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

Construction of a curved Kakeya set

Tongou Yang Department of Mathematics, University of California
Los Angeles, CA 90095, United States
[email protected]
 and  Yue Zhong Department of Mathematics, Sun Yat-sen University
Guangzhou, 510275, P.R. China
[email protected]
Abstract.

We construct a compact set in 2\mathbb{R}^{2} of measure 0 containing a piece of a parabola of every aperture between 11 and 22. As a consequence, we improve lower bounds for the LpL^{p}-LqL^{q} norm of the corresponding maximal operator for a range of p,qp,q. Moreover, our construction can be generalised from parabolas to a family of curves with cinematic curvature.

1. Introduction

Consider Wolff’s circular maximal Kakeya function introduced in [KW99]:

𝒞f(r)=supx2C(x,r)|f(y)|𝑑y,\mathcal{C}f(r)=\sup_{x\in\mathbb{R}^{2}}\int_{C(x,r)}|f(y)|dy,

initially defined for continuous functions f:2f:\mathbb{R}^{2}\to\mathbb{C} with compact support, where r[1,2]r\in[1,2] and C(x,r)C(x,r) denotes the circle centred at xx of radius rr. For δ>0\delta>0, we can also consider a δ\delta-thickened version of the maximal function, defined by

𝒞δf(r)=supx21|Cδ(x,r)|Cδ(x,r)|f(y)|𝑑y,\mathcal{C}_{\delta}f(r)=\sup_{x\in\mathbb{R}^{2}}\frac{1}{|C_{\delta}(x,r)|}\int_{C_{\delta}(x,r)}|f(y)|dy,

where Cδ(x,r)C_{\delta}(x,r) denotes the annulus centred at xx of radius rr and thickness δ\delta, namely, Cδ(x,r)={y2:rδ|yx|r+δ}C_{\delta}(x,r)=\{y\in\mathbb{R}^{2}:r-\delta\leq|y-x|\leq r+\delta\}. For Lebesgue exponents p,q[1,]p,q\in[1,\infty], we are interested in the Lp(2)Lq([1,2])L^{p}(\mathbb{R}^{2})\to L^{q}([1,2]) mapping property of 𝒞\mathcal{C} and 𝒞δ\mathcal{C}_{\delta}. Wolff [Wol97] proved the bound

𝒞δL3L3εδε,ε>0,\left\lVert\mathcal{C}_{\delta}\right\rVert_{L^{3}\to L^{3}}\lesssim_{\varepsilon}\delta^{-\varepsilon},\quad\forall\varepsilon>0,

which is sharp except for the ε\varepsilon-loss. Using this, he concluded that every compact subset of 2\mathbb{R}^{2} containing a circle of every radius between 11 and 22 must have Hausdorff dimension 22.

A closely related analogue of 𝒞\mathcal{C} is a parabolic maximal function defined by

𝒫f(a)=sup(x1,x2)201|f(x1+t,x2+at2)|𝑑t.\mathcal{P}f(a)=\sup_{(x_{1},x_{2})\in\mathbb{R}^{2}}\int_{0}^{1}|f(x_{1}+t,x_{2}+at^{2})|dt.

Similarly, for each δ>0\delta>0, we can define the δ\delta-thickened version

𝒫δf(a)=sup(x1,x2)2(2δ)101δδ|f(x1+t,x2+at2+s)|𝑑s𝑑t.\mathcal{P}_{\delta}f(a)=\sup_{(x_{1},x_{2})\in\mathbb{R}^{2}}(2\delta)^{-1}\int_{0}^{1}\int_{-\delta}^{\delta}|f(x_{1}+t,x_{2}+at^{2}+s)|dsdt.

1.1. Curves of cinematic curvature

Both maximal functions can be thought of as special cases of a family of curves with cinematic curvature, introduced by Sogge [Sog91]. See also [KW99][Zah12a][Zah12b][PYZ22][CGY23] [Zah23][CG24] for related discussions on maximal operator bounds related to curves of cinematic curvature. There are many different but essentially equivalent formulations of a family of curves ua(t)u_{a}(t) satisfying the cinematic curvature condition; for instance, in [CGY23] it is formulated as

det[uttuatutttuatt]0,\det\begin{bmatrix}u_{tt}&u_{at}\\ u_{ttt}&u_{att}\end{bmatrix}\neq 0, (1.1)

which one can check to be true for the family of parabolas {(t,at2):t[0,1]}\{(t,at^{2}):t\in[0,1]\} with a[1,2]a\in[1,2].

Thus, one has the following upper bounds for the parabolic maximal operator.

𝒫δfL3([1,2])εδεfL3(2)ε>0,\displaystyle\left\lVert\mathcal{P}_{\delta}f\right\rVert_{L^{3}([1,2])}\lesssim_{\varepsilon}\delta^{-\varepsilon}\left\lVert f\right\rVert_{L^{3}(\mathbb{R}^{2})}\quad\forall\varepsilon>0, (1.2)
𝒫δfLq([1,2])δ1232pfLp(2),p<83,q2pp1.\displaystyle\left\lVert\mathcal{P}_{\delta}f\right\rVert_{L^{q}([1,2])}\lesssim\delta^{\frac{1}{2}-\frac{3}{2p}}\left\lVert f\right\rVert_{L^{p}(\mathbb{R}^{2})},\quad p<\frac{8}{3},\quad q\geq\frac{2p}{p-1}. (1.3)

Indeed, (1.2) follows from [Zah12a] or [Zah12b], and (1.3) follows from [KW99]. It is worth noting that (1.3) has no ε\varepsilon-losses in the power of δ\delta.

1.2. Lower bounds

In order to make things simple, we first restrict ourselves to the parabolic maximal operator 𝒫δ\mathcal{P}_{\delta}. Define

B(p,q)=supf0𝒫fLq([1,2])fLp(2),B(p,q,δ)=supf0𝒫δfLq([1,2])fLp(2).B(p,q)=\sup_{f\neq 0}\frac{\left\lVert\mathcal{P}f\right\rVert_{L^{q}([1,2])}}{\left\lVert f\right\rVert_{L^{p}(\mathbb{R}^{2})}},\quad B(p,q,\delta)=\sup_{f\neq 0}\frac{\left\lVert\mathcal{P}_{\delta}f\right\rVert_{L^{q}([1,2])}}{\left\lVert f\right\rVert_{L^{p}(\mathbb{R}^{2})}}.

The inequalities (1.2) and (1.3) give upper bounds for B(p,q,δ)B(p,q,\delta). On the other hand, the existence of some Kakeya sets provides some related lower bounds of B(p,q,δ)B(p,q,\delta). For clarity, we first introduce the notion of (vertical) δ\delta-thickening S(δ)S(\delta) for a subset S2S\subseteq\mathbb{R}^{2}:

S(δ):={x+(0,ε):xS,δεδ}.S(\delta):=\{x+(0,\varepsilon):x\in S,-\delta\leq\varepsilon\leq\delta\}. (1.4)
Theorem 1.1 (Parabolic variant of Kolasa-Wolff construction [KW99]).

There exists a compact subset KK of 2\mathbb{R}^{2} of Lebesgue measure 0 that contains a piece of length 1\sim 1 of a parabola of every aperture between 11 and 22. Moreover, its δ\delta-thickening K(δ)K(\delta) has measure (logδ1)2(loglogδ1)2\lesssim(\log\delta^{-1})^{-2}(\log\log\delta^{-1})^{2}. Thus we have the lower bound

B(p,q,δ)(logδ1)2/p(loglogδ1)2/p,B(p,q,\delta)\gtrsim(\log\delta^{-1})^{2/p}(\log\log\delta^{-1})^{-2/p}, (1.5)

where the implicit constant is independent of δ\delta. In particular, B(p,q)=B(p,q)=\infty if p<p<\infty.

Proof.

This follows from an easy adaptation of the main construction of Proposition 1.1 in [KW99] for circles to the case of parabolas. ∎

We also encourage the reader to check other curved Kakeya set constructions, such as [BR68][Kin68][Dav72][Tal80][HKLO23][CYZ23].

The main theorem of this paper is an improvement of Theorem 1.1 as follows.

Theorem 1.2 (Main theorem).

There exists a compact subset KK of 2\mathbb{R}^{2} of Lebesgue measure 0 that contains a piece of length 1\sim 1 of a parabola of every aperture between 11 and 22. Moreover, its δ\delta-thickening K(δ)K(\delta) has measure (logδ1)2\lesssim(\log\delta^{-1})^{-2}. Thus we have the lower bound

B(p,q,δ)(logδ1)2/p.B(p,q,\delta)\gtrsim(\log\delta^{-1})^{2/p}. (1.6)

Namely, by refining the main construction in [KW99], we are able to remove the loglogδ1\log\log\delta^{-1} factor in the lower bound.

1.3. Kakeya set with cinematic curvature

More generally, we can generalise the construction in Theorem 1.2 with parabolas replaced by a family of functions obeying cinematic curvature conditions.

We start with a C2C^{2}-function f:[0,1]f:[0,1]\to\mathbb{R} that satisfies the following assumptions:

f(0)0,inff′′>0,\displaystyle f^{\prime}(0)\geq 0,\quad\inf f^{\prime\prime}>0, (1.7)
f′′′ exists and is bounded in (0,1),\displaystyle f^{\prime\prime\prime}\text{ exists and is bounded in }(0,1), (1.8)
ff′′′(f′′)20 in (0,1).\displaystyle f^{\prime}f^{\prime\prime\prime}-(f^{\prime\prime})^{2}\leq 0\text{ in }(0,1). (1.9)

Then one can check that the family of functions ua(t)=af(t)u_{a}(t)=af(t), 1a21\leq a\leq 2 satisfies (1.1). Also, the case of parabolas corresponds to f(t)=t2f(t)=t^{2}.

Define the corresponding maximal operators as

g(a)\displaystyle\mathcal{R}g(a) =sup(x1,x2)201|g(x1+t,x2+af(t))|𝑑t\displaystyle=\sup_{(x_{1},x_{2})\in\mathbb{R}^{2}}\int_{0}^{1}|g(x_{1}+t,x_{2}+af(t))|dt (1.10)
δg(a)\displaystyle\mathcal{R}_{\delta}g(a) =sup(x1,x2)2(2δ)101δδ|g(x1+t,x2+af(t)+s)|𝑑s𝑑t,\displaystyle=\sup_{(x_{1},x_{2})\in\mathbb{R}^{2}}(2\delta)^{-1}\int_{0}^{1}\int_{-\delta}^{\delta}|g(x_{1}+t,x_{2}+af(t)+s)|dsdt,

and define the corresponding operator norms

R(p,q)=supg0gLq([1,2])gLp(2),R(p,q,δ)=supg0δgLq([1,2])gLp(2).R(p,q)=\sup_{g\neq 0}\frac{\left\lVert\mathcal{R}g\right\rVert_{L^{q}([1,2])}}{\left\lVert g\right\rVert_{L^{p}(\mathbb{R}^{2})}},\quad R(p,q,\delta)=\sup_{g\neq 0}\frac{\left\lVert\mathcal{R}_{\delta}g\right\rVert_{L^{q}([1,2])}}{\left\lVert g\right\rVert_{L^{p}(\mathbb{R}^{2})}}.
Theorem 1.3.

Let f:[0,1]f:[0,1]\to\mathbb{R} be a C2C^{2} function obeying (1.7)(1.8)(1.9). Then there exists a compact subset KK of 2\mathbb{R}^{2} of Lebesgue measure 0 that contains a translated copy of a piece of length 1\sim 1 of the graph of a function of the form af(x)af(x) where 1a21\leq a\leq 2. Moreover, its δ\delta-thickening K(δ)K(\delta) has measure (logδ1)2\lesssim(\log\delta^{-1})^{-2}. Thus we have the lower bound

R(p,q,δ)(logδ1)2/p.R(p,q,\delta)\gtrsim(\log\delta^{-1})^{2/p}. (1.11)

In particular, R(p,q)=R(p,q)=\infty for p<p<\infty.

In the following of this article, we prove Theorem 1.3, from which Theorem 1.2 follows as a corollary.

1.4. Outline of the article

In Section 2 we construct the MM-th stage KMK_{M} of the curved Kakeya set KK. In Section 3 we iterate the previous construction to obtain an actual curved Kakeya set of zero measure. In the appendix we give a brief summary of current best upper and lower bounds for R(p,q,δ)R(p,q,\delta).

1.5. Acknowledgements

Tongou Yang is supported by the Croucher Fellowships for Postdoctoral Research. Yue Zhong is supported in part by the National Key R&D Program of China (No. 2022YFA1005700) and the NNSF of China (No. 12371105). Both authors would like to thank Sanghyuk Lee and Shaoming Guo for bringing this problem to our attention, and Lixin Yan, Xianghong Chen and Mingfeng Chen for helpful suggestions.

2. Compression by forcing tangencies

Let M2M\in 2\mathbb{N} be large enough. In this section, we are going to present the MM-th building block FMF_{M} of the construction of the curved Kakeya set KK in Theorem 1.3. This is done using a “cut-and-slide” procedure. The idea at each step jj is to create many tangencies at a fixed xx-coordinate xjx_{j} by translations, so that the curved rectangles are compressed near xjx_{j}. The following is the precise construction.

Unless otherwise specified, all implicit constants are allowed to depend on ff only; more precisely, they depend on fC2\left\lVert f\right\rVert_{C^{2}}, inff′′\inf f^{\prime\prime} and sup|f′′′|\sup|f^{\prime\prime\prime}| given in (1.7)(1.8)(1.9).

2.1. Step 0

We start with a C2C^{2}-function f:[0,1]f:[0,1]\to\mathbb{R} obeying (1.7)(1.8) (1.9). Fix 1a021\leq a_{0}\leq 2 and 0<δ02a00<\delta_{0}\leq 2-a_{0}. Consider the initial “curved rectangle”

T(0)(a0,δ0):={(x,af(x)):x[0,1],a[a0,a0+δ0]}.T^{(0)}(a_{0},\delta_{0}):=\{(x,af(x)):x\in[0,1],a\in[a_{0},a_{0}+\delta_{0}]\}. (2.1)

Fix M4M\in 4\mathbb{N}. We divide T0(a0,δ0)T_{0}(a_{0},\delta_{0}) into 2M2^{M} “curved rectangles” of the form

Tn(0):=T(0)(a0+nδ02M,δ02M),n=0,1,,2M1.T^{(0)}_{n}:=T^{(0)}(a_{0}+n\delta_{0}2^{-M},\delta_{0}2^{-M}),\quad n=0,1,\dots,2^{M}-1. (2.2)

The curve at the bottom of Tn(0)T^{(0)}_{n} is given by

Ln(0)(x):=a0f(x).L^{(0)}_{n}(x):=a_{0}f(x). (2.3)

Partition [0,1][0,1] uniformly into M/2M/2 intervals of length 2M12M^{-1}, where the partitioning points are given by

xj:=2jM1,j=0,1,,M/2.x_{j}:=2jM^{-1},\quad j=0,1,\dots,M/2. (2.4)

2.2. Step 1

For odd n=1,3,,2M1n=1,3,\dots,2^{M}-1, we now apply different translations to Tn(0)T^{(0)}_{n} so that its bottom curve Ln(0)L^{(0)}_{n} will be tangent to the bottom curve Ln1(0)L^{(0)}_{n-1} at x1x_{1}. That is, we need to find translations u:=un(1)u:=u_{n}^{(1)}, v:=vn(1)v:=v_{n}^{(1)} such that

{af(x1u)+v=a~f(x1)af(x1u)=a~f(x1)\left\{\begin{aligned} &af(x_{1}-u)+v=\tilde{a}f(x_{1})\\ &af^{\prime}(x_{1}-u)=\tilde{a}f^{\prime}(x_{1})\\ \end{aligned}\right.

where a=a0+nδ02Ma=a_{0}+n\delta_{0}2^{-M} and a~=a0+(n1)δ02M\tilde{a}=a_{0}+(n-1)\delta_{0}2^{-M}.

To find the solutions, we first focus on the second equation. First, using the implicit function theorem and the fact that inff′′>0\inf f^{\prime\prime}>0, we see that for MM large enough, such uu always exists and 0<uδ02M0<u\lesssim\delta_{0}2^{-M}. To find the expression of uu, by the mean value theorem, there exists some ξ=ξn(1)\xi=\xi_{n}^{(1)} such that

f(x1)f(x1u)=uf′′(ξ).f^{\prime}(x_{1})-f^{\prime}(x_{1}-u)=uf^{\prime\prime}(\xi).

Moreover, such ξ\xi must be unique since f′′>0f^{\prime\prime}>0. Thus the second equation gives

u=un(1)=δ02Maf(x1)f′′(ξ).u=u_{n}^{(1)}=\frac{\delta_{0}2^{-M}}{a}\frac{f^{\prime}(x_{1})}{f^{\prime\prime}(\xi)}. (2.5)

Plugging into the first equation, we can find vv, which is also positive. Also, by Lemma 2.1 below, the translated curve Ln(0)+(u,v)L_{n}^{(0)}+(u,v) is still strictly above Ln1(0)L_{n-1}^{(0)} except at the tangent point.

After Step 1, we obtain 2M12^{M-1} larger curved figures

Tn(1):=Tn(0)(Tn+1(0)+(un+1(1),vn+1(1))),T^{(1)}_{n}:=T_{n}^{(0)}\bigcup(T_{n+1}^{(0)}+(u_{n+1}^{(1)},v_{n+1}^{(1)})), (2.6)

where n=0,2,,2M2n=0,2,\dots,2^{M}-2, that are compressed well at x1x_{1}. Moreover, for each even nn, the curve Ln(0)L_{n}^{(0)} is still the bottom curve of Tn(1)T^{(1)}_{n}. Denote

T(1):=n:2|nTn(1).T^{(1)}:=\bigcup_{n:2|n}T^{(1)}_{n}.

2.3. Step 22

We continue in a similar way, this time compressing Tn(1)T^{(1)}_{n}, n=0,2,,2M2n=0,2,\dots,2^{M}-2 at the point x2x_{2}. More precisely, for n=2,6,10,,2M2n=2,6,10,\dots,2^{M}-2, we translate Tn(1)T^{(1)}_{n} further by some (un(2),vn(2))(u_{n}^{(2)},v_{n}^{(2)}) so that its bottom curve Ln(0)L_{n}^{(0)} is tangent to Ln2(0)L^{(0)}_{n-2} at x2x_{2}. Similarly to (2.6), we obtain 2M22^{M-2} larger curved figures Tn(2)T^{(2)}_{n}, n=0,4,,2M4n=0,4,\dots,2^{M}-4, whose bottom curve is Ln(0)L_{n}^{(0)} by Lemma 2.1 below, and they are compressed well at x2x_{2}. Denote

T(2):=n:4|nTn(2).T^{(2)}:=\bigcup_{n:4|n}T^{(2)}_{n}.

Refer to Figure 1, which shows two steps of translations when 2M=162^{M}=16.

Refer to caption
Figure 1. Figure after two steps when 2M=162^{M}=16

2.4. Step jj

Now we describe a general Step jj. For each nn of the form 2jk+2j12^{j}k+2^{j-1}, k=0,1,,2Mj1k=0,1,\dots,2^{M-j}-1, we translate Tn(j1)T_{n}^{(j-1)} by some (un(j),vn(j))(u^{(j)}_{n},v^{(j)}_{n}) so that its bottom curve Ln(0)L^{(0)}_{n} is tangent to Ln2j1(0)L_{n-2^{j-1}}^{(0)} at xjx_{j}. By the same computation as in Step 1, we have

un(j)=δ02j1Maf(xj)f′′(ξ),u_{n}^{(j)}=\frac{\delta_{0}2^{j-1-M}}{a}\frac{f^{\prime}(x_{j})}{f^{\prime\prime}(\xi)}, (2.7)

where a=a0+nδ02Ma=a_{0}+n\delta_{0}2^{-M}, and ξ=ξn(j)\xi=\xi_{n}^{(j)} is the unique number such that

f′′(ξ)=f(xj)f(xjun(j))un(j),f^{\prime\prime}(\xi)=\frac{f^{\prime}(x_{j})-f^{\prime}(x_{j}-u^{(j)}_{n})}{u^{(j)}_{n}}, (2.8)

and its existence is guaranteed by the implicit function theorem. Moreover, a direct computation using Taylor’s theorem gives for some ζ=ζn(j)\zeta=\zeta_{n}^{(j)} that

vn(j)=δ02j1M[f(xj)2f′′(ξ)f(xj)]af′′(ζ)2(un(j))2.v_{n}^{(j)}=\delta_{0}2^{j-1-M}\left[\frac{f^{\prime}(x_{j})^{2}}{f^{\prime\prime}(\xi)}-f(x_{j})\right]-a\frac{f^{\prime\prime}(\zeta)}{2}(u_{n}^{(j)})^{2}. (2.9)

Thus, similarly to (2.6), we obtain 2Mj2^{M-j} larger curved figures Tn(j)T_{n}^{(j)}, n=0,2j,,2M2jn=0,2^{j},\dots,2^{M}-2^{j}, which are compressed well at xjx_{j}. Also, by the following lemma, the bottom of Tn(j)T_{n}^{(j)} is still Ln(0)L_{n}^{(0)}. Denote

T(j):=n:2j|nTn(j).T^{(j)}:=\bigcup_{n:2^{j}|n}T^{(j)}_{n}.
Lemma 2.1.

If f:[0,1]f:[0,1]\to\mathbb{R} is a C2C^{2} function obeying (1.7)(1.8)(1.9), x0[0,1]x_{0}\in[0,1], and (u,v)(u,v) is the solution of the equation {af(x0u)+v=a~f(x0)af(x0u)=a~f(x0),\left\{\begin{aligned} &af(x_{0}-u)+v=\tilde{a}f(x_{0})\\ &af^{\prime}(x_{0}-u)=\tilde{a}f^{\prime}(x_{0})\\ \end{aligned},\right. then we have af(xu)+va~f(x)af(x-u)+v\geq\tilde{a}f(x) for any xx while aa~a\geq\tilde{a} (such that u,vu,v exist).

Proof.

Fix a~,x,x0\tilde{a},x,x_{0}. Let

F(a)=a(f(xu(a))f(x0u(a)))a~(f(x)f(x0)),F(a)=a(f(x-u(a))-f(x_{0}-u(a)))-\tilde{a}(f(x)-f(x_{0})),

where we regard u,vu,v as functions of aa. We have F(a~)=0F(\tilde{a})=0 since u(a~)=v(a~)=0u(\tilde{a})=v(\tilde{a})=0. We want to show F(a)0F(a)\geq 0 while aa~a\geq\tilde{a}, and it suffices to show that F(a)0F^{\prime}(a)\geq 0 for aa~a\geq\tilde{a}.

Since uu is the solution of the equation af(x0u)=a~f(x0)af^{\prime}(x_{0}-u)=\tilde{a}f^{\prime}(x_{0}), we have

a(af(x0u))=a(a~f(x0)),\frac{\partial}{\partial a}(af^{\prime}(x_{0}-u))=\frac{\partial}{\partial a}(\tilde{a}f^{\prime}(x_{0})),

which means that

u(a)=f(x0u)af′′(x0u).u^{\prime}(a)=\frac{f^{\prime}(x_{0}-u)}{af^{{}^{\prime\prime}}(x_{0}-u)}.

Denote X=xuX=x-u and X0=x0uX_{0}=x_{0}-u. Then by direct computation,

F(a)\displaystyle F^{\prime}(a) =(f(X)f(X0))a(f(X)f(X0))u(a)\displaystyle=(f(X)-f(X_{0}))-a(f^{\prime}(X)-f^{\prime}(X_{0}))\cdot u^{\prime}(a)
=(f(X)f(X0))f′′(X0)(f(X)f(X0))f(X0)f′′(X0)\displaystyle=\dfrac{(f(X)-f(X_{0}))f^{\prime\prime}(X_{0})-(f^{\prime}(X)-f^{\prime}(X_{0}))f^{\prime}(X_{0})}{f^{\prime\prime}(X_{0})}
=G(X)G(X0)f′′(X0),\displaystyle=\frac{G(X)-G(X_{0})}{f^{\prime\prime}(X_{0})},

where we have denoted

G(X):=f(X)f′′(X0)f(X)f(X0).G(X):=f(X)f^{\prime\prime}(X_{0})-f^{\prime}(X)f^{\prime}(X_{0}).

Then it suffices to show G(X)G(X0)G(X)\geq G(X_{0}) for all XX, which is true if we can show G(X)0G^{\prime}(X)\geq 0 for XX0X\geq X_{0} and G(X)0G^{\prime}(X)\leq 0 for XX0X\leq X_{0}. To this end, we consider

H(X):=G(X)f(X)=f(X)f′′(X0)f′′(X)f(X0)f(X).H(X):=\frac{G^{\prime}(X)}{f^{\prime}(X)}=\frac{f^{\prime}(X)f^{\prime\prime}(X_{0})-f^{\prime\prime}(X)f^{\prime}(X_{0})}{f^{\prime}(X)}.

We note that X=xu>0X=x-u>0 since x2/Mux\geq 2/M\ll u. Thus by (1.7), f(X)>0f^{\prime}(X)>0. Thus it suffices to show that H(X)0H(X)\geq 0 for XX0X\geq X_{0} and H(X)0H(X)\leq 0 for XX0X\leq X_{0}. But H(X0)=0H(X_{0})=0, so it further suffices to show H(X)0H^{\prime}(X)\geq 0 for all XX. But direct computation gives

H(X)=f(X0)[f′′(X)2f′′′(X)f(X)]f(X)2,H^{\prime}(X)=\frac{f^{\prime}(X_{0})[f^{\prime\prime}(X)^{2}-f^{\prime\prime\prime}(X)f^{\prime}(X)]}{f^{\prime}(X)^{2}},

which is nonnegative by (1.9). This finishes the proof. ∎

2.5. End of construction

Denote

m:=M/2.m:=M/2. (2.10)

We perform the above procedures for m1m-1 times at each tangent point xj=2j/Mx_{j}=2j/M, j=1,,m1j=1,\dots,m-1, arriving at the set T(m)T^{(m)}. We now define our required set

FM:=T(m)([4logMM,1]×).F_{M}:=T^{(m)}\bigcap\left(\left[\frac{4\log M}{M},1\right]\times\mathbb{R}\right). (2.11)

Note that this means we throw away the part of T(m)T^{(m)} over [0,xj][0,x_{j}] where j<2logMj<2\log M. The choice of the cutoff 4logMM\frac{4\log M}{M} will be clear later in (2.22) in the proof of Theorem 2.3.

2.6. Computation of translations

Our first task is to control the sum (un,vn)(u_{n},v_{n}) of all translations (un(j),vn(j))(u_{n}^{(j)},v_{n}^{(j)}) that have been performed to the original curved rectangle Tn(0)T^{(0)}_{n} at Steps j=1,2,,mj=1,2,\dots,m.

Given n=0,1,,2M1n=0,1,\dots,2^{M}-1, by binary expansion, we know there exist unique integers εj=εj(n){0,1}\varepsilon_{j}=\varepsilon_{j}(n)\in\{0,1\}, 1jM1\leq j\leq M such that n=j=1Mεj2j1n=\sum_{j=1}^{M}\varepsilon_{j}2^{j-1}.

For convenience, we introduce the notation

nj:=n(nmod  2j1),\left\lfloor n\right\rfloor_{j}:=n-(n\,\,\mathrm{mod}\,\,2^{j-1}), (2.12)

which means the “integral part” of nn in in 2j12^{j-1}\mathbb{Z}. Then we note that εj(n)=1\varepsilon_{j}(n)=1 if and only if unju_{\left\lfloor n\right\rfloor_{j}}, vnjv_{\left\lfloor n\right\rfloor_{j}} are defined by Step jj.

Proposition 2.2.

For each n=0,1,,2M1n=0,1,\dots,2^{M}-1, we have the relations

un=j=1mεj(n)unj(j),vn=j=1mεj(n)vnj(j).u_{n}=\sum_{j=1}^{m}\varepsilon_{j}(n)u^{(j)}_{\left\lfloor n\right\rfloor_{j}},\quad v_{n}=\sum_{j=1}^{m}\varepsilon_{j}(n)v^{(j)}_{\left\lfloor n\right\rfloor_{j}}. (2.13)

In particular, we have

0unCδ02M/2,0vnCδ02M/2.0\leq u_{n}\leq C\delta_{0}2^{-M/2},\quad 0\leq v_{n}\leq C\delta_{0}2^{-M/2}. (2.14)

More generally, denote the partial sums

Un(j):=i=1jεi(n)uni(i),Vn(j):=i=1jεi(n)vni(i),U_{n}^{(j)}:=\sum_{i=1}^{j}\varepsilon_{i}(n)u^{(i)}_{\left\lfloor n\right\rfloor_{i}},\quad V_{n}^{(j)}:=\sum_{i=1}^{j}\varepsilon_{i}(n)v^{(i)}_{\left\lfloor n\right\rfloor_{i}}, (2.15)

then we have

0Un(j)Cδ02jM,0Vn(j)Cδ02jM.0\leq U_{n}^{(j)}\leq C\delta_{0}2^{j-M},\quad 0\leq V_{n}^{(j)}\leq C\delta_{0}2^{j-M}. (2.16)

Here CC is a large constant depending on ff only.

Proof.

The proof of (2.13) is by inspection. For example, if m>100m>100 and n=27n=27, then Tn(0)T_{n}^{(0)} is translated according to the bottoms of T27(0)T_{27}^{(0)}, T26(0)T_{26}^{(0)}, T24(0)T_{24}^{(0)} and T16(0)T_{16}^{(0)} at Steps 1,2,4,51,2,4,5, respectively; it remains unchanged at all other steps. Note that εj(27)=1\varepsilon_{j}(27)=1 if and only if j=1,2,4,5j=1,2,4,5, whence 27(27mod  2j1)=27,26,24,1627-(27\,\,\mathrm{mod}\,\,2^{j-1})=27,26,24,16, respectively.

The relations (2.14) then follow from (2.7) and (2.9) and the choice m=M/2m=M/2. The estimates of the partial sums are similar. ∎

This proposition ensures that for large MM, the total distance of translations is tiny; in particular, it can be less than 4logMM\frac{4\log M}{M}, so that the projections of the translated curved rectangles onto the xx-axis all contain [4logMM,1][\frac{4\log M}{M},1].

For future reference, we denote by LnL_{n} the bottom curves of Tn(0)T_{n}^{(0)} after all steps of translations. Namely,

Ln:=Ln(0)+(un,vn).L_{n}:=L_{n}^{(0)}+(u_{n},v_{n}). (2.17)

2.7. Upper bound of measure

In this subsection, we control the measure of the set FMF_{M} we constructed.

Theorem 2.3.

The set FMF_{M} satisfies

|FM|δ0M2.|F_{M}|\lesssim\delta_{0}M^{-2}. (2.18)
Corollary 2.4.

The lower bound (1.11) holds.

Proof of corollary assuming Theorem 2.3.

Let δ=2M\delta=2^{-M} and take g=1FMg=1_{F_{M}} corresponding to a0=1,δ0=1a_{0}=1,\delta_{0}=1. Then the construction gives δg(a)1\mathcal{R}_{\delta}g(a)\sim 1 for every a[1,2]a\in[1,2]. Then the result follows from Theorem 2.3. ∎

Proof of Theorem 2.3.

It suffices to show that for each x0[4logMM,1]x_{0}\in[\frac{4\log M}{M},1],

|{y:(x0,y)FM}|δ0M2.|\{y\in\mathbb{R}:(x_{0},y)\in F_{M}\}|\lesssim\delta_{0}M^{-2}. (2.19)

Fix j[logM,m1]j\in[\log M,m-1] and assume x0[xj,xj+1]x_{0}\in[x_{j},x_{j+1}]. For each n=0,1,,2M1n=0,1,\dots,2^{M}-1, we write

n=p+i=1jεi2i1(p=0,2j,,2M2j).n=p+\sum_{i=1}^{j}\varepsilon_{i}\cdot 2^{i-1}\quad(p=0,2^{j},\cdots,2^{M}-2^{j}). (2.20)

In words, for each jj, we group the translated curved rectangles Tn(m)T_{n}^{(m)} into 2Mj2^{M-j} groups, and Tn(m)T_{n}^{(m)} belongs to the p2jp2^{-j}-th group, whose bottom curve is given by LpL_{p}.

By the triangle inequality, it suffices to show that the thickness of the p2jp2^{-j}-th group is δ0M22jM\lesssim\delta_{0}M^{-2}2^{j-M}. More precisely, we need to show

Ln(x0)Lp(x0)+δ02Mδ0M22jM.L_{n}(x_{0})-L_{p}(x_{0})+\delta_{0}2^{-M}\lesssim\delta_{0}M^{-2}2^{j-M}. (2.21)

Here, Ln(x0)Lp(x0)L_{n}(x_{0})-L_{p}(x_{0}) is the distance between the bottoms, and δ02M\delta_{0}2^{-M} is the thickness of one smallest curved rectangle. But by our choice that j2logMj\geq 2\log M, we have

2MM22jM.2^{-M}\lesssim M^{-2}2^{j-M}. (2.22)

Thus our task reduces to showing

Ln(x0)Lp(x0)δ0M22jM.L_{n}(x_{0})-L_{p}(x_{0})\lesssim\delta_{0}M^{-2}2^{j-M}. (2.23)

We trace back to the configuration right after Step jj. That is, we let

x:=x0i=j+1m1uni(i),x:=x_{0}-\sum_{i=j+1}^{m-1}u_{\left\lfloor n\right\rfloor_{i}}^{(i)}, (2.24)

which lies within [xj1,xj+1][x_{j-1},x_{j+1}], by (2.7). Thus

Ln(x0)Lp(x0)=Ln(j)(x)Lp(j)(x),\displaystyle L_{n}(x_{0})-L_{p}(x_{0})=L_{n}^{(j)}(x)-L_{p}^{(j)}(x),

where Ln(j)L_{n}^{(j)} stands for the bottom curve of Tn(0)T_{n}^{(0)} after jj steps of translations.

Recall that

{Ln(j)(x)=(a0+δ0n2M)f(xUn(j))+Vn(j)Lpj(x)=(a0+δ0p2M)f(xUp(j))+Vp(j).\left\{\begin{aligned} &L_{n}^{(j)}(x)=(a_{0}+\delta_{0}n2^{-M})f(x-U_{n}^{(j)})+V_{n}^{(j)}\\ &L_{p}^{j}(x)=(a_{0}+\delta_{0}p2^{-M})f(x-U_{p}^{(j)})+V_{p}^{(j)}\\ \end{aligned}\right..

According to the definition of Un(j)U_{n}^{(j)} and Vn(j)V_{n}^{(j)}, we know that

Up(j)=Vp(j)=0.U_{p}^{(j)}=V_{p}^{(j)}=0.

We then Taylor expand ff:

f(xUn(j))=f(x)f(x)Un(j)+O(Un(j))2,f(x-U_{n}^{(j)})=f(x)-f^{\prime}(x)U_{n}^{(j)}+O(U_{n}^{(j)})^{2},

and by (2.16), we have |Un(j)|2δ0222j2Mδ0M22jM|U_{n}^{(j)}|^{2}\lesssim\delta_{0}^{2}2^{-2j-2M}\ll\delta_{0}M^{-2}2^{j-M}. Thus we need to show

|(a0+δ0n2M)(f(x)f(x)Un(j))+Vn(j)(a0+δ0p2M)f(x)|δ0M22jM.|(a_{0}+\delta_{0}n2^{-M})(f(x)-f^{\prime}(x)U_{n}^{(j)})+V_{n}^{(j)}-(a_{0}+\delta_{0}p2^{-M})f(x)|\lesssim\delta_{0}M^{-2}2^{j-M}. (2.25)

We compute the left hand side using (2.20), (2.15), (2.7) and (2.9):

LHS of (2.25)
=i=1j(δ0εi2i1Mf(x)(a0+δ0n2M)f(x)uni(i)+vni(i))\displaystyle=\sum_{i=1}^{j}\left(\delta_{0}\varepsilon_{i}2^{i-1-M}f(x)-(a_{0}+\delta_{0}n2^{-M})f^{\prime}(x)u_{\left\lfloor n\right\rfloor_{i}}^{(i)}+v_{\left\lfloor n\right\rfloor_{i}}^{(i)}\right)
=δ02Mi=1j2i1(f(x)f(x)f(xi)f′′(ξi)+f(xi)2f′′(ξi)f(xi)+O(2iM))\displaystyle=\delta_{0}2^{-M}\sum_{i=1}^{j}2^{i-1}\left(f(x)-\frac{f^{\prime}(x)f^{\prime}(x_{i})}{f^{\prime\prime}(\xi_{i})}+\frac{f^{\prime}(x_{i})^{2}}{f^{\prime\prime}(\xi_{i})}-f(x_{i})+O(2^{i-M})\right)
=δ02Mi=1j2i1(f(x)f(xi)(f(x)f(xi))f(xi)f′′(ξi)+O(2iM)),\displaystyle=\delta_{0}2^{-M}\sum_{i=1}^{j}2^{i-1}\left(f(x)-f(x_{i})-(f^{\prime}(x)-f^{\prime}(x_{i}))\frac{f^{\prime}(x_{i})}{f^{\prime\prime}(\xi_{i})}+O(2^{i-M})\right),

where we abbreviated ξi:=ξni(i)\xi_{i}:=\xi^{(i)}_{\left\lfloor n\right\rfloor_{i}}. The sum of the quadratic error terms obeys

i=1j2i12iM22jMM22j,\sum_{i=1}^{j}2^{i-1}2^{i-M}\sim 2^{2j-M}\ll M^{-2}2^{j},

so it suffices to prove

i=1j2i|f(x)f(xi)(f(x)f(xi))f(xi)f′′(ξi)|2jM2.\sum_{i=1}^{j}2^{i}\left|f(x)-f(x_{i})-(f^{\prime}(x)-f^{\prime}(x_{i}))\frac{f^{\prime}(x_{i})}{f^{\prime\prime}(\xi_{i})}\right|\lesssim 2^{j}M^{-2}. (2.26)

To this end, we fix ii and let g(x)=f(x)f(x)f(xi)f′′(ξi)g(x)=f(x)-f^{\prime}(x)\frac{f^{\prime}(x_{i})}{f^{\prime\prime}(\xi_{i})}, so that we need to bound g(x)g(xi)g(x)-g(x_{i}). But by direct computation using (1.8),

|g(xi)|=|f(xi)f′′(ξi)(f′′(ξi)f′′(xi))||xxi|,|g^{\prime}(x_{i})|=\left|\frac{f^{\prime}(x_{i})}{f^{\prime\prime}(\xi_{i})}(f^{\prime\prime}(\xi_{i})-f^{\prime\prime}(x_{i}))\right|\lesssim|x-x_{i}|,

and so by Taylor expansion of gg, we have

|g(x)g(xi)||xxi|2.|g(x)-g(x_{i})|\lesssim|x-x_{i}|^{2}.

Thus we have reduced the problem to proving

i=1j2i|xxi|22jM2.\sum_{i=1}^{j}2^{i}|x-x_{i}|^{2}\lesssim 2^{j}M^{-2}.

However, using x[xj1,xj+1]x\in[x_{j-1},x_{j+1}] and the definition that xi=2i/Mx_{i}=2i/M, we obtain the bound. This finishes the proof.

3. Construction of zero measure Kakeya set

In this section, we use a routine method to construct the curved Kakeya set KK with zero measure mentioned in Theorem 1.3, based on the sets FMF_{M} constructed in the previous section.

Start with T(0)(a0=1,δ0=1)T^{(0)}(a_{0}=1,\delta_{0}=1) as defined in (2.1). Pick a large integer M2M\in 2\mathbb{N} such that we have all the results in Section 2. Denote

Mn:=Mn.M_{n}:=M^{n}. (3.1)

We then construct the set

K1:=FM1=FM,K_{1}:=F_{M_{1}}=F_{M}, (3.2)

which is a compact subset of [4logMM,1]×[C0,C0][\frac{4\log M}{M},1]\times[-C_{0},C_{0}] where C0=C0(f)C_{0}=C_{0}(f) is a fixed large constant. Denote A1:=4logMMA_{1}:=\frac{4\log M}{M}.

Recall K1K_{1} consists of 2M2^{M} smaller curved rectangles which we denote by T(1)T(1), each of the form T(0)(a,2M)+(u,v)T^{(0)}(a,2^{-M})+(u,v) for some a[1,2]a\in[1,2],

|u|C2M/2,|v|C2M/2|u|\leq C2^{-M/2},|v|\leq C2^{-M/2}

by (2.14). We then apply the same procedures in Section 2 to each T(1)T(1), this time with a0a_{0} taken to be this aa, MM taken to be M2M_{2}, and δ0\delta_{0} taken to be 2M12^{-M_{1}}. (More precisely, to fit the notation of Section 2 perfectly, we should first reverse the translation by (un,vn)(u_{n},v_{n}) to T(1)T(1), rescale so that it is over [0,1][0,1], apply the construction, and then rescale back in the end.) This gives us the set K2K_{2}, whose projection onto the xx-axis is equal to [A2,1][A_{2},1] where

A2:=1(14logM1M1)(14logM2M2).A_{2}:=1-\left(1-\frac{4\log M_{1}}{M_{1}}\right)\left(1-\frac{4\log M_{2}}{M_{2}}\right). (3.3)

Note that by (2.14) (and recall (1.4)), we have

K2(2M1M2)K1(2C2M12M2/2)K1(2M1).K_{2}(2^{-M_{1}-M_{2}})\subseteq K_{1}(2C2^{-M_{1}}2^{-M_{2}/2})\subseteq K_{1}(2^{-M_{1}}). (3.4)

We continue this process. Now K2K_{2} consists of 2M1+M22^{M_{1}+M_{2}} even smaller curved rectangles which we denote by T(2)T(2). We then apply the same procedures in Section 2 to each such T(2)T(2), this time with MM taken to be M3M_{3} and δ0\delta_{0} taken to be 2M1M22^{-M_{1}-M_{2}}. This gives us the set K3K_{3}, whose projection onto the xx-axis is equal to [A3,1][A_{3},1] where

A3:=1(14logM1M1)(14logM2M2)(14logM3M3).A_{3}:=1-\left(1-\frac{4\log M_{1}}{M_{1}}\right)\left(1-\frac{4\log M_{2}}{M_{2}}\right)\left(1-\frac{4\log M_{3}}{M_{3}}\right). (3.5)

By (2.14), we have

K3(2M1M2M3)K2(2C2M1M2M3/2)K2(2M1M2).K_{3}(2^{-M_{1}-M_{2}-M_{3}})\subseteq K_{2}(2C2^{-M_{1}-M_{2}-M_{3}/2})\subseteq K_{2}(2^{-M_{1}-M_{2}}). (3.6)

Continuing this process, we obtain a nested sequence of nonempty compact sets:

K1(2M1)K2(2M2M1)K3(2M3M2M1).K_{1}(2^{-M_{1}})\supseteq K_{2}(2^{-M_{2}-M_{1}})\supseteq K_{3}(2^{-M_{3}-M_{2}-M_{1}})\supseteq\cdots. (3.7)

We are now ready to define

K:=n=1Kn(2M1M2Mn),K:=\bigcap_{n=1}^{\infty}K_{n}(2^{-M_{1}-M_{2}-\cdots-M_{n}}), (3.8)

which is a nonempty compact set.

Theorem 3.1.

The following statements hold for KK.

  1. (1)

    KK has zero two dimensional Lebesgue measure.

  2. (2)

    The projection of KK onto the xx-axis contains an interval of positive length. Therefore, KK contains a translation of a piece of length 1\sim 1 of the graph of a function of the form af(x)af(x) where 1a21\leq a\leq 2.

  3. (3)

    |K(δ)|(logδ1)2|K(\delta)|\lesssim(\log\delta^{-1})^{-2}.

Proof.
  1. (1)

    We recall that KnK_{n} is the union of 2M1++Mn2^{M_{1}+\cdots+M_{n}} curved rectangles. By Theorem 2.3 and induction, we have

    |Kn|Mn2.|K_{n}|\lesssim M_{n}^{-2}.

    Thus

    |Kn(2C2M1MnMn+1/2)|Mn2+2Mn+1/22Mn2,\displaystyle|K_{n}(2C2^{-M_{1}-\cdots-M_{n}-M_{n+1}/2})|\lesssim M_{n}^{-2}+2^{-M_{n+1}/2}\leq 2M_{n}^{-2},

    which converges to 0 as nn\to\infty. Thus |K|=0|K|=0.

  2. (2)

    After the nn-th step, the projection of KnK_{n} onto the xx-axis is equal to [An,1][A_{n},1] where

    An:=1l=1n(14logMlMl).A_{n}:=1-\prod_{l=1}^{n}\left(1-\frac{4\log M_{l}}{M_{l}}\right). (3.9)

    Thus it suffices to prove that limnAn<1\lim_{n}A_{n}<1, which is equivalent to proving n4logMnMn<\sum_{n}\frac{4\log M_{n}}{M_{n}}<\infty. But by our choice of Mn=MnM_{n}=M^{n}, this follows.

  3. (3)

    Given δ(0,1)\delta\in(0,1), pick the unique nn such that

    2M1Mn<δ2M1Mn1.2^{-M_{1}-\cdots-M_{n}}<\delta\leq 2^{-M_{1}-\cdots-M_{n-1}}.

    Then we have

    |K(δ)||K(3C2M1Mn2Mn1/2)|Mn22.|K(\delta)|\leq|K(3C2^{-M_{1}-\cdots-M_{n-2}-M_{n-1}/2})|\lesssim M_{n-2}^{-2}. (3.10)

    On the other hand, using 2M1Mn<δ2^{-M_{1}-\cdots-M_{n}}<\delta, we have Mnlogδ1M^{n}\gtrsim\log\delta^{-1}. Hence we have Mn22(logδ1)2M_{n-2}^{-2}\lesssim(\log\delta^{-1})^{-2}.

4. Appendix

In this appendix, we provide a brief summary of the LpL^{p} to LqL^{q} boundedness of the maximal operator δ\mathcal{R}_{\delta} defined in (1.10), under an additional assumption that ff is smooth on [0,1][0,1].

Refer to caption
Figure 2. Interpolation diagram of δ\mathcal{R}_{\delta}
Theorem 4.1.

In the (1/p,1/q)(1/p,1/q)-interpolation diagram (see Figure 2), let OO be the origin, and

A=(0,1),B=(13,1),C=(38,1),D=(1,1),\displaystyle A=\left(0,1\right),\quad B=\left(\frac{1}{3},1\right),\quad C=\left(\frac{3}{8},1\right),\quad D=\left(1,1\right),
E=(1,0),F=(38,0),G=(38,516),H=(13,13).\displaystyle E=\left(1,0\right),\quad F=\left(\frac{3}{8},0\right),\quad G=\left(\frac{3}{8},\frac{5}{16}\right),\quad H=\left(\frac{1}{3},\frac{1}{3}\right).

Then we have the following estimates. Here all line segments and polygons below include their boundaries, and the implicit constants are allowed to depend on p,qp,q but not δ\delta.

R(p,q,δ)(1/p,1/q)1OA{εδε(logδ1)2/pOABH\OA{εδε+1232pδ1232pBCGH\BHδ1232pCDEG\CGδ1q1pEFG\FG{εδε+1q1pδ1q1pOFGH\OFδ1/pOE\begin{array}[]{|c|c|}\hline\cr R(p,q,\delta)&\forall(1/p,1/q)\in\\ \hline\cr\sim 1&OA\\ \hline\cr\begin{cases}\lesssim_{\varepsilon}\delta^{-\varepsilon}\\ \gtrsim(\log\delta^{-1})^{2/p}\end{cases}&OABH\backslash OA\\ \hline\cr\begin{cases}\lesssim_{\varepsilon}\delta^{-\varepsilon+\frac{1}{2}-\frac{3}{2p}}\\ \gtrsim\delta^{\frac{1}{2}-\frac{3}{2p}}\end{cases}&BCGH\backslash BH\\ \hline\cr\sim\delta^{\frac{1}{2}-\frac{3}{2p}}&CDEG\backslash CG\\ \hline\cr\sim\delta^{\frac{1}{q}-\frac{1}{p}}&EFG\backslash FG\\ \hline\cr\begin{cases}\lesssim_{\varepsilon}\delta^{-\varepsilon+\frac{1}{q}-\frac{1}{p}}\\ \gtrsim\delta^{\frac{1}{q}-\frac{1}{p}}\end{cases}&OFGH\backslash OF\\ \hline\cr\sim\delta^{-1/p}&OE\\ \hline\cr\end{array}
Proof.

We first come to the case p=p=\infty, namely, (1/p,1/q)OA(1/p,1/q)\in OA. Here the upper bound is trivial, and the lower bound follows from taking f=1B(0,1)f=1_{B(0,1)}.

For the case q=q=\infty, namely, (1/p,1/q)OE(1/p,1/q)\in OE, the upper bound follows from interpolating between p=1p=1 and p=p=\infty, and the lower bound follows from taking f=1Sf=1_{S} where

S={(x,af(x)):x[0,1],a[1,1+δ]}.S=\{(x,af(x)):x\in[0,1],a\in[1,1+\delta]\}. (4.1)

When (1/p,1/q)OABH\OA(1/p,1/q)\in OABH\backslash OA, to prove the upper bound, by Hölder’s inequality, it suffices to prove the bound on OHOH only. By [Zah12b], it holds at H=(1/3,1/3)H=(1/3,1/3). Then this follows from interpolating between (0,0)(0,0) and (1/3,1/3)(1/3,1/3). The lower bound follows from Theorem 1.3.

When (1/p,1/q)CDEG\CG(1/p,1/q)\in CDEG\backslash CG, the upper bound follows from [KW99]. The lower bound follows from taking f=1Tf=1_{T}, where TT is a rectangle of dimensions δ1/2×δ\delta^{1/2}\times\delta.

When (1/p,1/q)EFG\FG(1/p,1/q)\in EFG\backslash FG, the upper bound follows from interpolating between the points (1/p,0)(1/p,0) and (1p,1212p)(\frac{1}{p},\frac{1}{2}-\frac{1}{2p}). The lower bound follows from taking f=1Sf=1_{S}, where SS is given by (4.1).

When (1/p,1/q)BCGH\BH(1/p,1/q)\in BCGH\backslash BH, the upper bound follows from interpolating between BHBH and CGCG. The lower bound follows from taking f=1Tf=1_{T}, where TT is a rectangle of dimensions δ1/2×δ\delta^{1/2}\times\delta.

When (1/p,1/q)OFGH\OF(1/p,1/q)\in OFGH\backslash OF, the upper bound follows from interpolating between OFOF and the segments OH,HGOH,HG. The lower bound follows from taking f=1Sf=1_{S}, where SS is given by (4.1). ∎

4.1. Open problems

It is very natural to ask whether we can replace the ε\varepsilon-loss on the upper bounds of R(p,q,δ)R(p,q,\delta) to logarithmic loss (when we have a logarithmic lower bound), or to a constant loss (when we have a lower bound of the form δα\delta^{\alpha}, α>0\alpha>0, without ε\varepsilon-loss). There are two key open problems to consider.

  1. (1)

    Improving the upper bound of R(3,3,δ)R(3,3,\delta). One may work through all the details of [Sch03] and [Wol00] to see if each time a loss of the form δε\delta^{-\varepsilon} can be upgraded to just logδ1\log\delta^{-1}; if yes, then we can improve the bound to (logδ1)O(1)(\log\delta^{-1})^{O(1)}. However, improving to the lower bound (logδ1)2/p(\log\delta^{-1})^{2/p} proved in Theorem 1.3 seems to require a new method.

  2. (2)

    Removing the ε\varepsilon-loss in the range 13<1p38\frac{1}{3}<\frac{1}{p}\leq\frac{3}{8}. The condition p<8/3p<8/3 is needed in the proof in the combinatorial argument in [KW99], but as far as the authors know, there are no counterexample showing why p<8/3p<8/3 is necessary.

References

  • [BR68] A. S. Besicovitch and R. Rado. A plane set of measure zero containing circumferences of every radius. J. London Math. Soc., 43:717–719, 1968.
  • [CG24] M. Chen and S. Guo. The dichotomy of Nikodym sets and local smoothing estimates for wave equations. 2024. arXiv:2402.15476.
  • [CGY23] M. Chen, S. Guo, and T. Yang. A multi-parameter cinematic curvature. 2023. arXiv:2306.01606.
  • [CYZ23] X. Chen, L. Yan, and Y. Zhong. On the generalized Hausdorff dimension of Besicovitch sets. 2023. arXiv:2304.03633.
  • [Dav72] R. O. Davies. Another thin set of circles. J. London Math. Soc. (2), 5:191–192, 1972.
  • [HKLO23] S. Ham, H. Ko, S. Lee, and S. Oh. Remarks on dimension of unions of curves. Nonlinear Anal., 229:Paper No. 113207, 14, 2023.
  • [Kin68] J. R. Kinney. A thin set of circles. The American Mathematical Monthly, 75(10):1077–1081, 1968.
  • [KW99] L. Kolasa and T. Wolff. On some variants of the Kakeya problem. Pacific J. Math., 190(1):111–154, 1999.
  • [PYZ22] M. Pramanik, T. Yang, and J. Zahl. A Furstenberg-type problem for circles, and a Kaufman-type restricted projection theorem in 3\mathbb{R}^{3}. 2022. arXiv:2207.02259.
  • [Sch03] W. Schlag. On continuum incidence problems related to harmonic analysis. J. Funct. Anal., 201(2):480–521, 2003.
  • [Sog91] C. D. Sogge. Propagation of singularities and maximal functions in the plane. Invent. Math., 104(2):349–376, 1991.
  • [Tal80] M. Talagrand. Sur la mesure de la projection d’un compact et certaines familles de cercles. Bull. Sci. Math. (2), 104(3):225–231, 1980.
  • [Wol97] T. Wolff. A Kakeya-type problem for circles. Amer. J. Math., 119(5):985–1026, 1997.
  • [Wol00] T. Wolff. Local smoothing type estimates on LpL^{p} for large pp. Geom. Funct. Anal., 10(5):1237–1288, 2000.
  • [Zah12a] J. Zahl. L3L^{3} estimates for an algebraic variable coefficient Wolff circular maximal function. Rev. Mat. Iberoam., 28(4):1061–1090, 2012.
  • [Zah12b] J. Zahl. On the Wolff circular maximal function. Illinois J. Math., 56(4):1281–1295, 2012.
  • [Zah23] J. Zahl. On maximal functions associated to families of curves in the plane. 2023. arXiv:2307.05894.