1: \documentstyle[12pt, amsmath, amsfonts, epsfig, amssymb, rotating]{article}
2: \begin{document}
3: \def\be{\begin{equation}}
4: \def\eps{\epsilon}
5: \def\cala{{\mathcal A}}
6: \def\calm{{\mathcal M}}
7: \def\calb{{\mathcal B}}
8: \def\calc{{\mathcal C}}
9: \def\calh{{\mathcal H}}
10: \def\rm{\mathbb{R}}
11: \def\nm{\mathbb{N}}
12: \def\cm{\mathbb{C}}
13: \def\zm{\mathbb{Z}}
14: \def\hm{\mathbb{H}}
15: \def\pd{\pi/2}
16: \def\ua{\underline{a}}
17: \def\ub{\underline{b}}
18: \def\tr{\textrm{tr}}
19: \def\im{\textrm{Im}}
20: \def\Ul{\overleftarrow{U}}
21: \def\Ur{\overrightarrow{U}}
22: \def\Dl{\overleftarrow{D}}
23: \def\Dr{\overrightarrow{D}}
24:
25:
26:
27:
28: %%%%%%%%%%%%%%%%%%%%%%%%%% Format de la page %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
29:
30: \textwidth= 16cm
31: \oddsidemargin= 0.5cm
32: \evensidemargin=-0.5cm
33: \topmargin=-1cm
34: \textheight= 21cm
35:
36:
37: \title{\bf Diffractive orbits in isospectral billiards}
38: \author{O. Giraud\\
39: H. H. Wills Physics Laboratory\\
40: Tyndall Avenue\\
41: Bristol BS8 1TL, UK}
42:
43: \maketitle
44:
45: \begin{abstract}
46:
47: Isospectral domains are non-isometric regions of space for which the spectra
48: of the Laplace-Beltrami operator coincide. In the two-dimensional Euclidean space,
49: instances of such domains have been given. It has been proved for these examples that the
50: length spectrum, that is the set of the lengths of all periodic trajectories, coincides as well.
51: However there is no one-to-one correspondence between the diffractive trajectories.
52: It will be shown here how the diffractive contributions to the Green functions match nevertheless
53: in a ``one-to-three'' correspondence.
54:
55: \end{abstract}
56:
57: \pagebreak
58:
59: \section{Introduction}
60:
61: The quantum-mechanical problem of finding isospectral domains, that
62: is two non-isometric regions for which the sets $\{E_n, n\in\nm\}$ of solutions of the
63: stationary Schr\"odinger equation
64: \begin{equation}
65: \label{helmholtz}
66: (\Delta+E)\Psi=0,
67: \end{equation}
68: with $\Psi|_{\textrm{boundary}}=0$,
69: are identical, has been formulated in a synthetic way by Mark Ka{\v c} in 1966 in his famous paper
70: ``Can one hear the shape of a drum'' \cite{Kac66}. Negative answers to this problem have been given
71: for domains on Riemannian manifolds \cite{Be89, Br88}. The answer for two-dimensional Euclidean domains
72: was finally given
73: by Gordon {\it et al.} \cite{GorWebWol92}, who provides explicitly a pair of simply
74: connected non-isometric Euclidean isospectral domains. The two billiards considered in
75: \cite{GorWebWol92} are represented in Figure \ref{chapman}a.
76: Using a paper-folding method, Chapman \cite{Cha95} showed that in this case
77: isospectrality arises from the existence of a map between the two domains. This method allowed him to
78: produce more examples of isospectral billiards (by billiard we mean a two-dimensional Euclidean
79: connected compact domain): he showed that any triangle or even rectangle could replace the base right angled
80: isosceles triangle used to build the billiards in \ref{chapman}a; one just has to glue
81: together 7 copies of the chosen base shape obtained by symmetry with respect to its edges, in the
82: same order as to build the billiards of \ref{chapman}a. We will work here with the pair of billiards
83: $\calb_{1}$ and $\calb_{2}$ of Figure \ref{chapman}b, where each billiard is made of seven rectangles.
84: Later on, Buser {\it et al} \cite{BusConDoy94} produced more examples of planar isospectral domains made
85: of more building blocks; all of them are based on the same principle of gluing together copies of a
86: base triangle.
87:
88: \begin{figure}[ht]
89: \begin{center}
90: \epsfig{file=billards.eps,width=9cm}
91: \end{center}
92: \caption{a) Two isospectral billiards with a triangular base shape; b) The same with a rectangular base shape.}
93: \label{chapman}
94: \end{figure}
95:
96: A natural question arises when one considers isospectral billiards: is there any relation between their
97: periodic orbits, and is there any relation between their diffractive orbits?
98: It is well-known that the quantum density of states
99: \begin{equation}
100: d(E)=\sum_n\delta(E-E_n)
101: \end{equation}
102: can be expressed by means of the advanced Green function as
103: \begin{equation}
104: \label{densitegreen}
105: d(E)=-\frac{1}{\pi}\im\int d\ua G(\ua, \ua),
106: \end{equation}
107: where the integral is performed over the domain.
108: In the case of two isospectral billiards, one might naturally expect the integrals of the Green functions
109: of the two systems to be equal. Since the Green function can be expanded as a sum over all Feynman
110: paths, the correspondence between the spectra should be associated to a correspondence between
111: the periodic orbits and between the diffractive orbits.
112: Moreover, there exists an exact expression for the Green function in a two-dimensional polygonal billiard: following
113: Sommerfeld \cite{Som96, Som54}, who provides the exact Green function for a half infinite straight mirror in 2
114: dimensions, Stovicek \cite{Sto89, Sto91b} expressed the Green function for a collection of magnetic flux lines
115: on a plane (the multi-flux Aharonov-Bohm effect) as a sum over all possible scattering paths. This
116: method has been generalized in \cite{HanTha03} to provide the exact Green function for the scalar
117: wave equation in a plane with any set of perfectly reflecting straight mirrors joined by diffractive corners.
118: The Green function is given as a scattering series involving all classical trajectories
119: and all scattering contributions; a semi-classical series expansion shows that one expects the Green
120: functions of the two system to correspond order by order.
121:
122: The correspondence between the periodic orbits of two isospectral billiards
123: (which contribute to the lowest order of the scattering series)
124: has been discussed in \cite{Gor86} for isospectral domains on Riemannian
125: manifolds.
126: The Laplace spectrum versus the length spectrum is also discussed by Gordon in \cite{GorWebWol92}
127: for Euclidean isospectral billiards. A proof of the one-to-one correspondence between
128: the length spectra (referred to as ''iso-length spectrality'') in the case of the two celebrated
129: isospectral billiards considered by Gordon is given in \cite{Tha03}, based on simple mathematical
130: tools. More generally, iso-length spectrality has been proved in \cite{OkaShu01} for all pairs of isospectral billiards
131: having a ''transplantation'' property (which corresponds, roughly speaking, to the existence
132: of a map between the two domains, such as the one that will be given in Section \ref{formalism}).
133: However, what \cite{Tha03}
134: underlines is that there is no one-to-one correspondence between
135: the diffractive trajectories of the two billiards. More precisely, one can find straight lines between
136: diffractive corners in one of the billiards that do not have any counterpart in the other. Still, the
137: equality of the spectra, and hence the equality of the trace of the Green function, imposes
138: some correspondence.
139: In Section \ref{deux} we will study the map that exists between the two billiards and show how this
140: bijection allows to prove a one-to-one correspondence of the periodic orbits. In Section \ref{trois}
141: we establish a relation between the Green function of the two studied domains. Finally in Section \ref{quatre}
142: we answer the following question: what is the correspondence between the diffractive orbits.
143: The conclusion briefly shows how the method presented here can be naturally extended
144: to other pairs of isospectral billiards.
145:
146:
147: \section{Isospectrality and periodic orbits}
148: \label{deux}
149:
150: \subsection{The translation surfaces}
151: \label{formalismematriciel}
152:
153: Instead of studying directly the billiards $\calb_1$ and $\calb_2$ of Figure \ref{chapman}b, we
154: will consider the equivalent problem of studying the translation surfaces \cite{GutJud00} associated to
155: these billiards. $\calb_1$ and $\calb_2$
156: are polygons with angles ($\pd$, $\pd$, $\pd$, $\pd$, $\pd$, $\pd$, $\pd$, $3\pd$, $3\pd$, $2\pi$).
157: A construction due to Zemlyakov and Katok \cite{ZemKat76} shows that the translation surface
158: associated to a generic rational polygonal billiard is obtained by unfolding the polygon
159: with respect to each of its sides, which means gluing to the initial polygon its images by reflexion
160: with respect to each of its sides and repeating the operation. If $\alpha_i=\pi m_i/n_i$ are the angles
161: of the polygon and $N$ the least common multiple of the $n_i$, then $2N$ copies of the initial billiard
162: are needed. In the case of the billiards $\calb_{1}$ and $\calb_{2}$, since all the angles are multiples
163: of $\pi/2$, only 4 copies are needed,
164: and the translation surfaces $\calm_{1}$ and $\calm_{2}$ obtained by this construction are represented
165: in Figure \ref{structures}. In these surfaces, all opposite sides are identified.
166: Note that each structure has 4 singularities: 2 angles of measure $6\pi$ (dot and circle), and 2 angles
167: of measure $4\pi$ (cross and star). Moreover, each singularity with angle $2(k+1)\pi$ brings the
168: contribution $k$ to the quantity $2g-2$, where $g$ is the genus of the surface \cite{BerRic81}. Therefore
169: the surfaces $\calm_{1}$ and $\calm_{2}$ both are of genus 4. Any path drawn on $\calm_\nu$ corresponds
170: to an unfolded path on $\calb_\nu$.
171: \begin{figure}[ht]
172: \begin{center}
173: \epsfig{file=structures.eps,width=10cm}
174: \end{center}
175: \caption{4 copies of the two isospectral billiards $\calb_1$ and $\calb_2$ are glued together
176: to make two isospectral translation surfaces $\calm_1$ and $\calm_2$ made from 7 tiles.}
177: \label{structures}
178: \end{figure}
179:
180:
181: Each surface is made of 7 tiles, which give to each translation surface the structure of a 7-fold
182: torus cover. If the tiles are numbered as in Figure \ref{structures}, the way these tiles
183: are glued together to form a surface of genus 4 can be expressed, following \cite{Tha03},
184: through three $7\times 7$ matrices: for
185: each structure $\calm_{\nu},\ \nu=1,2$, we introduce the matrices $R^{(\nu)}$, $\Ul^{(\nu)}$ and
186: $\Ur^{(\nu)}$ such that $R^{(\nu)}_{i,j}=1$ if the right edge of tile $i$ is glued to
187: the left edge of tile $j$ and 0 otherwise,
188: $\Ul^{(\nu)}_{i,j}=1$ if the left half of the upper edge of tile $i$ is glued to
189: the left half of the lower edge of tile $j$ and 0 otherwise,
190: and of course $\Ur^{(\nu)}_{i,j}=1$ if the right half of the upper edge of tile $i$ is glued to
191: the right half of the lower edge of tile $j$ and 0 otherwise.
192: Just looking at Figure \ref{structures}, we can obtain these matrices easily. They are
193: given in Appendix \ref{matricesRUU}. For instance, in $\calm_1$, the right neighbour
194: of tile 5 is tile 1, the right neighbour of tile 3 is tile 7, etc...
195: We can also define the matrices $L^{(\nu)}$, $\Dl^{(\nu)}$ and
196: $\Dr^{(\nu)}$ which indicate which tile is glued to the left, the bottom left or the bottom right of
197: a given tile. These three matrices are nothing but the transposes of respectively $R^{(\nu)}$, $\Ul^{(\nu)}$ and
198: $\Ur^{(\nu)}$.
199:
200:
201:
202: \subsection{Mapping between translation surfaces}
203: \label{formalism}
204:
205: The isospectrality of the two billiards arises from the existence of a mapping between $\calm_{1}$ and
206: $\calm_{2}$, provided by Gordon \cite{GorWebWol92} and made explicit by Chapman \cite{Cha95}. The
207: following section proves the isospectrality between $\calm_1$ and $\calm_2$.
208:
209: By convention, we will label any point $\ua$ in $\calm_1$ or $\calm_2$ (which are made of 7 tiles
210: ore tori) by its position $a$ in the
211: tile and the number $i$ ($1\leq i \leq 7$) of the tile it is in; we will write alternatively $\ua$ or
212: $(a,i)$. Let us define the ''transplantation matrix'' $T$ as
213: \begin{equation}
214: \label{matriceT}
215: T=\left(
216: \begin{array}{ccccccc} 1&0&0&1&0&0&1\cr 0&1&0&0&1&0&1\cr 0&0&1&0&0&1&1\cr
217: 1&0&0&0&1&1&0\cr 0&1&0&1&0&1&0\cr 0&0&1&1&1&0&0\cr 1&1&1&0&0&0&0\cr\end{array}\right).
218: \end{equation}
219: The isospectrality arises from the fact that for any given eigenstate $\phi_n$ of $\calm_1$, we can construct
220: an eigenstate $\psi_n$ in $\calm_2$ defined by
221: \begin{equation}
222: \label{psiphi}
223: \psi_n(a, i)=\frac{1}{\cala_n}\sum_j T_{i, j}\phi_n(a,j),
224: \end{equation}
225: where $\cala_n$ is a normalization factor that we will discuss in Section \ref{trois}.
226: For instance, for $a$ in tile 1 of $\calm_2$, we have
227: \begin{equation}
228: \psi_n(a,1)=\frac{1}{\cala_n}\left(\phi_n(a,1)+\phi_n(a,4)+\phi_n(a,7)\right).
229: \end{equation}
230: We call the tiles 1,4,7 in $\calm_1$ the ''pre-images'' of tile 1 in $\calm_2$, and we say that 1
231: is ''made of'' tiles 1,4 and 7.
232: The fact that the functions $\psi_n$ are the eigenstates of
233: $\calm_2$ comes from the following relations:
234: \begin{equation}
235: \label{commutation}
236: R^{(2)}T=TR^{(1)},\ \ \Ul^{(2)}T=T\Ul^{(1)},\ \ \Ur^{(2)}T=T\Ur^{(1)},
237: \end{equation}
238: which can be verified directly on the matrices given in Appendix \ref{matricesRUU}.
239: Each of these commutation relation has a natural interpretation. For instance the
240: commutation relation for $R$ means that for any pair $(i,j)$ of tiles, $i$ on $\calm_2$ and $j$
241: on $\calm_1$, we have
242: \begin{equation}
243: \label{krt}
244: \sum_{k}R^{(2)}_{i,k}T_{k,j}=\sum_{k'}T_{i,k'}R^{(1)}_{k',j},
245: \end{equation}
246: which means that
247: if $k$ is the tile on the right of tile $i$ in $\calm_2$ (i.e. $R^{(2)}_{i,k}=1$) and if we call
248: $i_1, i_2, i_3$ the pre-images of $i$ (i.e. $T_{i,i_1}=T_{i,i_2}=T_{i,i_3}=1$), then
249: by (\ref{krt})
250: \be
251: \label{trr}
252: T_{k,j}=R^{(1)}_{i_1,j}+R^{(1)}_{i_2,j}+R^{(1)}_{i_3,j}.
253: \end{equation}
254: Among the 7 possible values of $j$, the right-hand side of Equation (\ref{trr}) will be 1
255: if and only if $j$ is on the right of $i_1$, $i_2$ or $i_3$ in $\calm_1$, and 0 otherwise.
256: This implies that if $k$ is on the right of $i$,
257: the three pre-images $k_1, k_2, k_3$ of $k$ (verifying $T_{k,k_i}=1$)
258: are the three tiles on the right of
259: the three pre-images of $i$. Concretely, since tile 1 in $\calm_2$ is made of tiles 1,4,7, then
260: its right neighbour tile 5 is made of the right neighbours 2,4,6 of tiles 1,4,7. Since the
261: commutation relation (\ref{commutation}) is valid for all matrices $A\in\{R,L,\Ur,\Ul,\Dr,\Dl\}$
262: all properties of continuity between tiles in $\calm_1$ are preserved by transplantation.
263: Therefore all the functions $\psi_n$ constructed by (\ref{psiphi}) are continuous on
264: $\calm_2$. Obviously these functions verify the Helmholtz equation (\ref{helmholtz})
265: as a linear combination of solution, and with the same eigenvalues. Finally, one has to
266: note that since $T$ is an invertible matrix,
267: only $\phi_n=0$ would give $\psi_n=0$. This proves the isospectrality between $\calm_1$
268: and $\calm_2$.
269:
270:
271:
272:
273:
274:
275: \subsection{Trajectories on the translation surfaces}
276: \label{principeessentiel}
277:
278: The matrix formalism introduced in subsection \ref{formalismematriciel} allows us to construct
279: a ''movement matrix'' $M$ from any path drawn on the translation surfaces $\calm_1$ or $\calm_2$,
280: adapting the method explained in \cite{Tha03}.
281: Each tile has 6 neighbours, and any path is drawn on a sequence $(i_1,i_2,...,i_n)$ of tiles
282: such that $i_{k+1}$ is a neighbour of $i_k$, provided it does not hit any vertex (we call ''vertex''
283: a point on the surface where 4 tiles join, or a point in the middle of a horizontal edge;
284: there are 4 scattering vertices, the others are non-scattering ones).
285: Let us call $A_k\in\{R,L,\Ur,\Ul,\Dr,\Dl\}$ the matrix corresponding
286: to the movement from $i_k$ to $i_{k+1}$, i.e. the matrix verifying $(A_k)_{i_k, i_{k+1}}=1$.
287: If we define the movement matrix $M$ as $\prod_{k=1}^{n} A_k$, it will verify $M_{i_1, i_n}=1$.
288: As a product of permutation matrices,
289: $M$ is a permutation matrix, and it maps tile $i_1$ onto tile $i_n$.
290: For instance the path drawn in Figure \ref{exemplepath}
291: corresponds to a sequence of tiles $(1,6,4,4,6,5)$ and to a sequence of matrices $M=R\Ul L \Ur R$.
292: \begin{figure}[ht]
293: \begin{center}
294: \epsfig{file=exemplepath.eps,width=7cm}
295: \end{center}
296: \caption{A path drawn on $\calm_1$}
297: \label{exemplepath}
298: \end{figure}
299: Reciprocally, for any product $M=A_1 A_2...A_n$ of matrices belonging
300: to $\{R,L,\Ur,\Ul,\Dr,\Dl\}$, if $M_{i, j}=1$ then there is a sequence $(i_1=i,i_2,...,i_n=j)$
301: of tiles such that $i_{k+1}$ is a neighbour of $i_k$ and $(A_k)_{i_k, i_{k+1}}=1$. Note
302: that infinitely many sequences $(i_1,i_2,...,i_n)$ give rise to the same movement matrix,
303: because there is an infinite number of sequences and only a finite number (7!=5040) of
304: permutations.
305: \\
306: \\
307: Let us now consider a sequence of matrices $(A^{(1)}_1, A^{(1)}_2,..., A^{(1)}_n)$, and the movement matrix
308: $M^{(1)}\equiv A^{(1)}_1 A^{(1)}_2... A^{(1)}_n$; then for any $i$
309: it defines a unique sequence of tiles $(i=i_1, i_2,...,i_n)$ such that $(A^{(1)}_k)_{i_k, i_{k+1}}=1$.
310: This sequence has the property that $M^{(1)}_{i_1, i_n}=1$. According to the relation
311: (\ref{commutation}), each $A^{(1)}_k$ verifies $A^{(1)}_k=T^{-1} A^{(2)}_k T$, which implies that
312: the matrix $M^{(2)}\equiv A^{(2)}_1 A^{(2)}_2... A^{(2)}_n$ verifies
313: \begin{equation}
314: \label{commutationM}
315: T M^{(1)}= M^{(2)} T.
316: \end{equation}
317: The interpretation of this commutation relation is the same as the interpretation of (\ref{commutation}) for the
318: individual $A_k$ (see Equations (\ref{krt}) and (\ref{trr}) with $R$ replaced by $M$):
319: if one can go from $i_1$ to $i_n$ by a sequence of tiles glued in a specific way, then
320: the three pre-images of $i_n$ are the tiles obtained by starting from the three pre-images of $i_1$
321: and following a sequence of tiles glued in exactly the same way.
322: To any path drawn on $\calm_2$ not hitting corners it is possible to associate the sequence of tiles
323: on which it is drawn, starting from a tile $i$ and finishing on a tile $j$; it is therefore
324: possible to draw an identical path starting from any of the three pre-images of $i$; this path will
325: necessarily arrive to one of the three pre-images of $j$.
326:
327:
328:
329:
330:
331: \subsection{Equality of the length spectrum}
332:
333: The two surfaces
334: $\calm_1$ and $\calm_2$ are pseudo-integrable, therefore their periodic orbits occur in families
335: of parallel orbits of same length \cite{BerRic81}.
336: It has been shown in \cite{Tha03} that there is an exact one-to-one correspondence between
337: the pencils of periodic orbits of $\calm_1$ and those of $\calm_2$. The main argument is the
338: principle explained in section \ref{principeessentiel} that for any path drawn on one of
339: the surfaces there is a corresponding sequence of tiles and a
340: corresponding sequence of matrices. The product of these matrices, the ''movement matrix'' $M$,
341: has the property that $M_{i,j}=1$ if and only if the sequence of tiles goes from $i$ to
342: $j$. For any periodic orbit on $\calm_1$ the last tile has to be equal to the first one, and
343: a closed path going from tile $i$ to itself has a movement matrix verifying
344: $M^{(1)}_{i,i}=1$. The quantity tr$(M^{(1)})$ is therefore the number of tiles from which
345: one can start and come back to oneself after a sequence of tiles giving the movement matrix
346: $M^{(1)}$.
347: Since the commutation relation (\ref{commutationM}) implies that $\tr(M^{(1)})=\tr(M^{(2)})$,
348: there is the same number of tiles having this property in $\calm_2$. So for any periodic pencil
349: drawn on this sequence of tiles there will be tr$M^{(1)}$ identical copies of it
350: (in particular with same length and same width) on $\calm_1$
351: and the same number on $\calm_2$, hence the bijection between the periodic orbits.
352: As an illustration, Figure \ref{2op} shows the two pencils of periodic orbits in a given
353: direction on $\calm_1$ and $\calm_2$: the grey orbit has same length and same width on both
354: surfaces, and so does the white orbit.
355: \begin{figure}[ht]
356: \begin{center}
357: \epsfig{file=2op.eps,width=7cm}
358: \end{center}
359: \caption{Two identical strips of periodic orbits on the surfaces $\calm_1$ and $\calm_2$}
360: \label{2op}
361: \end{figure}
362:
363:
364:
365:
366:
367:
368: \section{The Green function}
369: \label{trois}
370:
371: Relation (\ref{densitegreen}) implies that the imaginary part of the trace of the (retarded)
372: Green function is identical for two isospectral billiards. Here we will be more precise and
373: express the Green function of $\calm_2$ in terms of the Green function of $\calm_1$.
374:
375:
376:
377:
378: \subsection{Tile modes and normalization}
379:
380: Each surface $\calm_1$ and $\calm_2$ is made out of 7 tiles glued together. We will call
381: tile modes the solutions of the Helmholtz equation (\ref{helmholtz}) on a tile with periodic boundary
382: conditions, that is a torus of size $u\times v$. The eigenvalues corresponding
383: to these tile modes will be denoted $E_t$, and the corresponding eigenfunctions $\chi_t$.
384:
385: There is a one-to-one correspondence between these tile modes $E_t$ and a subset of
386: the spectrum $E_n$ common to the surfaces $\calm_1$ and $\calm_2$.
387: First, any eigenfunction $\chi_t$ corresponds to a solution $\psi_t(a,i)$ (or $\phi_t(a,i)$)
388: of the Helmholtz equation on $\calm_1$ (or $\calm_2$) by simply taking the function equal
389: to (or proportional to) $\chi_t(a)$ on each tile: the periodic boundary conditions for
390: the tile eigenfunctions $\chi_t$ will make $\psi_t$ and $\phi_t$ continuous. In order to normalize
391: correctly to 1 these eigenfunctions, one has to set
392: \begin{equation}
393: \label{tilepsiphi}
394: \psi_t(a,i)=\phi_t(a,i)=\frac{1}{\sqrt{7}}\chi_t(a).
395: \end{equation}
396: Reciprocally, for any eigenstate $\phi_n$ of the surface $\calm_1$, the function defined on
397: a tile by
398: \begin{equation}
399: \chi_n(a)=\sum_{i=1}^{7}\phi_n(a, i)
400: \end{equation}
401: is either the function 0 or an eigenstate $\chi_t$ of the tile. We will use the subscript
402: $s$ when the state $\phi_n$ verifies the condition
403: \begin{equation}
404: \label{sommezero}
405: \sum_{i=1}^{7}\phi_n(a, i)=0
406: \end{equation}
407: and the subscript $t$ when it does not. The set of eigenstates of $\calm_1$
408: (and, in the same way, the set of eigenstates of $\calm_2$)
409: can then be partitioned into two sets:
410: the eigenstates $\phi_t$ which are also the eigenstates of the tile
411: and do not have the property (\ref{sommezero}), and
412: the eigenstates $\phi_s$ which have the property (\ref{sommezero}) (pure ''surface'' states).
413:
414: The normalization constant $\cala_n$ in (\ref{psiphi}) will take a different value whether
415: the function $\phi_n$ belongs to the set of $\{\phi_t\}$ or $\{\phi_s\}$, as we will see now.
416: Equation (\ref{psiphi}) expresses each eigenfunction $\psi_n$ of $\calm_2$ as a sum of
417: an eigenfunction $\phi_n$ of $\calm_1$ with the same eigenvalue, taken at three
418: different points of $\calm_1$ (or rather at a similar point on three different tiles).
419: The normalization condition on $\psi$ can be written
420: \begin{eqnarray}
421: 1=\int_{\calm_2}\left|\psi_n(x)\right|^2 dx
422: &=&\sum_{k=1}^{7}\int_{\textrm{tile}}\left|\psi_n(a,k)\right|^2 da\nonumber\\
423: &=&\frac{1}{\cala_n^2}\sum_{k=1}^{7}\sum_{i,j=1}^{7}T_{k,i}T_{k,j}\int_{\textrm{tile}}
424: \overline{\phi_n(a,i)}\phi_n(a,j)da\nonumber\\
425: &=&\frac{1}{\cala_n^2}\sum_{i,j=1}^{7}T^2_{i,j}\int_{\textrm{tile}}\overline{\phi_n(a,i)}\phi_n(a,j)da
426: \end{eqnarray}
427: where we have used the fact that $T$ is symmetric and summed over $k$. Since
428: the matrix $T^2$ is equal to $(T^2)_{i,j}=1+2\delta_{ij}$ ($\delta_{ij}$ is the
429: Kronecker symbol), we get
430: \begin{eqnarray}
431: \label{norm}
432: \cala_n^2&=&\sum_{i,j=1}^{7}\int_{\textrm{tile}}\overline{\phi_n(a,i)}\phi_n(a,j)da
433: +2\sum_{i=1}^{7}\int_{\textrm{tile}}|\phi_n(a,i)|^2da\nonumber\\
434: &=&\int_{\textrm{tile}}\left|\sum_{i=1}^{7}\phi_n(a,i)\right|^2 da+2\int_{\calm_1}|\phi_n(a,i)|^2 da.
435: \end{eqnarray}
436: For the tile modes $\psi_t$, we can replace $\phi$ by $\chi/\sqrt{7}$ from Equation (\ref{tilepsiphi}).
437: This yields
438: \begin{equation}
439: \cala_t^2=\frac{49}{7}\int_{\textrm{tile}}\left|\chi_t(a)\right|^2 da
440: +\frac{14}{7}\int_{\textrm{tile}}\left|\chi_t(a)\right|^2 da=9
441: \end{equation}
442: because the functions $\chi_t$ are normalized to 1 on the tile. For the non-tile modes $\phi_s$, which
443: verify Equation (\ref{sommezero}), Equation (\ref{norm}) gives $\cala_s^2=2$. Finally we have
444: \begin{eqnarray}
445: \label{tilenontile}
446: \psi_t(a, i)=\frac{1}{3}\sum_j T_{i, j}\phi_n(a,j)&\ \ \ \ \ \ \ \ \textrm{(tile modes)}\nonumber\\
447: \psi_s(a, i)=\frac{1}{\sqrt{2}}\sum_j T_{i, j}\phi_n(a,j)&\ \ \ \ \ \ \ \ \textrm{(surface modes)}.\nonumber\\
448: \end{eqnarray}
449:
450:
451:
452: \subsection{The Green function of the translation surfaces}
453:
454: In this section, we want to express the Green function on the surface $\calm_2$ in
455: terms of the Green function of the surface $\calm_1$. We will use the expansion over
456: eigenstates for the advanced Green function of $\calm_2$:
457: \begin{equation}
458: G^{(2)}(\ua,\ub)=\sum_{n}\frac{\overline{\psi_n(\ua)}\psi_n(\ub)}{E-E_n+i\eps}
459: \end{equation}
460: where $\psi_n$ and $E_n$ are respectively the eigenfunctions and the eigenvalues of $\calm_2$.
461: We have to split the sum over $n$ into a sum over the tile modes $\psi_t$ and the non-tile modes
462: $\psi_s$, and replace $\psi$ by its expression (\ref{tilenontile}):
463: \begin{eqnarray}
464: G^{(2)}(a,i;b,j)
465: =\frac{1}{9}\sum_t\frac{\sum_{i',j'}T_{i,i'}T_{j,j'}\overline{\phi_t(a,i')}\phi_t(b,j')}{E-E_t+i\eps}
466: \nonumber\\
467: +\frac{1}{2}\sum_s\frac{\sum_{i',j'}T_{i,i'}T_{j,j'}\overline{\phi_s(a,i')}\phi_s(b,j')}{E-E_s+i\eps}.
468: \end{eqnarray}
469: If we add and subtract $(1/2)\sum_t$, we get
470: \begin{eqnarray}
471: G^{(2)}(a,i;b,j)
472: =\frac{1}{2}\sum_{i',j'}T_{i,i'}T_{j,j'}\sum_n\frac{\overline{\phi_n(a,i')}\phi_n(b,j')}{E-E_n+i\eps}
473: \nonumber\\
474: -\frac{7}{18}\sum_{i',j'}T_{i,i'}T_{j,j'}\sum_t\frac{\overline{\phi_t(a,i')}\phi_t(b,j')}{E-E_t+i\eps}.
475: \end{eqnarray}
476: The first sum is a sum over
477: Green functions of $\calm_1$; in the sum over the remaining $t$ modes, we replace $\phi_t$
478: by its value given by Equation (\ref{tilepsiphi}).
479: This gives
480: \begin{equation}
481: G^{(2)}(a,i;b,j)=\frac{1}{2}\sum_{i',j'}T_{i,i'}T_{j,j'}G^{(1)}(a,i';b,j')
482: - \frac{1}{18}\sum_{i',j'}T_{i,i'}T_{j,j'}G^{(t)}(a;b)
483: \end{equation}
484: where $G^{(t)}(a;b)$ is the Green function on the tile. This Green function
485: does not depend on $i'$ or $j'$: therefore
486: the sum over $i'$ and $j'$ can be performed: $\sum_{i',j'}T_{i,i'}T_{j,j'}=9$ since each row and each column
487: of $T$ has three 1's. Finally the Green function of the surface $\calm_2$ is
488: \begin{equation}
489: \label{relationG}
490: G^{(2)}(a,i;b,j)=\frac{1}{2}\sum_{i',j'}T_{i,i'}T_{j,j'}G^{(1)}(a,i';b,j')
491: - \frac{1}{2}G^{(t)}(a;b).
492: \end{equation}
493: Obviously, a similar relation can be obtained that expresses $G^{(1)}$ as a sum over functions $G^{(2)}$.
494:
495:
496:
497:
498: \section{Diffractive orbits}
499: \label{quatre}
500:
501:
502: \subsection{Stovicek's formalism}
503:
504: We know from \cite{HanTha03} that the exact expression for the Green function can be obtained
505: for billiards with a polygonal enclosure as a scattering series, where each term is
506: a sum over scattering paths made of straight lines and scatters on the singular corners.
507: Each scattering path contributing to $G(\ua, \ub)$ is made
508: of a starting leg of length $r_0$ going from the initial point $\ua$ to a scatterer, then an
509: alternating series of diffractions with angles $\varphi_i\in\rm$ followed by legs of length $r_i$, $1\leq i \leq n$,
510: the last leg of length $r_n$ going from the last scatterer to the final point $\ub$. The expression given in
511: \cite{HanTha03} for scattering on straight reflectors can be adapted here and gives
512: \begin{eqnarray}
513: \label{greenstovicek}
514: G(\ua, \ub)&=&\sum_{n=0}^{\infty}\frac{1}{(2\pi)^n}\sum_{\genfrac{}{}{0pt}{}{n\ \textrm{vertex}}{\textrm{paths}}}
515: \frac{1}{2i}\int_{-\infty}^{\infty}ds_1 ds_2 ... ds_n H_0^{(1)}\left[k R(s_1, s_2, ..., s_n)\right]\nonumber\\
516: &\times&\prod_{k=1}^{n}\frac{2\pi}{(\gamma_k M_k+\theta_k+i s_k)^2-\pi^2},
517: \end{eqnarray}
518: where
519: \begin{eqnarray}
520: R^2(s_1, s_2, ..., s_n)=\left(r_0+r_1 e^{s_1}+r_2 e^{s_1+s_2}+\cdots+r_n e^{s_1+s_2+\cdots+s_n}\right)\nonumber\\
521: \times\left(r_0+r_1 e^{-s_1}+r_2 e^{-s_1-s_2}+\cdots+r_n e^{-s_1-s_2-\cdots-s_n}\right).
522: \end{eqnarray}
523: The $k$-th diffraction angle $\varphi_k$ is equal to $M_k \gamma_k+\theta_k$; here $\gamma_k$
524: is the measure of the angle at the singularity ($4\pi$ or $6\pi$ in our case),
525: $M_k$ is the winding number (i.e. the number
526: of times the path winds around the singularity), and $0\leq \theta_k<\gamma_k$.
527: A schematic example of such a scattering orbit is provided at Figure \ref{pathinfini}.
528:
529:
530: \subsection{Diffractive orbits}
531:
532: Let us now look at the ''saddle-connexions'' \cite{EskMasSch01} which we define as the geodesics joining
533: two diffracting vertices. If $u$ and $v$ are the width and the height of a tile,
534: the vertices (scattering and non-scattering)
535: are in the directions $(m u/2, n v)$ with $m,n\in\zm$. Each pair $(m,n)$ with $m$ and $n$ co-prime
536: will give the direction of a saddle-connexion. If we define
537: \begin{equation}
538: \label{lp}
539: l_p=\sqrt{(m u/2)^2+(n v)^2}
540: \end{equation}
541: for $p=(m,n)\in {\zm}^2$, $m$ and $n$ co-prime, the lengths of the saddle-connexions will be
542: of the form $k l_p$, $k\in\nm$. These saddle-connexions will be the legs of the scattering paths
543: in the expansion (\ref{greenstovicek}).
544:
545: But there is no one-to-one correspondence between
546: the saddle-connexions of $\calm_1$ and those of $\calm_2$. In fact, there are saddle-connexions
547: in $\calm_2$ that cannot be found anywhere in $\calm_1$. For instance, the dashed saddle-connexion
548: on the surface $\calm_2$ in Figure \ref{2op} in the direction $(m=1, n=2)$ has a length
549: $4 l_p$, which is twice as long as any of the saddle-connexions
550: existing in $\calm_1$. A way of understanding how this difference between the sets of
551: diffractive orbits still allows isospectrality is to prove the relation (\ref{relationG})
552: between the Green functions, using their expression (\ref{greenstovicek}) as a definition.
553: This will be done in the following section, by establishing a certain
554: correspondence between the scattering trajectories.
555:
556:
557:
558: \subsection{Correspondence between diffractive orbits}
559:
560: Let us consider a contribution to (\ref{greenstovicek}) for any of the two surfaces.
561: It is a succession of scatters and saddle-connexions of lengths
562: $(k_{p_1} l_{p_1}, k_{p_2} l_{p_2}, ..., k_{p_n} l_{p_n})$,
563: where the $k_i$ are integers and the $l_i$ are defined by (\ref{lp}).
564: In the case of forward diffraction ($\theta_k=\pi\ [2\pi]$), the integral over $s_k$ has a pole
565: at $s_k=0$. But the discontinuity arising from this singularity is canceled by an opposite discontinuity
566: in the term of order $(n-1)$ in (\ref{greenstovicek}), as it should for physical reasons
567: of continuity. This cancellation is discussed in \cite{HanTha03}.
568: This means that in the expression (\ref{greenstovicek}) one has to
569: interpret any term containing a forward scattering $\theta_k=\pi\ [2\pi]$ between $a$ and $b$
570: as the limit for $\epsilon\to 0$ of a term corresponding to a path from $a_\eps$ to $b_\eps$,
571: with $a_\eps\to a$ and $b_\eps\to b$. This shifted path from $a_\eps$ to $b_\eps$
572: will have two contributions: a straight path
573: from $a_\eps$ to $b_\eps$ missing the singularity, plus a sum over all the scattering contributions
574: starting from $a_\eps$ and winding any number of times around the singularity (see Figure \ref{abscat}).
575: \begin{figure}[ht]
576: \begin{center}
577: \epsfig{file=abscat.eps,width=9cm}
578: \end{center}
579: \caption{A contribution to the Green function in case of forward diffraction}
580: \label{abscat}
581: \end{figure}
582:
583:
584: Furthermore, this still holds even if there is no scatterer at the vertex because in that case
585: the series of diffractive terms adds up to zero: when $\gamma_k=2\pi$,
586: \begin{equation}
587: \sum_{M_k=-\infty}^{\infty}\frac{2\pi}{(\gamma_k M_k+\pi+i s_k)^2-\pi^2}=0.
588: \end{equation}
589: Therefore any saddle-connexion of length $k l_p$ can be replaced
590: (as in Figure \ref{abscat} but imagining now that $a$ and $b$ are scattering vertices and that
591: the $\times$ corresponds to a non-diffracting vertex),
592: by a sum over all possible paths made of straight lines of length $k_i l_p$
593: parallel to the saddle-connexion and by
594: windings around the non-scattering vertices any number of times, with the condition that $\sum k_i=k$.
595: So, any contribution to (\ref{greenstovicek}) can be decomposed into an infinite sum of paths
596: made of a fixed number of windings around scattering vertices and any number of windings around
597: non-scattering vertices. This is illustrated at Figure \ref{pathinfini}.
598: \begin{figure}[ht]
599: \begin{center}
600: \epsfig{file=pathinfini.eps,width=11cm}
601: \end{center}
602: \caption{A contribution to the Green function in case of forward diffraction. The filled circles
603: are scattering vertices, the empty ones are non-scattering vertices.}
604: \label{pathinfini}
605: \end{figure}
606:
607: An orbit on the torus going from $a$ to $b$ consists only of a straight path between $a$ and $b$, since
608: there is only one vertex and it is non-diffractive. But again, one can decompose this straight path into
609: a sum over paths going from the vertex to itself and winding any number of times around the vertex. These
610: contributions will add up to zero.
611: \\
612: \\
613:
614: Let us now fix two points $a$ and $b$ on the torus and consider a path of
615: the type described at the left-hand side of Figure \ref{pathinfini},
616: with any given number of windings around the vertices. All the contributions to
617: $G^{(2)}(a, i;b,j)$, $G^{(1)}(a, i;b,j)$ or $G^{(t)}(a;b)$ are of this type.
618: Reciprocally, one can write each Green function as a sum over such contributions weighted
619: by 1 if such a path exists between $(a,i)$ and $(b,j)$ and by 0 otherwise.
620:
621: Since such a path does not hit any vertex, the sequence of matrices $(A_1,..., A_n)$
622: corresponding to the movement is well-defined (see section \ref{principeessentiel}),
623: and one can construct the corresponding movement matrix $M=A_1 A_2... A_n$.
624: This path always exists on the torus, which means that it is always a contribution to
625: $G^{(t)}(a,b)$. According to section \ref{principeessentiel}, it
626: will exist on the surface $\calm_\nu$, $\nu=1,2$, if and only if $a$ and $b$ are on tiles
627: $i$ and $j$ such that $M^{(\nu)}_{i,j}=1$. Therefore any such path contributes to the
628: right-hand side of Equation (\ref{relationG}) with a weight -1/2 coming from $G^{(t)}(a,b)$
629: and with a weight $\sum(1/2)(T_{i, i'}T_{j,j'})$, where the sum runs over all the pairs
630: $i', j'$ such that the orbit
631: exists between tile $i'$ and tile $j'$, i.e. such that $M^{(1)}_{i',j'}=1$. The total
632: weight associated to this path in the right-hand side of Equation (\ref{relationG}) is
633: therefore
634: \begin{equation}
635: \label{poidsdroite}
636: \frac{1}{2}\sum_{i', j'}T_{i, i'}T_{j,j'}M^{(1)}_{i',j'}-\frac{1}{2}.
637: \end{equation}
638: Using the commutation relation (\ref{commutationM}), we have
639: \begin{equation}
640: \label{eq26}
641: \sum_{i', j'}T_{i, i'}T_{j,j'}M^{(1)}_{i',j'}=\left(M^{(2)} T^2\right)_{i,j}.
642: \end{equation}
643: But it can be computed directly
644: from the expression (\ref{matriceT}) for $T$ that $(T^2)_{ij}=1+2\delta_{ij}$. Therefore
645: \begin{equation}
646: \label{eq27}
647: \left(M^{(2)} T^2\right)_{i,j}=\sum_{k}M^{(2)}_{i,k}+2 M^{(2)}_{i,j}.
648: \end{equation}
649: Since $M^{(2)}$ is a permutation matrix, the sum over a column is equal to $1$. We get, from
650: equations (\ref{eq26}) and (\ref{eq27}),
651: \begin{equation}
652: \label{identite}
653: \sum_{i', j'}T_{i, i'}T_{j,j'}M^{(1)}_{i',j'}=1+2 M^{(2)}_{i,j}.
654: \end{equation}
655: The weight (\ref{poidsdroite}) of a path in the right-hand side of Equation (\ref{relationG}) is
656: therefore equal to $M^{(2)}_{i,j}$, which is exactly its weight
657: in the expression of $G^{(2)}(a,i;b,j)$. This is an alternative way of proving the equality (\ref{relationG}),
658: using the formula (\ref{greenstovicek}) for the Green function instead of the isospectrality
659: of the surfaces.
660: \\
661:
662: In fact, Equation (\ref{identite}) provides a more precise result: it states
663: that there is a ``one-to-three'' correspondence between the diffracting orbits of $\calm_2$
664: and those of $\calm_1$. The main point is that for $(i,j)$ given tiles on $\calm_2$,
665: $\sum_{i', j'}T_{i, i'}T_{j,j'}M^{(1)}_{i',j'}$ is the number of times a path
666: of movement matrix $M^{(1)}$ appears among the 9 Green functions $G^{(1)}(a,i';b,j')$
667: with $i'$ pre-image of $i$ and $j'$ pre-image of $j$.
668: So if a path exists on $\calm_2$ between tile $i$ and tile $j$ (which means $M^{(2)}_{i,j}=1$),
669: then 3 copies of it exist in $\calm_1$ (and this will necessarily be between
670: the 3 pre-images of $i$ and the three pre-images of $j$), whereas if
671: a path does not exist in $\calm_2$, then by Equation (\ref{identite})
672: only 1 copy of it exists in $\calm_1$ (between
673: one of the three pre-images of $i$ and one of the three pre-images of $j$).
674:
675:
676: \section{Back to periodic orbits}
677:
678: A final question one might want to ask is the following: why is there a one-to-one
679: correspondence between periodic orbits, whilst each periodic orbit in $\calm_2$
680: should correspond to 3 orbits in $\calm_1$? The answer is that these 3 orbits are not
681: necessarily periodic.
682: Consider a periodic orbit in $\calm_2$, not hitting any vertex
683: (almost all of them verify this condition), going from tile $i$ to itself with a
684: movement matrix $M^{(2)}$. Let us call $\{i_1, i_2, i_3\}$ the three pre-images of $i$.
685: Then, according to what we said in the previous section,
686: there are 3 copies of this orbit in $\calm_1$ going from
687: $k\in\{i_1, i_2, i_3\}$ to $k'\in\{i_1, i_2, i_3\}$. But these copies in $\calm_1$
688: are not periodic if $k\neq k'$. Moreover, we know that there are in fact 7 copies
689: of this orbit in $\calm_1$, if we do not restrict ourselves to the pre-images of $i$ and $j$
690: as starting and ending tiles. Therefore for periodic
691: orbits the correspondence is more global: the condition of periodicity imposes that
692: we take into account not only orbits from pre-images to pre-images, but orbits from
693: any tile to any tile. Then, as we already said,
694: the number of such orbits is $\tr M^{(1)}$, which is equal to $\tr M^{(2)}$. We should
695: therefore speak of a $\tr M^{(1)}$-to-$\tr M^{(1)}$ correspondence between periodic
696: orbits, the value of $\tr M^{(1)}$ depending on the periodic orbit
697: considered.
698:
699:
700:
701:
702:
703: \section{Conclusion}
704: For the pair of billiards given in Figure \ref{structures}, which are made of rectangular tiles glued
705: together, there exists a set of neighbour matrices
706: $\{R^{(\nu)}$, $L^{(\nu)}$, $\Ur^{(\nu)}$, $\Ul^{(\nu)}$, $\Dr^{(\nu)}$, $\Dl^{(\nu)}\}$
707: describing the way these tiles are glued together. The essential feature which accounts both for
708: isospectrality and for the correspondence between paths on the surface is the existence
709: of a ''transplantation matrix'' $T$ which has 3 properties:
710: \begin{itemize}
711: \item it is invertible (otherwise one of the spectra would just be a subset of the other)
712: \item it is not a permutation matrix itself (otherwise the two domains would just be
713: congruent)
714: \item it has the commutation property $T M^{(1)}=M^{(2)} T$ for all neighbour matrices $M$,
715: which assures that smoothness at all the segments between the tiles and boundary conditions
716: will be satisfied.
717: \end{itemize}
718:
719: It is therefore possible to generalise the previous analysis to all the pairs of isospectral
720: billiards made out of base tiles glued together, provided one can find a transplantation
721: matrix $T$ having these 3 properties. It turns out that all the examples of pairs of isospectral
722: billiards, as far as is known (see \cite{BusConDoy94} and \cite{OkaShu01}), are constructed by the same
723: application of a theorem by Sunada \cite{Sun85} and consist of tiles of a given base shape
724: glued together. Moreover, Sunada's theorem implies the existence of a transplantation matrix
725: in each case \cite{OkaShu01}. Therefore Equation (\ref{eq26}) always holds, and
726: the same arguments can be adapted to establish a correspondence
727: between diffractive orbits in all these other known cases. However, the correspondence depends
728: on the entries of the matrix $T$ and $T^2$, and might be more complicated in the general case.
729: The same results could also be worked out for pairs of isospectral billiards which would not
730: be based on Sunada's theorem but would nevertheless have a transplantation matrix. It is not
731: known however if such billiards exist \cite{OkaShu01}.
732:
733:
734:
735:
736:
737:
738:
739:
740:
741: \section*{Acknowledgments}
742: J. H. Hannay is warmly thanked for his help all along this project.
743:
744:
745:
746:
747:
748: \appendix
749:
750:
751:
752:
753: \section{Appendix}
754: \label{matricesRUU}
755:
756: The matrices that describe the gluings between the plates of the translation surfaces
757: $\calm_1$ and $\calm_2$ are obtained by Figure (\ref{structures}). For the structure $\calm_1$
758: they read
759: \begin{eqnarray}
760: \label{matrices}
761: &R^{(1)}=\left(
762: \begin{array}{ccccccc} 0&0&0&0&0&1&0\cr 0&0&1&0&0&0&0\cr 0&0&0&0&0&0&1\cr
763: 0&0&0&1&0&0&0\cr 1&0&0&0&0&0&0\cr 0&0&0&0&1&0&0\cr 0&1&0&0&0&0&0\cr\end{array}\right),
764: \Ul^{(1)}=\left(
765: \begin{array}{ccccccc} 0&0&1&0&0&0&0\cr 0&1&0&0&0&0&0\cr 1&0&0&0&0&0&0\cr
766: 0&0&0&0&0&1&0\cr 0&0&0&0&1&0&0\cr 0&0&0&1&0&0&0\cr 0&0&0&0&0&0&1\cr\end{array}\right),\nonumber\\
767: &\Ur^{(1)}=\left(
768: \begin{array}{ccccccc} 1&0&0&0&0&0&0\cr 0&1&0&0&0&0&0\cr 0&0&1&0&0&0&0\cr
769: 0&0&0&0&0&1&0\cr 0&0&0&0&0&0&1\cr 0&0&0&1&0&0&0\cr 0&0&0&0&1&0&0\cr\end{array}\right).
770: \end{eqnarray}
771:
772: For the structure $\calm_2$
773: they read
774: \begin{eqnarray}
775: \label{matricesRU}
776: &R^{(2)}=\left(
777: \begin{array}{ccccccc} 0&0&0&0&1&0&0\cr 0&0&0&0&0&0&1\cr 0&1&0&0&0&0&0\cr
778: 0&0&0&1&0&0&0\cr 0&0&0&0&0&1&0\cr 1&0&0&0&0&0&0\cr 0&0&1&0&0&0&0\cr\end{array}\right),
779: \Ul^{(2)}=\left(
780: \begin{array}{ccccccc} 0&0&1&0&0&0&0\cr 0&1&0&0&0&0&0\cr 1&0&0&0&0&0&0\cr
781: 0&0&0&0&0&1&0\cr 0&0&0&0&1&0&0\cr 0&0&0&1&0&0&0\cr 0&0&0&0&0&0&1\cr\end{array}\right), \nonumber\\
782: &\Ur^{(2)}=\left(
783: \begin{array}{ccccccc} 0&0&0&1&0&0&0\cr 0&1&0&0&0&0&0\cr 0&0&0&0&0&1&0\cr
784: 1&0&0&0&0&0&0\cr 0&0&0&0&1&0&0\cr 0&0&1&0&0&0&0\cr 0&0&0&0&0&0&1\cr\end{array}\right).
785: \end{eqnarray}
786:
787: One can verify explicitly that the commutation relations (\ref{commutation}) hold for these
788: matrices.
789: \\
790: \\
791: It is of some mathematical interest to notice that the groups $\Gamma^{(1)}$ and $\Gamma^{(2)}$
792: generated by the neighbour
793: matrices $\{R^{(\nu)}, L^{(\nu)}, \Ur^{(\nu)}, \Ul^{(\nu)}, \Dr^{(\nu)}, \Dl^{(\nu)}\}$ for $\nu=1,2$,
794: are subgroups of $S_7$ (the group of permutations of 7 elements) of order 168. These groups
795: turn out to be isomorphic to the linear group $L_2(7)$ (also called $PSL(2,7)$), which is
796: the group of automorphisms of the finite projective plane of order 2, or Fano plane (see
797: \cite{Kar76} for a definition). The matrix $T$ is nothing but the incidence matrix of the
798: graph corresponding to the Fano plane.
799:
800:
801: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
802: % REFERENCES BIBLIOGRAPHIQUES
803: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
804:
805:
806: \begin{thebibliography}{10}
807:
808: \bibitem{Kac66}
809: Ka{\v c} M 1966 {\it Am. Math. Monthly} {\bf 73} 1
810:
811: \bibitem{Be89}
812: B\'erard P 1989 {\it Ast\'erisque} {\bf 177-178} 127
813:
814: \bibitem{Br88}
815: Brooks R 1988 {\it Am. Math. Mon.} {\bf 95} 823
816:
817: \bibitem{GorWebWol92}
818: Gordon C, Webb D and Wolpert S 1992 {\it Invent. math.} {\bf 110} 1
819:
820: \bibitem{Cha95}
821: Chapman S J 1995 {\it Am. Math. Monthly} {\bf 102} 124
822:
823: \bibitem{BusConDoy94}
824: Buser P, Conway J and Doyle P 1994 {\it International Mathematics Research Notices} {\bf 9} 391
825:
826: \bibitem{Som96}
827: Sommerfeld A 1896 {\it Math. Ann.} {\bf 47} 317
828:
829: \bibitem{Som54}
830: Sommerfeld A 1954 {\em Optics} (Academic, New York, 1954).
831:
832: \bibitem{Sto89}
833: {\v S}tov{\'\i}{\v c}ek P 1989 {\it Phys. Lett. A} {\bf 142} 5
834:
835: \bibitem{Sto91b}
836: {\v S}tov{\'\i}{\v c}ek P 1991 {\it J. Math. Phys} {\bf 32} 2114
837:
838: \bibitem{HanTha03}
839: Hannay J H and Thain A 2003 {\it J. Phys. A: Math. Gen.} {\bf 36} 4063
840:
841: \bibitem{Gor86}
842: Gordon C S 1986 {\it Contemp. Math.} {\bf 51} 63
843:
844: \bibitem{Tha03}
845: Thain A, to be published (unpublished).
846:
847: \bibitem{GutJud00}
848: Gutkin E and Judge C 2000 {\it Duke Math. J.} {\bf 103} 191
849:
850: \bibitem{ZemKat76}
851: Zemlyakov A B and Katok A N 1976 {\it Math. Notes} {\bf 18} 760
852:
853: \bibitem{BerRic81}
854: Berry M V and Richens P J 1981 {\it Physica D} {\bf 2} 495
855:
856: \bibitem{EskMasSch01}
857: Eskin A, Masur H and Schmoll M math.DS/0107204
858:
859: \bibitem{OkaShu01}
860: Okada Y and Shudo A 2001 {\it J Phys A: Math Gen} {\bf 34} 5911
861:
862: \bibitem{Sun85}
863: Sunada T 1985 {\it Ann Math} {\bf 121} 169
864:
865:
866: \bibitem{Kar76}
867: K\'arteszi F 1976 {\it Introduction to finite geometries} (North-Holland - Amsterdam - Oxford)
868:
869:
870: \end{thebibliography}
871:
872:
873:
874: \end{document}
875: