math0610408/pin.tex
1: \documentclass[11pt]{amsart}
2: \usepackage{amsmath}
3: \usepackage{amssymb,bbm}
4: \usepackage{epsfig}
5: \usepackage{a4wide}
6: \usepackage{pifont}
7: 
8: \newcommand{\N}{\ensuremath{\mathbb{N}^{}_0}}
9: \newcommand{\Z}{\ensuremath{\mathbb{Z}}}
10: \newcommand{\R}{\ensuremath{\mathbb{R}}}
11: \newcommand{\Q}{\ensuremath{\mathbb{Q}}}
12: \newcommand{\C}{\ensuremath{\mathbb{C}}}
13: \newcommand{\Sb}{\ensuremath{\mathbb{S}}}
14: \newcommand{\X}{\ensuremath{\mathbb{X}}}
15: \newcommand{\Sc}{\ensuremath{{\mathcal S}}}
16: \newcommand{\D}{\ensuremath{{\mathcal D}}}
17: \newcommand{\F}{\ensuremath{{\mathcal F}}}
18: \newcommand{\gL}{\varLambda}
19: \newcommand{\gG}{\varGamma}
20: \newcommand{\gT}{\varTheta}
21: \newcommand{\smc}{\scriptscriptstyle\square}
22: 
23: \newcommand{\todo}[1]{[\textnormal{\ding{46}\ding{55}} #1]}
24: 
25: \DeclareMathOperator{\dd}{d\!}
26: \DeclareMathOperator{\dens}{dens}
27: \DeclareMathOperator{\vol}{vol}
28: 
29: \newtheorem{thm}{Theorem}
30: \newtheorem{prop}[thm]{Proposition}
31: \newtheorem{lem}[thm]{Lemma}
32: \newtheorem{cor}[thm]{Corollary}
33: 
34: \newtheorem{claim}{Claim}
35: \newtheorem{conj}{Observation}
36: 
37: \theoremstyle{definition}
38: 
39: \renewcommand{\baselinestretch}{1.06}
40: 
41: \begin{document}
42: 
43: \title[A radial analogue of Poisson's summation formula]{A radial
44:   analogue of Poisson's summation formula\\[2mm] with applications to
45:   powder diffraction\\[2mm] and pinwheel patterns}  
46: 
47: \author{Michael Baake}
48: \address{Fakult\"at f\"ur Mathematik, Universit\"at Bielefeld, 
49: Postfach 100131, 33501 Bielefeld,  Germany}
50: \email{\texttt{mbaake@math.uni-bielefeld.de}, 
51: \texttt{dirk.frettloeh@math.uni-bielefeld.de}}
52: \urladdr{\texttt{http://www.math.uni-bielefeld.de/baake},
53: \texttt{http://www.math.uni-bielefeld.de/baake/frettloe/}}
54: 
55: \author{Dirk Frettl\"oh}
56: % \address{Fakult\"at f\"ur Mathematik, Universit\"at
57: %   Bielefeld, 33501 Bielefeld, Germany}
58: % \email{\texttt{dirk.frettloeh@math.uni-bielefeld.de}}
59: % \urladdr{\texttt{http://www.math.uni-bielefeld.de/baake/frettloe/}}
60: 
61: \author{Uwe Grimm}
62: \address{
63: Department of Mathematics, The Open University, Walton Hall,
64: Milton Keynes MK7 6AA, UK}
65: \email{u.g.grimm@open.ac.uk}
66: \urladdr{http://mcs.open.ac.uk/ugg2/}
67: 
68: 
69: \begin{abstract} 
70: Diffraction images with continuous rotation symmetry arise from
71: amorphous systems, but also from regular crystals when investigated
72: by powder diffraction. On the theoretical side, pinwheel patterns and
73: their higher dimensional generalisations display such symmetries as
74: well, in spite of being perfectly ordered. We present first steps and
75: results towards a general frame to investigate such systems, with
76: emphasis on statistical properties that are helpful to understand and
77: compare the diffraction images. An alternative substitution rule for
78: the pinwheel tiling, with two different prototiles, permits the
79: derivation of several combinatorial and spectral properties of this
80: still somewhat enigmatic example. These results are compared with 
81: properties of the square lattice and its powder diffraction.
82: 
83: \end{abstract} 
84: 
85: \maketitle
86: 
87: \section{Introduction} \label{seq:intro}
88: 
89: Since the discovery of quasicrystals some 20 years ago, mathematicians
90: and physicists have gained a reasonable understanding of aperiodically
91: ordered systems, in particular of those obtained from the projection
92: method. Such sets are called cut and project sets, or model sets
93: \cite{moo}. These are Delone sets of finite local complexity with
94: respect to translations, which also means that any finite patch occurs
95: in finitely many orientations only.
96: 
97: Much less is known about aperiodically ordered systems with local
98: patches occurring in infinitely many orientations, such as the
99: pinwheel tiling of the plane or its three-dimensional counterpart,
100: see \cite{rad} and references therein. Arguably, these are closer to
101: amorphous systems, but still perfectly ordered. To our knowledge, no
102: mathematically satisfactory frame for the analysis of radially
103: symmetric systems and the comparison of their spectral properties has
104: been developed so far. It is the aim of this contribution to show
105: first steps in this direction, by combining results and methods from
106: discrete geometry with the more recent approach of mathematical
107: diffraction theory.
108: 
109: Our guiding examples are the square lattice and the pinwheel tiling of
110: the plane. The diffraction of the square lattice is well understood,
111: and it is not difficult to get some insight into its powder
112: diffraction. The latter emerges from the presence of many grains in
113: random mutual orientations, and thus requires a setting with circular
114: symmetry.
115: 
116: Circular symmetry is also a fundamental property of the pinwheel
117: tiling, resp.\ its compact hull. This tiling has recently been
118: reinvestigated from the autocorrelation and diffraction point of view
119: \cite{mps}. However, hardly any explicit calculation exists in the
120: literature, the reason being the enigmatic nature of the substitution
121: generated pinwheel tiling.  New insight is gained by means of an
122: alternative construction on the basis of a substitution rule with two
123: distinct prototiles. This gives access to quantities such as
124: frequencies (hence also to the frequency module), distance sets and
125: the ring structure of the diffraction measure.
126: 
127: In Section~\ref{sec:rpsf}, we derive a radial analogue of Poisson's
128: summation formula for tempered distributions, which is needed in the
129: following analysis. In the last section, we outline the general
130: structure by presenting a number of results. They are clearly
131: distinguished according to their present status into theorems (with
132: proofs or references), claims (with a sketch of the idea and a
133: reference to future work) and observations (based on numerical or
134: preliminary evidence). Since our alternative substitution will no
135: doubt enable other developments as well, we hope that further progress
136: is stimulated by the results presented in this paper.
137: 
138: \subsection{Notation and preliminaries}
139: 
140: A rotation through $\alpha$ about the origin is denoted by
141: $R^{}_{\alpha}$, and $B^{}_r(x)$ is the closed ball of radius $r$
142: centred in $x$.  Whenever we speak of an \emph{absolute frequency}
143: (e.g., of a point set), we mean the average number per unit volume,
144: whereas \emph{relative frequency} of a subset of objects is used with
145: respect to the entire number of objects. The absolute frequency of the
146: points in a point set $X$ (if it exists) is also called the
147: \emph{density} of $X$, denoted by $\dens(X)$. The set of non-negative
148: integers is called $\N$, and the unit circle is
149: $\Sb^{1}=\{z\in\C:|z|=1\}$.  If $M$ is a locally finite point set, the
150: corresponding \emph{Dirac comb} $\delta^{}_M := \sum_{x \in M}
151: \delta^{}_x$, where $\delta^{}_x$ is the normalised point (or Dirac)
152: measure in $x$, is a well-defined measure. The Fourier transform of
153: $g$ is denoted by $\widehat{g}$. Whenever we use Fourier transform for
154: measures below, we are working in the framework of tempered
155: distributions \cite{sch}.
156: 
157: \section{A radial analogue of Poisson's summation formula}
158: \label{sec:rpsf}
159: 
160: The diffraction pattern of an ideal crystal, supported on a point
161: lattice $\gG \subset \R^d$, can be obtained by the Poisson
162: summation formula 
163: (PSF) for lattice measures or Dirac combs \cite{sch,cor1,cor2} 
164: \begin{equation} \label{eq:psf}
165: \widehat{\delta}^{}_{\gG} = \dens(\gG) \cdot
166: \delta^{}_{\gG^{\ast}},
167: \end{equation}
168: where $\gG^{\ast} := \{ x \in \R^d : x \!\cdot\! y \in \Z$ for all $y
169: \in \gG \}$ is the dual lattice. The distribution-valued (or
170: measure-valued) version follows from the classical PSF, compare
171: \cite{cor1,ik}, via applying it to a (compactly supported) Schwartz
172: function. It is often used to derive the diffraction measure
173: $\widehat{\gamma}^{}_{\omega}$ of a lattice periodic measure $\omega =
174: \varrho \ast \delta^{}_{\gG}$ (with $\varrho$ a finite measure). One
175: obtains the autocorrelation measure
176: \begin{equation}  \label{eq:autocor} 
177: \gamma^{}_{\omega} = \dens(\gG) \cdot (\varrho \ast
178: \widetilde{\varrho}\, ) \ast \delta^{}_{\gG} ,
179: \end{equation}
180: where $\widetilde{\varrho}(g) := \overline{\varrho(\widetilde{g})}$
181: with complex conjugation $\overline{\vphantom{g}.}$ and
182: $\widetilde{g}(x)=\overline{g(-x)}$, and the diffraction measure
183: \[
184: \widehat{\gamma}^{}_{\omega} = ( \dens(\gG))^2 \cdot |
185: \widehat{\varrho} \, |^2 \cdot \delta^{}_{\gG^{\ast}} ,
186: \] 
187: see \cite{hof,baa} for details. Our interest is to extend this
188: approach to situations with circular (or spherical) symmetry. 
189:  
190: Let us first consider the square lattice $\Z^2$, and recall that its
191: circular shells have radii precisely in the set 
192: \begin{equation}\label{defrad}
193:  \D^{}_{\smc} := \{ r \ge 0 : r^2 = m^2
194: + n^2 \; \mbox{with $m,n \in \Z$} \} =
195: \{0,1,\sqrt{2},2,\sqrt{5},2 \sqrt{2}, 3, \ldots \} . 
196: \end{equation}
197: This is the set of non-negative numbers whose squares are integers
198: that contain primes $p \equiv 3 \mod 4$ only to even
199: powers. Moreover, on a shell of radius 
200: $r \in \D^{}_{\smc}$, one 
201: finds finitely many lattice points, their number being given by 
202: \begin{equation} \label{etaq}
203:  \eta^{}_{\smc} (r) =   
204: \begin{cases} 
205: 1, & r=0, \\
206: 4 a(r^2), & r \in \D^{}_{\smc}\setminus\{0\}. 
207: \end{cases}
208: \end{equation}
209: Here, $a(n)$ is the number of ideals of norm $n$ in $\Z[i]$, the ring
210: of Gaussian integers. This is a multiplicative arithmetic function,
211: thus specified completely by its values at prime powers (see
212: \cite{bg,haw}).  They are given by
213: \[ 
214: a(p^{\ell}) = \begin{cases} 1, & \text{if $p=2$}, \\
215: \ell+1, &  \text{if $p \equiv 1\bmod 4$}, \\ 
216: 0, &  \text{if $p\equiv 3 \bmod 4$ and $\ell$ odd}, \\ 
217: 1, &  \text{if $p\equiv 3 \bmod 4$ and $\ell$ even}. \end{cases}
218: \] 
219: As $\Z^2$ is self-dual as a lattice, with $\dens(\Z^2)=1$, the PSF
220: \eqref{eq:psf} simplifies to $\widehat{\delta}^{}_{\Z^2} =
221: \delta^{}_{\Z^2}$.
222: 
223: Choose an irrational number $\alpha \in (0,1)$. By Weyl's lemma
224: \cite{kn}, the sequence $(n \alpha \bmod 1)_{n \ge 1}$ is uniformly
225: distributed in $(0,1)$. Consider the sequence 
226: $(\omega^{}_N)^{}_{N \ge  1}$ of measures defined by
227: \begin{equation} \label{eq:sumrg} 
228:  \omega^{}_N = \frac{1}{N} \sum_{n=1}^N \delta^{}_{R^n \Z^2}, 
229: \end{equation}
230: where $R=R^{}_{2 \pi \alpha}$ is the rotation through $2 \pi \alpha$.
231: If $|x|=r>0$, the sequence $(R^n x)^{}_{n \ge 1}$ is uniformly
232: distributed on $\partial B^{}_r(0)$, again by Weyl's lemma. Observe
233: that all lattices $R^n \Z^2$ share the same set of possible shell
234: radii, namely $\D^{}_{\smc}$ of \eqref{defrad}.  This implies that,
235: given an arbitrary compactly supported continuous function $\varphi$,
236: one has
237: \[ 
238: \lim_{N \to \infty} \frac{1}{N} \sum_{n=1}^N \delta^{}_{R^n \Z^2}
239: (\varphi) = \sum_{r \in \D^{}_{\smc}}
240: \eta^{}_{\smc} (r) \mu^{}_r (\varphi),
241: \]  
242: where the measure $\mu^{}_r$ is the normalised uniform distribution on
243: $\partial B^{}_r (0) = \{ x \in \R^2 : |x| =r \}$, with
244: $\mu^{}_0=\delta^{}_0$. This establishes the following result.
245: 
246: \begin{prop}
247: The sequence $(\omega^{}_N)^{}_{N \ge 1}$ of \eqref{eq:sumrg}
248: converges in the vague topology, and
249: \[ 
250: \omega:= \lim_{N \to \infty} \omega^{}_N  = \sum_{r \in
251:     \D^{}_{\smc}} 
252:    \eta^{}_{\smc} (r) \, \mu^{}_r, 
253: \]
254: with shelling numbers $\eta^{}_{\smc}(r)$ and
255: probability measures $\mu^{}_r$ as introduced above.\qed
256: \end{prop}
257: 
258: 
259: It is obvious that the limit is, at the same time, also a limit of
260: tempered distributions, i.e., a limit in $\Sc'(\R^2)$. As the Fourier
261: transform is continuous on $\Sc'(\R^2)$, one has
262: \[ 
263: \widehat{\omega} = \Bigl( \lim_{N \to \infty} \omega^{}_N
264: \Bigr)^{\widehat{}}  = \lim_{N \to \infty} \widehat{\omega}^{}_N. 
265: \] 
266: Employing the ordinary PSF \eqref{eq:psf}, one finds
267: \[  
268: \widehat{\omega}^{}_N = \frac{1}{N} \sum_{n=1}^N
269: \widehat{\delta}^{}_{R^n \Z^2} = \frac{1}{N} \sum_{n=1}^N
270: \delta^{}_{(R^n \Z^2)^\ast} = \frac{1}{N} \sum_{n=1}^N \delta^{}_{R^n
271: \Z^2} = \omega^{}_N 
272: \] 
273: where we used that $(R \gG)^{\ast} = R \gG^{\ast}$ for an
274: isometry $R$, and $\dens(R^n \Z^2)=\dens(\Z^2)=1$.  Combining the last
275: two equations, one sees that
276: \[ 
277: \widehat{\omega} = \lim_{N \to \infty} \widehat{\omega}^{}_N = \lim_{N
278:   \to \infty} \omega^{}_N = \omega. 
279: \]
280: 
281: Using polar coordinates, the Fourier transform of $\mu^{}_r$ --- which
282: is an analytic function because $\mu^{}_r$ has compact support --- can
283: be expressed by a Bessel function of the first kind via the following
284: calculation,
285: \begin{equation}\label{eq:foubessel}
286: \begin{split} 
287: \widehat{\mu}^{}_r(k) & = \int_{\R^2} e^{-2\pi i k\cdot x} \dd \mu^{}_r(x) 
288:  = \int_{0}^{\infty} \frac{1}{2\pi} \int_{0}^{2\pi} e^{-2\pi i
289:    |k| \varrho \cos{\varphi}} \dd \varphi \dd \delta^{}_r(\varrho)  \\[2mm]
290: & =  \int_{0}^{\infty} J^{}_0(2\pi|k|\varrho) \dd \delta^{}_r(\varrho) =  
291: J^{}_0 ( 2\pi |k| r),  
292: \end{split}
293: \end{equation}
294: where $J^{}_0(z) = \sum_{\ell=0}^{\infty} \frac{(-1)^\ell}{(\ell!)^2}
295: (\frac{z}{2})^{2\ell}$. This yields the identity
296: \begin{equation} \label{eq:sumbess}
297: \widehat{\omega} = \sum_{ r \in \D^{}_{\smc}}
298: \eta^{}_{\smc}(r) J^{}_0(2 \pi |k|r),  
299: \end{equation}
300: to be understood in the distribution sense. It has the
301: following consequence. 
302: 
303: \begin{cor} \label{corsumbess}
304: With $\eta^{}_{\smc}(r)$ and $\mu^{}_r$ as
305: introduced above, one has
306: \[  
307:     \sum_{r \in \D^{}_{\smc}} \eta^{}_{\smc} (r) \, \mu^{}_r =
308: \Bigl(  \sum_{r \in \D^{}_{\smc}}
309: \eta^{}_{\smc} (r) \, \mu^{}_r \Bigr)^{\widehat{}} =
310: \sum_{r \in \D^{}_{\smc}}
311: \eta^{}_{\smc} (r) J^{}_0(2 \pi |k| r ),
312: \] 
313: where the last expression is to be understood in the distribution
314: sense. \qed
315: \end{cor}
316: 
317: Identities of this type can be viewed as measure-valued
318: generalisations of classic Hardy-Landau-Voronoi summation formulae,
319: see \cite[Sec.\ 4.4]{ik} and references given there for
320: details. However, in view of the rather delicate convergence
321: properties, a direct verification via the PSF \eqref{eq:psf} seems a
322: simpler approach. 
323: 
324: Observe that one can alternatively reduce the problem to one dimension
325: and employ a Hankel transform, compare \cite[Sec.\ 4.4]{ik}. This
326: requires a separate treatment of the transformed measure at the origin
327: in $k$-space. As this looks slightly artificial from the point of view
328: of diffraction, we stick to ordinary Fourier transform.
329: 
330: The version for a general lattice $\gG \subset \R^d$ reads as
331: follows, where, in analogy to $\eta^{}_{\smc}(r)$, 
332: $\eta^{}_{\gG}(r)$ and $\eta^{}_{\gG^{\ast}}(r)$ denote
333: the number of points of $\gG$ and $\gG^{\ast}$ on centred
334: shells $\partial B^{}_r(0)$ of radius $r$. 
335: 
336: \begin{thm}[Radial PSF] \label{radpsf}
337: Let $\gG$ be a lattice of full rank in $\R^d$, with dual lattice
338: $\gG^{\ast}$. If the sets of radii for non-empty shells are
339: $\D^{}_{\gG}$ and $\D^{}_{\gG^{\ast}}$, with shelling
340: numbers $\eta^{}_{\gG}(r)$ and $\eta^{}_{\gG^{\ast}}(r)$,
341: the classical PSF \eqref{eq:psf} has the radial
342: analogue  
343: \begin{equation} \label{eq:radpsf}
344: \Bigl( \sum_{ r \in \D^{}_{\gG}}  \eta^{}_{\gG}(r) \, \mu^{}_r
345: \Bigr)^{\widehat{}} 
346: = \sum_{r \in \D^{}_{\gG} } \eta^{}_{\gG}(r)
347:   \, \widehat{\mu}^{}_r
348: = \dens(\gG) \sum_{r \in  \D^{}_{\gG^{\ast}} }
349: \eta^{}_{\gG^{\ast}}(r) \, \mu^{}_r ,
350: \end{equation}     
351: where $\mu^{}_r$ denotes the uniform probability measure on the sphere of
352: radius $r$ around the origin. 
353: \end{thm}
354: 
355: \begin{proof}
356: Select a sequence of isometries $(R^{}_n)^{}_{n \ge 0}, \, R^{}_n \in \mathrm{SO}(d)$,
357: such that $(R^{}_n x)^{}_{n \ge 0}$, for a fixed $x$ of length 1, is
358: uniformly distributed on the unit sphere ${\mathbb S}^d$. Consider
359: then the sequence of measures defined by 
360: \[ \omega^{}_N = \frac{1}{N} \sum_{n=1}^N \delta^{}_{R^{}_n \gG} . \]
361: The claim now follows from Weyl's lemma and the classical PSF
362: \eqref{eq:psf} in complete analogy to our previous planar example. 
363: \end{proof}
364: 
365: The formula for general $d$ can also be expressed in terms of Bessel
366: functions of the first kind. Here, by standard calculations with
367: spherical coordinates and integral representations of Bessel
368: functions, one obtains
369: \begin{equation} \label{ftmur}
370: \widehat{\mu^{}}_r(k) = \int_{\R^d} e^{-2 \pi i kx} \dd \mu^{}_r(x) =
371: \Gamma\bigl(\frac{d}{2}\bigr)\, \frac{ J^{}_{\frac{d}{2}-1} (2 \pi |k|r)}{(\pi
372:   |k|r)^{\frac{d}{2} -1}},  
373: \end{equation}
374: where $\Gamma(x)$ is the gamma function and $J^{}_{\nu}(z) = (
375: \frac{z}{2})^{\nu} \sum\limits_{\ell=0}^{\infty} 
376: \frac{(-1)^\ell}{\ell! \Gamma(\nu + \ell+1)} (\frac{z}{2})^{2\ell}$.
377: In particular, 
378: \[ 
379: \widehat{\mu}^{}_r(k) = \begin{cases} 
380: \cos(2\pi k r), & \text{if $d=1$}, \\[1mm]
381: J^{}_0(2 \pi |k| r), & \text{if $d=2$}, \\[1mm]
382: \frac{\sin(2 \pi |k| r)}{2 \pi |k| r}, & \text{if $d=3$}. 
383: \end{cases} 
384: \]
385: The analogue of Corollary~\ref{corsumbess} for $d=1$ and 
386: $\gG = \Z$ thus reads 
387: \[ \sum_{m \in \Z} \cos(2 \pi km) = \widehat{\delta}^{}_{\Z} =
388: \delta^{}_{\Z}, \]
389: which is just another form of the ordinary PSF in this case (as radial
390: averaging is trivial here). Figure~\ref{j123} shows \eqref{ftmur} for
391: $r=1$ and various values of $d$. 
392: 
393: \begin{figure} 
394: \epsfig{file=murd12345.eps, height=40mm}
395: \caption{\label{j123}
396: A plot of the radial structure of the function $\widehat{\mu}^{}_1(k)$ 
397: of \eqref{ftmur} for dimensions $d=1$ (light grey), $2$, $3$, $4$, and $5$ (black). }
398: \end{figure}
399:  
400: 
401: \section{A simplistic approach to powder diffraction} 
402: 
403: The intensity distribution in powder diffraction emerges from a
404: collection of grains in random and mutually uncorrelated
405: orientations. Its precise theoretical description is difficult,
406: compare \cite{war} and references therein.
407: 
408: Here, we look into a rather simplistic approach that nevertheless
409: captures the essence of the diffraction image. For simplicity, and for
410: comparison with related pinwheel patterns, we explain this for the
411: square lattice $\Z^2$. Instead of working with grains of finite size,
412: we consider the superposition of entire lattices, with appropriate
413: weights. Moreover, we make the restriction that there is a common
414: rotation centre for all lattices, which we choose to be the origin. As
415: a first step, let us take a look at $\Z^2 \cup R \Z^2$, where $R \in
416: \mathrm{SO}(2)$ is a generic rotation (by which we mean that it is
417: \emph{not} an element of the group of coincidence rotations
418: $\mathrm{SOC}(\Z^2) = \mathrm{SO}(2,\Q)$, see \cite{bg} for more on
419: this concept).
420: 
421: \begin{lem} \label{diff2gitter}
422: Let $R \in \mathrm{SO}(2)$ be a generic rotation, so that $\Z^2 \cap R\Z^2 =
423: \{0\}$. Then, the autocorrelation of\/ 
424: $\omega = \frac{1}{2} (\delta^{}_{\Z^2} + \delta^{}_{R \Z^2})$ is
425: \begin{equation} \label {eq:autocor2}
426: \gamma^{}_{\omega} = \frac{1}{4} \delta^{}_{\Z^2} + \frac{1}{4}
427: \delta^{}_{R\Z^2} + \frac{1}{2} \lambda, 
428: \end{equation}
429: where $\lambda$ is the Lebesgue measure in $\R^2$. 
430: The diffraction measure of $\omega$ is 
431: \begin{equation} \label {eq:autocor3}
432: \widehat{\gamma}^{}_{\omega} = \sum_{x \in \Z^2 \cup R\Z^2} I(x)
433: \delta^{}_x 
434: \end{equation}   
435: with $I(0)=1$ and $I(x)= \frac{1}{4}$ for all \ $0 \ne x \in \Z^2 \cup R
436: \Z^2$. 
437: \end{lem}
438: \begin{proof} 
439: The first two terms in \eqref{eq:autocor2} are the autocorrelations of
440: $\frac{1}{2}\Z^2$ and of $\frac{1}{2}R\Z^2$. The third term originates
441: from the cross-correlation between them, where we used the identity
442: \[ 
443: \lambda = \lim_{r \to \infty} \frac{1}{\pi r^2} \sum_{\substack{
444:     x \in \Z^2 \cap B^{}_r(0) \\ y \in R\Z^2 \cap B^{}_r(0)}}
445: \delta^{}_{x-y}, 
446: \] 
447: with the limit being taken in the vague topology. This identity can be
448: derived as follows.  It is easy to see that each square in the square
449: lattice is hit by $x-y$ exactly once, if $x$ is arbitrary but fixed
450: and $y$ runs through $R\Z^2$. (The exception is $y=0$, but we can
451: neglect it since it plays no role in the limit.)  Thus each square in
452: $\Z^2$ is hit with the same frequency in the limit. Since $R$ is a
453: generic rotation, the sequence $(x-y \bmod (1,1))$ is uniformly
454: distributed in the fundamental domain $[0,1)^2$ of $\Z^2$, if $x$ is
455: arbitrary but fixed and $y$ runs through $R \Z^2$. This is a
456: consequence of the uniform distribution of $(n\alpha \bmod 1)^{}_{n \ge
457: 1}$ in $(0,1)$ if $\alpha$ is irrational. Thus, the points in the sum
458: above are uniformly distributed in $\R^2$ in the limit, and the series
459: converges in the vague topology to the Lebesgue measure.
460: 
461: Now, Equation \eqref{eq:autocor3} follows immediately from
462: $\widehat{\lambda} = \delta^{}_0$ and the PSF \eqref{eq:psf}.  
463: \end{proof}
464: 
465: For completeness, let us briefly comment on the situation that
466: $R\in\mathrm{SOC}(\Z^2)$. In this case, $\gT:=\Z^2\cap R\Z^2$ is a
467: sublattice of $\Z^2$ of finite index, which is $1/\dens(\gT)$.  It is
468: possible to show that $\omega=\frac{1}{2}(\delta^{}_{\Z^2}
469: +\delta^{}_{R\Z^2})$ has diffraction
470: \[
471: \widehat{\gamma}^{}_{\omega} = 
472: \delta^{}_{\gT} + \frac{1}{4}\delta^{}_{\Z^2\setminus\gT} +  
473: \frac{1}{4}\delta^{}_{R\Z^2\setminus\gT}.
474: \]
475: The difference to the diffraction formula in Lemma~\ref{diff2gitter}
476: is concentrated on $\gT\setminus\{0\}$, and hence plays no role in our
477: further discussion when $\dens(\gT)\to 0$, which happens under
478: multiple coincidence intersections \cite{bg}. We may thus disregard
479: coincidence rotations for our present purposes.
480: 
481: \begin{figure} 
482: \epsfig{file=sqexact.eps,width=0.7\textwidth}
483: \caption{\label{sqexact}
484: Radial dependence of ring intensities of square lattice powder diffraction.
485: The central intensity is not shown, see text for details.}
486: \end{figure}
487: 
488: \begin{figure} 
489: \epsfig{file=sq25plot.eps,width=0.7\textwidth}
490: \caption{\label{sqnum}
491: Numerical approximation of Figure~\ref{sqexact}, 
492: based on summing Bessel functions, for
493: radii $r\le 25$. The required radial autocorrelation coefficients
494: $\eta^{}_{\smc}(r)$ are taken from Equation~\eqref{etaq}.
495: }
496: \end{figure}
497: 
498: Let us continue by considering multiple intersections.  If all
499: rotations are generic (in the sense that the lattices $R_{i}\Z^2$ and
500: $R_{j}\Z^2$ are distinct apart from the origin) and satisfy a uniform
501: distribution property, we obtain the following result.
502: 
503: \begin{prop}
504: Let $\gG = \Z^2$, $\gG^{}_j = R^{}_j \Z^2$ with
505: generic $R^{}_j$, i.e., $\gG^{}_j \cap \gG^{}_k = \{ 0 \}$
506: for $j \ne k$, and define $\omega = \frac{1}{N} \sum_{j=1}^N
507: \delta^{}_{\gG^{}_j}$. 
508: Then, one has the identities
509: \[ 
510: \gamma^{}_{\omega} = 
511: \frac{N-1}{N} \lambda \; +
512: \; \sum\limits_{j=1}^N \frac{1}{N^2}
513: \delta^{}_{\gG^{}_j}  \quad \mbox{ and }
514: \quad \widehat{\gamma}^{}_{\omega} = \frac{N-1}{N} \delta^{}_0 +
515: \frac{1}{N} \Bigl( \frac{1}{N} \sum\limits_{j=1}^N
516: \widehat{\delta}^{}_{\gG^{}_j}   \Bigr). \]
517: If $( R^{}_j x )^{}_{j \ge 0}$ is uniformly distributed on $\Sb^1$ for
518: some fixed $x \in \Sb^1$, we obtain 
519: \[ \lim_{N\to \infty} \frac{1}{N} \sum_{j=1}^N
520: \widehat{\delta}^{}_{\gG^{}_j} = 
521: \sum_{r \ge 0} \eta^{}_{\smc}(r) \mu^{}_r,   \] 
522: with $\eta^{}_{\smc}(r)$ as in \eqref{etaq}.
523: \end{prop}
524:  
525: \begin{proof}
526: The first statement --- about $\gamma^{}_{\omega}$ --- follows 
527: as in Lemma~\ref{diff2gitter}.  
528: Observing $(\gG^{}_j)^{\ast} = R_j \gG^{\ast}$, this implies 
529: \[ 
530: \widehat{\gamma}^{}_{\omega} =  \sum_{x \in \bigcup_{j}
531:   R_j \gG^{\ast}} I(x) \delta^{}_x, 
532: \]
533: where $I(0)=1$ and $I(x)=\frac{1}{N^2}$ otherwise. Consequently, one
534: finds 
535: \[ 
536: \widehat{\gamma}^{}_{\omega} = \frac{N-1}{N} \delta^{}_0 +
537: \frac{1}{N} \Bigl( \frac{1}{N} \sum_{j=1}^N
538: \widehat{\delta}^{}_{\gG^{}_j} \Bigr). 
539: \] 
540: By Theorem~\ref{radpsf} in connection with Weyl's lemma and the
541: self-duality of $\Z^2$, the term in the brackets converges to $\sum_{r
542: \ge 0} \eta^{}_{\smc}(r) \mu^{}_r$ as $N \to \infty$.
543: \end{proof}
544: 
545: 
546: This simple argument shows that, after discarding the central
547: intensity and multiplying the remainder by $N$, one is left with a
548: circular diffraction pattern in the spirit of the radial PSF in
549: Theorem~\ref{radpsf}. Disregarding the central intensity, the shelling
550: numbers $\eta^{}_{\smc}(r)$ reflect the total intensity, integrated
551: over the rings of radius $r$, in this idealized approach to powder
552: diffraction. Consequently, in a measurement that displays the
553: intensity along a given direction, the resulting radial dependence is
554: given by $\frac{\eta^{}_{\smc}(r)}{2 \pi r}$, compare
555: Figure~\ref{sqexact}. The numerical approximation in
556: Figure~\ref{sqnum}, included for later comparison, shows a strong
557: oscillatory behaviour, as a result of summing Bessel functions.  Note
558: that this approximation disregards positivity in favour of including
559: circular symmetry. The comparison gives a good impression on the
560: overshooting that originates from this approach.
561: 
562: 
563: \section{Application to tilings with statistical circular symmetry} 
564: 
565: \subsection{The pinwheel tiling}
566: The prototile of the pinwheel tiling is a rectangular triangle with
567: side length $1$, $2$ and $\sqrt{5}$.  The smallest angle in $T$ is
568: $\arctan(\frac{1}{2})$.  Here, we choose the prototile $T$ with
569: vertices $(-\frac{1}{2},-\frac{1}{2})$, $(\frac{1}{2},-\frac{1}{2})$,
570: $(-\frac{1}{2},\frac{3}{2})$, and equip $T$ with a control point at
571: $(0,0)$.  Every tile in a pinwheel tiling is either of the form
572: $R^{}_{\alpha} T + x$, or of the form $R^{}_{\alpha} S T +x$ for some
573: $x \in \R^2$, where $S$ denotes the reflection in the horizontal axis
574: and $R^{}_{\alpha}$ rotation through $\alpha$. The substitution
575: $\sigma$ for the pinwheel tiling is shown in Figure~\ref{fig:subst}
576: (left). By the action of $\sigma$, $T$ is expanded, rotated by
577: $-\arctan(\frac{1}{2})$ and dissected into five triangles that are
578: congruent to $T$. Formally, we define $\sigma(T) := \{ Q_1(T)+x_1,
579: Q_2(T)+x_2, Q_3(T)+x_3, Q_4(T)+x_4, Q_5(T)+x_5\}$, with $Q_i \in
580: \mathrm{O}(2)$ and $x_i \in \R^2$. The appropriate choices of $Q_i$
581: and $x_i$ can be derived from the figure. In particular, $\sigma(T)$
582: contains a triangle \emph{equal} to $T$, thus we may choose $Q_1 =
583: \mathbbm{1}$ and $x_1=0$. The substitution $\sigma$ extends in a
584: natural way to all isometric copies $QT+x$ (where $Q \in
585: \mathrm{O}(2)$ and $x \in \R^2$) by
586: $\sigma(QT+x):=Q\sigma(T)+\sqrt{5}x$. Since one of the triangles in
587: $\sigma(T)$ equals $T$, the tiling $PW$ can be defined as a fixed
588: point of the substitution $\sigma$. This results in $\sigma(PW)=PW=
589: \bigcup_{n \ge 0} \sigma^n(T)$. This convention follows \cite{mps} but
590: deviates from \cite{rad}, because it is advantageous for us to include
591: the rotation in the definition of $\sigma$.
592: 
593: \begin{figure}
594: \epsfig{width=\textwidth,file=subst2.eps}
595: \caption{The substitution for the pinwheel tiling (left) and for the
596: kite domino tiling (right). The dashed lines indicate how the dominos
597: have to be dissected to obtain a pinwheel tiling. They are also needed
598: to turn the global substitution rule for $KD$ into a local one for 
599: its hull.
600: \label{fig:subst}}
601: \end{figure}
602: 
603: \subsection{The kite domino tiling}
604: It is advantageous to consider a second substitution which yields a
605: closely related tiling. Consider the two prototiles $K$ and $D$ (for
606: 'kite' and 'domino'), where $D$ is the rectangle with vertices
607: $(-\frac{1}{2},-\frac{1}{2})$, $(\frac{1}{2},-\frac{1}{2})$,
608: $(\frac{1}{2},\frac{3}{2})$, $(-\frac{1}{2},\frac{3}{2})$, and $K$ is
609: the quadrangle with vertices $(-\frac{1}{2},-\frac{1}{2})$,
610: $(\frac{1}{2},-\frac{1}{2})$, $(\frac{11}{10},\frac{3}{10})$,
611: $(-\frac{1}{2},\frac{3}{2})$. Both tiles consist of two copies of the
612: pinwheel triangle $T$, glued together along their long edges. Every
613: pinwheel tiling gives rise to a kite domino tiling by deleting all
614: edges of length $\sqrt{5}$. The substitution for the kite domino
615: tiling is shown in Figure~\ref{fig:subst} (right). Essentially, it is
616: $\sigma^2$ applied to the new prototiles, where $\sigma$ is the
617: substitution for the pinwheel tiling. If $\varrho$ denotes this new
618: substitution, a fixed point is given by $KD = \bigcup_{n \ge 0}
619: \varrho^n(D)$. This is a \emph{global} substitution that works on the
620: specific $D$ defined above, compare \cite{f} for a discussion of the
621: general substitution concept. The two tilings, $PW$ and $KD$, are
622: mutually locally derivable (MLD) in the sense of
623: \cite{bsj}. Essentially, this means that one is obtained from the
624: other by local replacement rules and vice versa.  These rules are
625: evident from Figure~\ref{fig:subst}. 
626: 
627: As $PW$ and $KD$ are MLD, the global substitution $\varrho$ also
628: induces a \emph{local} substitution in the sense of \cite{bs}, where
629: two (oriented) copies of the domino have to be distinguished. Loosely
630: speaking, this follows from the observation that the local surrounding
631: of any domino in $KD$ or one of the other elements in the hull defined
632: by $KD$ determines the type of the domino, and hence how to apply
633: the substitution to it.
634: 
635: 
636: We equip $K$ with control points at $(0,0),(\frac{2}{5},\frac{1}{5})$,
637: and $D$ with control points at $(0,0),(0,1)$. Then, the set of all
638: control points in $KD$ is equal to the set of control points in
639: $PW$. This specific discrete point set is a Delone set, denoted by
640: $\gL$ in the sequel. It is not hard to see that $\gL$ is MLD with both
641: $PW$ and $KD$, even though one direction of the replacement rule is
642: less obvious than the one linking $PW$ and $KD$.
643: 
644: Observe that the relative and absolute frequencies of triangles $T$ in
645: the pinwheel tiling are both equal to $1$. For the relative
646: frequencies, this is clear since there is only one prototile. The
647: absolute frequency then follows from the fact that this prototile has
648: area $1$. Every triangle of $PW$ carries exactly one control point, so
649: $\dens(\gL)=1,$ too. For the kite domino tiling, standard
650: Perron-Frobenius theory yields the relative frequencies of kite
651: (resp.\ domino) as $\frac{5}{11}$ (resp.\ $\frac{6}{11}$). Thus, the
652: absolute frequency of a kite (resp.\ domino) is $\frac{5}{22}$
653: (resp.\ $\frac{6}{22}$). The substitution matrix can be extracted from
654: Figure~\ref{fig:subst}.
655: 
656: \begin{figure}
657: \epsfig{file=kdpatch.eps, width=150mm}
658: \caption{A patch from a kite domino tiling. The control points of the
659:   white tiles are contained in $\Z^2$. The control points in the other
660:   tiles are contained in rotated copies of $\Z^2$, where the common
661:   rotation centre is indicated in the figure. The possible rotations
662:   are described in Claim~\ref{rotgitter}. \label{fig:kdpatch}}
663: \end{figure}
664: 
665: Discrete structures that are MLD lead to dynamical systems that are
666: topologically conjugate \cite{kel}. Therefore, we formulate the hull on
667: the basis of the Delone sets of the control points. Let $\gL$ be one
668: such set, e.g., the set $\gL$ of control points of $PW$ as defined
669: above. Define the orbit closure
670: \[ 
671: \X(\gL) = \overline{ \R^2 + \gL}^{\mathsf{LRT}}  
672: \]
673: in the local rubber topology (LRT) \cite{bl}, which is compact. This
674: topology is slightly different from the one introduced in
675: \cite{rad}. But in the present case, both topologies yield the same
676: hull and are equivalent on $\X(\gL)$.
677: 
678: The following result is well-known \cite{rad,mps}. 
679: 
680: \begin{thm} \label{autoconst}
681: The hull $\, \X(\gL)$ is \ $\Sb^1$-symmetric. Moreover, $(\X(\gL),
682: \R^2)$ is a strictly ergodic dynamical system, i.e., it is uniquely
683: ergodic and minimal.  All elements of\/ $\X(\gL)$ possess the same
684: autocorrelation measure $\gamma = \gamma^{}_{\gL}$ and the same
685: diffraction measure $\widehat{\gamma}^{}_{\gL}$. Both measures
686: are\/ $\Sb^1$-symmetric. \qed
687: \end{thm}
688: 
689: In this sense, speaking of the diffraction of the pinwheel tiling or
690: its control points has a unique meaning. At this point, we know that
691: \[ 
692: \widehat{\gamma} = \left( \dens(\gL) \right)^2 \delta^{}_0 +
693: (\widehat{\gamma})^{}_{\mathsf{cont}} = \delta^{}_0 +
694: (\widehat{\gamma})^{}_{\mathsf{cont}} 
695: \]
696: because a translation bounded circularly symmetric measure cannot
697: contain Bragg peaks other than the trivial one at $k=0$.  That the
698: intensity coefficient of $\delta^{}_0$ is $\left( \dens(\gL)
699: \right)^2$, hence $1$ in our case, is a consequence of the equation
700: \[ 
701: \widehat{\gamma}(\{0\}) = \lim_{r \to \infty} \frac{1}{\pi r^2}
702: \, \gamma( B^{}_r (0)) = \left( \lim_{r \to \infty} \frac{1}{\pi r^2}
703: \, \delta^{}_{\gL} ( B^{}_r (0)) \right)^2 = \left( \dens( \gL )
704: \right)^2. 
705: \] 
706: It can be proved by means of a Fourier concentration argument in
707: connection with results from \cite{al}; alternatively, see \cite{hof}
708: for another approach and \cite{mps} for details on the case at hand.
709: It remains to determine the structure of
710: $(\widehat{\gamma})^{}_{\mathsf{cont}}$ more closely, in particular
711: the separation into singular and absolutely continuous components.
712: 
713: \subsection{Results and observations}
714: 
715: The following claims and conjectures are stated for the particular
716: tiling $KD$ or for its set of control points $\gL$. Since mutual local
717: derivability extends to entire hulls, we remain in the situation of
718: Theorem~\ref{autoconst}. Proofs of the claims are either outlined here
719: or will appear in \cite{bfg}. Let us start with our main observation.
720: 
721: \begin{conj} \label{mainconj}
722: The diffraction measure $\widehat{\gamma}$ of the pinwheel tiling is of the
723: form  
724: \[ 
725: \widehat{\gamma} = \delta^{}_0 + (\widehat{\gamma})^{}_{\mathsf{sc}} +
726: (\widehat{\gamma})^{}_{\mathsf{ac}}.
727: \]
728: There is a countable set $\D^*\subset\R_{\ge 0}$ such that
729: \[ 
730: (\widehat{\gamma})^{}_{\mathsf{sc}} = \sum_{r \in \D^*\setminus\{0\}} 
731: I(r) \, \mu^{}_r. 
732: \]
733: The set $\D^*$ seems to be locally finite, i.e., discrete and closed.
734: Moreover, $(\widehat{\gamma})^{}_{\mathsf{ac}}$ seems to be 
735: non-vanishing.
736: \end{conj}
737: 
738: The set $\D^*$ is a subset of another set, $\D= \D^{}_{\gL}$, which
739: shows up in the determination of the autocorrelation and is described
740: in more detail below.
741: 
742: \begin{claim} \label{rotgitter}
743: Let $\gL$ be the set of all control points of $KD$. Then, $\gL$ is a
744: Delone subset of a countable union of rotated square lattices. More
745: precisely, with $\theta = 2 \arctan(\frac{1}{2})$,   
746: \[ \gL \subset  \bigcup_{k \in \Z}  R^{}_{k \theta}(\Z^2). \]
747: \end{claim}
748: 
749: Using the geometry of the kite domino tiling, it is not difficult to
750: establish this property. Figure~\ref{fig:kdpatch} may serve as a
751: visualisation on a small scale.  This structure suggests that the
752: diffraction of the pinwheel tiling might share some features with that
753: of the powder diffraction of $\Z^{2}$, see the comment after the
754: proof of Lemma~\ref{diff2gitter}. Since $R^{}_{\theta}$ acts as
755: multiplication by $\frac{1}{5} \bigl( \begin{smallmatrix} 3 & -4 \\ 4
756: & 3 \end{smallmatrix} \bigr)$, the next result is immediate.  Note
757: that $R^{}_{\theta}$ is well known from the coincidence site lattice
758: problem of the square lattice, compare \cite{bg}, and has played a
759: crucial rule in the explicit calculations in \cite{mps}.
760:  
761: \begin{claim} \label{distsupset}
762: Let $\gL$ be the set of all control points of $KD$. It satisfies
763: \[ 
764: \gL \subset \bigl\{ (\textstyle{\frac{n}{5^{k}}, \frac{m}{5^{k}}}) :
765: m,n \in \Z, \; k\in\N \bigr\}, 
766: \]  
767: and the distance set $\D = \D^{}_{\gL} :=\{ |x-y| : x, y \in
768: \gL \}$ is a subset of  $\{ \sqrt{ \frac{p^2+q^2}{5^{\ell}} } :
769: p,q,\ell \in \N \}$. It is the same set for all elements of\/
770: $\X(\gL)$. \qed
771: \end{claim}
772: 
773: We believe that this subset relation is quite sharp, i.e., that the
774: difference between the two sets is not too large, in the following sense. 
775: 
776: \begin{claim} \label{allvalues}
777: All values $r = \sqrt{p^2+q^2}$ with $p,q \in \N$ occur in
778: $\D$. Moreover, for each $\ell \in \N$, there are infinitely many $p,q
779: \in \N$ such that $\sqrt{\frac{p^2+q^2}{5^{\ell}}}$ is contained in
780: $\D$.
781: \end{claim}
782: 
783: In fact, numerical computations suggest the following, stronger
784: property. 
785: 
786: \begin{conj}\label{conj2}
787: For each $\ell \in \N$, $\D$ contains all but finitely many
788: values of the form $\sqrt{\frac{p^2+q^2}{5^{\ell}}}$.
789: \end{conj}
790: 
791: In the sequel, the absolute frequencies of distances are of interest,
792: wherefore we define the \emph{radial autocorrelation function}
793: \begin{equation} \label{eq:etar}
794: \eta(r) = \lim_{R \to \infty} \frac{1}{\vol(B^{}_R)} \sum_{
795:   \substack{ x,y \in \gL\cap B^{}_R (0) \\ |x-y| =r  } } 1. 
796: \end{equation}
797: The limit exists due to unique ergodicity, see Theorem
798: \ref{autoconst}. This permits us to write the autocorrelation of $\gL$
799: --- and hence of $\X(\gL)$ --- as
800: \begin{equation} \label{eq:sumrind} 
801:  \gamma = \sum_{r \in \D} \eta(r) \, \mu^{}_r,   
802: \end{equation} 
803: with $\mu^{}_r$ as in Section~\ref{sec:rpsf}. This follows from
804: Theorem~\ref{autoconst} together with Claim~\ref{allvalues}.  Note
805: that $\gL$ is repetitive \cite{mps} (defined up to congruence), so
806: that all $\eta(r)$ in \eqref{eq:sumrind} are strictly positive, by the
807: definition of $\D$ in Claim~\ref{distsupset}.  Clearly, $\gamma$ is
808: both a positive and positive definite measure, and a tempered
809: distribution on $\R^2$.
810: 
811: The following theorem holds in more generality than the context
812: of this paper. In fact, it holds for all substitution tilings which are
813: of \emph{finite local complexity} (FLC) and \emph{self-similar}. The
814: latter means that $\lambda T^{}_i = \bigcup_{T \in \sigma(T^{}_i)} T$. Roughly
815: speaking, this means that the support of the substitution of each
816: prototile $T^{}_i$ is similar to $T^{}_i$. Finite local complexity means
817: that, for some $R>0$, the tiling contains only finitely many local
818: patches of diameter less than $R$, up to congruence. Note that FLC is
819: frequently defined with respect to translations, whereas we define FLC
820: here with respect to congruence, because of the nature of the pinwheel
821: tiling. Both properties, FLC and self-similarity,
822: hold for the majority of substitution tilings in the literature
823: \cite{fh}.
824: 
825: \begin{thm} \label{qlambdad}
826: In any self-similar substitution tiling of finite local complexity
827: {\rm (}w.r.t.\ congruence{\rm )} with substitution factor $\lambda$,
828: all relative frequencies are contained in $\Q(\lambda^d)$.  Moreover,
829: the tiling can be scaled such that all absolute frequencies are
830: contained in $\Q(\lambda^d)$ as well.  In particular, if $\lambda^d$
831: is an integer, and if the tiling is appropriately scaled, all
832: relative and absolute frequencies are rational.
833: \end{thm}
834: 
835: For a proof, we refer to \cite{bfg}. 
836: This theorem, applied to the pinwheel tiling or the kite domino
837: tiling, yields that all frequencies are rational. 
838: 
839: \begin{claim} All values of $r^2\le 5$ together with 
840:   $\eta(r)$ are given in the following table. Values marked with an
841:   asterisk are numerically based conjectures, all
842:   other values are exact. 
843: \[ 
844: \begin{array}{l||c|c|c|c|c|c|c|c|c|c|c|c|c}
845: \rule[-2mm]{0mm}{7mm} r^2 & 0 & \frac{1}{5} & 1 & \frac{8}{5} &
846: \frac{9}{5} & \frac{49}{25} & 2 & \frac{13}{5} & \frac{81}{25} &
847: \frac{17}{5} & 4 & \frac{113}{25} & 5 \\ \hline 
848: \rule[-2mm]{0mm}{7mm} \eta(r) & 1 & \frac{5}{11} &
849: \frac{439}{165} & \frac{1}{2} & \frac{67}{165} & \frac{4}{165}
850: &  \frac{7}{2}^{\ast} & \frac{142}{165} & \frac{4}{165} &
851: \frac{10}{11}^{\ast} &  3^{\ast} &  \frac{8}{165}^{\ast} &
852: \frac{73}{15}^{\ast} 
853: \end{array}
854: \] 
855: \end{claim}\smallskip
856: 
857: The value $\eta(0)$ is the absolute frequency of the control points.
858: Since each triangle in the pinwheel tiling has area $1$, we have
859: $\eta(0)=1$.  The distance $|x-y|=1/\sqrt{5}$ occurs precisely once
860: within each kite. Kites have absolute frequency $\frac{5}{22}$, but
861: the distance $|x-y|$ has to be counted twice in view of formula
862: \eqref{eq:etar}, hence $\eta(1/\sqrt{5})=\frac{5}{11}$. The other
863: exact values contained in the table can be established similarly, but
864: require more sophistication.  With some further effort, one can
865: determine the \emph{frequency module} of $\gL$, which is the $\Z$-span
866: of the absolute frequencies of all finite subsets of $\gL$, the latter
867: standardised to having density $1$ (which is our natural setting here).
868: 
869: \begin{figure}
870: \epsfig{file=vertstarfreq.eps}
871: \caption{The 11 vertex stars of the pinwheel tiling (up to
872:   congruence) and their absolute frequencies (i.e., frequencies per unit
873:   area). \label{fig:vertstars}} 
874: \end{figure}
875: 
876: \begin{claim}
877: The frequency module of $\gL$, and hence of\/ $\X(\gL)$, is 
878: $\F= \{ \frac{m}{264 \, \cdot \,  5^{\ell}} \, | \, m \in \Z, \; \ell
879: \in \N \}$. In particular, it is countably, though not finitely generated. 
880: \end{claim}
881: 
882: This result is derived from the absolute frequencies of the vertex
883: stars in the pinwheel tiling, see Figure~\ref{fig:vertstars}. These,
884: in turn, can be derived from the frequencies of kites and darts in the
885: kite domino tiling. Details will be given in \cite{bfg}.  
886: 
887: We proceed by considering the Fourier transform in relation to the
888: distance set $\D$. Since the Fourier transform is continuous
889: on the space $\Sc'(\R^2)$  of tempered distributions,
890: \eqref{eq:sumrind} becomes  
891: \begin{equation}\label{eq:pinbessel}
892: \widehat{\gamma}^{}_{\gL} = \Bigl( \sum_{ r \in \D}  \eta(r)
893: \, \mu^{}_r \Bigr)^{\widehat{}} = \sum_{r \in \D }
894: \eta(r) \, \widehat{\mu}^{}_r, 
895: \end{equation} 
896: where the sum is to be understood in the distribution sense. By
897: Bochner's theorem, $\widehat{\gamma}^{}_{\gL}$ is again a positive
898: and positive definite measure, and the equation can be understood as a
899: vague limit as well. 
900: 
901: Let us apply Theorem~\ref{radpsf} to the pinwheel pattern $\gL$.
902: According to Claim~\ref{distsupset}, the distances in $\D$ arise from
903: lattices of the form $\frac{1}{5^{\ell/2}}\Z^2$ with $\ell\in\N$.  The
904: corresponding dual lattices, which are relevant for the diffraction
905: analysis, are $5^{\ell/2}\Z^2$, which have distance sets
906: $5^{\ell/2}\D^{}_{\smc}$. Since $\Z^2$ contains a lattice congruent to
907: $\sqrt{5} \Z^2$, it follows that $5^{\ell / 2} \D^{}_{\smc} \subset
908: \D^{}_{\smc}$ for all $\ell \in \N$, and that $\D^{}_{\smc}$ contains
909: all distances that seem relevant from this point of view. At this
910: stage, we have not found a compelling argument to exclude the
911: possibility that certain subsets of $\gL$ contribute relevant distances
912: in a coherent fashion, and thus --  by unique ergodicity -- further
913: rings $I(r) \mu_r$ (with $r\not\in\D^{}_{\smc}$) to
914: $\widehat{\gamma}$. However, on the basis of
915: Theorem~\ref{radpsf}, it is at least plausible that the set $\D^*$ is
916: closely related with $\D^{}_{\smc}$, or even equal to it. In this
917: case, $\widehat{\gamma}^{}_{\gL}$ would show rings $I(r) \, \mu^{}_r$
918: for $r \in \D^{}_{\smc}$ only. According to Claim \ref{allvalues}, all
919: distances have positive frequencies, so that this assumption would
920: lead to
921: \[ 
922: ( \widehat{\gamma})^{}_{\mathsf{sing}} = \sum_{r \in
923:   \D^{}_{\smc}} I(r) \, \mu^{}_r = \delta^{}_0 +
924: ( \widehat{\gamma})^{}_{\mathsf{sc}} = \delta^{}_0 +  \sum_{0 < r \in
925:   \D^{}_{\smc}} I(r) \, \mu^{}_r 
926: \]
927: with $I(0)=1$ and $I(r)>0$ for all $r \in \D^{}_{\smc}$. 
928: 
929: Let us continue our discussion on the basis of this hypothesis.
930: If Observation~\ref{conj2} holds, then, for each $\ell \ge 0$, only
931: finitely many distances of type $\sqrt{\frac{p^2+q^2}{5^{\ell}}}$ are
932: missing, each one contributing an absolutely continuous part to
933: $\widehat{\gamma}$. Since $\widehat{\gamma}$ exists as a translation
934: bounded measure, and since $\widehat{\gamma}^{}_{\mathsf{sing}}$ is
935: already covered by the results above, every additional contribution
936: has to contribute to $\widehat{\gamma}^{}_{\mathsf{ac}}$. This
937: motivates the conjecture that $(\widehat{\gamma})^{}_{\mathsf{ac}} \ne
938: 0$.
939: 
940: \begin{figure} 
941: \epsfig{file=pin5plot.eps,width=0.7\textwidth}
942: \caption{\label{pin5} Numerical approximation of the radial intensity
943: dependence of the pinwheel diffraction pattern. It is based on Equations
944: \eqref{eq:foubessel} and \eqref{eq:pinbessel}, disregarding contributions
945: to the central intensity. The radial autocorrelation coefficients $\eta(r)$ 
946: are estimated from a patch of radius of about $56$ (thus effectively using the
947: fifth iteration of the substitution $\sigma$). The vertical scale is arbitrary
948: in the sense that it has no meaning without local averaging and integration.}
949: \end{figure}
950: 
951: \begin{figure} 
952: \epsfig{file=zoom6plot.eps,width=0.5\textwidth}
953: \caption{\label{zoom6} Detailed view of the shoulder to the left of
954: the peak at $k=1$, calculated from the sixth iteration of the
955: substitution rule, with the large peak intensity truncated.}
956: \end{figure}
957: 
958: Observation~\ref{mainconj} is also supported by numerical computations
959: of $\widehat{\gamma}$. A naive numerical analysis of a large finite
960: portion of $\gL$ would yield no relevant result due to the nature of
961: the pinwheel pattern. In particular, the number of orientations grows
962: only logarithmically with the radius, while the number of tiles grows
963: quadratically. The results above allow a more meaningful
964: computation. In particular, by employing the $\Sb^{1}$-symmetry, the
965: problem becomes one-dimensional, and by \eqref{eq:sumbess}, the
966: Fourier transform is expressed explicitly as a sum of weighted Bessel
967: functions.  As before, this approach disregards positivity of the
968: intensity function, and the result displayed in Figure~\ref{pin5}
969: shows strong oscillations with significant overshooting, similar in
970: kind to the ones observed in Figure~\ref{sqnum}. No smoothing of any
971: kind was used.
972: 
973: 
974: A comparison of the two diffraction images (Figures \ref{sqnum} versus
975: \ref{pin5}) also supports our claim about the possible radii of
976: pinwheel diffraction rings. One noticeable difference is the higher
977: peak at $k=\sqrt{5}$, which is due to the pairs of points in $\gL$
978: with distance $1/\sqrt{5}$. The existence of positive shoulders in
979: Figure~\ref{pin5} (such as that to the left of the first peak at
980: $k=1$, a blow-up of which is shown in Figure~\ref{zoom6}) is another
981: significant difference to Figure~\ref{sqnum}, and is one of the
982: reasons why we expect a non-vanishing radially continuous
983: contribution, hence giving an absolutely continuous component to
984: $\widehat{\gamma}$.
985: 
986: 
987: 
988: \section*{Acknowledgements}
989: It is a pleasure to thank Friedrich G\"otze, Robert V.\ Moody, Thomas
990: Proffen and Anthony Quas for helpful discussions. This work was
991: supported by the German Research Council (DFG) within the
992: Collaborative Research Center 701, and by EPSRC via Grant EP/D058465.
993: 
994: \begin{thebibliography}{99}
995: 
996: \bibitem{baa}
997: M.\ Baake:\
998: \emph{Mathematical Diffraction Theory in Euclidean Spaces},
999: lecture notes, EPFL, Lausanne (2005).
1000: 
1001: \bibitem{bfg}
1002: M.\ Baake, D. Frettl\"oh, U.\ Grimm:\
1003: in preparation.
1004: 
1005: \bibitem{bg}
1006: M.\ Baake, U.\ Grimm:\
1007: Multiple planar coincidences with $N$-fold symmetry,
1008: \emph{Z.\ Krist.} {\bf 221} (2006) 571--581;\newline 
1009: {\tt math.MG/0511306}.
1010: 
1011: \bibitem{bl}
1012: M.\ Baake, D.\ Lenz:\
1013: Dynamical systems on translation bounded measures:\ pure point
1014: dynamical and diffraction spectra,
1015: \emph{Ergodic Theory Dyn.\ Syst.} {\bf 24} (2004) 1867--1893;
1016: {\tt math.DS/0302061}.
1017: 
1018: \bibitem{bs}
1019: M.\ Baake and M.\ Schlottmann:\
1020: Geometric aspects of tilings and equivalence concepts,
1021: in: Proc.\ of the 5th Intern.\ Conf.\ on Quasicrystals,
1022: eds.\ C.\ Janot and R.\ Mosseri, World Scientific,
1023: Singapore (1995), pp.\ 15--21.
1024: 
1025: \bibitem{bsj}
1026: M.\ Baake, M.\ Schlottmann, P.\thinspace D.\ Jarvis:\ 
1027: Quasiperiodic tilings with tenfold symmetry and equivalence with
1028: respect to local derivability, 
1029: \emph{J.\ Phys.\ A:\ Math.\ Gen.} {\bf 24} (1991) 4637--4654.
1030: 
1031: \bibitem{cor1}
1032: A.\ C\'ordoba:\
1033: La formule sommatoire de Poisson,
1034: \emph{C.\ R.\ Acad.\ Sci.\ Paris, S\'er.\ I:\ Math.} {\bf 306} (1988)
1035: 373--376. 
1036: 
1037: \bibitem{cor2}
1038: A.\ C\'ordoba:\
1039: Dirac combs,
1040: \emph{Lett.\ Math.\ Phys.} {\bf 17} (1989) 191--196.
1041: 
1042: \bibitem{f}
1043: D.\ Frettl\"oh:\
1044: Duality of model sets generated by substitutions, 
1045: \emph{Rev.\ Roumaine Math.\ Pures Appl.} {\bf 50} (2005) 619--639;
1046: {\tt math.MG/0601064}.
1047: 
1048: \bibitem{fh}
1049: D.\ Frettl\"oh and E.\ Harriss,
1050: \emph{Tilings Encyclopedia}, available online at:\newline
1051: {\tt http://tilings.math.uni-bielefeld.de/}
1052: 
1053: \bibitem{haw}
1054: G.\thinspace H.\ Hardy, E.\thinspace M.\ Wright:\
1055: \emph{An Introduction to the Theory of Numbers}, 5th ed.,
1056: Clarendon Press, Oxford (1979).
1057: 
1058: \bibitem{hof} 
1059: A.\ Hof:\
1060: On diffraction by aperiodic structures,
1061: \emph{Commun.\ Math.\ Phys.} {\bf 169} (1995) 25--43.  
1062: 
1063: \bibitem{ik}
1064: H.\ Iwaniec, E.\ Kowalski:\
1065: \emph{Analytic Number Theory},
1066: AMS, Providence, RI (2004). 
1067: 
1068: \bibitem{kel}
1069: J.\ Kellendonk:\
1070: Topological equivalence of tilings,
1071: \emph{J.\ Math.\ Phys.} {\bf 38} (1997) 1823--1842;\newline
1072: {\tt cond-mat/9609254}.
1073: 
1074: \bibitem{kn}
1075: L.\ Kuipers, H.\ Niederreiter:\
1076: \emph{Uniform Distribution of Sequences},
1077: Wiley, New York (1974).
1078: 
1079: \bibitem{al}
1080: J.\ Gil de Lamadrid, L.\thinspace N.\ Argabright:\
1081: \emph{Almost Periodic Measures},
1082: Memoirs AMS, vol.\ 85, no.\ 428, Providence, RI (1990). 
1083: 
1084: \bibitem{moo}
1085: R.\thinspace V.\ Moody:\ 
1086: Model sets:\ A survey, in:\
1087: \emph{From Quasicrystals to More Complex Systems},
1088: eds.\ F.~Axel, F.~D\'enoyer, J.\thinspace P.~Gazeau,
1089: EDP Sciences, Les Ulis, and Springer, Berlin (2000), pp.\ 145--166;\newline
1090: {\tt math.MG/0002020}.
1091: 
1092: \bibitem{mps}
1093: R.\thinspace V.\ Moody, D.\ Postnikoff and N.\ Strungaru:\ 
1094: Circular symmetry of pinwheel diffraction,
1095: \emph{Ann.\ Henri Poincar\'e} {\bf 7} (2006) 711--730.
1096: 
1097: \bibitem{rad}
1098: C.\ Radin:\
1099: Aperiodic tilings, ergodic theory and rotations, in:\
1100: \emph{The Mathematics of Long-Range Aperiodic Order}, 
1101: ed.\ R.\thinspace V.\ Moody,
1102: NATO ASI C 489, Kluwer, Dordrecht (1997), pp.\ 499--519.
1103: 
1104: \bibitem{sch}
1105: L.~Schwartz:\
1106: \emph{Th\'eorie des Distributions}, 
1107: rev.\ ed., Hermann, Paris (1998).
1108: 
1109: \bibitem{war}
1110: B.\thinspace E.\ Warren:\
1111: \emph{X-ray Diffraction}, reprint, Dover, New York (1990).
1112: 
1113: \end{thebibliography}
1114: \bigskip
1115: 
1116: \parindent 0pt
1117: 
1118: \end{document}
1119: 
1120: 
1121: 
1122: 
1123: