Asymptotic estimates of large gaps between directions in certain planar quasicrystals
Abstract.
For quasicrystals of cut-and-project type in , 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 -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.
Dedicated to Gustav Hammarhjelm (1992–2022).
1. Introduction
1.1. Gaps between directions in planar point sets
Given a locally finite point set in the plane , we consider its set of directions, that is, the set of points on the unit circle as runs through . For each , let be the finite multi-set of directions to points in within distance from the origin, i.e.,
(1.1) |
Throughout the paper we will assume that has an asymptotic density , meaning that for any bounded set with boundary of measure zero,
(1.2) |
It then follows that the multi-set has cardinality as , and furthermore that this multi-set becomes asymptotically equidistributed along the unit circle, in the sense that for any arc ,
where denotes the length of (in particular ).
In this situation, it is natural to consider finer questions about the local statistics of the points in . A particular statistics which has been much studied is the distribution of normalized gaps between the points in , as . For example, when is the set of primitive lattice points in , the limiting distribution of normalized gaps was explicitly determined by Boca, Cobeli and Zaharescu [9]. More generally, if is an arbitrary lattice (possibly translated) in , for any , 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 by an arbitrary lattice in ; furthermore, when is an arbitrary ’irrational’ translate of a lattice in , the limiting distribution of gaps between directions coincides with the gap distribution for the fractional parts of 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 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 -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 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 , for 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 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 , set , and let be a lattice of full rank in . We refer to and as the physical space and internal space, respectively. In the present paper the dimension of the physical space will always be . We write and for the orthogonal projection of onto the first coordinates and last coordinates. Let be the closure of ; this is a closed abelian subgroup of . Let be the identity component of ; this is a linear subspace of , and set . Let , the window, be a bounded subset of with nonempty interior (with respect to the topology of ). We will always assume that and are such that
(1.3) |
We now define the cut-and-project set associated to and to be
(1.4) |
We will always assume that is regular, meaning that has measure zero with respect to the Haar measure of . Under these assumptions, it is known that has an asymptotic density, i.e. (1.2) holds, with being given by a simple explicit expression in terms of and [26, Prop. 3.2]; see also (1.11) below.
Let us view the unit circle as the set of with , and for any let us call the number the normalized angle of ; this gives an identification between and . Let us order the normalized angles of the points in in an increasing list as
(1.5) |
where . Also set .
Theorem 1.
We recall the explicit formula for from [27] in Section 3.3 below. Figure 1 shows conjectural graphs of the function , 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 in to be a translate of a lattice in , whereas in the present paper we always require to be a genuine lattice, i.e. . In fact, in Theorem 2 below we will also assume that , and so .
We also point out that the proof of [27, Cor. 3] immediately extends (by utilizing the freedom of choice of the measure “” in [27, Thm. 2]) to show that the limiting gap distribution for remains the same if we restrict attention to the directions lying in any fixed subinterval of . That is, the following more general version of (1.6) holds: For any fixed ,
(1.8) |
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 which are ’invisible’ from the origin; thus we allow to be a multi-set, and we allow equalities in the list in (1.5). Also, our function equals the function in the right hand side of [27, (1.15)], and is thus in general not the same as “” in [27, Cor. 3]. Clearly by (1.6), our function satisfies . In general has a jump discontinuity at ; it follows from [27, Cor. 3] that where is the relative density of visible points in , a quantity which was defined and proved to exist in [27]. Note that it is immediate to translate between the two versions of “”, except that it requires knowledge of the constant , which is non-trivial to compute. See [20] for explicit formulas for 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 in Theorem 1, for a particular class of cut-and-project sets in , 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 be a real quadratic field; let be its ring of integers; let be the unique non-trivial automorphism of , and let be the fundamental unit (thus and ). Let be the Minkowski embedding of in , viz.,
(1.10) |
We set , that is, we view as the product of a 2-dimensional physical space and a 2-dimensional internal space, and and denote the orthogonal projection of onto the first 2 coordinates and the last 2 coordinates, respectively. Note that in this case is dense in , i.e. we have , and so we take the window to be a bounded subset of with nonempty interior. (In the present case the restriction of to is injective, and so the property (1.3) is automatically fulfilled.) We will always assume that is regular, i.e. that has Lebesgue measure zero.
For this class of cut-and-project sets, the formula for the asymptotic density of , [26, (1.7)], becomes (see also Section 2 below):
(1.11) |
where is the discriminant of .
As we prove in Remark 3.29 below, the limiting gap distribution function is unchanged if is modified by any measure zero set. In particular, without loss of generality we may assume is open (indeed, otherwise replace by its interior). Finally, in the present paper we will make the key assumption that .
In our approach we actually work with the integral of the distribution function , i.e. with
(1.12) |
As we will note below (see Lemma 3.2), using the fact that is decreasing, an easy interpolation argument allows us to deduce an asymptotic formula for once we know an asymptotic formula for .
Our main result is the following:
Theorem 2.
Let where is a real quadratic field and is a bounded open subset of such that and has Lebesgue measure zero. Let be the associated limiting gap distribution function as in Theorem 1, and let . Then there exists a positive constant such that
(1.13) |
If we further assume to be convex, then we have
(1.14) |
Remark 1.15.
As mentioned above, in order to prove Theorem 2, it suffices to prove the two asymptotic estimates for in (1.13) and (1.14). Both these estimates will in fact follow from a more precise, general asymptotic formula for with an explicit error term; see (3.7) below. We also mention that our analysis leads to an explicit formula for the leading coefficient ; see (3.8).
1.4. Formula for
As we will now describe, after imposing one more assumption on the window set, we can further evaluate the aforementioned formula for . To state our result, we need to introduce some more notation. For we write . Then for any subset we let be the projection of on the -axis. On top of the assumptions in Theorem 2, we will now assume that for each , is an interval. Note that this assumption is always satisfied if is connected, but it also holds for many non-connected window sets .
Note that since is open and , is an open subset of containing . Hence if is assumed to be an interval, then we can parametrize it by two positive numbers via
(1.16) |
Theorem 3.
Retain the notation and assumptions in Theorem 2 and further assume that is of class number one, and that is an interval for each . Then there exist a finite partition of into intervals, and non-negative constants depending only on , such that
(1.17) |
where is the Dedekind zeta function attached to , , and and are defined by (1.16).
Remark 1.18.
We stress that the intervals are allowed to be open, closed, or half-open, and may be degenerate, i.e. of the form for some .
Remark 1.19.
Our analysis applies for any real quadratic field 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 and the constants 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 , there is one that is always non-empty, namely, the unique such that . In certain cases, this equals the whole interval while all other ’s are empty. In this case, the formula for can be simplified into
where is some positive constant depending only on , and is the polar set of , i.e.
(1.21) |
Here is the standard scalar product in . See Section 4.1 for more details. If is also convex and centrally symmetric, the product appearing in the above formula is known as the Mahler volume of .
Remark 1.22.
For any as in Theorem 3, replacing by its convex hull does not affect the intervals , hence does not affect the functions and . Therefore, in the explicit formula (1.17), only the factor is affected when replacing by its convex hull. In this connection it should also be noted that the polar set of the convex hull of equals the polar set of .
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.
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 , and let be the open regular octagon centered at the origin of edge length , oriented so that four of the edges are perpendicular to a coordinate axis. For each we set
(1.23) |
where and . Then for any with the property that , the cut-and-project set is the vertex set of an Ammann-Beenker tiling with one vertex at , and conversely, the vertex set of any Ammann-Beenker tiling having a vertex at is (up to scaling and rotation) either equal to such a point set , or can be obtained as a limit of such point sets 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 for any . Here the constant is given by a quite simple formula, also valid for much more general window sets:
Proposition 1.1.
Note that Proposition 1.1 applies when for any , and the formula (1.25) holds whenever lies sufficiently near . In particular, (1.25) implies that
(1.26) |
In Section 5.4, we present a comparison, for a few examples of points , between the exact values of computed using Proposition 1.1, and numerically computed approximate values of for large .
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 , and let be the open regular dodecagon centered at the origin of edge length , so that four of the edges are perpendicular to a coordinate axis. For each we set
(1.27) |
where and . Then for any with the property that , the cut-and-project set is the vertex set of a Gähler’s shield tiling with one vertex at . (See [5, Ch. 7.3] and Section 5.2 below.)
Proposition 1.2.
Note that Proposition 1.2 applies when for any , and the formula (1.29) holds whenever lies sufficiently near . Again see Section 5.4 for numerical computations related to the values of for .
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 and , and let be the open regular decagon centered at the origin of edge length , oriented so that two of the edges are perpendicular to the first coordinate axis. For each we set
(1.30) |
where and . Then for any with the property that , the cut-and-project set is the vertex set of a Tübingen triangle tiling with one vertex at . (See [5, Ch. 7.3], [6, Sec. 4] and Section 5.3 below.)
Proposition 1.3.
Proposition 1.3 applies when for any , and the formula (1.32) holds whenever lies sufficiently near . 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 be a real quadratic field with square-free. Let be its ring of integers, where
Let be the (abelian) group of fractional ideals of , and let be the subgroup of principal ideals. We let be the ideal class group of , and we call elements in ideal classes of . For any nonzero , we denote by the fractional ideal generated by .
Let be the unique non-trivial automorphism of and consider the embedding
(2.1) |
We note that is a lattice in for any . In particular, is the lattice generated by and , which is of covolume , where
(2.4) |
is the discriminant of . For any fractional ideal , its absolute norm is defined by
(2.5) |
Note that Nr is multiplicative in , that is, for any . Moreover, for any , we have , where is the standard norm on given by .
2.2. Hilbert modular group
Let be the upper half plane. The group acts on via the Möbius transformation: for any and . Let (which we view as a subgroup of via the block diagonal embedding) and let be the product of two upper half planes. The group naturally acts on via
Denote by which can be identified with the boundary of . The action of on naturally extends to via the same formula. We also denote by . It embeds into via the natural extension of the embedding (by sending to ). With slight abuse of notation, we also denote this embedding from to by . We will also write for the group homomorphism given by
(2.6) |
The Hilbert modular group for the field is given by
We write for the embedding of in :
The discreteness of in implies that is a discrete subgroup of . Indeed, it is a non-uniform lattice in , that is, the homogeneous space is non-compact and has finite volume with respect to a Haar measure of . As a subgroup of , naturally acts on , and this action preserves . The orbits of under are called the cusps of . We will often represent a cusp by an element in the corresponding -orbit, or by an element in (via the natural identification between and ). The following lemma, together with the obvious identification between and , shows that the number of cusps of equals the number of ideal classes of .
Lemma 2.1 ([16, Lemma 3.5]).
The map sending to induces a bijection between and , where is the ideal generated by if and if . In particular,
Throughout the remainder of this paper, we fix to be a complete list of cusps of . Fix , and , so that for each ; then set . Then are integral ideals which by Lemma 2.1 form a system of representatives of the ideal classes of . (Note that depends on the choice of ; however the class depends only on .)
For each , let
(2.7) |
be the isotropy group of the cusp . Below we give a more precise description of these isotropy groups. First, the isotropy group of is easy to compute: For , by direct computation if and only if , that is
where is the subgroup of upper triangular matrices in .
To compute the other isotropy groups, we will translate the cusp to . For any integral ideal let
(2.8) |
For each since , we can find such that . Throughout the paper we fix the scaling matrix
(2.9) |
Note that satisfies . We then have the following description of .
Lemma 2.2.
We have for each ,
(2.10) |
Proof.
By definition, if any only if . Since , this is equivalent to . This shows that
By direction computation we have and , implying that . Hence
We can then finish the proof by conjugating both sides of the above equation by . ∎
2.3. Coordinates and measures
Let
be the Minkowski embedding of in . Let be the space of lattices of the form with . Since if and only if , can be parameterized by the homogeneous space via . We thus identify with . Let be the unique invariant probability measure on .
Since we will be working with this measure extensively when computing later in Section 3, here we give a more explicit description of it in terms of coordinates from an Iwasawa decomposition of . It is well known that the group has an Iwasawa decomposition saying that any element can be written uniquely as for some and , where , and . This thus induces an Iwasawa decomposition of : Every can be written uniquely in the form with , and , where
Under these coordinates, the Haar measure of (up to scalars) is given by
(2.11) |
Note that essentially comes from a Haar measure of . Indeed, we may identify with a fundamental domain inside ; then is the restriction of a certain Haar measure to this fundamental domain normalized to be a probability measure. Thus is given by
(2.12) |
for some normalizing factor . This normalizing factor was computed by Siegel [42, 43] and is given by the following formula (see [45, p. 59]):
(2.13) |
where () is the Dedekind zeta function attached to . Here the summation is over all the nonzero integral ideals of .
Remark 2.14.
To compare our formula for in (1.17) with numerical computations, we need to express in more explicit terms. Let be the Dirichlet character defined by with the Kronecker symbol. Note that is an even primitive quadratic character of modulus ; see [32, p. 296-297]. Using the fact that for any prime number , the ideal is inert, ramified or split if and only if equals respectively, we have the following well-known formula for :
where is the Riemann zeta function and is the Dirichlet -function associated to . Now by [46, Prop. 4.1 and Thm. 4.2] we have for any ,
where is the -th Bernoulli polynomial given by the formula with the -th Bernoulli number. Combined with the functional equation of (see [45, p. 59]), we get
(2.15) |
For instance when we have and is the quadratic character of modulus with and . We also have and . Plugging all these relations we get in this case . Similarly, one can apply (2.15) to get and .
2.4. Siegel domain and cusp neighborhoods
In this section we give a more precise description of the homogeneous space 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 the set
(2.16) |
is a fundamental domain for . Here is a fundamental domain for . (Recall that is a lattice in for any fractional ideal of .)
Now for each and any define
(2.17) |
It follows from Shimizu’s lemma [41, Lemma 5] that there exists some constant depending only on such that
(2.18) |
Moreover, by the reduction theory of Borel and Harish-Chandra [11] we have a Siegel fundamental domain of the form
(2.19) |
Here is compact and the natural map from to is surjective and finite-to-one. Note, on the other hand, it follows from (2.18) that the projection from to is injective.
2.5. Geometry of lattices avoiding large balls
In this section we give necessary conditions for lattices avoiding large balls. These lattices will be the main objects we deal with when computing . We have the following description of these lattices.
Proposition 2.3.
There exists some depending only on such that for any and for any lattice avoiding a ball of radius , we have for some . Moreover, we have with the bounding constants depending only on .
Remark 2.20.
Regarding the notation “”, “” and “”: For two positive quantities and , we write or to mean that there is a constant such that , and we will write for . We will sometimes use subscripts to indicate the dependence of the bounding constant on parameters.
Proof.
Let , , and . Note that is a basis for . Let be sufficiently large such that
(2.21) |
We note that depends only on the compact set (hence only depends on ). Now take and suppose avoids a ball of radius . To prove the first claim, in view of the description of the Siegel fundamental domain in (2.19), we want to show there does not exist any such that . We prove by contradiction. Suppose there exists such . Then is a basis for . Let be the center of and write () as an -linear combination of these basis vectors. For each , take the closest integer to so that we have . Consider the lattice vector . Since , we have
implying that there exists such that , contradicting (2.21). This proves the first claim, i.e. for some . For the second claim, without loss of generality we may assume . To show , first note that and . Thus
Here denotes the operator norm (with respect to the Euclidean norm). On the other hand, for any ,
This, together with the bound implies that as claimed. ∎
For later reference, note that for each ,
(2.22) |
with as before. Let . We denote by
(2.23) |
and
(2.24) |
By Proposition 2.3 if avoids some ball of radius , then with for some and with . Then the two vectors
are of length . Note that and are linearly independent and is the primitive rank two sublattice of containing them. Thus lies -densely in the plane . For this reason, we call the filled plane of .
3. Asymptotics of
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 and its integral , 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 , and also an explicit formula for the leading coefficient . As we mentioned in the introduction, the asymptotics for will be derived as a consequence of those for ; see Lemma 3.2 below.
In order to state Theorem 4, we introduce some further notation. For any we write . Then for any and any subset , we set
(3.1) |
It follows that , which we defined in Section 1.4, can be expressed as
(3.2) |
Note also that satisfies the relation for any . For any integral ideal , any bounded Lebesgue measurable set with non-empty interior, and any , we set
(3.3) |
For later reference we record here a useful identity for the function which can be checked directly from its definition: For any with and for any ,
(3.4) |
Let us write for the Lebesgue measure on . For any Lebesgue measurable set and , we define to be the infimum of when ranges over all intervals of length :
(3.5) |
Note that always, and whenever contains some interval of length . Let us also define
(3.6) |
Recall from below Lemma 2.1 that we have made a fixed choice of ideals of , and that these form a system of representatives of the ideal classes of .
Theorem 4.
Let where is a real quadratic field and is a bounded open subset of such that and has Lebesgue measure zero. Let be the associated limiting gap distribution function as in Theorem 1, and let . Then there exists a positive constant such that for all sufficiently large ,
(3.7) |
where
(3.8) |
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 of , we have
(3.10) |
and furthermore, the product
(3.11) |
depends only on the ideal class of .
To prove (3.10), it suffices to note that
the differential form is invariant under the map sending to ,
a fact which follows by applying (3.4) for the special case when .
Next, to prove the invariance of (3.11),
assume that is another integral ideal belonging to the same ideal class as .
Then for some nonzero , which we may take to satisfy (otherwise replace by ).
Then the absolute norm scales by when replacing by .
On the other hand, using (3.4) and the fact that for any ,
we have
from which it follows (via the obvious substitution and then using (3.10) with in place of ) that the double integral in (3.11) scales by a factor when replacing by . Hence the product in (3.11) is invariant as claimed.
Remark 3.12.
The right hand side of (3.8) is invariant under replacing by , for any .
Proof.
We will prove the claim by showing that for any integral ideal of , and any ,
(3.13) |
Write , . For fixed , define the smooth maps and through
(3.14) |
Note that is a diffeomorphism of onto itself. Differentiating both sides of (3.14) with respect to , and using the fact that , we have
(3.15) |
It also follows from (3.15) and (3.14) that
(3.16) |
where in the fourth equality we used the fact that for any and in . Note that (3.15) and (3.16) imply that if then and for all , while if then and for all .
Using (3.14) and (3.15) we have for any , and hence via (3.1) and (3.2) one gets . We also have , , and , and hence:
Using also the simple relation (), it follows that
(3.17) |
where we substituted and then used (3.10). The two formulas for in (3.15) and (3.16) imply that . Using this in the last expression in (3.17), and then taking as a new variable of integration, we obtain the identity (3.13). ∎
Let us also note that for any , 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 , in order to extend it to all it suffices to verify that it also holds for and (); 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 in (1.13) and (1.14); then, by a standard general argument using the convexity of , we will deduce the tail asymptotics of .
We start with (1.13). We need to show that for any as in Theorem 2, the error bound in (3.7) is as . Take so that , where is the open disc with center and radius . Then for every with , and for every , we have and hence for all . It follows that the integral over in (3.7) can be replaced by . Hence, if we take as in Theorem 4, and set
then our task is to prove that as . Using we have for all , and ; hence by the Lebesgue bounded convergence theorem, it suffices to prove that as for any fixed . But if then for all , by the definition (3.6), while if then since is open, contains a non-empty interval, and so for all sufficiently large ; therefore as . This concludes the proof that the error bound in (3.7) is as , i.e. we have proved that , as stated in (1.13).
For the proof of (1.14) we need the following simple lemma regarding the function .
Lemma 3.1.
Let . Then for any interval of length and for any ,
Proof.
Note that for and for . Also . Hence for :
while for :
The desired estimate then follows from these two estimates noting that for and for . ∎
Now let us further assume that is convex. Then for all and , is either empty or an interval. Hence, using Lemma 3.1 and the fact that , it follows that the error bound in (3.7) is
(3.18) |
Choose such that . Let us temporarily fix , and set ; then the function vanishes for , but is positive and concave for . Hence for our fixed , the inner integral in (3.18) equals . We also have and , and so, by the concavity, for all . Hence
and by a simple computation, the last integral is seen to equal for all . Since this holds for every , we conclude that the expression in (3.18) is for all large , i.e. we have proved the asymptotics for stated in (1.14).
The next lemma shows how to use the convexity of to deduce the tail asymptotics of from that of .
Lemma 3.2.
Let be a differentiable function with continuous and decreasing. Suppose further that there exist constants and and a function , such that
(3.19) |
and
(3.20) |
Then we have
(3.21) |
where the implied constant depends only on and .
Proof.
Since by assumption is continuous and decreasing, we have for all , . Hence, assuming , writing , and using (3.19) and (3.20), we have
From this, using also , we get
Here we optimize by choosing ; note that this choice of is admissible, i.e. yields , because of our assumption . The conclusion is that holds for any .
Similarly, using , we have for all :
Here choose ; then , so that our choice is admissible for any . For these , we obtain . Combining this with the bound proved above, we have proved (3.21) (with an implied constant ). ∎
Let us now use Lemma 3.2 to conclude the proof of (1.13). Recall that in the situation of Theorem 2, is continuous and decreasing for (see Theorem 1), and . Furthermore, we have proved the first half of (1.13), i.e. that as . Hence, for any given constant , Lemma 3.2 applies with , , and an appropriate constant (depending on ), to yield as . Here the implied constant is independent of , and by taking arbitrarily small we conclude that . 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 and with an appropriate . This concludes the proof of Theorem 2.
Remark 3.22.
Let us note that there exist quite “nice” (but non-convex) windows for which the relative error bound in (3.7) in Theorem 4 tends to zero more slowly than any prescribed rate, as . One way in which this can happen is if has a cusp of an appropriate asymptotic shape.
To give a concrete statement, let us define
(3.23) |
Then (3.7) says that as , viz., is the relative error bound in this asymptotics, and our proof of the relation in (1.13) shows that as . Now let be an arbitrary decreasing function with . Choose a sequence satisfying both and ,111For example, one may choose . and then let be a continuous, increasing function satisfying , , and for all . (Note that we may even take to be smooth for .) Set
and with being a fixed vector in chosen so that . Finally let be the cut-and-project set , for any fixed real quadratic field . We claim that in this case, , viz., tends to zero more slowly than the given function !
To prove the claim, we first note that for every , writing for the first coordinate of , one easily verifies that for every , is a non-empty open interval of length . Hence by Lemma 3.1, for each such ,
This implies that
(3.24) |
Clearly is a decreasing function of . Also, for each we have for all , and so . Hence by our choice of , we have . Note also that for every and every , we have ; hence . In view of (3.24), this implies that .
As an interesting open problem, we mention that it seems plausible to us that, by using windows as above (perhaps with some further restrictions), also the actual relative difference, , can be proved to tend to zero more slowly than any prescribed rate.
Remark 3.25.
On the other hand, there exist with fractal boundary for which (3.7) in Theorem 4 implies as strong an error bound as for convex window sets, i.e. as in (1.14) in Theorem 2. For example, this holds if 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 contains an interval of length then for all . Hence by the argument leading to (3.18), the error bound in (3.7) is
where for we write for the supremum of the lengths of all intervals contained in . Now temporarily fix and write . One verifies that, if is an open set (with ) such that is the standard Koch snowflake, then there exists a constant independent of such that for all . This leads to , as claimed.
3.3. A preliminary integral formula for
The remainder of Section 3 is devoted to proving Theorem 4. We will start by recalling from [27] an expression for the function 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 is as in (1.4) with ; in particular is a lattice in where . Set and , and choose and so that . Let be the embedding of in given by
It follows from Ratner’s work [36, 37] that there exists a unique closed connected subgroup of such that is a lattice in , , and the closure of in equals . Let be the Haar measure on normalized so that . Now in Theorem 1 is given by [27, (3.6) and (11.1)]:
(3.26) |
with
(3.27) |
where is the open triangle in with vertices at and . (Recall that denotes the asymptotic density of ; see (1.2).) It is known that is on , and convex; hence is continuous and decreasing [27, Sec. 11].
Remark 3.28.
Remark 3.29.
It follows from the formula (3.27) that (and hence also ) remains unchanged if is modified by a null set (with respect to the Haar measure of ). Indeed, let be any bounded subset of with boundary of measure zero, satisfying , and let be given by (3.27) but with in place of . Let be given, and let be the indicator function of the set . Then
(3.30) |
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 in (3.30) by requiring , since . since .
Because of the -invariance of , the formula (3.27) remains valid if in the right hand side we replace “” by “”, for any given ; and applying this with we obtain:
(3.31) |
where is the open triangle with vertices at and .
Next we specialize to the setting of Theorem 4. Thus we take and let be as in (1.10); this means that , and , and that and are such that . Then by [26, Sec. 2.2.1] we have , where is as in Section 2.2. Furthermore,
(3.32) |
Indeed, we have , and every satisfies , hence and so . Conversely, assume , and write . Then implies that , and so . It follows in particular that and lie in , and this in turn implies that and , , , . Hence , and (3.32) is proved.
It follows that the map () is a diffeomorphism from onto , carrying to . Hence (3.31) can be rewritten as
(3.33) |
Here let us also note that is equivalent with , where
Hence (3.33) can be rewritten as follows:
(3.34) |
This formula will be the starting point of our proof of Theorem 4. Since we will be only concerned about the asymptotics of for large s, in the remainder of this section we will always assume .
Remark 3.35.
Using (3.34) and the fact that for any , it is easy to verify that is unchanged if is replaced by for any .
3.4. Separating the main term and error term
We will apply results from Section 2.5 to further analyze the condition . For this we first renormalize the sets to produce sets containing large balls (for large ).
Lemma 3.3.
For let be such that , and let
Then for any , holds if and only if .
Proof.
Let . The lemma then follows immediately by noting that and . ∎
Since contains a ball of radius , Proposition 2.3 gives that for sufficiently large, the condition forces for some , that is, the point in belongs to the -th cuspidal neighborhood. Using also the fact that these cuspidal neighborhoods are pairwise disjoint, by (2.18), it follows that for sufficiently large we have
where for each ,
(3.36) |
For the remainder of this section, we fix an index , and seek an asymptotic formula for the function as .
First we note that since projects injectively into (by (2.18) applied with ), we may express as an integral over . In view of (2.12), (2.16) and (2.17), we get:
(3.37) |
where as in (2.13),
and
(3.38) |
Now for sufficiently large, let be such that . Then by Proposition 2.3 we have . As discussed in Section 2.5, since avoids a large ball and , it contains a rank two sublattice (cf. (2.24)) which sits densely in the corresponding filled plane (cf. (2.23)). Let us write
(3.39) |
Then , and is the union of all -translates of the filled plane . Here is as before and the second equality in (3.39) follows from (2.22) and (2.23). Our strategy of computing is to replace the condition by the slightly stronger (since is very close to when is sufficiently large) and more manageable condition . Thus we define, for :
(3.40) |
and
(3.41) |
where in both integrals we use the notation . It is clear from the definitions that .
We start with a few auxiliary lemmas. For , let be the projection of on the -axis. Recall also that is the projection of on the -axis.
Lemma 3.4.
Fix and let be such that as before. For , the condition holds if and only if
(3.42) |
where is the Minkowski embedding of as before.
Proof.
Note that
and recall that . Hence holds if and only if
which is easily seen to be equivalent to (3.42). ∎
Remark 3.43.
Similarly, using (see (2.22)), one verifies that holds if and only if there exist and such that
(3.44) |
As a consequence of Lemma 3.4 we have a simple necessary condition for .
Lemma 3.5.
Keep the assumptions as in Lemma 3.4. If , then .
Proof.
Recall that we are assuming , i.e. that contains a small disc centered at the origin; therefore for every . Now if , then , and hence , and so by Lemma 3.4. ∎
Because of Lemma 3.5, we will often be able to reduce our discussion to the case , where
In other words, we will be able to reduce the domain for to be
3.5. Computing the main term
The goal of this section is to obtain simpler formulas for the function for large . 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 .
Lemma 3.6.
For any fractional ideal of , there exists a constant such that for any two intervals with , we have . Here denotes the length of the interval .
Proof.
This is a direct consequence of the fact that is invariant under for any , so that we can renormalize the rectangle to be such that . ∎
We next give a first explicit formula for . Recall that denotes the projection of on the -axis.
Proposition 3.7.
For all sufficiently large we have
(3.45) |
where .
Proof.
Using (3.40), Lemma 3.4 and Lemma 3.5, and the fact that is invariant under the action of , we have
(3.46) |
Because of , there is an such that for all . Furthermore, is an open interval of length for all . Hence by Lemma 3.6, there is a constant such that holds for all and all subject to . It follows that if is sufficiently large, then the formula in (3.46) remains valid if the domain of integration (defined in (3.38)) is replaced by the larger set
i.e. we have (after also changing the order of integration)
(3.47) |
Going through the same arguments but replacing the condition by in the definition of in (2.16) (this amounts to choosing a different fundamental domain for ), one verifies that (3.47) remains valid for all sufficiently large with replaced by the set
Equivalently, we have for sufficiently large
(3.48) |
where
Next, using the fact that the lattice is invariant under the action of , it follows that the integrand in (3.48) is invariant under , for each fixed . Also the measure is invariant under this map and both and are fundamental domains for . Hence the formula (3.48) remains valid if the domain of integration is replaced by , 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 defined in (3.3).
Proposition 3.8.
For all sufficiently large we have
(3.49) |
In particular, we have for all sufficiently large ,
(3.50) |
where is as in (3.8).
Proof.
The formula in Proposition 3.7 can be expressed as
(3.51) |
where
Here recall that equals the projection of on the -axis, which we can compute to be:
(3.52) |
Recall the definition of in (3.3). We also give a companion definition here. For any integral ideal and any bounded Lebesgue measurable set with non-empty interior, we set for any ,
(3.53) |
With these definitions the formula for can be rewritten as
implying that
Now note that from which we can deduce that . Hence
implying that
Plugging this back to (3.51) we get the desired formula for , (3.49).
3.6. Estimating the error term
Recall that for all sufficiently large , and by definition we have ; 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 is majorized by the bound in (3.7) for all sufficiently large .
We first prove two auxiliary lemmas.
Lemma 3.9.
For any fractional ideal of , there exists a constant such that for any two intervals with , and for any Lebesgue measurable subset , we have
where , and is the quotient map.
(Recall that denotes Lebesgue measure on ; also for an interval we write for its length.)
Proof.
Let us fix a fundamental parallelogram for the lattice in , and then fix intervals such that . We claim that the statement of the lemma holds with . Indeed, assume that are intervals with . Then there exists some such that and , viz., . This means that a translate of contains ; and hence contains a translate of . But using the fact that is invariant under , it follows that is again an fundamental parallelogram for the lattice in . Hence
Using and , the above is
∎
Lemma 3.10.
For any fixed set we have
(3.54) |
Proof.
Let be given, and let be an arbitrary interval of length . Take so that ; then there exist pairwise disjoint (open) intervals which all have length . Now , and since this holds for all intervals of length , it follows that . ∎
We now turn to bounding the error term , which we defined in (3.41), for large .
First of all, note that if then contains the line segment between and . Recall also that since , there is an such that . It follows that if then the condition (3.44) is satisfied with whenever belongs to the rectangle . This rectangle has area , since . Hence by Lemma 3.6, for sufficiently large there always exists some such that belongs to the rectangle, and so, by Remark 3.43, . This proves that the contribution from all to the integral in (3.41) is zero. Thus, for sufficiently large, we may replace the range of integration for in (3.41) by .
Next, as in the discussion in Section 3.4 (near (3.38)), because of the condition in the integrand, any which contribute to the integral in (3.41) must satisfy . Furthermore, by Lemma 3.4, given any which makes the integrand in (3.41) nonzero, there exists some such that belongs to the set
(3.55) |
Since we have restricted to , we have , and hence . Note that and are contained in bounded intervals independent of ; hence for , the set is contained in a fixed bounded region . Set
This is a finite subset of , and our discussion shows that for any which makes the integrand in (3.41) nonzero, there must exist some satisfying , or equivalently, . Using now the fact that both the minima of and for are bounded away from zero, and also the fact that both and are bounded away from zero independently of , it follows that we must have . We have thus proved that there exist some constants (which depend on and , but not on ), such that for all sufficiently large , every which makes the integrand in (3.41) nonzero, must satisfy .
In view of the above discussion, we have for sufficiently large:
(3.56) |
Next, the condition in (3.56) implies by Remark 3.43 that (3.44) fails for all . This is equivalent to saying that for every we have , where
(3.57) |
with and defined by
(3.58) |
Hence from (3.56) we get, using also the fact that if holds for every , then in particular it holds for the which is our summation variable:
(3.59) | |||
Note here that , and recall that is a fundamental domain for . Note also that the map induces a diffeomorphism from onto , carrying d to . Hence the integral over in (3.59) equals
(3.60) |
Let be the quotient map from onto the torus . Note that the condition is equivalent with . Also, since , is a sublattice of ; hence induces to a covering map from onto , which preserves Lebesgue measure. Using (2.5) it follows that this covering map has degree . Hence the integral in (3.60) equals
But we have , where is an open interval which is non-empty if and only if . This last condition is guaranteed to hold because of the condition appearing in (3.59). It now follows from Lemma 3.9 that there is a constant such that, for any interval of length , the integral in (3.60) is bounded from above by . Furthermore, comparing (3.58) and (3.1) we see that . Hence, writing with and , it follows that the integral in (3.60) is bounded from above by
This is true for every . Hence, using the notation introduced in (3.5), we conclude that the integral in (3.60) is bounded from above by
Note also that holds if and only if both and , and we recall that the first of these conditions is equivalent with , while, by (3.2), the second condition is equivalent with . Using these facts together with our bound on the integral in (3.60), we obtain:
(3.61) |
Here it should be recalled that ; see (3.58). Now note that for any appearing in the above integral, we have ; hence if we set then , and so, using Lemma 3.10 and then substituting , we obtain (since is a fixed finite subset of ):
(3.62) |
where we use the notation defined in (3.5).
Now we have
and it follows from the definition in (3.58) that for any , and for which holds, we have with , where is a certain absolute constant. Furthermore, with . Hence, using Lemma 3.10,
Here the integrand is independent of and . Hence, using the fact that is a fixed finite set, and substituting , we conclude:
(3.63) |
Hence we have proved the error bound in (3.7) (with ). As we have already noted, because of and Proposition 3.8, the bound (3.63) completes the proof of Theorem 4.
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 in Theorem 2, when we further assume that is such that is an interval for every . Recall here that for any , is the projection of on the -axis.
Theorem 5.
Remark 4.2.
For each , the partition and the constants are all computable from our analysis; see Section 4.4 for three explicit examples. We stress that the intervals are allowed to be open, closed, or half-open, and may be degenerate, i.e. of the form for some .
4.1. Relations to the polar set of
Let be as above. The integral appearing in (4.1) is closely related to the polar set of , which is defined by (1.21). Note that is closed and convex; furthermore, since is bounded and contains in its interior, is also bounded and contains in its interior. Moreover, if is symmetric about the origin, then so is ; and if is a convex polygon, then also is a convex polygon.
The following lemma shows that can be parameterized in polar coordinates in terms of the function .
Lemma 4.1.
Let be as in Theorem 5. Then is parameterized in polar coordinates by
Proof.
Note that for any ,
Thus
It then follows from the definition of polar set and the relation , that the set
equals the interval . The lemma then follows by a substitution . ∎
Remark 4.3.
As a direct consequence of Lemma 4.1 and the relation
(4.4) |
(which in turn follows from the symmetry ), we have
(4.5) |
For each , let be the unique index such that . In the special case when for every (this happens if 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) |
with . More generally, the sum of integrals appearing in (4.1) can be understood as a weighted area of .
The remainder of this section is devoted to proving Theorem 5.
4.2. Some preliminary computations
We start our proof with the formula for given in Theorem 4 which states that with
(4.7) |
It thus suffices to compute the above double integral which we denote by . It is immediate from the definition (3.3) that for all . Thus, defining for and applying (3.10), we have:
(4.8) |
We wish to compute the innermost integral in the last expression.
4.3. Extremal points
Fix an integral ideal . As mentioned above, we are interested in computing the integral
(4.9) |
for any . For this we introduce a notion which captures the possible values of the function . For any and , let if and if . Note that for any , if and only if . We say is -extremal with respect to if
where
We denote by the set of all -extremal points with respect to . Below we will simply call elements in -extremal when there is no ambiguity. Since is a lattice in , is always nonempty. Indeed, is always infinite, as it is invariant under a certain multiplication map as shown in the following lemma.
Lemma 4.2.
The set is invariant under the map defined by if and if .
Proof.
Suppose that is -extremal, i.e. with the rectangle given as above. Let and note that is invariant under the right multiplication action of . The lemma then follows by noting that if and if . ∎
Remark 4.10.
The set is invariant under the map even if . This is because we always have , since .
Note that is -extremal if and only if
(4.11) | for any with . |
In particular, if are two -extremal points, then . Moreover, one can check from (4.11) that takes values only on -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 , let be two consecutive elements in . Then for any .
Remark 4.12.
By discreteness of one sees that for any ; thus we can take two consecutive elements from as in the above lemma.
Proof of Lemma 4.3.
Take any and let . In particular, is -extremal and , and the last condition is equivalent to . Moreover, implies that . Thus . If , then since are two consecutive -extremal points, we must have and therefore , which contradicts the fact that . Hence as desired. ∎
We can now compute the integral in terms of the set .
Lemma 4.4.
For any and , if is the complete list of points in , then
where
Proof.
Remark 4.13.
If are two consecutive -extremal points, then so are for any integer . Using the equality one easily sees that both and are independent of the choice of .
Next, we study how varies in . Our goal is to show there are only finitely many possibilities for (and hence also for the coefficients and ). For this we introduce another related notion. We say is positively (resp. negatively) extremal with respect to if (resp. ) and (resp. ). We denote by and the set of positively and negatively extremal points in . Recall that is the map if and the map if . It is clear that are invariant under , and from the discreteness of we see that . Note that a necessary condition for to be -extremal for some is that . Indeed this is also a sufficient condition in the sense that for any there exists some such that ; see Lemma 4.6 below.
For any , set so that . Note that (resp. ) is decreasing (resp. increasing) in , and there is a constant such that for all and for all , because of Lemma 3.6 and the fact that
To understand how varies in we study how varies in which is easier as it is monotone. We first study this property for a single element in . The following lemma is immediate from our definitions.
Lemma 4.5.
For any let
Then is -extremal if and only if . Similarly, for any define
Then is -extremal if and only if .
Remark 4.14.
For each , when attains the critical value , fails to be -extremal. This fact, together with the finiteness of and the fact that (resp. ) is decreasing (resp. increasing) in implies that (resp. ) is right (resp. left) continuous in the sense that for any , there exists some such that for any (resp. ). Note also that
From this formula we see that the attaining the above infimum (resp. supremum) is an element in (resp. ). This gives very strict restrictions on the possible values of these ’s. In particular, can not be since otherwise we would have , contradicting . Moreover, the set of values is finite, since and for every .
Lemma 4.6.
There exist finitely many constants such that and
Similarly, there exist such that and
Proof.
We noted in Remark 4.14 that the set is finite; let us order its elements as . Then if , it follows from Lemma 4.5 that every is -extremal, i.e. ; similarly, if then . Finally if for some , then for every the condition is equivalent with ; hence by Lemma 4.5, . We have thus proved the part of the lemma which concerns ; the proof of the part concerning 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 with . Explicitly, and .
As a corollary we have the following.
Corollary 4.7.
There exist a finite partition of into intervals, and pairwise distinct subsets of , such that for any .
(The intervals 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 . The pairwise disjointness of follows again from the relation and also the fact that and are decreasing and increasing in respectively. ∎
Remark 4.16.
We see from the above proof that if , then is an interior point of one of the intervals . In particular, since , one of the intervals contains as an interior point.
We can now give the proof of Theorem 5.
Proof.
For each (), let be the partition as in Corollary 4.7 applied with , and define and for as in Lemma 4.4. For each , it follows from Corollary 4.7 that both and are constant as varies over ; thus we may define and for any . Then by (4.8) and Lemma 4.4 we have
where . Plugging this formula into (4.7) and applying the relation we get the desired formula for . ∎
4.4. Examples; proofs of Propositions 1.1–1.3
In this section we work out the formula (4.1) explicitly in the three cases , and , thus proving Propositions 1.1–1.3. Note that in all cases, is of class number one; thus and , and we set . For details of the following computations, see also the program [21, explicit_aP.mpl].
First we treat the case when . Then , and one verifies that
with the associated values given by
Hence
and
Using also the relation , we get the corresponding partition where with and . The corresponding constants and are
Plugging these values into (4.1) and using also and (see Remark 2.14), we get
(4.17) |
Moreover, because of (4.4) and , we have the general symmetry relation that for any bounded measurable set ,
(4.18) |
Using (4.18) and the fact that for , we see that (4.17) is equivalent to the formula (1.24) in Proposition 1.1. Finally, if then the sum inside the parenthesis in (1.24) equals by (4.5), and hence (1.25) holds. This completes the proof of Proposition 1.1.
Next, we consider the case when . Then , and one verifies that
with the associated values given by
Hence
and
The corresponding partition of is where with , , and . The corresponding constants and are
and
Plugging these formulas into (4.1) and using also and (see Remark 2.14), we get
(4.19) |
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 . Then , and one verifies that
with the associated values given by
Hence
The corresponding partition of is where with , and . The corresponding constants and are
Plugging these formulas into (4.1) and using also and (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 5.1–5.3 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 in Section 1.5. Let (an 8th root of unity), and let be the automorphism of the cyclotomic field given by . From now on we will use the standard identification of with , ; in particular we consider the ring as a subset of . Set
(5.1) |
this is the set of admissible “generic” translates of . Next, for any , set
(5.2) |
Then for every , is the vertex set of an Ammann-Beenker tiling (see [7] and [5, Ch. 7.3; Ex. 7.8]); and we note that . In fact every Ammann-Beenker tiling (appropriately scaled and rotated, and translated so that is a vertex), has vertex set either equal to for some , 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 . (In the limit case, it follows that there exists some such that the vertex set agrees with “up to density zero”, viz., the symmetric difference set has asymptotic density zero in . This implies that and 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 .)
In order to rewrite (5.2) as in (1.23), let ; then . Also set and . Using , one verifies that . Note that for any , we have under our identification of with , and in particular, as a subset of . Furthermore, using , we have for any :
and this point lies in if and only if . It follows that , i.e. the formula in (1.23) holds. (In the case , the formula (1.23) was noted in [20, Sec. 3.2].)
5.2. Gähler’s shield tilings
Recall the definition of in Section 1.5. Let (a 12th root of unity), and let be the automorphism of the cyclotomic field given by . As before, we identify with in the standard way. For any , set
(5.3) |
Then for every satisfying , 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 ; then . Also set and . Using one verifies that . Note that for any , we have under our identification of with , and in particular, as a subset of . Furthermore, using , we have for any :
and this point lies in if and only if . It follows that , i.e. the formula in (1.27) holds.
5.3. Tübingen triangle tilings
Recall the definition of in Section 1.5. Let (a 5th root of unity), and let be the automorphism of the cyclotomic field given by . As in Section 5.1, we identify with in the standard way. For any , set
(5.4) |
Then for every satisfying , 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 ; then , where we recall that . Also set and . Using one verifies that . Note that for any , we have under our identification of with , and in particular, as a subset of . Furthermore, using , we have for any :
and this point lies in if and only if . It follows that , 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, , and . In Table 1 below, for a few examples of points , we present conjectural approximate values of , where is the limiting gap distribution function associated to , and, for comparison, the exact values of . Tables 2 and 3 present similar types of data for and .
Experimental approximate values of for
0.24638… | 0.2810 | 0.2628 | 0.2525 | 0.2497 | 0.250 | 0.246 | 0.23 | |
---|---|---|---|---|---|---|---|---|
0.21809… | 0.2460 | 0.2317 | 0.2236 | 0.2212 | 0.220 | 0.219 | 0.21 | |
0.20416… | 0.2286 | 0.2160 | 0.2090 | 0.2067 | 0.206 | 0.205 | 0.19 | |
0.17444… | 0.2211 | 0.2048 | 0.1815 | 0.1767 | 0.177 | 0.178 | 0.18 | |
Error | 0.0005 | 0.0006 | 0.0011 | 0.0023 | 0.005 | 0.015 | 0.05 |
Experimental approximate values of for
0.17790… | 0.1974 | 0.1873 | 0.1817 | 0.180 | 0.180 | 0.176 | 0.19 | |
---|---|---|---|---|---|---|---|---|
0.21374… | 0.2460 | 0.2289 | 0.2199 | 0.216 | 0.217 | 0.215 | 0.23 | |
0.19210… | 0.2143 | 0.2027 | 0.1964 | 0.194 | 0.195 | 0.190 | 0.20 | |
0.20870… | 0.2341 | 0.2208 | 0.2136 | 0.211 | 0.211 | 0.205 | 0.23 | |
Error | 0.0004 | 0.0007 | 0.001 | 0.004 | 0.006 | 0.019 | 0.05 |
Experimental approximate values of for
0.26135… | 0.2919 | 0.2762 | 0.2673 | 0.2643 | 0.263 | 0.265 | 0.26 | |
---|---|---|---|---|---|---|---|---|
0.28875… | 0.3283 | 0.3080 | 0.2965 | 0.2927 | 0.291 | 0.294 | 0.30 | |
0.29103… | 0.3329 | 0.3116 | 0.2995 | 0.2954 | 0.294 | 0.297 | 0.30 | |
0.18732… | 0.2195 | 0.1977 | 0.1907 | 0.1892 | 0.188 | 0.190 | 0.19 | |
Error | 0.0004 | 0.0004 | 0.0007 | 0.0017 | 0.004 | 0.014 | 0.03 |
The approximations of in these tables were obtained by collecting all the points of (where or or ) in fairly large discs and evaluating the ratio in the left hand side of Theorem 1; by repeating this for several large -values we also obtained conjectural error bounds; see below for a more detailed discussion. Recall that by Theorem 2, in fact with a rate . We certainly expect that the absolute difference typically decreases as runs through the values , and it should be noted that the data in Tables 1–3 is consistent with this hypothesis, when considering also the conjectural error bounds displayed in the last line of the tables.
Precise numerical values of such as those presented in Tables 1–3 (but also to much higher precision), are quite easy to compute using Propositions 1.1–1.3. To describe how to do this, first note that in the case , it follows from (1.23), the last statement in Theorem 1 and Remark 3.12, that this is the same as for ; similarly, for , is the same as for (see (1.27)), and for , is the same as for (see (1.30)). Now, somewhat more generally, assume that where is the open regular -gon with vertices at the points (), for some integer and some , and . (We obtain by taking and , by taking and , and by taking and .) For such a window , the function in (1.16) is given by
where for each , and are defined by . Using also , it is then easy to verify that all the sets in Propositions 1.1–1.3 are finite unions of intervals. For given and , 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 in Tables 1–3 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) |
Then , and so for sufficiently large, is a good approximation of . Note that computing involves collecting all the points of in the disc . When carrying out the computations, we noted that for fixed , the values typically oscillate as a function of ; hence if is a finite set of several sufficiently large -values of the same order of magnitude, it seems reasonable to consider the difference
as an indication of the size of the error for any . To lessen the probability of getting a small difference by coincidence, we took the experimental error bound given in the last line of Tables 1–3 to be the maximum of over values of roughly equispaced in the interval , and over all four choices . Because of the factor , it is not surprising that these experimental error bounds get worse as increases.
In order to produce Table 1 (i.e. the case of Ammann-Beenker tilings) we used
(5.6) |
and for each we chose as approximate value of the mid-point . We remark that computing the functions in this case involve collecting a bit over points from each point set . For example, for we found the number of points in to be (this may be compared with ).
Similarly, for Table 2 (Gähler’s shield tilings) we used
(5.7) |
and computing the functions involved collecting a bit over points from each point set . For example, for and we found the number of points in to be (which may be compared with ).
Similarly, for Table 3 (Tübingen triangle tilings) we used
(5.8) |
and computing the functions involved collecting a bit over points from each point set . For example, for and we found the number of points in to be (which may be compared with ).
Let us next discuss the graphs in Figure 1. These were generated by distributing the normalized gaps , – discarding the vanishing gaps – in bins of width , 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) |
where while for all .333The fact that we use in place of corresponds exactly to the fact that we discard those gaps which vanish; without this modification the graph would have a huge peak for ; see the discussion below. We point out that the graphs of , as they are scaled in Figure 1, look identical to the naked eye for each , i.e. the picture looks the same if we plot all four graphs in the same coordinate system; in fact the maximum over of the difference was in each of the six cases found to be less than , and the average of that same difference was less than .
These findings support the conjecture that, in the cases considered, the derivative 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 is
Furthermore, as , tends to , the relative density of visible points in , a quantity which was defined and proved to exist for general regular cut-and-project sets in [27]. (The proof of is immediate by comparing with the definition of in [27].) The values which we obtained for for varying in consistently agreed to within ca ; hence we expect that these values approximate to within an absolute error of similar order of magnitude. In particular for with and , we obtained .444Similar numerics, but for considerably smaller -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) |
For the other panels in Figure 1 we obtained the following values of (viz., area under the curve): II. . III. . IV. . V. . VI. .
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 “” 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 . In fact, in our notation, the plots in [4, Fig. 9] and [20, Fig. 2] are experimental graphs of the function , with 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 -values than those used for our Figure 1.
Index of notation
leading coefficient of the limiting gap distribution function of a planar point set | 2 | |
matrix element with | 2.3 | |
closure of with a lattice in and the natural projection to the internal space | 1.2 | |
identity component of | 1.2 | |
the set | 3.55 | |
the function with an integral ideal | ||
and bounded measurable with non-empty interior | 3.3 | |
the function with as above | 3.53 | |
if and if | 4.3 | |
open disc in with center and radius | 3.2 | |
the set | 3.57 | |
the vector such that is the -cusp of | 2.5 | |
ideal class group of | 2.1 | |
normalizing factor of a Haar measure of which equals | 2.13 | |
asymptotic density of a planar point set | 1.2 | |
discriminant of | 1.11, 2.4 | |
directions of points of length bounded by in a planar point set | 1.1 | |
set of positively or negatively extremal points with respect to an integral ideal | 4.3 | |
set of -extremal points with respect to an integral ideal | 4.3 | |
set of positive or negative -extremal points with respect to | 4.3 | |
limiting distribution function of gaps of directions of a planar cut-and-project set | 1.6 | |
the function with and Lebesgue measurable | 3.5 | |
either or depending on whether is empty or not | 3.6 | |
a Siegel fundamental domain for | 2.19 | |
a fixed fundamental domain for | 2.16 | |
the set | 2.17 | |
a fixed fundamental domain for | 2.4 | |
integral of the limiting gap distribution function | 1.12 | |
contribution of the -th cusp of to the function for sufficiently large | 3.36 | |
error term of | 3.41 | |
main term of | 3.40 | |
the group | 3.3 | |
isotropy group of the -th cusp of | 2.7 | |
the group | 2.2 | |
the group | 2.2 | |
the group with | 3.3 | |
ideal generated by with such that | 2.1 | |
the natural embedding either from to or from to | 2.1, 2.6 | |
the set | 3.5 | |
open interval with | 4.2 | |
the set | 3.58 | |
the set | 3.58 | |
real quadratic field with square-free | 1.3, 2.1 | |
the set | 2.2 | |
-th cusp of | 2.1 | |
matrix element with | 2.3 | |
number of cusps of | 2.1 | |
relative density of visible points in a planar point set | 1.9, 5.4 | |
the probability -invariant measure on | 2.12 | |
the set | 2.24 | |
the set | 3.39 | |
Minkowski embedding of in | 1.10 | |
projection of on the -axis with | 3.3 | |
projection of on the -axis | 3.3 | |
fundamental unit of | 1.3 | |
Lebesgue measure on | 3.1 | |
standard norm on defined by | 2.1 | |
Nr | absolute norm of ideals | 2.5 |
matrix element with | 2.3 | |
ring of integers of | 1.3 | |
unit group of | 1.3 | |
critical value where the -extremality of changes | 4.5 | |
parameter in the expression | 1.16 | |
the Ammann-Beenker tiling corresponding to the translate | 1.23, 5.2 | |
the Gähler’s shield tiling corresponding to the translate | 1.27, 5.3 | |
cut-and-project set associated to a window set and a lattice | 1.4 | |
natural projection from to the physical space | 1.2 | |
natural projection from to the internal space | 1.2 | |
the plane | 2.23 | |
a rectangle depending on and with an integral ideal | 4.3 | |
the set | 3.1 | |
parameter in the expression | 1.16 | |
the unit circle | 1.1 | |
the -th interval in the partition of corresponding to the -th cusp of | 5 | |
the set | 5 | |
Hilbert modular group | 2.2 | |
the unique nontrivial automorphism of the real quadratic field | 2.1 | |
an absolute positive number depending only on | 2.4 | |
a generator of which is if and if | 2.1 | |
open triangle with vertices at and | 3.3 | |
the set with such that | 3.3 | |
the set | 3.3 | |
window set of the cut-and-project set | 1.2 | |
an open regular octagon with particular size and position | 2 | |
an open regular dodecagon with particular size and position | 1.1 | |
the set | 1.23 | |
the set | 1.29 | |
polar set of | 4.1 | |
scaling matrix with respect to the -th cusp of | 2.9 | |
the set | 3.7 | |
the set | 3.38 | |
the set | 3.4 | |
the set | 3.4 | |
Dedekind zeta function corresponding to | 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 -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 -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 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/astrombe/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.