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

\settrims

0pt0pt \setheadfoot0.5in0.5in \setulmarginsandblock1.0in1.0in* \setlrmarginsandblock1.0in1.0in* \checkandfixthelayout \setsecheadstyle \setsecnumdepthsubsubsection \setsecnumformat. \setsubsecheadstyle \setaftersubsecskip-0.5em \setsubsubsecheadstyle \setaftersubsubsecskip-0.5em \OnehalfSpacing

Uniform distribution of saddle connection lengths

Jon Chaika    Donald Robertson with an appendix by Daniel El-Baz    Bingrong Huang
Abstract

For any 𝖲𝖫(2,)\mathsf{SL}(2,\mathbb{R}) invariant and ergodic probability measure on any stratum of flat surfaces, almost every flat surface has the property that its non-decreasing sequence of saddle connection lengths is uniformly distributed mod one.

1 Introduction

By a flat surface we mean a pair (X,ω)(X,\omega) where XX is a closed, compact Riemann surface of genus g2g\geq 2 and ω\omega is a non-zero holomorphic one form on XX. Every non-zero holomorphic one form on XX has 2g22g-2 zeros where gg is the genus of XX. Fixing the genus of XX the set {(X,ω):ω0}\{(X,\omega):\omega\neq 0\} is partitioned into strata according to the possible partitions of 2g22g-2. Each connected component of a stratum carries a natural volume measure that is invariant under the action of 𝖲𝖫(2,)\mathsf{SL}(2,\mathbb{R}) on pairs (X,ω)(X,\omega). The 𝖲𝖫(2,)\mathsf{SL}(2,\mathbb{R}) action preserves the area of (X,ω)(X,\omega). Masur [Mas82] and Veech [Vee82] showed independently that the natural volume on a connected component induces on its subset \mathcal{H} of unit area flat surfaces a probability measure 𝖬\mathsf{M} that is ergodic for the 𝖲𝖫(2,)\mathsf{SL}(2,\mathbb{R}) actions.

Fix a flat surface (X,ω)(X,\omega). Every non-zero holomorphic one form gives, away from its finite set Σ\Sigma of zeroes, an atlas of charts to 2\mathbb{R}^{2} whose transition maps are translations in 2\mathbb{R}^{2}. For every 0θ<π0\leq\theta<\pi we can use such an atlas to induce a foliation on XΣX\setminus\Sigma given locally by the lines in 2\mathbb{R}^{2} making an angle of θ\theta with the horiontal axis. We can also use such an atlas to measure lengths of curves on XX. In particular we can calculate the lengths of the leaves in the above foliations. A saddle connection of (X,ω)(X,\omega) is any leaf of any of these foliations that starts and ends at a point of Σ\Sigma. The holonomy of a saddle connection vv is the vector

𝗁(v)=(vRe(ω),vIm(ω))\mathsf{h}(v)=\left(\int\limits_{v}\mathrm{Re}(\omega),\int\limits_{v}\mathrm{Im}(\omega)\right)

in 2\mathbb{R}^{2}.

In this paper we are interested in the uniform distribution of saddle connection lengths. Recall that a sequence nxnn\mapsto x_{n} in \mathbb{R} is uniformly distributed mod 1 when

limN|{1nN:axnmod1<b}|N=ba\lim_{N\to\infty}\frac{|\{1\leq n\leq N:a\leq x_{n}\bmod 1<b\}|}{N}=b-a

for all intervals [a,b)[0,1)[a,b)\subset[0,1). The well-known Weyl criterion is that uniform distribution of nxnn\mapsto x_{n} is equivalent to

limN1Nn=1Nexp(2πipxn)=0\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\exp(2\pi ipx_{n})=0

for all p{0}p\in\mathbb{Z}\setminus\{0\} and one can use it to prove, for example, that nnαn\mapsto n\alpha is uniformly distributed for all irrational α\alpha.

Returning to flat surfaces (and suppressing XX from our notation) we write Λ(ω)\Lambda(\omega) for the set of saddle connections of ω\omega. For each R>0R>0 write Λ(ω;R)\Lambda(\omega;R) for the set of saddle connections whose holonomy vectors have length at most RR. Given also S>0S>0 write Λ(ω;R,S)\Lambda(\omega;R,S) for the set of saddle connections whose holonomy vectors lie within the ellipse (Ry)2+(Sx)2=(RS)2(Ry)^{2}+(Sx)^{2}=(RS)^{2}.

It follows from minor modifications to work of Vorobets [Vor05, Theorem 1.9] that the projection of Λ(ω;R)\Lambda(\omega;R) to the unit circle along rays through the origin is uniformly distributed as RR\to\infty for almost every ω\omega. Specifically, by [Vor05, Proposition 4.5] it suffices to show that the set of saddle connection holonomies satisfy axioms (0), (A), (B), (C) and (E) therein. That saddle connections doe satisfy these axioms follows from the definitions and results in [EM01] just as it does for the sets Vorobets considers.

Our main result is that lengths of saddle connection holonomy vectors are also almost always uniformly distributed mod 1. To make sense of this we fix, for each flat surface ω\omega, an enumeration nvnn\mapsto v_{n} of Λ(ω)\Lambda(\omega) such that n||𝗁(vn)||n\mapsto|\!|_{\mathsf{}}\mathsf{h}(v_{n})|\!|_{\mathsf{}} is non-decreasing.

Theorem 1.

For any 𝖲𝖫(2,)\mathsf{SL}(2,\mathbb{R}) invariant and ergodic probability measure 𝗆\mathsf{m} on \mathcal{H} and 𝗆\mathsf{m} almost every flat surface ω\omega the sequence n||𝗁(vn)||n\mapsto|\!|_{\mathsf{}}\mathsf{h}(v_{n})|\!|_{\mathsf{}} is uniformly distributed.

Masur [Mas88, Mas90] proved that for every surface ω\omega there exists c1c_{1}, c2c_{2} (depending on ω\omega) so that for all large RR we have

c1R2|Λ(ω;R)|<c2R2c_{1}R^{2}\leq|\Lambda(\omega;R)|<c_{2}R^{2} (2)

(see also Vorobets [Vor03]). Veech [Vee98] introduced what is now called the Siegel-Veech transform to show that there exists cc depending only on ν\nu and RnR_{n}\to\infty so that

limiΛ(ω;Ri)Ri2=c\lim_{i\to\infty}\frac{\Lambda(\omega;R_{i})}{R_{i}^{2}}=c

for 𝖬\mathsf{M} almost every ω\omega. He also showed the limit exists for all surfaces in certain suborbifolds of \mathcal{H}. Building on Veech’s approach, Eskin-Masur [EM01] proved that for every 𝖲𝖫(2,)\mathsf{SL}(2,\mathbb{R}) invariant and 𝖲𝖫(2,)\mathsf{SL}(2,\mathbb{R}) ergodic probability measure 𝗆\mathsf{m} there exists cc so that we have

limRΛ(ω;R)R2=c\underset{R\to\infty}{\lim}\,\frac{\Lambda(\omega;R)}{R^{2}}=c (3)

for 𝗆\mathsf{m} almost every ω\omega. Dozier [Doz17a] proved analogues of these results for saddle connections with holonomies in certain sectors of directions.

The best known results for asymptotic counting that hold on all surfaces are Eskin, Mirzakhani and Mohammadi’s result [EMM15, Theorem 2.12] that every surface has an asymptotic growth of saddle connections on average, and Dozier’s result [Doz17] building on techniques in [EM11] that the constants c1,c2c_{1},c_{2} in (2) can be chosen to depend only on the connected component of the stratum in question. The best almost everywhere counting result is the following theorem of Nevo, Rühr and Weiss.

Theorem 4 ([NRW17, Theorem 1.1]).

For every 𝖲𝖫(2,)\mathsf{SL}(2,\mathbb{R}) invariant and ergodic probability measure 𝗆\mathsf{m} on \mathcal{H} there are constants c,κ>0c,\kappa>0 such that 𝗆\mathsf{m} almost every flat surface ω\omega satisfies

|Λ(ω;R)|=cR2+𝖮ω(R2(1κ))|\Lambda(\omega;R)|=cR^{2}+\mathsf{O}_{\omega}(R^{2(1-\kappa)}) (5)

for all R>0R>0.

The gaps of saddle connection directions has also been studied by Athreya, the first named author, Lelievre, Uyanik and Work [AC12], [ACL15] and [UW16]. Recently Athreya, Cheung and Masur [ACM17] have begun investigating the L2\mathrm{L}^{2} properties of the Siegel-Veech transform.

Most of these papers to a greater or lesser extent adhere to the philosophy of renormalization dynamics, using the 𝖲𝖫(2,)\mathsf{SL}(2,\mathbb{R}) action and especially the geodesic flow

𝗀t=[et00et]\mathsf{g}^{t}=\begin{bmatrix}e^{t}&0\\ 0&e^{-t}\end{bmatrix}

to translate questions about long saddle connections to moderate saddle connections on different surfaces. Our approach is a little different: we still use the 𝖲𝖫(2,)\mathsf{SL}(2,\mathbb{R}) action, but only as a means to perturb. Namely, we do not apply the geodesic flow 𝗀t\mathsf{g}^{t} long enough to make a saddle connection 𝖮(1)\mathsf{O}(1) in length. Instead, writing also

𝗋θ=[cosθsinθsinθcosθ]\mathsf{r}^{\theta}=\begin{bmatrix}\cos\theta&-\sin\theta\\ \sin\theta&\cos\theta\end{bmatrix}

we prove that for almost every ω\omega and for almost every θ,t\theta,t the lengths of saddle connections on 𝗀t𝗋θω\mathsf{g}^{t}\mathsf{r}^{\theta}\omega are uniformly distributed mod 1. However, we emphasize that we do use Theorem 4, which does use renormalization dynamics.

We remark that even if the conclusion of [NRW17, Theorem 1.1] were known for every flat surface, our methods would not give uniform distribution of saddle connection lengths for every flat surface. Therefore, the following question is in general open.

Question 6.

Is the sequence n||𝗁(vn)||n\mapsto|\!|_{\mathsf{}}\mathsf{h}(v_{n})|\!|_{\mathsf{}} uniformly distributed for every flat surface?

However, in the presence a very strong error term for the saddle connection counting function |Λ(ω;R)||\Lambda(\omega;R)| one can obtain uniform distribution of saddle connection lengths everywhere with a much simpler proof. The appendix by Daniel El-Baz and Bingrong Huang proves using Huxley and Nowak’s bound [HN96] on the counting function for primitive lattice points that the conclusion of Theorem 1 holds for every torus.

The research of J. Chaika was supported in part by NSF grants DMS-135500 and DMS-1452762, the Sloan foundation, a Poincarè chair, and a Warnock chair. J. Chaika thanks Alex Eskin for encouraging him to pursue this question. The research of D. Robertson was supported by NSF grant DMS-1703597.

2 Proof of Theorem 1

In this section we prove Theorem 1. Let \mathcal{H} be the set of unit-area flat surfaces in the connected component of a stratum and write 𝗆\mathsf{m} for the Masur-Veech probability measure 𝗆\mathsf{m} on \mathcal{H}. Let c>0c>0 and κ>0\kappa>0 be as in Theorem 4. Our proof of Theorem 1 has the following four ingredients.

  1. 1.

    The Weyl criterion.

  2. 2.

    An exhaustion by compact sets.

  3. 3.

    A relation with orbit averages.

  4. 4.

    A linear approximation to handle exponential sums.

These steps are covered in the following four subsections. In Subsection 2.5 we combine the steps to prove Theorem 1.

2.1 The Weyl criterion

By the Weyl criterion for uniform distrubtion the following result implies Theorem 1. Recall that n𝗁(vn)n\mapsto\mathsf{h}(v_{n}) is an enumeration of Λ(ω)\Lambda(\omega) such that n||𝗁(vn)||n\mapsto|\!|_{\mathsf{}}\mathsf{h}(v_{n})|\!|_{\mathsf{}} is non-decreasing.

Theorem 7.

For 𝗆\mathsf{m} almost every ω\omega we have

limN1Nn=1Nexp(2πip||𝗁(vn)||)=0\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\exp(2\pi ip|\!|_{\mathsf{}}\mathsf{h}(v_{n})|\!|_{\mathsf{}})=0 (8)

for every pp\in\mathbb{N}.

Fix pp\in\mathbb{N} and write χp(x)=exp(2πipx)\chi_{p}(x)=\exp(2\pi ipx) for any xx\in\mathbb{R}. To prove that (8) holds almost surely we use the following theorem.

Lemma 9.

Fix a flat surface ω\omega. If there is a sequence τ1>τ2>τ3>1\tau_{1}>\tau_{2}>\tau_{3}>\cdots\to 1 such that

limJ1τiJn=1τiJχp(||𝗁(vn)||)=0\lim_{J\to\infty}\frac{1}{\lceil{\tau_{i}^{J}}\rceil}\sum_{n=1}^{\lceil{\tau_{i}^{J}}\rceil}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v_{n})|\!|_{\mathsf{}})=0 (10)

for all ii\in\mathbb{N} then (8) holds.

Proof.

This is a fact about Cesàro convergence. Fix a sequence τ1>τ2>τ3>1\tau_{1}>\tau_{2}>\tau_{3}>\cdots\to 1. Suppose that a bounded sequence nbnn\mapsto b_{n} has the property that

limJ1τiJn=1τiJbn=β\lim_{J\to\infty}\frac{1}{\lceil{\tau_{i}^{J}}\rceil}\sum_{n=1}^{\lceil{\tau_{i}^{J}}\rceil}b_{n}=\beta

for all ii\in\mathbb{N}. We verify that nbnn\mapsto b_{n} Cesàro converges to β\beta. Indeed, fix ϵ>0\epsilon>0 and ii\in\mathbb{N} such that

1<τi<1+ϵ4||b||1<\tau_{i}<1+\frac{\epsilon}{4|\!|_{\mathsf{}}b|\!|_{\mathsf{\infty}}} (11)

and choose JJ\in\mathbb{N} so large that

max{|1τijn=1τijbnβ|,2||b||τij}<ϵ3\max\left\{\left|\frac{1}{\lceil{\tau_{i}^{j}}\rceil}\sum_{n=1}^{\lceil{\tau_{i}^{j}}\rceil}b_{n}-\beta\right|,\frac{2|\!|_{\mathsf{}}b|\!|_{\mathsf{\infty}}}{\lceil{\tau_{i}^{j}}\rceil}\right\}<\frac{\epsilon}{3}

holds for all jJj\geq J. Put N=τiJN=\lceil{\tau_{i}^{J}}\rceil. Given M>NM>N write M=τij+kM=\lceil{\tau_{i}^{j}}\rceil+k with kk\in\mathbb{N} and jj maximal. Now

τij+kτij+k<τij+1+1<(1+ϵ4||b||)τij+1\tau_{i}^{j}+k\leq\lceil{\tau_{i}^{j}}\rceil+k<\tau_{i}^{j+1}+1<\left(1+\frac{\epsilon}{4|\!|_{\mathsf{}}b|\!|_{\mathsf{\infty}}}\right)\tau_{i}^{j}+1

giving 4k||b||ϵτij+4||b||4k|\!|_{\mathsf{}}b|\!|_{\mathsf{\infty}}\leq\epsilon\tau_{i}^{j}+4|\!|_{\mathsf{}}b|\!|_{\mathsf{\infty}}. So

|1τijn=1τijbn1τij+kn=1τij+kbn|2k||b||τij+kϵ2τijτij+k+2||b||τij+kϵ\left|\frac{1}{\lceil{\tau_{i}^{j}}\rceil}\sum_{n=1}^{\lceil{\tau_{i}^{j}}\rceil}b_{n}-\frac{1}{\lceil{\tau_{i}^{j}}\rceil+k}\sum_{n=1}^{\lceil{\tau_{i}^{j}}\rceil+k}b_{n}\right|\leq\frac{2k|\!|_{\mathsf{}}b|\!|_{\mathsf{\infty}}}{\lceil{\tau_{i}^{j}}\rceil+k}\leq\frac{\epsilon}{2}\frac{\tau_{i}^{j}}{\lceil{\tau_{i}^{j}}\rceil+k}+\frac{2|\!|_{\mathsf{}}b|\!|_{\mathsf{\infty}}}{\lceil{\tau_{i}^{j}}\rceil+k}\leq\epsilon

as desired. ∎

We conclude this subsection by replacing (10) with an average over all saddle connections whose holonomy vectors have length at most τJ/c\sqrt{\lceil{\tau^{J}}\rceil/c}.

Lemma 12.

If

limJcτJvΛ(ω;τJ/c)χp(||𝗁(v)||)=0\lim_{J\to\infty}\frac{c}{\lceil{\tau^{J}}\rceil}\sum_{v\in\Lambda\left(\omega;\sqrt{\lceil{\tau^{J}}\rceil/c}\right)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}})=0 (13)

for 𝗆\mathsf{m} almost every ω\omega then

limJ1τJn=1τJχp(||𝗁(vn)||)=0\lim_{J\to\infty}\frac{1}{\lceil{\tau^{J}}\rceil}\sum_{n=1}^{\lceil{\tau^{J}}\rceil}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v_{n})|\!|_{\mathsf{}})=0

for 𝗆\mathsf{m} almost every ω\omega.

Proof.

For almost every ω\omega there is from Theorem 4 a constant Cω>0C_{\omega}>0 with

|n=1τJχp(𝗁||vn||)vΛ(ω;τJ/c)χp(𝗁(||v||))|Cωc1κτJ1κ\left|\sum_{n=1}^{\lceil{\tau^{J}}\rceil}\chi_{p}(\mathsf{h}|\!|_{\mathsf{}}v_{n}|\!|_{\mathsf{}})-\sum_{v\in\Lambda\left(\omega;\sqrt{\lceil{\tau^{J}}\rceil/c}\right)}\chi_{p}(\mathsf{h}(|\!|_{\mathsf{}}v|\!|_{\mathsf{}}))\right|\leq\frac{C_{\omega}}{c^{1-\kappa}}\lceil{\tau^{J}}\rceil^{1-\kappa}

for all JJ\in\mathbb{N}. Dividing by τJ\lceil{\tau^{J}}\rceil gives the desired result. ∎

2.2 An exhaustion by compact sets

To relieve notation slightly we now consider the expression

1R2vΛ(ω;R)χp(||𝗁(v)||)\frac{1}{R^{2}}\sum_{v\in\Lambda(\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}) (14)

for R>0R>0. Our goal is to estimate the size of this average. We do so on compact sets that exhaust a full-measure subset of \mathcal{H}. To define our compact sets write (ω)\ell(\omega) for the length of the shortest saddle connection of ω\omega and

ξ(ω)=sup{||Λ(ω;R)|cR2|R2(1κ):R>0}\xi(\omega)=\sup\left\{\frac{\Big{|}|\Lambda(\omega;R)|-cR^{2}\Big{|}}{R^{2(1-\kappa)}}:R>0\right\}

for the 𝗆\mathsf{m} almost-surely defined optimal constant in (5). Fix r>0r>0 small and Ξ>0\Xi>0 large. Let KK be a compact subset of 1[r,)\ell^{-1}[r,\infty) on which both

𝗆(1[r,)K)<1r\mathsf{m}(\ell^{-1}[r,\infty)\setminus K)<\frac{1}{r}

and

sup{ξ(ω):ωK}Ξ\sup\{\xi(\omega):\omega\in K\}\leq\Xi (15)

hold. We can find such a compact set because ξ\xi is measurable. As r0r\to 0 and Ξ\Xi\to\infty our compact set KK exhausts \mathcal{H} up to a set of 𝗆\mathsf{m} measure zero.

To describe the behavior of (14) on KK we introduce some constants. Fix 0<η<10<\eta<1 and choose ρ\rho such that

1>ρ>max{1+η2,12κ}1>\rho>\max\left\{\frac{1+\eta}{2},1-2\kappa\right\}

with κ\kappa coming from Theorem 4. There is δ>0\delta>0 such that

|exp(2πis)12πis1|<ϵ\left|\frac{\exp(2\pi is)-1}{2\pi is}-1\right|<\epsilon (16)

whenever |s|<δ|s|<\delta. We only consider below values of RR large enough that

R1+ηmax{5ϵp242δ2,ϵ42}R^{1+\eta}\geq\max\left\{\frac{5\epsilon p^{2}}{42\delta^{2}},\frac{\epsilon}{42}\right\} (17)

and fix the relationship

S=ϵ42R1+ηS=\sqrt{\frac{\epsilon}{42R^{1+\eta}}} (18)

between SS and RR. Note that (17) implies S1S\leq 1. Although SS and RR are related we think of SS as the small amount of time we will geodesic flow for, and RR as a large bound on the lengths of our saddle connections.

2.3 A relation with orbit averages

The main result of this section is a relation between (14) on KK and its 𝖲𝖫(2,)\mathsf{SL}(2,\mathbb{R}) average over a small disc.

Theorem 19.

We have

𝗆({ωK:|1R2vΛ(ω;R)χp(||𝗁(v)||)|21Rσ})2\displaystyle\mathsf{m}\left(\left\{\omega\in K:\left|\frac{1}{R^{2}}\sum_{v\in\Lambda(\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}})\right|^{2}\geq\frac{1}{R^{\sigma}}\right\}\right)^{2} (20)
=\displaystyle={} 𝖮(RσK12π02π1S0S|1R2vΛ(𝗀t𝗋θω;R)χp(||𝗁(v)||)|2𝖽t𝖽θ𝖽𝗆(ω))\displaystyle\mathsf{O}\Bigg{(}R^{\sigma}\int\limits_{K}\frac{1}{2\pi}\int\limits_{0}^{2\pi}\frac{1}{S}\int\limits_{0}^{S}\left|\frac{1}{R^{2}}\sum_{v\in\Lambda(\mathsf{g}^{t}\mathsf{r}^{\theta}\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}})\right|^{2}\,\mathsf{d}t\,\mathsf{d}\theta\,\mathsf{d}\mathsf{m}(\omega)\Bigg{)}

for all RR satisfying (17) and all σ>0\sigma>0.

Proof.

Let

N={ωK:|1R2vΛ(ω;R)χp(||𝗁(v)||)|21Rσ}N=\left\{\omega\in K:\left|\frac{1}{R^{2}}\sum_{v\in\Lambda(\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}})\right|^{2}\geq\frac{1}{R^{\sigma}}\right\}

be the set whose measure we wish to bound. Write μ\mu for Haar measure on 𝖲𝖫(2,)\mathsf{SL}(2,\mathbb{R}), which is left and right invariant because 𝖲𝖫(2,)\mathsf{SL}(2,\mathbb{R}) is unimodular. For each s>0s>0 write

D(s)={[abcd]𝖲𝖫(2,):𝖽(i,ai+bci+d)<s}D(s)=\{[\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}]\in\mathsf{SL}(2,\mathbb{R}):\mathsf{d}(i,\tfrac{ai+b}{ci+d})<s\}

where 𝖽\mathsf{d} is the hyperbolic distance on the upper half-plane determined by the metric of constant curvature 4-4. In this metric the area of a hyperbolic disc of radius rr is π(sinh(r2))2\pi(\sinh(\frac{r}{2}))^{2} so whenever s<1s<1 we have

1μ(D(2s))μ(D(s))=(2cosh(s2))291\leq\frac{\mu(D(2s))}{\mu(D(s))}=(2\cosh(\tfrac{s}{2}))^{2}\leq 9 (21)

using the hyperbolic double angle formula.

The defining property of NN gives

1RσN1μ(D(S))D(S)1N(gω)𝖽μ(g)𝖽𝗆(ω)\displaystyle\frac{1}{R^{\sigma}}\int\limits_{N}\frac{1}{\mu(D(S))}\int\limits_{D(S)}1_{N}(g\omega)\,\mathsf{d}\mu(g)\,\mathsf{d}\mathsf{m}(\omega)
\displaystyle\leq K1μ(D(S))D(S)|1R2vΛ(gω;R)χp(||𝗁(v)||)|2𝖽μ(g)𝖽𝗆(ω)\displaystyle\int\limits_{K}\frac{1}{\mu(D(S))}\int\limits_{D(S)}\left|\frac{1}{R^{2}}\sum_{v\in\Lambda(g\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}})\right|^{2}\,\mathsf{d}\mu(g)\,\mathsf{d}\mathsf{m}(\omega)
=\displaystyle= K1(sinh(S2))202π02π0Ssinh(t)|1R2vΛ(𝗋θ1𝗀t𝗋θ2ω;R)χp(||𝗁(v)||)|2𝖽t𝖽θ1𝖽θ2𝖽𝗆(ω)\displaystyle\int\limits_{K}\frac{1}{(\sinh(\frac{S}{2}))^{2}}\int\limits_{0}^{2\pi}\int\limits_{0}^{2\pi}\int\limits_{0}^{S}\sinh(t)\left|\frac{1}{R^{2}}\sum_{v\in\Lambda(\mathsf{r}^{\theta_{1}}\mathsf{g}^{t}\mathsf{r}^{\theta_{2}}\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}})\right|^{2}\,\mathsf{d}t\,\mathsf{d}\theta_{1}\,\mathsf{d}\theta_{2}\,\mathsf{d}\mathsf{m}(\omega)

after using the Cartan integral formula [Kna01, Proposition 5.28] with μ\mu normalized appropriately. We can eliminate the integral over θ1\theta_{1} because the rotation 𝗋θ1\mathsf{r}^{\theta_{1}} doesn’t change the sum. Together with some simple estimates we arrive at

1RσN1μ(D(S))D(S)1N(gω)𝖽μ(g)𝖽𝗆(ω)CK12π02π1S0S|1R2vΛ(𝗀t𝗋θω;R)χp(||𝗁(v)||)|2𝖽t𝖽θ𝖽𝗆(ω)\frac{1}{R^{\sigma}}\int\limits_{N}\frac{1}{\mu(D(S))}\int\limits_{D(S)}1_{N}(g\omega)\,\mathsf{d}\mu(g)\,\mathsf{d}\mathsf{m}(\omega)\leq C\int\limits_{K}\frac{1}{2\pi}\int\limits_{0}^{2\pi}\frac{1}{S}\int\limits_{0}^{S}\left|\frac{1}{R^{2}}\sum_{v\in\Lambda(\mathsf{g}^{t}\mathsf{r}^{\theta}\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}})\right|^{2}\,\mathsf{d}t\,\mathsf{d}\theta\,\mathsf{d}\mathsf{m}(\omega)

where CC is some absolute positive constant because sinh(S)S\sinh(S)\sim S. It therefore suffices by (21) to prove

N1μ(D(S/4))D(S)1N(gω)𝖽μ(g)𝖽𝗆(ω)𝗆(N)24\int\limits_{N}\frac{1}{\mu(D(S/4))}\int\limits_{D(S)}1_{N}(g\omega)\,\mathsf{d}\mu(g)\,\mathsf{d}\mathsf{m}(\omega)\geq\frac{\mathsf{m}(N)^{2}}{4} (22)

holds.

Given EE\subset\mathcal{H} and ω\omega\in\mathcal{H} write Eω1={g𝖲𝖫(2,):gωE}E\omega^{-1}=\{g\in\mathsf{SL}(2,\mathbb{R}):g\omega\in E\}. Put

h(ω)=μ(Nω1D(S/4))μ(D(S/4))h(\omega)=\frac{\mu(N\omega^{-1}\cap D(S/4))}{\mu(D(S/4))}

and note that

𝗆(N)\displaystyle\mathsf{m}(N) =1μ(D(S/4))D(S/4)1g1N(ω)𝖽𝗆(ω)𝖽μ(g)\displaystyle=\frac{1}{\mu(D(S/4))}\int\limits_{D(S/4)}\int\limits_{\mathcal{H}}1_{g^{-1}N}(\omega)\,\mathsf{d}\mathsf{m}(\omega)\,\mathsf{d}\mu(g) (23)
=1μ(D(S/4))D(S/4)1N(gω)𝖽μ(g)𝖽𝗆(ω)=h𝖽𝗆\displaystyle=\int\limits_{\mathcal{H}}\frac{1}{\mu(D(S/4))}\int\limits_{D(S/4)}1_{N}(g\omega)\,\mathsf{d}\mu(g)\,\mathsf{d}\mathsf{m}(\omega)=\int\limits_{\mathcal{H}}h\,\mathsf{d}\mathsf{m}

by invariance of 𝗆\mathsf{m} and Fubini. Write

Uq=h(ω)qD(S/2)ωU_{q}=\bigcup_{h(\omega)\geq q}D(S/2)\cdot\omega

for any 0<q<10<q<1. As in (23) we have

𝗆(NUq)=1μ(D(S/4))D(S/4)1NUq(gω)𝖽μ(g)𝖽𝗆(ω)\mathsf{m}(N\cap U_{q})=\int\limits_{\mathcal{H}}\frac{1}{\mu(D(S/4))}\int\limits_{D(S/4)}1_{N\cap U_{q}}(g\omega)\,\mathsf{d}\mu(g)\,\mathsf{d}\mathsf{m}(\omega)

for every 0<q<10<q<1 and it follows for every 0<q<10<q<1 that

𝗆(NUq)h1[q,1]μ(Nω1Uqω1D(S/4))μ(D(S/4))𝖽𝗆(ω)=h1[q,1]h(ω)𝖽𝗆(ω)q𝗆(h1[q,1])\mathsf{m}(N\cap U_{q})\geq\int\limits_{h^{-1}[q,1]}\frac{\mu(N\omega^{-1}\cap U_{q}\omega^{-1}\cap D(S/4))}{\mu(D(S/4))}\,\mathsf{d}\mathsf{m}(\omega)=\int\limits_{h^{-1}[q,1]}h(\omega)\,\mathsf{d}\mathsf{m}(\omega)\geq q\mathsf{m}(h^{-1}[q,1]) (24)

because Uqω1D(S/2)U_{q}\omega^{-1}\supset D(S/2) whenever h(ω)qh(\omega)\geq q.

Claim.

𝗆(NUq)𝗆({ωN:μ(Nω1D(S))μ(D(S/4))q})\displaystyle\mathsf{m}(N\cap U_{q})\leq\mathsf{m}\left(\left\{\omega\in N:\frac{\mu(N\omega^{-1}\cap D(S))}{\mu(D(S/4))}\geq q\right\}\right)

Proof.

If ωNUq\omega\in N\cap U_{q} then there is η\eta with h(η)qh(\eta)\geq q and ωD(S/2)η\omega\in D(S/2)\cdot\eta. Fix aD(S/2)a\in D(S/2) with aη=ωa\cdot\eta=\omega. We get

qμ((Nη1)a1D(S/4)a1)μ(D(S/4))=μ(Nω1D(S/4)a1)μ(D(S/4))μ(Nω1D(S))μ(D(S/4))q\leq\frac{\mu((N\eta^{-1})a^{-1}\cap D(S/4)a^{-1})}{\mu(D(S/4))}=\frac{\mu(N\omega^{-1}\cap D(S/4)a^{-1})}{\mu(D(S/4))}\leq\frac{\mu(N\omega^{-1}\cap D(S))}{\mu(D(S/4))}

since μ\mu is invariant on both sides and D(S/4)a1D(S)D(S/4)a^{-1}\subset D(S)

The claim, combined with (24), gives

N1μ(D(S/4))D(S)1N(gω)𝖽μ(g)𝖽𝗆(ω)\displaystyle\int\limits_{N}\frac{1}{\mu(D(S/4))}\int\limits_{D(S)}1_{N}(g\omega)\,\mathsf{d}\mu(g)\,\mathsf{d}\mathsf{m}(\omega)
=\displaystyle={} 0𝗆({ωN:μ(Nω1D(S))μ(D(S/4))q})𝖽q\displaystyle\int\limits_{0}^{\infty}\mathsf{m}\left(\left\{\omega\in N:\frac{\mu(N\omega^{-1}\cap D(S))}{\mu(D(S/4))}\geq q\right\}\right)\,\mathsf{d}q
\displaystyle\geq{} 𝗆(N)21q𝗆(h1[q,1])𝖽q\displaystyle\int\limits_{\frac{\mathsf{m}(N)}{2}}^{1}q\mathsf{m}(h^{-1}[q,1])\,\mathsf{d}q
\displaystyle\geq{} 𝗆(N)2𝗆(N)21𝗆(h1[q,1])𝖽q𝗆(N)24\displaystyle\frac{\mathsf{m}(N)}{2}\int\limits_{\frac{\mathsf{m}(N)}{2}}^{1}\mathsf{m}(h^{-1}[q,1])\,\mathsf{d}q\geq\frac{\mathsf{m}(N)^{2}}{4}

where the last inequality follows from

h𝖽𝗆=0𝗆(N)2𝗆(h1[q,1])𝖽q+𝗆(N)21𝗆(h1[q,1])𝖽q𝗆(N)2+𝗆(N)21𝗆(h1[q,1])𝖽q\int\limits_{\mathcal{H}}h\,\mathsf{d}\mathsf{m}=\int\limits_{0}^{\frac{\mathsf{m}(N)}{2}}\mathsf{m}(h^{-1}[q,1])\,\mathsf{d}q+\int\limits_{\frac{\mathsf{m}(N)}{2}}^{1}\mathsf{m}(h^{-1}[q,1])\,\mathsf{d}q\leq\frac{\mathsf{m}(N)}{2}+\int\limits_{\frac{\mathsf{m}(N)}{2}}^{1}\mathsf{m}(h^{-1}[q,1])\,\mathsf{d}q

and (23). This establishes (22) as desired. ∎

The following lemma, another application of Theorem 4, will allow us to move the action of 𝖲𝖫(2,)\mathsf{SL}(2,\mathbb{R}) from ω\omega to the holonomy vectors in the summation.

Lemma 25.

For every ωK\omega\in K we have

|1R2vΛ(𝗀tω;R)χp(||𝗁(v)||)1R2vΛ(ω;R)χp(||𝗀t𝗁(v)||)|8ΞR2κ+8cϵ42R1+η\left|\frac{1}{R^{2}}\sum_{v\in\Lambda(\mathsf{g}^{t}\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}})-\frac{1}{R^{2}}\sum_{v\in\Lambda(\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{g}^{t}\mathsf{h}(v)|\!|_{\mathsf{}})\right|\leq\frac{8\Xi}{R^{2\kappa}}+8c\sqrt{\frac{\epsilon}{42R^{1+\eta}}}

all 0tS0\leq t\leq S.

Proof.

Fix R>0R>0 and ωK\omega\in K and t0t\geq 0. First note that vv belongs to Λ(𝗀tω;R)\Lambda(\mathsf{g}^{t}\omega;R) if and only if 𝗀t𝗁(v)\mathsf{g}^{-t}\mathsf{h}(v) belongs to the ellipse of width 2etR2e^{-t}R and height 2etR2e^{t}R. Thus

vΛ(𝗀tω;R)χp(||𝗁(v)||)=vΛ(ω;etR,etR)χp(||𝗀t𝗁(v)||)\sum_{v\in\Lambda(\mathsf{g}^{t}\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}})=\sum_{v\in\Lambda(\omega;e^{-t}R,e^{t}R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{g}^{t}\mathsf{h}(v)|\!|_{\mathsf{}})

holds for all R>0R>0. Next we estimate |Λ(ω;etR,etR)Λ(ω;R)||\Lambda(\omega;e^{t}R,e^{-t}R)\operatorname{\triangle}\Lambda(\omega;R)|. For every ω\omega we have

|Λ(ω;etR,etR)Λ(ω;R)|\displaystyle|\Lambda(\omega;e^{t}R,e^{-t}R)\operatorname{\triangle}\Lambda(\omega;R)|
\displaystyle\leq{} |Λ(ω;etR)||Λ(ω;etR)|\displaystyle|\Lambda(\omega;e^{t}R)|-|\Lambda(\omega;e^{-t}R)|
\displaystyle\leq{} (etR)2(1κ)Ξ+c(etR)2+(etR)2(1κ)Ξc(etR)2\displaystyle(e^{t}R)^{2(1-\kappa)}\Xi+c(e^{t}R)^{2}+(e^{-t}R)^{2(1-\kappa)}\Xi-c(e^{-t}R)^{2}
\displaystyle\leq{} 2Ξcosh(2t(1κ))R2(1κ)+2csinh(2t)R2\displaystyle 2\Xi\cosh(2t(1-\kappa))R^{2(1-\kappa)}+2c\sinh(2t)R^{2}

for all t0t\geq 0 using (15), as we may since ωK\omega\in K. It follows from the above that

|vΛ(𝗀tω;R)χp(||𝗁(v)||)vΛ(ω;R)χp(||𝗀t𝗁(v)||)|8ΞR2(1κ)+8cSR2\left|\sum_{v\in\Lambda(\mathsf{g}^{t}\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}})-\sum_{v\in\Lambda(\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{g}^{t}\mathsf{h}(v)|\!|_{\mathsf{}})\right|\leq 8\Xi R^{2(1-\kappa)}+8cSR^{2}

for all 0tS0\leq t\leq S. The conclusion follows from (18). ∎

Corollary 26.

For every ωK\omega\in K we have

|1R2vΛ(𝗀t𝗋θω;R)χp(||𝗁(v)||)1R2vΛ(ω;R)χp(||𝗀t𝗋θ𝗁(v)||)|8ΞR2κ+8cϵ42R1+η\left|\frac{1}{R^{2}}\sum_{v\in\Lambda(\mathsf{g}^{t}\mathsf{r}^{\theta}\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}})-\frac{1}{R^{2}}\sum_{v\in\Lambda(\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{g}^{t}\mathsf{r}^{\theta}\mathsf{h}(v)|\!|_{\mathsf{}})\right|\leq\frac{8\Xi}{R^{2\kappa}}+8c\sqrt{\frac{\epsilon}{42R^{1+\eta}}} (27)

for all 0θ<2π0\leq\theta<2\pi and all 0tS0\leq t\leq S.

Proof.

Apply the lemma with 𝗋θω\mathsf{r}^{\theta}\omega in place of ω\omega. Then note that when t=0t=0 the two sums are equal. ∎

2.4 A linear approximation

In this section we estimate the right-hand side of (20), proving the following result.

Theorem 28.

There is 0<γ<10<\gamma<1 such that

K12π02π1S0S1R4vΛ(ω;R)wΛ(ω;R)χp(||𝗀t𝗋θ𝗁(v)||)χp(||𝗀t𝗋θ𝗁(w)||)¯𝖽t𝖽θ𝖽𝗆(ω)=𝖮(1Rγ)\int\limits_{K}\frac{1}{2\pi}\int\limits_{0}^{2\pi}\frac{1}{S}\int\limits_{0}^{S}\frac{1}{R^{4}}\sum_{v\in\Lambda(\omega;R)}\sum_{w\in\Lambda(\omega;R)}\chi_{p}(|\!|_{\mathsf{}}\mathsf{g}^{t}\mathsf{r}^{\theta}\mathsf{h}(v)|\!|_{\mathsf{}})\,\overline{\chi_{p}(|\!|_{\mathsf{}}\mathsf{g}^{t}\mathsf{r}^{\theta}\mathsf{h}(w)|\!|_{\mathsf{}})}\,\mathsf{d}t\,\mathsf{d}\theta\,\mathsf{d}\mathsf{m}(\omega)=\mathsf{O}\left(\frac{1}{R^{\gamma}}\right) (29)

for almost all ωK\omega\in K.

For the proof of Theorem 28 we will need several lemmas. Write

α(u)=u12u22||u||β(u)=2u1u2||u||\alpha(u)=\frac{u_{1}^{2}-u_{2}^{2}}{|\!|_{\mathsf{}}u|\!|_{\mathsf{}}}\qquad\qquad\beta(u)=\frac{2u_{1}u_{2}}{|\!|_{\mathsf{}}u|\!|_{\mathsf{}}}

for any u=(v1,v2)u=(v_{1},v_{2}) in 2\mathbb{R}^{2}. Certainly |α(u)|||u|||\alpha(u)|\leq|\!|_{\mathsf{}}u|\!|_{\mathsf{}} and |β(u)|||u|||\beta(u)|\leq|\!|_{\mathsf{}}u|\!|_{\mathsf{}} for all u2u\in\mathbb{R}^{2}.

Lemma 30.

We have

|||u||+α(u)t||𝗀tu|||<ϵRη\Big{|}|\!|_{\mathsf{}}u|\!|_{\mathsf{}}+\alpha(u)t-|\!|_{\mathsf{}}\mathsf{g}^{t}u|\!|_{\mathsf{}}\Big{|}<\frac{\epsilon}{R^{\eta}} (31)

whenever 0tS0\leq t\leq S and u2u\in\mathbb{R}^{2} satisfies ||u||R|\!|_{\mathsf{}}u|\!|_{\mathsf{}}\leq R.

Proof.

Fix u2u\in\mathbb{R}^{2} with ||u||R|\!|_{\mathsf{}}u|\!|_{\mathsf{}}\leq R. Writing fu(t)=||𝗀tu||f_{u}(t)=|\!|_{\mathsf{}}\mathsf{g}^{t}u|\!|_{\mathsf{}} whenever 0tS0\leq t\leq S note that

et||u||fu(t)et||u||e^{-t}|\!|_{\mathsf{}}u|\!|_{\mathsf{}}\leq f_{u}(t)\leq e^{t}|\!|_{\mathsf{}}u|\!|_{\mathsf{}} (32)

and

fu′′(t)=2fu(t)(e2tu12e2tu22)2fu(t)3=fu(t)+4u12u22fu(t)3f_{u}^{\prime\prime}(t)=2f_{u}(t)-\frac{(e^{2t}u_{1}^{2}-e^{-2t}u_{2}^{2})^{2}}{f_{u}(t)^{3}}=f_{u}(t)+\frac{4u_{1}^{2}u_{2}^{2}}{f_{u}(t)^{3}}

for all tt. Fix ϵ>0\epsilon>0. From (18) and (32) we have

|fu′′(t)2t2|t22(et+4e3t)||u||42t2||u||42S2R=ϵRη\left|\frac{f^{\prime\prime}_{u}(t)}{2}t^{2}\right|\leq\frac{t^{2}}{2}(e^{t}+4e^{3t})|\!|_{\mathsf{}}u|\!|_{\mathsf{}}\leq 42t^{2}|\!|_{\mathsf{}}u|\!|_{\mathsf{}}\leq 42S^{2}R=\frac{\epsilon}{R^{\eta}}

for all 0<t<S0<t<S. Therefore (31) follows from the Lagrange form of the remainder in Taylor’s theorem. ∎

We now focus on

12π02π1S0Sχp(||𝗋θ𝗁(v)||+tα(𝗋θ𝗁(v)))χp(||𝗋θ𝗁(w)||+tα(𝗋θ𝗁(w)))¯𝖽t𝖽θ\frac{1}{2\pi}\int\limits_{0}^{2\pi}\frac{1}{S}\int\limits_{0}^{S}\chi_{p}(|\!|_{\mathsf{}}\mathsf{r}^{\theta}\mathsf{h}(v)|\!|_{\mathsf{}}+t\alpha(\mathsf{r}^{\theta}\mathsf{h}(v)))\,\overline{\chi_{p}(|\!|_{\mathsf{}}\mathsf{r}^{\theta}\mathsf{h}(w)|\!|_{\mathsf{}}+t\alpha(\mathsf{r}^{\theta}\mathsf{h}(w)))}\,\mathsf{d}t\,\mathsf{d}\theta (33)

for fixed v,wΛ(ω;R)v,w\in\Lambda(\omega;R). Note that

|(33)|=|12π02π1S0Sχp(tα(𝗋θ𝗁(v)))χp(tα(𝗋θ𝗁(w)))¯𝖽t𝖽θ||\eqref{eqn:afterEllipse}|=\left|\frac{1}{2\pi}\int\limits_{0}^{2\pi}\frac{1}{S}\int\limits_{0}^{S}\chi_{p}(t\alpha(\mathsf{r}^{\theta}\mathsf{h}(v)))\,\overline{\chi_{p}(t\alpha(\mathsf{r}^{\theta}\mathsf{h}(w)))}\,\mathsf{d}t\,\mathsf{d}\theta\right| (34)

because |||||\!|_{\mathsf{}}\cdot|\!|_{\mathsf{}} is 𝗋\mathsf{r} invariant.

Lemma 35.

If v,wΛ(ω;R)v,w\in\Lambda(\omega;R) satisfy

|||𝗁(v)||||𝗁(w)|||Rρ\Big{|}|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}-|\!|_{\mathsf{}}\mathsf{h}(w)|\!|_{\mathsf{}}\Big{|}\geq R^{\rho} (36)

then

|(33)|41+ϵ2πR+4π242Rηϵ1Rρ(logR+logπ4)|\eqref{eqn:afterEllipse}|\leq 4\cdot\frac{1+\epsilon}{2\pi R}+\frac{4}{\pi^{2}}\sqrt{\frac{42R^{\eta}}{\epsilon}}\frac{1}{R^{\rho}}\left(\log R+\log\frac{\pi}{4}\right)

holds.

Proof.

In absolute value (33) is equal to

|12π02πexp(2πiSp(α(𝗋θ𝗁(v))α(𝗋θ𝗁(w))))12πiSp(α(𝗋θ𝗁(v))α(𝗋θ𝗁(w)))𝖽θ|\left|\frac{1}{2\pi}\int\limits_{0}^{2\pi}\frac{\exp(2\pi iSp(\alpha(\mathsf{r}^{\theta}\mathsf{h}(v))-\alpha(\mathsf{r}^{\theta}\mathsf{h}(w))))-1}{2\pi iSp(\alpha(\mathsf{r}^{\theta}\mathsf{h}(v))-\alpha(\mathsf{r}^{\theta}\mathsf{h}(w)))}\,\mathsf{d}\theta\right|

by using (34) then integrating over tt. First, by making a substitution, we can assume 𝗁(w)\mathsf{h}(w) is horizontal. We have β(𝗁(w))=0\beta(\mathsf{h}(w))=0 and α(𝗁(w))=||𝗁(w)||\alpha(\mathsf{h}(w))=|\!|_{\mathsf{}}\mathsf{h}(w)|\!|_{\mathsf{}}. So

α(𝗋θ𝗁(v))α(𝗋θ𝗁(w))\displaystyle\alpha(\mathsf{r}^{\theta}\mathsf{h}(v))-\alpha(\mathsf{r}^{\theta}\mathsf{h}(w))
=\displaystyle={} (α(𝗁(v))cos(2θ)β(𝗁(v))sin(2θ))(α(𝗁(w))cos(2θ)β(𝗁(w))sin(2θ))\displaystyle\Big{(}\alpha(\mathsf{h}(v))\cos(2\theta)-\beta(\mathsf{h}(v))\sin(2\theta)\Big{)}-\Big{(}\alpha(\mathsf{h}(w))\cos(2\theta)-\beta(\mathsf{h}(w))\sin(2\theta)\Big{)}
=\displaystyle={} (α(𝗁(v))||𝗁(w)||)cos(2θ)β(𝗁(v))sin(2θ)\displaystyle\Big{(}\alpha(\mathsf{h}(v))-|\!|_{\mathsf{}}\mathsf{h}(w)|\!|_{\mathsf{}}\Big{)}\cos(2\theta)-\beta(\mathsf{h}(v))\sin(2\theta)
=\displaystyle={} A(v,w)sin(ϕ2θ)\displaystyle A(v,w)\sin(\phi-2\theta)

where A(v,w)2=(α(𝗁(v))||𝗁(w)||)2+β(𝗁(v))2A(v,w)^{2}=\left(\alpha(\mathsf{h}(v))-|\!|_{\mathsf{}}\mathsf{h}(w)|\!|_{\mathsf{}}\right)^{2}+\beta(\mathsf{h}(v))^{2} and ϕ\phi is chosen appropriately using the angle addition formula. We are therefore interested in

|12π02πexp(2πiSpA(v,w)sin(ϕ2θ))12πiSpA(v,w)sin(ϕ2θ)𝖽θ|\left|\frac{1}{2\pi}\int\limits_{0}^{2\pi}\frac{\exp(2\pi iSpA(v,w)\sin(\phi-2\theta))-1}{2\pi iSpA(v,w)\sin(\phi-2\theta)}\,\mathsf{d}\theta\right|

and consider separately the integral over small intervals centered at zeros of sin(ϕ2θ)\sin(\phi-2\theta) and what remains. Precisely, if II is an interval of radius 1/R1/R centered at a zero of sin(ϕ2θ)\sin(\phi-2\theta) then

|SpA(v,w)sin(ϕ2θ)|p5ϵ42R1(1+η)/2|sin(ϕ2θ)|p5ϵ421R(1+η)/2<δ|SpA(v,w)\sin(\phi-2\theta)|\leq p\sqrt{\frac{5\epsilon}{42}}R^{1-(1+\eta)/2}|\sin(\phi-2\theta)|\leq p\sqrt{\frac{5\epsilon}{42}}\frac{1}{R^{(1+\eta)/2}}<\delta

for all θI\theta\in I by (17), giving

|exp(2πiSpA(v,w)sin(ϕ2θ))12πiSpA(v,w)sin(ϕ2θ)1|ϵ\left|\frac{\exp(2\pi iSpA(v,w)\sin(\phi-2\theta))-1}{2\pi iSpA(v,w)\sin(\phi-2\theta)}-1\right|\leq\epsilon

for every θI\theta\in I by definition of δ\delta (see (16)). There being four such intervals, we can estimate

|12π02πexp(2πiSpA(v,w)sin(ϕ2θ))12πiSpA(v,w)sin(ϕ2θ)𝖽θ|\displaystyle\left|\frac{1}{2\pi}\int\limits_{0}^{2\pi}\frac{\exp(2\pi iSpA(v,w)\sin(\phi-2\theta))-1}{2\pi iSpA(v,w)\sin(\phi-2\theta)}\,\mathsf{d}\theta\right| (37)
\displaystyle\leq{} 41+ϵ2πR+42πϕ2+1Rϕ2+π21R22πSA(v,w)|sin(ϕ2θ)|𝖽θ\displaystyle 4\cdot\frac{1+\epsilon}{2\pi R}+\frac{4}{2\pi}\int\limits_{\frac{\phi}{2}+\frac{1}{R}}^{\frac{\phi}{2}+\frac{\pi}{2}-\frac{1}{R}}\frac{2}{2\pi SA(v,w)|\sin(\phi-2\theta)|}\,\mathsf{d}\theta

by trivially estimating the numerator on the complement of the four intervals. So (after substituting away ϕ\phi) we wish to make

1πSA(v,w)4π1Rπ41sin(2θ)𝖽θ\frac{1}{\pi SA(v,w)}\frac{4}{\pi}\int\limits_{\frac{1}{R}}^{\frac{\pi}{4}}\frac{1}{\sin(2\theta)}\,\mathsf{d}\theta (38)

small. Now 4πθsin(2θ)\frac{4}{\pi}\theta\leq\sin(2\theta) on [0,π4][0,\frac{\pi}{4}] so

4π1Rπ41sin(2θ)𝖽θ1Rπ41θ𝖽θ=logπ4+logR\frac{4}{\pi}\int\limits_{\frac{1}{R}}^{\frac{\pi}{4}}\frac{1}{\sin(2\theta)}\,\mathsf{d}\theta\leq\int\limits_{\frac{1}{R}}^{\frac{\pi}{4}}\frac{1}{\theta}\,\mathsf{d}\theta=\log\frac{\pi}{4}+\log R

holds. With (18) the quantity (38) becomes

4π242R1+ηϵ1A(v,w)(logR+logπ4)\frac{4}{\pi^{2}}\sqrt{\frac{42R^{1+\eta}}{\epsilon}}\frac{1}{A(v,w)}\left(\log R+\log\frac{\pi}{4}\right)

so we wish to determine when A(v,w)A(v,w) not too small. So we estimate the size of A(v,w)A(v,w) recalling that 𝗁(w)\mathsf{h}(w) is horizontal. Let θv,w\theta_{v,w} be the angle between 𝗁(v)\mathsf{h}(v) and 𝗁(w)\mathsf{h}(w) so that 𝗁(v)=||𝗁(v)||(cosθv,w,sinθv,w)\mathsf{h}(v)=|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}(\cos\theta_{v,w},\sin\theta_{v,w}). Then

A(v,w)2\displaystyle A(v,w)^{2} =(||𝗁(v)||cos(2θv,w)||𝗁(w)||)2+(||𝗁(v)||sin(2θv,w))2\displaystyle=(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}\cos(2\theta_{v,w})-|\!|_{\mathsf{}}\mathsf{h}(w)|\!|_{\mathsf{}})^{2}+(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}\sin(2\theta_{v,w}))^{2}
=||𝗁(v)||2+||𝗁(w)||22||𝗁(v)||||𝗁(w)||cos(2θv,w)\displaystyle=|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}^{2}+|\!|_{\mathsf{}}\mathsf{h}(w)|\!|_{\mathsf{}}^{2}-2|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}|\!|_{\mathsf{}}\mathsf{h}(w)|\!|_{\mathsf{}}\cos(2\theta_{v,w})
(||𝗁(v)||||𝗁(w)||)2\displaystyle\geq(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}-|\!|_{\mathsf{}}\mathsf{h}(w)|\!|_{\mathsf{}})^{2}

and (36) concludes the proof. ∎

Lemma 39.

Fix ωK\omega\in K. We have

|{uΛ(ω;R):|||𝗁(v)||||𝗁(u)|||Rρ}|(4c+8Ξ)R1+ρ\left|\left\{u\in\Lambda(\omega;R):\Big{|}|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}-|\!|_{\mathsf{}}\mathsf{h}(u)|\!|_{\mathsf{}}\Big{|}\leq R^{\rho}\right\}\right|\leq(4c+8\Xi)R^{1+\rho}

for every vΛ(ω;R)v\in\Lambda(\omega;R).

Proof.

The cardinality of the set is the same as the number of saddle connections of ω\omega in the annulus 𝖲1×[||𝗁(v)||Rρ,||𝗁(v)||+Rρ]\mathsf{S}^{1}\times[|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}-R^{\rho},|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}+R^{\rho}]. By Theorem 4 this cardinality is at most

|c(||𝗁(v)||+Rρ)2c(||𝗁(v)||Rρ)2+Ξ(||𝗁(v)||+Rρ)2(1κ)+Ξ(||𝗁(v)||Rρ)2(1κ)|\displaystyle|c(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}+R^{\rho})^{2}-c(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}-R^{\rho})^{2}+\Xi(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}+R^{\rho})^{2(1-\kappa)}+\Xi(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}-R^{\rho})^{2(1-\kappa)}|
\displaystyle\leq{} 4c||𝗁(v)||Rρ+2Ξ(||𝗁(v)||+Rρ)2(1κ)\displaystyle 4c|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}R^{\rho}+2\Xi(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}}+R^{\rho})^{2(1-\kappa)}
\displaystyle\leq{} 4cR1+ρ+8ΞR2(1κ)\displaystyle 4cR^{1+\rho}+8\Xi R^{2(1-\kappa)}

by choice of ρ\rho. ∎

We can now give the proof of Theorem 28.

Proof of Theorem 28.

It suffices to prove that

K1R4vΛ(ω;R)wΛ(ω;R)12π02π1S0Sχp(tα(𝗋θ𝗁(v))χp(tα(𝗋θ𝗁(w))¯𝖽θ𝖽t𝖽𝗆(ω)=𝖮(1Rγ)\int\limits_{K}\frac{1}{R^{4}}\sum_{v\in\Lambda(\omega;R)}\sum_{w\in\Lambda(\omega;R)}\frac{1}{2\pi}\int\limits_{0}^{2\pi}\frac{1}{S}\int\limits_{0}^{S}\chi_{p}(t\alpha(\mathsf{r}^{\theta}\mathsf{h}(v))\,\overline{\chi_{p}(t\alpha(\mathsf{r}^{\theta}\mathsf{h}(w))}\,\mathsf{d}\theta\,\mathsf{d}t\,\mathsf{d}\mathsf{m}(\omega)=\mathsf{O}\left(\frac{1}{R^{\gamma}}\right)

for some 0<γ<10<\gamma<1 by Lemma 30. For each ωK\omega\in K consider the inner sum over ww above in two parts, according to whether ww satisfies (36). By Lemma 39 and the trivial bound the sum over those ww not satisfying (36) is 𝖮(R1+ρ)\mathsf{O}(R^{1+\rho}) which is good enough because ρ<1\rho<1. For ww that do satisfy (36) we apply Lemma 35. In combination we obtain the estimate

|1R4vΛ(α;R)wΛ(α;R)12π02π1T0Tχp(||𝗋θv||𝟤+tα(𝗋θv))χp(||𝗋θw||𝟤+tα(𝗋θw))𝖽t𝖽θ|\displaystyle\left|\frac{1}{R^{4}}\sum_{v\in\Lambda(\alpha;R)}\sum_{w\in\Lambda(\alpha;R)}\frac{1}{2\pi}\int\limits_{0}^{2\pi}\frac{1}{T}\int\limits_{0}^{T}\chi_{p}(|\!|_{\mathsf{}}\mathsf{r}^{\theta}v|\!|_{\mathsf{2}}+t\alpha(\mathsf{r}^{\theta}v))\chi_{p}(|\!|_{\mathsf{}}\mathsf{r}^{\theta}w|\!|_{\mathsf{2}}+t\alpha(\mathsf{r}^{\theta}w))\,\mathsf{d}t\,\mathsf{d}\theta\right|
\displaystyle\leq{} 1R4(c4R2)(c3R1+ρ)+1R4(c4R2)2(41+ϵ2πR+8π26|p|ϵ(logR+logπ4)R12Rρ)\displaystyle\frac{1}{R^{4}}(c_{4}R^{2})(c_{3}R^{1+\rho})+\frac{1}{R^{4}}(c_{4}R^{2})^{2}\left(4\cdot\frac{1+\epsilon}{2\pi R}+\frac{8}{\pi^{2}}\sqrt{\frac{6|p|}{\epsilon}}\left(\log R+\log\frac{\pi}{4}\right)\frac{R^{\frac{1}{2}}}{R^{\rho}}\right)

which gives the theorem because 1>ρ>121>\rho>\frac{1}{2}. ∎

2.5 Proof of Theorem 1

Here we combine the preceding subsections to prove Theorem 1.

Proof of Theorem 1.

Fix a sequence τ1>τ2>1\tau_{1}>\tau_{2}>\cdots\to 1. The compact sets KK exhause almost all of \mathcal{H} as Ξ\Xi\to\infty and r0r\to 0. It therefore suffices by Lemma 9 and Lemma 12 to prove for every ii\in\mathbb{N} that (13) holds for 𝗆\mathsf{m} almost every ωK\omega\in K. Fix ii\in\mathbb{N} and write τ\tau for τi\tau_{i}. Let γ\gamma be as in Theorem 28. Taking σ\sigma small enough and R2=τJ/cR^{2}=\lceil{\tau^{J}}\rceil/c in Theorem 19 and applying using both Corollary 26 and Theorem 28 gives ψ>0\psi>0 such that

𝗆({ωK:|cτJvΛ(ω;τJ/c)χp(||𝗁(v)||)|2(cτJ)σ/2})=𝖮(1τJψ)\mathsf{m}\left(\left\{\omega\in K:\left|\frac{c}{\lceil{\tau^{J}}\rceil}\sum_{v\in\Lambda(\omega;\sqrt{\lceil{\tau^{J}}\rceil/c})}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}})\right|^{2}\geq\left(\frac{c}{\lceil{\tau^{J}}\rceil}\right)^{\sigma/2}\right\}\right)=\mathsf{O}\left(\frac{1}{\lceil{\tau^{J}}\rceil^{\psi}}\right)

for all JJ\in\mathbb{N} large enough. Since the right-hand side is summable we conclude from the Borel-Cantelli lemma that

limJcτJvΛ(ω;τJ/c)χp(||𝗁(v)||)=0\lim_{J\to\infty}\frac{c}{\lceil{\tau^{J}}\rceil}\sum_{v\in\Lambda(\omega;\sqrt{\lceil{\tau^{J}}\rceil/c})}\chi_{p}(|\!|_{\mathsf{}}\mathsf{h}(v)|\!|_{\mathsf{}})=0

as desired. ∎

Appendix A Equidistribution of the lengths of the primitive vectors in integer lattices by Daniel El-Baz and Bingrong Huang

For a real number xx, let x\lfloor x\rfloor be the largest integer less than or equal to xx, and let {x}=xx\{x\}=x-\lfloor x\rfloor be its fractional part. Let \|\cdot\| be the usual Euclidean norm in 2\mathbb{R}^{2}. Let prim2={(a,b)2:gcd(a,b)=1}\mathbb{Z}_{\mathrm{prim}}^{2}=\{(a,b)\in\mathbb{Z}^{2}:\gcd(a,b)=1\} and 𝒮R={𝐯prim2:g𝐯R}\mathcal{S}_{R}=\{\mathbf{v}\in\mathbb{Z}_{\mathrm{prim}}^{2}\,:\,\|g\mathbf{v}\|\leq R\}.

Theorem 40.

For every gGL2()g\in\mathrm{GL}_{2}(\mathbb{R}), the sequence ({g𝐯})𝐯𝒮R(\{\|g\mathbf{v}\|\})_{\mathbf{v}\in\mathcal{S}_{R}} is uniformly distributed as RR\rightarrow\infty.

Proof.

We want to estimate, for every α(0,1)\alpha\in(0,1), the following quantity:

NR(α)=#{𝐯𝒮R:{g𝐯}α}.N_{R}(\alpha)=\#\{\mathbf{v}\in\mathcal{S}_{R}\,:\,\{\|g\mathbf{v}\|\}\leq\alpha\}.

We rewrite

NR(α)\displaystyle N_{R}(\alpha) =dRdg𝐯d+αg𝐯R1.\displaystyle=\sum_{d\leq\lfloor R\rfloor}\sum_{\begin{subarray}{c}d\leq\|g\mathbf{v}\|\leq d+\alpha\\ \|g\mathbf{v}\|\leq R\end{subarray}}1.

Let 𝒟g={𝐱2:g𝐱1}\mathcal{D}_{g}=\{\mathbf{x}\in\mathbb{R}^{2}:\|g\mathbf{x}\|\leq 1\}. By the prime number theorem estimate for the primitive lattice point problem for a compact convex domain with smooth boundary (which our elliptic domain 𝒟g\mathcal{D}_{g} certainly is) [HN96, Equation 1.6], we have

P(y):=#{𝐯prim2:g𝐯y}=𝐯prim2g𝐯y1=6π2area(𝒟g)y2+o(y).P(y):=\#\{\mathbf{v}\in\mathbb{Z}_{\mathrm{prim}}^{2}\,:\,\|g\mathbf{v}\|\leq y\}=\sum_{\begin{subarray}{c}\mathbf{v}\in\mathbb{Z}_{\mathrm{prim}}^{2}\\ \|g\mathbf{v}\|\leq y\end{subarray}}1=\frac{6}{\pi^{2}}\mathrm{area}(\mathcal{D}_{g})y^{2}+o(y).

Hence we obtain

NR(α)\displaystyle N_{R}(\alpha) =dR1(P(d+α)P(d))+O(R)\displaystyle=\sum_{d\leq\lfloor R\rfloor-1}(P(d+\alpha)-P(d))+O(R)
=dR1(6π2area(𝒟g)2dα+o(d))+O(R)\displaystyle=\sum_{d\leq\lfloor R\rfloor-1}\left(\frac{6}{\pi^{2}}\mathrm{area}(\mathcal{D}_{g})2d\alpha+o(d)\right)+O(R)
=6π2area(𝒟g)R2α+o(R2).\displaystyle=\frac{6}{\pi^{2}}\mathrm{area}(\mathcal{D}_{g})R^{2}\alpha+o(R^{2}).

This implies our claim. ∎

References

  • [AC12] J.. Athreya and J. Chaika “The distribution of gaps for saddle connection directions” In Geom. Funct. Anal. 22.6, 2012, pp. 1491–1516 DOI: 10.1007/s00039-012-0164-9
  • [ACL15] Jayadev S. Athreya, Jon Chaika and Samuel Lelièvre “The gap distribution of slopes on the golden L” In Recent trends in ergodic theory and dynamical systems 631, Contemp. Math. Amer. Math. Soc., Providence, RI, 2015, pp. 47–62 DOI: 10.1090/conm/631/12595
  • [ACM17] Jayadev Athreya, Yitwah Cheung and Howard Masur “Siegel-Veech transforms are in L2\mathrm{L}^{2}”, 2017 arXiv:1711.08537 [math.DS]
  • [Doz17] Benjamin Dozier “Convergence of Siegel-Veech constants”, 2017 arXiv:1701.00175 [math.DS]
  • [Doz17a] Benjamin Dozier “Equidistribution of saddle connections on translation surfaces”, 2017 arXiv:1705.10847 [math.DS]
  • [EM01] Alex Eskin and Howard Masur “Asymptotic formulas on flat surfaces” In Ergodic Theory Dynam. Systems 21.2, 2001, pp. 443–478 URL: https://doi.org/10.1017/S0143385701001225
  • [EM11] Alex Eskin and Maryam Mirzakhani “Counting closed geodesics in moduli space” In J. Mod. Dyn. 5.1, 2011, pp. 71–105 DOI: 10.3934/jmd.2011.5.71
  • [EMM15] Alex Eskin, Maryam Mirzakhani and Amir Mohammadi “Isolation, equidistribution, and orbit closures for the SL(2,)\mathrm{SL}(2,\mathbb{R}) action on moduli space” In Ann. of Math. (2) 182.2, 2015, pp. 673–721 DOI: 10.4007/annals.2015.182.2.7
  • [HN96] Martin N. Huxley and Werner Georg Nowak “Primitive lattice points in convex planar domains” In Acta Arith. 76.3, 1996, pp. 271–283 DOI: 10.4064/aa-76-3-271-283
  • [Kna01] Anthony W. Knapp “Representation theory of semisimple groups”, Princeton Landmarks in Mathematics Princeton University Press, Princeton, NJ, 2001, pp. xx+773
  • [Mas82] Howard Masur “Interval exchange transformations and measured foliations” In Ann. of Math. (2) 115.1, 1982, pp. 169–200 DOI: 10.2307/1971341
  • [Mas88] Howard Masur “Lower bounds for the number of saddle connections and closed trajectories of a quadratic differential” In Holomorphic functions and moduli, Vol. I (Berkeley, CA, 1986) 10, Math. Sci. Res. Inst. Publ. Springer, New York, 1988, pp. 215–228 DOI: 10.1007/978-1-4613-9602-4_20
  • [Mas90] Howard Masur “The growth rate of trajectories of a quadratic differential” In Ergodic Theory Dynam. Systems 10.1, 1990, pp. 151–176 DOI: 10.1017/S0143385700005459
  • [NRW17] Amos Nevo, Rene Rühr and Barack Weiss “Effective counting on translation surfaces”, 2017 arXiv:1708.06263 [math.DS]
  • [UW16] Caglar Uyanik and Grace Work “The distribution of gaps for saddle connections on the octagon” In Int. Math. Res. Not. IMRN, 2016, pp. 5569–5602 DOI: 10.1093/imrn/rnv317
  • [Vee82] William A. Veech “Gauss measures for transformations on the space of interval exchange maps” In Ann. of Math. (2) 115.1, 1982, pp. 201–242 DOI: 10.2307/1971391
  • [Vee98] William A. Veech “Siegel measures” In Ann. of Math. (2) 148.3, 1998, pp. 895–944 DOI: 10.2307/121033
  • [Vor03] Yaroslav Vorobets “Periodic geodesics on translation surfaces”, 2003 arXiv:math/0307249 [math.DS]
  • [Vor05] Yaroslav Vorobets “Periodic geodesics on generic translation surfaces” In Algebraic and topological dynamics 385, Contemp. Math. Amer. Math. Soc., Providence, RI, 2005, pp. 205–258 DOI: 10.1090/conm/385/07199