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

Asymptotic estimates of large gaps between directions in certain planar quasicrystals

Gustav Hammarhjelm Andreas Strömbergsson Department of Mathematics, Uppsala University, Box 480, SE-75106, Uppsala, Sweden [email protected]  and  Shucheng Yu School of Mathematical Sciences, University of Science and Technology of China (USTC), 230026, Hefei, China [email protected]
(Date: October 1, 2024.)
Abstract.

For quasicrystals of cut-and-project type in d\mathbb{R}^{d}, it was proved by Marklof and Strömbergsson [27] that the limit local statistical properties of the directions to the points in the set are described by certain SLd()\operatorname{SL}_{d}(\mathbb{R})-invariant point processes. In the present paper we make a detailed study of the tail asymptotics of the limiting gap statistics of the directions, for certain specific classes of planar quasicrystals.

All three authors were supported by the Knut and Alice Wallenberg Foundation

Dedicated to Gustav Hammarhjelm (1992–2022).

1. Introduction

1.1. Gaps between directions in planar point sets

Given a locally finite point set 𝒫{\mathcal{P}} in the plane 2\mathbb{R}^{2}, we consider its set of directions, that is, the set of points 𝒗^:=𝒗1𝒗\widehat{{\text{\boldmath$v$}}}:=\|{\text{\boldmath$v$}}\|^{-1}{\text{\boldmath$v$}} on the unit circle S11\operatorname{S{}}_{1}^{1} as 𝒗v runs through 𝒫{𝟎}{\mathcal{P}}\smallsetminus\{\mathbf{0}\}. For each R>0R>0, let ΔR\Delta_{R} be the finite multi-set of directions to points in 𝒫{\mathcal{P}} within distance RR from the origin, i.e.,

(1.1) ΔR:={𝒗^:𝒗𝒫, 0<𝒗R}.\displaystyle\Delta_{R}:=\{\widehat{{\text{\boldmath$v$}}}\>:\>{\text{\boldmath$v$}}\in{\mathcal{P}},\>0<\|{\text{\boldmath$v$}}\|\leq R\}.

Throughout the paper we will assume that 𝒫{\mathcal{P}} has an asymptotic density c𝒫>0c_{\mathcal{P}}>0, meaning that for any bounded set 𝒟2{\mathcal{D}}\subset\mathbb{R}^{2} with boundary of measure zero,

(1.2) limR#(𝒫R𝒟)R2=c𝒫Area(𝒟).\displaystyle\lim_{R\to\infty}\frac{\#({\mathcal{P}}\cap R{\mathcal{D}})}{R^{2}}=c_{{\mathcal{P}}}\operatorname{Area}({\mathcal{D}}).

It then follows that the multi-set ΔR\Delta_{R} has cardinality c𝒫πR2\sim c_{\mathcal{P}}\hskip 1.0pt\pi R^{2} as RR\to\infty, and furthermore that this multi-set becomes asymptotically equidistributed along the unit circle, in the sense that for any arc IS11I\subset\operatorname{S{}}_{1}^{1},

limR#(ΔRI)#ΔR=|I|2π,\displaystyle\lim_{R\to\infty}\frac{\#(\Delta_{R}\cap I)}{\#\Delta_{R}}=\frac{|I|}{2\pi},

where |I||I| denotes the length of II (in particular |S11|=2π|\operatorname{S{}}_{1}^{1}|=2\pi).

In this situation, it is natural to consider finer questions about the local statistics of the points in ΔR\Delta_{R}. A particular statistics which has been much studied is the distribution of normalized gaps between the points in ΔR\Delta_{R}, as RR\to\infty. For example, when 𝒫{\mathcal{P}} is the set of primitive lattice points in 2\mathbb{Z}^{2}, the limiting distribution of normalized gaps was explicitly determined by Boca, Cobeli and Zaharescu [9]. More generally, if 𝒫{\mathcal{P}} is an arbitrary lattice (possibly translated) in d\mathbb{R}^{d}, for any d2d\geq 2, the limiting distribution of general local statistics of directions was proved to exist by Marklof and Strömbergsson [25]; see also [14] and [23] regarding convergence of related moments. In particular it was noted in [25] that the explicit limiting distribution computed in [9] remains valid when replacing 2\mathbb{Z}^{2} by an arbitrary lattice in 2\mathbb{R}^{2}; furthermore, when 𝒫{\mathcal{P}} is an arbitrary ’irrational’ translate of a lattice in 2\mathbb{R}^{2}, the limiting distribution of gaps between directions coincides with the gap distribution for the fractional parts of n\sqrt{n} calculated by Elkies and McMullen [15]; see [25, Sec. 1.2].

Athreya and Chaika [2, Prop. 3.10] have proved that the limiting distribution of gaps between directions exists in the case when 𝒫{\mathcal{P}} is the set of holonomy vectors of either the saddle connections or periodic cylinders on a generic translation surface; see also [3], [44], [24], [40], [8] and [35] for later studies with more precise results in specific cases. Also in hyperbolic nn-space, for any point set which is the orbit of a lattice within the group of orientation preserving isometries, the analogous limiting gap distribution and also more general local statistics of directions, has been proved to exist by Marklof and Vinogradov [31]; for related work see [10], [22], [38].

The present paper concerns the case of 𝒫{\mathcal{P}} being a regular cut-and-project set (also referred to as a Euclidean model set). For this case, the limit distribution of normalized gaps between directions was proved to exist, and given a complicated but explicit description, by Marklof and Strömbergsson in [27]; see Theorem 1 below. See also Rühr, Smilansky and Weiss [39] for closely related investigations, and El-Baz [13] for an extension to so called adelic model sets. The results of [27] answered some questions which had been raised by Baake, Götze, Huck and Jakobi in [4], where a numerical investigation was carried out of the normalized gap distribution between directions for several vertex sets coming from aperiodic tilings, some of these being of cut-and-project type and others not. Building on the results of [27], Hammarhjelm [20] explicitly determined the limit of the minimal normalized gap between the directions to the points in 𝒫{\mathcal{P}}, for 𝒫{\mathcal{P}} belonging to either of two families of planar quasicrystals, including both the Ammann-Beenker point set and the vertex sets of some rhombic Penrose tilings. Also in [20], the asymptotic density of visible points was explicitly determined for several families of planar quasicrystals, including the two families just mentioned.

The main purpose of the present paper is to continue the explicit study begun in [20] of the limit distribution of normalized gaps between directions in the case of 𝒫{\mathcal{P}} belonging to certain families of planar cut-and-project sets. Our focus will be on asymptotics for large gaps between directions. The present work grew out of the initial study carried out by Hammarhjelm [19].

We remark that the methods developed in the present paper can be expected to be useful also for questions related to the Lorentz gas on a cut-and-project scatterer configuration – namely, for the task of obtaining asymptotic estimates for the transition probabilities in the transport (Markov) process which arises when considering such a Lorentz gas in the limit of low scatterer density [26], [29]. For the case of a lattice scatterer configuration, such asymptotic estimates were obtained in [28], and found an important application in [30].

1.2. The limit distribution of gaps for cut-and-project sets

To recall the precise definition of cut-and-project sets considered in [27], let d,m1d,m\geq 1, set n=d+mn=d+m, and let {\mathcal{L}} be a lattice of full rank in n=d×m\mathbb{R}^{n}=\mathbb{R}^{d}\times\mathbb{R}^{m}. We refer to d\mathbb{R}^{d} and m\mathbb{R}^{m} as the physical space and internal space, respectively. In the present paper the dimension of the physical space will always be d=2d=2. We write π\pi and πint\pi_{\operatorname{int}} for the orthogonal projection of n\mathbb{R}^{n} onto the first dd coordinates and last mm coordinates. Let 𝒜{\mathcal{A}} be the closure of πint()\pi_{\operatorname{int}}({\mathcal{L}}); this is a closed abelian subgroup of m\mathbb{R}^{m}. Let 𝒜{\mathcal{A}}^{\circ} be the identity component of 𝒜{\mathcal{A}}; this is a linear subspace of m\mathbb{R}^{m}, and set m=dim𝒜=dim𝒜m^{\prime}=\dim{\mathcal{A}}=\dim{\mathcal{A}}^{\circ}. Let 𝒲{\mathcal{W}}, the window, be a bounded subset of 𝒜{\mathcal{A}} with nonempty interior (with respect to the topology of 𝒜{\mathcal{A}}). We will always assume that 𝒲{\mathcal{W}} and {\mathcal{L}} are such that

(1.3) 𝒚1,𝒚2:if πint(𝒚1)𝒲 and πint(𝒚2)𝒲 and π(𝒚1)=π(𝒚2), then 𝒚1=𝒚2.\displaystyle\forall\,{\text{\boldmath$y$}}_{1},{\text{\boldmath$y$}}_{2}\in{\mathcal{L}}:\quad\text{if $\pi_{\operatorname{int}}({\text{\boldmath$y$}}_{1})\in{\mathcal{W}}$ and $\pi_{\operatorname{int}}({\text{\boldmath$y$}}_{2})\in{\mathcal{W}}$ and $\pi({\text{\boldmath$y$}}_{1})=\pi({\text{\boldmath$y$}}_{2})$, then ${\text{\boldmath$y$}}_{1}={\text{\boldmath$y$}}_{2}$.}

We now define the cut-and-project set associated to 𝒲{\mathcal{W}} and {\mathcal{L}} to be

(1.4) 𝒫=𝒫(𝒲,)={π(𝒚):𝒚,πint(𝒚)𝒲}d.\displaystyle{\mathcal{P}}={\mathcal{P}}({\mathcal{W}},{\mathcal{L}})=\bigl{\{}\pi({\text{\boldmath$y$}})\>:\>{\text{\boldmath$y$}}\in{\mathcal{L}},\>\pi_{\operatorname{int}}({\text{\boldmath$y$}})\in{\mathcal{W}}\bigr{\}}\subset\mathbb{R}^{d}.

We will always assume that 𝒫(𝒲,){\mathcal{P}}({\mathcal{W}},{\mathcal{L}}) is regular, meaning that 𝒲\partial{\mathcal{W}} has measure zero with respect to the Haar measure of 𝒜{\mathcal{A}}. Under these assumptions, it is known that 𝒫{\mathcal{P}} has an asymptotic density, i.e. (1.2) holds, with c𝒫c_{\mathcal{P}} being given by a simple explicit expression in terms of {\mathcal{L}} and 𝒲{\mathcal{W}} [26, Prop. 3.2]; see also (1.11) below.

Let us view the unit circle S11\operatorname{S{}}_{1}^{1} as the set of zz\in\mathbb{C} with |z|=1|z|=1, and for any zS11z\in\operatorname{S{}}_{1}^{1} let us call the number 12πarg(z)\frac{1}{2\pi}\arg(z) the normalized angle of zz; this gives an identification between S11\operatorname{S{}}_{1}^{1} and /\mathbb{R}/\mathbb{Z}. Let us order the normalized angles of the points in ΔR\Delta_{R} in an increasing list as

(1.5) 12<ξR,1ξR,2ξR,N(R)12,\displaystyle-\tfrac{1}{2}<\xi_{R,1}\leq\xi_{R,2}\leq\cdots\leq\xi_{R,N(R)}\leq\tfrac{1}{2},

where N(R)=#ΔRN(R)=\#\Delta_{R}. Also set ξR,0=ξR,N(R)1\xi_{R,0}=\xi_{R,N(R)}-1.

The following result was proved in [27] (see also Remark 1.7 below).

Theorem 1.

[27, Cor. 3] Let 𝒫{\mathcal{P}} be a regular cut-and-project set in 2\mathbb{R}^{2}. Then there exists a decreasing function F:0[0,1]F:\mathbb{R}_{\geq 0}\to[0,1], which is continuous on >0\mathbb{R}_{>0}, such that for every s0s\geq 0,

(1.6) limR#{1jN(R):N(R)(ξR,jξR,j1)s}N(R)=F(s).\displaystyle\lim_{R\to\infty}\frac{\#\{1\leq j\leq N(R)\>:\>N(R)(\xi_{R,j}-\xi_{R,j-1})\geq s\}}{N(R)}=F(s).

Furthermore, this limit relation (1.6) remains true, with the function FF unchanged, if 𝒫{\mathcal{P}} is replaced by 𝒫T={𝐯T:𝐯𝒫}{\mathcal{P}}\,T=\{{\text{\boldmath$v$}}T\>:\>{\text{\boldmath$v$}}\in{\mathcal{P}}\} for any fixed TGL2()T\in\operatorname{GL}_{2}(\mathbb{R}).

We recall the explicit formula for F(s)F(s) from [27] in Section 3.3 below. Figure 1 shows conjectural graphs of the function F(s)-F^{\prime}(s), i.e. the limiting density of normalized gaps, for some examples of vertex sets of Ammann-Beenker, Gähler’s shield, and Tübingen triangle tilings.

Remark 1.7.

Theorem 1 is a special case of [27, Cor. 3], since in [27, Cor. 3] we allow {\mathcal{L}} in 𝒫=𝒫(𝒲,){\mathcal{P}}={\mathcal{P}}({\mathcal{W}},{\mathcal{L}}) to be a translate of a lattice in n\mathbb{R}^{n}, whereas in the present paper we always require {\mathcal{L}} to be a genuine lattice, i.e. 𝟎\mathbf{0}\in{\mathcal{L}}. In fact, in Theorem 2 below we will also assume that 𝟎𝒲\mathbf{0}\in{\mathcal{W}}, and so 𝟎𝒫\mathbf{0}\in{\mathcal{P}}.

We also point out that the proof of [27, Cor. 3] immediately extends (by utilizing the freedom of choice of the measure “λ\lambda” in [27, Thm. 2]) to show that the limiting gap distribution for 𝒫{\mathcal{P}} remains the same if we restrict attention to the directions lying in any fixed subinterval of S11\operatorname{S{}}_{1}^{1}. That is, the following more general version of (1.6) holds: For any fixed 12α1<α212-\frac{1}{2}\leq\alpha_{1}<\alpha_{2}\leq\frac{1}{2},

(1.8) limR#{1jN(R):ξR,j(α1,α2],N(R)(ξR,jξR,j1)s}(α2α1)N(R)=F(s).\displaystyle\lim_{R\to\infty}\frac{\#\{1\leq j\leq N(R)\>:\>\xi_{R,j}\in(\alpha_{1},\alpha_{2}],\>N(R)(\xi_{R,j}-\xi_{R,j-1})\geq s\}}{(\alpha_{2}-\alpha_{1})\cdot N(R)}=F(s).

The last statement of Theorem 1, regarding the invariance of F(s)F(s) when replacing 𝒫{\mathcal{P}} by 𝒫T{\mathcal{P}}\,T, is easily derived from (1.8) by a limit argument letting α2α10\alpha_{2}-\alpha_{1}\to 0. (In the special cases when TT is simply a scaling by a constant or a rotation or a reflection, the statement even follows directly from (1.6).)

Refer to captionI. II.Refer to caption III.Refer to caption IV.Refer to caption V.Refer to caption VI.Refer to caption

Figure 1. Experimental graphs of the limiting density of normalized gaps, F(s)-F^{\prime}(s), for the vertex sets of Ammann-Beenker tilings 𝒫AB,𝒘{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}} with (I) 𝒘=𝟎{\text{\boldmath$w$}}=\mathbf{0} and (II) 𝒘=(0.9,0.3){\text{\boldmath$w$}}=(0.9,0.3); Gähler’s shield tilings 𝒫Gh,𝒘{\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}} with (III) 𝒘=(1.7,0.6){\text{\boldmath$w$}}=(1.7,0.6) and (IV) 𝒘=(0.1,0){\text{\boldmath$w$}}=(0.1,0); and Tübingen triangle tilings 𝒫TT,𝒘{\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},{\text{\boldmath$w$}}} with (V) 𝒘=(0.5,0.4){\text{\boldmath$w$}}=(0.5,0.4) and (VI) 𝒘=(1.3,0.4){\text{\boldmath$w$}}=(1.3,0.4) (notation as in Section 1.5). These graphs were obtained from numerical computations described in Section 5.4.
Remark 1.9.

Our notation differs from the notation used in [4], [27] and [20]: In the present paper we do not remove the points in 𝒫{\mathcal{P}} which are ’invisible’ from the origin; thus we allow ΔR\Delta_{R} to be a multi-set, and we allow equalities in the list in (1.5). Also, our function F(s)F(s) equals the function in the right hand side of [27, (1.15)], and is thus in general not the same as “F(s)F(s)” in [27, Cor. 3]. Clearly by (1.6), our function F(s)F(s) satisfies F(0)=1F(0)=1. In general FF has a jump discontinuity at 0; it follows from [27, Cor. 3] that lims0+F(s)=κ𝒫\lim_{s\to 0+}F(s)=\kappa_{{\mathcal{P}}} where κ𝒫\kappa_{{\mathcal{P}}} is the relative density of visible points in 𝒫{\mathcal{P}}, a quantity which was defined and proved to exist in [27]. Note that it is immediate to translate between the two versions of “F(s)F(s)”, except that it requires knowledge of the constant κ𝒫\kappa_{{\mathcal{P}}}, which is non-trivial to compute. See [20] for explicit formulas for κ𝒫\kappa_{{\mathcal{P}}} in special cases. See also the discussion at the end of Section 5.4 below regarding the effect of the different normalizations when comparing the graph in the top left panel of Figure 1 with the graphs in [4, Fig. 9] and [20, Fig. 2].

1.3. Main result; tail asymptotics

Our main goal in the present paper is to describe the tail asymptotics of the distribution function F(s)F(s) in Theorem 1, for a particular class of cut-and-project sets in 2\mathbb{R}^{2}, which includes several classical examples. We expect that the methods which we develop can be extended to more general cases as well (in particular see Remark 1.33 below). We now give the description of the class which we will consider. Let KK be a real quadratic field; let 𝒪K{\mathcal{O}}_{K} be its ring of integers; let σ\sigma be the unique non-trivial automorphism of KK, and let λ𝒪K×\lambda\in\mathcal{O}_{K}^{\times} be the fundamental unit (thus λ>1\lambda>1 and 𝒪K×={±λn:n}\mathcal{O}_{K}^{\times}=\left\{\pm\lambda^{n}\>:\>n\in\mathbb{Z}\right\}). Let K{\mathcal{L}}_{K} be the Minkowski embedding of 𝒪K2{\mathcal{O}}_{K}^{2} in 4\mathbb{R}^{4}, viz.,

(1.10) K:={(α,β,σ(α),σ(β))4:(α,β)𝒪K2}.\displaystyle{\mathcal{L}}_{K}:=\left\{(\alpha,\beta,\sigma(\alpha),\sigma(\beta))\in\mathbb{R}^{4}\>:\>(\alpha,\beta)\in\mathcal{O}_{K}^{2}\right\}.

We set d=m=2d=m=2, that is, we view 4\mathbb{R}^{4} as the product of a 2-dimensional physical space and a 2-dimensional internal space, and π\pi and πint\pi_{\operatorname{int}} denote the orthogonal projection of 4\mathbb{R}^{4} onto the first 2 coordinates and the last 2 coordinates, respectively. Note that in this case πint(K)\pi_{\operatorname{int}}({\mathcal{L}}_{K}) is dense in 2\mathbb{R}^{2}, i.e. we have 𝒜=2{\mathcal{A}}=\mathbb{R}^{2}, and so we take the window 𝒲{\mathcal{W}} to be a bounded subset of 2\mathbb{R}^{2} with nonempty interior. (In the present case the restriction of π\pi to K{\mathcal{L}}_{K} is injective, and so the property (1.3) is automatically fulfilled.) We will always assume that 𝒫:=𝒫(𝒲,K){\mathcal{P}}:={\mathcal{P}}({\mathcal{W}},{\mathcal{L}}_{K}) is regular, i.e. that 𝒲\partial{\mathcal{W}} has Lebesgue measure zero.

For this class of cut-and-project sets, the formula for the asymptotic density of 𝒫{\mathcal{P}}, [26, (1.7)], becomes (see also Section 2 below):

(1.11) c𝒫=ΔK1Area(𝒲),\displaystyle c_{\mathcal{P}}=\Delta_{K}^{-1}\operatorname{Area}({\mathcal{W}}),

where ΔK\Delta_{K} is the discriminant of KK.

As we prove in Remark 3.29 below, the limiting gap distribution function is unchanged if 𝒲\mathcal{W} is modified by any measure zero set. In particular, without loss of generality we may assume 𝒲\mathcal{W} is open (indeed, otherwise replace 𝒲\mathcal{W} by its interior). Finally, in the present paper we will make the key assumption that 𝟎𝒲\mathbf{0}\in{\mathcal{W}}.

In our approach we actually work with the integral of the distribution function FF, i.e. with

(1.12) G(s):=sF(t)dt.\displaystyle G(s):=\int_{s}^{\infty}F(t)\,\text{d}t.

As we will note below (see Lemma 3.2), using the fact that FF is decreasing, an easy interpolation argument allows us to deduce an asymptotic formula for FF once we know an asymptotic formula for GG.

Our main result is the following:

Theorem 2.

Let 𝒫=𝒫(𝒲,K){\mathcal{P}}={\mathcal{P}}({\mathcal{W}},{\mathcal{L}}_{K}) where KK is a real quadratic field and 𝒲{\mathcal{W}} is a bounded open subset of 2\mathbb{R}^{2} such that 𝟎𝒲\mathbf{0}\in{\mathcal{W}} and 𝒲\partial{\mathcal{W}} has Lebesgue measure zero. Let F(s)F(s) be the associated limiting gap distribution function as in Theorem 1, and let G(s)=sF(t)dtG(s)=\int_{s}^{\infty}F(t)\,\textup{d}t. Then there exists a positive constant a𝒫a_{\mathcal{P}} such that

(1.13) G(s)a𝒫s1andF(s)a𝒫s2as s.\displaystyle G(s)\sim a_{{\mathcal{P}}}s^{-1}\quad\text{and}\quad F(s)\sim a_{{\mathcal{P}}}s^{-2}\qquad\text{as }\>s\to\infty.

If we further assume 𝒲\mathcal{W} to be convex, then we have

(1.14) G(s)=a𝒫s1+O(s2logs)andF(s)=a𝒫s2+O(s52logs)as s.\displaystyle G(s)=a_{{\mathcal{P}}}s^{-1}+O\bigl{(}s^{-2}\log s\bigr{)}\quad\text{and}\quad F(s)=a_{{\mathcal{P}}}s^{-2}+O\Bigl{(}s^{-\frac{5}{2}}\sqrt{\log s}\Bigr{)}\qquad\text{as }\>s\to\infty.
Remark 1.15.

As mentioned above, in order to prove Theorem 2, it suffices to prove the two asymptotic estimates for G(s)G(s) in (1.13) and (1.14). Both these estimates will in fact follow from a more precise, general asymptotic formula for G(s)G(s) with an explicit error term; see (3.7) below. We also mention that our analysis leads to an explicit formula for the leading coefficient a𝒫a_{\mathcal{P}}; see (3.8).

1.4. Formula for a𝒫a_{\mathcal{P}}

As we will now describe, after imposing one more assumption on the window set, we can further evaluate the aforementioned formula for a𝒫a_{\mathcal{P}}. To state our result, we need to introduce some more notation. For θ\theta\in\mathbb{R} we write kθ:=(cosθsinθsinθcosθ)SL2()\text{k}_{\theta}:=\left(\begin{smallmatrix}\cos\theta&-\sin\theta\\ \sin\theta&\cos\theta\end{smallmatrix}\right)\in\operatorname{SL}_{2}(\mathbb{R}). Then for any subset 𝒲2\mathcal{W}\subset\mathbb{R}^{2} we let 𝒲(θ)\ell_{\mathcal{W}}(\theta) be the projection of 𝒲kθ\mathcal{W}\text{k}_{-\theta} on the xx-axis. On top of the assumptions in Theorem 2, we will now assume that for each θ\theta, 𝒲(θ)\ell_{\mathcal{W}}(\theta) is an interval. Note that this assumption is always satisfied if 𝒲{\mathcal{W}} is connected, but it also holds for many non-connected window sets 𝒲{\mathcal{W}}.

Note that since 𝒲\mathcal{W} is open and 𝟎𝒲\mathbf{0}\in\mathcal{W}, 𝒲(θ)\ell_{\mathcal{W}}(\theta) is an open subset of \mathbb{R} containing 0. Hence if 𝒲(θ)\ell_{\mathcal{W}}(\theta) is assumed to be an interval, then we can parametrize it by two positive numbers r(θ),ν(θ)r(\theta),\nu(\theta) via

(1.16) 𝒲(θ)=r(θ)(ν(θ),1).\displaystyle\ell_{\mathcal{W}}(\theta)=r(\theta)(-\nu(\theta),1).
Theorem 3.

Retain the notation and assumptions in Theorem 2 and further assume that KK is of class number one, and that 𝒲(θ)\ell_{\mathcal{W}}(\theta) is an interval for each θ\theta. Then there exist a finite partition >0=j=1lSj\mathbb{R}_{>0}=\bigsqcup_{j=1}^{\,l}S_{j} of >0\mathbb{R}_{>0} into intervals, and non-negative constants Aj,BjA_{j},B_{j} (1jl)(1\leq j\leq l) depending only on KK, such that

(1.17) a𝒫=Area(𝒲)4ΔK2ζK(2)j=1lS~jr(θ)2(Aj+Bjν(θ)2)dθ,\displaystyle a_{{\mathcal{P}}}=\frac{\operatorname{Area}({\mathcal{W}})}{4\Delta_{K}^{2}\zeta_{K}(2)}\sum_{j=1}^{l}\int_{\tilde{S}_{j}}r(\theta)^{-2}\left(A_{j}+B_{j}\nu(\theta)^{-2}\right)\,\mathrm{d}\theta,

where ζK\zeta_{K} is the Dedekind zeta function attached to KK, S~j:={θ[0,2π):ν(θ)Sj}\tilde{S}_{j}:=\{\theta\in[0,2\pi)\>:\>\nu(\theta)\in S_{j}\}, and r(θ)r(\theta) and ν(θ)\nu(\theta) are defined by (1.16).

Remark 1.18.

We stress that the intervals SjS_{j} are allowed to be open, closed, or half-open, and may be degenerate, i.e. of the form [a,a]={a}[a,a]=\{a\} for some a>0a>0.

Remark 1.19.

Our analysis applies for any real quadratic field KK and the class number one assumption in Theorem 3 is only for simplicity of presentation; see Theorem 5 in Section 4 for the most general version. We also mention that the partition >0=j=1lSj\mathbb{R}_{>0}=\bigsqcup_{j=1}^{l}S_{j} and the constants Aj,BjA_{j},B_{j} in the above theorem are all computable and we will illustrate it in the next section for three well-known classes of quasicrystals, namely, the vertex sets of the Ammann-Beenker, Gähler’s shield, and Tübingen triangle tilings.

Remark 1.20.

Among the sets S~j\tilde{S}_{j}, there is one that is always non-empty, namely, the unique S~i\tilde{S}_{i} such that 1Si1\in S_{i}. In certain cases, this S~i\tilde{S}_{i} equals the whole interval [0,2π)[0,2\pi) while all other S~j\tilde{S}_{j}’s are empty. In this case, the formula for a𝒫a_{\mathcal{P}} can be simplified into

a𝒫=CKArea(𝒲)Area(𝒲),\displaystyle a_{\mathcal{P}}=C_{K}\operatorname{Area}(\mathcal{W})\operatorname{Area}(\mathcal{W}^{*}),

where CKC_{K} is some positive constant depending only on KK, and 𝒲{\mathcal{W}}^{*} is the polar set of 𝒲\mathcal{W}, i.e.

(1.21) 𝒲:={𝒛2:𝒛𝒘1𝒘𝒲}.\displaystyle{\mathcal{W}}^{*}:=\{{\text{\boldmath$z$}}\in\mathbb{R}^{2}\>:\>{\text{\boldmath$z$}}\cdot{\text{\boldmath$w$}}\leq 1\>\>\forall\>{\text{\boldmath$w$}}\in{\mathcal{W}}\}.

Here \cdot is the standard scalar product in 2\mathbb{R}^{2}. See Section 4.1 for more details. If 𝒲\mathcal{W} is also convex and centrally symmetric, the product Area(𝒲)Area(𝒲)\operatorname{Area}(\mathcal{W})\operatorname{Area}(\mathcal{W}^{*}) appearing in the above formula is known as the Mahler volume of 𝒲\mathcal{W}.

Remark 1.22.

For any 𝒲\mathcal{W} as in Theorem 3, replacing 𝒲\mathcal{W} by its convex hull does not affect the intervals 𝒲(θ)\ell_{\mathcal{W}}(\theta), hence does not affect the functions r(θ)r(\theta) and ν(θ)\nu(\theta). Therefore, in the explicit formula (1.17), only the factor Area(𝒲)\operatorname{Area}(\mathcal{W}) is affected when replacing 𝒲\mathcal{W} by its convex hull. In this connection it should also be noted that the polar set of the convex hull of 𝒲\mathcal{W} equals the polar set of 𝒲\mathcal{W}.

1.5. Examples

We next illustrate how our results apply in three cases of well-known planar quasicrystals. The three propositions below are all proved in Section 4.4. More details, and comparison with numerics, is provided in Section 5.

Refer to caption    Refer to caption

Refer to caption    Refer to caption

Figure 2. Patches of an Ammann-Beenker tiling with vertex set 𝒫AB,(0.8,0.1){\mathcal{P}}_{\operatorname{AB},(0.8,0.1)} (top left panel), a Gähler’s shield tiling with vertex set 𝒫Gh,(1.7,0.6){\mathcal{P}}_{\operatorname{Gh},(1.7,0.6)} (top right panel), and Tübingen triangle tilings with vertex sets 𝒫TT,(0.001,0){\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},(0.001,0)} (bottom left panel) and 𝒫TT,(1.3,0.4){\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},(1.3,0.4)} (bottom right panel). In each patch, the origin is marked by a small black disc.

We first consider the Ammann-Beenker tiling, which was discovered by Robert Ammann in the 70s and first described in [17] and [1]. Specifically, we consider the “A5 set, variant (b)”, in the notation of [1]; see Figure 2 (top left panel) above for a small patch of this tiling. It is well-known that the set of vertices of an arbitrary Ammann-Beenker tiling can be generated using the cut-and-project construction. Specifically, let K=(2)K=\mathbb{Q}(\sqrt{2}), and let 𝒲AB2{\mathcal{W}}_{\operatorname{AB}}\subset\mathbb{R}^{2} be the open regular octagon centered at the origin of edge length 11, oriented so that four of the edges are perpendicular to a coordinate axis. For each 𝒘2{\text{\boldmath$w$}}\in\mathbb{R}^{2} we set

(1.23) 𝒲𝒘(AB):=(𝒲AB+𝒘)g12and𝒫AB,𝒘:=𝒫(𝒲𝒘(AB),K)g2,\displaystyle{\mathcal{W}}_{{\text{\boldmath$w$}}}^{(\operatorname{AB})}:=({\mathcal{W}}_{\operatorname{AB}}+{\text{\boldmath$w$}})g_{1}\subset\mathbb{R}^{2}\qquad\text{and}\qquad{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}}:={\mathcal{P}}({\mathcal{W}}^{(\operatorname{AB})}_{{\text{\boldmath$w$}}},{\mathcal{L}}_{K})g_{2},

where g1:=(1012)g_{1}:=\left(\begin{smallmatrix}1&0\\ 1&\sqrt{2}\end{smallmatrix}\right) and g2:=(101/21/2)g_{2}:=\left(\begin{smallmatrix}1&0\\ 1/\sqrt{2}\>&1/\sqrt{2}\end{smallmatrix}\right). Then for any 𝒘𝒲AB{\text{\boldmath$w$}}\in{\mathcal{W}}_{\operatorname{AB}} with the property that πint(K)𝒲𝒘(AB)=\pi_{\operatorname{int}}({\mathcal{L}}_{K})\cap\partial{\mathcal{W}}^{(\operatorname{AB})}_{{\text{\boldmath$w$}}}=\emptyset, the cut-and-project set 𝒫AB,𝒘{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}} is the vertex set of an Ammann-Beenker tiling with one vertex at 𝟎\mathbf{0}, and conversely, the vertex set of any Ammann-Beenker tiling having a vertex at 𝟎\mathbf{0} is (up to scaling and rotation) either equal to such a point set 𝒫AB,𝒘{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}}, or can be obtained as a limit of such point sets 𝒫AB,𝒘{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}} in an appropriate topology. (See [5, Ch. 7.3] and Section 5.1 below.)

Note that because of (1.23) and the last statement in Theorem 1, the formula (1.14) in Theorem 2 applies to the cut-and-project set 𝒫AB,𝒘{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}} for any 𝒘𝒲AB{\text{\boldmath$w$}}\in{\mathcal{W}}_{\operatorname{AB}}. Here the constant a𝒫a_{{\mathcal{P}}} is given by a quite simple formula, also valid for much more general window sets:

Proposition 1.1.

Let 𝒫=𝒫(𝒲,K){\mathcal{P}}={\mathcal{P}}({\mathcal{W}},{\mathcal{L}}_{K}) with K=(2)K=\mathbb{Q}(\sqrt{2}) and with 𝒲{\mathcal{W}} as in Theorem 3, and let 𝒲(θ)\ell_{{\mathcal{W}}}(\theta) be parametrized as in (1.16). Then

a𝒫=3216π4Area(𝒲)((7+82)S~1r(θ)2dθ+(4+62)S~2r(θ)2dθ\displaystyle a_{\mathcal{P}}=\frac{3\sqrt{2}}{16\pi^{4}}\operatorname{Area}(\mathcal{W})\biggl{(}(7+8\sqrt{2})\int_{\tilde{S}_{1}}r(\theta)^{-2}\,\mathrm{d}\theta+(4+6\sqrt{2})\int_{\tilde{S}_{2}}r(\theta)^{-2}\,\mathrm{d}\theta\hskip 100.0pt
(1.24) +(2+22)S~3r(θ)2dθ+S~4r(θ)2dθ),\displaystyle+(2+2\sqrt{2})\int_{\tilde{S}_{3}}r(\theta)^{-2}\,\mathrm{d}\theta+\int_{\tilde{S}_{4}}r(\theta)^{-2}\,\mathrm{d}\theta\biggr{)},

where S~j:={θ[0,2π):ν(θ)Sj}\tilde{S}_{j}:=\{\theta\in[0,2\pi)\>:\>\nu(\theta)\in S_{j}\}, with

S1=(0,21],S2=(21,122),S3=[122,2]andS4=(2,1+2).\displaystyle S_{1}=\bigl{(}0,\sqrt{2}-1\bigr{]},\quad S_{2}=\bigl{(}\sqrt{2}-1,\tfrac{1}{2}\sqrt{2}\bigr{)},\quad S_{3}=\bigl{[}\tfrac{1}{2}\sqrt{2},\sqrt{2}\bigr{]}\quad\text{and}\quad S_{4}=\bigl{(}\sqrt{2},1+\sqrt{2}\bigr{)}.

When S~3=[0,2π)\tilde{S}_{3}=[0,2\pi), the above formula simplifies to

(1.25) a𝒫\displaystyle a_{\mathcal{P}} =32(1+2)4π4Area(𝒲)Area(𝒲).\displaystyle=\frac{3\sqrt{2}(1+\sqrt{2})}{4\pi^{4}}\operatorname{Area}({\mathcal{W}})\operatorname{Area}({\mathcal{W}}^{*}).

Note that Proposition 1.1 applies when 𝒲=𝒲𝒘(AB){\mathcal{W}}={\mathcal{W}}^{(\operatorname{AB})}_{{\text{\boldmath$w$}}} for any 𝒘𝒲AB{\text{\boldmath$w$}}\in{\mathcal{W}}_{\operatorname{AB}}, and the formula (1.25) holds whenever 𝒘w lies sufficiently near 𝟎\mathbf{0}. In particular, (1.25) implies that

(1.26) a𝒫AB,𝟎=24/π4=0.24638.\displaystyle a_{{\mathcal{P}}_{\operatorname{AB},\mathbf{0}}}=24/\pi^{4}=0.24638\ldots.

In Section 5.4, we present a comparison, for a few examples of points 𝒘𝒲AB{\text{\boldmath$w$}}\in{\mathcal{W}}_{\operatorname{AB}}, between the exact values of a𝒫a_{{\mathcal{P}}} computed using Proposition 1.1, and numerically computed approximate values of s2F(s)s^{2}F(s) for large ss.

Next we consider Gähler’s shield tiling, which was discovered in [18]; these tilings are built up of a certain triple of tiles (a triangle, a square, and a hexagon called a ’shield’), equipped with local matching rules. The vertex set of a Gähler’s shield tiling can be obtained using the cut-and-project construction in the following way: Let K=(3)K=\mathbb{Q}(\sqrt{3}), and let 𝒲Gh2{\mathcal{W}}_{\operatorname{Gh}}\subset\mathbb{R}^{2} be the open regular dodecagon centered at the origin of edge length 11, so that four of the edges are perpendicular to a coordinate axis. For each 𝒘2{\text{\boldmath$w$}}\in\mathbb{R}^{2} we set

(1.27) 𝒲𝒘(Gh):=(𝒲Gh+𝒘)g12and𝒫Gh,𝒘:=𝒫(𝒲𝒘(Gh),K)g2,\displaystyle{\mathcal{W}}_{{\text{\boldmath$w$}}}^{(\operatorname{Gh})}:=({\mathcal{W}}_{\operatorname{Gh}}+{\text{\boldmath$w$}})g_{1}\subset\mathbb{R}^{2}\qquad\text{and}\qquad{\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}}:={\mathcal{P}}({\mathcal{W}}_{{\text{\boldmath$w$}}}^{(\operatorname{Gh})},{\mathcal{L}}_{K})g_{2},

where g1:=(1032)g_{1}:=\left(\begin{smallmatrix}1&0\\ \sqrt{3}&2\end{smallmatrix}\right) and g2:=(103/21/2)g_{2}:=\left(\begin{smallmatrix}1&0\\ \sqrt{3}/2&1/2\end{smallmatrix}\right). Then for any 𝒘𝒲Gh{\text{\boldmath$w$}}\in{\mathcal{W}}_{\operatorname{Gh}} with the property that πint(K)𝒲𝒘(Gh)=\pi_{\operatorname{int}}({\mathcal{L}}_{K})\cap\partial{\mathcal{W}}^{(\operatorname{Gh})}_{{\text{\boldmath$w$}}}=\emptyset, the cut-and-project set 𝒫Gh,𝒘{\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}} is the vertex set of a Gähler’s shield tiling with one vertex at 𝟎\mathbf{0}. (See [5, Ch. 7.3] and Section 5.2 below.)

Again, (1.14) in Theorem 2 applies to 𝒫Gh,𝒘{\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}} for any 𝒘𝒲Gh{\text{\boldmath$w$}}\in{\mathcal{W}}_{\operatorname{Gh}}, and regarding a𝒫a_{{\mathcal{P}}} we have:

Proposition 1.2.

Let 𝒫=𝒫(𝒲,K){\mathcal{P}}={\mathcal{P}}({\mathcal{W}},{\mathcal{L}}_{K}) with K=(3)K=\mathbb{Q}(\sqrt{3}) and with 𝒲{\mathcal{W}} as in Theorem 3, and let 𝒲(θ)\ell_{{\mathcal{W}}}(\theta) be parametrized as in (1.16). Then

a𝒫=316π4Area(𝒲)((17+163)S~1r(θ)2dθ+(9+123)S~2r(θ)2dθ\displaystyle a_{\mathcal{P}}=\frac{\sqrt{3}}{16\pi^{4}}\operatorname{Area}(\mathcal{W})\biggl{(}(17+16\sqrt{3})\int_{\tilde{S}_{1}}r(\theta)^{-2}\,\mathrm{d}\theta+(9+12\sqrt{3})\int_{\tilde{S}_{2}}r(\theta)^{-2}\,\mathrm{d}\theta\hskip 120.0pt
(1.28) +(6+63)S~3r(θ)2dθ+(3+43)S~4r(θ)2dθ+(3+23)S~5r(θ)2dθ+2S~6r(θ)2dθ),\displaystyle+(6+6\sqrt{3})\int_{\tilde{S}_{3}}r(\theta)^{-2}\,\mathrm{d}\theta+(3+4\sqrt{3})\int_{\tilde{S}_{4}}r(\theta)^{-2}\,\mathrm{d}\theta+(3+2\sqrt{3})\int_{\tilde{S}_{5}}r(\theta)^{-2}\,\mathrm{d}\theta+2\int_{\tilde{S}_{6}}r(\theta)^{-2}\,\mathrm{d}\theta\biggr{)},

where S~j:={θ[0,2π):ν(θ)Sj}\tilde{S}_{j}:=\{\theta\in[0,2\pi)\>:\>\nu(\theta)\in S_{j}\}, with S1=(0,11+3]S_{1}=\bigl{(}0,\tfrac{1}{1+\sqrt{3}}\bigr{]}, S2=(11+3,13)S_{2}=\bigl{(}\tfrac{1}{1+\sqrt{3}},\tfrac{1}{\sqrt{3}}\bigr{)}, S3=[13,31]S_{3}=\bigl{[}\tfrac{1}{\sqrt{3}},\sqrt{3}-1\bigr{]}, S4=(31,131)S_{4}=\bigl{(}\sqrt{3}-1,\tfrac{1}{\sqrt{3}-1}\bigr{)}, S5=[131,3]S_{5}=\bigl{[}\tfrac{1}{\sqrt{3}-1},\sqrt{3}\bigr{]} and S6=(3,1+3)S_{6}=\bigl{(}\sqrt{3},1+\sqrt{3}\bigr{)}. When S~4=[0,2π)\tilde{S}_{4}=[0,2\pi), the above formula simplifies to

(1.29) a𝒫\displaystyle a_{\mathcal{P}} =12+338π4Area(𝒲)Area(𝒲).\displaystyle=\frac{12+3\sqrt{3}}{8\pi^{4}}\operatorname{Area}({\mathcal{W}})\operatorname{Area}({\mathcal{W}}^{*}).

Note that Proposition 1.2 applies when 𝒲=𝒲𝒘(Gh){\mathcal{W}}={\mathcal{W}}_{{\text{\boldmath$w$}}}^{(\operatorname{Gh})} for any 𝒘𝒲Gh{\text{\boldmath$w$}}\in{\mathcal{W}}_{\operatorname{Gh}}, and the formula (1.29) holds whenever 𝒘w lies sufficiently near 𝟎\mathbf{0}. Again see Section 5.4 for numerical computations related to the values of a𝒫a_{{\mathcal{P}}} for 𝒫=𝒫Gh,𝒘{\mathcal{P}}={\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}}.

Finally, we consider the Tübingen triangle tiling, which was discovered and studied in [6]; these tilings are built up of a certain pair of isosceles triangles. The set of vertices of a Tübingen triangle tiling can be obtained using the cut-and-project construction in the following way. Let K=(5)K=\mathbb{Q}(\sqrt{5}) and τ=12(1+5)\tau=\frac{1}{2}(1+\sqrt{5}), and let 𝒲TT2{\mathcal{W}}_{\operatorname{\hskip 1.0ptTT}}\subset\mathbb{R}^{2} be the open regular decagon centered at the origin of edge length (2+τ)/5\sqrt{(2+\tau)/5}, oriented so that two of the edges are perpendicular to the first coordinate axis. For each 𝒘2{\text{\boldmath$w$}}\in\mathbb{R}^{2} we set

(1.30) 𝒲𝒘(TT):=(𝒲TT+𝒘)g12and𝒫TT,𝒘:=𝒫(𝒲𝒘(TT),K)g2,\displaystyle{\mathcal{W}}_{{\text{\boldmath$w$}}}^{(\operatorname{\hskip 1.0ptTT})}:=({\mathcal{W}}_{\operatorname{\hskip 1.0ptTT}}+{\text{\boldmath$w$}})g_{1}\subset\mathbb{R}^{2}\qquad\text{and}\qquad{\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},{\text{\boldmath$w$}}}:={\mathcal{P}}({\mathcal{W}}^{(\operatorname{\hskip 1.0ptTT})}_{{\text{\boldmath$w$}}},{\mathcal{L}}_{K})g_{2},

where g1:=(100(2+τ)/5)(10τ2)g_{1}:=\left(\begin{smallmatrix}1&0\\ 0&\sqrt{(2+\tau)/5}\end{smallmatrix}\right)\left(\begin{smallmatrix}1&0\\ \tau&2\end{smallmatrix}\right) and g2:=(10(τ1)/22+τ/2)g_{2}:=\left(\begin{smallmatrix}1&0\\ (\tau-1)/2\>&\sqrt{2+\tau}/2\end{smallmatrix}\right). Then for any 𝒘𝒲TT{\text{\boldmath$w$}}\in{\mathcal{W}}_{\operatorname{\hskip 1.0ptTT}} with the property that 15πint(K)𝒲𝒘(TT)=\frac{1}{5}\pi_{\operatorname{int}}({\mathcal{L}}_{K})\cap\partial{\mathcal{W}}^{(\operatorname{\hskip 1.0ptTT})}_{{\text{\boldmath$w$}}}=\emptyset, the cut-and-project set 𝒫TT,𝒘{\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},{\text{\boldmath$w$}}} is the vertex set of a Tübingen triangle tiling with one vertex at 𝟎\mathbf{0}. (See [5, Ch. 7.3], [6, Sec. 4] and Section 5.3 below.)

Again, (1.14) in Theorem 2 applies to 𝒫TT,𝒘{\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},{\text{\boldmath$w$}}} for any 𝒘𝒲TT{\text{\boldmath$w$}}\in{\mathcal{W}}_{\operatorname{\hskip 1.0ptTT}}, and regarding a𝒫a_{{\mathcal{P}}} we have:

Proposition 1.3.

Let 𝒫=𝒫(𝒲,K){\mathcal{P}}={\mathcal{P}}({\mathcal{W}},{\mathcal{L}}_{K}) with K=(5)K=\mathbb{Q}(\sqrt{5}) and with 𝒲{\mathcal{W}} as in Theorem 3, and let 𝒲(θ)\ell_{{\mathcal{W}}}(\theta) be parametrized as in (1.16). Then

(1.31) a𝒫=3516π4Area(𝒲)(\displaystyle a_{\mathcal{P}}=\frac{3\sqrt{5}}{16\pi^{4}}\operatorname{Area}(\mathcal{W})\biggl{(} (5+35)S~1r(θ)2dθ+(1+5)S~2r(θ)2dθ),\displaystyle(5+3\sqrt{5})\int_{\tilde{S}_{1}}r(\theta)^{-2}\,\mathrm{d}\theta+(1+\sqrt{5})\int_{\tilde{S}_{2}}r(\theta)^{-2}\,\mathrm{d}\theta\biggr{)},

where S~j:={θ[0,2π):ν(θ)Sj}\tilde{S}_{j}:=\{\theta\in[0,2\pi)\>:\>\nu(\theta)\in S_{j}\}, with S1=(0,512]S_{1}=(0,\frac{\sqrt{5}-1}{2}] and S2=(512,1+52)S_{2}=(\frac{\sqrt{5}-1}{2},\frac{1+\sqrt{5}}{2}). When S~2=[0,2π)\tilde{S}_{2}=[0,2\pi), the above formula simplifies to

(1.32) a𝒫\displaystyle a_{\mathcal{P}} =3(5+5)8π4Area(𝒲)Area(𝒲).\displaystyle=\frac{3(5+\sqrt{5})}{8\pi^{4}}\operatorname{Area}({\mathcal{W}})\operatorname{Area}({\mathcal{W}}^{*}).

Proposition 1.3 applies when 𝒲=𝒲𝒘(TT){\mathcal{W}}={\mathcal{W}}_{{\text{\boldmath$w$}}}^{(\operatorname{\hskip 1.0ptTT})} for any 𝒘𝒲TT{\text{\boldmath$w$}}\in{\mathcal{W}}_{\operatorname{\hskip 1.0ptTT}}, and the formula (1.32) holds whenever 𝒘w lies sufficiently near 𝟎\mathbf{0}. Again see Section 5.4 for related numerical computations.

Remark 1.33.

As a concluding remark of the introduction, we mention that our main result, Theorem 2, does not apply to the vertex sets of the classical rhombic Penrose tilings, as these can only be realized as unions of (four) translates of the cut-and-project type sets considered in Theorem 2; see [26, Sec. 2.5]. Nevertheless, in preliminary work, using similar methods as in the present paper we have proved analogous tail asymptotic formulas for the limiting gap distribution function for point sets in this generality, thus in particular covering the rhombic Penrose tilings.

2. Preliminaries

2.1. Real quadratic fields

In this section we give a brief review on backgrounds on real quadratic fields. Let K=(d)K=\mathbb{Q}(\sqrt{d}) be a real quadratic field with dd\in\mathbb{N} square-free. Let 𝒪K=[τ]K\mathcal{O}_{K}=\mathbb{Z}[\tau]\subset K be its ring of integers, where

τ={1+d2d1(mod 4),dd2,3(mod 4).\displaystyle\tau=\left\{\begin{array}[]{ll}\frac{1+\sqrt{d}}{2}&d\equiv 1\ (\mathrm{mod}\ 4),\\[4.0pt] \sqrt{d}&d\equiv 2,3\ (\mathrm{mod}\ 4).\end{array}\right.

Let JKJ_{K} be the (abelian) group of fractional ideals of 𝒪K\mathcal{O}_{K}, and let PKP_{K} be the subgroup of principal ideals. We let CK=JK/PKC_{K}=J_{K}/P_{K} be the ideal class group of KK, and we call elements in CKC_{K} ideal classes of KK. For any nonzero a1,,amKa_{1},\ldots,a_{m}\in K, we denote by a1,,am\langle a_{1},\ldots,a_{m}\rangle the fractional ideal generated by a1,,ama_{1},\ldots,a_{m}.

Let σ\sigma be the unique non-trivial automorphism of KK and consider the embedding

(2.1) ι:K2,k(k,σ(k)).\displaystyle\iota:K\to\mathbb{R}^{2},\qquad k\mapsto(k,\sigma(k)).

We note that ι(I)\iota(I) is a lattice in 2\mathbb{R}^{2} for any IJKI\in J_{K}. In particular, ι(𝒪K)\iota(\mathcal{O}_{K}) is the lattice generated by (1,1)(1,1) and (τ,σ(τ))(\tau,\sigma(\tau)), which is of covolume ΔK1/2\Delta_{K}^{1/2}, where

(2.4) ΔK:={dd1(mod 4),4dd2,3(mod 4),\displaystyle\Delta_{K}:=\left\{\begin{array}[]{ll}d&d\equiv 1\ (\mathrm{mod}\ 4),\\ 4d&d\equiv 2,3\ (\mathrm{mod}\ 4),\end{array}\right.

is the discriminant of KK. For any fractional ideal IJKI\in J_{K}, its absolute norm is defined by

(2.5) Nr(I):=covol(ι(I))covol(ι(𝒪K))=ΔK1/2covol(ι(I)).\displaystyle\text{Nr}(I):=\frac{\operatorname{covol}(\iota(I))}{\operatorname{covol}(\iota(\mathcal{O}_{K}))}=\Delta_{K}^{-1/2}\operatorname{covol}(\iota(I)).

Note that Nr is multiplicative in JKJ_{K}, that is, Nr(I1I2)=Nr(I1)Nr(I2)\text{Nr}(I_{1}I_{2})=\text{Nr}(I_{1})\text{Nr}(I_{2}) for any I1,I2JKI_{1},I_{2}\in J_{K}. Moreover, for any αK{0}\alpha\in K\setminus\{0\}, we have Nr(αI)=|N(α)|Nr(I)\text{Nr}(\alpha I)=|\mathrm{N}(\alpha)|\text{Nr}(I), where N:K\mathrm{N}:K\to\mathbb{R} is the standard norm on KK given by N(α):=ασ(α)\mathrm{N}(\alpha):=\alpha\sigma(\alpha).

2.2. Hilbert modular group

Let \mathbb{H} be the upper half plane. The group SL2()\operatorname{SL}_{2}(\mathbb{R}) acts on \mathbb{H} via the Möbius transformation: gz=az+bcz+dg\cdot z=\frac{az+b}{cz+d} for any g=(abcd)SL2()g=\left(\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}\right)\in\operatorname{SL}_{2}(\mathbb{R}) and zz\in\mathbb{H}. Let H=SL2()×SL2()\mathrm{H}=\operatorname{SL}_{2}(\mathbb{R})\times\operatorname{SL}_{2}(\mathbb{R}) (which we view as a subgroup of SL4()\operatorname{SL}_{4}(\mathbb{R}) via the block diagonal embedding) and let 2\mathbb{H}^{2} be the product of two upper half planes. The group H\mathrm{H} naturally acts on 2\mathbb{H}^{2} via

(h1,h2)(z1,z2):=(h1z2,h2z2),h1,h2SL2(),z1,z2.\displaystyle(h_{1},h_{2})(z_{1},z_{2}):=(h_{1}\cdot z_{2},h_{2}\cdot z_{2}),\qquad\forall\,h_{1},h_{2}\in\operatorname{SL}_{2}(\mathbb{R}),\,z_{1},z_{2}\in\mathbb{H}.

Denote by ¯:={}\overline{\mathbb{R}}:=\mathbb{R}\cup\{\infty\} which can be identified with the boundary of \mathbb{H}. The action of H\mathrm{H} on 2\mathbb{H}^{2} naturally extends to ¯2\overline{\mathbb{R}}^{2} via the same formula. We also denote by K¯:=K{}\overline{K}:=K\cup\{\infty\}. It embeds into ¯2\overline{\mathbb{R}}^{2} via the natural extension of the embedding ι:K2\iota:K\to\mathbb{R}^{2} (by sending K¯\infty\in\overline{K} to (,)¯2(\infty,\infty)\in\overline{\mathbb{R}}^{2}). With slight abuse of notation, we also denote this embedding from K¯\overline{K} to ¯2\overline{\mathbb{R}}^{2} by ι\iota. We will also write ι\iota for the group homomorphism SL2(K)H\operatorname{SL}_{2}(K)\to\mathrm{H} given by

(2.6) ι((abcd))=((abcd),(σ(a)σ(b)σ(c)σ(d))).\displaystyle\iota\left(\left(\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}\right)\right)=\left(\left(\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}\right),\left(\begin{smallmatrix}\sigma(a)&\sigma(b)\\ \sigma(c)&\sigma(d)\end{smallmatrix}\right)\right).

The Hilbert modular group for the field KK is given by

SL2(𝒪K):={(abcd)SL2(K):a,b,c,d𝒪K}.\displaystyle\operatorname{SL}_{2}({\mathcal{O}}_{K}):=\left\{\left(\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}\right)\in\operatorname{SL}_{2}(K)\>:\>a,b,c,d\in\mathcal{O}_{K}\right\}.

We write ΓK\Gamma_{K} for the embedding of SL2(𝒪K)\operatorname{SL}_{2}({\mathcal{O}}_{K}) in H\mathrm{H}:

ΓK:=ι(SL2(𝒪K)).\displaystyle\Gamma_{K}:=\iota\bigl{(}\operatorname{SL}_{2}({\mathcal{O}}_{K})\bigr{)}.

The discreteness of ι(𝒪K)\iota(\mathcal{O}_{K}) in 2\mathbb{R}^{2} implies that ΓK\Gamma_{K} is a discrete subgroup of H\mathrm{H}. Indeed, it is a non-uniform lattice in H\mathrm{H}, that is, the homogeneous space ΓK\H\Gamma_{K}\backslash\mathrm{H} is non-compact and has finite volume with respect to a Haar measure of H\mathrm{H}. As a subgroup of H\mathrm{H}, ΓK\Gamma_{K} naturally acts on ¯2\overline{\mathbb{R}}^{2}, and this action preserves ι(K¯)\iota\bigl{(}\overline{K}\bigr{)}. The orbits of ι(K¯)\iota\bigl{(}\overline{K}\bigr{)} under ΓK\Gamma_{K} are called the cusps of ΓK\Gamma_{K}. We will often represent a cusp by an element in the corresponding ΓK\Gamma_{K}-orbit, or by an element in K¯\overline{K} (via the natural identification between ι(K¯)\iota(\overline{K}) and K¯\overline{K}). The following lemma, together with the obvious identification between ι(K¯)/ΓK\iota(\overline{K})/\Gamma_{K} and K¯/SL2(𝒪K)\overline{K}/\operatorname{SL}_{2}(\mathcal{O}_{K}), shows that the number of cusps of ΓK\Gamma_{K} equals the number of ideal classes of KK.

Lemma 2.1 ([16, Lemma 3.5]).

The map sending kK¯k\in\overline{K} to [k,1]CK[\langle k,1\rangle]\in C_{K} induces a bijection between K¯/SL2(𝒪K)\overline{K}/\operatorname{SL}_{2}(\mathcal{O}_{K}) and CKC_{K}, where k,1\langle k,1\rangle is the ideal generated by k,1k,1 if kKk\in K and k,1:=𝒪K\langle k,1\rangle:=\mathcal{O}_{K} if k=k=\infty. In particular,

κ:=#CK=number of cusps of ΓK.\displaystyle\kappa:=\#C_{K}=\text{number of cusps of $\Gamma_{K}$}.

Throughout the remainder of this paper, we fix k1=,k2,,kκK¯k_{1}=\infty,k_{2},\ldots,k_{\kappa}\in\overline{K} to be a complete list of cusps of ΓK\Gamma_{K}. Fix a1=1a_{1}=1, a2,,aκ𝒪Ka_{2},\ldots,a_{\kappa}\in{\mathcal{O}}_{K} and c1=0c_{1}=0, c2,,cκ𝒪Kc_{2},\ldots,c_{\kappa}\in{\mathcal{O}}_{K} so that ki=aicik_{i}=\frac{a_{i}}{c_{i}} for each ii; then set Ii:=ai,ciI_{i}:=\langle a_{i},c_{i}\rangle. Then I1=𝒪K,I2,,IκI_{1}=\mathcal{O}_{K},I_{2},\ldots,I_{\kappa} are integral ideals which by Lemma 2.1 form a system of representatives of the ideal classes of KK. (Note that IiI_{i} depends on the choice of ai,cia_{i},c_{i}; however the class [Ii][I_{i}] depends only on kik_{i}.)

For each 1iκ1\leq i\leq\kappa, let

(2.7) Γi:={γΓK:γι(ki)=ι(ki)}\displaystyle\Gamma_{i}:=\left\{\gamma\in\Gamma_{K}\>:\>\gamma\iota(k_{i})=\iota(k_{i})\right\}

be the isotropy group of the cusp kik_{i}. Below we give a more precise description of these isotropy groups. First, the isotropy group of \infty is easy to compute: For γ=ι((abcd))ΓK\gamma=\iota\left(\left(\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}\right)\right)\in\Gamma_{K}, by direct computation γι()=ι()\gamma\cdot\iota(\infty)=\iota(\infty) if and only if c=0c=0, that is

Γ1=ΓKB={ι((ab0a1)):a𝒪K×,b𝒪K},\displaystyle\Gamma_{1}=\Gamma_{K}\cap B=\left\{\iota\left(\left(\begin{smallmatrix}a&b\\ 0&a^{-1}\end{smallmatrix}\right)\right)\>:\>a\in\mathcal{O}_{K}^{\times},\ b\in\mathcal{O}_{K}\right\},

where B<HB<\mathrm{H} is the subgroup of upper triangular matrices in H\mathrm{H}.

To compute the other isotropy groups, we will translate the cusp kik_{i} to \infty. For any integral ideal I𝒪KI\subset\mathcal{O}_{K} let

(2.8) ΓK,I:={ι((abcd)):a,d𝒪K,bI2,cI2,adbc=1}.\displaystyle\Gamma_{K,I}:=\left\{\iota\left(\left(\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}\right)\right)\>:\>a,d\in\mathcal{O}_{K},\ b\in I^{-2},\ c\in I^{2},\ ad-bc=1\right\}.

For each 1iκ1\leq i\leq\kappa since Ii=ai,ciI_{i}=\langle a_{i},c_{i}\rangle, we can find bi,diIi1b_{i},d_{i}\in I_{i}^{-1} such that aidibici=1a_{i}d_{i}-b_{i}c_{i}=1. Throughout the paper we fix the scaling matrix

(2.9) ξi=ι((aibicidi))H.\displaystyle{\xi}_{i}=\iota\left(\left(\begin{smallmatrix}a_{i}&b_{i}\\ c_{i}&d_{i}\end{smallmatrix}\right)\right)\in\mathrm{H}.

Note that ξi\xi_{i} satisfies ξiι()=ι(ki)\xi_{i}\iota(\infty)=\iota(k_{i}). We then have the following description of Γi\Gamma_{i}.

Lemma 2.2.

We have for each 1iκ1\leq i\leq\kappa,

(2.10) Γi=ξi(ΓK,IiB)ξi1=ξi{ι((ab0a1)):a𝒪K×,bIi2}ξi1.\displaystyle\Gamma_{i}=\xi_{i}\left(\Gamma_{K,I_{i}}\cap B\right)\xi_{i}^{-1}=\xi_{i}\left\{\iota\left(\left(\begin{smallmatrix}a&b\\ 0&a^{-1}\end{smallmatrix}\right)\right)\>:\>a\in\mathcal{O}_{K}^{\times},\ b\in I_{i}^{-2}\right\}\xi_{i}^{-1}.
Proof.

By definition, γΓi\gamma\in\Gamma_{i} if any only if γι(ki)=ι(ki)\gamma\iota(k_{i})=\iota(k_{i}). Since ξiι()=ι(ki)\xi_{i}\iota(\infty)=\iota(k_{i}), this is equivalent to ξi1γξiι()=ι()\xi_{i}^{-1}\gamma\xi_{i}\iota(\infty)=\iota(\infty). This shows that

ξi1Γiξi=(ξi1ΓKξi)B.{\xi}_{i}^{-1}\Gamma_{i}{\xi}_{i}=({\xi}^{-1}_{i}\Gamma_{K}{\xi}_{i})\cap B.

By direction computation we have ξi1ΓKξiΓK,Ii{\xi}^{-1}_{i}\Gamma_{K}{\xi}_{i}\subset\Gamma_{K,I_{i}} and ξiΓK,Iiξi1ΓK\xi_{i}\Gamma_{K,I_{i}}\xi_{i}^{-1}\subset\Gamma_{K}, implying that ξi1ΓKξi=ΓK,Ii{\xi}^{-1}_{i}\Gamma_{K}{\xi}_{i}=\Gamma_{K,I_{i}}. Hence

ξi1Γiξi=ΓK,IiB={ι((ab0a1)):a𝒪K×,bIi2}.\displaystyle{\xi}_{i}^{-1}\Gamma_{i}{\xi}_{i}=\Gamma_{K,I_{i}}\cap B=\left\{\iota\left(\left(\begin{smallmatrix}a&b\\ 0&a^{-1}\end{smallmatrix}\right)\right)\>:\>a\in\mathcal{O}_{K}^{\times},\ b\in I_{i}^{-2}\right\}.

We can then finish the proof by conjugating both sides of the above equation by ξi{\xi}_{i}. ∎

2.3. Coordinates and measures

Let

K:={(α,β,σ(α),σ(β))4:(α,β)𝒪K2}\displaystyle\mathcal{L}_{K}:=\left\{(\alpha,\beta,\sigma(\alpha),\sigma(\beta))\in\mathbb{R}^{4}\>:\>(\alpha,\beta)\in\mathcal{O}_{K}^{2}\right\}

be the Minkowski embedding of 𝒪K2\mathcal{O}_{K}^{2} in 4\mathbb{R}^{4}. Let XX be the space of lattices of the form Kh\mathcal{L}_{K}h with hHh\in\mathrm{H}. Since Kh=K\mathcal{L}_{K}h=\mathcal{L}_{K} if and only if hΓKh\in\Gamma_{K}, XX can be parameterized by the homogeneous space ΓK\H\Gamma_{K}\backslash\mathrm{H} via KhXΓKhΓK\H\mathcal{L}_{K}h\in X\leftrightarrow\Gamma_{K}h\in\Gamma_{K}\backslash\mathrm{H}. We thus identify XX with ΓK\H\Gamma_{K}\backslash\mathrm{H}. Let μK\mu_{K} be the unique invariant probability measure on X=ΓK\HX=\Gamma_{K}\backslash\mathrm{H}.

Since we will be working with this measure extensively when computing G(s)G(s) later in Section 3, here we give a more explicit description of it in terms of coordinates from an Iwasawa decomposition of H\mathrm{H}. It is well known that the group SL2()\operatorname{SL}_{2}(\mathbb{R}) has an Iwasawa decomposition saying that any element gSL2()g\in\operatorname{SL}_{2}(\mathbb{R}) can be written uniquely as g=nxaykθg=\text{n}_{x}\text{a}_{y}\text{k}_{\theta} for some x,y>0x\in\mathbb{R},y>0 and θ/2π\theta\in\mathbb{R}/2\pi\mathbb{Z}, where nx:=(1x01)\text{n}_{x}:=\left(\begin{smallmatrix}1&x\\ 0&1\end{smallmatrix}\right), ay:=(y00y1)\text{a}_{y}:=\left(\begin{smallmatrix}y&0\\ 0&y^{-1}\end{smallmatrix}\right) and kθ:=(cosθsinθsinθcosθ)\text{k}_{\theta}:=\left(\begin{smallmatrix}\cos\theta&-\sin\theta\\ \sin\theta&\cos\theta\end{smallmatrix}\right). This thus induces an Iwasawa decomposition of H\mathrm{H}: Every hHh\in\mathrm{H} can be written uniquely in the form h=n𝒙a𝒚k𝜽h=\text{n}_{\bm{x}}\text{a}_{\bm{y}}\text{k}_{\bm{\theta}} with 𝒙2\bm{x}\in\mathbb{R}^{2}, 𝒚(>0)2\bm{y}\in(\mathbb{R}_{>0})^{2} and 𝜽(/2π)2\bm{\theta}\in(\mathbb{R}/2\pi\mathbb{Z})^{2}, where

n𝒙:=(nx1,nx2),a𝒚:=(ay1,ay2)andk𝜽:=(kθ1,kθ2).\text{n}_{\bm{x}}:=\left(\text{n}_{x_{1}},\text{n}_{x_{2}}\right),\quad\text{a}_{\bm{y}}:=\left(\text{a}_{y_{1}},\text{a}_{y_{2}}\right)\quad\text{and}\quad\text{k}_{\bm{\theta}}:=(\text{k}_{\theta_{1}},\text{k}_{\theta_{2}}).

Under these coordinates, the Haar measure of H\mathrm{H} (up to scalars) is given by

(2.11) dμH(h)=y13y23d𝒙d𝒚d𝜽.\displaystyle\text{d}\mu_{\mathrm{H}}(h)=y_{1}^{-3}y_{2}^{-3}\,\text{d}\bm{x}\,\text{d}\bm{y}\,\text{d}\bm{\theta}.

Note that μK\mu_{K} essentially comes from a Haar measure of H\mathrm{H}. Indeed, we may identify XX with a fundamental domain inside H\mathrm{H}; then μK\mu_{K} is the restriction of a certain Haar measure to this fundamental domain normalized to be a probability measure. Thus μK\mu_{K} is given by

(2.12) dμK(h)=cKy13y23d𝒙d𝒚d𝜽\displaystyle\text{d}\mu_{K}(h)=c_{K}y_{1}^{-3}y_{2}^{-3}\,\text{d}\bm{x}\,\text{d}\bm{y}\,\text{d}\bm{\theta}

for some normalizing factor cK>0c_{K}>0. This normalizing factor was computed by Siegel [42, 43] and is given by the following formula (see [45, p. 59]):

(2.13) cK=ΔK3/2ζK(2)1,\displaystyle c_{K}=\Delta_{K}^{-3/2}\zeta_{K}(2)^{-1},

where ζK(s):=I𝒪KNr(I)s\zeta_{K}(s):=\sum_{I\subset\mathcal{O}_{K}}\text{Nr}(I)^{-s} (𝔢(s)>1{\mathfrak{Re}}(s)>1) is the Dedekind zeta function attached to KK. Here the summation is over all the nonzero integral ideals of 𝒪K\mathcal{O}_{K}.

Remark 2.14.

To compare our formula for a𝒫a_{\mathcal{P}} in (1.17) with numerical computations, we need to express ζK(2)\zeta_{K}(2) in more explicit terms. Let χK\chi_{K} be the Dirichlet character defined by χK(n):=(ΔKn)\chi_{K}(n):=\left(\frac{\Delta_{K}}{n}\right) with ()\left(\frac{\cdot}{\cdot}\right) the Kronecker symbol. Note that χK\chi_{K} is an even primitive quadratic character of modulus ΔK\Delta_{K}; see [32, p. 296-297]. Using the fact that for any prime number pp, the ideal (p)𝒪K(p)\subset\mathcal{O}_{K} is inert, ramified or split if and only if χK(p)\chi_{K}(p) equals 1,0,1-1,0,1 respectively, we have the following well-known formula for ζK\zeta_{K}:

ζK(s)=ζ(s)L(s,χK),\displaystyle\zeta_{K}(s)=\zeta(s)L(s,\chi_{K}),

where ζ(s)\zeta(s) is the Riemann zeta function and L(s,χK)L(s,\chi_{K}) is the Dirichlet LL-function associated to χK\chi_{K}. Now by [46, Prop. 4.1 and Thm. 4.2] we have for any nn\in\mathbb{N},

L(1n,χK)=ΔKn1na=1ΔKχK(a)Bn(aΔK),\displaystyle L(1-n,\chi_{K})=-\frac{\Delta_{K}^{n-1}}{n}\sum_{a=1}^{\Delta_{K}}\chi_{K}(a)B_{n}\left(\tfrac{a}{\Delta_{K}}\right),

where Bn(X)B_{n}(X) is the nn-th Bernoulli polynomial given by the formula Bn(X)=i=0n(ni)BiXniB_{n}(X)=\sum_{i=0}^{n}\binom{n}{i}B_{i}X^{n-i} with BiB_{i} the ii-th Bernoulli number. Combined with the functional equation of ζK(s)\zeta_{K}(s) (see [45, p. 59]), we get

(2.15) ζK(2)=4π4ΔK32ζK(1)=4π4ΔK32ζ(1)L(1,χK)=π46ΔKa=1ΔKχK(a)B2(aΔK).\displaystyle\zeta_{K}(2)=4\pi^{4}\Delta_{K}^{-\frac{3}{2}}\zeta_{K}(-1)=4\pi^{4}\Delta_{K}^{-\frac{3}{2}}\zeta(-1)L(-1,\chi_{K})=\frac{\pi^{4}}{6\sqrt{\Delta_{K}}}\sum_{a=1}^{\Delta_{K}}\chi_{K}(a)B_{2}\left(\tfrac{a}{\Delta_{K}}\right).

For instance when K=(2)K=\mathbb{Q}(\sqrt{2}) we have ΔK=8\Delta_{K}=8 and χK(n)=(8n)\chi_{K}(n)=\left(\frac{8}{n}\right) is the quadratic character of modulus 88 with χK(1)=χK(7)=1\chi_{K}(1)=\chi_{K}(7)=1 and χK(3)=χK(5)=1\chi_{K}(3)=\chi_{K}(5)=-1. We also have B2(1/8)=B2(7/8)=11192B_{2}(1/8)=B_{2}(7/8)=\frac{11}{192} and B2(3/8)=B2(5/8)=13192B_{2}(3/8)=B_{2}(5/8)=-\frac{13}{192}. Plugging all these relations we get in this case ζK(2)=π4482\zeta_{K}(2)=\frac{\pi^{4}}{48\sqrt{2}}. Similarly, one can apply (2.15) to get ζ(3)(2)=π4363\zeta_{\mathbb{Q}(\sqrt{3})}(2)=\frac{\pi^{4}}{36\sqrt{3}} and ζ(5)(2)=2π4755\zeta_{\mathbb{Q}(\sqrt{5})}(2)=\frac{2\pi^{4}}{75\sqrt{5}}.

2.4. Siegel domain and cusp neighborhoods

In this section we give a more precise description of the homogeneous space X=ΓK\HX=\Gamma_{K}\backslash\mathrm{H} in terms of the coordinates given in the previous section.

First, note that under these coordinates and in view of Lemma 2.2, for each 1iκ1\leq i\leq\kappa the set

(2.16) i:={ξin𝒙a𝒚k𝜽H:𝒙𝔉i,y1/y2[1,λ2),𝜽[0,π)×[0,2π)}\displaystyle\mathcal{F}_{i}:=\left\{{\xi}_{i}\text{n}_{\bm{x}}\text{a}_{\bm{y}}\text{k}_{\bm{\theta}}\in\mathrm{H}\>:\>\bm{x}\in\mathfrak{F}_{i},\ y_{1}/y_{2}\in[1,\lambda^{2}),\ \bm{\theta}\in[0,\pi)\times[0,2\pi)\right\}

is a fundamental domain for Γi\H\Gamma_{i}\backslash\mathrm{H}. Here 𝔉i2\mathfrak{F}_{i}\subset\mathbb{R}^{2} is a fundamental domain for 2/ι(Ii2)\mathbb{R}^{2}/\iota(I_{i}^{-2}). (Recall that ι(I)\iota(I) is a lattice in 2\mathbb{R}^{2} for any fractional ideal II of 𝒪K\mathcal{O}_{K}.)

Now for each 1iκ1\leq i\leq\kappa and any t>0t>0 define

(2.17) i(t):={ξin𝒙a𝒚k𝜽i:y1y2t}.\displaystyle\mathcal{F}_{i}(t):=\left\{\xi_{i}\text{n}_{\bm{x}}\text{a}_{\bm{y}}\text{k}_{\bm{\theta}}\in\mathcal{F}_{i}\>:\>y_{1}y_{2}\geq t\right\}.

It follows from Shimizu’s lemma [41, Lemma 5] that there exists some constant t1>0t_{1}>0 depending only on KK such that

(2.18) γΓK,1i,jκ:γi(t1)j(t1)i=jandγ=id.\displaystyle\forall\,\gamma\in\Gamma_{K},1\leq i,j\leq\kappa\>:\>\gamma\mathcal{F}_{i}(t_{1})\cap\mathcal{F}_{j}(t_{1})\neq\emptyset\quad\Leftrightarrow\quad i=j\ \text{and}\ \gamma=\text{id}.

Moreover, by the reduction theory of Borel and Harish-Chandra [11] we have a Siegel fundamental domain of the form

(2.19) ΓK:=(i=1κi(t1)).\displaystyle\mathcal{F}_{\Gamma_{K}}:=\mathfrak{C}\bigcup\left(\bigsqcup_{i=1}^{\kappa}\mathcal{F}_{i}(t_{1})\right).

Here H\mathfrak{C}\subset\mathrm{H} is compact and the natural map from ΓK\mathcal{F}_{\Gamma_{K}} to ΓK\H\Gamma_{K}\backslash\mathrm{H} is surjective and finite-to-one. Note, on the other hand, it follows from (2.18) that the projection from i=1κi(t1)\bigsqcup_{i=1}^{\kappa}\mathcal{F}_{i}(t_{1}) to ΓK\H\Gamma_{K}\backslash\mathrm{H} is injective.

2.5. Geometry of lattices avoiding large balls

In this section we give necessary conditions for lattices KhX\mathcal{L}_{K}h\in X avoiding large balls. These lattices will be the main objects we deal with when computing G(s)G(s). We have the following description of these lattices.

Proposition 2.3.

There exists some R0>0R_{0}>0 depending only on KK such that for any R>R0R>R_{0} and for any lattice KhX\mathcal{L}_{K}h\in X avoiding a ball of radius RR, we have Kh=Kξin𝐱a𝐲k𝛉ΓKi(t1)\mathcal{L}_{K}h=\mathcal{L}_{K}\xi_{i}\text{n}_{\bm{x}}\text{a}_{\bm{y}}\text{k}_{\bm{\theta}}\in\Gamma_{K}\mathcal{F}_{i}(t_{1}) for some 1iκ1\leq i\leq\kappa. Moreover, we have y1y2Ry_{1}\asymp y_{2}\gg R with the bounding constants depending only on KK.

Remark 2.20.

Regarding the notation “\asymp”, “\gg” and “\ll”: For two positive quantities AA and BB, we write ABA\ll B or BAB\gg A to mean that there is a constant C>0C>0 such that ACBA\leq CB, and we will write ABA\asymp B for ABAA\ll B\ll A. We will sometimes use subscripts to indicate the dependence of the bounding constant CC on parameters.

Proof.

Let 𝒗1=(1,0,1,0)\bm{v}_{1}=(1,0,1,0), 𝒗2=(τ,0,σ(τ),0)\bm{v}_{2}=(\tau,0,\sigma(\tau),0), 𝒗3=(0,1,0,1)\bm{v}_{3}=(0,1,0,1) and 𝒗4=(0,τ,0,σ(τ))\bm{v}_{4}=(0,\tau,0,\sigma(\tau)). Note that {𝒗j:j=1,2,3,4}\{\bm{v}_{j}\>:\>j=1,2,3,4\} is a basis for K\mathcal{L}_{K}. Let R0>0R_{0}>0 be sufficiently large such that

(2.21) 2𝒗jhR0, 1j4,h.\displaystyle 2\|\bm{v}_{j}h\|\leq R_{0},\qquad\forall\,1\leq j\leq 4,\ h\in\mathfrak{C}.

We note that R0R_{0} depends only on the compact set \mathfrak{C} (hence only depends on KK). Now take R>R0R>R_{0} and suppose xXx\in X avoids a ball BRB_{R} of radius RR. To prove the first claim, in view of the description of the Siegel fundamental domain ΓK\mathcal{F}_{\Gamma_{K}} in (2.19), we want to show there does not exist any hh\in\mathfrak{C} such that x=Khx=\mathcal{L}_{K}h. We prove by contradiction. Suppose there exists such hh\in\mathfrak{C}. Then {𝒗jh: 1j4}\{\bm{v}_{j}h\>:\>1\leq j\leq 4\} is a basis for x=Khx=\mathcal{L}_{K}h. Let 𝒙4\bm{x}\in\mathbb{R}^{4} be the center of BRB_{R} and write 𝒙=j=14aj𝒗jh\bm{x}=\sum_{j=1}^{4}a_{j}\bm{v}_{j}h (aja_{j}\in\mathbb{R}) as an \mathbb{R}-linear combination of these basis vectors. For each 1j41\leq j\leq 4, take njn_{j}\in\mathbb{Z} the closest integer to aja_{j} so that we have |njaj|12|n_{j}-a_{j}|\leq\frac{1}{2}. Consider the lattice vector 𝒙:=j=14nj𝒗jhKh\bm{x}^{\prime}:=\sum_{j=1}^{4}n_{j}\bm{v}_{j}h\in\mathcal{L}_{K}h. Since KhBR=\mathcal{L}_{K}h\cap B_{R}=\emptyset, we have

R𝒙𝒙j=14|njaj|𝒗jh12i=j4𝒗jh,\displaystyle R\leq\|\bm{x}^{\prime}-\bm{x}\|\leq\sum_{j=1}^{4}|n_{j}-a_{j}|\|\bm{v}_{j}h\|\leq\frac{1}{2}\sum_{i=j}^{4}\|\bm{v}_{j}h\|,

implying that there exists 1j41\leq j\leq 4 such that 𝒗jhR2>R02\|\bm{v}_{j}h\|\geq\frac{R}{2}>\frac{R_{0}}{2}, contradicting (2.21). This proves the first claim, i.e. KhΓKi(t1)\mathcal{L}_{K}h\in\Gamma_{K}\mathcal{F}_{i}(t_{1}) for some 1iκ1\leq i\leq\kappa. For the second claim, without loss of generality we may assume h=ξin𝒙a𝒚k𝜽i(t1)h=\xi_{i}\text{n}_{\bm{x}}\text{a}_{\bm{y}}\text{k}_{\bm{\theta}}\in\mathcal{F}_{i}(t_{1}). To show y1y2Ry_{1}\asymp y_{2}\asymp R, first note that y1λy2y_{1}\asymp_{\lambda}y_{2} and y1y2t11y_{1}y_{2}\geq t_{1}\gg 1. Thus

y1y2max{y1,y11,y2,y21}=a𝒚op.\displaystyle y_{1}\asymp y_{2}\asymp\max\{y_{1},y_{1}^{-1},y_{2},y_{2}^{-1}\}=\|\text{a}_{\bm{y}}\|_{\rm op}.

Here op\|\cdot\|_{\rm op} denotes the operator norm (with respect to the Euclidean norm). On the other hand, for any 1j41\leq j\leq 4,

𝒗jh𝒗jξiopn𝒙opa𝒚op𝔉i,ξia𝒚opy1y2.\displaystyle\|\bm{v}_{j}h\|\leq\|\bm{v}_{j}\|\|\xi_{i}\|_{\rm op}\|\text{n}_{\bm{x}}\|_{\rm op}\|\text{a}_{\bm{y}}\|_{\rm op}\ll_{\mathfrak{F}_{i},\xi_{i}}\|\text{a}_{\bm{y}}\|_{\rm op}\asymp y_{1}\asymp y_{2}.

This, together with the bound j=14𝒗jh2R\sum_{j=1}^{4}\|\bm{v}_{j}h\|\geq 2R implies that y1y2Ry_{1}\asymp y_{2}\gg R as claimed. ∎

For later reference, note that for each 1iκ1\leq i\leq\kappa,

(2.22) K={(α,β,σ(α),σ(β))ξi1:αIi,βIi1}\displaystyle\mathcal{L}_{K}=\left\{(\alpha,\beta,\sigma(\alpha),\sigma(\beta))\xi_{i}^{-1}\>:\>\alpha\in I_{i},\ \beta\in I_{i}^{-1}\right\}

with Ii=ai,ciI_{i}=\langle a_{i},c_{i}\rangle as before. Let 𝖈i=(ci,ai)𝒪K2\bm{\mathfrak{c}}_{i}=(-c_{i},a_{i})\in\mathcal{O}_{K}^{2}. We denote by

(2.23) Πi:=ι(𝖈i)ι(τ𝖈i)=((0,1,0,1)(0,τ,0,σ(τ)))ξi1=({0}××{0}×)ξi1,\displaystyle\Pi_{i}:=\mathbb{R}\iota(\bm{\mathfrak{c}}_{i})\oplus\mathbb{R}\iota(\tau\bm{\mathfrak{c}}_{i})=\left(\mathbb{R}(0,1,0,1)\oplus\mathbb{R}(0,\tau,0,\sigma(\tau))\right)\xi^{-1}_{i}=(\{0\}\times\mathbb{R}\times\{0\}\times\mathbb{R})\xi^{-1}_{i},

and

(2.24) i:=ΠiK\displaystyle\mathcal{L}_{i}:=\Pi_{i}\cap\mathcal{L}_{K} =(({0}××{0}×)Kξi)ξi1={(0,β,0,σ(β))ξi1:βIi1}.\displaystyle=\left((\{0\}\times\mathbb{R}\times\{0\}\times\mathbb{R})\cap\mathcal{L}_{K}\xi_{i}\right)\xi^{-1}_{i}=\left\{(0,\beta,0,\sigma(\beta))\xi^{-1}_{i}\>:\>\beta\in I_{i}^{-1}\right\}.

By Proposition 2.3 if xXx\in X avoids some ball of radius R>R0R>R_{0}, then x=Khx=\mathcal{L}_{K}h with h=ξin𝒙a𝒚k𝜽i(t1)h=\xi_{i}\text{n}_{\bm{x}}\text{a}_{\bm{y}}\text{k}_{\bm{\theta}}\in\mathcal{F}_{i}(t_{1}) for some 1iκ1\leq i\leq\kappa and y1,y2>0y_{1},y_{2}>0 with y1y2Ry_{1}\asymp y_{2}\gg R. Then the two vectors

ι(𝖈i)h=(0,y11,0,y21)k𝜽andι(τ𝖈i)h=(0,τy11,0,σ(τ)y21)k𝜽\displaystyle\iota(\bm{\mathfrak{c}}_{i})h=(0,y_{1}^{-1},0,y_{2}^{-1})\text{k}_{\bm{\theta}}\quad\text{and}\quad\iota(\tau\bm{\mathfrak{c}}_{i})h=(0,\tau y_{1}^{-1},0,\sigma(\tau)y_{2}^{-1})\text{k}_{\bm{\theta}}

are of length R1\ll R^{-1}. Note that ι(𝖈i)h\iota(\bm{\mathfrak{c}}_{i})h and ι(τ𝖈i)h\iota(\tau\bm{\mathfrak{c}}_{i})h are linearly independent and ih\mathcal{L}_{i}h is the primitive rank two sublattice of xx containing them. Thus ih\mathcal{L}_{i}h lies O(R1)O(R^{-1})-densely in the plane Πih\Pi_{i}h. For this reason, we call Πih\Pi_{i}h the filled plane of x=Khx=\mathcal{L}_{K}h.

3. Asymptotics of G(s)G(s)

3.1. Main result

The main goal of this Section 3 is to prove Theorem 2, which gives the tail asymptotics of both the limiting gap distribution function F(s)F(s) and its integral G(s)G(s), for the particular class of cut-and-project sets which we consider. In fact we will prove Theorem 4 below, which will be shown (in Section 3.2) to imply Theorem 2. This Theorem 4 gives an explicit bound on the error term in the asymptotics for G(s)G(s), and also an explicit formula for the leading coefficient a𝒫a_{{\mathcal{P}}}. As we mentioned in the introduction, the asymptotics for F(s)F(s) will be derived as a consequence of those for G(s)G(s); see Lemma 3.2 below.

In order to state Theorem 4, we introduce some further notation. For any θ\theta\in\mathbb{R} we write kθ:=(cosθsinθsinθcosθ)SL2()\text{k}_{\theta}:=\left(\begin{smallmatrix}\cos\theta&-\sin\theta\\ \sin\theta&\cos\theta\end{smallmatrix}\right)\in\operatorname{SL}_{2}(\mathbb{R}). Then for any θ,x\theta,x\in\mathbb{R} and any subset 𝒲2{\mathcal{W}}\subset\mathbb{R}^{2}, we set

(3.1) R𝒲(θ,x):={y:(x,y)kθ𝒲}.\displaystyle R_{{\mathcal{W}}}(\theta,x):=\{y\in\mathbb{R}\>:\>(x,y)\text{k}_{\theta}\in{\mathcal{W}}\}.

It follows that 𝒲(θ)\ell_{\mathcal{W}}(\theta), which we defined in Section 1.4, can be expressed as

(3.2) 𝒲(θ)={x:R𝒲(θ,x)}.\displaystyle\ell_{{\mathcal{W}}}(\theta)=\bigl{\{}x\in\mathbb{R}\>:\>R_{{\mathcal{W}}}(\theta,x)\neq\emptyset\bigr{\}}.

Note also that 𝒲(θ)\ell_{\mathcal{W}}(\theta) satisfies the relation 𝒲(θ)=𝒲(π+θ)\ell_{\mathcal{W}}(\theta)=-\ell_{\mathcal{W}}(\pi+\theta) for any θ\theta\in\mathbb{R}. For any integral ideal I𝒪KI\subset\mathcal{O}_{K}, any bounded Lebesgue measurable set JJ\subset\mathbb{R} with non-empty interior, and any y>0y>0, we set

(3.3) αI,J(y):=min{αI>0:yσ(α)J}.\displaystyle\alpha_{I,J}(y):=\min\bigl{\{}\alpha\in I\cap\mathbb{R}_{>0}\>:\>y\,\sigma(\alpha)\in J\bigr{\}}.

For later reference we record here a useful identity for the function αI,J\alpha_{I,J} which can be checked directly from its definition: For any βI1\beta\in I^{-1} with σ(β)>0\sigma(\beta)>0 and for any y>0y>0,

(3.4) αβI,J(y)=|β|αI,sgn(β)J(σ(β)y).\displaystyle\alpha_{\beta I,J}(y)=|\beta|\alpha_{I,\operatorname{sgn}(\beta)J}(\sigma(\beta)y).

Let us write mm for the Lebesgue measure on \mathbb{R}. For any Lebesgue measurable set AA\subset\mathbb{R} and a>0a>0, we define 𝔣(A,a)\mathfrak{f}(A,a) to be the infimum of a1m(JA)a^{-1}m(J\setminus A) when JJ ranges over all intervals of length aa:

(3.5) 𝔣(A,a):\displaystyle\mathfrak{f}(A,a): =a1inf{m((x,x+a)A):x}.\displaystyle=a^{-1}\inf\bigl{\{}m\bigl{(}(x,x+a)\setminus A\bigr{)}\>:\>x\in\mathbb{R}\bigr{\}}.

Note that 0𝔣(A,a)10\leq\mathfrak{f}(A,a)\leq 1 always, and 𝔣(A,a)=0\mathfrak{f}(A,a)=0 whenever AA contains some interval of length aa. Let us also define

(3.6) 𝔣~(A,a):={𝔣(A,a)if A,0if A=.\displaystyle\widetilde{\mathfrak{f}}(A,a):=\begin{cases}\,\mathfrak{f}(A,a)&\text{if }\>A\neq\emptyset,\\ 0&\text{if }\>A=\emptyset.\end{cases}

Recall from below Lemma 2.1 that we have made a fixed choice of ideals I1,,IκI_{1},\ldots,I_{\kappa} of 𝒪K{\mathcal{O}}_{K}, and that these form a system of representatives of the ideal classes of KK.

Theorem 4.

Let 𝒫=𝒫(𝒲,K){\mathcal{P}}={\mathcal{P}}({\mathcal{W}},{\mathcal{L}}_{K}) where KK is a real quadratic field and 𝒲{\mathcal{W}} is a bounded open subset of 2\mathbb{R}^{2} such that 𝟎𝒲\mathbf{0}\in{\mathcal{W}} and 𝒲\partial{\mathcal{W}} has Lebesgue measure zero. Let F(s)F(s) be the associated limiting gap distribution function as in Theorem 1, and let G(s)=sF(t)dtG(s)=\int_{s}^{\infty}F(t)\,\textup{d}t. Then there exists a positive constant C=C𝒫C=C_{\mathcal{P}} such that for all sufficiently large ss,

(3.7) 0G(s)a𝒫ss20Cs02π𝔣~(R𝒲(θ,y),u1)dydθdu,\displaystyle 0\leq G(s)-\frac{a_{{\mathcal{P}}}}{s}\ll s^{-2}\int_{0}^{Cs}\int_{0}^{2\pi}\int_{\mathbb{R}}\widetilde{\mathfrak{f}}\biggl{(}R_{{\mathcal{W}}}(\theta,y),u^{-1}\biggr{)}\,\mathrm{d}y\,\mathrm{d}\theta\,\mathrm{d}u,

where

(3.8) a𝒫=Area(𝒲)4ΔK2ζK(2)i=1κNr(Ii)202π1λ2αIi,𝒲(θ)(y)2dyy3dθ.\displaystyle a_{{\mathcal{P}}}=\frac{\operatorname{Area}({\mathcal{W}})}{4\Delta_{K}^{2}\zeta_{K}(2)}\sum_{i=1}^{\kappa}\operatorname{Nr}(I_{i})^{-2}\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}\alpha_{I_{i},\ell_{\mathcal{W}}(\theta)}(y)^{2}\,\frac{\mathrm{d}y}{y^{3}}\,\mathrm{d}\theta.

We here make a couple of remarks regarding Theorem 4; for two more remarks see the end of Section 3.2.

Remark 3.9.

In the formula (3.8), it should be noted that for any integral ideal II of 𝒪K{\mathcal{O}}_{K}, we have

(3.10) 1λ2αI,𝒲(θ)(y)2dyy3=aaλ2αI,𝒲(θ)(y)2dyy3,a>0,θ,\displaystyle\int_{1}^{\lambda^{2}}\alpha_{I,\ell_{\mathcal{W}}(\theta)}(y)^{2}\,\frac{\text{d}y}{y^{3}}=\int_{a}^{a\lambda^{2}}\alpha_{I,\ell_{\mathcal{W}}(\theta)}(y)^{2}\,\frac{\text{d}y}{y^{3}},\qquad\forall\,a>0,\>\theta\in\mathbb{R},

and furthermore, the product

(3.11) Nr(I)202π1λ2αI,𝒲(θ)(y)2dyy3dθ\displaystyle\operatorname{Nr}(I)^{-2}\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}\alpha_{I,\ell_{\mathcal{W}}(\theta)}(y)^{2}\,\frac{\text{d}y}{y^{3}}\,\text{d}\theta

depends only on the ideal class [I][I] of II.
To prove (3.10), it suffices to note that the differential form αI,𝒲(θ)(y)2dyy3\alpha_{I,\ell_{\mathcal{W}}(\theta)}(y)^{2}\,\frac{\text{d}y}{y^{3}} is invariant under the map sending yy to λ2y\lambda^{2}y, a fact which follows by applying (3.4) for the special case when β=λ2\beta=\lambda^{-2}. Next, to prove the invariance of (3.11), assume that II^{\prime} is another integral ideal belonging to the same ideal class as II. Then I=aII^{\prime}=aI for some nonzero aI1a\in I^{-1}, which we may take to satisfy σ(a)>0\sigma(a)>0 (otherwise replace aa by a-a). Then the absolute norm Nr(Ii)2\operatorname{Nr}(I_{i})^{-2} scales by N(a)2\mathrm{N}(a)^{-2} when replacing II by I=aII^{\prime}=aI. On the other hand, using (3.4) and the fact that 𝒲(θ+π)=𝒲(θ)\ell_{\mathcal{W}}(\theta+\pi)=-\ell_{\mathcal{W}}(\theta) for any θ\theta\in\mathbb{R}, we have

αI,𝒲(θ)(y)={aαI,𝒲(θ)(σ(a)y)if a>0,|a|αI,𝒲(θ+π)(σ(a)y)if a<0,\displaystyle\alpha_{I^{\prime},\ell_{\mathcal{W}}(\theta)}(y)=\begin{cases}a\alpha_{I,\ell_{\mathcal{W}}(\theta)}(\sigma(a)y)&\text{if $a>0$},\\[2.0pt] |a|\alpha_{I,\ell_{\mathcal{W}}(\theta+\pi)}(\sigma(a)y)&\text{if $a<0$},\end{cases}

from which it follows (via the obvious substitution and then using (3.10) with σ(a)\sigma(a) in place of aa) that the double integral in (3.11) scales by a factor N(a)2\mathrm{N}(a)^{2} when replacing II by I=aII^{\prime}=aI. Hence the product in (3.11) is invariant as claimed.

Remark 3.12.

The right hand side of (3.8) is invariant under replacing 𝒲{\mathcal{W}} by 𝒲g{\mathcal{W}}g, for any gGL2()g\in\operatorname{GL}_{2}(\mathbb{R}).

Proof.

We will prove the claim by showing that for any integral ideal II of 𝒪K{\mathcal{O}}_{K}, and any gGL2()g\in\operatorname{GL}_{2}(\mathbb{R}),

(3.13) Area(𝒲g)02π1λ2αI,𝒲g(θ)(y)2dyy3dθ=Area(𝒲)02π1λ2αI,𝒲(θ)(y)2dyy3dθ.\displaystyle\operatorname{Area}({\mathcal{W}}g)\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}\alpha_{I,\ell_{\mathcal{W}g}(\theta)}(y)^{2}\,\frac{\text{d}y}{y^{3}}\,\text{d}\theta=\operatorname{Area}({\mathcal{W}})\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}\alpha_{I,\ell_{\mathcal{W}}(\theta)}(y)^{2}\,\frac{\text{d}y}{y^{3}}\,\text{d}\theta.

Write 𝒆1=(1,0){\text{\boldmath$e$}}_{1}=(1,0), 𝒆2=(0,1){\text{\boldmath$e$}}_{2}=(0,1). For fixed gGL2()g\in\operatorname{GL}_{2}(\mathbb{R}), define the smooth maps ω:/2π/2π\omega:\mathbb{R}/2\pi\mathbb{Z}\to\mathbb{R}/2\pi\mathbb{Z} and ρ:/2π>0\rho:\mathbb{R}/2\pi\mathbb{Z}\to\mathbb{R}_{>0} through

(3.14) 𝒆2kθg1=ρ(θ)𝒆2kω(θ).\displaystyle{\text{\boldmath$e$}}_{2}\text{k}_{\theta}g^{-1}=\rho(\theta){\text{\boldmath$e$}}_{2}\text{k}_{\omega(\theta)}.

Note that ω\omega is a diffeomorphism of /2π\mathbb{R}/2\pi\mathbb{Z} onto itself. Differentiating both sides of (3.14) with respect to θ\theta, and using the fact that ddθkθ=kθ+π2\frac{\text{d}}{\text{d}\theta}\text{k}_{\theta}=\text{k}_{\theta+\frac{\pi}{2}}, we have

(3.15) 𝒆1kθg1=(f(θ),ρ(θ))kω(θ),where f(θ):=ρ(θ)ω(θ).\displaystyle{\text{\boldmath$e$}}_{1}\text{k}_{\theta}g^{-1}=\bigl{(}f(\theta),\rho^{\prime}(\theta)\bigr{)}\text{k}_{\omega(\theta)},\qquad\text{where }\>f(\theta):=\rho(\theta)\omega^{\prime}(\theta).

It also follows from (3.15) and (3.14) that

f(θ)=𝒆1kω(θ)𝒆1kθg1=𝒆2kω(θ)kπ2𝒆1kθg1\displaystyle f(\theta)={\text{\boldmath$e$}}_{1}\text{k}_{\omega(\theta)}\cdot{\text{\boldmath$e$}}_{1}\text{k}_{\theta}g^{-1}={\text{\boldmath$e$}}_{2}\text{k}_{\omega(\theta)}\text{k}_{\frac{\pi}{2}}\cdot{\text{\boldmath$e$}}_{1}\text{k}_{\theta}g^{-1} =ρ(θ)1𝒆2kθg1kπ2𝒆1kθg1\displaystyle=\rho(\theta)^{-1}{\text{\boldmath$e$}}_{2}\text{k}_{\theta}g^{-1}\text{k}_{\frac{\pi}{2}}\cdot{\text{\boldmath$e$}}_{1}\text{k}_{\theta}g^{-1}
(3.16) =ρ(θ)1det(kθg1)=ρ(θ)1det(g)1,\displaystyle=\rho(\theta)^{-1}\det(\text{k}_{\theta}g^{-1})=\rho(\theta)^{-1}\det(g)^{-1},

where in the fourth equality we used the fact that 𝒘kπ2𝒗=v1w2v2w1{\text{\boldmath$w$}}\text{k}_{\frac{\pi}{2}}\cdot{\text{\boldmath$v$}}=v_{1}w_{2}-v_{2}w_{1} for any 𝒗=(v1,v2){\text{\boldmath$v$}}=(v_{1},v_{2}) and 𝒘=(w1,w2){\text{\boldmath$w$}}=(w_{1},w_{2}) in 2\mathbb{R}^{2}. Note that (3.15) and (3.16) imply that if det(g)>0\det(g)>0 then f(θ)>0f(\theta)>0 and ω(θ)>0\omega^{\prime}(\theta)>0 for all θ\theta, while if det(g)<0\det(g)<0 then f(θ)<0f(\theta)<0 and ω(θ)<0\omega^{\prime}(\theta)<0 for all θ\theta.

Using (3.14) and (3.15) we have (x,y)kθg1=(xf(θ),xρ(θ)+yρ(θ))kω(θ)(x,y)\text{k}_{\theta}g^{-1}=\bigl{(}xf(\theta),x\rho^{\prime}(\theta)+y\rho(\theta)\bigr{)}\text{k}_{\omega(\theta)} for any (x,y)2(x,y)\in\mathbb{R}^{2}, and hence via (3.1) and (3.2) one gets 𝒲g(θ)=f(θ)1𝒲(ω(θ))\ell_{{\mathcal{W}}g}(\theta)=f(\theta)^{-1}\ell_{{\mathcal{W}}}(\omega(\theta)). We also have ω(θ+π)=ω(θ)+π\omega(\theta+\pi)=\omega(\theta)+\pi, f(θ+π)=f(θ)f(\theta+\pi)=f(\theta), and 𝒲(ω(θ)+π)=𝒲(ω(θ))\ell_{{\mathcal{W}}}(\omega(\theta)+\pi)=-\ell_{{\mathcal{W}}}(\omega(\theta)), and hence:

𝒲g(θ)=|f(θ~)|1𝒲(ω(θ~)),where θ~:=θ if det(g)>0θ~:=θ+π if det(g)<0.\displaystyle\ell_{{\mathcal{W}}g}(\theta)=|f(\tilde{\theta})|^{-1}\ell_{{\mathcal{W}}}(\omega(\tilde{\theta})),\qquad\text{where $\tilde{\theta}:=\theta$ if $\det(g)>0$, $\tilde{\theta}:=\theta+\pi$ if $\det(g)<0$.}

Using also the simple relation αI,aJ(y)=αI,J(a1y)\alpha_{I,aJ}(y)=\alpha_{I,J}(a^{-1}y) (a>0\forall\,a>0), it follows that

Area(𝒲g)02π1λ2αI,𝒲g(θ)(y)2dyy3dθ=Area(𝒲g)/2π1λ2αI,𝒲(ω(θ~))(|f(θ~)|y)2dyy3dθ~\displaystyle\operatorname{Area}({\mathcal{W}}g)\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}\alpha_{I,\ell_{\mathcal{W}g}(\theta)}(y)^{2}\,\frac{\text{d}y}{y^{3}}\,\text{d}\theta=\operatorname{Area}({\mathcal{W}}g)\int_{\mathbb{R}/2\pi\mathbb{Z}}\int_{1}^{\lambda^{2}}\alpha_{I,\ell_{\mathcal{W}}(\omega(\tilde{\theta}))}\bigl{(}|f(\tilde{\theta})|y\bigr{)}^{2}\,\frac{\text{d}y}{y^{3}}\,\text{d}\tilde{\theta}
(3.17) =|detg|Area(𝒲)/2πf(θ~)21λ2αI,𝒲(ω(θ~))(u)2duu3dθ~,\displaystyle=|\det g|\operatorname{Area}({\mathcal{W}})\int_{\mathbb{R}/2\pi\mathbb{Z}}f(\tilde{\theta})^{2}\int_{1}^{\lambda^{2}}\alpha_{I,\ell_{\mathcal{W}}(\omega(\tilde{\theta}))}(u)^{2}\,\frac{\text{d}u}{u^{3}}\,\text{d}\tilde{\theta},

where we substituted u=|f(θ~)|yu=|f(\tilde{\theta})|y and then used (3.10). The two formulas for f(θ)f(\theta) in (3.15) and (3.16) imply that f(θ~)2=det(g)1ω(θ~)f(\tilde{\theta})^{2}=\det(g)^{-1}\omega^{\prime}(\tilde{\theta}). Using this in the last expression in (3.17), and then taking ω\omega as a new variable of integration, we obtain the identity (3.13). ∎

Let us also note that for any gSL2()g\in\operatorname{SL}_{2}(\mathbb{R}), the invariance statement in Remark 3.12 may alternatively be deduced as a consequence of (3.7) in Theorem 4 and Remark 3.35 below. Once the invariance is known for all gSL2()g\in\operatorname{SL}_{2}(\mathbb{R}), in order to extend it to all gGL2()g\in\operatorname{GL}_{2}(\mathbb{R}) it suffices to verify that it also holds for g=diag(1,1)g=\operatorname{diag}(1,-1) and g=diag(a,a)g=\operatorname{diag}(a,a) (a>0\forall\,a>0); for these cases the above proof applies in a significantly simplified form.

3.2. Proof of Theorem 2 using Theorem 4

In this section we prove Theorem 2 assuming Theorem 4. We will first prove the two tail asymptotics of G(s)G(s) in (1.13) and (1.14); then, by a standard general argument using the convexity of G(s)G(s), we will deduce the tail asymptotics of F(s)F(s).

We start with (1.13). We need to show that for any 𝒲{\mathcal{W}} as in Theorem 2, the error bound in (3.7) is o(s1)o(s^{-1}) as ss\to\infty. Take R>0R>0 so that 𝒲R2{\mathcal{W}}\subset{\mathcal{B}}_{R}^{2}, where R22{\mathcal{B}}_{R}^{2}\subset\mathbb{R}^{2} is the open disc with center 𝟎\bm{0} and radius RR. Then for every yy\in\mathbb{R} with |y|R|y|\geq R, and for every θ\theta, we have R𝒲(θ,y)=R_{{\mathcal{W}}}(\theta,y)=\emptyset and hence 𝔣~(R𝒲(θ,y),u1)=0\widetilde{\mathfrak{f}}\bigl{(}R_{{\mathcal{W}}}(\theta,y),u^{-1}\bigr{)}=0 for all u>0u>0. It follows that the integral over \mathbb{R} in (3.7) can be replaced by RR\int_{-R}^{R}. Hence, if we take C=C𝒫C=C_{\mathcal{P}} as in Theorem 4, and set

Fθ,y(s):=s10Cs𝔣~(R𝒲(θ,y),u1)du,\displaystyle F_{\theta,y}(s):=s^{-1}\int_{0}^{Cs}\widetilde{\mathfrak{f}}\bigl{(}R_{{\mathcal{W}}}(\theta,y),u^{-1}\bigr{)}\,\text{d}u,

then our task is to prove that 02πRRFθ,y(s)dydθ0\int_{0}^{2\pi}\int_{-R}^{R}F_{\theta,y}(s)\,\text{d}y\,\text{d}\theta\to 0 as ss\to\infty. Using 0𝔣~(R𝒲(θ,y),u1)10\leq\widetilde{\mathfrak{f}}\bigl{(}R_{{\mathcal{W}}}(\theta,y),u^{-1}\bigr{)}\leq 1 we have 0Fθ,y(s)C0\leq F_{\theta,y}(s)\leq C for all θ\theta, yy and ss; hence by the Lebesgue bounded convergence theorem, it suffices to prove that Fθ,y(s)0F_{\theta,y}(s)\to 0 as ss\to\infty for any fixed θ,y\theta,y. But if R𝒲(θ,y)=R_{{\mathcal{W}}}(\theta,y)=\emptyset then Fθ,y(s)=0F_{\theta,y}(s)=0 for all ss, by the definition (3.6), while if R𝒲(θ,y)R_{{\mathcal{W}}}(\theta,y)\neq\emptyset then since 𝒲{\mathcal{W}} is open, R𝒲(θ,y)R_{{\mathcal{W}}}(\theta,y) contains a non-empty interval, and so 𝔣~(R𝒲(θ,y),u1)=0\widetilde{\mathfrak{f}}\bigl{(}R_{{\mathcal{W}}}(\theta,y),u^{-1}\bigr{)}=0 for all sufficiently large uu; therefore Fθ,y(s)0F_{\theta,y}(s)\to 0 as ss\to\infty. This concludes the proof that the error bound in (3.7) is o(s1)o(s^{-1}) as ss\to\infty, i.e. we have proved that G(s)a𝒫s1G(s)\sim a_{{\mathcal{P}}}s^{-1}, as stated in (1.13).

For the proof of (1.14) we need the following simple lemma regarding the function 𝔣~(A,a)\widetilde{\mathfrak{f}}(A,a).

Lemma 3.1.

Let C>0C>0. Then for any interval AA of length b>0b>0 and for any s>0s>0,

0Cs𝔣~(A,u1)duCmin(s,b1).\displaystyle\int_{0}^{Cs}\widetilde{\mathfrak{f}}(A,u^{-1})\,\mathrm{d}u\asymp_{C}\min(s,b^{-1}).
Proof.

Note that 𝔣(A,a)=0\mathfrak{f}(A,a)=0 for a(0,b]a\in(0,b] and 𝔣(A,a)=1(b/a)\mathfrak{f}(A,a)=1-(b/a) for a[b,)a\in[b,\infty). Also 𝔣~(A,a)=𝔣(A,a)\widetilde{\mathfrak{f}}(A,a)=\mathfrak{f}(A,a). Hence for s(Cb)1s\leq(Cb)^{-1}:

0Cs𝔣~(A,u1)du=0Cs(1bu)duCsCs,\displaystyle\int_{0}^{Cs}\widetilde{\mathfrak{f}}(A,u^{-1})\,\text{d}u=\int_{0}^{Cs}\bigl{(}1-bu\bigr{)}\,\text{d}u\asymp Cs\asymp_{C}s,

while for s(Cb)1s\geq(Cb)^{-1}:

0Cs𝔣~(A,u1)du=0b1(1bu)du+0b1.\displaystyle\int_{0}^{Cs}\widetilde{\mathfrak{f}}(A,u^{-1})\,\text{d}u=\int_{0}^{b^{-1}}\bigl{(}1-bu\bigr{)}\,\text{d}u+0\asymp b^{-1}.

The desired estimate then follows from these two estimates noting that min(s,b1)Cs\min(s,b^{-1})\asymp_{C}s for s(Cb)1s\leq(Cb)^{-1} and min(s,b1)Cb1\min(s,b^{-1})\asymp_{C}b^{-1} for s(Cb)1s\geq(Cb)^{-1}. ∎

Now let us further assume that 𝒲\mathcal{W} is convex. Then for all θ\theta and yy, R𝒲(θ,y)R_{{\mathcal{W}}}(\theta,y) is either empty or an interval. Hence, using Lemma 3.1 and the fact that m(R𝒲(θ+π,y))=m(R𝒲(θ,y))m(R_{{\mathcal{W}}}(\theta+\pi,y))=m(R_{{\mathcal{W}}}(\theta,-y)), it follows that the error bound in (3.7) is

(3.18) s202π0I(R𝒲(θ,y))min(s,m(R𝒲(θ,y))1)dydθ.\displaystyle\asymp s^{-2}\int_{0}^{2\pi}\int_{0}^{\infty}I\bigl{(}R_{{\mathcal{W}}}(\theta,y)\neq\emptyset\bigr{)}\cdot\min\Bigl{(}s,m(R_{{\mathcal{W}}}(\theta,y))^{-1}\Bigr{)}\,\text{d}y\,\text{d}\theta.

Choose 0<r0R00<r_{0}\leq R_{0} such that r02𝒲R02{\mathcal{B}}_{r_{0}}^{2}\subset\mathcal{W}\subset{\mathcal{B}}_{R_{0}}^{2}. Let us temporarily fix θ[0,2π]\theta\in[0,2\pi], and set y0:=sup{y>0:R𝒲(θ,y)}y_{0}:=\sup\{y>0\>:\>R_{{\mathcal{W}}}(\theta,y)\neq\emptyset\}; then the function f(y):=m(R𝒲(θ,y))f(y):=m(R_{{\mathcal{W}}}(\theta,y)) vanishes for y>y0y>y_{0}, but is positive and concave for y[0,y0)y\in[0,y_{0}). Hence for our fixed θ\theta, the inner integral in (3.18) equals 0y0min(s,f(y)1)dy\int_{0}^{y_{0}}\min\bigl{(}s,f(y)^{-1}\bigr{)}\,\text{d}y. We also have r0y0R0r_{0}\leq y_{0}\leq R_{0} and f(0)r0f(0)\geq r_{0}, and so, by the concavity, f(y)(r0/R0)(y0y)f(y)\geq(r_{0}/R_{0})\cdot(y_{0}-y) for all y[0,y0]y\in[0,y_{0}]. Hence

0y0min(s,f(y)1)dy0R0min(s,R0r0t)dt,\displaystyle\int_{0}^{y_{0}}\min\bigl{(}s,f(y)^{-1}\bigr{)}\,\text{d}y\leq\int_{0}^{R_{0}}\min\Bigl{(}s,\frac{R_{0}}{r_{0}\,t}\Bigr{)}\,\text{d}t,

and by a simple computation, the last integral is seen to equal R0r0(log(r0s)+1){\displaystyle\frac{R_{0}}{r_{0}}}\bigl{(}\log(r_{0}s)+1\bigr{)} for all s1/r0s\geq 1/r_{0}. Since this holds for every θ[0,2π]\theta\in[0,2\pi], we conclude that the expression in (3.18) is s2logs\ll s^{-2}\log s for all large ss, i.e. we have proved the asymptotics for G(s)G(s) stated in (1.14).

The next lemma shows how to use the convexity of G(s)G(s) to deduce the tail asymptotics of F(s)=G(s)F(s)=-G^{\prime}(s) from that of G(s)G(s).

Lemma 3.2.

Let G:>0>0G:\mathbb{R}_{>0}\to\mathbb{R}_{>0} be a differentiable function with F(s):=G(s)F(s):=-G^{\prime}(s) continuous and decreasing. Suppose further that there exist constants a>0a>0 and C1,C2>1C_{1},C_{2}>1 and a function E:(C1,)>0E:(C_{1},\infty)\to\mathbb{R}_{>0}, such that

(3.19) |G(s)as1|E(s)1s,s>C1,\displaystyle\bigl{|}G(s)-as^{-1}\bigr{|}\leq E(s)\leq\frac{1}{s},\qquad\forall\,s>C_{1},

and

(3.20) C21E(s)E(s)C2whenever C1<ss2s.\displaystyle C_{2}^{-1}\leq\frac{E(s)}{E(s^{\prime})}\leq C_{2}\qquad\text{whenever $C_{1}<s\leq s^{\prime}\leq 2s$.}

Then we have

(3.21) F(s)=as2+O(s32E(s)12)as s,\displaystyle F(s)=as^{-2}+O(s^{-\frac{3}{2}}E(s)^{\frac{1}{2}})\quad\text{as $s\to\infty$},

where the implied constant depends only on aa and C2C_{2}.

Proof.

Since by assumption F(s)F(s) is continuous and decreasing, we have for all 0<s1<s20<s_{1}<s_{2}, G(s1)G(s2)=s1s2F(s)ds(s2s1)F(s1)G(s_{1})-G(s_{2})=\int_{s_{1}}^{s_{2}}F(s)\,\mathrm{d}s\leq(s_{2}-s_{1})F(s_{1}). Hence, assuming C1<s1<s22s1C_{1}<s_{1}<s_{2}\leq 2s_{1}, writing h:=s2s1h:=s_{2}-s_{1}, and using (3.19) and (3.20), we have

F(s1)a(s11s21)hE(s1)+E(s2)hC2+1hE(s1).\displaystyle F(s_{1})-\frac{a(s_{1}^{-1}-s_{2}^{-1})}{h}\geq-\frac{E(s_{1})+E(s_{2})}{h}\geq-\frac{C_{2}+1}{h}E(s_{1}).

From this, using also s11s21=h/(s1s2)s_{1}^{-1}-s_{2}^{-1}=h/(s_{1}s_{2}), we get

F(s1)as1s2C2+1hE(s1)as12ahs13C2+1hE(s1).\displaystyle F(s_{1})\geq\frac{a}{s_{1}s_{2}}-\frac{C_{2}+1}{h}E(s_{1})\geq\frac{a}{s_{1}^{2}}-\frac{ah}{s_{1}^{3}}-\frac{C_{2}+1}{h}E(s_{1}).

Here we optimize by choosing h=E(s1)s13h=\sqrt{E(s_{1})s_{1}^{3}}; note that this choice of hh is admissible, i.e. yields s22s1s_{2}\leq 2s_{1}, because of our assumption E(s)s1E(s)\leq s^{-1}. The conclusion is that F(s1)as12(a+C2+1)s132E(s1)12F(s_{1})-as_{1}^{-2}\geq-(a+C_{2}+1)s_{1}^{-\frac{3}{2}}E(s_{1})^{\frac{1}{2}} holds for any s1>C1s_{1}>C_{1}.

Similarly, using G(s1)G(s2)=s1s2F(s)ds(s2s1)F(s2)G(s_{1})-G(s_{2})=\int_{s_{1}}^{s_{2}}F(s)\,\mathrm{d}s\geq(s_{2}-s_{1})F(s_{2}), we have for all C1<s1<s22s1C_{1}<s_{1}<s_{2}\leq 2s_{1}:

F(s2)as1s2+E(s1)+E(s2)has22+2ahs23+C2+1hE(s2).\displaystyle F(s_{2})\leq\frac{a}{s_{1}s_{2}}+\frac{E(s_{1})+E(s_{2})}{h}\leq\frac{a}{s_{2}^{2}}+\frac{2ah}{s_{2}^{3}}+\frac{C_{2}+1}{h}E(s_{2}).

Here choose h=41E(s2)s23h=\sqrt{4^{-1}E(s_{2})s_{2}^{3}}; then h41s22=21s2h\leq\sqrt{4^{-1}s_{2}^{2}}=2^{-1}s_{2}, so that our choice is admissible for any s2>2C1s_{2}>2C_{1}. For these s2s_{2}, we obtain F(s2)as22(a+2C2+2)s232E(s2)12F(s_{2})-as_{2}^{-2}\leq(a+2C_{2}+2)s_{2}^{-\frac{3}{2}}E(s_{2})^{\frac{1}{2}}. Combining this with the bound proved above, we have proved (3.21) (with an implied constant C=a+2C2+2C=a+2C_{2}+2). ∎

Let us now use Lemma 3.2 to conclude the proof of (1.13). Recall that in the situation of Theorem 2, F(s)F(s) is continuous and decreasing for s>0s>0 (see Theorem 1), and G(s)=sF(t)dtG(s)=\int_{s}^{\infty}F(t)\,\text{d}t. Furthermore, we have proved the first half of (1.13), i.e. that G(s)a𝒫s1G(s)\sim a_{{\mathcal{P}}}s^{-1} as ss\to\infty. Hence, for any given constant 0<c<10<c<1, Lemma 3.2 applies with E(s)=cs1E(s)=cs^{-1}, C2=2C_{2}=2, and an appropriate constant C1C_{1} (depending on cc), to yield F(s)=a𝒫s2+O(c1/2s2)F(s)=a_{\mathcal{P}}s^{-2}+O(c^{1/2}s^{-2}) as ss\to\infty. Here the implied constant is independent of cc, and by taking cc arbitrarily small we conclude that F(s)=a𝒫s2+o(s2)F(s)=a_{\mathcal{P}}s^{-2}+o(s^{-2}). This completes the proof of (1.13).

Similarly, the last part of (1.14) follows from the first part of (1.14), by Lemma 3.2 applied with appropriate constants C1,C2C_{1},C_{2} and E(s)=Cs2logsE(s)=Cs^{-2}\log s with an appropriate C>0C>0. This concludes the proof of Theorem 2. \square

Remark 3.22.

Let us note that there exist quite “nice” (but non-convex) windows 𝒲{\mathcal{W}} for which the relative error bound in (3.7) in Theorem 4 tends to zero more slowly than any prescribed rate, as ss\to\infty. One way in which this can happen is if 𝒲{\mathcal{W}} has a cusp of an appropriate asymptotic shape.

To give a concrete statement, let us define

(3.23) H𝒲(s):=1s0C𝒫s/2π𝔣~(R𝒲(θ,y),u1)dydθdu.\displaystyle H_{{\mathcal{W}}}(s):=\frac{1}{s}\int_{0}^{C_{{\mathcal{P}}}s}\int_{\mathbb{R}/2\pi\mathbb{Z}}\int_{\mathbb{R}}\widetilde{\mathfrak{f}}\bigl{(}R_{{\mathcal{W}}}(\theta,y),u^{-1}\bigr{)}\,\text{d}y\,\text{d}\theta\,\text{d}u.

Then (3.7) says that G(s)=a𝒫s1(1+O(H𝒲(s)))G(s)=a_{{\mathcal{P}}}s^{-1}\bigl{(}1+O(H_{{\mathcal{W}}}(s))\bigr{)} as ss\to\infty, viz., H𝒲(s)H_{\mathcal{W}}(s) is the relative error bound in this asymptotics, and our proof of the relation G(s)a𝒫s1G(s)\sim a_{{\mathcal{P}}}s^{-1} in (1.13) shows that H𝒲(s)0H_{\mathcal{W}}(s)\to 0 as ss\to\infty. Now let H0:[1,)(0,)H_{0}:[1,\infty)\to(0,\infty) be an arbitrary decreasing function with limsH0(s)=0\lim_{s\to\infty}H_{0}(s)=0. Choose a sequence 1>t2>t3>1>t_{2}>t_{3}>\cdots satisfying both limntn+1/H0(n)=\lim_{n\to\infty}t_{n+1}/H_{0}(n)=\infty and limntn=0\lim_{n\to\infty}t_{n}=0,111For example, one may choose tn+1:=31(1+n1)H0(n)/H0(1)t_{n+1}:=3^{-1}(1+n^{-1})\sqrt{H_{0}(n)/H_{0}(1)}. and then let f:[0,1][0,1]f:[0,1]\to[0,1] be a continuous, increasing function satisfying f(0)=0f(0)=0, f(1)=1f(1)=1, and f(tn)=n1f(t_{n})=n^{-1} for all n2n\geq 2. (Note that we may even take f(s)f(s) to be smooth for 0<s10<s\leq 1.) Set

𝒲:=𝒲0+𝒘with𝒲0:={(w1,w2)2: 0<w1<1, 0<w2<f(w1)},\displaystyle{\mathcal{W}}:={\mathcal{W}}_{0}+{\text{\boldmath$w$}}\quad\text{with}\quad{\mathcal{W}}_{0}:=\{(w_{1},w_{2})\in\mathbb{R}^{2}\>:\>0<w_{1}<1,\>0<w_{2}<f(w_{1})\},

and with 𝒘w being a fixed vector in 2\mathbb{R}^{2} chosen so that 𝟎𝒲=𝒲\mathbf{0}\in{\mathcal{W}}={\mathcal{W}}^{\circ}. Finally let 𝒫{\mathcal{P}} be the cut-and-project set 𝒫=𝒫(𝒲,K){\mathcal{P}}={\mathcal{P}}({\mathcal{W}},{\mathcal{L}}_{K}), for any fixed real quadratic field KK. We claim that in this case, limsH𝒲(s)/H0(s)=\lim_{s\to\infty}H_{{\mathcal{W}}}(s)/H_{0}(s)=\infty, viz., H𝒲(s)H_{\mathcal{W}}(s) tends to zero more slowly than the given function H0(s)H_{0}(s)!

To prove the claim, we first note that for every π4<θ<0-\frac{\pi}{4}<\theta<0, writing wθ,1w_{\theta,1} for the first coordinate of 𝒘kθ{\text{\boldmath$w$}}\text{k}_{-\theta}, one easily verifies that for every 0<y<1/20<y<1/2, R𝒲(θ,y+wθ,1)R_{{\mathcal{W}}}(\theta,y+w_{\theta,1}) is a non-empty open interval of length (cosθ)1f(y/cosθ)2f(y/cosθ)\leq(\cos\theta)^{-1}f(y/\cos\theta)\leq\sqrt{2}f(y/\cos\theta). Hence by Lemma 3.1, for each such θ\theta,

0C𝒫s𝔣~(R𝒲(θ,y),u1)dudy𝒫01/2min(s,f(y/cosθ)1)dy01/2min(s,f(y)1)dy.\displaystyle\int_{\mathbb{R}}\int_{0}^{C_{{\mathcal{P}}}s}\widetilde{\mathfrak{f}}(R_{{\mathcal{W}}}(\theta,y),u^{-1})\,\text{d}u\,\text{d}y\gg_{{\mathcal{P}}}\int_{0}^{1/2}\min(s,f(y/\cos\theta)^{-1})\,\text{d}y\gg\int_{0}^{1/2}\min(s,f(y)^{-1})\,\text{d}y.

This implies that

(3.24) H𝒲(s)𝒫1s01/2min(s,f(y)1)dy=01/2min(1,1sf(y))dy=:H1(s).\displaystyle H_{\mathcal{W}}(s)\gg_{{\mathcal{P}}}\>\frac{1}{s}\int_{0}^{1/2}\min(s,f(y)^{-1})\,\text{d}y=\int_{0}^{1/2}\min\Bigl{(}1,\frac{1}{sf(y)}\Bigr{)}\,\text{d}y=:H_{1}(s).

Clearly H1(s)H_{1}(s) is a decreasing function of s>0s>0. Also, for each n2n\geq 2 we have f(y)f(tn)=n1f(y)\leq f(t_{n})=n^{-1} for all y[0,tn]y\in[0,t_{n}], and so H1(n)0tn1dy=tnH_{1}(n)\geq\int_{0}^{t_{n}}1\,\text{d}y=t_{n}. Hence by our choice of {tn}\{t_{n}\}, we have limnH1(n+1)/H0(n)=\lim_{n\to\infty}H_{1}(n+1)/H_{0}(n)=\infty. Note also that for every n1n\geq 1 and every s[n,n+1]s\in[n,n+1], we have H1(s)/H0(s)H1(n+1)/H0(n)H_{1}(s)/H_{0}(s)\geq H_{1}(n+1)/H_{0}(n); hence limsH1(s)/H0(s)=\lim_{s\to\infty}H_{1}(s)/H_{0}(s)=\infty. In view of (3.24), this implies that limsH𝒲(s)/H0(s)=\lim_{s\to\infty}H_{\mathcal{W}}(s)/H_{0}(s)=\infty. \square

As an interesting open problem, we mention that it seems plausible to us that, by using windows 𝒲{\mathcal{W}} as above (perhaps with some further restrictions), also the actual relative difference, G(s)sa𝒫G(s)s-a_{{\mathcal{P}}}, can be proved to tend to zero more slowly than any prescribed rate.

Remark 3.25.

On the other hand, there exist 𝒲{\mathcal{W}} with fractal boundary for which (3.7) in Theorem 4 implies as strong an error bound as for convex window sets, i.e. G(s)=a𝒫s1+O(s2logs)G(s)=a_{{\mathcal{P}}}s^{-1}+O\bigl{(}s^{-2}\log s\bigr{)} as in (1.14) in Theorem 2. For example, this holds if 𝒲\partial{\mathcal{W}} is the standard Koch snowflake; and it appears to also hold for the window sets appearing in, e.g., [12], [33] and for several of the specific examples in [34]. Indeed, the proof of Lemma 3.1 shows that if the set AA\subset\mathbb{R} contains an interval of length b>0b>0 then 0Cs𝔣~(A,u1)duCmin(s,b1)\int_{0}^{Cs}\widetilde{\mathfrak{f}}(A,u^{-1})\,\text{d}u\ll_{C}\min(s,b^{-1}) for all s>0s>0. Hence by the argument leading to (3.18), the error bound in (3.7) is

s202π0I(m(R𝒲(θ,y))>0)min(s,j(R𝒲(θ,y))1)dydθ,\displaystyle\ll s^{-2}\int_{0}^{2\pi}\int_{0}^{\infty}I\bigl{(}m(R_{{\mathcal{W}}}(\theta,y))>0\bigr{)}\cdot\min\Bigl{(}s,j(R_{{\mathcal{W}}}(\theta,y))^{-1}\Bigr{)}\,\text{d}y\,\text{d}\theta,

where for AA\subset\mathbb{R} we write j(A)j(A) for the supremum of the lengths of all intervals contained in AA. Now temporarily fix θ[0,2π]\theta\in[0,2\pi] and write y0:=sup{y>0:R𝒲(θ,y)}y_{0}:=\sup\{y>0\>:\>R_{{\mathcal{W}}}(\theta,y)\neq\emptyset\}. One verifies that, if 𝒲{\mathcal{W}} is an open set (with 𝟎𝒲\mathbf{0}\in{\mathcal{W}}) such that 𝒲\partial{\mathcal{W}} is the standard Koch snowflake, then there exists a constant c>0c>0 independent of θ\theta such that j(R𝒲(θ,y))>c(y0y)j(R_{{\mathcal{W}}}(\theta,y))>c\cdot(y_{0}-y) for all y[0,y0]y\in[0,y_{0}]. This leads to G(s)=a𝒫s1+O(s2logs)G(s)=a_{{\mathcal{P}}}s^{-1}+O\bigl{(}s^{-2}\log s\bigr{)}, as claimed.

3.3. A preliminary integral formula for G(s)G(s)

The remainder of Section 3 is devoted to proving Theorem 4. We will start by recalling from [27] an expression for the function G(s):=sF(t)dtG(s):=\int_{s}^{\infty}F(t)\,\text{d}t as the Haar measure of a certain set in a homogeneous space.

We first work in a slightly more general situation, namely, we assume that the given cut-and-project set 𝒫{\mathcal{P}} is as in (1.4) with d=2d=2; in particular {\mathcal{L}} is a lattice in n\mathbb{R}^{n} where n=2+mn=2+m. Set G=SLn()\mathrm{G}=\operatorname{SL}_{n}(\mathbb{R}) and Γ=SLn()\Gamma=\operatorname{SL}_{n}(\mathbb{Z}), and choose gGg\in\mathrm{G} and δ>0\delta>0 so that =δ1/nng{\mathcal{L}}=\delta^{1/n}\mathbb{Z}^{n}g. Let φg\varphi_{g} be the embedding of SL2()\operatorname{SL}_{2}(\mathbb{R}) in G\mathrm{G} given by

φg(A)=g(A00Im)g1.\displaystyle\varphi_{g}(A)=g\left(\begin{matrix}A&0\\ 0&I_{m}\end{matrix}\right)g^{-1}.

It follows from Ratner’s work [36, 37] that there exists a unique closed connected subgroup Hg\mathrm{H}_{g} of G\mathrm{G} such that ΓHg\Gamma\cap\mathrm{H}_{g} is a lattice in Hg\mathrm{H}_{g}, φg(SL2())Hg\varphi_{g}(\operatorname{SL}_{2}(\mathbb{R}))\subset\mathrm{H}_{g}, and the closure of Γ\Γφg(SL2())\Gamma\backslash\Gamma\varphi_{g}(\operatorname{SL}_{2}(\mathbb{R})) in Γ\G\Gamma\backslash\mathrm{G} equals Γ\ΓHg\Gamma\backslash\Gamma\mathrm{H}_{g}. Let μg\mu_{g} be the Haar measure on Hg\mathrm{H}_{g} normalized so that μg(Γ\ΓHg)=1\mu_{g}(\Gamma\backslash\Gamma\mathrm{H}_{g})=1. Now F(s)F(s) in Theorem 1 is given by [27, (3.6) and (11.1)]:

(3.26) F(s)=G(s)(s>0)\displaystyle F(s)=-G^{\prime}(s)\qquad(s>0)

with

(3.27) G(s)=μg({ΓhΓ\ΓHg:𝒫(𝒲,δ1/nnhg)(s)=}),\displaystyle G(s)=\mu_{g}\bigl{(}\bigl{\{}\Gamma h\in\Gamma\backslash\Gamma\mathrm{H}_{g}\>:\>{\mathcal{P}}({\mathcal{W}},\delta^{1/n}\mathbb{Z}^{n}hg)\cap\mathfrak{C}(s)=\emptyset\bigr{\}}\bigr{)},

where (s)\mathfrak{C}(s) is the open triangle in 2\mathbb{R}^{2} with vertices at (0,0)(0,0) and (1,±s/c𝒫)(1,\pm s/c_{{\mathcal{P}}}). (Recall that c𝒫c_{{\mathcal{P}}} denotes the asymptotic density of 𝒫{\mathcal{P}}; see (1.2).) It is known that G(s)G(s) is C1C^{1} on >0\mathbb{R}_{>0}, and convex; hence F(s)F(s) is continuous and decreasing [27, Sec. 11].

Remark 3.28.

Recall from Remark 1.9 that our function F(s)F(s) equals the function in the right hand side of [27, (1.15)]. This is the reason why we have (s)\mathfrak{C}(s) in (3.27), and not (κ𝒫1s)\mathfrak{C}(\kappa_{{\mathcal{P}}}^{-1}s) as in [27, (3.6)].

Remark 3.29.

It follows from the formula (3.27) that G(s)G(s) (and hence also F(s)F(s)) remains unchanged if 𝒲\mathcal{W} is modified by a null set (with respect to the Haar measure μ𝒜\mu_{\mathcal{A}} of 𝒜=πint()¯m\mathcal{A}=\overline{\pi_{\operatorname{int}}({\mathcal{L}})}\subset\mathbb{R}^{m}). Indeed, let 𝒲1𝒜\mathcal{W}_{1}\subset\mathcal{A} be any bounded subset of 𝒜\mathcal{A} with boundary of measure zero, satisfying μ𝒜(𝒲𝒲1)=0\mu_{\mathcal{A}}(\mathcal{W}\triangle\mathcal{W}_{1})=0, and let G1(s)G_{1}(s) be given by (3.27) but with 𝒲1\mathcal{W}_{1} in place of 𝒲\mathcal{W}. Let s>0s>0 be given, and let fs:d×𝒜{0,1}f_{s}:\mathbb{R}^{d}\times{\mathcal{A}}\to\{0,1\} be the indicator function of the set δ1n((s)×(𝒲𝒲1))\delta^{-\frac{1}{n}}\bigl{(}\mathfrak{C}(s)\times(\mathcal{W}\triangle\mathcal{W}_{1})\bigr{)}. Then

|G1(s)G(s)|\displaystyle|G_{1}(s)-G(s)|\hskip 390.0pt
μg({Γh:exactly one of the sets 𝒫(𝒲,δ1/nnhg) and 𝒫(𝒲1,δ1/nnhg) intersects (s)})\displaystyle\hskip 10.0pt\leq\mu_{g}\bigl{(}\bigl{\{}\Gamma h\>:\>\text{exactly one of the sets ${\mathcal{P}}({\mathcal{W}},\delta^{1/n}\mathbb{Z}^{n}hg)$ and ${\mathcal{P}}({\mathcal{W}}_{1},\delta^{1/n}\mathbb{Z}^{n}hg)$ intersects $\mathfrak{C}(s)$}\bigr{\}}\bigr{)}
(3.30) Γ\ΓHg𝒎nhgfs(𝒎)dμg(h)=0,\displaystyle\leq\int_{\Gamma\backslash\Gamma\mathrm{H}_{g}}\sum_{\begin{subarray}{c}{\text{\boldmath$m$}}\in\mathbb{Z}^{n}hg\end{subarray}}f_{s}({\text{\boldmath$m$}})\,\text{d}\mu_{g}(h)=0,\hskip 230.0pt

where the final equality holds by the Siegel formula [26, Thm. 5.1],222Note the correction of this formula given in the erratum to [26]; note also that we may restrict the summation over 𝒎m in (3.30) by requiring π(𝒎)𝟎\pi({\text{\boldmath$m$}})\neq\mathbf{0}, since 𝟎(s)\mathbf{0}\notin\mathfrak{C}(s). since μ𝒜(𝒲𝒲1)=0\mu_{\mathcal{A}}(\mathcal{W}\triangle\mathcal{W}_{1})=0.

Because of the Hg\mathrm{H}_{g}-invariance of μg\mu_{g}, the formula (3.27) remains valid if in the right hand side we replace “δ1/nnhg\delta^{1/n}\mathbb{Z}^{n}hg” by “δ1/nnhφg(A)g\delta^{1/n}\mathbb{Z}^{n}h\varphi_{g}(A)g”, for any given ASL2()A\in\operatorname{SL}_{2}(\mathbb{R}); and applying this with A=diag[(s/c𝒫)1/2,(s/c𝒫)1/2]A=\operatorname{diag}[(s/c_{\mathcal{P}})^{-1/2},(s/c_{\mathcal{P}})^{1/2}] we obtain:

(3.31) G(s)=μg({ΓhΓ\ΓHg:𝒫(𝒲,δ1/nnhg)T(s)=}),\displaystyle G(s)=\mu_{g}\bigl{(}\bigl{\{}\Gamma h\in\Gamma\backslash\Gamma\mathrm{H}_{g}\>:\>{\mathcal{P}}({\mathcal{W}},\delta^{1/n}\mathbb{Z}^{n}hg)\cap T(s)=\emptyset\bigr{\}}\bigr{)},

where T(s)2T(s)\subset\mathbb{R}^{2} is the open triangle with vertices at (0,0)(0,0) and (s/c𝒫)1/2(1,±1)(s/c_{\mathcal{P}})^{1/2}(1,\pm 1).

Next we specialize to the setting of Theorem 4. Thus we take m=2m=2 and let K{\mathcal{L}}_{K} be as in (1.10); this means that n=4n=4, G=SL4()\mathrm{G}=\operatorname{SL}_{4}(\mathbb{R}) and Γ=SL4()\Gamma=\operatorname{SL}_{4}(\mathbb{Z}), and that gGg\in\mathrm{G} and δ>0\delta>0 are such that K=δ1/44g{\mathcal{L}}_{K}=\delta^{1/4}\mathbb{Z}^{4}g. Then by [26, Sec. 2.2.1] we have Hg=gHg1\mathrm{H}_{g}=g\mathrm{H}g^{-1}, where H\mathrm{H} is as in Section 2.2. Furthermore,

(3.32) ΓK=g1ΓgH.\displaystyle\Gamma_{K}=g^{-1}\Gamma g\cap\mathrm{H}.

Indeed, we have ΓKH\Gamma_{K}\subset\mathrm{H}, and every γΓK\gamma\in\Gamma_{K} satisfies Kγ=K{\mathcal{L}}_{K}\gamma={\mathcal{L}}_{K}, hence 4gγ=4g\mathbb{Z}^{4}g\gamma=\mathbb{Z}^{4}g and so γg1Γg\gamma\in g^{-1}\Gamma g. Conversely, assume γg1ΓgH\gamma\in g^{-1}\Gamma g\cap\mathrm{H}, and write γ=((abcd),(abcd))\gamma=\left(\left(\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}\right),\left(\begin{smallmatrix}a^{\prime}&b^{\prime}\\ c^{\prime}&d^{\prime}\end{smallmatrix}\right)\right). Then γg1Γg\gamma\in g^{-1}\Gamma g implies that 4gγg1=4\mathbb{Z}^{4}g\gamma g^{-1}=\mathbb{Z}^{4}, and so Kγ=K{\mathcal{L}}_{K}\gamma={\mathcal{L}}_{K}. It follows in particular that (1,0,1,0)γ(1,0,1,0)\gamma and (0,1,0,1)γ(0,1,0,1)\gamma lie in K{\mathcal{L}}_{K}, and this in turn implies that a,b,c,d𝒪Ka,b,c,d\in{\mathcal{O}}_{K} and a=σ(a)a^{\prime}=\sigma(a), b=σ(b)b^{\prime}=\sigma(b), c=σ(c)c^{\prime}=\sigma(c), d=σ(d)d^{\prime}=\sigma(d). Hence γΓK\gamma\in\Gamma_{K}, and (3.32) is proved.

It follows that the map ΓKhΓghg1\Gamma_{K}h\mapsto\Gamma ghg^{-1} (hHh\in\mathrm{H}) is a diffeomorphism from Γ\ΓHg\Gamma\backslash\Gamma\mathrm{H}_{g} onto ΓK\H\Gamma_{K}\backslash\mathrm{H}, carrying μg\mu_{g} to μK\mu_{K}. Hence (3.31) can be rewritten as

(3.33) G(s)=μK({ΓKhΓK\H:𝒫(𝒲,Kh)T(s)=}).\displaystyle G(s)=\mu_{K}\bigl{(}\bigl{\{}\Gamma_{K}h\in\Gamma_{K}\backslash\mathrm{H}\>:\>{\mathcal{P}}({\mathcal{W}},{\mathcal{L}}_{K}h)\cap T(s)=\emptyset\bigr{\}}\bigr{)}.

Here let us also note that 𝒫(𝒲,Kh)T(s)={\mathcal{P}}({\mathcal{W}},{\mathcal{L}}_{K}h)\cap T(s)=\emptyset is equivalent with Kh𝒯(s)={\mathcal{L}}_{K}h\cap{\mathcal{T}}^{\prime}(s)=\emptyset, where

𝒯(s):=T(s)×𝒲4.\displaystyle{\mathcal{T}}^{\prime}(s):=T(s)\times{\mathcal{W}}\subset\mathbb{R}^{4}.

Hence (3.33) can be rewritten as follows:

(3.34) G(s)\displaystyle G(s) :=XI(Kh𝒯(s)=)dμK(h).\displaystyle:=\int_{X}I\left(\mathcal{L}_{K}h\cap\mathcal{T}^{\prime}(s)=\emptyset\right)\,\text{d}\mu_{K}(h).

This formula will be the starting point of our proof of Theorem 4. Since we will be only concerned about the asymptotics of G(s)G(s) for large s, in the remainder of this section we will always assume s>1s>1.

Remark 3.35.

Using (3.34) and the fact that (I2,g)H(I_{2},g)\in\mathrm{H} for any gSL2()g\in\operatorname{SL}_{2}(\mathbb{R}), it is easy to verify that G(s)G(s) is unchanged if 𝒲\mathcal{W} is replaced by 𝒲g\mathcal{W}g for any gSL2()g\in\operatorname{SL}_{2}(\mathbb{R}).

3.4. Separating the main term and error term

We will apply results from Section 2.5 to further analyze the condition Kh𝒯(s)=\mathcal{L}_{K}h\cap\mathcal{T}^{\prime}(s)=\emptyset. For this we first renormalize the sets 𝒯(s)\mathcal{T}^{\prime}(s) to produce sets containing large balls (for large ss).

Lemma 3.3.

For s>1s>1 let rr\in\mathbb{N} be such that λ2rs1/4<λ2r+2\lambda^{2r}\leq s^{1/4}<\lambda^{2r+2}, and let

𝒯(s):=(λ2rT(s))×(λ2r𝒲).\displaystyle\mathcal{T}(s):=(\lambda^{-2r}T(s))\times(\lambda^{2r}\mathcal{W}).

Then for any hHh\in\mathrm{H}, Kh𝒯(s)=\mathcal{L}_{K}h\cap\mathcal{T}^{\prime}(s)=\emptyset holds if and only if Kh𝒯(s)=\mathcal{L}_{K}h\cap\mathcal{T}(s)=\emptyset.

Proof.

Let d2r=diag(λ2r,λ2r,λ2r,λ2r)d_{2r}=\operatorname{diag}(\lambda^{-2r},\lambda^{-2r},\lambda^{2r},\lambda^{2r}). The lemma then follows immediately by noting that 𝒯(s)d2r=𝒯(s)\mathcal{T}^{\prime}(s)d_{2r}=\mathcal{T}(s) and Khd2r=Kd2rh=Kh\mathcal{L}_{K}hd_{2r}=\mathcal{L}_{K}d_{2r}h=\mathcal{L}_{K}h. ∎

Since 𝒯(s)\mathcal{T}(s) contains a ball of radius s1/4\gg s^{1/4}, Proposition 2.3 gives that for ss sufficiently large, the condition Kh𝒯(s)=\mathcal{L}_{K}h\cap\mathcal{T}(s)=\emptyset forces ΓKhΓKi(t1)\Gamma_{K}h\in\Gamma_{K}\mathcal{F}_{i}(t_{1}) for some 1iκ1\leq i\leq\kappa, that is, the point ΓKh\Gamma_{K}h in XX belongs to the ii-th cuspidal neighborhood. Using also the fact that these cuspidal neighborhoods are pairwise disjoint, by (2.18), it follows that for ss sufficiently large we have

G(s)=i=1κGi(s)\displaystyle G(s)=\sum_{i=1}^{\kappa}G_{i}(s)

where for each 1iκ1\leq i\leq\kappa,

(3.36) Gi(s):=XI(Kh𝒯(s)= and KhΓKi(t1))dμK(h).\displaystyle G_{i}(s):=\int_{X}I\bigl{(}\mathcal{L}_{K}h\cap\mathcal{T}(s)=\emptyset\,\text{ and }\,\mathcal{L}_{K}h\in\Gamma_{K}\mathcal{F}_{i}(t_{1})\bigr{)}\,\text{d}\mu_{K}(h).

For the remainder of this section, we fix an index i{1,,κ}i\in\{1,\ldots,\kappa\}, and seek an asymptotic formula for the function Gi(s)G_{i}(s) as ss\to\infty.

First we note that since i(t1){\mathcal{F}}_{i}(t_{1}) projects injectively into XX (by (2.18) applied with j=ij=i), we may express Gi(s)G_{i}(s) as an integral over i(t1){\mathcal{F}}_{i}(t_{1}). In view of (2.12), (2.16) and (2.17), we get:

(3.37) Gi(s)=cKZYt1𝔉iI(Kξin𝒙a𝒚k𝜽𝒯(s)=)y13y23d𝒙d𝒚d𝜽,\displaystyle G_{i}(s)=c_{K}\int_{Z}\int_{Y_{t_{1}}}\int_{\mathfrak{F}_{i}}I\left(\mathcal{L}_{K}\xi_{i}\text{n}_{\bm{x}}\text{a}_{\bm{y}}\text{k}_{\bm{\theta}}\cap\mathcal{T}(s)=\emptyset\right)y_{1}^{-3}y_{2}^{-3}\,\text{d}\bm{x}\,\text{d}\bm{y}\,\text{d}\bm{\theta},

where cK=ΔK3/2ζK(2)1c_{K}=\Delta_{K}^{-3/2}\zeta_{K}(2)^{-1} as in (2.13),

Z:=(0,π)×(0,2π),\displaystyle Z:=(0,\pi)\times(0,2\pi),

and

(3.38) Yt1:={(y1,y2)(>0)2: 1y1/y2<λ2,y1y2>t1}.\displaystyle Y_{t_{1}}:=\left\{(y_{1},y_{2})\in(\mathbb{R}_{>0})^{2}\>:\>1\leq y_{1}/y_{2}<\lambda^{2},\ y_{1}y_{2}>t_{1}\right\}.

Now for ss sufficiently large, let h=ξin𝒙a𝒚k𝜽i(t1)h=\xi_{i}\text{n}_{\bm{x}}\text{a}_{\bm{y}}\text{k}_{\bm{\theta}}\in\mathcal{F}_{i}(t_{1}) be such that Kh𝒯(s)=\mathcal{L}_{K}h\cap\mathcal{T}(s)=\emptyset. Then by Proposition 2.3 we have y1,y2s1/4y_{1},y_{2}\gg s^{1/4}. As discussed in Section 2.5, since Kh\mathcal{L}_{K}h avoids a large ball and hi(t1)h\in\mathcal{F}_{i}(t_{1}), it contains a rank two sublattice ih\mathcal{L}_{i}h (cf. (2.24)) which sits densely in the corresponding filled plane Πih\Pi_{i}h (cf. (2.23)). Let us write

(3.39) ~i:=𝒗K(Πi+𝒗)={(α,t1,σ(α),t2):αIi,t1,t2}ξi1.\displaystyle\widetilde{\mathcal{L}}_{i}:=\bigcup_{\bm{v}\in\mathcal{L}_{K}}(\Pi_{i}+\bm{v})=\left\{(\alpha,t_{1},\sigma(\alpha),t_{2})\>:\>\alpha\in I_{i},\ t_{1},t_{2}\in\mathbb{R}\right\}\xi^{-1}_{i}.

Then Kh~ih\mathcal{L}_{K}h\subset\widetilde{\mathcal{L}}_{i}h, and ~ih\widetilde{\mathcal{L}}_{i}h is the union of all Kh\mathcal{L}_{K}h-translates of the filled plane Πih\Pi_{i}h. Here Ii=ai,ciI_{i}=\langle a_{i},c_{i}\rangle is as before and the second equality in (3.39) follows from (2.22) and (2.23). Our strategy of computing Gi(s)G_{i}(s) is to replace the condition Kh𝒯(s)=\mathcal{L}_{K}h\cap\mathcal{T}(s)=\emptyset by the slightly stronger (since ~ih\widetilde{\mathcal{L}}_{i}h is very close to Kh\mathcal{L}_{K}h when ss is sufficiently large) and more manageable condition ~ih𝒯(s)=\widetilde{\mathcal{L}}_{i}h\cap\mathcal{T}(s)=\emptyset. Thus we define, for s>1s>1:

(3.40) GM,i(s):=cKZYt1𝔉iI(~ih𝒯(s)=)y13y23d𝒙d𝒚d𝜽,\displaystyle G_{M,i}(s):=c_{K}\int_{Z}\int_{Y_{t_{1}}}\int_{\mathfrak{F}_{i}}I\left(\widetilde{\mathcal{L}}_{i}h\cap\mathcal{T}(s)=\emptyset\right)y_{1}^{-3}y_{2}^{-3}\,\text{d}\bm{x}\,\text{d}\bm{y}\,\text{d}\bm{\theta},

and

(3.41) GE,i(s):=cKZYt1𝔉iI(Kh𝒯(s)=,~ih𝒯(s))y13y23d𝒙d𝒚d𝜽,\displaystyle G_{E,i}(s):=c_{K}\int_{Z}\int_{Y_{t_{1}}}\int_{\mathfrak{F}_{i}}I\left(\mathcal{L}_{K}h\cap\mathcal{T}(s)=\emptyset,\ \widetilde{\mathcal{L}}_{i}h\cap\mathcal{T}(s)\neq\emptyset\right)y_{1}^{-3}y_{2}^{-3}\,\text{d}\bm{x}\,\text{d}\bm{y}\,\text{d}\bm{\theta},

where in both integrals we use the notation h=ξin𝒙a𝒚k𝜽h=\xi_{i}\text{n}_{\bm{x}}\text{a}_{\bm{y}}\text{k}_{\bm{\theta}}. It is clear from the definitions that Gi(s)=GM,i(s)+GE,i(s)G_{i}(s)=G_{M,i}(s)+G_{E,i}(s).

We start with a few auxiliary lemmas. For θ\theta\in\mathbb{R}, let (θ)\ell(\theta) be the projection of T(1)kθT(1)\text{k}_{-\theta} on the xx-axis. Recall also that 𝒲(θ)\ell_{{\mathcal{W}}}(\theta) is the projection of 𝒲kθ\mathcal{W}\text{k}_{-\theta} on the xx-axis.

Lemma 3.4.

Fix s>1s>1 and let rr\in\mathbb{N} be such that λ2rs1/4<λ2r+2\lambda^{2r}\leq s^{1/4}<\lambda^{2r+2} as before. For h=ξin𝐱a𝐲k𝛉i(t1)h=\xi_{i}\text{n}_{\bm{x}}\text{a}_{\bm{y}}\text{k}_{\bm{\theta}}\in\mathcal{F}_{i}(t_{1}), the condition ~ih𝒯(s)=\widetilde{\mathcal{L}}_{i}h\cap\mathcal{T}(s)=\emptyset holds if and only if

(3.42) ι(Ii)(y11s1/2λ2r(θ1))×(y21λ2r𝒲(θ2))=,\displaystyle\iota(I_{i})\cap(y_{1}^{-1}s^{1/2}\lambda^{-2r}\ell(\theta_{1}))\times(y_{2}^{-1}\lambda^{2r}\ell_{{\mathcal{W}}}(\theta_{2}))=\emptyset,

where ι(Ii)2\iota(I_{i})\subset\mathbb{R}^{2} is the Minkowski embedding of IiI_{i} as before.

Proof.

Note that

~ih={(αy1,t1,σ(α)y2,t2)k𝜽:αIi,t1,t2},\displaystyle\widetilde{\mathcal{L}}_{i}h=\left\{(\alpha y_{1},t_{1},\sigma(\alpha)y_{2},t_{2})\text{k}_{\bm{\theta}}\>:\>\alpha\in I_{i},\ t_{1},t_{2}\in\mathbb{R}\right\},

and recall that 𝒯(s)=(λ2rT(s))×(λ2r𝒲)\mathcal{T}(s)=(\lambda^{-2r}T(s))\times(\lambda^{2r}\mathcal{W}). Hence ~ih𝒯(s)=\widetilde{\mathcal{L}}_{i}h\cap\mathcal{T}(s)=\emptyset holds if and only if

αIi,(t1,t2)2:(αy1,t1,σ(α)y2,t2)s1/2λ2rT(1)kθ1×λ2r𝒲kθ2,\displaystyle\forall\,\alpha\in I_{i},\,\forall\,(t_{1},t_{2})\in\mathbb{R}^{2}:\quad(\alpha y_{1},t_{1},\sigma(\alpha)y_{2},t_{2})\notin s^{1/2}\lambda^{-2r}T(1)\text{k}_{-\theta_{1}}\times\lambda^{2r}\mathcal{W}\text{k}_{-\theta_{2}},

which is easily seen to be equivalent to (3.42). ∎

Remark 3.43.

Similarly, using K={(α,β,σ(α),σ(β))ξi1:αIi,βIi1}\mathcal{L}_{K}=\left\{(\alpha,\beta,\sigma(\alpha),\sigma(\beta))\xi^{-1}_{i}\>:\>\alpha\in I_{i},\ \beta\in I_{i}^{-1}\right\} (see (2.22)), one verifies that Kh𝒯(s)\mathcal{L}_{K}h\cap\mathcal{T}(s)\neq\emptyset holds if and only if there exist αIi\alpha\in I_{i} and βIi1\beta\in I_{i}^{-1} such that

(3.44) (y1α,y11(αx1+β),y2σ(α),y21(σ(α)x2+σ(β)))(s1/2λ2rT(1)kθ1)×(λ2r𝒲kθ2).\displaystyle(y_{1}\alpha,y_{1}^{-1}(\alpha x_{1}+\beta),y_{2}\sigma(\alpha),y_{2}^{-1}(\sigma(\alpha)x_{2}+\sigma(\beta)))\in(s^{1/2}\lambda^{-2r}T(1)\text{k}_{-\theta_{1}})\times(\lambda^{2r}\mathcal{W}\text{k}_{-\theta_{2}}).

As a consequence of Lemma 3.4 we have a simple necessary condition for ~ih𝒯(s)=\widetilde{\mathcal{L}}_{i}h\cap\mathcal{T}(s)=\emptyset.

Lemma 3.5.

Keep the assumptions as in Lemma 3.4. If (θ1,θ2)(π4,3π4)×[0,2π)(\theta_{1},\theta_{2})\in(\frac{\pi}{4},\frac{3\pi}{4})\times[0,2\pi), then ~ih𝒯(s)\widetilde{\mathcal{L}}_{i}h\cap\mathcal{T}(s)\neq\emptyset.

Proof.

Recall that we are assuming 𝟎𝒲=𝒲\mathbf{0}\in{\mathcal{W}}^{\circ}=\mathcal{W}, i.e. that 𝒲\mathcal{W} contains a small disc centered at the origin; therefore 0𝒲(θ2)0\in\ell_{{\mathcal{W}}}(\theta_{2}) for every θ2\theta_{2}. Now if θ1(π4,3π4)\theta_{1}\in(\frac{\pi}{4},\frac{3\pi}{4}), then 0(θ1)0\in\ell(\theta_{1}), and hence 𝟎ι(Ii)(y11s1/2λ2r(θ1))×(y21λ2r𝒲(θ2))\bm{0}\in\iota(I_{i})\cap(y_{1}^{-1}s^{1/2}\lambda^{-2r}\ell(\theta_{1}))\times(y_{2}^{-1}\lambda^{2r}\ell_{{\mathcal{W}}}(\theta_{2})), and so ~ih𝒯(s)\widetilde{\mathcal{L}}_{i}h\cap\mathcal{T}(s)\neq\emptyset by Lemma 3.4. ∎

Because of Lemma 3.5, we will often be able to reduce our discussion to the case θ1J\theta_{1}\in J, where

J:=(0,π)[π4,3π4]=(0,π4)(3π4,π).\displaystyle J:=(0,\pi)\setminus\bigl{[}\tfrac{\pi}{4},\tfrac{3\pi}{4}\bigr{]}=\bigl{(}0,\tfrac{\pi}{4}\bigr{)}\cup\bigl{(}\tfrac{3\pi}{4},\pi\bigr{)}.

In other words, we will be able to reduce the domain for 𝜽\theta to be

ZJ:=J×(0,2π)Z.\displaystyle Z_{J}:=J\times(0,2\pi)\subset Z.

3.5. Computing the main term

The goal of this section is to obtain simpler formulas for the function GM,i(s)G_{M,i}(s) for large ss. We first record the following useful and simple sufficient condition for a rectangle in the plane to intersect a lattice generated by an ideal of 𝒪K\mathcal{O}_{K}.

Lemma 3.6.

For any fractional ideal II of 𝒪K\mathcal{O}_{K}, there exists a constant C=C(I)>0C=C(I)>0 such that for any two intervals R1,R2R_{1},R_{2}\subset\mathbb{R} with |R1||R2|C|R_{1}|\cdot|R_{2}|\geq C, we have ι(I)(R1×R2)\iota(I)\cap(R_{1}\times R_{2})\neq\emptyset. Here |Ri||R_{i}| denotes the length of the interval RiR_{i}.

Proof.

This is a direct consequence of the fact that ι(I)\iota(I) is invariant under diag(u,σ(u))\operatorname{diag}(u,\sigma(u)) for any u𝒪K×u\in\mathcal{O}_{K}^{\times}, so that we can renormalize the rectangle R1×R2R_{1}\times R_{2} to be such that |R1|/|R2|[1,λ)|R_{1}|/|R_{2}|\in[1,\lambda). ∎

We next give a first explicit formula for GM,i(s)G_{M,i}(s). Recall that (θ)\ell(\theta) denotes the projection of T(1)kθT(1)\text{k}_{-\theta} on the xx-axis.

Proposition 3.7.

For all sufficiently large s>1s>1 we have

(3.45) GM,i(s)=cKArea(𝔉i)2sZJYI(ι(Ii)(y11(θ1)×y21𝒲(θ2))=)y13y23d𝒚d𝜽,\displaystyle G_{M,i}(s)=\frac{c_{K}\operatorname{Area}(\mathfrak{F}_{i})}{2s}\int_{Z_{J}}\int_{Y}I\left(\iota(I_{i})\cap(y_{1}^{-1}\ell(\theta_{1})\times y_{2}^{-1}\ell_{{\mathcal{W}}}(\theta_{2}))=\emptyset\right)y_{1}^{-3}y_{2}^{-3}\,\mathrm{d}\bm{y}\,\mathrm{d}\bm{\theta},

where Y:={(y1,y2)(>0)2:y2[1,λ2)}Y:=\left\{(y_{1},y_{2})\in(\mathbb{R}_{>0})^{2}\>:\>y_{2}\in[1,\lambda^{2})\right\}.

Proof.

Using (3.40), Lemma 3.4 and Lemma 3.5, and the fact that ι(Ii)\iota(I_{i}) is invariant under the action of diag(λ2r,λ2r)\operatorname{diag}(\lambda^{-2r},\lambda^{2r}), we have

(3.46) GM,i(s):=cKArea(𝔉i)ZJYt1I(ι(Ii)(y11s1/2(θ1))×(y21𝒲(θ2))=)y13y23d𝒚d𝜽.\displaystyle G_{M,i}(s):=c_{K}\operatorname{Area}(\mathfrak{F}_{i})\int_{Z_{J}}\int_{Y_{t_{1}}}I\Bigl{(}\iota(I_{i})\cap(y_{1}^{-1}s^{1/2}\ell(\theta_{1}))\times(y_{2}^{-1}\ell_{{\mathcal{W}}}(\theta_{2}))=\emptyset\Bigr{)}\,y_{1}^{-3}y_{2}^{-3}\,\text{d}\bm{y}\,\text{d}\bm{\theta}.

Because of 𝟎𝒲=𝒲\mathbf{0}\in{\mathcal{W}}^{\circ}=\mathcal{W}, there is an r𝒲>0r_{\mathcal{W}}>0 such that (r𝒲,r𝒲)𝒲(θ2)(-r_{\mathcal{W}},r_{\mathcal{W}})\subset\ell_{{\mathcal{W}}}(\theta_{2}) for all θ2\theta_{2}. Furthermore, (θ1)\ell(\theta_{1}) is an open interval of length |(θ1)|1|\ell(\theta_{1})|\asymp 1 for all θ1\theta_{1}. Hence by Lemma 3.6, there is a constant C>0C>0 such that ι(Ii)(y11s1/2(θ1))×(y21𝒲(θ2))\iota(I_{i})\cap(y_{1}^{-1}s^{1/2}\ell(\theta_{1}))\times(y_{2}^{-1}\ell_{{\mathcal{W}}}(\theta_{2}))\neq\emptyset holds for all 𝜽ZJ{\text{\boldmath$\theta$}}\in Z_{J} and all s,y1,y2>0s,y_{1},y_{2}\in\mathbb{R}_{>0} subject to y11s1/2y21>Cy_{1}^{-1}s^{1/2}\cdot y_{2}^{-1}>C. It follows that if ss is sufficiently large, then the formula in (3.46) remains valid if the domain of integration Yt1Y_{t_{1}} (defined in (3.38)) is replaced by the larger set

Y:={(y1,y2)(>0)2:y1/y2[1,λ2)},\displaystyle Y^{\prime}:=\left\{(y_{1},y_{2})\in(\mathbb{R}_{>0})^{2}\>:\>y_{1}/y_{2}\in[1,\lambda^{2})\right\},

i.e. we have (after also changing the order of integration)

(3.47) GM,i(s):=cKArea(𝔉i)YZJI(ι(Ii)(y11s1/2(θ1))×(y21𝒲(θ2))=)d𝜽dy1y13dy2y23.\displaystyle G_{M,i}(s):=c_{K}\operatorname{Area}(\mathfrak{F}_{i})\int_{Y^{\prime}}\int_{Z_{J}}I\Bigl{(}\iota(I_{i})\cap(y_{1}^{-1}s^{1/2}\ell(\theta_{1}))\times(y_{2}^{-1}\ell_{{\mathcal{W}}}(\theta_{2}))=\emptyset\Bigr{)}\,\text{d}{\text{\boldmath$\theta$}}\,\frac{\text{d}y_{1}}{y_{1}^{3}}\,\frac{\text{d}y_{2}}{y_{2}^{3}}.

Going through the same arguments but replacing the condition y1/y2[1,λ2)y_{1}/y_{2}\in[1,\lambda^{2}) by y1/y2[λ2,λ4)y_{1}/y_{2}\in[\lambda^{2},\lambda^{4}) in the definition of i\mathcal{F}_{i} in (2.16) (this amounts to choosing a different fundamental domain for Γi\H\Gamma_{i}\backslash\mathrm{H}), one verifies that (3.47) remains valid for all sufficiently large s>1s>1 with YY^{\prime} replaced by the set

{(y1,y2)(>0)2:y1/y2[λ2,λ4)}.\displaystyle\left\{(y_{1},y_{2})\in(\mathbb{R}_{>0})^{2}\>:\>y_{1}/y_{2}\in[\lambda^{2},\lambda^{4})\right\}.

Equivalently, we have for s>1s>1 sufficiently large

(3.48) GM,i(s)\displaystyle G_{M,i}(s) =12cKArea(𝔉i)Y′′ZJI(ι(Ii)(y11s1/2(θ1))×(y21𝒲(θ2))=)d𝜽dy1y13dy2y23,\displaystyle=\frac{1}{2}c_{K}\operatorname{Area}(\mathfrak{F}_{i})\int_{Y^{\prime\prime}}\int_{Z_{J}}I\Bigl{(}\iota(I_{i})\cap(y_{1}^{-1}s^{1/2}\ell(\theta_{1}))\times(y_{2}^{-1}\ell_{{\mathcal{W}}}(\theta_{2}))=\emptyset\Bigr{)}\,\text{d}{\text{\boldmath$\theta$}}\,\frac{\text{d}y_{1}}{y_{1}^{3}}\,\frac{\text{d}y_{2}}{y_{2}^{3}},

where

Y′′:={(y1,y2)(>0)2:y1/y2[1,λ4)}.\displaystyle Y^{\prime\prime}:=\left\{(y_{1},y_{2})\in(\mathbb{R}_{>0})^{2}\>:\>y_{1}/y_{2}\in[1,\lambda^{4})\right\}.

Next, using the fact that the lattice ι(Ii)\iota(I_{i}) is invariant under the action of S=diag(λ2,λ2)S=\operatorname{diag}(\lambda^{2},\lambda^{-2}), it follows that the integrand in (3.48) is invariant under 𝒚𝒚S{\text{\boldmath$y$}}\mapsto{\text{\boldmath$y$}}S, for each fixed 𝜽\theta. Also the measure y13y23d𝒚y_{1}^{-3}y_{2}^{-3}\,\text{d}{\text{\boldmath$y$}} is invariant under this map and both Y′′Y^{\prime\prime} and YY are fundamental domains for (>0)2/𝒚𝒚S(\mathbb{R}_{>0})^{2}/\langle{\text{\boldmath$y$}}\mapsto{\text{\boldmath$y$}}S\rangle. Hence the formula (3.48) remains valid if the domain of integration Y′′Y^{\prime\prime} is replaced by YY, giving us (3.45). ∎

Next we further simplify the above empty intersection condition to get an even more explicit formula in terms of the function αI,J\alpha_{I,J} defined in (3.3).

Proposition 3.8.

For all sufficiently large s>1s>1 we have

(3.49) GM,i(s)=cKArea(𝔉i)c𝒫4s02π1λ2αIi,𝒲(θ)(y)2dyy3dθ.\displaystyle G_{M,i}(s)=\frac{c_{K}\operatorname{Area}(\mathfrak{F}_{i})\,c_{{\mathcal{P}}}}{4s}\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}\alpha_{I_{i},\ell_{\mathcal{W}}(\theta)}(y)^{2}\,\frac{\mathrm{d}y}{y^{3}}\,\mathrm{d}\theta.

In particular, we have for all sufficiently large s>1s>1,

(3.50) i=1κGM,i(s)=a𝒫s,\displaystyle\sum_{i=1}^{\kappa}G_{M,i}(s)=\frac{a_{\mathcal{P}}}{s},

where a𝒫a_{\mathcal{P}} is as in (3.8).

Proof.

The formula in Proposition 3.7 can be expressed as

(3.51) GM,i(s)=cKArea(𝔉i)2s02π1λ2(y2,θ2)dy2y23dθ2,\displaystyle G_{M,i}(s)=\frac{c_{K}\operatorname{Area}(\mathfrak{F}_{i})}{2s}\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}{\mathcal{I}}(y_{2},\theta_{2})\,\frac{\text{d}y_{2}}{y_{2}^{3}}\,\text{d}\theta_{2},

where

(y2,θ2):=J0I(ι(Ii)(y11(θ1)×y21𝒲(θ2))=)dy1y13dθ1.\displaystyle{\mathcal{I}}(y_{2},\theta_{2}):=\int_{J}\int_{0}^{\infty}I\left(\iota(I_{i})\cap(y_{1}^{-1}\ell(\theta_{1})\times y_{2}^{-1}\ell_{{\mathcal{W}}}(\theta_{2}))=\emptyset\right)\,\frac{\text{d}y_{1}}{y_{1}^{3}}\,\text{d}\theta_{1}.

Here recall that (θ1)\ell(\theta_{1}) equals the projection of T(1)kθ1T(1)\text{k}_{-\theta_{1}} on the xx-axis, which we can compute to be:

(3.52) (θ1)={(0,c𝒫1/2(cosθ1+sinθ1))if  0θ1π/4,(c𝒫1/2(cosθ1sinθ1),0)if  3π/4θ1π.\displaystyle\ell(\theta_{1})=\begin{cases}\bigl{(}0,c_{{\mathcal{P}}}^{-1/2}(\cos\theta_{1}+\sin\theta_{1})\bigr{)}&\text{if }\>0\leq\theta_{1}\leq\pi/4,\\ \bigl{(}c_{{\mathcal{P}}}^{-1/2}(\cos\theta_{1}-\sin\theta_{1}),0\bigr{)}&\text{if }\>3\pi/4\leq\theta_{1}\leq\pi.\end{cases}

Recall the definition of αI,J(y)\alpha_{I,J}(y) in (3.3). We also give a companion definition here. For any integral ideal I𝒪KI\subset\mathcal{O}_{K} and any bounded Lebesgue measurable set JJ\subset\mathbb{R} with non-empty interior, we set for any y>0y>0,

(3.53) α~I,J(y):=max{αI<0:yσ(α)J}.\displaystyle\tilde{\alpha}_{I,J}(y):=\max\bigl{\{}\alpha\in I\cap\mathbb{R}_{<0}\>:\>y\,\sigma(\alpha)\in J\bigr{\}}.

With these definitions the formula for (y2,θ2){\mathcal{I}}(y_{2},\theta_{2}) can be rewritten as

0π/40I(αIi,𝒲(θ2)(y2)y11(θ1))dy1y13dθ1+3π/4π0I(α~Ii,𝒲(θ2)(y2)y11(θ1))dy1y13dθ1\displaystyle\int_{0}^{\pi/4}\int_{0}^{\infty}I\left(\alpha_{I_{i},\ell_{\mathcal{W}}(\theta_{2})}(y_{2})\notin y_{1}^{-1}\ell(\theta_{1})\right)\,\frac{\text{d}y_{1}}{y_{1}^{3}}\,\text{d}\theta_{1}+\int_{3\pi/4}^{\pi}\int_{0}^{\infty}I\left(\tilde{\alpha}_{I_{i},\ell_{\mathcal{W}}(\theta_{2})}(y_{2})\notin y_{1}^{-1}\ell(\theta_{1})\right)\,\frac{\text{d}y_{1}}{y_{1}^{3}}\,\text{d}\theta_{1}
=12c𝒫αIi,𝒲(θ2)(y2)20π/4dθ1(cosθ1+sinθ1)2+12c𝒫α~Ii,𝒲(θ2)(y2)23π/4πdθ1(cosθ1sinθ1)2\displaystyle=\frac{1}{2}c_{{\mathcal{P}}}\alpha_{I_{i},\ell_{\mathcal{W}}(\theta_{2})}(y_{2})^{2}\int_{0}^{\pi/4}\frac{\text{d}\theta_{1}}{(\cos\theta_{1}+\sin\theta_{1})^{2}}+\frac{1}{2}c_{{\mathcal{P}}}\tilde{\alpha}_{I_{i},\ell_{\mathcal{W}}(\theta_{2})}(y_{2})^{2}\int_{3\pi/4}^{\pi}\frac{\text{d}\theta_{1}}{(\cos\theta_{1}-\sin\theta_{1})^{2}}
=c𝒫4(αIi,𝒲(θ2)(y2)2+α~Ii,𝒲(θ2)(y2)2),\displaystyle=\frac{c_{{\mathcal{P}}}}{4}\bigl{(}\alpha_{I_{i},\ell_{\mathcal{W}}(\theta_{2})}(y_{2})^{2}+\tilde{\alpha}_{I_{i},\ell_{\mathcal{W}}(\theta_{2})}(y_{2})^{2}\bigr{)},

implying that

02π1λ2(y2,θ2)dy2y23dθ2\displaystyle\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}{\mathcal{I}}(y_{2},\theta_{2})\,\frac{\text{d}y_{2}}{y_{2}^{3}}\,\text{d}\theta_{2} =c𝒫402π1λ2(αIi,𝒲(θ)(y)2+α~Ii,𝒲(θ)(y)2)dyy3dθ.\displaystyle=\frac{c_{\mathcal{P}}}{4}\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}\bigl{(}\alpha_{I_{i},\ell_{\mathcal{W}}(\theta)}(y)^{2}+\tilde{\alpha}_{I_{i},\ell_{\mathcal{W}}(\theta)}(y)^{2}\bigr{)}\,\frac{\text{d}y}{y^{3}}\,\text{d}\theta.

Now note that 𝒲(θ+π)=𝒲(θ)\ell_{\mathcal{W}}(\theta+\pi)=-\ell_{\mathcal{W}}(\theta) from which we can deduce that αIi,𝒲(θ+π)(y)=α~Ii,𝒲(θ)(y)\alpha_{I_{i},\ell_{\mathcal{W}}(\theta+\pi)}(y)=-\tilde{\alpha}_{I_{i},\ell_{\mathcal{W}}(\theta)}(y). Hence

02π1λ2α~Ii,𝒲(θ)(y)2dyy3dθ\displaystyle\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}\tilde{\alpha}_{I_{i},\ell_{\mathcal{W}}(\theta)}(y)^{2}\,\frac{\text{d}y}{y^{3}}\,\text{d}\theta =02π1λ2αIi,𝒲(θ+π)(y)2dyy3dθ\displaystyle=\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}\alpha_{I_{i},\ell_{\mathcal{W}}(\theta+\pi)}(y)^{2}\,\frac{\text{d}y}{y^{3}}\,\text{d}\theta
=02π1λ2αIi,𝒲(θ)(y)2dyy3dθ,\displaystyle=\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}\alpha_{I_{i},\ell_{\mathcal{W}}(\theta)}(y)^{2}\,\frac{\text{d}y}{y^{3}}\,\text{d}\theta,

implying that

02π1λ2(y2,θ2)dy2y23dθ2\displaystyle\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}{\mathcal{I}}(y_{2},\theta_{2})\,\frac{\text{d}y_{2}}{y_{2}^{3}}\,\text{d}\theta_{2} =c𝒫202π1λ2αIi,𝒲(θ)(y)2dyy3dθ.\displaystyle=\frac{c_{\mathcal{P}}}{2}\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}\alpha_{I_{i},\ell_{\mathcal{W}}(\theta)}(y)^{2}\,\frac{\text{d}y}{y^{3}}\,\text{d}\theta.

Plugging this back to (3.51) we get the desired formula for GM,i(s)G_{M,i}(s), (3.49).

The formula (3.50) follows directly from (3.49) by using

Area(𝔉i)=covol(ι(Ii2))=Nr(Ii)2ΔK1/2\operatorname{Area}(\mathfrak{F}_{i})=\operatorname{covol}(\iota(I^{-2}_{i}))=\text{Nr}(I_{i})^{-2}\Delta_{K}^{1/2}

(which holds by (2.5)) and since 𝔉i\mathfrak{F}_{i} is a fundamental domain for 2/ι(Ii2)\mathbb{R}^{2}/\iota(I_{i}^{-2})), and the formulas for cKc_{K} and c𝒫c_{\mathcal{P}} respectively; see (2.13) and (1.11). ∎

3.6. Estimating the error term

Recall that G(s)=i=1κGi(s)=i=1κ(GM,i(s)+GE,i(s))G(s)=\sum_{i=1}^{\kappa}G_{i}(s)=\sum_{i=1}^{\kappa}\bigl{(}G_{M,i}(s)+G_{E,i}(s)\bigr{)} for all sufficiently large ss, and by definition we have GE,i(s)0G_{E,i}(s)\geq 0; see (3.41). Recalling also Proposition 3.8, it follows that, in order to complete the proof of Theorem 4, it only remains to prove that GE,i(s)G_{E,i}(s) is majorized by the bound in (3.7) for all sufficiently large ss.

We first prove two auxiliary lemmas.

Lemma 3.9.

For any fractional ideal II of 𝒪K\mathcal{O}_{K}, there exists a constant C=C(I)>0C=C(I)>0 such that for any two intervals R1,R2R_{1},R_{2}\subset\mathbb{R} with |R1||R2|C|R_{1}|\cdot|R_{2}|\geq C, and for any Lebesgue measurable subset R2R_{2}^{\prime}\subset\mathbb{R}, we have

Area(XIπ(R1×R2))|R1|m(R2R2),\displaystyle\operatorname{Area}(X_{I}\setminus\pi(R_{1}\times R_{2}^{\prime}))\leq|R_{1}|\cdot m(R_{2}\setminus R_{2}^{\prime}),

where XI:=2/ι(I)X_{I}:=\mathbb{R}^{2}/\iota(I), and π:2XI\pi:\mathbb{R}^{2}\to X_{I} is the quotient map.

(Recall that mm denotes Lebesgue measure on \mathbb{R}; also for an interval RR we write |R|=m(R)|R|=m(R) for its length.)

Proof.

Let us fix a fundamental parallelogram 𝔉\mathfrak{F} for the lattice ι(I)\iota(I) in 2\mathbb{R}^{2}, and then fix intervals J1,J2J_{1},J_{2}\subset\mathbb{R} such that 𝔉J1×J2\mathfrak{F}\subset J_{1}\times J_{2}. We claim that the statement of the lemma holds with C:=λ|J1||J2|C:=\lambda|J_{1}||J_{2}|. Indeed, assume that R1,R2R_{1},R_{2} are intervals with |R1||R2|λ|J1||J2||R_{1}|\cdot|R_{2}|\geq\lambda|J_{1}||J_{2}|. Then there exists some mm\in\mathbb{Z} such that λm|R1||J1|\lambda^{m}|R_{1}|\geq|J_{1}| and λm|R2||J2|\lambda^{-m}|R_{2}|\geq|J_{2}|, viz., |σ(λm)R2||J2||\sigma(\lambda^{m})R_{2}|\geq|J_{2}|. This means that a translate of (R1×R2)diag(λ,σ(λ))m(R_{1}\times R_{2})\operatorname{diag}(\lambda,\sigma(\lambda))^{m} contains FF; and hence R1×R2R_{1}\times R_{2} contains a translate 𝔉\mathfrak{F}^{\prime} of 𝔉diag(λ,σ(λ))m\mathfrak{F}\operatorname{diag}(\lambda,\sigma(\lambda))^{-m}. But using the fact that ι(I)\iota(I) is invariant under diag(λ,σ(λ))\operatorname{diag}(\lambda,\sigma(\lambda)), it follows that 𝔉\mathfrak{F}^{\prime} is again an fundamental parallelogram for the lattice ι(I)\iota(I) in 2\mathbb{R}^{2}. Hence

Area(XIπ(R1×R2))=Area(𝔉π1(π(R1×R2))).\displaystyle\operatorname{Area}(X_{I}\setminus\pi(R_{1}\times R_{2}^{\prime}))=\operatorname{Area}(\mathfrak{F}^{\prime}\setminus\pi^{-1}(\pi(R_{1}\times R_{2}^{\prime}))).

Using 𝔉R1×R2\mathfrak{F}^{\prime}\subset R_{1}\times R_{2} and π1(π(R1×R2))R1×R2\pi^{-1}(\pi(R_{1}\times R_{2}^{\prime}))\supset R_{1}\times R_{2}^{\prime}, the above is

Area((R1×R2)(R1×R2))=Area(R1×(R2R2))=|R1|m(R2R2).\displaystyle\leq\operatorname{Area}((R_{1}\times R_{2})\setminus(R_{1}\times R_{2}^{\prime}))=\operatorname{Area}(R_{1}\times(R_{2}\setminus R_{2}^{\prime}))=|R_{1}|\cdot m(R_{2}\setminus R_{2}^{\prime}).

Lemma 3.10.

For any fixed set AA\subset\mathbb{R} we have

(3.54)  0<a1a2:𝔣(A,a1)2𝔣(A,a2).\displaystyle\forall\>0<a_{1}\leq a_{2}:\qquad\mathfrak{f}(A,a_{1})\leq 2\mathfrak{f}(A,a_{2}).
Proof.

Let 0<a1a20<a_{1}\leq a_{2} be given, and let JJ be an arbitrary interval of length a2a_{2}. Take k>0k\in\mathbb{Z}_{>0} so that ka1a2<(k+1)a1ka_{1}\leq a_{2}<(k+1)a_{1}; then there exist kk pairwise disjoint (open) intervals J1,,JkJJ_{1},\ldots,J_{k}\subset J which all have length a1a_{1}. Now m(JA)=1km(JA)ka1𝔣(A,a1)m(J\setminus A)\geq\sum_{\ell=1}^{k}m(J_{\ell}\setminus A)\geq ka_{1}\mathfrak{f}(A,a_{1}), and since this holds for all intervals JJ of length a2a_{2}, it follows that 𝔣(A,a2)a21ka1𝔣(A,a1)kk+1𝔣(A,a1)12𝔣(A,a1)\mathfrak{f}(A,a_{2})\geq a_{2}^{-1}\cdot ka_{1}\mathfrak{f}(A,a_{1})\geq\frac{k}{k+1}\mathfrak{f}(A,a_{1})\geq\frac{1}{2}\mathfrak{f}(A,a_{1}). ∎

We now turn to bounding the error term GE,i(s)G_{E,i}(s), which we defined in (3.41), for large ss.

First of all, note that if θ1(π4,3π4)\theta_{1}\in(\frac{\pi}{4},\frac{3\pi}{4}) then T(1)kθ1T(1)\text{k}_{-\theta_{1}} contains the line segment between (0,0)(0,0) and (0,c𝒫1/2)(0,c_{{\mathcal{P}}}^{-1/2}). Recall also that since 𝟎𝒲=𝒲\mathbf{0}\in{\mathcal{W}}^{\circ}=\mathcal{W}, there is an r𝒲>0r_{\mathcal{W}}>0 such that r𝒲2𝒲{\mathcal{B}}_{r_{\mathcal{W}}}^{2}\subset{\mathcal{W}}. It follows that if θ1(π4,3π4)\theta_{1}\in(\frac{\pi}{4},\frac{3\pi}{4}) then the condition (3.44) is satisfied with α=0\alpha=0 whenever (β,σ(β))(\beta,\sigma(\beta)) belongs to the rectangle (0,y1s1/2λ2rc𝒫1/2)×(y2λ2rr𝒲,y2λ2rr𝒲)\bigl{(}0,y_{1}s^{1/2}\lambda^{-2r}c_{{\mathcal{P}}}^{-1/2}\bigr{)}\times\bigl{(}-y_{2}\lambda^{2r}r_{\mathcal{W}},y_{2}\lambda^{2r}r_{\mathcal{W}}\bigr{)}. This rectangle has area s1/2\gg s^{1/2}, since 𝒚Yt1{\text{\boldmath$y$}}\in Y_{t_{1}}. Hence by Lemma 3.6, for ss sufficiently large there always exists some βIi1\beta\in I_{i}^{-1} such that (β,σ(β))(\beta,\sigma(\beta)) belongs to the rectangle, and so, by Remark 3.43, Kh𝒯(s){\mathcal{L}}_{K}h\cap{\mathcal{T}}(s)\neq\emptyset. This proves that the contribution from all θ1(π4,3π4)\theta_{1}\in(\frac{\pi}{4},\frac{3\pi}{4}) to the integral in (3.41) is zero. Thus, for ss sufficiently large, we may replace the range of integration for 𝜽\theta in (3.41) by ZJZ_{J}.

Next, as in the discussion in Section 3.4 (near (3.38)), because of the condition Kh𝒯(s)=\mathcal{L}_{K}h\cap\mathcal{T}(s)=\emptyset in the integrand, any 𝒚Yt1{\text{\boldmath$y$}}\in Y_{t_{1}} which contribute to the integral in (3.41) must satisfy y1,y2s1/4y_{1},y_{2}\gg s^{1/4}. Furthermore, by Lemma 3.4, given any 𝜽,𝒚,𝒙\langle{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},{\text{\boldmath$x$}}\rangle which makes the integrand in (3.41) nonzero, there exists some αIi\alpha\in I_{i} such that (α,σ(α))(\alpha,\sigma(\alpha)) belongs to the set

(3.55) 𝔄𝜽,𝒚(s):=(y11s1/2λ2r(θ1))×(y21λ2r𝒲(θ2)).\displaystyle\mathfrak{A}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}}}(s):=(y_{1}^{-1}s^{1/2}\lambda^{-2r}\ell(\theta_{1}))\times(y_{2}^{-1}\lambda^{2r}\ell_{{\mathcal{W}}}(\theta_{2})).

Since we have restricted to θ1J\theta_{1}\in J, we have 0(θ1)0\notin\ell(\theta_{1}), and hence α0\alpha\neq 0. Note that (θ1)\ell(\theta_{1}) and 𝒲(θ2)\ell_{{\mathcal{W}}}(\theta_{2}) are contained in bounded intervals independent of θ1,θ2\theta_{1},\theta_{2}; hence for y1,y2s1/4y_{1},y_{2}\gg s^{1/4}, the set 𝔄𝜽,𝒚(s)\mathfrak{A}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}}}(s) is contained in a fixed bounded region B2B\subset\mathbb{R}^{2}. Set

Λ:={αIi{0}:(α,σ(α))B}.\displaystyle\Lambda:=\{\alpha\in I_{i}\setminus\{0\}\>:\>(\alpha,\sigma(\alpha))\in B\}.

This is a finite subset of Ii{0}I_{i}\setminus\{0\}, and our discussion shows that for any 𝜽,𝒚,𝒙ZJ×Yt1×𝔉i\langle{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},{\text{\boldmath$x$}}\rangle\in Z_{J}\times Y_{t_{1}}\times\mathfrak{F}_{i} which makes the integrand in (3.41) nonzero, there must exist some αΛ\alpha\in\Lambda satisfying (α,σ(α))𝔄𝜽,𝒚(s)(\alpha,\sigma(\alpha))\in\mathfrak{A}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}}}(s), or equivalently, ι(Λ)𝔄𝜽,𝒚(s)\iota(\Lambda)\cap\mathfrak{A}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}}}(s)\neq\emptyset. Using now the fact that both the minima of |α||\alpha| and |σ(α)||\sigma(\alpha)| for αΛ\alpha\in\Lambda are bounded away from zero, and also the fact that both |(θ1)||\ell(\theta_{1})| and |𝒲(θ2)||\ell_{{\mathcal{W}}}(\theta_{2})| are bounded away from zero independently of θ1,θ2\theta_{1},\theta_{2}, it follows that we must have y1,y2s1/4y_{1},y_{2}\ll s^{1/4}. We have thus proved that there exist some constants c2>c1>0c_{2}>c_{1}>0 (which depend on KK and 𝒲{\mathcal{W}}, but not on ss), such that for all sufficiently large ss, every 𝜽,𝒚,𝒙ZJ×Yt1×𝔉i\langle{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},{\text{\boldmath$x$}}\rangle\in Z_{J}\times Y_{t_{1}}\times\mathfrak{F}_{i} which makes the integrand in (3.41) nonzero, must satisfy c1s1/4y1,y2c2s1/4c_{1}s^{1/4}\leq y_{1},y_{2}\leq c_{2}s^{1/4}.

In view of the above discussion, we have for ss sufficiently large:

GE,i(s)s3/2ZJc1s1/4c2s1/4c1s1/4c2s1/4I(ι(Λ)𝔄𝜽,𝒚(s))𝔉iI(Kh𝒯(s)=)d𝒙d𝒚d𝜽\displaystyle G_{E,i}(s)\asymp s^{-3/2}\int_{Z_{J}}\int_{c_{1}s^{1/4}}^{c_{2}s^{1/4}}\int_{c_{1}s^{1/4}}^{c_{2}s^{1/4}}I\Bigl{(}\iota(\Lambda)\cap\mathfrak{A}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}}}(s)\neq\emptyset\Bigr{)}\int_{\mathfrak{F}_{i}}I\left(\mathcal{L}_{K}h\cap\mathcal{T}(s)=\emptyset\right)\,\text{d}\bm{x}\,\text{d}\bm{y}\,\text{d}\bm{\theta}
(3.56) s3/2αΛZJc1s1/4c2s1/4c1s1/4c2s1/4I((α,σ(α))𝔄𝜽,𝒚(s))𝔉iI(Kh𝒯(s)=)d𝒙d𝒚d𝜽.\displaystyle\asymp s^{-3/2}\sum_{\alpha\in\Lambda}\int_{Z_{J}}\int_{c_{1}s^{1/4}}^{c_{2}s^{1/4}}\int_{c_{1}s^{1/4}}^{c_{2}s^{1/4}}I\Bigl{(}(\alpha,\sigma(\alpha))\in\mathfrak{A}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}}}(s)\Bigr{)}\int_{\mathfrak{F}_{i}}I\left(\mathcal{L}_{K}h\cap\mathcal{T}(s)=\emptyset\right)\,\text{d}\bm{x}\,\text{d}\bm{y}\,\text{d}\bm{\theta}.

Next, the condition Kh𝒯(s)=\mathcal{L}_{K}h\cap\mathcal{T}(s)=\emptyset in (3.56) implies by Remark 3.43 that (3.44) fails for all (α,β)Ii×Ii1(\alpha,\beta)\in I_{i}\times I_{i}^{-1}. This is equivalent to saying that for every αIi\alpha\in I_{i} we have ι(Ii1)𝔅𝜽,𝒚,𝒙,α(s)=\iota(I_{i}^{-1})\cap\mathfrak{B}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},{\text{\boldmath$x$}},\alpha}(s)=\emptyset, where

(3.57) 𝔅𝜽,𝒚,𝒙,α(s):=(y1𝔍1αx1)×(y2𝔍2σ(α)x2)\displaystyle\mathfrak{B}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},{\text{\boldmath$x$}},\alpha}(s):=\left(y_{1}\mathfrak{J}_{1}-\alpha x_{1}\right)\times\left(y_{2}\mathfrak{J}_{2}-\sigma(\alpha)x_{2}\right)

with 𝔍1=𝔍1(s,y1,θ1,α)\mathfrak{J}_{1}=\mathfrak{J}_{1}(s,y_{1},\theta_{1},\alpha) and 𝔍2=𝔍2(s,y2,θ2,α)\mathfrak{J}_{2}=\mathfrak{J}_{2}(s,y_{2},\theta_{2},\alpha) defined by

(3.58) 𝔍1:={t:(y1α,t)s1/2λ2rT(1)kθ1}and𝔍2:={t:(y2σ(α),t)λ2r𝒲kθ2}.\displaystyle\mathfrak{J}_{1}:=\left\{t\in\mathbb{R}\>:\>(y_{1}\alpha,t)\in s^{1/2}\lambda^{-2r}T(1)\text{k}_{-\theta_{1}}\right\}\quad\text{and}\quad\mathfrak{J}_{2}:=\left\{t\in\mathbb{R}\>:\>(y_{2}\sigma(\alpha),t)\in\lambda^{2r}\mathcal{W}\text{k}_{-\theta_{2}}\right\}.

Hence from (3.56) we get, using also the fact that if ι(Ii1)𝔅𝜽,𝒚,𝒙,α(s)=\iota(I_{i}^{-1})\cap\mathfrak{B}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},{\text{\boldmath$x$}},\alpha}(s)=\emptyset holds for every αIi\alpha\in I_{i}, then in particular it holds for the αΛ\alpha\in\Lambda which is our summation variable:

(3.59) GE,i(s)s3/2αΛZJc1s1/4c2s1/4c1s1/4c2s1/4I((α,σ(α))𝔄𝜽,𝒚(s))\displaystyle G_{E,i}(s)\ll s^{-3/2}\sum_{\alpha\in\Lambda}\int_{Z_{J}}\int_{c_{1}s^{1/4}}^{c_{2}s^{1/4}}\int_{c_{1}s^{1/4}}^{c_{2}s^{1/4}}I\Bigl{(}(\alpha,\sigma(\alpha))\in\mathfrak{A}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}}}(s)\Bigr{)}\hskip 100.0pt
×𝔉iI(ι(Ii1)𝔅𝜽,𝒚,𝒙,α(s)=)d𝒙d𝒚d𝜽.\displaystyle\times\int_{\mathfrak{F}_{i}}I\Bigl{(}\iota(I_{i}^{-1})\cap\mathfrak{B}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},{\text{\boldmath$x$}},\alpha}(s)=\emptyset\Bigr{)}\,\text{d}\bm{x}\,\text{d}\bm{y}\,\text{d}\bm{\theta}.

Note here that 𝔅𝜽,𝒚,𝒙,α(s)=𝔅𝜽,𝒚,𝟎,α(s)(αx1,σ(α)x2)\mathfrak{B}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},{\text{\boldmath$x$}},\alpha}(s)=\mathfrak{B}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},\mathbf{0},\alpha}(s)-(\alpha x_{1},\sigma(\alpha)x_{2}), and recall that 𝔉i\mathfrak{F}_{i} is a fundamental domain for 2/ι(Ii2)\mathbb{R}^{2}/\iota(I_{i}^{-2}). Note also that the map 𝒙𝒗:=(αx1,σ(α)x2){\text{\boldmath$x$}}\mapsto{\text{\boldmath$v$}}:=(\alpha x_{1},\sigma(\alpha)x_{2}) induces a diffeomorphism from 2/ι(Ii2)\mathbb{R}^{2}/\iota(I_{i}^{-2}) onto 2/ι(αIi2)\mathbb{R}^{2}/\iota(\alpha I_{i}^{-2}), carrying d𝒙x to |N(α)|1d𝒗|\mathrm{N}(\alpha)|^{-1}\text{d}{\text{\boldmath$v$}}. Hence the integral over 𝒙x in (3.59) equals

(3.60) 1|N(α)|2/ι(αIi2)I(ι(Ii1)(𝔅𝜽,𝒚,𝟎,α(s)𝒗)=)d𝒗.\displaystyle\frac{1}{|\mathrm{N}(\alpha)|}\int_{\mathbb{R}^{2}/\iota(\alpha I_{i}^{-2})}I\Bigl{(}\iota(I_{i}^{-1})\cap(\mathfrak{B}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},\mathbf{0},\alpha}(s)-{\text{\boldmath$v$}})=\emptyset\Bigr{)}\,\text{d}{\text{\boldmath$v$}}.

Let π\pi be the quotient map from 2\mathbb{R}^{2} onto the torus X:=2/ι(Ii1)X:=\mathbb{R}^{2}/\iota(I_{i}^{-1}). Note that the condition ι(Ii1)(𝔅𝜽,𝒚,𝟎,α(s)𝒗)=\iota(I_{i}^{-1})\cap(\mathfrak{B}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},\mathbf{0},\alpha}(s)-{\text{\boldmath$v$}})=\emptyset is equivalent with π(𝒗)π(𝔅𝜽,𝒚,𝟎,α(s))\pi({\text{\boldmath$v$}})\notin\pi(\mathfrak{B}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},\mathbf{0},\alpha}(s)). Also, since αΛIi\alpha\in\Lambda\subset I_{i}, ι(αIi2)\iota(\alpha I_{i}^{-2}) is a sublattice of ι(Ii1)\iota(I_{i}^{-1}); hence π\pi induces to a covering map from 2/ι(αIi2)\mathbb{R}^{2}/\iota(\alpha I_{i}^{-2}) onto XX, which preserves Lebesgue measure. Using (2.5) it follows that this covering map has degree |N(α)|/Nr(Ii)|\mathrm{N}(\alpha)|/\text{Nr}(I_{i}). Hence the integral in (3.60) equals

Nr(Ii)1Area(Xπ(𝔅𝜽,𝒚,𝟎,α(s))).\displaystyle\operatorname{Nr}(I_{i})^{-1}\operatorname{Area}(X\setminus\pi(\mathfrak{B}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},\mathbf{0},\alpha}(s))).

But we have 𝔅𝜽,𝒚,𝟎,α(s)=y1𝔍1×y2𝔍2\mathfrak{B}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},\mathbf{0},\alpha}(s)=y_{1}\mathfrak{J}_{1}\times y_{2}\mathfrak{J}_{2}, where 𝔍1\mathfrak{J}_{1} is an open interval which is non-empty if and only if αy11s1/2λ2r(θ1)\alpha\in y_{1}^{-1}s^{1/2}\lambda^{-2r}\ell(\theta_{1}). This last condition is guaranteed to hold because of the condition (α,σ(α))𝔄𝜽,𝒚(s)(\alpha,\sigma(\alpha))\in\mathfrak{A}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}}}(s) appearing in (3.59). It now follows from Lemma 3.9 that there is a constant C=C(Ii1)C=C(I_{i}^{-1}) such that, for any interval R2R_{2} of length |R2|=C/|y1𝔍1||R_{2}|=C/|y_{1}\mathfrak{J}_{1}|, the integral in (3.60) is bounded from above by Nr(Ii)1|y1𝔍1|m(R2y2𝔍2)\operatorname{Nr}(I_{i})^{-1}|y_{1}\mathfrak{J}_{1}|\cdot m\bigl{(}R_{2}\setminus y_{2}\mathfrak{J}_{2}\bigr{)}. Furthermore, comparing (3.58) and (3.1) we see that 𝔍2=λ2rR𝒲(θ2,λ2rσ(α)y2)\mathfrak{J}_{2}=\lambda^{2r}R_{{\mathcal{W}}}(\theta_{2},\lambda^{-2r}\sigma(\alpha)y_{2}). Hence, writing R2=y2λ2r(x,x+a)R_{2}=y_{2}\lambda^{2r}\cdot(x,x+a) with xx\in\mathbb{R} and a=C/(λ2ry1y2|𝔍1|)a=C/\bigl{(}\lambda^{2r}y_{1}y_{2}|\mathfrak{J}_{1}|\bigr{)}, it follows that the integral in (3.60) is bounded from above by

Nr(Ii)1λ2ry1y2|𝔍1|m((x,x+a)R𝒲(θ2,λ2rσ(α)y2)).\displaystyle\operatorname{Nr}(I_{i})^{-1}\lambda^{2r}y_{1}y_{2}|\mathfrak{J}_{1}|\cdot m\bigl{(}(x,x+a)\setminus R_{{\mathcal{W}}}(\theta_{2},\lambda^{-2r}\sigma(\alpha)y_{2})\bigr{)}.

This is true for every xx\in\mathbb{R}. Hence, using the notation introduced in (3.5), we conclude that the integral in (3.60) is bounded from above by

Nr(Ii)1C𝔣(R𝒲(θ2,λ2rσ(α)y2),Cλ2ry1y2|𝔍1|).\displaystyle\operatorname{Nr}(I_{i})^{-1}C\cdot\mathfrak{f}\biggl{(}R_{{\mathcal{W}}}(\theta_{2},\lambda^{-2r}\sigma(\alpha)y_{2}),\frac{C}{\lambda^{2r}y_{1}y_{2}|\mathfrak{J}_{1}|}\biggr{)}.

Note also that (α,σ(α))𝔄𝜽,𝒚(s)(\alpha,\sigma(\alpha))\in\mathfrak{A}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}}}(s) holds if and only if both αy11s1/2λ2r(θ1)\alpha\in y_{1}^{-1}s^{1/2}\lambda^{-2r}\ell(\theta_{1}) and σ(α)y21λ2r𝒲(θ2)\sigma(\alpha)\in y_{2}^{-1}\lambda^{2r}\ell_{{\mathcal{W}}}(\theta_{2}), and we recall that the first of these conditions is equivalent with 𝔍1\mathfrak{J}_{1}\neq\emptyset, while, by (3.2), the second condition is equivalent with R𝒲(θ2,λ2ry2σ(α))R_{{\mathcal{W}}}(\theta_{2},\lambda^{-2r}y_{2}\sigma(\alpha))\neq\emptyset. Using these facts together with our bound on the integral in (3.60), we obtain:

GE,i(s)s3/2αΛZJc1s1/4c2s1/4I(𝔍1)c1s1/4c2s1/4I(R𝒲(θ2,λ2ry2σ(α)))\displaystyle G_{E,i}(s)\ll s^{-3/2}\sum_{\alpha\in\Lambda}\int_{Z_{J}}\int_{c_{1}s^{1/4}}^{c_{2}s^{1/4}}I(\mathfrak{J}_{1}\neq\emptyset)\int_{c_{1}s^{1/4}}^{c_{2}s^{1/4}}I\bigl{(}R_{{\mathcal{W}}}(\theta_{2},\lambda^{-2r}y_{2}\sigma(\alpha))\neq\emptyset\bigr{)}\hskip 80.0pt
(3.61) ×𝔣(R𝒲(θ2,λ2rσ(α)y2),Cλ2ry1y2|𝔍1|)dy2dy1d𝜽.\displaystyle\times\mathfrak{f}\biggl{(}R_{{\mathcal{W}}}(\theta_{2},\lambda^{-2r}\sigma(\alpha)y_{2}),\frac{C}{\lambda^{2r}y_{1}y_{2}|\mathfrak{J}_{1}|}\biggr{)}\,\text{d}y_{2}\,\text{d}y_{1}\,\text{d}\bm{\theta}.

Here it should be recalled that 𝔍1=𝔍1(s,y1,θ1,α)\mathfrak{J}_{1}=\mathfrak{J}_{1}(s,y_{1},\theta_{1},\alpha); see (3.58). Now note that for any y1,y2y_{1},y_{2} appearing in the above integral, we have λ2ry1y2c12λ2s3/4\lambda^{2r}y_{1}y_{2}\geq c_{1}^{2}\lambda^{-2}s^{3/4}; hence if we set C:=Cλ2c12C^{\prime}:=C\lambda^{2}c_{1}^{-2} then Cλ2ry1y2|J1|Cs3/4|J1|{\displaystyle\frac{C}{\lambda^{2r}y_{1}y_{2}|J_{1}|}\leq\frac{C^{\prime}}{s^{3/4}|J_{1}|}}, and so, using Lemma 3.10 and then substituting y2=λ2rσ(α)1yy_{2}=\lambda^{2r}\sigma(\alpha)^{-1}y, we obtain (since Λ\Lambda is a fixed finite subset of Ii{0}I_{i}\setminus\{0\}):

(3.62) GE,i(s)s5/4αΛZJc1s1/4c2s1/4I(𝔍1)𝔣~(R𝒲(θ2,y),Cs3/4|𝔍1|)dydy1d𝜽,\displaystyle G_{E,i}(s)\ll s^{-5/4}\sum_{\alpha\in\Lambda}\int_{Z_{J}}\int_{c_{1}s^{1/4}}^{c_{2}s^{1/4}}I(\mathfrak{J}_{1}\neq\emptyset)\int_{\mathbb{R}}\widetilde{\mathfrak{f}}\biggl{(}R_{{\mathcal{W}}}(\theta_{2},y),\frac{C^{\prime}}{s^{3/4}|\mathfrak{J}_{1}|}\biggr{)}\,\text{d}y\,\text{d}y_{1}\,\text{d}\bm{\theta},

where we use the notation 𝔣~\widetilde{\mathfrak{f}} defined in (3.5).

Now we have

𝔍1{0<y1α<s12λ2rc𝒫12(cosθ1+sinθ1)if θ1(0,π4),s12λ2rc𝒫12(cosθ1sinθ1)<y1α<0if θ1(3π4,π),\displaystyle\mathfrak{J}_{1}\neq\emptyset\quad\Leftrightarrow\quad\left\{\begin{array}[]{ll}0<y_{1}\alpha<s^{\frac{1}{2}}\lambda^{-2r}c_{{\mathcal{P}}}^{-\frac{1}{2}}(\cos\theta_{1}+\sin\theta_{1})&\text{if $\theta_{1}\in(0,\tfrac{\pi}{4})$},\\ s^{\frac{1}{2}}\lambda^{-2r}c_{{\mathcal{P}}}^{-\frac{1}{2}}(\cos\theta_{1}-\sin\theta_{1})<y_{1}\alpha<0&\text{if $\theta_{1}\in(\tfrac{3\pi}{4},\pi)$},\end{array}\right.

and it follows from the definition in (3.58) that for any αΛ\alpha\in\Lambda, θ1J\theta_{1}\in J and y1c1s1/4y_{1}\geq c_{1}s^{1/4} for which 𝔍1\mathfrak{J}_{1}\neq\emptyset holds, we have |𝔍1|C′′u1|\mathfrak{J}_{1}|\geq C^{\prime\prime}u_{1} with u1:=s12λ2rc𝒫12(|cosθ1|+sinθ1)y1|α|>0u_{1}:=s^{\frac{1}{2}}\lambda^{-2r}c_{{\mathcal{P}}}^{-\frac{1}{2}}(|\cos\theta_{1}|+\sin\theta_{1})-y_{1}|\alpha|\in\mathbb{R}_{>0}, where C′′>0C^{\prime\prime}>0 is a certain absolute constant. Furthermore, u1<c3s1/4u_{1}<c_{3}s^{1/4} with c3:=2c𝒫12λ2c_{3}:=2c_{{\mathcal{P}}}^{-\frac{1}{2}}\lambda^{2}. Hence, using Lemma 3.10,

GE,i(s)\displaystyle G_{E,i}(s) s5/4αΛZJ0c3s1/4𝔣~(R𝒲(θ2,y),CC′′s3/4u1)dydu1d𝜽.\displaystyle\ll s^{-5/4}\sum_{\alpha\in\Lambda}\int_{Z_{J}}\int_{0}^{c_{3}s^{1/4}}\int_{\mathbb{R}}\widetilde{\mathfrak{f}}\biggl{(}R_{{\mathcal{W}}}(\theta_{2},y),\frac{C^{\prime}}{C^{\prime\prime}s^{3/4}u_{1}}\biggr{)}\,\text{d}y\,\text{d}u_{1}\,\text{d}\bm{\theta}.

Here the integrand is independent of θ1\theta_{1} and α\alpha. Hence, using the fact that Λ\Lambda is a fixed finite set, and substituting u1=(C/C′′)s3/4uu_{1}=(C^{\prime}/C^{\prime\prime})s^{-3/4}u, we conclude:

(3.63) GE,i(s)s20(c3C′′/C)s02π𝔣~(R𝒲(θ2,y),u1)dydθ2du.\displaystyle G_{E,i}(s)\ll s^{-2}\int_{0}^{(c_{3}C^{\prime\prime}/C^{\prime})s}\int_{0}^{2\pi}\int_{\mathbb{R}}\widetilde{\mathfrak{f}}\bigl{(}R_{{\mathcal{W}}}(\theta_{2},y),u^{-1}\bigr{)}\,\text{d}y\,\text{d}\theta_{2}\,\text{d}u.

Hence we have proved the error bound in (3.7) (with C:=c3C′′/CC:=c_{3}C^{\prime\prime}/C^{\prime}). As we have already noted, because of G(s)=i=1κGi(s)=i=1κ(GM,i(s)+GE,i(s))G(s)=\sum_{i=1}^{\kappa}G_{i}(s)=\sum_{i=1}^{\kappa}\bigl{(}G_{M,i}(s)+G_{E,i}(s)\bigr{)} and Proposition 3.8, the bound (3.63) completes the proof of Theorem 4. \square

4. More explicit formula for the leading coefficient

The main goal of this section is to prove the following more explicit formula for the leading coefficient a𝒫a_{\mathcal{P}} in Theorem 2, when we further assume that 𝒲{\mathcal{W}} is such that 𝒲(θ)\ell_{{\mathcal{W}}}(\theta) is an interval for every θ\theta. Recall here that for any θ/2π\theta\in\mathbb{R}/2\pi\mathbb{Z}, 𝒲(θ)\ell_{\mathcal{W}}(\theta) is the projection of 𝒲kθ\mathcal{W}\text{k}_{-\theta} on the xx-axis.

Theorem 5.

Keep the notation and assumptions as in Theorem 2 and further assume that 𝒲(θ)\ell_{\mathcal{W}}(\theta) is an interval for each θ\theta, parametrized as in (1.16), i.e. 𝒲(θ)=r(θ)(ν(θ),1)\ell_{\mathcal{W}}(\theta)=r(\theta)(-\nu(\theta),1). Then for each IiI_{i} (1iκ1\leq i\leq\kappa), there exist a finite partition >0=j=1lIiSIi,j\mathbb{R}_{>0}=\bigsqcup_{j=1}^{l_{I_{i}}}S_{I_{i},j} of >0\mathbb{R}_{>0} into intervals, and non-negative constants AIi,j,BIi,jA_{I_{i},j},B_{I_{i},j} (1jlIi)(1\leq j\leq l_{I_{i}}), such that

(4.1) a𝒫=Area(𝒲)4ΔK2ζK(2)i=1κNr(Ii)2j=1lIiS~Ii,jr(θ)2(AIi,j+BIi,jν(θ)2)dθ,\displaystyle a_{{\mathcal{P}}}=\frac{\operatorname{Area}({\mathcal{W}})}{4\Delta_{K}^{2}\zeta_{K}(2)}\sum_{i=1}^{\kappa}\operatorname{Nr}(I_{i})^{-2}\sum_{j=1}^{l_{I_{i}}}\int_{\tilde{S}_{I_{i},j}}r(\theta)^{-2}\left(A_{I_{i},j}+B_{I_{i},j}\nu(\theta)^{-2}\right)\,\mathrm{d}\theta,

where S~Ii,j:={θ[0,2π):ν(θ)SIi,j}\tilde{S}_{I_{i},j}:=\{\theta\in[0,2\pi)\>:\>\nu(\theta)\in S_{I_{i},j}\}.

Remark 4.2.

For each IiI_{i}, the partition >0=j=1lIiSIi,j\mathbb{R}_{>0}=\bigsqcup_{j=1}^{l_{I_{i}}}S_{I_{i},j} and the constants AIi,j,BIi,jA_{I_{i},j},B_{I_{i},j} are all computable from our analysis; see Section 4.4 for three explicit examples. We stress that the intervals SIi,jS_{I_{i},j} are allowed to be open, closed, or half-open, and may be degenerate, i.e. of the form [a,a]={a}[a,a]=\{a\} for some a>0a>0.

4.1. Relations to the polar set of 𝒲\mathcal{W}

Let 𝒲2\mathcal{W}\subset\mathbb{R}^{2} be as above. The integral appearing in (4.1) is closely related to the polar set 𝒲\mathcal{W}^{*} of 𝒲\mathcal{W}, which is defined by (1.21). Note that 𝒲\mathcal{W}^{*} is closed and convex; furthermore, since 𝒲\mathcal{W} is bounded and contains 𝟎\mathbf{0} in its interior, 𝒲\mathcal{W}^{*} is also bounded and contains 𝟎\mathbf{0} in its interior. Moreover, if 𝒲{\mathcal{W}} is symmetric about the origin, then so is 𝒲{\mathcal{W}}^{*}; and if 𝒲{\mathcal{W}} is a convex polygon, then also 𝒲{\mathcal{W}}^{*} is a convex polygon.

The following lemma shows that 𝒲\mathcal{W}^{*} can be parameterized in polar coordinates in terms of the function r(θ)r(\theta).

Lemma 4.1.

Let 𝒲{\mathcal{W}} be as in Theorem 5. Then 𝒲\mathcal{W}^{*} is parameterized in polar coordinates by

𝒲={(t,θ)0×/2π: 0tr(θ)1}.\displaystyle\mathcal{W}^{*}=\left\{(t,\theta)\in\mathbb{R}_{\geq 0}\times\mathbb{R}/2\pi\mathbb{Z}\>:\>0\leq t\leq r(-\theta)^{-1}\right\}.
Proof.

Note that for any θ/2π\theta\in\mathbb{R}/2\pi\mathbb{Z},

𝒲kθ={(w1cosθw2sinθ,w1sinθ+w2cosθ):(w1,w2)𝒲}.\displaystyle\mathcal{W}\text{k}_{-\theta}=\left\{(w_{1}\cos\theta-w_{2}\sin\theta,w_{1}\sin\theta+w_{2}\cos\theta)\>:\>(w_{1},w_{2})\in\mathcal{W}\right\}.

Thus

𝒲(θ)={w1cosθw2sinθ:(w1,w2)𝒲}.\displaystyle\ell_{\mathcal{W}}(\theta)=\left\{w_{1}\cos\theta-w_{2}\sin\theta\>:\>(w_{1},w_{2})\in\mathcal{W}\right\}.

It then follows from the definition of polar set and the relation 𝒲(θ)0=[0,r(θ))\ell_{\mathcal{W}}(\theta)\cap\mathbb{R}_{\geq 0}=[0,r(\theta)), that the set

{t0:t(cosθ,sinθ)𝒲}\displaystyle\left\{t\in\mathbb{R}_{\geq 0}\>:\>t(\cos\theta,-\sin\theta)\in{\mathcal{W}}^{*}\right\}

equals the interval [0,r(θ)1][0,r(\theta)^{-1}]. The lemma then follows by a substitution θθ\theta\mapsto-\theta. ∎

Remark 4.3.

As a direct consequence of Lemma 4.1 and the relation

(4.4) r(θ+π)=r(θ)ν(θ),θ/2π,\displaystyle r(\theta+\pi)=r(\theta)\nu(\theta),\quad\forall\,\theta\in\mathbb{R}/2\pi\mathbb{Z},

(which in turn follows from the symmetry 𝒲(θ+π)=𝒲(θ)\ell_{\mathcal{W}}(\theta+\pi)=-\ell_{\mathcal{W}}(\theta)), we have

(4.5) 2Area(𝒲)=02πr(θ)2dθ=02πν(θ)2r(θ)2dθ.\displaystyle 2\operatorname{Area}({\mathcal{W}}^{*})=\int_{0}^{2\pi}r(\theta)^{-2}\,\text{d}\theta=\int_{0}^{2\pi}\nu(\theta)^{-2}r(\theta)^{-2}\,\text{d}\theta.

For each 1iκ1\leq i\leq\kappa, let 1jilIi1\leq j_{i}\leq l_{I_{i}} be the unique index such that 1SIi,ji1\in S_{I_{i},j_{i}}. In the special case when S~Ii,ji=[0,2π)\tilde{S}_{I_{i},j_{i}}=[0,2\pi) for every 1iκ1\leq i\leq\kappa (this happens if 𝒲\mathcal{W} is symmetric with respect to the origin, or a sufficiently small translate of such a set), then by (4.5) the formula (4.1) simplifies to

(4.6) a𝒫\displaystyle a_{\mathcal{P}} =Area(𝒲)Area(𝒲)ΔK2ζK(2)i=1κCIiNr(Ii)2,\displaystyle=\frac{\operatorname{Area}({\mathcal{W}})\operatorname{Area}({\mathcal{W}}^{*})}{\Delta_{K}^{2}\,\zeta_{K}(2)}\sum_{i=1}^{\kappa}C_{I_{i}}\text{Nr}(I_{i})^{-2},

with CIi:=12(AIi,ji+BIi,ji)C_{I_{i}}:=\frac{1}{2}(A_{I_{i},j_{i}}+B_{I_{i},j_{i}}). More generally, the sum of integrals appearing in (4.1) can be understood as a weighted area of 𝒲\mathcal{W}^{*}.

The remainder of this section is devoted to proving Theorem 5.

4.2. Some preliminary computations

We start our proof with the formula for a𝒫a_{\mathcal{P}} given in Theorem 4 which states that a𝒫=i=1κa𝒫,ia_{\mathcal{P}}=\sum_{i=1}^{\kappa}a_{\mathcal{P},i} with

(4.7) a𝒫,i\displaystyle a_{\mathcal{P},i} =Area(𝒲)4ΔK2ζK(2)Nr(Ii)202π1λ2αIi,𝒲(θ)(y)2dyy3dθ.\displaystyle=\frac{\operatorname{Area}({\mathcal{W}})}{4\Delta_{K}^{2}\zeta_{K}(2)}\text{Nr}(I_{i})^{-2}\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}\alpha_{I_{i},\ell_{\mathcal{W}}(\theta)}(y)^{2}\,\frac{\text{d}y}{y^{3}}\,\text{d}\theta.

It thus suffices to compute the above double integral which we denote by 𝒫,i\mathcal{I}_{\mathcal{P},i}. It is immediate from the definition (3.3) that αI,aJ(ay)=αI,J(y)\alpha_{I,aJ}(ay)=\alpha_{I,J}(y) for all a,y>0a,y>0. Thus, defining Jν:=(ν,1)J_{\nu}:=(-\nu,1) for ν>0\nu>0 and applying (3.10), we have:

(4.8) 𝒫,i\displaystyle\mathcal{I}_{\mathcal{P},i} =02π1λ2αIi,Jν(θ)(yr(θ))2dyy3dθ=02πr(θ)21λ2αIi,Jν(θ)(y)2dyy3dθ.\displaystyle=\int_{0}^{2\pi}\int_{1}^{\lambda^{2}}\alpha_{I_{i},J_{\nu(\theta)}}\Bigl{(}\frac{y}{r(\theta)}\Bigr{)}^{2}\,\frac{\text{d}y}{y^{3}}\,\text{d}\theta=\int_{0}^{2\pi}r(\theta)^{-2}\int_{1}^{\lambda^{2}}\alpha_{I_{i},J_{\nu(\theta)}}(y)^{2}\,\frac{\text{d}y}{y^{3}}\,\text{d}\theta.

We wish to compute the innermost integral in the last expression.

4.3. Extremal points

Fix an integral ideal I𝒪KI\subset\mathcal{O}_{K}. As mentioned above, we are interested in computing the integral

(4.9) 1λ2αI,Jν(y)2dyy3\displaystyle\int_{1}^{\lambda^{2}}\alpha_{I,J_{\nu}}(y)^{2}\,\frac{\text{d}y}{y^{3}}

for any ν>0\nu>0. For this we introduce a notion which captures the possible values of the function αI,Jν\alpha_{I,J_{\nu}}. For any ν>0\nu>0 and αI>0\alpha\in I\cap\mathbb{R}_{>0}, let bα,ν:=σ(α)b_{\alpha,\nu}:=\sigma(\alpha) if σ(α)>0\sigma(\alpha)>0 and bα,ν:=ν1σ(α)b_{\alpha,\nu}:=\nu^{-1}\sigma(\alpha) if σ(α)<0\sigma(\alpha)<0. Note that for any y>0y>0, yσ(α)Jνy\sigma(\alpha)\in J_{\nu} if and only if y<|bα,ν|1y<|b_{\alpha,\nu}|^{-1}. We say αI>0\alpha\in I\cap\mathbb{R}_{>0} is ν\nu-extremal with respect to II if

ι(I)Rα,ν=,\iota(I)\cap R_{\alpha,\nu}=\emptyset,

where

Rα,ν:={(0,α)×[νσ(α),σ(α)]if σ(α)>0,(0,α)×[σ(α),ν1|σ(α)|]if σ(α)<0.\displaystyle R_{\alpha,\nu}:=\left\{\begin{array}[]{ll}(0,\alpha)\times[-\nu\sigma(\alpha),\sigma(\alpha)]&\text{if $\sigma(\alpha)>0$},\\ (0,\alpha)\times[\sigma(\alpha),\nu^{-1}|\sigma(\alpha)|]&\text{if $\sigma(\alpha)<0$}.\end{array}\right.

We denote by EI,νI>0E_{I,\nu}\subset I\cap\mathbb{R}_{>0} the set of all ν\nu-extremal points with respect to II. Below we will simply call elements in EI,νE_{I,\nu} ν\nu-extremal when there is no ambiguity. Since ι(I)\iota(I) is a lattice in 2\mathbb{R}^{2}, EI,νE_{I,\nu} is always nonempty. Indeed, EI,νE_{I,\nu} is always infinite, as it is invariant under a certain multiplication map as shown in the following lemma.

Lemma 4.2.

The set EI,νE_{I,\nu} is invariant under the map S:>0>0S:\mathbb{R}_{>0}\to\mathbb{R}_{>0} defined by S(y):=λyS(y):=\lambda y if σ(λ)>0\sigma(\lambda)>0 and S(y):=λ2yS(y):=\lambda^{2}y if σ(λ)<0\sigma(\lambda)<0.

Proof.

Suppose that αI>0\alpha\in I\cap\mathbb{R}_{>0} is ν\nu-extremal, i.e. ι(I)Rα,ν=\iota(I)\cap R_{\alpha,\nu}=\emptyset with Rα,νR_{\alpha,\nu} the rectangle given as above. Let dλ=diag(λ,σ(λ))d_{\lambda}=\operatorname{diag}(\lambda,\sigma(\lambda)) and note that ι(I)\iota(I) is invariant under the right multiplication action of dλd_{\lambda}. The lemma then follows by noting that Rα,νdλ=Rλα,νR_{\alpha,\nu}d_{\lambda}=R_{\lambda\alpha,\nu} if σ(λ)>0\sigma(\lambda)>0 and Rα,νdλ2=Rλ2α,νR_{\alpha,\nu}d^{2}_{\lambda}=R_{\lambda^{2}\alpha,\nu} if σ(λ)<0\sigma(\lambda)<0. ∎

Remark 4.10.

The set EI,1E_{I,1} is invariant under the map yλyy\mapsto\lambda y even if σ(λ)<0\sigma(\lambda)<0. This is because we always have Rα,1dλ=Rλα,1R_{\alpha,1}d_{\lambda}=R_{\lambda\alpha,1}, since Rα,1=(0,α)×[|σ(α)|,|σ(α)|]R_{\alpha,1}=(0,\alpha)\times[-|\sigma(\alpha)|,|\sigma(\alpha)|].

Note that αI>0\alpha\in I\cap\mathbb{R}_{>0} is ν\nu-extremal if and only if

(4.11) |bβ,ν|>|bα,ν||b_{\beta,\nu}|>|b_{\alpha,\nu}| for any βI>0\beta\in I\cap\mathbb{R}_{>0} with β<α\beta<\alpha.

In particular, if α<β\alpha<\beta are two ν\nu-extremal points, then |bα,ν|>|bβ,ν||b_{\alpha,\nu}|>|b_{\beta,\nu}|. Moreover, one can check from (4.11) that αI,Jν\alpha_{I,J_{\nu}} takes values only on ν\nu-extremal points. Using these facts, we can now prove the following lemma which enables us to compute the integral (4.9).

Lemma 4.3.

For any ν>0\nu>0, let 0<β<α0<\beta<\alpha be two consecutive elements in EI,νE_{I,\nu}. Then αI,Jν(y)=α\alpha_{I,J_{\nu}}(y)=\alpha for any |bβ,ν|1y<|bα,ν|1|b_{\beta,\nu}|^{-1}\leq y<|b_{\alpha,\nu}|^{-1}.

Remark 4.12.

By discreteness of ι(I)\iota(I) one sees that #(EI,ν/S)<\#(E_{I,\nu}/S)<\infty for any ν>0\nu>0; thus we can take two consecutive elements from EI,νE_{I,\nu} as in the above lemma.

Proof of Lemma 4.3.

Take any |bβ,ν|1y<|bα,ν|1|b_{\beta,\nu}|^{-1}\leq y<|b_{\alpha,\nu}|^{-1} and let β=αI,Jν(y)\beta^{\prime}=\alpha_{I,J_{\nu}}(y). In particular, β\beta^{\prime} is ν\nu-extremal and yσ(β)Jνy\sigma(\beta^{\prime})\in J_{\nu}, and the last condition is equivalent to y<|bβ,ν|1y<|b_{\beta^{\prime},\nu}|^{-1}. Moreover, y<|bα,ν|1y<|b_{\alpha,\nu}|^{-1} implies that yσ(α)Jνy\sigma(\alpha)\in J_{\nu}. Thus βα\beta^{\prime}\leq\alpha. If β<α\beta^{\prime}<\alpha, then since β<α\beta<\alpha are two consecutive ν\nu-extremal points, we must have ββ\beta^{\prime}\leq\beta and therefore |bβ,ν||bβ,ν||b_{\beta^{\prime},\nu}|\geq|b_{\beta,\nu}|, which contradicts the fact that |bβ,ν|1y<|bβ,ν|1|b_{\beta,\nu}|^{-1}\leq y<|b_{\beta^{\prime},\nu}|^{-1}. Hence β=α\beta^{\prime}=\alpha as desired. ∎

We can now compute the integral 1λ2αI,Jν(y)2dyy3\int_{1}^{\lambda^{2}}\alpha_{I,J_{\nu}}(y)^{2}\ \frac{\text{d}y}{y^{3}} in terms of the set EI,νE_{I,\nu}.

Lemma 4.4.

For any ν>0\nu>0 and α0EI,ν\alpha_{0}\in E_{I,\nu}, if α0<α1<<αl=λ2α0\alpha_{0}<\alpha_{1}<\cdots<\alpha_{l}=\lambda^{2}\alpha_{0} is the complete list of points in EI,ν[α0,λ2α0]E_{I,\nu}\cap[\alpha_{0},\lambda^{2}\alpha_{0}], then

1λ2αI,Jν(y)2dyy3\displaystyle\int_{1}^{\lambda^{2}}\alpha_{I,J_{\nu}}(y)^{2}\,\frac{\mathrm{d}y}{y^{3}} =AI,ν+BI,νν2,\displaystyle=A_{I,\nu}+B_{I,\nu}\nu^{-2},

where

AI,ν:=12j=0(σ(αj)>0)l1σ(αj)2(αj+12αj2)andBI,ν:=12j=0(σ(αj)<0)l1σ(αj)2(αj+12αj2).A_{I,\nu}:=\frac{1}{2}\sum_{\begin{subarray}{c}j=0\\ (\sigma(\alpha_{j})>0)\end{subarray}}^{l-1}\sigma(\alpha_{j})^{2}\bigl{(}\alpha_{j+1}^{2}-\alpha_{j}^{2}\bigr{)}\quad\text{and}\quad B_{I,\nu}:=\frac{1}{2}\sum_{\begin{subarray}{c}j=0\\ (\sigma(\alpha_{j})<0)\end{subarray}}^{l-1}\sigma(\alpha_{j})^{2}\bigl{(}\alpha_{j+1}^{2}-\alpha_{j}^{2}\bigr{)}.
Proof.

We have |bα0,ν|1<|bα1,ν|1<<|bαl,ν|1=|bα0,ν|1λ2|b_{\alpha_{0},\nu}|^{-1}<|b_{\alpha_{1},\nu}|^{-1}<\cdots<|b_{\alpha_{l},\nu}|^{-1}=|b_{\alpha_{0},\nu}|^{-1}\lambda^{2}. Hence by (3.10) and Lemma 4.3,

1λ2αI,Jν(y)2dyy3\displaystyle\int_{1}^{\lambda^{2}}\alpha_{I,J_{\nu}}(y)^{2}\,\frac{\text{d}y}{y^{3}} =j=1l|bαj1,ν|1|bαj,ν|1αj2dyy3=12j=1lαj2(bαj1,ν2bαj,ν2).\displaystyle=\sum_{j=1}^{l}\int_{|b_{\alpha_{j-1},\nu}|^{-1}}^{|b_{\alpha_{j},\nu}|^{-1}}\alpha_{j}^{2}\,\frac{\text{d}y}{y^{3}}=\frac{1}{2}\sum_{j=1}^{l}\alpha_{j}^{2}\left({b^{2}_{\alpha_{j-1},\nu}}-{b^{2}_{\alpha_{j},\nu}}\right).

The lemma then follows by rewriting the above as (using the fact that αl2bαl,ν2=α02bα0,ν2\alpha_{l}^{2}b_{\alpha_{l},\nu}^{2}=\alpha_{0}^{2}b_{\alpha_{0},\nu}^{2}):

12j=0l1bαj,ν2(αj+12αj2)=12j=0(σ(αj)>0)l1σ(αj)2(αj+12αj2)+ν22j=0(σ(αj)<0)l1σ(αj)2(αj+12αj2).\displaystyle\frac{1}{2}\sum_{j=0}^{l-1}b_{\alpha_{j},\nu}^{2}\bigl{(}\alpha_{j+1}^{2}-\alpha_{j}^{2}\bigr{)}=\frac{1}{2}\sum_{\begin{subarray}{c}j=0\\ (\sigma(\alpha_{j})>0)\end{subarray}}^{l-1}\sigma(\alpha_{j})^{2}\bigl{(}\alpha_{j+1}^{2}-\alpha_{j}^{2}\bigr{)}+\frac{\nu^{-2}}{2}\sum_{\begin{subarray}{c}j=0\\ (\sigma(\alpha_{j})<0)\end{subarray}}^{l-1}\sigma(\alpha_{j})^{2}\bigl{(}\alpha_{j+1}^{2}-\alpha_{j}^{2}\bigr{)}.

Remark 4.13.

If α<β\alpha<\beta are two consecutive ν\nu-extremal points, then so are λ2nα<λ2nβ\lambda^{2n}\alpha<\lambda^{2n}\beta for any integer nn. Using the equality σ(α)2(β2α2)=σ(λ2nα)2((λ2nβ)2(λ2nα)2)\sigma({\alpha})^{2}(\beta^{2}-\alpha^{2})=\sigma({\lambda^{2n}\alpha})^{2}\left((\lambda^{2n}\beta)^{2}-(\lambda^{2n}\alpha)^{2}\right) one easily sees that both AI,νA_{I,\nu} and BI,νB_{I,\nu} are independent of the choice of α0\alpha_{0}.

Next, we study how EI,νE_{I,\nu} varies in ν\nu. Our goal is to show there are only finitely many possibilities for EI,νE_{I,\nu} (and hence also for the coefficients AI,νA_{I,\nu} and BI,νB_{I,\nu}). For this we introduce another related notion. We say αI>0\alpha\in I\cap\mathbb{R}_{>0} is positively (resp. negatively) extremal with respect to II if σ(α)>0\sigma(\alpha)>0 (resp. σ(α)<0\sigma(\alpha)<0) and ι(I)(0,α)×[0,σ(α)]=\iota(I)\cap(0,\alpha)\times[0,\sigma(\alpha)]=\emptyset (resp. ι(I)(0,α)×[σ(α),0]=\iota(I)\cap(0,\alpha)\times[\sigma(\alpha),0]=\emptyset). We denote by EI+E^{+}_{I} and EIE_{I}^{-} the set of positively and negatively extremal points in II. Recall that SS is the map yλyy\mapsto\lambda y if σ(λ)>0\sigma(\lambda)>0 and the map yλ2yy\mapsto\lambda^{2}y if σ(λ)<0\sigma(\lambda)<0. It is clear that EI±E_{I}^{\pm} are invariant under SS, and from the discreteness of ι(I)\iota(I) we see that #(EI±/S)<\#(E_{I}^{\pm}/S)<\infty. Note that a necessary condition for αI>0\alpha\in I\cap\mathbb{R}_{>0} to be ν\nu-extremal for some ν>0\nu>0 is that αEI+EI\alpha\in E_{I}^{+}\sqcup E_{I}^{-}. Indeed this is also a sufficient condition in the sense that for any αEI+EI\alpha\in E_{I}^{+}\sqcup E_{I}^{-} there exists some ν>0\nu>0 such that αEI,ν\alpha\in E_{I,\nu}; see Lemma 4.6 below.

For any ν>0\nu>0, set EI,ν±:=EI,νEI±E_{I,\nu}^{\pm}:=E_{I,\nu}\cap E_{I}^{\pm} so that EI,ν=EI,ν+EI,νE_{I,\nu}=E^{+}_{I,\nu}\sqcup E_{I,\nu}^{-}. Note that EI,ν+E_{I,\nu}^{+} (resp. EI,νE_{I,\nu}^{-}) is decreasing (resp. increasing) in ν\nu, and there is a constant C=CI>0C=C_{I}>0 such that EI,ν+=E_{I,\nu}^{+}=\emptyset for all ν>C\nu>C and EI,ν=E_{I,\nu}^{-}=\emptyset for all ν<1/C\nu<1/C, because of Lemma 3.6 and the fact that

Area(Rα,ν)=(1+νsgn(σ(α)))|ασ(α)|>νsgn(σ(α))(αI{0}).\operatorname{Area}(R_{\alpha,\nu})=(1+\nu^{\operatorname{sgn}(\sigma(\alpha))})|\alpha\sigma(\alpha)|>\nu^{\operatorname{sgn}(\sigma(\alpha))}\qquad(\forall\,\alpha\in I\setminus\{0\}).

To understand how EI,νE_{I,\nu} varies in ν\nu we study how EI,ν±E_{I,\nu}^{\pm} varies in ν\nu which is easier as it is monotone. We first study this property for a single element in II. The following lemma is immediate from our definitions.

Lemma 4.5.

For any αEI+\alpha\in E_{I}^{+} let

να:=sup{ν>0:ι(I)(0,α)×[νσ(α),σ(α)]=}.\nu_{\alpha}:=\sup\{\nu>0\>:\>\iota(I)\cap(0,\alpha)\times[-\nu\sigma(\alpha),\sigma(\alpha)]=\emptyset\}.

Then α\alpha is ν\nu-extremal if and only if 0<ν<να0<\nu<\nu_{\alpha}. Similarly, for any αEI\alpha\in E_{I}^{-} define

να:=inf{ν>0:ι(I)(0,α)×[σ(α),ν1|σ(α)|]=}.\nu_{\alpha}:=\inf\{\nu>0\>:\>\iota(I)\cap(0,\alpha)\times[\sigma(\alpha),\nu^{-1}|\sigma(\alpha)|]=\emptyset\}.

Then α\alpha is ν\nu-extremal if and only if ν>να\nu>\nu_{\alpha}.

Remark 4.14.

For each αEI+EI\alpha\in E_{I}^{+}\sqcup E_{I}^{-}, when ν\nu attains the critical value να\nu_{\alpha}, α\alpha fails to be να\nu_{\alpha}-extremal. This fact, together with the finiteness of (EI+EI)/S(E_{I}^{+}\sqcup E_{I}^{-})/S and the fact that EI,ν+E_{I,\nu}^{+} (resp. EI,νE_{I,\nu}^{-}) is decreasing (resp. increasing) in ν\nu implies that EI,ν+E_{I,\nu}^{+} (resp. EIE_{I}^{-}) is right (resp. left) continuous in the sense that for any ν>0\nu>0, there exists some ϵ>0\epsilon>0 such that EI,ν+=EI,ν+E_{I,\nu^{\prime}}^{+}=E_{I,\nu}^{+} for any ν(ν,ν+ϵ)\nu^{\prime}\in(\nu,\nu+\epsilon) (resp. ν(νϵ,ν)\nu^{\prime}\in(\nu-\epsilon,\nu)). Note also that

να={inf{σ(β)σ(α):βI, 0<β<α,σ(β)<0}if αEI+,sup{σ(α)σ(β):βI,0<β<α,σ(β)>0}if αEI.\displaystyle\nu_{\alpha}=\left\{\begin{array}[]{ll}\inf\left\{\tfrac{-\sigma(\beta)}{\sigma(\alpha)}\>:\>\beta\in I,\,0<\beta<\alpha,\,\sigma(\beta)<0\right\}&\text{if $\alpha\in E_{I}^{+}$},\\ \sup\left\{\tfrac{-\sigma(\alpha)}{\sigma(\beta)}\>:\>\beta\in I,0<\beta<\alpha,\,\sigma(\beta)>0\right\}&\text{if $\alpha\in E_{I}^{-}$}.\end{array}\right.

From this formula we see that the β\beta attaining the above infimum (resp. supremum) is an element in EIE_{I}^{-} (resp. EI+E_{I}^{+}). This gives very strict restrictions on the possible values of these να\nu_{\alpha}’s. In particular, να\nu_{\alpha} can not be 11 since otherwise we would have β=α\beta=-\alpha, contradicting 0<β<α0<\beta<\alpha. Moreover, the set of values {να:αEI+EI}\{\nu_{\alpha}\>:\>\alpha\in E_{I}^{+}\sqcup E_{I}^{-}\} is finite, since #(EI+EI)/S<\#(E_{I}^{+}\sqcup E_{I}^{-})/S<\infty and νλ2α=να\nu_{\lambda^{2}\alpha}=\nu_{\alpha} for every αEI+EI\alpha\in E_{I}^{+}\sqcup E_{I}^{-}.

Lemma 4.6.

There exist finitely many constants 0<ν1+<<νl+<0<\nu_{1}^{+}<\cdots<\nu_{l}^{+}<\infty such that EI+EI,ν1++EI,ν2++EI,νl++=E_{I}^{+}\supsetneq E_{I,\nu_{1}^{+}}^{+}\supsetneq E_{I,\nu_{2}^{+}}^{+}\supsetneq\cdots\supsetneq E_{I,\nu_{l}^{+}}^{+}=\emptyset and

EI,ν+\displaystyle E_{I,\nu}^{+} ={EI+if 0<ν<ν1+,EI,νi++if νi+ν<νi+1+ (1i<l),if ννl+.\displaystyle=\left\{\begin{array}[]{ll}E_{I}^{+}&\text{if $0<\nu<\nu_{1}^{+}$},\\ E^{+}_{I,\nu_{i}^{+}}&\text{if $\nu_{i}^{+}\leq\nu<\nu_{i+1}^{+}$\quad($1\leq i<l$)},\\ \emptyset&\text{if $\nu\geq\nu_{l}^{+}$}.\end{array}\right.

Similarly, there exist >ν1>>νl>0\infty>\nu_{1}^{-}>\cdots>\nu_{l^{\prime}}^{-}>0 such that EIEI,ν1EI,ν2EI,νl=E_{I}^{-}\supsetneq E_{I,\nu_{1}^{-}}^{-}\supsetneq E_{I,\nu_{2}^{-}}^{-}\supsetneq\cdots\supsetneq E_{I,\nu_{l^{\prime}}^{-}}^{-}=\emptyset and

EI,ν\displaystyle E_{I,\nu}^{-} ={EIif ν>ν1,EI,νiif νi+1<ννi (1i<l),if 0<ννl.\displaystyle=\left\{\begin{array}[]{ll}E_{I}^{-}&\text{if $\nu>\nu_{1}^{-}$},\\ E^{-}_{I,\nu_{i}^{-}}&\text{if $\nu_{i+1}^{-}<\nu\leq\nu_{i}^{-}$\quad($1\leq i<l^{\prime}$)},\\ \emptyset&\text{if $0<\nu\leq\nu_{l^{\prime}}^{-}$}.\end{array}\right.
Proof.

We noted in Remark 4.14 that the set {να:αEI+}\{\nu_{\alpha}\>:\>\alpha\in E_{I}^{+}\} is finite; let us order its elements as 0<ν1+<<νl+0<\nu_{1}^{+}<\cdots<\nu_{l}^{+}. Then if 0<ν<ν1+0<\nu<\nu_{1}^{+}, it follows from Lemma 4.5 that every αEI+\alpha\in E_{I}^{+} is ν\nu-extremal, i.e. EI,ν+=EI+E_{I,\nu}^{+}=E_{I}^{+}; similarly, if ννl+\nu\geq\nu_{l}^{+} then EI,ν+=E_{I,\nu}^{+}=\emptyset. Finally if νi+ν<νi+1+\nu_{i}^{+}\leq\nu<\nu_{i+1}^{+} for some 1i<l1\leq i<l, then for every αEI+\alpha\in E_{I}^{+} the condition ν<να\nu<\nu_{\alpha} is equivalent with να{νi+1+,,νl+}\nu_{\alpha}\in\{\nu_{i+1}^{+},\ldots,\nu_{l}^{+}\}; hence by Lemma 4.5, EI,ν+={αEI+:να{νi+1+,,νl+}}=EI,νi++E_{I,\nu}^{+}=\{\alpha\in E_{I}^{+}\>:\>\nu_{\alpha}\in\{\nu_{i+1}^{+},\ldots,\nu_{l}^{+}\}\}=E_{I,\nu_{i}^{+}}^{+}. We have thus proved the part of the lemma which concerns EI,ν+E_{I,\nu}^{+}; the proof of the part concerning EI,νE_{I,\nu}^{-} is entirely similar. ∎

Remark 4.15.

From the proof we see that the break-points in the statement of Lemma 4.6 are exactly the values να\nu_{\alpha} with αEI+EI\alpha\in E_{I}^{+}\sqcup E_{I}^{-}. Explicitly, {νi+: 1il}={να:αEI+}\{\nu_{i}^{+}\>:\>1\leq i\leq l\}=\{\nu_{\alpha}\>:\>\alpha\in E_{I}^{+}\} and {νi: 1il}={να:αEI}\{\nu_{i}^{-}\>:\>1\leq i\leq l^{\prime}\}=\{\nu_{\alpha}\>:\>\alpha\in E_{I}^{-}\}.

As a corollary we have the following.

Corollary 4.7.

There exist a finite partition of >0=j=1lISI,j\mathbb{R}_{>0}=\sqcup_{j=1}^{l_{I}}S_{I,j} into intervals, and lIl_{I} pairwise distinct subsets B1=EI+,B2,,BlI1,BlI=EIB_{1}=E_{I}^{+},B_{2},\ldots,B_{l_{I}-1},B_{l_{I}}=E_{I}^{-} of EI+EIE_{I}^{+}\sqcup E_{I}^{-}, such that EI,ν=BjE_{I,\nu}=B_{j} for any νSI,j\nu\in S_{I,j}.

(The intervals SI,jS_{I,j} are allowed to be open, closed, half-open, and degenerate; see Remark 4.2.)

Proof.

The existence of such a partition follows from Lemma 4.6 and the relation EI,ν=EI,ν+EI,νE_{I,\nu}=E_{I,\nu}^{+}\sqcup E_{I,\nu}^{-}. The pairwise disjointness of B1,,BlIEI+EIB_{1},\ldots,B_{l_{I}}\subset E_{I}^{+}\sqcup E_{I}^{-} follows again from the relation EI,ν=EI,ν+EI,νE_{I,\nu}=E_{I,\nu}^{+}\sqcup E_{I,\nu}^{-} and also the fact that EI,ν+E_{I,\nu}^{+} and EI,νE_{I,\nu}^{-} are decreasing and increasing in ν\nu respectively. ∎

Remark 4.16.

We see from the above proof that if ν{να:αEI+EI}\nu\notin\{\nu_{\alpha}\>:\>\alpha\in E_{I}^{+}\sqcup E_{I}^{-}\}, then ν\nu is an interior point of one of the intervals SI,jS_{I,j}. In particular, since 1{να:αEI+EI}1\notin\{\nu_{\alpha}\>:\>\alpha\in E_{I}^{+}\sqcup E_{I}^{-}\}, one of the intervals SI,jS_{I,j} contains 11 as an interior point.

We can now give the proof of Theorem 5.

Proof.

For each IiI_{i} (1iκ1\leq i\leq\kappa), let >0=j=1lIiSIi,j\mathbb{R}_{>0}=\sqcup_{j=1}^{l_{I_{i}}}S_{I_{i},j} be the partition as in Corollary 4.7 applied with I=IiI=I_{i}, and define AIi,ν0A_{I_{i},\nu}\geq 0 and BIi,ν0B_{I_{i},\nu}\geq 0 for ν>0\nu>0 as in Lemma 4.4. For each 1jlIi1\leq j\leq l_{I_{i}}, it follows from Corollary 4.7 that both AIi,νA_{I_{i},\nu} and BIi,νB_{I_{i},\nu} are constant as ν\nu varies over SIi,jS_{I_{i},j}; thus we may define AIi,j:=AIi,νA_{I_{i},j}:=A_{I_{i},\nu} and BIi,j:=BIi,νB_{I_{i},j}:=B_{I_{i},\nu} for any νSIi,j\nu\in S_{I_{i},j}. Then by (4.8) and Lemma 4.4 we have

𝒫,i\displaystyle\mathcal{I}_{\mathcal{P},i} =j=1lIiS~Ii,jr(θ)2(AIi,j+BIi,jν(θ)2)dyy3dθ,\displaystyle=\sum_{j=1}^{l_{I_{i}}}\int_{\tilde{S}_{I_{i},j}}r(\theta)^{-2}\left(A_{I_{i},j}+B_{I_{i},j}\nu(\theta)^{-2}\right)\,\frac{\text{d}y}{y^{3}}\text{d}\theta,

where S~Ii,j:={θ[0,2π):ν(θ)SIi,j}\tilde{S}_{I_{i},j}:=\{\theta\in[0,2\pi)\>:\>\nu(\theta)\in S_{I_{i},j}\}. Plugging this formula into (4.7) and applying the relation a𝒫=i=1κa𝒫,ia_{\mathcal{P}}=\sum_{i=1}^{\kappa}a_{\mathcal{P},i} we get the desired formula for a𝒫a_{\mathcal{P}}. ∎

4.4. Examples; proofs of Propositions 1.11.3

In this section we work out the formula (4.1) explicitly in the three cases K=(2)K=\mathbb{Q}(\sqrt{2}), (3)\mathbb{Q}(\sqrt{3}) and (5)\mathbb{Q}(\sqrt{5}), thus proving Propositions 1.11.3. Note that in all cases, KK is of class number one; thus κ=1\kappa=1 and I1=𝒪KI_{1}=\mathcal{O}_{K}, and we set I:=I1I:=I_{1}. For details of the following computations, see also the program [21, explicit_aP.mpl].

First we treat the case when K=(2)K=\mathbb{Q}(\sqrt{2}). Then λ=1+2\lambda=1+\sqrt{2}, and one verifies that

EI+={1,2+2}λ2andEI={2,1+2}λ2,\displaystyle E_{I}^{+}=\{1,2+\sqrt{2}\}\cdot\lambda^{2\mathbb{Z}}\quad\text{and}\quad E_{I}^{-}=\{\sqrt{2},1+\sqrt{2}\}\cdot\lambda^{2\mathbb{Z}},

with the associated values να\nu_{\alpha} given by

ν1=1+2,ν2+2=122andν2=2,ν1+2=21.\displaystyle\nu_{1}=1+\sqrt{2},\qquad\nu_{2+\sqrt{2}}=\tfrac{1}{2}\sqrt{2}\qquad\text{and}\qquad\nu_{\sqrt{2}}=\sqrt{2},\quad\nu_{1+\sqrt{2}}=\sqrt{2}-1.

Hence

EI,ν+={{1,2+2}λ2if ν<122,λ2if 122ν<1+2,if  1+2ν,\displaystyle E_{I,\nu}^{+}=\begin{cases}\{1,2+\sqrt{2}\}\cdot\lambda^{2\mathbb{Z}}&\text{if }\>\nu<\tfrac{1}{2}\sqrt{2},\\ \lambda^{2\mathbb{Z}}&\text{if }\>\tfrac{1}{2}\sqrt{2}\leq\nu<1+\sqrt{2},\\ \emptyset&\text{if }\>1+\sqrt{2}\leq\nu,\end{cases}

and

EI,ν={if ν21,(1+2)λ2if 21<ν2,{2,1+2}λ2if 2<ν.\displaystyle E_{I,\nu}^{-}=\begin{cases}\emptyset&\text{if }\>\nu\leq\sqrt{2}-1,\\ (1+\sqrt{2})\lambda^{2\mathbb{Z}}&\text{if }\>\sqrt{2}-1<\nu\leq\sqrt{2},\\ \{\sqrt{2},1+\sqrt{2}\}\cdot\lambda^{2\mathbb{Z}}&\text{if }\>\sqrt{2}<\nu.\end{cases}

Using also the relation EI,ν=EI,ν+EI,νE_{I,\nu}=E_{I,\nu}^{+}\sqcup E_{I,\nu}^{-}, we get the corresponding partition >0=j=15SI,j\mathbb{R}_{>0}=\sqcup_{j=1}^{5}S_{I,j} where SI,j=SjS_{I,j}=S_{j} with S1=(0,21],S2=(21,122),S3=[122,2],S4=(2,1+2)S_{1}=(0,\sqrt{2}-1],S_{2}=(\sqrt{2}-1,\frac{1}{2}\sqrt{2}),S_{3}=[\frac{1}{2}\sqrt{2},\sqrt{2}],S_{4}=(\sqrt{2},1+\sqrt{2}) and S5=[1+2,)S_{5}=[1+\sqrt{2},\infty). The corresponding constants Aj:=AI,jA_{j}:=A_{I,j} and Bj:=BI,jB_{j}:=B_{I,j} are

(A1,B1)=(B5,A5)=(72+42,0),(A2,B2)=(B4,A4)=(2+32,12)andA3=B3=1+2.\displaystyle(A_{1},B_{1})=(B_{5},A_{5})=(\tfrac{7}{2}+4\sqrt{2},0),\quad(A_{2},B_{2})=(B_{4},A_{4})=(2+3\sqrt{2},\tfrac{1}{2})\quad\text{and}\quad A_{3}=B_{3}=1+\sqrt{2}.

Plugging these values into (4.1) and using also ΔK=8\Delta_{K}=8 and ζK(2)=π4482\zeta_{K}(2)=\frac{\pi^{4}}{48\sqrt{2}} (see Remark 2.14), we get

a𝒫\displaystyle a_{\mathcal{P}} =3216π4Area(𝒲)(S~2r(θ)2((2+32)+12ν(θ)2)dθ+S~4r(θ)2(12+(2+32)ν(θ)2)dθ\displaystyle=\frac{3\sqrt{2}}{16\pi^{4}}\operatorname{Area}(\mathcal{W})\biggl{(}\int_{\tilde{S}_{2}}r(\theta)^{-2}\left((2+3\sqrt{2})+\tfrac{1}{2}\nu(\theta)^{-2}\right)\,\text{d}\theta+\int_{\tilde{S}_{4}}r(\theta)^{-2}\left(\tfrac{1}{2}+(2+3\sqrt{2})\nu(\theta)^{-2}\right)\,\text{d}\theta
(4.17) +(72+42)(S~1r(θ)2dθ+S~5r(θ)2ν(θ)2dθ)+(1+2)S~3r(θ)2(1+ν(θ)2)dθ).\displaystyle\hskip 20.0pt+(\tfrac{7}{2}+4\sqrt{2})\left(\int_{\tilde{S}_{1}}r(\theta)^{-2}\,\text{d}\theta+\int_{\tilde{S}_{5}}r(\theta)^{-2}\nu(\theta)^{-2}\,\text{d}\theta\right)+(1+\sqrt{2})\int_{\tilde{S}_{3}}r(\theta)^{-2}\left(1+\nu(\theta)^{-2}\right)\,\text{d}\theta\biggr{)}.

Moreover, because of (4.4) and ν(θ+π)=ν(θ)1\nu(\theta+\pi)=\nu(\theta)^{-1}, we have the general symmetry relation that for any bounded measurable set S>0S\subset\mathbb{R}_{>0},

(4.18) {θ:ν(θ)S}r(θ)2ν(θ)2dθ={θ:ν(θ)S1}r(θ)2dθ.\displaystyle\int_{\{\theta\>:\>\nu(\theta)\in S\}}r(\theta)^{-2}\nu(\theta)^{-2}\,\text{d}\theta=\int_{\{\theta\>:\>\nu(\theta)\in S^{-1}\}}r(\theta)^{-2}\,\text{d}\theta.

Using (4.18) and the fact that Sj1=S6jS_{j}^{-1}=S_{6-j} for 1j51\leq j\leq 5, we see that (4.17) is equivalent to the formula (1.24) in Proposition 1.1. Finally, if S~3=[0,2π)\tilde{S}_{3}=[0,2\pi) then the sum inside the parenthesis in (1.24) equals 2(1+2)02πr(θ)2dθ=4(1+2)Area(𝒲)2(1+\sqrt{2})\int_{0}^{2\pi}r(\theta)^{-2}\,\mathrm{d}\theta=4(1+\sqrt{2})\operatorname{Area}({\mathcal{W}}^{*}) by (4.5), and hence (1.25) holds. This completes the proof of Proposition 1.1.

Next, we consider the case when K=(3)K=\mathbb{Q}(\sqrt{3}). Then λ=2+3\lambda=2+\sqrt{3}, and one verifies that

EI+={1,λ}λ2=λandEI={3,1+3,3λ,(1+3)λ}λ2={3,1+3}λ\displaystyle E_{I}^{+}=\{1,\lambda\}\cdot\lambda^{2\mathbb{Z}}=\lambda^{\mathbb{Z}}\quad\text{and}\quad E_{I}^{-}=\{\sqrt{3},1+\sqrt{3},\sqrt{3}\lambda,(1+\sqrt{3})\lambda\}\cdot\lambda^{2\mathbb{Z}}=\{\sqrt{3},1+\sqrt{3}\}\lambda^{\mathbb{Z}}

with the associated values να\nu_{\alpha} given by

ν1=νλ=1+3,ν3=ν3λ=3andν1+3=ν(1+3)λ=1+3.\displaystyle\nu_{1}=\nu_{\lambda}=1+\sqrt{3},\quad\nu_{\sqrt{3}}=\nu_{\sqrt{3}\lambda}=\sqrt{3}\quad\mathrm{and}\quad\nu_{1+\sqrt{3}}=\nu_{(1+\sqrt{3})\lambda}=-1+\sqrt{3}.

Hence

EI,ν+={λif ν<1+3,if ν1+3,\displaystyle E_{I,\nu}^{+}=\begin{cases}\lambda^{\mathbb{Z}}&\text{if }\>\nu<1+\sqrt{3},\\ \emptyset&\text{if }\>\nu\geq 1+\sqrt{3},\end{cases}

and

EI,ν={if ν31,{1+3}λif 31<ν3,{3,1+3}λif 3<ν.\displaystyle E_{I,\nu}^{-}=\begin{cases}\emptyset&\text{if }\>\nu\leq\sqrt{3}-1,\\ \{1+\sqrt{3}\}\lambda^{\mathbb{Z}}&\text{if }\>\sqrt{3}-1<\nu\leq\sqrt{3},\\ \{\sqrt{3},1+\sqrt{3}\}\lambda^{\mathbb{Z}}&\text{if }\>\sqrt{3}<\nu.\end{cases}

The corresponding partition of >0\mathbb{R}_{>0} is >0=j=14SI,j\mathbb{R}_{>0}=\sqcup_{j=1}^{4}S_{I,j} where SI,j=SjS_{I,j}=S_{j} with S1=(0,31]S_{1}=(0,\sqrt{3}-1], S2=(31,3]S_{2}=(\sqrt{3}-1,\sqrt{3}], S3=(3,3+1)S_{3}=(\sqrt{3},\sqrt{3}+1) and S4=[3+1,)S_{4}=[\sqrt{3}+1,\infty). The corresponding constants Aj:=AI,jA_{j}:=A_{I,j} and Bj:=BI,jB_{j}:=B_{I,j} are

(A1,B1)=(6+43,0),(A2,B2)=(3+23,23)\displaystyle(A_{1},B_{1})=(6+4\sqrt{3},0),\qquad(A_{2},B_{2})=(3+2\sqrt{3},2\sqrt{3})

and

(A3,B3)=(2,3+83),(A4,B4)=(0,11+123).\displaystyle(A_{3},B_{3})=(2,3+8\sqrt{3}),\qquad(A_{4},B_{4})=(0,11+12\sqrt{3}).

Plugging these formulas into (4.1) and using also ΔK=12\Delta_{K}=12 and ζK(2)=π4363\zeta_{K}(2)=\frac{\pi^{4}}{36\sqrt{3}} (see Remark 2.14), we get

a𝒫=316π4Area(𝒲)(\displaystyle a_{\mathcal{P}}=\frac{\sqrt{3}}{16\pi^{4}}\operatorname{Area}(\mathcal{W})\biggl{(} (6+43)S~1r(θ)2dθ+S~2r(θ)2((3+23)+23ν(θ)2)dθ\displaystyle(6+4\sqrt{3})\int_{\tilde{S}_{1}}r(\theta)^{-2}\,\text{d}\theta+\int_{\tilde{S}_{2}}r(\theta)^{-2}\left((3+2\sqrt{3})+2\sqrt{3}\nu(\theta)^{-2}\right)\,\text{d}\theta
(4.19) +S~3r(θ)2(2+(3+83)ν(θ)2)dθ+(11+123)S~4r(θ)2ν(θ)2dθ).\displaystyle+\int_{\tilde{S}_{3}}r(\theta)^{-2}\left(2+(3+8\sqrt{3})\nu(\theta)^{-2}\right)\,\text{d}\theta+(11+12\sqrt{3})\int_{\tilde{S}_{4}}r(\theta)^{-2}\nu(\theta)^{-2}\,\text{d}\theta\biggr{)}.

Again by the symmetry identity (4.18), we see that (4.19) is equivalent to the formula (1.28) in Proposition 1.2. The last statement of Proposition 1.2 is again verified using (4.5). This completes the proof of Proposition 1.2.

Finally we consider the case when K=(5)K=\mathbb{Q}(\sqrt{5}). Then λ=1+52\lambda=\frac{1+\sqrt{5}}{2}, and one verifies that

EI+={1}λ2=λ2andEI={λ}λ2=λ2+1\displaystyle E_{I}^{+}=\{1\}\cdot\lambda^{2\mathbb{Z}}=\lambda^{2\mathbb{Z}}\quad\text{and}\quad E_{I}^{-}=\{\lambda\}\cdot\lambda^{2\mathbb{Z}}=\lambda^{2\mathbb{Z}+1}

with the associated values να\nu_{\alpha} given by

ν1=1+52andνλ=512.\displaystyle\nu_{1}=\tfrac{1+\sqrt{5}}{2}\quad\mathrm{and}\quad\nu_{\lambda}=\tfrac{\sqrt{5}-1}{2}.

Hence

EI,ν+={λ2if ν<1+52,if ν1+52,andEI,ν={if ν512,λ2+1if 512<ν.\displaystyle E_{I,\nu}^{+}=\begin{cases}\lambda^{2\mathbb{Z}}&\text{if }\>\nu<\frac{1+\sqrt{5}}{2},\\ \emptyset&\text{if }\>\nu\geq\frac{1+\sqrt{5}}{2},\end{cases}\quad\text{and}\quad E_{I,\nu}^{-}=\begin{cases}\emptyset&\text{if }\>\nu\leq\frac{\sqrt{5}-1}{2},\\ \lambda^{2\mathbb{Z}+1}&\text{if }\>\frac{\sqrt{5}-1}{2}<\nu.\end{cases}

The corresponding partition of >0\mathbb{R}_{>0} is >0=j=13SI,j\mathbb{R}_{>0}=\sqcup_{j=1}^{3}S_{I,j} where SI,j=SjS_{I,j}=S_{j} with S1=(0,512]S_{1}=(0,\frac{\sqrt{5}-1}{2}], S2=(512,1+52)S_{2}=(\frac{\sqrt{5}-1}{2},\frac{1+\sqrt{5}}{2}) and S3=[1+52,)S_{3}=[\frac{1+\sqrt{5}}{2},\infty). The corresponding constants Aj:=AI,jA_{j}:=A_{I,j} and Bj:=BI,jB_{j}:=B_{I,j} are

(A1,B1)=(B3,A3)=(5+354,0)and(A2,B2)=(1+54,1+54).\displaystyle(A_{1},B_{1})=(B_{3},A_{3})=(\tfrac{5+3\sqrt{5}}{4},0)\quad\text{and}\quad(A_{2},B_{2})=(\tfrac{1+\sqrt{5}}{4},\tfrac{1+\sqrt{5}}{4}).

Plugging these formulas into (4.1) and using also ΔK=5\Delta_{K}=5 and ζK(2)=2π4755\zeta_{K}(2)=\frac{2\pi^{4}}{75\sqrt{5}} (see Remark 2.14) and the symmetry identity (4.18), we obtain (1.31). The last statement of Proposition 1.3 is again verified using (4.5). This completes the proof of Proposition 1.3.

5. Details on the examples and numerical computations

In Sections 5.15.3 below we show how the formulas (1.23), (1.27) and (1.30) follow from the cut-and-project presentation of these vertex sets in [5, Ch. 7.3]. Then in Section 5.4 we discuss numerical computations of the corresponding gap distributions.

5.1. Ammann-Beenker tilings

Recall the definition of 𝒲AB2{\mathcal{W}}_{\operatorname{AB}}\subset\mathbb{R}^{2} in Section 1.5. Let ξ8=eπi/4\xi_{8}=e^{\pi i/4} (an 8th root of unity), and let \star be the automorphism of the cyclotomic field (ξ8)\mathbb{Q}(\xi_{8}) given by ξ8=ξ83\xi_{8}^{\star}=\xi_{8}^{3}. From now on we will use the standard identification of \mathbb{C} with 2\mathbb{R}^{2}, z(𝔢(z),𝔪(z))z\leftrightarrow({\mathfrak{Re}}(z),{\mathfrak{Im}}(z)); in particular we consider the ring [ξ8]\mathbb{Z}[\xi_{8}] as a subset of 2\mathbb{R}^{2}. Set

(5.1) 𝒯AB:={𝒘𝒲AB:[ξ8](𝒲AB+𝒘)=};\displaystyle{\mathcal{T}}_{\operatorname{AB}}:=\{{\text{\boldmath$w$}}\in{\mathcal{W}}_{\operatorname{AB}}\>:\>\mathbb{Z}[\xi_{8}]\cap\partial({\mathcal{W}}_{\operatorname{AB}}+{\text{\boldmath$w$}})=\emptyset\};

this is the set of admissible “generic” translates of 𝒲AB{\mathcal{W}}_{\operatorname{AB}}. Next, for any 𝒘2{\text{\boldmath$w$}}\in\mathbb{R}^{2}, set

(5.2) 𝒫AB,𝒘:={z[ξ8]:z𝒲AB+𝒘}(2).\displaystyle{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}}:=\{z\in\mathbb{Z}[\xi_{8}]\>:\>z^{\star}\in{\mathcal{W}}_{\operatorname{AB}}+{\text{\boldmath$w$}}\}\qquad(\subset\mathbb{R}^{2}).

Then for every 𝒘𝒯AB{\text{\boldmath$w$}}\in{\mathcal{T}}_{\operatorname{AB}}, 𝒫AB,𝒘{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}} is the vertex set of an Ammann-Beenker tiling (see [7] and [5, Ch. 7.3; Ex. 7.8]); and we note that 𝟎𝒫AB,𝒘\mathbf{0}\in{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}}. In fact every Ammann-Beenker tiling (appropriately scaled and rotated, and translated so that 𝟎\mathbf{0} is a vertex), has vertex set either equal to 𝒫AB,𝒘{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}} for some 𝒘𝒯AB{\text{\boldmath$w$}}\in{\mathcal{T}}_{\operatorname{AB}}, or equal to a limit of a sequence of such point sets, with respect to an appropriate metric on the family of locally finite point sets in 2\mathbb{R}^{2}. (In the limit case, it follows that there exists some 𝒘𝒲AB¯{\text{\boldmath$w$}}\in\overline{{\mathcal{W}}_{\operatorname{AB}}} such that the vertex set 𝒫{\mathcal{P}} agrees with 𝒫AB,𝒘{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}} “up to density zero”, viz., the symmetric difference set 𝒫𝒫AB,𝒘{\mathcal{P}}\triangle{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}} has asymptotic density zero in 2\mathbb{R}^{2}. This implies that 𝒫{\mathcal{P}} and 𝒫AB,𝒘{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}} have the same limiting distribution of normalized gaps between directions, and therefore the methods of the present paper apply also to these vertex sets, unless 𝒘𝒲AB{\text{\boldmath$w$}}\in\partial{\mathcal{W}}_{\operatorname{AB}}.)

In order to rewrite (5.2) as in (1.23), let K=(2)K=\mathbb{Q}(\sqrt{2}); then 𝒪K=[2]{\mathcal{O}}_{K}=\mathbb{Z}[\sqrt{2}]. Also set g1=(1012)g_{1}=\left(\begin{smallmatrix}1&0\\ 1&\sqrt{2}\end{smallmatrix}\right) and g2:=(101/21/2)g_{2}:=\left(\begin{smallmatrix}1&0\\ 1/\sqrt{2}\>&1/\sqrt{2}\end{smallmatrix}\right). Using ξ82=2ξ81\xi_{8}^{2}=\sqrt{2}\xi_{8}-1, one verifies that [ξ8]=𝒪K𝒪Kξ8\mathbb{Z}[\xi_{8}]={\mathcal{O}}_{K}\oplus{\mathcal{O}}_{K}\xi_{8}. Note that for any x1,x2𝒪Kx_{1},x_{2}\in{\mathcal{O}}_{K}, we have x1+x2ξ8=(x1,x2)g2x_{1}+x_{2}\xi_{8}=(x_{1},x_{2})g_{2} under our identification of \mathbb{C} with 2\mathbb{R}^{2}, and in particular, [ξ8]=𝒪K2g2\mathbb{Z}[\xi_{8}]={\mathcal{O}}_{K}^{2}g_{2} as a subset of 2\mathbb{R}^{2}. Furthermore, using (2)=(ξ8+ξ81)=ξ83+ξ83=2(\sqrt{2})^{\star}=(\xi_{8}+\xi_{8}^{-1})^{\star}=\xi_{8}^{3}+\xi_{8}^{-3}=-\sqrt{2}, we have for any (x1,x2)𝒪K2(x_{1},x_{2})\in{\mathcal{O}}_{K}^{2}:

(x1+x2ξ8)=σ(x1)+σ(x2)ξ83=(σ(x1),σ(x2))g11,\displaystyle(x_{1}+x_{2}\xi_{8})^{\star}=\sigma(x_{1})+\sigma(x_{2})\xi_{8}^{3}=(\sigma(x_{1}),\sigma(x_{2}))g_{1}^{-1},

and this point lies in 𝒲AB+𝒘{\mathcal{W}}_{\operatorname{AB}}+{\text{\boldmath$w$}} if and only if (σ(x1),σ(x2))(𝒲AB+𝒘)g1(\sigma(x_{1}),\sigma(x_{2}))\in({\mathcal{W}}_{\operatorname{AB}}+{\text{\boldmath$w$}})g_{1}. It follows that 𝒫AB,𝒘=𝒫((𝒲AB+𝒘)g1,K)g2{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}}={\mathcal{P}}(({\mathcal{W}}_{\operatorname{AB}}+{\text{\boldmath$w$}})g_{1},{\mathcal{L}}_{K})g_{2}, i.e. the formula in (1.23) holds. (In the case 𝒘=𝟎{\text{\boldmath$w$}}=\mathbf{0}, the formula (1.23) was noted in [20, Sec. 3.2].)

5.2. Gähler’s shield tilings

Recall the definition of 𝒲Gh2{\mathcal{W}}_{\operatorname{Gh}}\subset\mathbb{R}^{2} in Section 1.5. Let ξ12=eπi/6\xi_{12}=e^{\pi i/6} (a 12th root of unity), and let \star be the automorphism of the cyclotomic field (ξ12)\mathbb{Q}(\xi_{12}) given by ξ12=ξ125\xi_{12}^{\star}=\xi_{12}^{5}. As before, we identify \mathbb{C} with 2\mathbb{R}^{2} in the standard way. For any 𝒘2{\text{\boldmath$w$}}\in\mathbb{R}^{2}, set

(5.3) 𝒫Gh,𝒘:={z[ξ12]:z𝒲Gh+𝒘}(2).\displaystyle{\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}}:=\{z\in\mathbb{Z}[\xi_{12}]\>:\>z^{\star}\in{\mathcal{W}}_{\operatorname{Gh}}+{\text{\boldmath$w$}}\}\qquad(\subset\mathbb{R}^{2}).

Then for every 𝒘𝒲Gh{\text{\boldmath$w$}}\in{\mathcal{W}}_{\operatorname{Gh}} satisfying [ξ12](𝒲Gh+𝒘)=\mathbb{Z}[\xi_{12}]\cap\partial({\mathcal{W}}_{\operatorname{Gh}}+{\text{\boldmath$w$}})=\emptyset, 𝒫Gh,𝒘{\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}} is the vertex set of a Gähler’s shield tiling [5, Ch. 7.3; Ex. 7.12].

In order to rewrite (5.3) as in (1.27), let K=(3)K=\mathbb{Q}(\sqrt{3}); then 𝒪K=[3]{\mathcal{O}}_{K}=\mathbb{Z}[\sqrt{3}]. Also set g1:=(1032)g_{1}:=\left(\begin{smallmatrix}1&0\\ \sqrt{3}&2\end{smallmatrix}\right) and g2:=(103/21/2)g_{2}:=\left(\begin{smallmatrix}1&0\\ \sqrt{3}/2&1/2\end{smallmatrix}\right). Using ξ122=3ξ121\xi_{12}^{2}=\sqrt{3}\xi_{12}-1 one verifies that [ξ12]=𝒪K𝒪Kξ12\mathbb{Z}[\xi_{12}]={\mathcal{O}}_{K}\oplus{\mathcal{O}}_{K}\xi_{12}. Note that for any x1,x2𝒪Kx_{1},x_{2}\in{\mathcal{O}}_{K}, we have x1+x2ξ12=(x1,x2)g2x_{1}+x_{2}\xi_{12}=(x_{1},x_{2})g_{2} under our identification of \mathbb{C} with 2\mathbb{R}^{2}, and in particular, [ξ12]=𝒪K2g2\mathbb{Z}[\xi_{12}]={\mathcal{O}}_{K}^{2}g_{2} as a subset of 2\mathbb{R}^{2}. Furthermore, using (3)=(ξ12+ξ121)=ξ125+ξ125=3(\sqrt{3})^{\star}=(\xi_{12}+\xi_{12}^{-1})^{\star}=\xi_{12}^{5}+\xi_{12}^{-5}=-\sqrt{3}, we have for any (x1,x2)𝒪K2(x_{1},x_{2})\in{\mathcal{O}}_{K}^{2}:

(x1+x2ξ12)=σ(x1)+σ(x2)ξ125,\displaystyle(x_{1}+x_{2}\xi_{12})^{\star}=\sigma(x_{1})+\sigma(x_{2})\xi_{12}^{5},

and this point lies in 𝒲Gh+𝒘{\mathcal{W}}_{\operatorname{Gh}}+{\text{\boldmath$w$}} if and only if (σ(x1),σ(x2))(𝒲Gh+𝒘)g1(\sigma(x_{1}),\sigma(x_{2}))\in({\mathcal{W}}_{\operatorname{Gh}}+{\text{\boldmath$w$}})g_{1}. It follows that 𝒫Gh,𝒘=𝒫((𝒲Gh+𝒘)g1,K)g2{\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}}={\mathcal{P}}(({\mathcal{W}}_{\operatorname{Gh}}+{\text{\boldmath$w$}})g_{1},{\mathcal{L}}_{K})g_{2}, i.e. the formula in (1.27) holds.

5.3. Tübingen triangle tilings

Recall the definition of 𝒲TT2{\mathcal{W}}_{\operatorname{\hskip 1.0ptTT}}\subset\mathbb{R}^{2} in Section 1.5. Let ξ5=e2πi/5\xi_{5}=e^{2\pi i/5} (a 5th root of unity), and let \star be the automorphism of the cyclotomic field (ξ5)\mathbb{Q}(\xi_{5}) given by ξ5=ξ52\xi_{5}^{\star}=\xi_{5}^{2}. As in Section 5.1, we identify \mathbb{C} with 2\mathbb{R}^{2} in the standard way. For any 𝒘2{\text{\boldmath$w$}}\in\mathbb{R}^{2}, set

(5.4) 𝒫TT,𝒘:={z[ξ5]:z𝒲TT+𝒘}(2).\displaystyle{\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},{\text{\boldmath$w$}}}:=\{z\in\mathbb{Z}[\xi_{5}]\>:\>z^{\star}\in{\mathcal{W}}_{\operatorname{\hskip 1.0ptTT}}+{\text{\boldmath$w$}}\}\qquad(\subset\mathbb{R}^{2}).

Then for every 𝒘𝒲TT{\text{\boldmath$w$}}\in{\mathcal{W}}_{\operatorname{\hskip 1.0ptTT}} satisfying 15[ξ5](𝒲TT+𝒘)=\frac{1}{5}\mathbb{Z}[\xi_{5}]\cap\partial({\mathcal{W}}_{\operatorname{\hskip 1.0ptTT}}+{\text{\boldmath$w$}})=\emptyset, 𝒫TT,𝒘{\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},{\text{\boldmath$w$}}} is the vertex set of a Tübingen triangle tiling [5, Ch. 7.3; Ex. 7.10], [6, Sec. 4].

In order to rewrite (5.4) as in (1.30), let K=(5)K=\mathbb{Q}(\sqrt{5}); then 𝒪K=[τ]{\mathcal{O}}_{K}=\mathbb{Z}[\tau], where we recall that τ=12(1+5)\tau=\frac{1}{2}(1+\sqrt{5}). Also set g1:=(100(2+τ)/5)(10τ2)g_{1}:=\left(\begin{smallmatrix}1&0\\ 0&\sqrt{(2+\tau)/5}\end{smallmatrix}\right)\left(\begin{smallmatrix}1&0\\ \tau&2\end{smallmatrix}\right) and g2:=(10(τ1)/22+τ/2)g_{2}:=\left(\begin{smallmatrix}1&0\\ (\tau-1)/2\>&\sqrt{2+\tau}/2\end{smallmatrix}\right). Using ξ52=(τ1)ξ51\xi_{5}^{2}=(\tau-1)\xi_{5}-1 one verifies that [ξ5]=𝒪K𝒪Kξ5\mathbb{Z}[\xi_{5}]={\mathcal{O}}_{K}\oplus{\mathcal{O}}_{K}\xi_{5}. Note that for any x1,x2𝒪Kx_{1},x_{2}\in{\mathcal{O}}_{K}, we have x1+x2ξ5=(x1,x2)g2x_{1}+x_{2}\xi_{5}=(x_{1},x_{2})g_{2} under our identification of \mathbb{C} with 2\mathbb{R}^{2}, and in particular, [ξ5]=𝒪K2g2\mathbb{Z}[\xi_{5}]={\mathcal{O}}_{K}^{2}g_{2} as a subset of 2\mathbb{R}^{2}. Furthermore, using (5)=(1+2ξ5+2ξ54)=1+2ξ52+2ξ53=(1+2ξ5+2ξ54)=5(\sqrt{5})^{\star}=(1+2\xi_{5}+2\xi_{5}^{4})^{\star}=1+2\xi_{5}^{2}+2\xi_{5}^{3}=-(1+2\xi_{5}+2\xi_{5}^{4})=-\sqrt{5}, we have for any (x1,x2)𝒪K2(x_{1},x_{2})\in{\mathcal{O}}_{K}^{2}:

(x1+x2ξ5)=σ(x1)+σ(x2)ξ52,\displaystyle(x_{1}+x_{2}\xi_{5})^{\star}=\sigma(x_{1})+\sigma(x_{2})\xi_{5}^{2},

and this point lies in 𝒲TT+𝒘{\mathcal{W}}_{\operatorname{\hskip 1.0ptTT}}+{\text{\boldmath$w$}} if and only if (σ(x1),σ(x2))(𝒲TT+𝒘)g1(\sigma(x_{1}),\sigma(x_{2}))\in({\mathcal{W}}_{\operatorname{\hskip 1.0ptTT}}+{\text{\boldmath$w$}})g_{1}. It follows that 𝒫TT,𝒘=𝒫((𝒲TT+𝒘)g1,K)g2{\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},{\text{\boldmath$w$}}}={\mathcal{P}}(({\mathcal{W}}_{\operatorname{\hskip 1.0ptTT}}+{\text{\boldmath$w$}})g_{1},{\mathcal{L}}_{K})g_{2}, i.e. the formula in (1.30) holds.

5.4. Numerical computations

Next we discuss numerical computations of the gaps between directions of points in the Ammann-Beenker, Gähler’s shield, and Tübingen triangle tilings, 𝒫AB,𝒘{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}}, 𝒫Gh,𝒘{\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}} and 𝒫TT,𝒘{\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},{\text{\boldmath$w$}}}. In Table 1 below, for a few examples of points 𝒘𝒲AB{\text{\boldmath$w$}}\in{\mathcal{W}}_{\operatorname{AB}}, we present conjectural approximate values of s2F(s)s^{2}F(s), where F(s)F(s) is the limiting gap distribution function associated to 𝒫AB,𝒘{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}}, and, for comparison, the exact values of a𝒫a_{{\mathcal{P}}}. Tables 2 and 3 present similar types of data for 𝒫Gh,𝒘{\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}} and 𝒫TT,𝒘{\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},{\text{\boldmath$w$}}}.

Experimental approximate values of s2F(s)s^{2}F(s) for 𝒫=𝒫AB,w{\mathcal{P}}={\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}}

𝒘w a𝒫a_{{\mathcal{P}}} s=10s=10 s=20s=20 s=50s=50 s=100s=100 s=200s=200 s=500s=500 s=1000s=1000
(0,0)(0,0) 0.24638… 0.2810 0.2628 0.2525 0.2497 0.250 0.246 0.23
(0.8,0.1)(0.8,0.1) 0.21809… 0.2460 0.2317 0.2236 0.2212 0.220 0.219 0.21
(0.9,0.3)(0.9,0.3) 0.20416… 0.2286 0.2160 0.2090 0.2067 0.206 0.205 0.19
(1.2,0.5)(1.2,0.5) 0.17444… 0.2211 0.2048 0.1815 0.1767 0.177 0.178 0.18
Error \lesssim 0.0005 0.0006 0.0011 0.0023 0.005 0.015 0.05
Table 1. Values of a𝒫a_{{\mathcal{P}}} and approximate values of s2F(s)s^{2}F(s), for 𝒫=𝒫AB,𝒘{\mathcal{P}}={\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}}. The last line gives a conjectural approximate upper bound on the absolute error of each value in the corresponding column. (Thus, for example, we expect that for 𝒘=𝟎{\text{\boldmath$w$}}=\mathbf{0} and s=50s=50, |s2F(s)0.2525|0.0011|s^{2}F(s)-0.2525|\lesssim 0.0011.)

Experimental approximate values of s2F(s)s^{2}F(s) for 𝒫=𝒫Gh,w{\mathcal{P}}={\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}}

𝒘w a𝒫a_{{\mathcal{P}}} s=10s=10 s=20s=20 s=50s=50 s=100s=100 s=200s=200 s=500s=500 s=1000s=1000
(1.7,0.6)(1.7,0.6) 0.17790… 0.1974 0.1873 0.1817 0.180 0.180 0.176 0.19
(0.1,0)(0.1,0) 0.21374… 0.2460 0.2289 0.2199 0.216 0.217 0.215 0.23
(1.5,0.3)(1.5,-0.3) 0.19210… 0.2143 0.2027 0.1964 0.194 0.195 0.190 0.20
(0.3,1.2)(0.3,1.2) 0.20870… 0.2341 0.2208 0.2136 0.211 0.211 0.205 0.23
Error \lesssim 0.0004 0.0007 0.001 0.004 0.006 0.019 0.05
Table 2. Values of a𝒫a_{{\mathcal{P}}} and approximate values of s2F(s)s^{2}F(s), for 𝒫=𝒫Gh,𝒘{\mathcal{P}}={\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}}.

Experimental approximate values of s2F(s)s^{2}F(s) for 𝒫=𝒫TT,w{\mathcal{P}}={\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},{\text{\boldmath$w$}}}

𝒘w a𝒫a_{{\mathcal{P}}} s=10s=10 s=20s=20 s=50s=50 s=100s=100 s=200s=200 s=500s=500 s=1000s=1000
(0.5,0.4)(0.5,0.4) 0.26135… 0.2919 0.2762 0.2673 0.2643 0.263 0.265 0.26
(0.4,0)(0.4,0) 0.28875… 0.3283 0.3080 0.2965 0.2927 0.291 0.294 0.30
(0.15,0.3)(0.15,0.3) 0.29103… 0.3329 0.3116 0.2995 0.2954 0.294 0.297 0.30
(1.3,0.4)(1.3,0.4) 0.18732… 0.2195 0.1977 0.1907 0.1892 0.188 0.190 0.19
Error \lesssim 0.0004 0.0004 0.0007 0.0017 0.004 0.014 0.03
Table 3. Values of a𝒫a_{{\mathcal{P}}} and approximate values of s2F(s)s^{2}F(s), for 𝒫=𝒫TT,𝒘{\mathcal{P}}={\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},{\text{\boldmath$w$}}}.

The approximations of F(s)F(s) in these tables were obtained by collecting all the points of 𝒫{\mathcal{P}} (where 𝒫=𝒫AB,𝒘{\mathcal{P}}={\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}} or 𝒫Gh,𝒘{\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}} or 𝒫TT,𝒘{\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},{\text{\boldmath$w$}}}) in fairly large discs R2{\mathcal{B}}_{R}^{2} and evaluating the ratio in the left hand side of Theorem 1; by repeating this for several large RR-values we also obtained conjectural error bounds; see below for a more detailed discussion. Recall that limss2F(s)=a𝒫\lim_{s\to\infty}s^{2}F(s)=a_{{\mathcal{P}}} by Theorem 2, in fact with a rate O(s12+ε)O(s^{-\frac{1}{2}+\varepsilon}). We certainly expect that the absolute difference |s2F(s)a𝒫||s^{2}F(s)-a_{{\mathcal{P}}}| typically decreases as ss runs through the values 10,20,50,100,200,500,100010,20,50,100,200,500,1000, and it should be noted that the data in Tables 13 is consistent with this hypothesis, when considering also the conjectural error bounds displayed in the last line of the tables.

Precise numerical values of a𝒫a_{{\mathcal{P}}} such as those presented in Tables 13 (but also to much higher precision), are quite easy to compute using Propositions 1.11.3. To describe how to do this, first note that in the case 𝒫=𝒫AB,𝒘{\mathcal{P}}={\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}}, it follows from (1.23), the last statement in Theorem 1 and Remark 3.12, that this a𝒫a_{{\mathcal{P}}} is the same as for 𝒫=𝒫(𝒲AB+𝒘,K){\mathcal{P}}={\mathcal{P}}({\mathcal{W}}_{\operatorname{AB}}+{\text{\boldmath$w$}},{\mathcal{L}}_{K}); similarly, for 𝒫=𝒫Gh,𝒘{\mathcal{P}}={\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}}, a𝒫a_{{\mathcal{P}}} is the same as for 𝒫=𝒫(𝒲Gh+𝒘,K){\mathcal{P}}={\mathcal{P}}({\mathcal{W}}_{\operatorname{Gh}}+{\text{\boldmath$w$}},{\mathcal{L}}_{K}) (see (1.27)), and for 𝒫=𝒫TT,𝒘{\mathcal{P}}={\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},{\text{\boldmath$w$}}}, a𝒫a_{{\mathcal{P}}} is the same as for 𝒫=𝒫(𝒲TT+𝒘,K){\mathcal{P}}={\mathcal{P}}({\mathcal{W}}_{\operatorname{\hskip 1.0ptTT}}+{\text{\boldmath$w$}},{\mathcal{L}}_{K}) (see (1.30)). Now, somewhat more generally, assume that 𝒲=𝒲0+𝒘{\mathcal{W}}={\mathcal{W}}_{0}+{\text{\boldmath$w$}} where 𝒲0{\mathcal{W}}_{0} is the open regular NN-gon with vertices at the points R0e2πi(12N+kN)R_{0}e^{2\pi i(\frac{1}{2N}+\frac{k}{N})} (k/Nk\in\mathbb{Z}/N\mathbb{Z}), for some integer N3N\geq 3 and some R0>0R_{0}>0, and 𝒘𝒲0{\text{\boldmath$w$}}\in{\mathcal{W}}_{0}. (We obtain 𝒲0=𝒲AB{\mathcal{W}}_{0}={\mathcal{W}}_{\operatorname{AB}} by taking N=8N=8 and R0=1+1/2R_{0}=\sqrt{1+\sqrt{1/2}},  𝒲0=𝒲Gh{\mathcal{W}}_{0}={\mathcal{W}}_{\operatorname{Gh}} by taking N=12N=12 and R0=2+3R_{0}=\sqrt{2+\sqrt{3}}, and 𝒲0=𝒲TT{\mathcal{W}}_{0}={\mathcal{W}}_{\operatorname{\hskip 1.0ptTT}} by taking N=10N=10 and R0=220(5+5)3/2R_{0}=\frac{\sqrt{2}}{20}(5+\sqrt{5})^{3/2}.) For such a window 𝒲{\mathcal{W}}, the function r(θ)r(\theta) in (1.16) is given by

r(θ)=𝔢(rjei(θj+θ))=rjcos(θj+θ),θ[(j1)2πN,j2πN]+2π,\displaystyle r(\theta)={\mathfrak{Re}}\bigl{(}r_{j}e^{i(\theta_{j}+\theta)}\bigr{)}=r_{j}\cos(\theta_{j}+\theta),\qquad\forall\theta\in\bigl{[}(j-1)\tfrac{2\pi}{N},j\tfrac{2\pi}{N}\bigr{]}+2\pi\mathbb{Z},

where for each j/Nj\in\mathbb{Z}/N\mathbb{Z}, rj>0r_{j}>0 and θj/2π\theta_{j}\in\mathbb{R}/2\pi\mathbb{Z} are defined by 𝒘+R0e2πi(12NjN)=rjeiθj{\text{\boldmath$w$}}+R_{0}e^{2\pi i(\frac{1}{2N}-\frac{j}{N})}=r_{j}e^{i\theta_{j}}. Using also ν(θ)=r(θ+π)/r(θ)\nu(\theta)=r(\theta+\pi)/r(\theta), it is then easy to verify that all the sets S~j\tilde{S}_{j} in Propositions 1.11.3 are finite unions of intervals. For given 𝒲0{\mathcal{W}}_{0} and 𝒘w, these intervals can be computed numerically, and with this, also the integrals in (1.24) and (1.28) can be computed. This is carried out in the program [21, explicit_aP.mpl].

We next describe more in detail how the experimental approximate values of F(s)F(s) in Tables 13 were computed. (This is similar as in [4] and [20], but we have carried out more extensive computations.) With notation as in Theorem 1, set

(5.5) FR(s):=#{1jN(R):N(R)(ξR,jξR,j1)s}N(R).\displaystyle F_{R}(s):=\frac{\#\{1\leq j\leq N(R)\>:\>N(R)(\xi_{R,j}-\xi_{R,j-1})\geq s\}}{N(R)}.

Then F(s)=limRFR(s)F(s)=\lim_{R\to\infty}F_{R}(s), and so for RR sufficiently large, FR(s)F_{R}(s) is a good approximation of F(s)F(s). Note that computing FR(s)F_{R}(s) involves collecting all the points of 𝒫{\mathcal{P}} in the disc R2{\mathcal{B}}_{R}^{2}. When carrying out the computations, we noted that for fixed ss, the values FR(s)F_{R}(s) typically oscillate as a function of RR; hence if \mathfrak{R} is a finite set of several sufficiently large RR-values of the same order of magnitude, it seems reasonable to consider the difference

ΔF(s):=maxRFR(s)minRFR(s)\displaystyle\Delta_{\mathfrak{R}}F(s):=\max_{R\in\mathfrak{R}}F_{R}(s)-\min_{R\in\mathfrak{R}}F_{R}(s)

as an indication of the size of the error |FR(s)F(s)||F_{R}(s)-F(s)| for any RR\in\mathfrak{R}. To lessen the probability of getting a small difference by coincidence, we took the experimental error bound given in the last line of Tables 13 to be the maximum of s~2ΔF(s~)\tilde{s}^{2}\Delta_{\mathfrak{R}}F(\tilde{s}) over 100100 values of s~\tilde{s} roughly equispaced in the interval [45s,65s][\frac{4}{5}s,\frac{6}{5}s], and over all four choices 𝒘w. Because of the factor s~2\tilde{s}^{2}, it is not surprising that these experimental error bounds get worse as ss increases.

In order to produce Table 1 (i.e. the case of Ammann-Beenker tilings) we used

(5.6) ={22000,23000,24000,25000},\displaystyle\mathfrak{R}=\{22000,23000,24000,25000\},

and for each ss we chose as approximate value of F(s)F(s) the mid-point 12(minRFR(s)+maxRFR(s))\frac{1}{2}\bigl{(}\min_{R\in\mathfrak{R}}F_{R}(s)+\max_{R\in\mathfrak{R}}F_{R}(s)\bigr{)}. We remark that computing the functions FRF_{R} in this case involve collecting a bit over 2.31092.3\cdot 10^{9} points from each point set 𝒫AB,𝒘{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}}. For example, for R=25000R=25000 we found the number of points in 𝒫AB,𝟎R2{𝟎}{\mathcal{P}}_{\operatorname{AB},\mathbf{0}}\cap{\mathcal{B}}_{R}^{2}\setminus\{\mathbf{0}\} to be 23701485922370148592 (this may be compared with c𝒫πR2=1+22πR2=2370148622.42c_{{\mathcal{P}}}\pi R^{2}=\frac{1+\sqrt{2}}{2}\pi R^{2}=2370148622.42\ldots).

Similarly, for Table 2 (Gähler’s shield tilings) we used

(5.7) ={12000,13000,14000,15000},\displaystyle\mathfrak{R}=\{12000,13000,14000,15000\},

and computing the functions FRF_{R} involved collecting a bit over 2.61092.6\cdot 10^{9} points from each point set 𝒫Gh,𝒘{\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}}. For example, for R=15000R=15000 and 𝒘=(1.7,0.6){\text{\boldmath$w$}}=(1.7,0.6) we found the number of points in 𝒫Gh,𝒘R2{𝟎}{\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}}\cap{\mathcal{B}}_{R}^{2}\setminus\{\mathbf{0}\} to be 26380314852638031485 (which may be compared with c𝒫πR2=(2+3)πR2=2638031264.97c_{{\mathcal{P}}}\pi R^{2}=(2+\sqrt{3})\pi R^{2}=2638031264.97\ldots).

Similarly, for Table 3 (Tübingen triangle tilings) we used

(5.8) ={17000,18000,19000,20000},\displaystyle\mathfrak{R}=\{17000,18000,19000,20000\},

and computing the functions FRF_{R} involved collecting a bit over 2.51092.5\cdot 10^{9} points from each point set 𝒫Gh,𝒘{\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}}. For example, for R=20000R=20000 and 𝒘=(0.4,0){\text{\boldmath$w$}}=(0.4,0) we found the number of points in 𝒫TT,𝒘R2{𝟎}{\mathcal{P}}_{\operatorname{\hskip 1.0ptTT},{\text{\boldmath$w$}}}\cap{\mathcal{B}}_{R}^{2}\setminus\{\mathbf{0}\} to be 25031184102503118410 (which may be compared with c𝒫πR2=25(τ+1)τ+2πR2=2503118771.19c_{{\mathcal{P}}}\pi R^{2}=\frac{2}{5}(\tau+1)\sqrt{\tau+2}\cdot\pi R^{2}=2503118771.19\ldots).

Let us next discuss the graphs in Figure 1. These were generated by distributing the normalized gaps N(R)(ξR,jξR,j1)N(R)(\xi_{R,j}-\xi_{R,j-1}), j=1,,N(R)j=1,\ldots,N(R) – discarding the vanishing gaps – in bins of width 0.020.02, and then drawing the corresponding histogram, appropriately scaled. In other words, using the notation in (5.5), each panel displays the graph of the function

(5.9) sER,h(s):=F~R(sh)FR(sh+h)h,with sh:=hs/h and h:=0.02,\displaystyle s\mapsto E_{R,h}(s):=\frac{\widetilde{F}_{R}(s_{h})-F_{R}(s_{h}+h)}{h},\qquad\text{with }\>s_{h}:=h\lfloor s/h\rfloor\text{ and }h:=0.02,

where F~R(0):=FR(0+):=lims0+FR(s)\widetilde{F}_{R}(0):=F_{R}(0+):=\lim_{s\to 0+}F_{R}(s) while F~R(s):=FR(s)\widetilde{F}_{R}(s):=F_{R}(s) for all s>0s>0.333The fact that we use F~R\widetilde{F}_{R} in place of FRF_{R} corresponds exactly to the fact that we discard those gaps N(R)(ξR,jξR,j1)N(R)(\xi_{R,j}-\xi_{R,j-1}) which vanish; without this modification the graph would have a huge peak for s[0,h)s\in[0,h); see the discussion below. We point out that the graphs of ER,hE_{R,h}, as they are scaled in Figure 1, look identical to the naked eye for each RR\in\mathfrak{R}, i.e. the picture looks the same if we plot all four graphs in the same coordinate system; in fact the maximum over s[0,6]s\in[0,6] of the difference maxRER,h(s)minRER,h(s)\max_{R\in\mathfrak{R}}E_{R,h}(s)-\min_{R\in\mathfrak{R}}E_{R,h}(s) was in each of the six cases found to be less than 0.00180.0018, and the average of that same difference was less than 0.00040.0004.

These findings support the conjecture that, in the cases considered, the derivative F(s)F^{\prime}(s) exists, is continuous, and is essentially correctly depicted by the graphs in Figure 1.

It is also of interest to consider the areas under the curves in Figure 1. Note that it follows from (5.9) that the area under the graph of ER,h(s)E_{R,h}(s) is

0ER,h(s)𝑑s=FR(0+).\displaystyle\int_{0}^{\infty}E_{R,h}(s)\,ds=F_{R}(0+).

Furthermore, as RR\to\infty, FR(0+)F_{R}(0+) tends to κ𝒫\kappa_{{\mathcal{P}}}, the relative density of visible points in 𝒫{\mathcal{P}}, a quantity which was defined and proved to exist for general regular cut-and-project sets in [27]. (The proof of limRFR(0+)=κ𝒫\lim_{R\to\infty}F_{R}(0+)=\kappa_{{\mathcal{P}}} is immediate by comparing with the definition of κ𝒫\kappa_{{\mathcal{P}}} in [27].) The values which we obtained for FR(0+)F_{R}(0+) for varying RR in RR\in\mathfrak{R} consistently agreed to within ca 31063\cdot 10^{-6}; hence we expect that these values approximate κ𝒫\kappa_{{\mathcal{P}}} to within an absolute error of similar order of magnitude. In particular for 𝒫AB,𝒘{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}} with 𝒘=𝟎{\text{\boldmath$w$}}=\mathbf{0} and R=25000R=25000, we obtained FR(0+)=13683158722370148592=0.57731227{\displaystyle F_{R}(0+)=\frac{1368315872}{2370148592}=0.57731227\ldots}.444Similar numerics, but for considerably smaller RR-values, also appears in [4, Table 2] and [20, Table 1]. Thus, this value is the area under the curve in panel I of Figure 1, and it may be compared with the known exact value [20, p. 10]

(5.10) κ𝒫=2|σ(λ)|ζK(2)=2(21)π4/(482)=0.57731262.\displaystyle\kappa_{{\mathcal{P}}}=\frac{2|\sigma(\lambda)|}{\zeta_{K}(2)}=\frac{2(\sqrt{2}-1)}{\pi^{4}/(48\sqrt{2})}=0.57731262\ldots.

For the other panels in Figure 1 we obtained the following values of FR(0+)F_{R}(0+) (viz., area under the curve): II. 0.710230.71023\ldots. III. 0.764470.76447\ldots. IV. 0.628630.62863\ldots. V. 0.601730.60173\ldots. VI. 0.735150.73515\ldots.

Finally, to avoid possible confusion, let us remind about the fact that, as explained in Remark 1.9, in the present paper we are using a different normalization of “F(s)F(s)” than what was used in [4], [27] and [20]. This has the effect that the graph in the top left panel of Figure 1 is quantitatively strongly different from the graphs in [4, Fig. 9] and [20, Fig. 2], despite the fact that each of these three graphs depict the gap distribution for directions in the point set 𝒫=𝒫AB,𝟎{\mathcal{P}}={\mathcal{P}}_{\operatorname{AB},\mathbf{0}}. In fact, in our notation, the plots in [4, Fig. 9] and [20, Fig. 2] are experimental graphs of the function sκ𝒫2F(κ𝒫1s)s\mapsto-\kappa_{{\mathcal{P}}}^{-2}F^{\prime}(\kappa_{{\mathcal{P}}}^{-1}s), with κ𝒫\kappa_{{\mathcal{P}}} as in (5.10). Taking this rescaling into account, the three graphs agree fairly well — one naturally expects that the more erratic small-scale behavior of the graphs in [4] and [20] is due to the fact that those were calculated using considerably smaller RR-values than those used for our Figure 1.

Index of notation

a𝒫a_{\mathcal{P}} leading coefficient of the limiting gap distribution function F(s)F(s) of a planar point set 𝒫\mathcal{P} 2
a𝒚\text{a}_{\bm{y}} matrix element (ay1,ay2)(\text{a}_{y_{1}},\text{a}_{y_{2}}) with ay=(y00y1)\text{a}_{y}=\left(\begin{smallmatrix}y&0\\ 0&y^{-1}\end{smallmatrix}\right) 2.3
𝒜\mathcal{A} closure of πint()\pi_{\operatorname{int}}(\mathcal{L}) with \mathcal{L} a lattice in d+m\mathbb{R}^{d+m} and πint\pi_{\operatorname{int}} the natural projection to the internal space 1.2
𝒜\mathcal{A}^{\circ} identity component of 𝒜\mathcal{A} 1.2
𝔄𝜽,𝒚(s)\mathfrak{A}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}}}(s) the set (y11s1/2λ2r(θ1))×(y21λ2r𝒲(θ2))(y_{1}^{-1}s^{1/2}\lambda^{-2r}\ell(\theta_{1}))\times(y_{2}^{-1}\lambda^{2r}\ell_{{\mathcal{W}}}(\theta_{2})) 3.55
αI,J(y)\alpha_{I,J}(y) the function min{αI>0:yσ(α)J}\min\left\{\alpha\in I\cap\mathbb{R}_{>0}\>:\>y\,\sigma(\alpha)\in J\right\} with I𝒪KI\subset\mathcal{O}_{K} an integral ideal
and JJ\subset\mathbb{R} bounded measurable with non-empty interior 3.3
α~I,J(y)\tilde{\alpha}_{I,J}(y) the function max{αI<0:yσ(α)J}\max\bigl{\{}\alpha\in I\cap\mathbb{R}_{<0}\>:\>y\,\sigma(\alpha)\in J\bigr{\}} with I,JI,J as above 3.53
bα,νb_{\alpha,\nu} bα,ν:=σ(α)b_{\alpha,\nu}:=\sigma(\alpha) if σ(α)>0\sigma(\alpha)>0 and bα,ν:=ν1σ(α)b_{\alpha,\nu}:=\nu^{-1}\sigma(\alpha) if σ(α)<0\sigma(\alpha)<0 4.3
R2{\mathcal{B}}_{R}^{2} open disc in 2\mathbb{R}^{2} with center 𝟎\bm{0} and radius RR 3.2
𝔅𝜽,𝒚,𝒙,α(s)\mathfrak{B}_{{\text{\boldmath$\theta$}},{\text{\boldmath$y$}},{\text{\boldmath$x$}},\alpha}(s) the set (y1𝔍1αx1)×(y2𝔍2σ(α)x2)\left(y_{1}\mathfrak{J}_{1}-\alpha x_{1}\right)\times\left(y_{2}\mathfrak{J}_{2}-\sigma(\alpha)x_{2}\right) 3.57
𝔠i\mathfrak{c}_{i} the vector (ci,ai)𝒪K2(-c_{i},a_{i})\in\mathcal{O}_{K}^{2} such that aici=ki\frac{a_{i}}{c_{i}}=k_{i} is the ii-cusp of ΓK\Gamma_{K} 2.5
CKC_{K} ideal class group of KK 2.1
cKc_{K} normalizing factor of a Haar measure of H\mathrm{H} which equals ΔK3/2ζK(2)1\Delta_{K}^{-3/2}\zeta_{K}(2)^{-1} 2.13
c𝒫c_{\mathcal{P}} asymptotic density of a planar point set 𝒫\mathcal{P} 1.2
ΔK\Delta_{K} discriminant of KK 1.11, 2.4
ΔR\Delta_{R} directions of points of length bounded by RR in a planar point set 𝒫\mathcal{P} 1.1
EI±E_{I}^{\pm} set of positively or negatively extremal points with respect to an integral ideal II 4.3
EI,νE_{I,\nu} set of ν\nu-extremal points with respect to an integral ideal II 4.3
EI,ν±E^{\pm}_{I,\nu} set of positive or negative ν\nu-extremal points with respect to II 4.3
F(s)F(s) limiting distribution function of gaps of directions of a planar cut-and-project set 1.6
𝔣(A,a)\mathfrak{f}(A,a) the function a1inf{m((x,x+a)A):x}a^{-1}\inf\bigl{\{}m\bigl{(}(x,x+a)\setminus A\bigr{)}\>:\>x\in\mathbb{R}\bigr{\}} with a>0a>0 and AA\subset\mathbb{R} Lebesgue measurable 3.5
𝔣~(A,a)\widetilde{\mathfrak{f}}(A,a) either 0 or 𝔣(A,a)\mathfrak{f}(A,a) depending on whether AA is empty or not 3.6
ΓK\mathcal{F}_{\Gamma_{K}} a Siegel fundamental domain for ΓK\H\Gamma_{K}\backslash\mathrm{H} 2.19
i\mathcal{F}_{i} a fixed fundamental domain for Γi\H\Gamma_{i}\backslash H 2.16
i(t)\mathcal{F}_{i}(t) the set {ξin𝒙a𝒚k𝜽i:y1y2t}\left\{\xi_{i}\text{n}_{\bm{x}}\text{a}_{\bm{y}}\text{k}_{\bm{\theta}}\in\mathcal{F}_{i}\>:\>y_{1}y_{2}\geq t\right\} 2.17
𝔉i\mathfrak{F}_{i} a fixed fundamental domain for 2/ι(Ii2)\mathbb{R}^{2}/\iota(I_{i}^{-2}) 2.4
G(s)G(s) integral of the limiting gap distribution function F(s)F(s) 1.12
Gi(s)G_{i}(s) contribution of the ii-th cusp of ΓK\Gamma_{K} to the function G(s)G(s) for sufficiently large ss 3.36
GE,i(s)G_{E,i}(s) error term of Gi(s)G_{i}(s) 3.41
GM,i(s)G_{M,i}(s) main term of Gi(s)G_{i}(s) 3.40
Γ\Gamma the group SL4()\operatorname{SL}_{4}(\mathbb{Z}) 3.3
Γi\Gamma_{i} isotropy group of the ii-th cusp of ΓK\Gamma_{K} 2.7
ΓK\Gamma_{K} the group ι(SL2(𝒪K))\iota(\operatorname{SL}_{2}(\mathcal{O}_{K})) 2.2
H\mathrm{H} the group SL2()×SL2()\operatorname{SL}_{2}(\mathbb{R})\times\operatorname{SL}_{2}(\mathbb{R}) 2.2
Hg\mathrm{H}_{g} the group gHg1g\mathrm{H}g^{-1} with gSL4()g\in\operatorname{SL}_{4}(\mathbb{R}) 3.3
IiI_{i} ideal generated by (ai,ci)𝒪K2(a_{i},c_{i})\in\mathcal{O}_{K}^{2} with ai,cia_{i},c_{i} such that ki=aicik_{i}=\frac{a_{i}}{c_{i}} 2.1
ι\iota the natural embedding either from KK to 2\mathbb{R}^{2} or from SL2(K)\operatorname{SL}_{2}(K) to H\mathrm{H} 2.1, 2.6
JJ the set (0,π4)(3π4,π)\bigl{(}0,\tfrac{\pi}{4}\bigr{)}\cup\bigl{(}\tfrac{3\pi}{4},\pi\bigr{)} 3.5
JνJ_{\nu} open interval (ν,1)(-\nu,1) with ν>0\nu>0 4.2
𝔍1\mathfrak{J}_{1} the set {t:(y1α,t)s1/2λ2rT(1)kθ1}\left\{t\in\mathbb{R}\>:\>(y_{1}\alpha,t)\in s^{1/2}\lambda^{-2r}T(1)\text{k}_{-\theta_{1}}\right\} 3.58
𝔍2\mathfrak{J}_{2} the set {t:(y2σ(α),t)λ2r𝒲kθ2}\left\{t\in\mathbb{R}\>:\>(y_{2}\sigma(\alpha),t)\in\lambda^{2r}\mathcal{W}\text{k}_{-\theta_{2}}\right\} 3.58
KK real quadratic field (d)\mathbb{Q}(\sqrt{d}) with d2d\geq 2 square-free 1.3, 2.1
K¯\overline{K} the set K{}K\cup\{\infty\} 2.2
kik_{i} ii-th cusp of ΓK\Gamma_{K} 2.1
k𝜽\text{k}_{\bm{\theta}} matrix element (kθ1,kθ2)(\text{k}_{\theta_{1}},\text{k}_{\theta_{2}}) with kθ=(cosθsinθsinθcosθ)\text{k}_{\theta}=\left(\begin{smallmatrix}\cos\theta&-\sin\theta\\ \sin\theta&\cos\theta\end{smallmatrix}\right) 2.3
κ\kappa number of cusps of ΓK\Gamma_{K} 2.1
κ𝒫\kappa_{\mathcal{P}} relative density of visible points in a planar point set 𝒫\mathcal{P} 1.9, 5.4
μK\mu_{K} the probability H\mathrm{H}-invariant measure on ΓK\H\Gamma_{K}\backslash\mathrm{H} 2.12
i\mathcal{L}_{i} the set ΠiK\Pi_{i}\cap\mathcal{L}_{K} 2.24
~i\widetilde{\mathcal{L}}_{i} the set {(α,t1,σ(α),t2):αIi,t1,t2}ξi1\left\{(\alpha,t_{1},\sigma(\alpha),t_{2})\>:\>\alpha\in I_{i},\ t_{1},t_{2}\in\mathbb{R}\right\}\xi^{-1}_{i} 3.39
K\mathcal{L}_{K} Minkowski embedding of 𝒪K2\mathcal{O}_{K}^{2} in 4\mathbb{R}^{4} 1.10
(θ)\ell(\theta) projection of T(1)kθT(1)\text{k}_{-\theta} on the xx-axis with θ/2π\theta\in\mathbb{R}/2\pi\mathbb{Z} 3.3
𝒲(θ)\ell_{\mathcal{W}}(\theta) projection of 𝒲kθ\mathcal{W}\text{k}_{-\theta} on the xx-axis 3.3
λ\lambda fundamental unit of KK 1.3
mm Lebesgue measure on \mathbb{R} 3.1
N\mathrm{N} standard norm on KK defined by N(α)=ασ(α)\mathrm{N}(\alpha)=\alpha\sigma(\alpha) 2.1
Nr absolute norm of ideals 2.5
n𝒙\text{n}_{\bm{x}} matrix element (nx1,nx2)(\text{n}_{x_{1}},\text{n}_{x_{2}}) with nx=(1x01)\text{n}_{x}=\left(\begin{smallmatrix}1&x\\ 0&1\end{smallmatrix}\right) 2.3
𝒪K\mathcal{O}_{K} ring of integers of KK 1.3
𝒪K×\mathcal{O}_{K}^{\times} unit group of 𝒪K\mathcal{O}_{K} 1.3
να\nu_{\alpha} critical value where the ν\nu-extremality of αEI+EI\alpha\in E_{I}^{+}\sqcup E_{I}^{-} changes 4.5
ν(θ)\nu(\theta) parameter in the expression 𝒲(θ)=r(θ)(ν(θ),1)\ell_{\mathcal{W}}(\theta)=r(\theta)(-\nu(\theta),1) 1.16
𝒫AB,𝒘{\mathcal{P}}_{\operatorname{AB},{\text{\boldmath$w$}}} the Ammann-Beenker tiling corresponding to the translate 𝒘w 1.23, 5.2
𝒫Gh,𝒘{\mathcal{P}}_{\operatorname{Gh},{\text{\boldmath$w$}}} the Gähler’s shield tiling corresponding to the translate 𝒘w 1.27, 5.3
𝒫(𝒲,)\mathcal{P}(\mathcal{W},\mathcal{L}) cut-and-project set associated to a window set 𝒲\mathcal{W} and a lattice \mathcal{L} 1.4
π\pi natural projection from d+m\mathbb{R}^{d+m} to the physical space d\mathbb{R}^{d} 1.2
πint\pi_{\operatorname{int}} natural projection from d+m\mathbb{R}^{d+m} to the internal space m\mathbb{R}^{m} 1.2
Πi\Pi_{i} the plane ({0}××{0}×)ξi1(\{0\}\times\mathbb{R}\times\{0\}\times\mathbb{R})\xi^{-1}_{i} 2.23
Rα,νR_{\alpha,\nu} a rectangle depending on αI>0\alpha\in I\cap\mathbb{R}_{>0} and ν>0\nu>0 with II an integral ideal 4.3
R𝒲(θ,x)R_{\mathcal{W}}(\theta,x) the set {y:(x,y)kθ𝒲}\{y\in\mathbb{R}\>:\>(x,y)\text{k}_{\theta}\in{\mathcal{W}}\} 3.1
r(θ)r(\theta) parameter in the expression 𝒲(θ)=r(θ)(ν(θ),1)\ell_{\mathcal{W}}(\theta)=r(\theta)(-\nu(\theta),1) 1.16
S11S_{1}^{1} the unit circle 1.1
SIi,jS_{I_{i},j} the jj-th interval in the partition of >0\mathbb{R}_{>0} corresponding to the ii-th cusp of ΓK\Gamma_{K} 5
S~Ii,j\tilde{S}_{I_{i},j} the set {θ[0,2π):ν(θ)SIi,j}\left\{\theta\in[0,2\pi)\>:\>\nu(\theta)\in S_{I_{i},j}\right\} 5
SL2(𝒪K)\operatorname{SL}_{2}(\mathcal{O}_{K}) Hilbert modular group 2.2
σ\sigma the unique nontrivial automorphism of the real quadratic field KK 2.1
t1t_{1} an absolute positive number depending only on KK 2.4
τ\tau a generator of 𝒪K\mathcal{O}_{K} which is 1+d2\frac{1+\sqrt{d}}{2} if d1(mod 4)d\equiv 1\ (\mathrm{mod}\ 4) and d\sqrt{d} if d2,3(mod 4)d\equiv 2,3\ (\mathrm{mod}\ 4) 2.1
T(s)T(s) open triangle with vertices at (0,0)(0,0) and (s/c𝒫)1/2(1,±1)(s/c_{\mathcal{P}})^{1/2}(1,\pm 1) 3.3
𝒯(s){\mathcal{T}}(s) the set λ2rT(s)×λ2r𝒲\lambda^{-2r}T(s)\times\lambda^{2r}\mathcal{W} with rr\in\mathbb{N} such that λ2rs1/4<λ2(r+1)\lambda^{2r}\leq s^{1/4}<\lambda^{2(r+1)} 3.3
𝒯(s){\mathcal{T}}^{\prime}(s) the set T(s)×𝒲T(s)\times{\mathcal{W}} 3.3
𝒲\mathcal{W} window set of the cut-and-project set 𝒫(𝒲,)\mathcal{P}(\mathcal{W},\mathcal{L}) 1.2
𝒲AB\mathcal{W}_{\operatorname{AB}} an open regular octagon with particular size and position 2
𝒲Gh\mathcal{W}_{\operatorname{Gh}} an open regular dodecagon with particular size and position 1.1
𝒲𝒘(AB)\mathcal{W}_{{\text{\boldmath$w$}}}^{(\operatorname{AB})} the set (𝒲AB+𝒘)(1012)({\mathcal{W}}_{\operatorname{AB}}+{\text{\boldmath$w$}})\left(\begin{smallmatrix}1&0\\ 1&\sqrt{2}\end{smallmatrix}\right) 1.23
𝒲𝒘(Gh)\mathcal{W}_{{\text{\boldmath$w$}}}^{(\operatorname{Gh})} the set (𝒲Gh+𝒘)(1032)({\mathcal{W}}_{\operatorname{Gh}}+{\text{\boldmath$w$}})\left(\begin{smallmatrix}1&0\\ \sqrt{3}&2\end{smallmatrix}\right) 1.29
𝒲\mathcal{W}^{*} polar set of 𝒲\mathcal{W} 4.1
ξi\xi_{i} scaling matrix with respect to the ii-th cusp of ΓK\Gamma_{K} 2.9
YY the set {(y1,y2)(>0)2:y2[1,λ2)}\left\{(y_{1},y_{2})\in(\mathbb{R}_{>0})^{2}\>:\>y_{2}\in[1,\lambda^{2})\right\} 3.7
Yt1Y_{t_{1}} the set {(y1,y2)(>0)2: 1y1/y2<λ2,y1y2>t1}\left\{(y_{1},y_{2})\in(\mathbb{R}_{>0})^{2}\>:\>1\leq y_{1}/y_{2}<\lambda^{2},\ y_{1}y_{2}>t_{1}\right\} 3.38
ZZ the set (0,π)×(0,2π)(0,\pi)\times(0,2\pi) 3.4
ZJZ_{J} the set J×(0,2π)J\times(0,2\pi) 3.4
ζK\zeta_{K} Dedekind zeta function corresponding to KK 2.3

References

  • [1] R. Ammann, B. Grünbaum, and G. C. Shephard. Aperiodic tiles. Discrete Comput. Geom., 8(1):1–25, 1992.
  • [2] J. S. Athreya and J. Chaika. The distribution of gaps for saddle connection directions. Geom. Funct. Anal., 22(6):1491–1516, 2012.
  • [3] J. S. Athreya, J. Chaika, and S. Lelièvre. The gap distribution of slopes on the golden L. In Recent trends in ergodic theory and dynamical systems, volume 631 of Contemp. Math., pages 47–62. Amer. Math. Soc., Providence, RI, 2015.
  • [4] M. Baake, F. Götze, C. Huck, and T. Jakobi. Radial spacing distributions from planar points sets. Acta Crystallographica Section a. Foundations of Crystallography, 70:472–482, 2014.
  • [5] M. Baake and U. Grimm. Aperiodic order. Vol. 1, volume 149 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 2013. A mathematical invitation, With a foreword by Roger Penrose.
  • [6] M. Baake, P. Kramer, M. Schlottmann, and D. Zeidler. Planar patterns with fivefold symmetry as sections of periodic structures in 44-space. Internat. J. Modern Phys. B, 4(15-16):2217–2268, 1990.
  • [7] F. P. M. Beenker. Algebraic theory of non-periodic tilings of the plane by two simple building blocks: a square and a rhombus. Technical report, Eindhoven University of Technology, 1982.
  • [8] J. Berman, T. McAdam, A. Miller-Murthy, C. Uyanik, and H. Wan. Slope gap distribution of saddle connections on the 2n2n-gon. Discrete Contin. Dyn. Syst., 43(1):1–56, 2023.
  • [9] F. P. Boca, C. Cobeli, and A. Zaharescu. Distribution of lattice points visible from the origin. Comm. Math. Phys., 213(2):433–470, 2000.
  • [10] F. P. Boca, A. A. Popa, and A. Zaharescu. Pair correlation of hyperbolic lattice angles. Int. J. Number Theory, 10(8):1955–1989, 2014.
  • [11] A. Borel and Harish-Chandra. Arithmetic subgroups of algebraic groups. Ann. of Math. (2), 75:485–535, 1962.
  • [12] T. Dotera, S. Bekku, and P. Ziherl. Bronze-mean hexagonal quasicrystal. Nature Materials, 16:987–992, 2017.
  • [13] D. El-Baz. Spherical equidistribution in adelic lattices and applications. arXiv preprint arXiv:1710.07944, 2017.
  • [14] D. El-Baz, J. Marklof, and I. Vinogradov. The distribution of directions in an affine lattice: two-point correlations and mixed moments. Int. Math. Res. Not. IMRN, (5):1371–1400, 2015.
  • [15] N. D. Elkies and C. T. McMullen. Gaps in nmod1{\sqrt{n}}\bmod 1 and ergodic theory. Duke Math. J., 123(1):95–139, 2004.
  • [16] E. Freitag. Hilbert modular forms. Springer-Verlag, Berlin, 1990.
  • [17] B. Grünbaum and G. C. Shephard. Tilings and patterns. W. H. Freeman and Company, New York, 1987.
  • [18] F. Gähler. Matching rules for quasicrystals: the composition-decomposition method. Journal of Non-Crystalline Solids, 153 and 154:160–164, 1993.
  • [19] G. Hammarhjelm. Gap asymptotics of the directions in an Ammann-Beenker-like quasicrystal. arXiv preprint arXiv:2106.07422, 2021.
  • [20] G. Hammarhjelm. The density and minimal gap of visible points in some planar quasicrystals. Discrete Math., 345(12):Paper No. 113074, 2022.
  • [21] G. Hammarhjelm, A. Strömbergsson, and S. Yu. Supplement; programs for numerical experiments. https://www2.math.uu.se/\simastrombe/gap_asymptotics/supplement.html, 2024.
  • [22] D. Kelmer and A. Kontorovich. On the pair correlation density for hyperbolic angles. Duke Math. J., 164(3):473–509, 2015.
  • [23] W. Kim and J. Marklof. Poissonian pair correlation for directions in multi-dimensional affine lattices and escape of mass estimates for embedded horospheres. To appear in Ergodic Theory Dynam. Systems.
  • [24] L. Kumanduri, A. Sanchez, and J. Wang. Slope gap distributions of Veech surfaces. Algebr. Geom. Topol., 24(2):951–980, 2024.
  • [25] J. Marklof and A. Strömbergsson. The distribution of free path lengths in the periodic Lorentz gas and related lattice point problems. Ann. of Math. (2), 172(3):1949–2033, 2010.
  • [26] J. Marklof and A. Strömbergsson. Free path lengths in quasicrystals. Comm. Math. Phys., 330(2):723–755, 2014; correction, ibid. 374: 367, 2020.
  • [27] J. Marklof and A. Strömbergsson. Visibility and directions in quasicrystals. Int. Math. Res. Not. IMRN, (15):6588–6617, 2015; erratum, ibid. 2020.
  • [28] J. Marklof and A. Strömbergsson. The periodic lorentz gas in the boltzmann-grad limit: asymptotic estimates. Geom. Funct. Anal., 21(3):560–646, 2011.
  • [29] J. Marklof and A. Strömbergsson. Kinetic theory for the low-density lorentz gas. Mem. Amer. Math. Soc., 294(1464):x+136, 2024.
  • [30] J. Marklof and B. Tóth. Superdiffusion in the periodic lorentz gas. Comm. Math. Phys., 347(3):933–981, 2016.
  • [31] J. Marklof and I. Vinogradov. Directions in hyperbolic lattices. J. Reine Angew. Math., 740:161–186, 2018.
  • [32] H. L. Montgomery and R. C. Vaughan. Multiplicative number theory. I. Classical theory, volume 97 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2007.
  • [33] K. Niizeki. A dodecagonal quasiperiodic tiling with a fractal window. Philosophical Magazine, 87(18-21):2855–2861, 2007.
  • [34] K. Niizeki. Self-similar quasilattices with windows having fractal boundaries. Journal of Physics A: Mathematical and Theoretical, 41(17):175208, 2008.
  • [35] T. Osman, J. Southerland, and J. Wang. An effective slope gap distribution for lattice surfaces. arXiv preprint arXiv:2409.15660, 2024.
  • [36] M. Ratner. On Raghunathan’s measure conjecture. Ann. of Math. (2), 134(3):545–607, 1991.
  • [37] M. Ratner. Raghunathan’s topological conjecture and distributions of unipotent flows. Duke Math. J., 63(1):235–280, 1991.
  • [38] M. S. Risager and A. Södergren. Angles in hyperbolic lattices: the pair correlation density. Trans. Amer. Math. Soc., 369(4):2807–2841, 2017.
  • [39] R. Rühr, Y. Smilansky, and B. Weiss. Classification and statistics of cut-and-project sets. J. Eur. Math. Soc. (JEMS), 26(9):3575–3638, 2024.
  • [40] A. Sanchez. Gaps of saddle connection directions for some branched covers of tori. Ergodic Theory Dynam. Systems, 42(10):3191–3245, 2022.
  • [41] H. Shimizu. On discontinuous groups operating on the product of the upper half planes. Ann. of Math. (2), 77:33–71, 1963.
  • [42] C. L. Siegel. The volume of the fundamental domain for some infinite groups. Trans. Amer. Math. Soc., 39(2):209–218, 1936. Correction in: Zur Bestimmung des Volumens des Fundamentalbereichs der unimodularen Gruppe. Math. Ann. 137 (1959), 427–432.
  • [43] C. L. Siegel. Über die analytische Theorie der quadratischen Formen. III. Ann. of Math. (2), 38(1):212–291, 1937.
  • [44] C. Uyanik and G. Work. The distribution of gaps for saddle connections on the octagon. Int. Math. Res. Not. IMRN, (18):5569–5602, 2016.
  • [45] G. van der Geer. Hilbert modular surfaces, volume 16 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1988.
  • [46] L. C. Washington. Introduction to cyclotomic fields, volume 83 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1997.