|<\frac{\delta}{2},i=1,2\}, and ff as the characteristic function over B(0,δ)B(0,\delta) with 0<δϵϵ10<\delta\ll\epsilon^{\prime}\ll\epsilon\ll 1.

We will prove Theorem 1.7 in the rest of this section.

Proof of Theorem 1.7: Our construction of γ(x1)\gamma(x_{1}) was inspired by Example 6.1.1 in [16]. Let χ\chi be a smooth cutoff function, 0χ10\leq\chi\leq 1, χ=1\chi=1 on [1,1][-1,1] and supp χ[2,2]\chi\subset[-2,2]. For each kk\in\mathbb{N}, set the dyadic interval Ik=[2k1,2k]I_{k}=[2^{-k-1},2^{-k}] with center c(Ik)=3/42kc(I_{k})=3/4\cdot 2^{-k}. Put

γ(x1)=k122kχ(x1c(Ik)2k10).\gamma(x_{1})=\sum_{k\geq 1}2^{-2^{k}}\chi\biggl{(}\frac{x_{1}-c(I_{k})}{2^{-k-10}}\biggl{)}.

It is clear that γ(x1)\gamma(x_{1}) is a smooth function and γ(j)(0)=0\gamma^{(j)}(0)=0 for every integer j0j\geq 0. We notice that for each kk\in\mathbb{N}, γ(x1)=22k\gamma(x_{1})=2^{-2^{k}} on Ik~=[2k10+c(Ik),2k10+c(Ik)]\tilde{I_{k}}=[-2^{-k-10}+c(I_{k}),2^{-k-10}+c(I_{k})].

For arbitrary small neighborhood UU of the origin, we choose kk sufficiently large such that Ik~×(0,δ)U\tilde{I_{k}}\times(0,\delta)\subset U, where δ22k\delta\ll 2^{-2^{k}} which is allowed to be sufficiently small. Let the vectors

e1=(1,0,0),e2=(0,1,m(22k)m1),e3=(0,m(22k)m1,1),e_{1}=(1,0,0),\quad\quad e_{2}=(0,1,m(-2^{-2^{k}})^{m-1}),\quad\quad e_{3}=(0,-m(-2^{-2^{k}})^{m-1},1),

and the rectangle

D={y3:|ye1|102k;|ye2|10δ;|ye3|10m22(m2)2kδ2}.D=\{y\in\mathbb{R}^{3}:|y\cdot e_{1}|\leq 10\cdot 2^{-k};|y\cdot e_{2}|\leq 10\cdot\delta;|y\cdot e_{3}|\leq 10m^{2}\cdot 2^{-(m-2)2^{k}}\delta^{2}\}.

For each t[1,2]t\in[1,2], set

Dt\displaystyle D_{t} ={y3:|[yt(0,0,(22k)m)]e1|2k;|[yt(0,0,(22k)m)]e2|δ;\displaystyle=\{y\in\mathbb{R}^{3}:|[y-t(0,0,(-2^{-2^{k}})^{m})]\cdot e_{1}|\leq 2^{-k};|[y-t(0,0,(-2^{-2^{k}})^{m})]\cdot e_{2}|\leq\delta;
|[yt(0,0,(22k)m)]e3|2(m2)2kδ2}.\displaystyle\quad|[y-t(0,0,(-2^{-2^{k}})^{m})]\cdot e_{3}|\leq 2^{-(m-2)2^{k}}\delta^{2}\}.

Notice that if t,t[1,2]t,t^{\prime}\in[1,2] and |tt|10222kδ2|t-t^{\prime}|\geq 10\cdot 2^{2\cdot 2^{k}}\delta^{2}, then DtDt=D_{t}\cap D_{t^{\prime}}=\emptyset.

Let f=χDf=\chi_{D} be the characteristic function over the set DD. Then

fLp(3)kδ3/p,\|f\|_{L^{p}(\mathbb{R}^{3})}\sim_{k}\delta^{3/p}, (5.17)

where the implied constant depends on kk (notice that kk is fixed). Moreover, for fixed t[1,2]t\in[1,2] and every yDty\in D_{t}, (x1,x2)Ik~×(0,δ)(x_{1},x_{2})\in\tilde{I_{k}}\times(0,\delta), there holds

|[yt(x1,x2,(x2γ(x1))m)]e1||[yt(0,0,(22k)m)]e1|+22k32k,|[y-t(x_{1},x_{2},(x_{2}-\gamma(x_{1}))^{m})]\cdot e_{1}|\leq|[y-t(0,0,(-2^{-2^{k}})^{m})]\cdot e_{1}|+2\cdot 2^{-k}\leq 3\cdot 2^{-k},
|[yt(x1,x2,(x2γ(x1))m)]e2||[yt(0,0,(22k)m)]e2|+2δ+2m2(m1)2kδ3δ,|[y-t(x_{1},x_{2},(x_{2}-\gamma(x_{1}))^{m})]\cdot e_{2}|\leq|[y-t(0,0,(-2^{-2^{k}})^{m})]\cdot e_{2}|+2\delta+2m2^{-(m-1)2^{k}}\delta\leq 3\cdot\delta,
|[yt(x1,x2,(x2γ(x1))m)]e3||[yt(0,0,(22k)m)]e3|+2m(m1)2(m2)2kδ22m22(m2)2kδ2.|[y-t(x_{1},x_{2},(x_{2}-\gamma(x_{1}))^{m})]\cdot e_{3}|\leq|[y-t(0,0,(-2^{-2^{k}})^{m})]\cdot e_{3}|+2m(m-1)2^{-(m-2)2^{k}}\delta^{2}\leq 2m^{2}\cdot 2^{-(m-2)2^{k}}\delta^{2}.

Hence,

Atf(y)Ik~×(0,δ)χD(y1tx1,y2tx2,y3t(x2γ(x1))m)dx1dx2kδ.A_{t}f(y)\geq\int_{\tilde{I_{k}}\times(0,\delta)}\chi_{D}(y_{1}-tx_{1},y_{2}-tx_{2},y_{3}-t(x_{2}-\gamma(x_{1}))^{m})dx_{1}dx_{2}\gtrsim_{k}\delta.

We choose a sequence {ti}i1[1,2]\{t_{i}\}_{i\geq 1}\subset[1,2] such that ti+1ti=10222kδ2t_{i+1}-t_{i}=10\cdot 2^{2\cdot 2^{k}}\delta^{2}, there holds

supt[1,2]|Atf(y)|Lp(iDti)k|iDti|1/pδkδ1+1/p.\displaystyle\biggl{\|}\sup_{t\in[1,2]}|A_{t}f(y)|\biggl{\|}_{L^{p}(\cup_{i}D_{t_{i}})}\gtrsim_{k}|\cup_{i}D_{t_{i}}|^{1/p}\delta\gtrsim_{k}\delta^{1+1/p}. (5.18)

Finally, Theorem 1.7 follows from inequality (5.17) and (5.18) since δ\delta can be sufficiently small.

6 Appendix

This appendix consists of two parts. Firstly, we provide some counterexamples related to Theorem 1.2 and Theorem 1.3, and discuss the necessary conditions for the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) boundedness of the corresponding local maximal operators. Secondly, for the completeness of the paper, we explain how to prove the Hölder continuity property starting from the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) bounded estimates of the local maximal operators in this paper. Once we obtain the Hölder continuity property of the local maximal operator, we can immediately obtain the weighted estimate for the corresponding global maximal operator by combining it with Theorem 1.20 and Corollary 1.21 in [12].

6.1 Some counterexamples and necessary conditions

Theorem 6.1.

Let ν1>ν22\nu_{1}>\nu_{2}\geq 2, cc\in\mathbb{R}, U2U\subset\mathbb{R}^{2} be a small neighborhood of the origin (0,0)(0,0), S={(x1,x2,x1ν1x2ν2+c):(x1,x2)U}S=\{(x_{1},x_{2},x_{1}^{\nu_{1}}x_{2}^{\nu_{2}}+c):(x_{1},x_{2})\in U\}. Then the local maximal operator associated with SS cannot be Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) bounded if (1p,1q)Δ0¯(\frac{1}{p},\frac{1}{q})\notin\overline{\Delta_{0}}. Moreover, when c=0c=0, ν1+1pν1+1q10\frac{\nu_{1}+1}{p}-\frac{\nu_{1}+1}{q}-1\leq 0 is necessary for the local maximal operator to be Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) bounded; when c0c\neq 0, ν1+1p1q10\frac{\nu_{1}+1}{p}-\frac{1}{q}-1\leq 0 is a necessary condition for the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) boundedness of the local maximal operator.

Proof. Let

Atf(y)=Uf(yt(x1,x2,x1ν1x2ν2+c))dx1dx2.A_{t}f(y)=\int_{U}f(y-t(x_{1},x_{2},x_{1}^{\nu_{1}}x_{2}^{\nu_{2}}+c))dx_{1}dx_{2}.

We first show that supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| cannot be Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) bounded unless 1p23\frac{1}{p}\leq\frac{2}{3}, 1q1p3q\frac{1}{q}\leq\frac{1}{p}\leq\frac{3}{q} and 1q2p1\frac{1}{q}\geq\frac{2}{p}-1. By selecting the function ff as the characteristic function on balls B(0,δ1)B(0,\delta^{-1}) and the set TT as the δ1\delta^{-1} neighborhood of SS, where δ>0\delta>0 can be sufficiently small, if supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| is Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) bounded, then we can obtain |T|1q|B(0,δ1)|1p|T|^{\frac{1}{q}}\lesssim|B(0,\delta^{-1})|^{\frac{1}{p}}, it follows the necessity of 1q1p\frac{1}{q}\leq\frac{1}{p}. In order to establish the necessary condition 1p3q\frac{1}{p}\leq\frac{3}{q}, we choose the function ff as the characteristic function on the set TT, where the set TT is defined as the δ\delta neighborhood of SS. It is important to note that for any yy belonging to the ball B(0,δ)B(0,\delta) and (x1,x2)U(x_{1},x_{2})\in U, y+(x1,x2,x1ν1x2ν2+c)y+(x_{1},x_{2},x_{1}^{\nu_{1}}x_{2}^{\nu_{2}}+c) is an element of TT. Hence the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) boundedness of supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| implies the validity of |B(0,δ)|1q|T|1p|B(0,\delta)|^{\frac{1}{q}}\lesssim|T|^{\frac{1}{p}}, which means that the condition 1p3q\frac{1}{p}\leq\frac{3}{q} is necessary.

To establish the necessary condition 1p23\frac{1}{p}\leq\frac{2}{3}, we define the function ff as the characteristic function on the ball B(0,δ)B(0,\delta), and the set T={t(x1,x2,x1ν1x2ν2+c):t[1,2],(x1,x2)(ϵ2,ϵ)×(ϵ2,ϵ)}T=\{t(x_{1},x_{2},x_{1}^{\nu_{1}}x_{2}^{\nu_{2}}+c):t\in[1,2],(x_{1},x_{2})\in(\frac{\epsilon}{2},\epsilon)\times(\frac{\epsilon}{2},\epsilon)\}, δϵ1\delta\ll\epsilon\ll 1. Then |T|ϵ2|T|\sim\epsilon^{2}, and for each yTy\in T, there exists a t(y)[1,2]t(y)\in[1,2] and (z1,z2)(ϵ2,ϵ)×(ϵ2,ϵ)(z_{1},z_{2})\in(\frac{\epsilon}{2},\epsilon)\times(\frac{\epsilon}{2},\epsilon) (which depend on yy) such that y=t(y)(z1,z2,z1ν1z2ν2+c)Ty=t(y)(z_{1},z_{2},z_{1}^{\nu_{1}}z_{2}^{\nu_{2}}+c)\in T, there holds

supt[1,2[|Atf(y)||At(y)f(y)||U(z1,z2)|,\sup_{t\in[1,2[}|A_{t}f(y)|\geq|A_{t(y)}f(y)|\geq|U(z_{1},z_{2})|,

where the set U(z1,z2)U(z_{1},z_{2}) is defined by U(z1,z2)={(x1,x2):|xizi|<δ,i=1,2}U(z_{1},z_{2})=\{(x_{1},x_{2}):|x_{i}-z_{i}|<\delta,i=1,2\} with measure δ2\delta^{2}. Then it follows from the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) boundedness of supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| that |T|1q|U(z1,z2)||B(0,δ)|1p|T|^{\frac{1}{q}}|U(z_{1},z_{2})|\lesssim|B(0,\delta)|^{\frac{1}{p}}, which requires 1p23\frac{1}{p}\leq\frac{2}{3} since δ\delta can be sufficiently small.

Next we will prove that the condition 1q2p1\frac{1}{q}\leq\frac{2}{p}-1 is necessary. We choose ff as the characteristic function on the set D={z3:|ze1|10δ,|ze2|10δ,|ze3|10δ2}D=\{z\in\mathbb{R}^{3}:|z\cdot e_{1}|\leq 10\delta,|z\cdot e_{2}|\leq 10\delta,|z\cdot e_{3}|\leq 10\delta^{2}\}, where e1=(1,0,ν1ϵν1+ν21)e_{1}=(1,0,\nu_{1}\epsilon^{\nu_{1}+\nu_{2}-1}), e2=(0,1,ν2ϵν1+ν21)e_{2}=(0,1,\nu_{2}\epsilon^{\nu_{1}+\nu_{2}-1}), e3=(ν1ϵν1+ν21,ν2ϵν1+ν21,1)e_{3}=(-\nu_{1}\epsilon^{\nu_{1}+\nu_{2}-1},-\nu_{2}\epsilon^{\nu_{1}+\nu_{2}-1},1), δϵ1\delta\ll\epsilon\ll 1. Let Rt={y3:|(yt(ϵ,ϵ,ϵν1+ν2+c))e1|δ,|(yt(ϵ,ϵ,ϵν1+ν2+c))e2|δ,|(yt(ϵ,ϵ,ϵν1+ν2+c))e3|δ2}R_{t}=\{y\in\mathbb{R}^{3}:|(y-t(\epsilon,\epsilon,\epsilon^{\nu_{1}+\nu_{2}}+c))\cdot e_{1}|\leq\delta,|(y-t(\epsilon,\epsilon,\epsilon^{\nu_{1}+\nu_{2}}+c))\cdot e_{2}|\leq\delta,|(y-t(\epsilon,\epsilon,\epsilon^{\nu_{1}+\nu_{2}}+c))\cdot e_{3}|\leq\delta^{2}\}, t[1,2]t\in[1,2]. Then for each (x1,x2)U(ϵ):={(x1,x2):|xiϵ|δ,i=1,2}(x_{1},x_{2})\in U(\epsilon):=\{(x_{1},x_{2}):|x_{i}-\epsilon|\leq\delta,i=1,2\} and yRty\in R_{t}, t[1,2]t\in[1,2], we have yt(x1,x2,x1ν1x2ν2+c)Dy-t(x_{1},x_{2},x_{1}^{\nu_{1}}x_{2}^{\nu_{2}}+c)\in D. Hence the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) boundedness of supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| yields |t[1,2]Rt|1q|U(ϵ)||D|1p|\cup_{t\in[1,2]}R_{t}|^{\frac{1}{q}}|U(\epsilon)|\lesssim|D|^{\frac{1}{p}}. This implies that 1+1q2p1+\frac{1}{q}\geq\frac{2}{p} since |t[1,2]Rt|ϵδ2|\cup_{t\in[1,2]}R_{t}|\sim_{\epsilon}\delta^{2}.

Finally, in order to complete the proof of the theorem, we choose ff as the characteristic function on the set D=(10δ,10δ)×(ϵ,ϵ)×(10δν1,10δν1)D=(-10\delta,10\delta)\times(-\epsilon,\epsilon)\times(-10\delta^{\nu_{1}},10\delta^{\nu_{1}}), and the set U=(0,δ)×(0,ϵ)U=(0,\delta)\times(0,\epsilon), δϵ1\delta\ll\epsilon\ll 1. When c=0c=0, for each (x1,x2)U(x_{1},x_{2})\in U, and yR:=(0,δ)×(0,ϵ)×(0,δν1)y\in R:=(0,\delta)\times(0,\epsilon)\times(0,\delta^{\nu_{1}}), we have y(x1,x2,x1ν1)Dy-(x_{1},x_{2},x_{1}^{\nu_{1}})\in D. Then it follows from the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) boundedness of supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| that |R|1q|U||D|1p|R|^{\frac{1}{q}}|U|\lesssim|D|^{\frac{1}{p}}. As a result, we have obtained the necessity of ν1+1pν1+1q10\frac{\nu_{1}+1}{p}-\frac{\nu_{1}+1}{q}-1\leq 0. When c0c\neq 0, we choose the set Rt=(0,δ)×(0,ϵ)×(tc,δν1+tc)R_{t}=(0,\delta)\times(0,\epsilon)\times(tc,\delta^{\nu_{1}}+tc), t[1,2]t\in[1,2]. Then for each yRty\in R_{t}, t[1,2]t\in[1,2] and (x1,x2)U(x_{1},x_{2})\in U, we have yt(x1,x2,x1ν1+c)Dy-t(x_{1},x_{2},x_{1}^{\nu_{1}}+c)\in D. Then the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) boundedness of supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| implies |t[1,2]Rt|1q|U||D|1p|\cup_{t\in[1,2]}R_{t}|^{\frac{1}{q}}|U|\lesssim|D|^{\frac{1}{p}}, which dedicates the necessity of ν1+1p1q10\frac{\nu_{1}+1}{p}-\frac{1}{q}-1\leq 0 since |t[1,2]Rt|ϵδ|\cup_{t\in[1,2]}R_{t}|\sim_{\epsilon}\delta. \Box

Theorem 6.2.

Let M6M\geq 6, ϵ1\epsilon\ll 1, S={(x1,x2,x1M+x2M+c):(x1,x2)U}S=\{(x_{1},x_{2},x_{1}^{M}+x_{2}^{M}+c):(x_{1},x_{2})\in U\}. Here, UU is a small neighborhood of the origin (0,0)(0,0). The local maximal operator along SS cannot be Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) bounded unless 1q1p3q\frac{1}{q}\leq\frac{1}{p}\leq\frac{3}{q}. Furthermore, when c0c\neq 0, we have an additional necessary condition that (1p,1q)(\frac{1}{p},\frac{1}{q}) satisfies M/2+1p1q10\frac{M/2+1}{p}-\frac{1}{q}-1\leq 0; when c=0c=0, 1p2MM+1\frac{1}{p}\leq\frac{2M}{M+1} and M+2pM+2q20\frac{M+2}{p}-\frac{M+2}{q}-2\leq 0 are also necessary for the local maximal operator along SS to be Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) bounded.

Proof. The argument on the necessity of condition 1q1p3q\frac{1}{q}\leq\frac{1}{p}\leq\frac{3}{q} in Theorem 6.1 applies to the maximal operator in Theorem 6.2, we prefer to omit its proof here. Let

Atf(y)=Uf(yt(x1,x2,x1M+x2M+c))dx1dx2.A_{t}f(y)=\int_{U}f(y-t(x_{1},x_{2},x_{1}^{M}+x_{2}^{M}+c))dx_{1}dx_{2}.

We first consider the case c0c\neq 0. Let ff be the characteristic function on the set D:=(10δ,10δ)×(10δ,10δ)×(10δM,10δM)D:=(-10\delta,10\delta)\times(-10\delta,10\delta)\times(-10\delta^{M},10\delta^{M}). For each yRt:=(δ,δ)×(δ,δ)×(δM+tc,δM+tc)y\in R_{t}:=(-\delta,\delta)\times(-\delta,\delta)\times(-\delta^{M}+tc,\delta^{M}+tc), (x1,x2)U~=(δ,δ)×(δ,δ)(x_{1},x_{2})\in\tilde{U}=(-\delta,\delta)\times(-\delta,\delta) and t[1,2]t\in[1,2], we have yt(x1,x2,x1M+x2M+c)y-t(x_{1},x_{2},x_{1}^{M}+x_{2}^{M}+c) belongs to DD. Hence the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) boundedness of supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| implies that |t[1,2]Rt|1q|U~||D|1p|\cup_{t\in[1,2]}R_{t}|^{\frac{1}{q}}|\tilde{U}|\lesssim|D|^{\frac{1}{p}}, which requires M/2+1p1q10\frac{M/2+1}{p}-\frac{1}{q}-1\leq 0.

Secondly, when c=0c=0, we will prove the necessity of conditions M+2pM+2q20\frac{M+2}{p}-\frac{M+2}{q}-2\leq 0 and 1pM+12M\frac{1}{p}\leq\frac{M+1}{2M}. Given any (x1,x2)(0,δ)×(0,δ)(x_{1},x_{2})\in(0,\delta)\times(0,\delta), y(0,δ)×(0,δ)×(0,δM)y\in(0,\delta)\times(0,\delta)\times(0,\delta^{M}) and t[1,2]t\in[1,2], we have yt(x1,x2,x1M+x2M)D=(10δ,10δ)×(10δ,10δ)×(10δM,10δM)y-t(x_{1},x_{2},x_{1}^{M}+x_{2}^{M})\in D=(-10\delta,10\delta)\times(-10\delta,10\delta)\times(-10\delta^{M},10\delta^{M}). Therefore, we choose ff as the characteristic function on the set D, then the Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) boundedness of supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| implies the necessity of the condition M+2pM+2q20\frac{M+2}{p}-\frac{M+2}{q}-2\leq 0. In order to illustrate the necessity of 1pM+12M\frac{1}{p}\leq\frac{M+1}{2M}, we choose ff as the characteristic function on the set D:=(10δ,10δ)×(10δM,10δM)×(10MδM,10MδM)D:=(-10\delta,10\delta)\times(-10\delta^{M},10\delta^{M})\times(-10M\delta^{M},10M\delta^{M}). For any given y=t(y)(z1,z2,z2M)T:={t(z1,z2,z2M):(z1,z2)(0,δ)×(ϵ2,ϵ),t[1,2]}y=t(y)(z_{1},z_{2},z_{2}^{M})\in T:=\{t(z_{1},z_{2},z_{2}^{M}):(z_{1},z_{2})\in(0,\delta)\times(\frac{\epsilon}{2},\epsilon),t\in[1,2]\}, δϵ1\delta\ll\epsilon\ll 1, and (x1,x2)U(z1,z2):=(z1δ,z1+δ)×(z2δM,z2+δM)(x_{1},x_{2})\in U(z_{1},z_{2}):=(z_{1}-\delta,z_{1}+\delta)\times(z_{2}-\delta^{M},z_{2}+\delta^{M}), we have yt(y)(x1,x2,x1M+x2M)Dy-t(y)(x_{1},x_{2},x_{1}^{M}+x_{2}^{M})\in D. Then the Lp(3)Lp(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{p}(\mathbb{R}^{3}) boundedness of supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| yields |T|1p|U(z1,z2)||D|1p|T|^{\frac{1}{p}}|U(z_{1},z_{2})|\lesssim|D|^{\frac{1}{p}}. This implies the necessity of 1pM+12M\frac{1}{p}\leq\frac{M+1}{2M}. \Box

6.2 Hölder continuity property

This subsection is a brief overview of how we can prove the Hölder continuity property of the local maximal operator considered in this article, that is to prove that supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| satisfies

supt[1,2]|Atf(y+z)Atf(y)|Lq(3)|z|ϵfLp(3)\biggl{\|}\sup_{t\in[1,2]}\biggl{|}A_{t}f(y+z)-A_{t}f(y)\biggl{|}\biggl{\|}_{L^{q}(\mathbb{R}^{3})}\lesssim|z|^{\epsilon}\|f\|_{L^{p}(\mathbb{R}^{3})} (6.1)

for some real number ϵ>0\epsilon>0 and any z3z\in\mathbb{R}^{3}, |z|1|z|\leq 1, whenever (1p,1q)(0,0)(\frac{1}{p},\frac{1}{q})\neq(0,0) and belongs to the region Δ\Delta where supt[1,2]|At|\sup_{t\in[1,2]}|A_{t}| is Lp(3)Lq(3)L^{p}(\mathbb{R}^{3})\rightarrow L^{q}(\mathbb{R}^{3}) bounded. Since the hypersurface studied in this paper are of finite type, the Fourier transform of their surface measures decays to a certain extent. Therefore, we can utilize the argument in Section 1.1.3 of reference [12] to deduce the problem to obtaining the inequality

supt[1,2],h[0,δ]|At+hf(y)Atf(y)|Lq(3)δϵ1fLp(3)\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}A_{t+h}f(y)-A_{t}f(y)\biggl{|}\biggl{\|}_{L^{q}(\mathbb{R}^{3})}\lesssim\delta^{\epsilon_{1}}\|f\|_{L^{p}(\mathbb{R}^{3})} (6.2)

for every δ(0,1)\delta\in(0,1) and some ϵ1>0\epsilon_{1}>0. Noting that for each (1p,1q)Δ(\frac{1}{p},\frac{1}{q})\in\Delta we have

supt[1,2],h[0,δ]|At+hf(y)Atf(y)|Lq(3)fLp(3),\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}A_{t+h}f(y)-A_{t}f(y)\biggl{|}\biggl{\|}_{L^{q}(\mathbb{R}^{3})}\lesssim\|f\|_{L^{p}(\mathbb{R}^{3})},

then we can get inequality (6.2) by interpolating with the results in the following theorem.

Theorem 6.3.

(1) If (1p,1q)=(12,12)(\frac{1}{p},\frac{1}{q})=(\frac{1}{2},\frac{1}{2}), then the estimate (6.2) holds true for AtA_{t} that is considered in Theorem 1.2 (1), Theorem 1.3 (1) and Theorem 1.6, respectively.
(2) Let AtA_{t} be the averaging operator described in Theorem 1.2 (2) or Theorem 1.3 (2). Then the estimate (6.2) holds true with (1p,1q)=(18hΦ,18hΦ)(\frac{1}{p},\frac{1}{q})=(\frac{1}{8h_{\Phi}},\frac{1}{8h_{\Phi}}).

Proof. We firstly prove the corresponding results for the averaging operator AtA_{t} considered in Theorem 1.2 (1) and Theorem 1.3 (1). After dyadic decomposition and isometric transformation, we have

supt[1,2],h[0,δ]|At+hf(y)Atf(y)|Lp(3)k122ksupt[1,2],h[0,δ]|At+h,k~f(y)At,k~f(y)|Lp(3),\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}A_{t+h}f(y)-A_{t}f(y)\biggl{|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})}\lesssim\sum_{k\geq 1}2^{-2k}\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A_{t+h,k}}f(y)-\widetilde{A_{t,k}}f(y)\biggl{|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})}, (6.3)

where At,k~\widetilde{A_{t,k}} is defined by equality (4.8), then the proof of this part can be finished if we can show that

supt[1,2],h[0,δ]|At+h,k~f(y)At,k~f(y)|L2(3)δϵ1fL2(3)\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A_{t+h,k}}f(y)-\widetilde{A_{t,k}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim\delta^{\epsilon_{1}}\|f\|_{L^{2}(\mathbb{R}^{3})} (6.4)

for some ϵ1>0\epsilon_{1}>0. We present the principal part of our proof, which is to demonstrate that the operator At,kλh0~\widetilde{A_{t,k}^{\lambda_{h_{0}}}} in Lemma 4.2 and the operator At,kτ~\widetilde{A_{t,k}^{\tau}} in Lemma 4.4 satisfy (6.4). In the context of operator At,kλh0~\widetilde{A_{t,k}^{\lambda_{h_{0}}}}, for any p>1p>1, we decompose At,kλh0~\widetilde{A_{t,k}^{\lambda_{h_{0}}}} such that

supt[1,2],h[0,δ]|Aλh0t+h,k~f(y)Aλh0t,k~f(y)|Lp(3)\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t+h,k}}f(y)-\widetilde{A^{\lambda_{h_{0}}}_{t,k}}f(y)\biggl{|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})}
llog1ϵ2lsupt[1,2],h[0,δ]|Aλh0t+h,k,l~f(y)Aλh0t,k,l~f(y)|Lp(3),\displaystyle\leq\sum_{l\geq\log\frac{1}{\epsilon}}2^{-l}\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t+h,k,l}}f(y)-\widetilde{A^{\lambda_{h_{0}}}_{t,k,l}}f(y)\biggl{|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})}, (6.5)

where Aλh0t,k,l~\widetilde{A^{\lambda_{h_{0}}}_{t,k,l}} denotes the averaging operator along the hypersurface Sh0k,lS^{h_{0}}_{k,l} defined by (4.5) (with c=0c=0) whose Gaussian curvature is non-vanishing. We further introduce the operator At,k,l,jλh0~\widetilde{A_{t,k,l,j}^{\lambda_{h_{0}}}} such that the Fourier transform of At,k,l,jλh0~f\widetilde{A_{t,k,l,j}^{\lambda_{h_{0}}}}f is supported in {ξ3:|ξ|2j}\{\xi\in\mathbb{R}^{3}:|\xi|\sim 2^{j}\}, j1j\geq 1, then Sobolev embedding and Plancherel’s theorem imply that

supt[1,2],h[0,δ]|Aλh0t+h,k,l,j~f(y)Aλh0t,k,l,j~f(y)|L2(3)δ12fL2(3)\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t+h,k,l,j}}f(y)-\widetilde{A^{\lambda_{h_{0}}}_{t,k,l,j}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim\delta^{\frac{1}{2}}\|f\|_{L^{2}(\mathbb{R}^{3})} (6.6)

and

supt[1,2],h[0,δ]|Aλh0t+h,k,l,j~f(y)Aλh0t,k,l,j~f(y)|L2(3)2j2fL2(3).\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t+h,k,l,j}}f(y)-\widetilde{A^{\lambda_{h_{0}}}_{t,k,l,j}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim 2^{-\frac{j}{2}}\|f\|_{L^{2}(\mathbb{R}^{3})}. (6.7)

Interpolation between (6.6) and (6.7) implies that for ϵ1(0,1)\epsilon_{1}\in(0,1), there holds

supt[1,2],h[0,δ]|Aλh0t+h,k,l,j~f(y)Aλh0t,k,l,j~f(y)|L2(3)δϵ122(1ϵ1)j2fL2(3).\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t+h,k,l,j}}f(y)-\widetilde{A^{\lambda_{h_{0}}}_{t,k,l,j}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim\delta^{\frac{\epsilon_{1}}{2}}2^{-\frac{(1-\epsilon_{1})j}{2}}\|f\|_{L^{2}(\mathbb{R}^{3})}. (6.8)

Then the estimate (6.4) for At,kλh0~\widetilde{A_{t,k}^{\lambda_{h_{0}}}} follows from (6.8) and (6.2).

Regarding the operator At,kτ~\widetilde{A_{t,k}^{\tau}}, we introduce the averaging operator At,k,jτ~\widetilde{A_{t,k,j}^{\tau}} along the hypersurface Sk~\tilde{S_{k}} defined by (4.15) with c=0c=0, and the Fourier transform of At,k,jτ~f\widetilde{A_{t,k,j}^{\tau}}f is supported in {ξ3:|ξ|2j}\{\xi\in\mathbb{R}^{3}:|\xi|\sim 2^{j}\}, j1j\geq 1. Sobolev embedding and the maximal estimate (2.23) yield

supt[1,2],h[0,δ]|Aτt+h,k,j~f(y)Aτt,k,j~f(y)|L2(3)δ122(121M)jfL2(3)\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\tau}_{t+h,k,j}}f(y)-\widetilde{A^{\tau}_{t,k,j}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim\delta^{\frac{1}{2}}2^{(\frac{1}{2}-\frac{1}{M})j}\|f\|_{L^{2}(\mathbb{R}^{3})} (6.9)

and

supt[1,2],h[0,δ]|Aτt+h,k,j~f(y)Aτt,k,j~f(y)|L2(3)2jMfL2(3).\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\tau}_{t+h,k,j}}f(y)-\widetilde{A^{\tau}_{t,k,j}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim 2^{-\frac{j}{M}}\|f\|_{L^{2}(\mathbb{R}^{3})}. (6.10)

Interpolating between (6.9) and (6.10), we obtain that for ϵ1(0,2M)\epsilon_{1}\in(0,\frac{2}{M}), there holds

supt[1,2],h[0,δ]|Aτt+h,k,j~f(y)Aτt,k,j~f(y)|L2(3)δϵ122(ϵ121M)jfL2(3).\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\tau}_{t+h,k,j}}f(y)-\widetilde{A^{\tau}_{t,k,j}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim\delta^{\frac{\epsilon_{1}}{2}}2^{(\frac{\epsilon_{1}}{2}-\frac{1}{M})j}\|f\|_{L^{2}(\mathbb{R}^{3})}. (6.11)

Then we have obtained the estimate (6.4) for At,kτ~\widetilde{A_{t,k}^{\tau}}.

Next we will show the results for AtA_{t} considered in Theorem 1.2 and Theorem 1.3 with c0c\neq 0. Let At,k~\widetilde{A_{t,k}} be defined by (4.8). Then the desired results follow if we can prove that

supt[1,2],h[0,δ]|At+h,k~f(y)At,k~f(y)|Lp(3)22mkpδϵ1pfLp(3)\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A_{t+h,k}}f(y)-\widetilde{A_{t,k}}f(y)\biggl{|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})}\lesssim 2^{\frac{2mk}{p}}\delta^{\frac{\epsilon_{1}}{p}}\|f\|_{L^{p}(\mathbb{R}^{3})} (6.12)

for p>2hΦp>2h_{\Phi}. We demonstrate how to prove this result for operators At,kλh0~\widetilde{A_{t,k}^{\lambda_{h_{0}}}} and At,kτ~\widetilde{A_{t,k}^{\tau}}. For the operator At,kλh0~\widetilde{A_{t,k}^{\lambda_{h_{0}}}}, we define At,k,lλh0~\widetilde{A_{t,k,l}^{\lambda_{h_{0}}}} and At,k,l,jλh0~\widetilde{A_{t,k,l,j}^{\lambda_{h_{0}}}} as in the case when c=0c=0. It follows from Sobolev embedding and Plancherel’s theorem that

supt[1,2],h[0,δ]|Aλh0t+h,k,l,j~f(y)Aλh0t,k,l,j~f(y)|L2(3)2mk+nh0lδ12fL2(3)\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t+h,k,l,j}}f(y)-\widetilde{A^{\lambda_{h_{0}}}_{t,k,l,j}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim 2^{mk+n_{h_{0}}l}\delta^{\frac{1}{2}}\|f\|_{L^{2}(\mathbb{R}^{3})} (6.13)

and

supt[1,2],h[0,δ]|Aλh0t+h,k,l,j~f(y)Aλh0t,k,l,j~f(y)|L2(3)2mk+nh0l22j2fL2(3).\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t+h,k,l,j}}f(y)-\widetilde{A^{\lambda_{h_{0}}}_{t,k,l,j}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim 2^{\frac{mk+n_{h_{0}}l}{2}}2^{-\frac{j}{2}}\|f\|_{L^{2}(\mathbb{R}^{3})}. (6.14)

Inequalities (6.13) and (6.14) imply that for ϵ1(0,1)\epsilon_{1}\in(0,1), there holds

supt[1,2],h[0,δ]|Aλh0t+h,k,l,j~f(y)Aλh0t,k,l,j~f(y)|L2(3)2mk+nh0lδϵ122(1ϵ1)j2fL2(3),\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t+h,k,l,j}}f(y)-\widetilde{A^{\lambda_{h_{0}}}_{t,k,l,j}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim 2^{mk+n_{h_{0}}l}\delta^{\frac{\epsilon_{1}}{2}}2^{-\frac{(1-\epsilon_{1})j}{2}}\|f\|_{L^{2}(\mathbb{R}^{3})}, (6.15)

this yields

supt[1,2],h[0,δ]|Aλh0t+h,k,l~f(y)Aλh0t,k,l~f(y)|L2(3)2mk+nh0lδϵ12fL2(3).\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t+h,k,l}}f(y)-\widetilde{A^{\lambda_{h_{0}}}_{t,k,l}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim 2^{mk+n_{h_{0}}l}\delta^{\frac{\epsilon_{1}}{2}}\|f\|_{L^{2}(\mathbb{R}^{3})}. (6.16)

Moreover, we have the following trivial estimate,

supt[1,2],h[0,δ]|Aλh0t+h,k,l~f(y)Aλh0t,k,l~f(y)|L(3)fL(3).\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t+h,k,l}}f(y)-\widetilde{A^{\lambda_{h_{0}}}_{t,k,l}}f(y)\biggl{|}\biggl{\|}_{L^{\infty}(\mathbb{R}^{3})}\lesssim\|f\|_{L^{\infty}(\mathbb{R}^{3})}. (6.17)

Interpolation between (6.16) and (6.17) yields

supt[1,2],h[0,δ]|Aλh0t+h,k,l~f(y)Aλh0t,k,l~f(y)|Lp(3)22mk+2nh0lpδϵ1pfLp(3).\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\lambda_{h_{0}}}_{t+h,k,l}}f(y)-\widetilde{A^{\lambda_{h_{0}}}_{t,k,l}}f(y)\biggl{|}\biggl{\|}_{L^{p}(\mathbb{R}^{3})}\lesssim 2^{\frac{2mk+2n_{h_{0}}l}{p}}\delta^{\frac{\epsilon_{1}}{p}}\|f\|_{L^{p}(\mathbb{R}^{3})}. (6.18)

Inequalities (6.18) and (6.2) imply that (6.12) holds for At,kτ~\widetilde{A_{t,k}^{\tau}} with p>2hΦp>2h_{\Phi}.

As for the operator At,kτ~\widetilde{A_{t,k}^{\tau}}, we define the operators At,k,jτ~\widetilde{A_{t,k,j}^{\tau}} in the same way as in the case when c=0c=0. Sobolev embedding and the maximal estimate (2.23) yield

supt[1,2],h[0,δ]|Aτt+h,k,j~f(y)Aτt,k,j~f(y)|L2(3)2mkδ122(121M)jfL2(3)\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\tau}_{t+h,k,j}}f(y)-\widetilde{A^{\tau}_{t,k,j}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim 2^{mk}\delta^{\frac{1}{2}}2^{(\frac{1}{2}-\frac{1}{M})j}\|f\|_{L^{2}(\mathbb{R}^{3})} (6.19)

and

supt[1,2],h[0,δ]|Aτt+h,k,j~f(y)Aτt,k,j~f(y)|L2(3)2mk22jMfL2(3).\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\tau}_{t+h,k,j}}f(y)-\widetilde{A^{\tau}_{t,k,j}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim 2^{\frac{mk}{2}}2^{-\frac{j}{M}}\|f\|_{L^{2}(\mathbb{R}^{3})}. (6.20)

Interpolating between (6.19) and (6.20), we obtain that for ϵ1(0,2M)\epsilon_{1}\in(0,\frac{2}{M}), there holds

supt[1,2],h[0,δ]|Aτt+h,k,j~f(y)Aτt,k,j~f(y)|L2(3)2mkδϵ122(ϵ121M)jfL2(3).\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\tau}_{t+h,k,j}}f(y)-\widetilde{A^{\tau}_{t,k,j}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim 2^{mk}\delta^{\frac{\epsilon_{1}}{2}}2^{(\frac{\epsilon_{1}}{2}-\frac{1}{M})j}\|f\|_{L^{2}(\mathbb{R}^{3})}. (6.21)

It follows that

supt[1,2],h[0,δ]|Aτt+h,k~f(y)Aτt,k~f(y)|L2(3)2mkδϵ12fL2(3).\displaystyle\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\tau}_{t+h,k}}f(y)-\widetilde{A^{\tau}_{t,k}}f(y)\biggl{|}\biggl{\|}_{L^{2}(\mathbb{R}^{3})}\lesssim 2^{mk}\delta^{\frac{\epsilon_{1}}{2}}\|f\|_{L^{2}(\mathbb{R}^{3})}. (6.22)

Then the estimate (6.12) for At,kτ~\widetilde{A_{t,k}^{\tau}} can be obtained by interpolating between (6.22) and the trivial estimate .

supt[1,2],h[0,δ]|Aτt+h,k~f(y)Aτt,k~f(y)|L(3)fL(3).\biggl{\|}\sup_{t\in[1,2],h\in[0,\delta]}\biggl{|}\widetilde{A^{\tau}_{t+h,k}}f(y)-\widetilde{A^{\tau}_{t,k}}f(y)\biggl{|}\biggl{\|}_{L^{\infty}(\mathbb{R}^{3})}\lesssim\|f\|_{L^{\infty}(\mathbb{R}^{3})}.

\Box

Finally, we tend to omit the proof of the corresponding conclusion for the operator AtA_{t} considered in Theorem 1.6, as the idea of its proof is very similar to that of the operators involved in Theorem 1.2 (1) and 1.3 (1).

References

  • [1] B. Buschenhenke, S. Dendrinos, I. A. Ikromov and D. Müller, Estimates for maximal functions associated to hypersurfaces in 3\mathbbm{R}^{3} with height h<2h<2: Part I, Trans. Amer. Math. Soc., 372(2) (2019), 1363-1406.
  • [2] B. Buschenhenke, I. A. Ikromov and D. Müller, Estimates for maximal functions associated to hypersurfaces in 3\mathbbm{R}^{3} with height h<2h<2: part II: A geometric conjecture and its proof for generic 2-surfaces, Ann. Sc. Norm. Super. Pisa Cl. Sci., (2023). https://doi.org/10.2422/2036-2145.202301-016.
  • [3] A. Greenleaf, Principal curvature and harmonic analysis, Indiana U. Math. J., 4 (1981), 519-537.
  • [4] I. A. Ikromov and D. Müller, On adapted coordinate systems, Trans. Amer. Math. Soc., 363(6) (2011), 2821-2848.
  • [5] I. A. Ikromov and D. Müller, Fourier restriction for hypersurfaces in three dimensions and Newton polyhedra, Annals of Mathematics Studies 194, Princeton University Press, Princeton and Oxford 2016, 260 pp.
  • [6] I. A. Ikromov, M. Kempe and D. Müller, Damped oscillatory integrals and boundedness of maximal operators associated to mixed homogeneous hypersurfaces, Duke Math. J., 126 (3) (2005), 471-490.
  • [7] I. A. Ikromov, M. Kempe and D. Müller, Estimate for maximal functions associated with hypersurfaces in 3{\mathbb{R}}^{3} and related problems of harmonic analysis, Acta Math., 204 (2010), 151-171.
  • [8] A. Iosevich, Maximal operators assciated to families of flat curves in the plane, Duke Math. J., 76 (1994), 633-644.
  • [9] A. Iosevich and E. Sawyer, Osillatory integrals and maximal averages over homogeneous surfaces, Duke Math. J., 82 (1996), 103-141.
  • [10] A. Iosevich and E. Sawyer, Maximal averages over surfaces, Adv. Math., 132 (1997), 46-119.
  • [11] W. Li, Maximal functions associated with non-isotropic dilations of hypersurfaces in 3\mathbb{R}^{3}, J. Math. Pures Appl., 113 (2018), 70-140.
  • [12] W. Li, H. Wang and Y. Zhai, LpL^{p}-improving bounds and weighted estimates for maximal functions associated with curvature, J. Fourier. Anal. Appl., 10 (2023), 1-63.
  • [13] G. Mockenhaupt, A. Seeger and C. D. Sogge, Local smoothing of Fourier integral operators and Carleson-Sjölin estimates, J. Amer. Math. Soc., 6 (1993), 65-130.
  • [14] W. Schlag and C. D. Sogge, Local smoothing estimates related to the circular maximal theorem, Math. Res. Lett., 4 (1997), 1-15.
  • [15] C. D. Sogge, Fourier integrals in classical analysis, Cambridge Tracts in Mathematics, vol. 105, Cambridge University Press, Cambridge, 1993.
  • [16] E. Zimmermann, On LpL^{p}-estimates for maximal average over hypersurfaces not satisfying the transversality condition, Phd thesis, Christian-Albrechts Universität Bibliothek Kiel, 2014.