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

Billiards in polyhedra: a method
to convert 2-dimensional uniformity
to 3-dimensional uniformity

J. Beck Department of Mathematics, Hill Center for the Mathematical Sciences, Rutgers University, Piscataway NJ 08854, USA [email protected] W.W.L. Chen School of Mathematical and Physical Sciences, Faculty of Science and Engineering, Macquarie University, Sydney NSW 2109, Australia [email protected]  and  Y. Yang School of Science, Beijing University of Posts and Telecommunications, Beijing 100876, China [email protected]
Abstract.

The class of 22-dimensional non-integrable flat dynamical systems has a rather extensive literature with many deep results, but the methods developed for this type of problems, both the traditional approach via Teichmüller geometry and our recent shortline-ancestor method, appear to be exclusively plane-specific. Thus we know very little of any real significance concerning 33-dimensional systems.

Our purpose here is to describe some very limited extensions of uniformity in 22 dimensions to uniformity in 33 dimensions. We consider a 33-manifold which is the cartesian product of the regular octagonal surface with the unit torus. This is a restricted system, in the sense that one of the directions is integrable. However, this restriction also allows us to make use of a transference theorem for arithmetic progressions established earlier by Beck, Donders and Yang.

Key words and phrases:
geodesics, billiards, uniformity
2010 Mathematics Subject Classification:
11K38, 37E35

1. Introduction

For rational polygons where every angle is a rational multiple of π\pi, we have the following fundamental result of Kerckhoff, Masur and Smillie [4] in 1986.

Theorem A.

Let PP be a rational polygon. For almost every initial direction and for every non-pathological starting point for this direction, the half-infinite billiard orbit in PP is uniformly distributed.

Given any initial direction, a point 𝐩0P\mathbf{p}_{0}\in P is called a pathological starting point for this direction if the half-infinite billiard orbit starting from 𝐩0\mathbf{p}_{0} and with this direction hits a singularity of PP. Otherwise the point 𝐩0\mathbf{p}_{0} is called a non-pathological starting point for this direction. It is easy to see that for any given direction, almost every point in PP is a non-pathological starting point.

The proof of Theorem A consists of essentially three steps. The first step is to establish the ergodicity of the corresponding interval exchange transformation. The second step is to use the well known Birkhoff ergodic theorem. The final step is to extend ergodicity to unique ergodicity.

Our aim is to convert Theorem A to a result concerning equidistribution of 33-dimensional billiard in some polyhedra. However, we need to restrict our discussion to rational polygonal right prisms. A rational polygonal right prism is a region in 33-dimensional cartesian space of the form

M=P×I={(x,y,z)3:(x,y)P and zI},M=P\times I=\{(x,y,z)\in\mathbb{R}^{3}:(x,y)\in P\mbox{ and }z\in I\}, (1.1)

where PP is a rational polygon and I=[0,z0]I=[0,z_{0}] is an interval.

As the rational polygonal right prism M=P×IM=P\times I is integrable in the direction of the interval II, our extension is somewhat limited.

Theorem 1.

Let MM be a rational polygonal right prism of the form (1.1), where PP is a rational polygon and I=[0,z0]I=[0,z_{0}] is an interval. For almost every pair of initial direction and starting point, the half-infinite billiard orbit in MM is uniformly distributed.

For illustration, we consider a special case where PP is a right triangle. It is well known that the right triangle billiard with angle π/4\pi/4 and the right triangle billiard with angle π/6\pi/6 are the only right triangle billiards that are integrable, exhibiting stable and predictable behaviour. Perhaps the simplest non-integrable billiard is the right triangle billiard with angle π/8\pi/8. It is also well known that unfolding in the spirit of König and Szücs [5] leads to a 1616-fold covering of the triangle and shows that this billiard is equivalent to geodesic flow on the regular octagon surface 𝒫\mathcal{P} where parallel edges are identified in pairs. On the other hand, unfolding also leads to a 22-fold covering of the interval I=[0,z0]I=[0,z_{0}]. Thus billiard in the rational polygonal right prism M=P×IM=P\times I, where PP is the right triangle with angle π/8\pi/8 and I=[0,z0]I=[0,z_{0}], is equivalent to geodesic flow in the translation 33-manifold =𝒫×\mathcal{M}=\mathcal{P}\times\mathcal{I}, where 𝒫\mathcal{P} is the regular octagon translation surface and =[0,2z0]\mathcal{I}=[0,2z_{0}], treated as a torus. Figure 1 illustrates that =𝒫×\mathcal{M}=\mathcal{P}\times\mathcal{I} gives a 3232-fold covering of the rational polygonal right prism M=P×IM=P\times I. It has 22 octagonal faces which are identified with each other, and 88 rectangular faces, with pairs of parallel ones identified with each other, analogous to the edge identification of the regular octagon translation surface 𝒫\mathcal{P}.

[Uncaptioned image]Figure 1: the translation 3-manifold =𝒫×\begin{array}[]{c}\includegraphics[scale={0.8}]{figure-1-1.pdf}\vspace{3pt}\\ \mbox{Figure 1: the translation $3$-manifold $\mathcal{M}=\mathcal{P}\times\mathcal{I}$}\end{array}
Theorem 2.

Let =𝒫×\mathcal{M}=\mathcal{P}\times\mathcal{I} be a rational octagonal right prism translation 33-manifold, where 𝒫\mathcal{P} is the regular octagon translation surface and =[0,2z0]\mathcal{I}=[0,2z_{0}], treated as a torus. For almost every pair of direction and starting point, the half-infinite geodesic in \mathcal{M} is uniformly distributed.

2. Proof of Theorem 2

Suppose that a half-infinite geodesic

S0,𝐯(t)=(s1+v1t,s2+v2t,s3+v3t),t0,\mathcal{L}_{S_{0},\mathbf{v}}(t)=(s_{1}+v_{1}t,s_{2}+v_{2}t,s_{3}+v_{3}t),\quad t\geqslant 0,

in \mathcal{M} has a non-pathological starting point S0,𝐯(0)=S0=(s1,s2,s3)\mathcal{L}_{S_{0},\mathbf{v}}(0)=S_{0}=(s_{1},s_{2},s_{3}), and direction given by the unit vector

𝐯=(v1,v2,v3)3,wherev12+v22+v32=1,\mathbf{v}=(v_{1},v_{2},v_{3})\in\mathbb{R}^{3},\quad\mbox{where}\quad v_{1}^{2}+v_{2}^{2}+v_{3}^{2}=1,

with arc-length parametrization. The coordinates (s1+v1t,s2+v2t)(s_{1}+v_{1}t,s_{2}+v_{2}t) are modulo 𝒫\mathcal{P} and the coordinate s3+v3ts_{3}+v_{3}t is modulo \mathcal{I}.

We may assume without loss of generality that v3>0v_{3}>0. Then the geodesic S0,𝐯(t)\mathcal{L}_{S_{0},\mathbf{v}}(t), t0t\geqslant 0, hits the octagon face of \mathcal{M} for the very first time at time t=t0t=t_{0}, where s3+v3t0=2z0s_{3}+v_{3}t_{0}=2z_{0}, so that t0=(2z0s3)/v3t_{0}=(2z_{0}-s_{3})/v_{3}. Indeed, the geodesic hits the octagon face of \mathcal{M} for the (k+1)(k+1)-th time at time t=tkt=t_{k}, where

tk=2kz0+(2z0s3)v3=kθ+λ,k=0,1,2,3,,t_{k}=\frac{2kz_{0}+(2z_{0}-s_{3})}{v_{3}}=k\theta+\lambda,\quad k=0,1,2,3,\ldots, (2.1)

with parameters xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

θ=2z0v3andλ=2z0s3v3.\theta=\frac{2z_{0}}{v_{3}}\quad\mbox{and}\quad\lambda=\frac{2z_{0}-s_{3}}{v_{3}}.

This gives rise to an arithmetic progression

λ<θ+λ<2θ+λ<3θ+λ<,\lambda<\theta+\lambda<2\theta+\lambda<3\theta+\lambda<\ldots,

with common gap θ\theta between consecutive terms.

We need the following result on arithmetic progressions; see [1, Theorem 2.2.2].

Lemma 2.1 (transference theorem for arithmetic progressions).

Let 𝒮\mathcal{S}\subset\mathbb{R} be a measurable set. For every \ell\in\mathbb{Z}, there exists a constant c1()>0c_{1}(\ell)>0, dependent only on \ell, such that for almost every pair θ,λ\theta,\lambda satisfying 2θ<2+12^{\ell}\leqslant\theta<2^{\ell+1} and 0λ<θ0\leqslant\lambda<\theta, the inequality

|k0kθ+λ𝒮[0,n]11θmeas(𝒮[0,n])|c1()n3/4(logn)1/2\left|\sum_{\begin{subarray}{c}{k\geqslant 0}\\ {k\theta+\lambda\in\mathcal{S}\cap[0,n]}\end{subarray}}1-\frac{1}{\theta}\operatorname{meas}(\mathcal{S}\cap[0,n])\right|\leqslant c_{1}(\ell)n^{3/4}(\log n)^{1/2}

holds for every sufficiently large positive integer nn.

We have following immediate consequence.

Lemma 2.2.

Suppose that the set 𝒮\mathcal{S}\subset\mathbb{R} is measurable. Suppose further that 𝒮\mathcal{S} has asymptotic density d=d(𝒮)[0,1]d=d(\mathcal{S})\in[0,1], so that there exists a monotonic sequence ε(n)=ε(𝒮;n)0\varepsilon(n)=\varepsilon(\mathcal{S};n)\to 0 as nn\to\infty such that

|1nmeas(𝒮[0,n])d(𝒮)|<ε(n),n=1,2,3,.\left|\frac{1}{n}\operatorname{meas}(\mathcal{S}\cap[0,n])-d(\mathcal{S})\right|<\varepsilon(n),\quad n=1,2,3,\ldots.

For every \ell\in\mathbb{Z}, there exists a constant c2()>0c_{2}(\ell)>0, dependent only on \ell, such that for almost every pair θ,λ\theta,\lambda satisfying 2θ<2+12^{\ell}\leqslant\theta<2^{\ell+1} and 0λ<θ0\leqslant\lambda<\theta, the inequality

|θnk0kθ+λ𝒮[0,n]1d(𝒮)|ε(n)+c2()(logn)1/2n1/4\left|\frac{\theta}{n}\sum_{\begin{subarray}{c}{k\geqslant 0}\\ {k\theta+\lambda\in\mathcal{S}\cap[0,n]}\end{subarray}}1-d(\mathcal{S})\right|\leqslant\varepsilon(n)+c_{2}(\ell)\frac{(\log n)^{1/2}}{n^{1/4}} (2.2)

holds for every sufficiently large positive integer nn.

Meanwhile, every point in \mathcal{M} is of the form (x,y,z)(x,y,z), where (x,y)𝒫(x,y)\in\mathcal{P} and zz\in\mathcal{I}. We consider the projection

ϕ:𝒫:(x,y,z)(x,y).\phi:\mathcal{M}\to\mathcal{P}:(x,y,z)\mapsto(x,y). (2.3)

Then the image of the geodesic S0,𝐯(t)\mathcal{L}_{S_{0},\mathbf{v}}(t), t0t\geqslant 0, under this projection is a half-infinite geodesic xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

S0,𝐯(t)=(s1+v1t,s2+v2t),t0,\mathcal{H}_{S_{0},\mathbf{v}}(t)=(s_{1}+v_{1}t,s_{2}+v_{2}t),\quad t\geqslant 0, (2.4)

on the regular octagon translation surface 𝒫\mathcal{P}. Clearly the key parameters s3s_{3} and v3v_{3}, particularly concerning the hitting times given in (2.1), are lost under this projection (2.3). However, we know the arithmetic progression (2.1) of the time instances when the geodesic S0,𝐯(t)\mathcal{L}_{S_{0},\mathbf{v}}(t), t0t\geqslant 0, hits the octagon face of \mathcal{M}. This gives rise to an infinite sequence of points S0,𝐯(tk)\mathcal{H}_{S_{0},\mathbf{v}}(t_{k}), k=0,1,2,3,,k=0,1,2,3,\ldots, on 𝒫\mathcal{P}. For any S0S_{0} and 𝐯\mathbf{v}, if we can show that this sequence of points is uniformly distributed on 𝒫\mathcal{P}, then the half-infinite geodesic S0,𝐯(t)\mathcal{L}_{S_{0},\mathbf{v}}(t), t0t\geqslant 0, is uniformly distributed in \mathcal{M}.

Consider a typical half-infinite geodesic 𝐰(τ)\mathcal{H}_{\mathbf{w}}(\tau), τ0\tau\geqslant 0, on 𝒫\mathcal{P}, with direction given by the unit vector 𝐰=(w1,w2)2\mathbf{w}=(w_{1},w_{2})\in\mathbb{R}^{2}. Suppose that this geodesic is the image on 𝒫\mathcal{P} of S0,𝐯(t)\mathcal{L}_{S_{0},\mathbf{v}}(t), t0t\geqslant 0, under the projection (2.3). Then

𝐰(τ)=(s1+w1τ,s2+w2τ),τ0.\mathcal{H}_{\mathbf{w}}(\tau)=(s_{1}+w_{1}\tau,s_{2}+w_{2}\tau),\quad\tau\geqslant 0. (2.5)

In view of the different parametrizations of (2.4) and (2.5), we have

S0,𝐯(t)=𝐰(τ)if and only ifτ=(v12+v22)1/2t.\mathcal{H}_{S_{0},\mathbf{v}}(t)=\mathcal{H}_{\mathbf{w}}(\tau)\quad\mbox{if and only if}\quad\tau=(v_{1}^{2}+v_{2}^{2})^{1/2}t.

Corresponding to the arithmetic progression tkt_{k}, k=0,1,2,3,,k=0,1,2,3,\ldots, of hitting times given by (2.1) is the arithmetic progression

τk=(v12+v22)1/2tk=(v12+v22)1/2(2kz0s3)v3=kθ+λ,k=0,1,2,3,,\tau_{k}=(v_{1}^{2}+v_{2}^{2})^{1/2}t_{k}=\frac{(v_{1}^{2}+v_{2}^{2})^{1/2}(2kz_{0}-s_{3})}{v_{3}}=k\theta+\lambda,\quad k=0,1,2,3,\ldots,

with parameters xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

θ=2z0(v12+v22)1/2v3andλ=(2z0s3)(v12+v22)1/2v3.\theta=\frac{2z_{0}(v_{1}^{2}+v_{2}^{2})^{1/2}}{v_{3}}\quad\mbox{and}\quad\lambda=\frac{(2z_{0}-s_{3})(v_{1}^{2}+v_{2}^{2})^{1/2}}{v_{3}}. (2.6)

By the geodesic analogue of Theorem A, for almost every direction 𝐰\mathbf{w} and for every non-pathological starting point for this direction, the geodesic 𝐰(τ)\mathcal{H}_{\mathbf{w}}(\tau), τ0\tau\geqslant 0, is uniformly distributed on 𝒫\mathcal{P}. Let RP\operatorname{RP} denote an arbitrary polygon on 𝒫\mathcal{P} where all the vertices have rational coordinates, and let the measurable set

𝒮=𝒮(𝐰;RP)={τ0:𝐰(τ)RP}\mathcal{S}=\mathcal{S}(\mathcal{H}_{\mathbf{w}};\operatorname{RP})=\{\tau\geqslant 0:\mathcal{H}_{\mathbf{w}}(\tau)\in\operatorname{RP}\} (2.7)

denote the set of time instances when this geodesic visits RP\operatorname{RP}. The uniformity of the geodesic then implies that 𝒮\mathcal{S} has asymptotic density

d=d(𝒮)=area(RP)area(𝒫)[0,1].d=d(\mathcal{S})=\frac{\operatorname{area}(\operatorname{RP})}{\operatorname{area}(\mathcal{P})}\in[0,1].

The uniformly distributed geodesic 𝐰(τ)\mathcal{H}_{\mathbf{w}}(\tau), τ0\tau\geqslant 0, is clearly the image on 𝒫\mathcal{P} under the projection (2.3) of infinitely many different geodesics S0,𝐯(t)\mathcal{L}_{S_{0},\mathbf{v}}(t), t0t\geqslant 0, in the 33-manifold \mathcal{M}, as there are only two requirements, namely 𝐰(0)=(s1,s2)\mathcal{H}_{\mathbf{w}}(0)=(s_{1},s_{2}) concerning the starting point, where S0=(s1,s2,s3)S_{0}=(s_{1},s_{2},s_{3}), and w1v2=w2v1w_{1}v_{2}=w_{2}v_{1} concerning equality of the relevant directions. Applying Lemma 2.2, we see that for every \ell\in\mathbb{Z} and for almost every pair θ,λ\theta,\lambda of the form (2.6) satisfying 2θ<2+12^{\ell}\leqslant\theta<2^{\ell+1} and 0λ<θ0\leqslant\lambda<\theta, the inequality (2.2) with 𝒮\mathcal{S} given by (2.7) holds for every sufficiently large positive integer nn.

This means that for almost every pair of starting point S0S_{0} and unit direction vector 𝐯\mathbf{v}, the infinite sequence

S0;𝐯(tk),k=0,1,2,3,,\mathcal{L}_{S_{0};\mathbf{v}}(t_{k}),\quad k=0,1,2,3,\ldots, (2.8)

of points, where the sequence tkt_{k}, k=0,1,2,3,,k=0,1,2,3,\ldots, of time instances is given by (2.1), is uniformly distributed on 𝒫\mathcal{P} relative to the single test set RP\operatorname{RP}.

The set of all polygons RP\operatorname{RP} on 𝒫\mathcal{P} where all the vertices have rational coordinates is countable. On the other hand, a countable union of sets of measure zero has measure zero. It follows that for almost every pair of starting point S0S_{0} and unit direction vector 𝐯\mathbf{v}, the infinite sequence (2.8) of points is uniformly distributed on 𝒫\mathcal{P} relative to every polygon RP\operatorname{RP} on 𝒫\mathcal{P} where all the vertices have rational coordinates. This guarantees uniformity in general, in the classical Weyl sense, and completes the proof of Theorem 2.

Remark.

Theorem 1 is a result on time-qualitative uniformity, and does not say anything about the speed of convergence to uniform distribution, as a key ingredient of the proof is the geodesic analogue of Theorem A which is also time-qualitative in nature. There are instances, however, when we can establish time-quantitative results. This happens, for example, when it is possible to establish time-quantitative uniform distribution results for some geodesics on the underlying rational polygonal translation surface 𝒫\mathcal{P}. We can establish such extensions of the geodesic analogue of Theorem A in [1, 2] for the L-surface and in [3] for finite polysquare translation surfaces and for regular polygonal translation surfaces. These in turn lead to extensions of various analogues of Theorem 2, and hence also Theorem 1, to time-quantitative results.

3. Proof of Lemma 2.1

Throughout the proof, the set 𝒮\mathcal{S}\subset\mathbb{R} is measurable.

Let a non-negative integer \ell be chosen and fixed.

Consider an infinite sequence N1,N2,N3,N_{1},N_{2},N_{3},\ldots of positive integers satisfying

1<N1<N2<N3<<Nh<,1<N_{1}<N_{2}<N_{3}<\ldots<N_{h}<\ldots,

and another infinite sequence M1,M2,M3,M_{1},M_{2},M_{3},\ldots of positive integers satisfying

1<Mh<Nh,h=1,2,3,,1<M_{h}<N_{h},\quad h=1,2,3,\ldots,

both to be specified later in terms of the parameter hh and the chosen integer \ell. For every positive integer hh, let S(h)[0,1]S(h)\subset[0,1] denote the contraction of 𝒮[0,Nh]\mathcal{S}\cap[0,N_{h}] to the unit interval, so that

xS(h)if and only ifNhx𝒮[0,Nh].x\in S(h)\quad\mbox{if and only if}\quad N_{h}x\in\mathcal{S}\cap[0,N_{h}]. (3.1)

Since the characteristic function

χS(h)(x)={1,if xS(h),0,if xS(h),\chi_{S(h)}(x)=\left\{\begin{array}[]{ll}1,&\mbox{if $x\in S(h)$},\\ 0,&\mbox{if $x\not\in S(h)$},\end{array}\right.

defined over [0,1][0,1] and extended periodically over the whole real line with period 11, is measurable, we can consider its Fourier series

χS(h)(x)=jaje2πijx,\chi_{S(h)}(x)=\sum_{j\in\mathbb{Z}}a_{j}\mathrm{e}^{2\pi\mathrm{i}jx}, (3.2)

with Fourier coefficients aja_{j}, jj\in\mathbb{Z}. In particular,

a0=λ1(S(h)).a_{0}=\lambda_{1}(S(h)).
Remark.

For a measurable set S(h)S(h), the infinite Fourier series (3.2) may diverge at some points. However, Lemma 2.1 is a measure theoretic statement which ignores sets of measure zero. So it suffices to have pointwise convergence almost everywhere. Fourier analysis provides at least two options to settle this issue. We can use the very deep Carleson theorem. Alternatively, we can use the much simpler Lebesgue theorem with Cesàro summability.

The Parseval formula gives

j|aj|2=λ1(S(h))1,\sum_{j\in\mathbb{Z}}|a_{j}|^{2}=\lambda_{1}(S(h))\leqslant 1,

so that

j{0}|aj|2=λ1(S(h))λ12(S(h))=λ1(S(h))(1λ1(S(h)))<1.\sum_{j\in\mathbb{Z}\setminus\{0\}}|a_{j}|^{2}=\lambda_{1}(S(h))-\lambda_{1}^{2}(S(h))=\lambda_{1}(S(h))(1-\lambda_{1}(S(h)))<1. (3.3)

Lemma 2.1 for the chosen value \ell concerns the arithmetic progression kθ+ηk\theta+\eta, k0k\geqslant 0, where 2θ<2+12^{\ell}\leqslant\theta<2^{\ell+1} and 0η<θ0\leqslant\eta<\theta. The contraction (3.1) leads to a new arithmetic progression kα+βk\alpha+\beta, k0k\geqslant 0, where 2/Nhα<2+1/Nh2^{\ell}/N_{h}\leqslant\alpha<2^{\ell+1}/N_{h} and 0β<α0\leqslant\beta<\alpha.

For any α[2/Nh,2+1/Nh)\alpha\in[2^{\ell}/N_{h},2^{\ell+1}/N_{h}), let K(α)K(\alpha) be the unique integer satisfying

(K(α)1)α<1K(α)α.(K(\alpha)-1)\alpha<1\leqslant K(\alpha)\alpha. (3.4)

Using the Fourier series (3.2), we have

k=0K(α)1χS(h)(kα+β)K(α)λ1(S(h))=j{0}ajk=0K(α)1e2πij(kα+β)\sum_{k=0}^{K(\alpha)-1}\chi_{S(h)}(k\alpha+\beta)-K(\alpha)\lambda_{1}(S(h))=\sum_{j\in\mathbb{Z}\setminus\{0\}}a_{j}\sum_{k=0}^{K(\alpha)-1}\mathrm{e}^{2\pi\mathrm{i}j(k\alpha+\beta)} (3.5)

for every α\alpha and β\beta satisfying 2/Nhα<2+1/Nh2^{\ell}/N_{h}\leqslant\alpha<2^{\ell+1}/N_{h} and 0β<α0\leqslant\beta<\alpha.

To study (3.5), we consider the integral

J(𝐚;Nh;Mh)=2/Nh2+1/Nh21/Mh21/Mh|j{0}ajk=0K(α)1e2πij(kα+γ)|2dγdα.J(\mathbf{a};N_{h};M_{h})=\int_{2^{\ell}/N_{h}}^{2^{\ell+1}/N_{h}}\!\!\int_{-2^{\ell-1}/M_{h}}^{2^{\ell-1}/M_{h}}\left|\sum_{j\in\mathbb{Z}\setminus\{0\}}a_{j}\sum_{k=0}^{K(\alpha)-1}\mathrm{e}^{2\pi\mathrm{i}j(k\alpha+\gamma)}\right|^{2}\mathrm{d}\gamma\,\mathrm{d}\alpha. (3.6)

To obtain a bound on this integral, we observe that for γ[21/Mh,21/Mh]\gamma\in[-2^{\ell-1}/M_{h},2^{\ell-1}/M_{h}], the inequality 2(1|γ|Mh/2)12(1-|\gamma|M_{h}/2^{\ell})\geqslant 1 holds. It then follows that

J(𝐚;Nh;Mh)2J(𝐚;Nh;Mh),J(\mathbf{a};N_{h};M_{h})\leqslant 2J^{*}(\mathbf{a};N_{h};M_{h}), (3.7)

where for integers NN and MM satisfying 1<M<N1<M<N,

J(𝐚;N;M)\displaystyle J^{*}(\mathbf{a};N;M)
=2/N2+1/N2/M2/M|j{0}ajk=0K(α)1e2πij(kα+γ)|2(1|γ|M2)dγdα.\displaystyle\quad=\int_{2^{\ell}/N}^{2^{\ell+1}/N}\!\!\int_{-2^{\ell}/M}^{2^{\ell}/M}\left|\sum_{j\in\mathbb{Z}\setminus\{0\}}a_{j}\sum_{k=0}^{K(\alpha)-1}\mathrm{e}^{2\pi\mathrm{i}j(k\alpha+\gamma)}\right|^{2}\left(1-\frac{|\gamma|M}{2^{\ell}}\right)\,\mathrm{d}\gamma\,\mathrm{d}\alpha. (3.8)

At the end of this section, we establish the following bound for this integral.

Lemma 3.1.

For any sequence 𝐚\mathbf{a} satisfying (3.3), the inequality

J(𝐚;N;M)2+11.J^{*}(\mathbf{a};N;M)\leqslant 2^{\ell+11}. (3.9)

holds uniformly for integers MM and NN satisfying 1<M<N1<M<N, where K(α)K(\alpha) is the integer defined by (3.4) for every α[2/N,2+1/N)\alpha\in[2^{\ell}/N,2^{\ell+1}/N).

Combining (3.6), (3.7) and (3.9), we deduce that

J(𝐚;Nh;Mh)2+12.J(\mathbf{a};N_{h};M_{h})\leqslant 2^{\ell+12}. (3.10)

The inequality (3.10) is a quadratic average result, from which we can derive information concerning the majority of pairs

(α,γ)[2Nh,2+1Nh)×[21Mh,21Mh].(\alpha,\gamma)\in\left[\frac{2^{\ell}}{N_{h}},\frac{2^{\ell+1}}{N_{h}}\right)\times\left[-\frac{2^{\ell-1}}{M_{h}},\frac{2^{\ell-1}}{M_{h}}\right]. (3.11)

Let xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

B(Nh;Mh)={(α,γ)[2Nh,2+1Nh)×[21Mh,21Mh]:(3.12) holds}B(N_{h};M_{h})=\left\{(\alpha,\gamma)\in\left[\frac{2^{\ell}}{N_{h}},\frac{2^{\ell+1}}{N_{h}}\right)\times\left[-\frac{2^{\ell-1}}{M_{h}},\frac{2^{\ell-1}}{M_{h}}\right]:\mbox{\eqref{eq3.12} holds}\right\}

denote the collection of pairs (α,γ)(\alpha,\gamma) satisfying (3.11) such that

|j{0}ajk=0K(α)1e2πij(kα+γ)|(hNhMh)1/2log(1+h).\left|\sum_{j\in\mathbb{Z}\setminus\{0\}}a_{j}\sum_{k=0}^{K(\alpha)-1}\mathrm{e}^{2\pi\mathrm{i}j(k\alpha+\gamma)}\right|\geqslant(hN_{h}M_{h})^{1/2}\log(1+h). (3.12)

Then it follows from (3.10) that

λ2(B(Nh;Mh))2+12hNhMhlog2(1+h),\lambda_{2}(B(N_{h};M_{h}))\leqslant\frac{2^{\ell+12}}{hN_{h}M_{h}\log^{2}(1+h)}, (3.13)

where λ2\lambda_{2} denotes 22-dimensional Lebesgue measure.

Next, note that as we move from (3.5) to (3.6), we replace the parameter β\beta over a short interval [0,α)[0,\alpha) by a parameter γ[21/Mh,21/Mh]\gamma\in[-2^{\ell-1}/M_{h},2^{\ell-1}/M_{h}] over a longer interval. For any pair xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

(α,β)[2Nh,2+1Nh)×[0,α),(\alpha,\beta)\in\left[\frac{2^{\ell}}{N_{h}},\frac{2^{\ell+1}}{N_{h}}\right)\times[0,\alpha), (3.14)

there are at least Nh/2MhN_{h}/2M_{h} values of γ[21/Mh,21/Mh]\gamma\in[-2^{\ell-1}/M_{h},2^{\ell-1}/M_{h}] where {γ/α}α=β\{\gamma/\alpha\}\alpha=\beta. For each of these values of γ\gamma, consider the two arithmetic progressions

kα+γ,k=0,1,2,3,,K(α)1,k\alpha+\gamma,\quad k=0,1,2,3,\ldots,K(\alpha)-1, (3.15)

and xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

kα+β,k=0,1,2,3,,K(α)1.k\alpha+\beta,\quad k=0,1,2,3,\ldots,K(\alpha)-1. (3.16)

Since xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

γ=[γα]α+β,\gamma=\left[\frac{\gamma}{\alpha}\right]\alpha+\beta,

the arithmetic progression (3.15) is obtained by simply advancing the arithmetic progression (3.16) by [γ/α][\gamma/\alpha] terms. More precisely, the arithmetic progression (3.15) is given by

kα+β,k=[γα],[γα]+1,[γα]+2,[γα]+3,,[γα]+K(α)1.k\alpha+\beta,\quad k=\left[\frac{\gamma}{\alpha}\right],\left[\frac{\gamma}{\alpha}\right]+1,\left[\frac{\gamma}{\alpha}\right]+2,\left[\frac{\gamma}{\alpha}\right]+3,\ldots,\left[\frac{\gamma}{\alpha}\right]+K(\alpha)-1. (3.17)
Lemma 3.2.

If a pair (α,β)(\alpha,\beta) such that (3.14) holds satisfies the inequality

|j{0}ajk=0K(α)1e2πij(kα+β)|2NhMh+(hNhMh)1/2log(1+h),\left|\sum_{j\in\mathbb{Z}\setminus\{0\}}a_{j}\sum_{k=0}^{K(\alpha)-1}\mathrm{e}^{2\pi\mathrm{i}j(k\alpha+\beta)}\right|\geqslant\frac{2N_{h}}{M_{h}}+(hN_{h}M_{h})^{1/2}\log(1+h), (3.18)

then each pair (α,γ)(\alpha,\gamma) such that (3.11) and {γ/α}α=β\{\gamma/\alpha\}\alpha=\beta hold satisfies the inequality (3.12).

Proof.

It clearly suffices to prove that

|j{0}ajk=0K(α)1e2πij(kα+β)j{0}ajk=0K(α)1e2πij(kα+γ)|2NhMh.\left|\sum_{j\in\mathbb{Z}\setminus\{0\}}a_{j}\sum_{k=0}^{K(\alpha)-1}\mathrm{e}^{2\pi\mathrm{i}j(k\alpha+\beta)}-\sum_{j\in\mathbb{Z}\setminus\{0\}}a_{j}\sum_{k=0}^{K(\alpha)-1}\mathrm{e}^{2\pi\mathrm{i}j(k\alpha+\gamma)}\right|\leqslant\frac{2N_{h}}{M_{h}}. (3.19)

Since xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

|[γα]|Nh2Mh,\left|\left[\frac{\gamma}{\alpha}\right]\right|\leqslant\frac{N_{h}}{2M_{h}},

it follows from (3.16) and (3.17) that those terms that belong to one of the arithmetic progressions (3.15) or (3.16) but not both then form two arithmetic progressions of the form xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

kα+ρ,k=0,1,2,3,,K1,k\alpha+\rho,\quad k=0,1,2,3,\ldots,K-1,

where KNh/2MhK\leqslant N_{h}/2M_{h}. Hence

k=0K(α)1e2πij(kα+β)k=0K(α)1e2πij(kα+γ)\sum_{k=0}^{K(\alpha)-1}\mathrm{e}^{2\pi\mathrm{i}j(k\alpha+\beta)}-\sum_{k=0}^{K(\alpha)-1}\mathrm{e}^{2\pi\mathrm{i}j(k\alpha+\gamma)}

is the sum of two sums of the form

k=0K1e2πij(kα+ρ),\sum_{k=0}^{K-1}\mathrm{e}^{2\pi\mathrm{i}j(k\alpha+\rho)},

where KNh/2MhK\leqslant N_{h}/2M_{h}. Now for each of the two sums, we have

|j{0}ajk=0K1e2πij(kα+ρ)|\displaystyle\left|\sum_{j\in\mathbb{Z}\setminus\{0\}}a_{j}\sum_{k=0}^{K-1}\mathrm{e}^{2\pi\mathrm{i}j(k\alpha+\rho)}\right| =|k=0K1j{0}aje2πij(kα+ρ)|\displaystyle=\left|\sum_{k=0}^{K-1}\sum_{j\in\mathbb{Z}\setminus\{0\}}a_{j}\mathrm{e}^{2\pi\mathrm{i}j(k\alpha+\rho)}\right|
=|k=0K1(jaje2πij(kα+ρ)a0)|\displaystyle=\left|\sum_{k=0}^{K-1}\left(\sum_{j\in\mathbb{Z}}a_{j}\mathrm{e}^{2\pi\mathrm{i}j(k\alpha+\rho)}-a_{0}\right)\right|
k=0K1|χS(h)(kα+ρ)λ1(Sh)|\displaystyle\leqslant\sum_{k=0}^{K-1}\left|\chi_{S(h)}(k\alpha+\rho)-\lambda_{1}(S_{h})\right|
2K.\displaystyle\leqslant 2K.

This clearly leads to (3.19) and completes the proof. ∎

Let xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

A(Nh;Mh)={(α,β)[2Nh,2+1Nh]×[0,α):(3.18) holds}.A(N_{h};M_{h})=\left\{(\alpha,\beta)\in\left[\frac{2^{\ell}}{N_{h}},\frac{2^{\ell+1}}{N_{h}}\right]\times[0,\alpha):\mbox{\eqref{eq3.18} holds}\right\}.

Then the above argument leads to the inequality

λ2(B(Nh;Mh))Nh2Mhλ2(A(Nh;Mh)).\lambda_{2}(B(N_{h};M_{h}))\geqslant\frac{N_{h}}{2M_{h}}\lambda_{2}(A(N_{h};M_{h})). (3.20)

Combining (3.13) and (3.20), we obtain the upper bound

λ2(A(Nh;Mh))2+13hNh2log2(1+h).\lambda_{2}(A(N_{h};M_{h}))\leqslant\frac{2^{\ell+13}}{hN_{h}^{2}\log^{2}(1+h)}.

Combining this with (3.5), it is not difficult to see that apart from a set of measure λ2(A(Nh;Mh))\lambda_{2}(A(N_{h};M_{h})), every pair (α,β)(\alpha,\beta) such that (3.14) holds satisfies the inequality

|k=0K(α)1χS(h)(kα+β)K(α)λ1(S(h))|<2NhMh+(hNhMh)1/2log(1+h).\left|\sum_{k=0}^{K(\alpha)-1}\chi_{S(h)}(k\alpha+\beta)-K(\alpha)\lambda_{1}(S(h))\right|<\frac{2N_{h}}{M_{h}}+(hN_{h}M_{h})^{1/2}\log(1+h).

Next, note that the two expressions

k0kα+βS(h)11αλ1(S(h))andk=0K(α)1χS(h)(kα+β)K(α)λ1(S(h))\sum_{\begin{subarray}{c}{k\geqslant 0}\\ {k\alpha+\beta\in S(h)}\end{subarray}}1-\frac{1}{\alpha}\lambda_{1}(S(h))\quad\mbox{and}\quad\sum_{k=0}^{K(\alpha)-1}\chi_{S(h)}(k\alpha+\beta)-K(\alpha)\lambda_{1}(S(h))

differ by at most 22, due to the possibility that (K(α)1)α+β>1(K(\alpha)-1)\alpha+\beta>1 and the difference |K(α)1/α|<1|K(\alpha)-1/\alpha|<1, in view of (3.4). It follows that on reversing the contraction, we see that apart from a set of measure at most

2+13hlog2(1+h),\frac{2^{\ell+13}}{h\log^{2}(1+h)},

every pair (θ,η)(\theta,\eta) such that 2θ2+12^{\ell}\leqslant\theta\leqslant 2^{\ell+1} and η[0,θ)\eta\in[0,\theta) satisfies the inequality

|k0kθ+η𝒮[0,Nh]11θλ1(𝒮[0,Nh])|<2NhMh+(hNhMh)1/2log(1+h)+2.\left|\sum_{\begin{subarray}{c}{k\geqslant 0}\\ {k\theta+\eta\in\mathcal{S}\cap[0,N_{h}]}\end{subarray}}1-\frac{1}{\theta}\lambda_{1}(\mathcal{S}\cap[0,N_{h}])\right|<\frac{2N_{h}}{M_{h}}+(hN_{h}M_{h})^{1/2}\log(1+h)+2. (3.21)
Lemma 3.3 (Borel–Cantelli lemma).

Let (X,Σ,μ)(X,\Sigma,\mu) be a measure space, and suppose that EhE_{h}, h=1,2,3,,h=1,2,3,\ldots, is a sequence of Σ\Sigma-measurable sets. If

h=1μ(Eh)<,\sum_{h=1}^{\infty}\mu(E_{h})<\infty,

then xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

μ(h=1i=hEi)=0.\mu\left(\bigcap_{h=1}^{\infty}\bigcup_{i=h}^{\infty}E_{i}\right)=0.

Since xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

h=11hlog2(1+h)<,\sum_{h=1}^{\infty}\frac{1}{h\log^{2}(1+h)}<\infty,

we conclude that for almost every pair (θ,η)(\theta,\eta) such that 2θ2+12^{\ell}\leqslant\theta\leqslant 2^{\ell+1} and η[0,θ)\eta\in[0,\theta), the inequality (3.21) holds for all sufficiently large positive integers hh.

Finally, we specify the integers NhN_{h} and MhM_{h} in terms of the parameter h1h\geqslant 1 and the chosen integer \ell. Choosing them to satisfy

Nh2(h4log2(1+h))<Nh+1andMh=2hN_{h}\leqslant 2^{\ell}(h^{4}\log^{2}(1+h))<N_{h}+1\quad\mbox{and}\quad M_{h}=2^{\ell}h (3.22)

ensures that the two dominant terms on the right hand side of (3.21) have the same order of magnitude in terms of hh. For an arbitrary sufficiently integer nn, we choose hh to satisfy Nhn<Nh+1N_{h}\leqslant n<N_{h+1}. Then it follows from (3.21) and (3.22) that

|k0kθ+η𝒮[0,n]11θλ1(𝒮[0,n])|\displaystyle\left|\sum_{\begin{subarray}{c}{k\geqslant 0}\\ {k\theta+\eta\in\mathcal{S}\cap[0,n]}\end{subarray}}1-\frac{1}{\theta}\lambda_{1}(\mathcal{S}\cap[0,n])\right|
<Nh+1Nhθ+2NhMh+(hNhMh)1/2log(1+h)+2\displaystyle\qquad<\frac{N_{h+1}-N_{h}}{\theta}+\frac{2N_{h}}{M_{h}}+(hN_{h}M_{h})^{1/2}\log(1+h)+2
2((h+1)4log2(2+h)h4log2(1+h))θ+(2+2)h3log2(1+h)+3\displaystyle\qquad\leqslant\frac{2^{\ell}((h+1)^{4}\log^{2}(2+h)-h^{4}\log^{2}(1+h))}{\theta}+(2+2^{\ell})h^{3}\log^{2}(1+h)+3
2+3h3log2(1+h)+O(h3log(1+h))c1()n3/4(logn)1/2,\displaystyle\qquad\leqslant 2^{\ell+3}h^{3}\log^{2}(1+h)+O_{\ell}(h^{3}\log(1+h))\leqslant c_{1}(\ell)n^{3/4}(\log n)^{1/2},

provided that nn, and hence also hh, is sufficiently large.

This completes the proof of Lemma 2.1.

Proof of Lemma 3.1.

For any fixed δ(0,1/2)\delta\in(0,1/2), we define the roof function Rδ:R_{\delta}:\mathbb{R}\to\mathbb{R} by writing xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

Rδ(x)={0,if |x|>δ,1(|x|/δ),if 0|x|δ.R_{\delta}(x)=\left\{\begin{array}[]{ll}0,&\mbox{if $|x|>\delta$},\\ 1-(|x|/\delta),&\mbox{if $0\leqslant|x|\leqslant\delta$}.\end{array}\right.

For every integer jj\in\mathbb{Z}, we consider the integral

I(δ;j)=1/21/2Rδ(x)e2πijxdx.I(\delta;j)=\int_{-1/2}^{1/2}R_{\delta}(x)\mathrm{e}^{2\pi\mathrm{i}jx}\,\mathrm{d}x. (3.23)

Then xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

I(δ;0)=δandI(δ;j)=δ(sinπjδπjδ)2,j{0}.I(\delta;0)=\delta\quad\mbox{and}\quad I(\delta;j)=\delta\left(\frac{\sin\pi j\delta}{\pi j\delta}\right)^{2},\quad j\in\mathbb{Z}\setminus\{0\}. (3.24)

For any integers j1,j2{0}j_{1},j_{2}\in\mathbb{Z}\setminus\{0\} and positive integer NN, let

𝔅(j1;j2;N)=2/N2+1/N(aj1k1=0K(α)1e2πij1k1α)(aj2¯k2=0K(α)1e2πij2k2α)dα\displaystyle\mathfrak{B}(j_{1};j_{2};N)=\int_{2^{\ell}/N}^{2^{\ell+1}/N}\left(a_{j_{1}}\sum_{k_{1}=0}^{K(\alpha)-1}\mathrm{e}^{2\pi\mathrm{i}j_{1}k_{1}\alpha}\right)\left(\overline{a_{j_{2}}}\sum_{k_{2}=0}^{K(\alpha)-1}\mathrm{e}^{-2\pi\mathrm{i}j_{2}k_{2}\alpha}\right)\,\mathrm{d}\alpha
=2/N2+1/N(aj1e2πij1K(α)α1e2πij1α1)(aj2¯e2πij2K(α)α1e2πij2α1)dα,\displaystyle\qquad=\int_{2^{\ell}/N}^{2^{\ell+1}/N}\left(a_{j_{1}}\frac{\mathrm{e}^{2\pi\mathrm{i}j_{1}K(\alpha)\alpha}-1}{\mathrm{e}^{2\pi\mathrm{i}j_{1}\alpha}-1}\right)\left(\overline{a_{j_{2}}}\,\frac{\mathrm{e}^{-2\pi\mathrm{i}j_{2}K(\alpha)\alpha}-1}{\mathrm{e}^{-2\pi\mathrm{i}j_{2}\alpha}-1}\right)\mathrm{d}\alpha, (3.25)

so that

|𝔅(j1;j2;N)|\displaystyle|\mathfrak{B}(j_{1};j_{2};N)|
122/N2+1/N(|aj1e2πij1K(α)α1e2πij1α1|2+|aj2e2πij2K(α)α1e2πij2α1|2)dα.\displaystyle\quad\leqslant\frac{1}{2}\int_{2^{\ell}/N}^{2^{\ell+1}/N}\left(\left|a_{j_{1}}\frac{\mathrm{e}^{2\pi\mathrm{i}j_{1}K(\alpha)\alpha}-1}{\mathrm{e}^{2\pi\mathrm{i}j_{1}\alpha}-1}\right|^{2}+\left|a_{j_{2}}\frac{\mathrm{e}^{2\pi\mathrm{i}j_{2}K(\alpha)\alpha}-1}{\mathrm{e}^{2\pi\mathrm{i}j_{2}\alpha}-1}\right|^{2}\right)\mathrm{d}\alpha. (3.26)

Then it follows from (3), (3.23) with δ=2/M\delta=2^{\ell}/M and j=j1j2j=j_{1}-j_{2} and from (3) that xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

J(𝐚;N;M)=j1{0}j2{0}I(2M;j1j2)𝔅(j1;j2;N).J^{*}(\mathbf{a};N;M)=\sum_{j_{1}\in\mathbb{Z}\setminus\{0\}}\sum_{j_{2}\in\mathbb{Z}\setminus\{0\}}I\left(\frac{2^{\ell}}{M};j_{1}-j_{2}\right)\mathfrak{B}(j_{1};j_{2};N).

We write xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

J(𝐚;N;M)=J1(𝐚;N;M)+J2(𝐚;N;M),J^{*}(\mathbf{a};N;M)=J^{*}_{1}(\mathbf{a};N;M)+J^{*}_{2}(\mathbf{a};N;M), (3.27)

where J1(𝐚;N;M)J^{*}_{1}(\mathbf{a};N;M) contains all the diagonal terms in J(𝐚;N;M)J^{*}(\mathbf{a};N;M) with j1=j2j_{1}=j_{2}, while J2(𝐚;N;M)J^{*}_{2}(\mathbf{a};N;M) contains all the off-diagonal terms in J(𝐚;N;M)J^{*}(\mathbf{a};N;M) with j1j2j_{1}\neq j_{2}. Noting that I(2/M;0)=2/MI(2^{\ell}/M;0)=2^{\ell}/M, we see that

J1(𝐚;N;M)=2Mj{0}|aj|2E(j;N),J^{*}_{1}(\mathbf{a};N;M)=\frac{2^{\ell}}{M}\sum_{j\in\mathbb{Z}\setminus\{0\}}|a_{j}|^{2}E(j;N), (3.28)

where xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

E(j;N)=2/N2+1/N|e2πijK(α)α1e2πijα1|2dα.E(j;N)=\int_{2^{\ell}/N}^{2^{\ell+1}/N}\left|\frac{\mathrm{e}^{2\pi\mathrm{i}jK(\alpha)\alpha}-1}{\mathrm{e}^{2\pi\mathrm{i}j\alpha}-1}\right|^{2}\mathrm{d}\alpha. (3.29)

Meanwhile, noting (3.24), we see that

J2(𝐚;N;M)=2Mj1{0}j2{0}j1j2(sinπ(j1j2)2M1π(j1j2)2M1)2𝔅(j1;j2;N).J^{*}_{2}(\mathbf{a};N;M)=\frac{2^{\ell}}{M}\mathop{\sum_{j_{1}\in\mathbb{Z}\setminus\{0\}}\sum_{j_{2}\in\mathbb{Z}\setminus\{0\}}}_{j_{1}\neq j_{2}}\left(\frac{\sin\pi(j_{1}-j_{2})2^{\ell}M^{-1}}{\pi(j_{1}-j_{2})2^{\ell}M^{-1}}\right)^{2}\mathfrak{B}(j_{1};j_{2};N). (3.30)

Combining (3) and (3.30), we deduce that

|J2(𝐚;N;M)|2Mj{0}ζ{0}(sinπζ2M1πζ2M1)2|aj|2E(j;N).|J^{*}_{2}(\mathbf{a};N;M)|\leqslant\frac{2^{\ell}}{M}\sum_{j\in\mathbb{Z}\setminus\{0\}}\sum_{\zeta\in\mathbb{Z}\setminus\{0\}}\left(\frac{\sin\pi\zeta 2^{\ell}M^{-1}}{\pi\zeta 2^{\ell}M^{-1}}\right)^{2}|a_{j}|^{2}E(j;N). (3.31)

It then follows from (3.27), (3.28) and (3.31) that

|J(𝐚;N;M)|2M(1+ζ{0}(sinπζ2M1πζ2M1)2)j{0}|aj|2E(j;N).\!\!\!|J^{*}(\mathbf{a};N;M)|\leqslant\frac{2^{\ell}}{M}\left(1+\sum_{\zeta\in\mathbb{Z}\setminus\{0\}}\left(\frac{\sin\pi\zeta 2^{\ell}M^{-1}}{\pi\zeta 2^{\ell}M^{-1}}\right)^{2}\right)\sum_{j\in\mathbb{Z}\setminus\{0\}}|a_{j}|^{2}E(j;N). (3.32)

Next, note that

2M(1+ζ{0}(sinπζ2M1πζ2M1)2)2M(|ζ|M/21+|ζ|>M/2(Mπζ2)2)\displaystyle\frac{2^{\ell}}{M}\left(1+\sum_{\zeta\in\mathbb{Z}\setminus\{0\}}\left(\frac{\sin\pi\zeta 2^{\ell}M^{-1}}{\pi\zeta 2^{\ell}M^{-1}}\right)^{2}\right)\leqslant\frac{2^{\ell}}{M}\left(\sum_{|\zeta|\leqslant M/2^{\ell}}1+\sum_{|\zeta|>M/2^{\ell}}\left(\frac{M}{\pi\zeta 2^{\ell}}\right)^{2}\right)
2M(2M2+1+2M2π24M/2dxx2)6.\displaystyle\qquad\leqslant\frac{2^{\ell}}{M}\left(\frac{2M}{2^{\ell}}+1+\frac{2M^{2}}{\pi^{2}4^{\ell}}\int_{M/2^{\ell}}^{\infty}\frac{\mathrm{d}x}{x^{2}}\right)\leqslant 6. (3.33)

To complete the proof of Lemma 3.1, in view of (3.3), (3.32) and (3), it suffices to show that for every j{0}j\in\mathbb{Z}\setminus\{0\} and integer N>1N>1, we have

E(j;N)2+8.E(j;N)\leqslant 2^{\ell+8}. (3.34)

Consider first small values of jj, where 1|j|N/2+21\leqslant|j|\leqslant N/2^{\ell+2}. Suppose first that j>0j>0. Since α[2/N,2+1/N)\alpha\in[2^{\ell}/N,2^{\ell+1}/N), it follows that 0jα1/20\leqslant j\alpha\leqslant 1/2, so that e2πijα\mathrm{e}^{2\pi\mathrm{i}j\alpha} is a point on the upper half circle of unit radius. On the other hand, (3.4) implies the inequality 0jK(α)αj<jα0\leqslant jK(\alpha)\alpha-j<j\alpha. Hence e2πijK(α)α=e2πi(jK(α)αj)\mathrm{e}^{2\pi\mathrm{i}jK(\alpha)\alpha}=\mathrm{e}^{2\pi\mathrm{i}(jK(\alpha)\alpha-j)} is a point on the circular arc of the upper half circle of unit radius joining the points 11 and e2πijα\mathrm{e}^{2\pi\mathrm{i}j\alpha}. As shown in Figure 4.1, we clearly have |e2πijK(α)α1|<|e2πijα1||\mathrm{e}^{2\pi\mathrm{i}jK(\alpha)\alpha}-1|<|\mathrm{e}^{2\pi\mathrm{i}j\alpha}-1|.

[Uncaptioned image]Figure 2: justifying the inequality (3.35)\begin{array}[]{c}\includegraphics[scale={0.8}]{figure-4-1.pdf}\vspace{3pt}\\ \mbox{Figure 2: justifying the inequality \eqref{eq3.35}}\end{array}

Meanwhile, replacing jj by j-j preserves this inequality. It follows that for every integer jj satisfying 1|j|N/2+21\leqslant|j|\leqslant N/2^{\ell+2}, we have

|e2πijK(α)α1e2πijα1|<1and soE(j;N)2N.\left|\frac{\mathrm{e}^{2\pi\mathrm{i}jK(\alpha)\alpha}-1}{\mathrm{e}^{2\pi\mathrm{i}j\alpha}-1}\right|<1\quad\mbox{and so}\quad E(j;N)\leqslant\frac{2^{\ell}}{N}. (3.35)

Next, for any fixed integer jj satisfying |j|>N/2+2|j|>N/2^{\ell+2}, we use the inequalities

|e2πijK(α)α1e2πijα1|min{K(α),1jα}min{N,1jα},\left|\frac{\mathrm{e}^{2\pi\mathrm{i}jK(\alpha)\alpha}-1}{\mathrm{e}^{2\pi\mathrm{i}j\alpha}-1}\right|\leqslant\min\left\{K(\alpha),\frac{1}{\|j\alpha\|}\right\}\leqslant\min\left\{N,\frac{1}{\|j\alpha\|}\right\}, (3.36)

where x\|x\| denotes the distance of a real number xx from the nearest integer, and the inequality K(α)NK(\alpha)\leqslant N follows from α[2/N,2+1/N)\alpha\in[2^{\ell}/N,2^{\ell+1}/N) and (3.4).

Let rr be the integer closest to jαj\alpha. Then

jα<1Nif and only if|αrj|<1|j|N,\|j\alpha\|<\frac{1}{N}\quad\mbox{if and only if}\quad\left|\alpha-\frac{r}{j}\right|<\frac{1}{|j|N},

so that xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

jα<1Nif and only ifα(rj1|j|N,rj+1|j|N).\|j\alpha\|<\frac{1}{N}\quad\mbox{if and only if}\quad\alpha\in\left(\frac{r}{j}-\frac{1}{|j|N},\frac{r}{j}+\frac{1}{|j|N}\right). (3.37)

For this fixed integer jj satisfying |j|>N/2+2|j|>N/2^{\ell+2}, there are at most

max{5,2+2|j|N}2+5|j|N\max\left\{5,\frac{2^{\ell+2}|j|}{N}\right\}\leqslant\frac{2^{\ell+5}|j|}{N} (3.38)

integers rr such that

(rj1|j|N,rj+1|j|N)[2N,2+1N).\left(\frac{r}{j}-\frac{1}{|j|N},\frac{r}{j}+\frac{1}{|j|N}\right)\cap\left[\frac{2^{\ell}}{N},\frac{2^{\ell+1}}{N}\right)\neq\emptyset. (3.39)

On the other hand, suppose that nn is a positive fixed integer satisfying 2nN2^{n}\leqslant N. Then analogous to (3.37), we have

2n1Njα2nNif and only ifα(j;N;n),\frac{2^{n-1}}{N}\leqslant\|j\alpha\|\leqslant\frac{2^{n}}{N}\quad\mbox{if and only if}\quad\alpha\in\mathfrak{I}(j;N;n), (3.40)

where xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

(j;N;n)=[rj2n|j|N,rj2n1|j|N][rj+2n1|j|N,rj+2n|j|N],\mathfrak{I}(j;N;n)=\left[\frac{r}{j}-\frac{2^{n}}{|j|N},\frac{r}{j}-\frac{2^{n-1}}{|j|N}\right]\cup\left[\frac{r}{j}+\frac{2^{n-1}}{|j|N},\frac{r}{j}+\frac{2^{n}}{|j|N}\right], (3.41)

except when 2n/N>1/22^{n}/N>1/2, in which case we have the modification

(j;N;n)=[rj12|j|,rj2n1|j|N][rj+2n1|j|N,rj+12|j|].\mathfrak{I}(j;N;n)=\left[\frac{r}{j}-\frac{1}{2|j|},\frac{r}{j}-\frac{2^{n-1}}{|j|N}\right]\cup\left[\frac{r}{j}+\frac{2^{n-1}}{|j|N},\frac{r}{j}+\frac{1}{2|j|}\right]. (3.42)

Similarly, for this fixed integer jj satisfying |j|>N/2+2|j|>N/2^{\ell+2} and fixed positive integer nn satisfying 2nN2^{n}\leqslant N, there are at most (3.38) integers rr such that

(j;N;n)[2N,2+1N).\mathfrak{I}(j;N;n)\cap\left[\frac{2^{\ell}}{N},\frac{2^{\ell+1}}{N}\right)\neq\emptyset. (3.43)

Combining (3.29) and (3.36)–(3.43), we see that for any fixed integer jj satisfying |j|>N/2+2|j|>N/2^{\ell+2}, we have

E(j;N)2/N2+1/N(min{N,1jα})2dα\displaystyle E(j;N)\leqslant\int_{2^{\ell}/N}^{2^{\ell+1}/N}\left(\min\left\{N,\frac{1}{\|j\alpha\|}\right\}\right)^{2}\mathrm{d}\alpha
N22+5|j|N2|j|N+n=12nN(N2n1)22+5|j|N2n|j|N2+5(2+4n=112n),\displaystyle\qquad\leqslant N^{2}\,\frac{2^{\ell+5}|j|}{N}\,\frac{2}{|j|N}+\sum_{\begin{subarray}{c}{n=1}\\ {2^{n}\leqslant N}\end{subarray}}^{\infty}\left(\frac{N}{2^{n-1}}\right)^{2}\frac{2^{\ell+5}|j|}{N}\,\frac{2^{n}}{|j|N}\leqslant 2^{\ell+5}\left(2+4\sum_{n=1}^{\infty}\frac{1}{2^{n}}\right),

confirming the assertion (3.34) and completing the proof. ∎

References

  • [1] J. Beck, M. Donders, Y. Yang. Quantitative behavior of non-integrable systems I. Acta Math. Hungar. 161 (2020), 66–184.
  • [2] J. Beck, M. Donders, Y. Yang. Quantitative behavior of non-integrable systems II. Acta Math. Hungar. 162 (2020), 220–324.
  • [3] J. Beck, W.W.L. Chen, Y. Yang. Quantitative behavior of non-integrable systems IV. Acta Math. Hungar. 167 (2022), 1–160.
  • [4] S. Kerckhoff, H. Masur, J. Smillie. Ergodicity of billiard flows and quadratic differentials. Ann. of Math. 124 (1986), 293–311.
  • [5] D. König, A. Szücs. Mouvement d’un point abondonne a l’interieur d’un cube. Rend. Circ. Mat. Palermo 36 (1913), 79–90.