1: \documentstyle[prl,twocolumn,aps,floats,epsfig]{revtex}
2: %\documentstyle[prl,twocolumn,aps,floats,epsfig]{article}
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: %\tolerance = 10000
5: %\parindent=4mm
6: %\addtolength{\textheight}{0.9truecm}
7:
8: \begin{document}
9: \draft
10: \flushbottom
11: \twocolumn[
12: \hsize\textwidth\columnwidth\hsize\csname @twocolumnfalse\endcsname
13:
14: \title{ A Finite Element Algorithm for High-Lying Eigenvalues and
15: Eigenfunctions with Homogeneous Neumann and Dirichlet Boundary Conditions}
16: \author{G. B\'{a}ez$^{1,3}$, F. Leyvraz$^{1,4}$,
17: R. A. M\'{e}ndez-S\'{a}nchez$^{1,2,5}$ and T. H. Seligman$^{1,6}$}
18: \address{$^1$Centro de Ciencias F\'{\i}sicas,Universidad Nacional Aut\'{o}noma
19: M\'{e}xico\\
20: A.P. 48-3, 62250, Cuernavaca, Morelos, MEXICO.}
21: \address{$^2$Universit\"at G.H. Essen, Fachbereich 7, Physik, 45117, Essen,
22: Germany.}
23: \address{$^3$e-mail:baez@fis.unam.mx $^4$e-mail:leyvraz@fis.unam.mx}
24: \address{$^5$e-mail:mendez@fis.unam.mx $^6$e-mail:seligman@fis.unam.mx}
25:
26: \maketitle
27:
28: %\tightenlines
29: %\widetext
30: %\advance\leftskip by 57pt
31: %\advance\rightskip by 57pt
32: \begin{abstract}
33: We present a finite element algorithm that computes eigenvalues and
34: eigenfunctions of the Laplace operator for two-dimensional problems with
35: homogeneous Neumann or Dirichlet boundary conditions or combinations of
36: either
37: for different parts of the boundary. In order to solve the generalized
38: eigenvalue problem, we use an inverse power plus Gauss-Seidel algorithm. For
39: Neumann boundary conditions the method is much more efficient than the
40: equivalent finite difference algorithm. We have cheked the algorithm
41: comparing the cumulative level density of the espectrum obtained numerically,
42: with the theoretical prediction given by the Weyl formula. A systematic
43: deviation was found. This deviation is due to the discretisation and not to the
44: algorithm. As an application we calculate the statistical properties of the
45: eigenvalues of the acoustic Bunimovich stadium and compare them with the
46: theoretical results given by random matrix theory.\\
47:
48: \vskip .5 cm
49:
50: Presentamos un algoritmo de elementos finitos que calcula eigenvalores
51: y eigenfunciones del operador de Laplace para problemas en dos dimensiones con
52: condiciones a la frontera de Neumann o Dirichlet o combinaciones de ambas en
53: distintas partes de la frontera. Para resolver el problema de eigenvalores
54: generalizado, usamos un algoritmo de potencias inverso m\'{a}s otro de
55: Gauss-Seidel. Para condiciones a la frontera de Neumann, el m\'{e}todo es
56: mucho m\'{a}s eficiente que el algoritmo equivalente de diferencias
57: finitas. Hemos probado el algoritmo comparando la densidad
58: acumulada de niveles del espectro obtenido
59: num\'ericamente, con la predicci\'on
60: te\'orica dada por la f\'{o}rmula de Weyl. Se encontr\'{o} una desviaci\'on
61: sistem\'{a}tica. Esta desviaci\'on es debida a la discretizaci\'{o}n y no al
62: algoritmo. Como una aplicaci\'{o}n, calculamos las propiedades
63: estad\'{\i}sticas de los eigenvalores del estadio de Bunimovich ac\'ustico
64: y las comparamos con los resultados te\'{o}ricos dados por la
65: teor\'{\i}a de matrices aleatorias.
66:
67: \end{abstract}
68:
69: \vskip .5cm
70: %\pacs{Subject Classification: 65P25, 81C06, 81C07}
71: {Subject Classification: 65P25, 81C06, 81C07}
72:
73: \vskip .5cm
74:
75: ]
76:
77: %\narrowtext
78: %\tightenlines
79: %\setcounter{equation}{0}
80:
81: \begin{center}
82: {\bf I Introduction}
83: \end{center}
84:
85: Several years ago Neuberger and Noid\cite{Neu1,Neu2} presented an algorithm
86: and a FORTRAN program for the successive computation of the high-lying
87: eigenvalues and eigenfunctions of a time independent Schr\"{o}dinger or
88: Helmholtz equation. They used an inverse power method with a Gauss-Seidel
89: procedure for the inversion, and solved the problem with finite differences
90: on successively finer grids. This program was often used and a
91: two-dimensional version thereof\cite{Neu3} was adapted for the case of
92: Laplace operators with homogeneous boundary conditions
93: \cite{Novaroetal,Mateosetal}. The case of Dirichlet conditions works very well
94: and requires
95: minimal adjustments. This is not the case for Neumann conditions.
96: Computation times increase by orders of magnitude compared to the Dirichlet
97: case\cite{Novaroetal}; this is not surprising as the treatment of an
98: irregular boundary, and particularly of corners, is very cumbersome.
99:
100: The need for such programs arises both in acoustic\cite{BaezG} and
101: earthquake research\cite{Mateosetal}, as well as for other wave phenomena.
102: For example, if we want to make statistics of eigenvalues, say for acoustic
103: systems\cite{BaezG}, we need large numbers of states and therefore efficient
104: algorithms. In particular, geometries whose high-frequency limit (ray
105: dynamics) is chaotic, are of great interest. The starting point in this new
106: field called acoustic chaos is that the time-independent wave equation
107: (Helmholtz equation) is the same for different systems such as: quantum
108: billiards\cite{BGS1,Berry}, membranes\cite{Arcosetal} and flat microwave
109: cavities \cite{Sridhar,Richteretal,Stockmann&Stein}. Thus the statistical
110: fluctuation measures developed in nuclear physics\cite{Metha,Brodyetal}
111: have been applied to those systems and to a wide variety of more complicated
112: systems such as: Chladni's plates\cite{Stein&Stockmann,Legrandetal}, quartz
113: crystals\cite{Guhr}, aluminum blocks\cite{Guhr2,Weaver}, quantum dots
114: \cite{Marcusetal}, quantum corrals\cite{Crommieetal,Crommieetal&Heller}, waves
115: in a ripple tank\cite{Blumeletal}, elastic media with ray
116: splitting\cite{Ottetal},microwave cavities with ray splitting\cite{Kochetal}
117: amog others. As rather fine details of the boundary of those systems are
118: believed to be important, a good representation of the boundary conditions is
119: essential. Similar arguments will hold if we wish to study the effect of
120: obstacles inside a cavity or in the old Tenochtitlan lake bed\cite{Baezetal}.
121:
122: We shall use the finite element method (FEM). It is based on the
123: minimization of the functional:
124:
125: \begin{equation}
126: {\cal F}\left[ \Psi \right] =\int_R\left( \nabla \Psi \right)
127: ^2ds-k^2\int_R\Psi ^2ds, \label {funcional}
128: \end{equation}
129: where $R$ is a two dimensional region and $\psi$ is the wave function.
130:
131: The possibility of solving mixed boundary conditions will be important in
132: systems with mirror symmetries, in which we may solve
133: each non-symmetric part using
134: mixed boundary conditions. An irregular boundary as well as corners can also
135: easily be
136: implemented. The FEM\ formalism is based on minimizing
137: the functional (\ref{funcional}) not in the complete Hilbert space but only
138: in a sub-space
139: spanned by a finite set of piece-wise linear functions, typically defined as
140: pyramids over hexagonal cells. We call an element the triangles that form
141: this cell.
142:
143: The finite difference method becomes very cumbersome for boundary conditions
144: involving normal derivatives at irregular boundaries. This is particularly
145: troublesome for Neumann conditions for which it leads to poor convergence.
146: Novaro et al. (\cite {Novaroetal}) found in a particular case that computations
147: would be two orders of magnitude slower for Neumann conditions than for
148: Dirichlet conditions.
149:
150: The minimization of the functional (\ref{funcional}) on the
151: subspace of Hilbert space and in the non-orthogonal basis
152: mentioned above yields the generalized eigenvalue
153: equation
154:
155: \begin{equation}
156: A{\bf x}=\lambda B{\bf x}, \label{generalised}
157: \end{equation}
158: with $A_{ij}=\int_R\nabla \psi _i\cdot \nabla \psi _jds$ and
159: $B_{ij}=\int_R\,\psi _i\psi _jds$, where $\psi _i$ and $\psi _j$ denote the
160: functions defined around the node $i$ and $j$ of the hexagonal grid
161: respectively. These functions for simplicity, are taken to be linear
162:
163: \begin{equation}
164: \psi _i^k=a_i^kx+b_i^ky+c_i^k \label{functions}
165: \end{equation}
166:
167: with $a_i^k,b_i^k$ and $c_i^k$ constants to be determined for each triangle
168: $\Delta _k$ of the i-th hexagonal element. The trial functions are then
169: defined as pyramids of unit height over each hexagonal cell, {\it i.e.} one
170: function corresponding to each node of the grid. The simplicity of a
171: piecewise linear basis gives as result a non-orthogonal basis.
172: The Neumann boundary conditions are obtained by allowing variations of
173: the trial functions on the boundary. The Dirichlet boundary conditions can
174: be obtained by putting the trial functions to zero in the desired part of
175: the boundary. We refer to the literature for a general and deeper discussion
176: of the finite element formalism\cite{Strang&Fix,Schwarz,Bethe&Wilson,Mori}.
177:
178: If the dimension $N$ of the matrices $A$ and $B$ is small enough that they can
179: be diagonalised, standard techniques for non-orthogonal bases may be used. But
180: in a typical application, the dimensions are of several thousand to
181: tenthousands. Yet we
182: are only interested in a fairly small number of eigenvalues and eigenfunctions
183: near the low end of the spectrum. We thus have to use some method that makes
184: explicit use of the sparseness of the matrices $A$ and $B$. We shall see in
185: the next section, that a combination of
186: the inverse power and Gauss-Seidel methods proposed by
187: Neuberger and Noid\cite{Neu1,Neu2} can be generalised to solve Eq.
188: (\ref{generalised}).
189:
190: Thus in the following section we show how the inverse power method can be
191: applied when a non-orthogonal basis is used. Next we discuss
192: how to implement this for finite differences as well as a number of
193: tricks that can be used to accelerate the numerical procedure and comment on
194: the performance of the program. In section IV we apply the program to the
195: acoustic stadium, analyze the resulting spectra and the corresponding
196: states. We give a correction to the spectral density based on an analysis of
197: the equations infinite elements for the exactly solvable rectangle discussed in the appendix, Finally we present some conclusions.
198:
199: \begin{center}
200: {\bf II The inverse power method in a non-orthogonal basis}
201: \end{center}
202:
203: We shall transform the Eq.\ (\ref
204: {generalised}) by left multiplication with $B^{-1}$ to the form
205: \begin{equation}
206: B^{-1}A{\bf x}=\lambda {\bf x}
207: \end{equation}
208: As usual a power $(B^{-1}A)^{-M}\equiv (A^{-1}B)^M$ applied to an arbitrary
209: initial vector ${\bf x}^0$ will successively select the lowest
210: eigenvector corresponding to eigenvalue $\lambda _1$ as it will appear with
211: a power $(1/\lambda _1)^M$.
212:
213: The problem now seems to be that $A^{-1}$ is no longer a sparse matrix, but
214: this can be averted in the procedure of applying $A^{-1}B$. Thus we need to
215: know the product
216: \begin{equation}
217: {\bf y}\equiv A^{-1}B{\bf x}^0. \label{defin}
218: \end{equation}
219:
220: In order to obtain this product we define $\widetilde{ {\bf x}}^0\equiv B%
221: {\bf x}^0$ giving for Eq. (\ref{defin})
222: \begin{equation}
223: {\bf y}=A^{-1}{\bf \widetilde{x}}^0
224: \end{equation}
225: Multiplying by A by the left we obtain
226:
227: \begin{equation}
228: A{\bf y}=\widetilde{{\bf x}}^0
229: \end{equation}
230: that can be solved alternatively by Jacobi or Gauss-Seidel procedures which
231: will work well as all matrices involved continue to be sparse. Summarising:
232: the Gauss-Seidel Method can be utilized because we do not need the inverse
233: matrix $A^{-1}$ but the product $A^{-1}\widetilde{{\bf x}}^0$.
234:
235: Up to here we
236: have only specified how to obtain the lowest state; for excited states
237: the usual procedure is to orthogonalize the space in which we carry out the
238: variation on all states that have already been calculated. Here again the
239: non-orthogonality of our basis has to be taken into account; indeed, the
240: eigenstates of the Laplace operator are orthogonal and we have to derive the
241: consequences this has for eigenvectors in our non-orthogonal basis. If $\Psi
242: _i$ and $\Psi _j$ are eigenstates of the Laplace operator and are expanded
243: as with coefficients $\alpha _m^l\,;\,l=i,j$ that form vectors ${\bf x}_l$
244: we can readily check that
245: \begin{equation}
246: \int_R\Psi _iB\Psi _jds=\delta _{ij} \label{orto}
247: \end{equation}
248: implies
249: \begin{equation}
250: {\bf x}_i^t B{\bf x}_j=\delta _{ij} \label{Borto}
251: \end{equation}
252: and vice-versa. Thus we replace the usual orthogonalization by what we may
253: call a $B$-orthogonalization, i.e. we require the new vectors to be
254: orthogonal on $B{\bf x}_j$ (where $j$ indicates the calculated
255: eigenstates) thus guaranteeing the validity of Eq. (\ref{Borto}) and by
256: consequence Eq. (\ref{orto}).
257:
258: \begin{center}
259: {\bf III Implementation of finite elements and convergence}
260: \end{center}
261:
262: Once we know the grid and the matrix elements of $A$ and $B$ both of which
263: will be evaluated in the end of this section, we are in principle ready to
264: write our program. As usual a program consists in part of an efficient
265: algorithm which we have presented, and in part of a bag of semi-empirical
266: tricks that tend to repeat themselves in different guises again and again.
267: The efficiency of the latter is quite problem-dependent and the
268: corresponding parameters must be adjusted to optimize operation of the
269: program in every case. We shall give some recommendations that ought to work
270: reasonably, but we urge the user to fiddle around with these parameters.
271:
272: When using Neumann boundary conditions, the lowest eigenvalue is zero and
273: its corresponding eigenfunction is a constant. Yet the Gauss-Seidel inversion
274: requires
275: positive definite eigenvalues. This is obtained by adding an arbitrary
276: constant $C$ multiplied by $B$ to the Laplace operator. If we choose this
277: constant large we will need few Gauss-Seidel iterations as the operator is
278: near diagonal. On the other hand, we will loose precision and convergence
279: speed in the inverse power process as neighbouring eigenvalues will have
280: their inverse very close to each other. A good balance seems to be to choose
281: the constant smaller than, but of the order of, the largest eigenvalue which
282: we want to obtain. In the case of Dirichlet conditions this parameter can
283: also be introduced as a means of improving convergence exclusively. Adequate
284: choices of this parameter may improve computation time by a factor of $\sim
285: 2 $.
286:
287: Superconvergence, or over-relaxation, is another standard tool to improve
288: convergence. We introduce a factor $1<\alpha <2$ and at each step of any
289: iteration from ${\bf x}_n\to {\bf x}_{n+1}$ we replace ${\bf x}
290: _{n+1}$ by ${\bf x}_n+\alpha ({\bf x}_{n+1}-{\bf x}_n)$ thus
291: correcting a little more than the iteration warrants. For the Gauss-Seidel
292: iterations values of $\alpha \approx 1.5$ yield improvements in computing
293: time of the order of $\sim 2$, which is consistent with what we found for
294: finite differences. But a careful analysis for the stadium shows that at
295: least in this case for a quite precise value of $\alpha =1.75$ we find an
296: acceleration by a factor of $\sim 3$. On the other hand, in the power
297: iteration, improvements are not very significant and only values of $\alpha $
298: near 1 seem acceptable. We do not use superconvergence in this context, but
299: again we must warn that it might be very significant in other cases. All
300: these parameters were carefully explored by M\'{e}ndez\cite{MendezR}.
301:
302: Another option is the stepwise introduction of finer grids, with
303: interpolated results from the rougher grid results as starting point. This
304: idea was extensively exploited in the finite difference programs of
305: Neuberger and Noid\cite{Neu1,Neu2} and gave excellent results both
306: shortening the computation time by giving a good trial function on the
307: finest grid and allowing a considerable improvement of the eigenvalue upon
308: extrapolation. Unfortunately these advantages diminish even in their case as
309: we go to higher states, for which only the finest grid is acceptable (finer
310: ones would yield too large matrices). For this reason we have not
311: implemented these procedures for finite elements at this point.
312:
313: Now we return with the problem of establishing the grid that defines our
314: finite elements and of calculating the matrix elements of $A$ and $B$. For
315: this we use a standard method\cite{Strang&Fix,Schwarz,Bethe&Wilson,Mori}
316: summarised in the following steps:
317:
318: \begin{enumerate}
319: \item We immerse the region $R$ in a quadratic grid. The triangles of the %
320: elements are defined using the sides of the squares and in addition one of %
321: the diagonals.
322:
323: \item We redefine the grid and triangles along the boundary by considering %
324: all points that lie just outside our contour. We then move the exterior %
325: points along the edges of the squares or the selected diagonal, to the %
326: boundary, so as not to change the topology of the grid and its connections.
327: \end{enumerate}
328:
329:
330: The corresponding integrals are evaluated using a
331: change of variable with a linear transformation. The transformation can be
332: found solving a linear system of 3 equations. Evaluating the constants of
333: Eq.~(\ref{functions}) and the Jacobian of the transformation (in this case
334: the one half area of the triangle, because the transformation is
335: linear) we obtain for the integrals:
336:
337: \begin{equation}
338: A_{ij}=\sum_kS(\Delta _k)(a_i^ka_j^k+b_i^kb_j^k)
339: \end{equation}
340: and
341: \begin{equation}
342: B_{ij}=\left\{ \matrix{\ \sum_k2S(\Delta _k)\frac 1{24},i\ne j,\cr\cr
343: \sum_k2S(\Delta _k)\frac 1{12},i=j}\right.
344: \end{equation}
345: Here $S(\Delta _k)$ is the area of triangle $\Delta _k$ and the sum is made
346: over the number of triangles common to both elements $i$ and $j$.
347:
348: The routines consist of two main parts. The first computes all elements
349: different from zero. The elements are put in two matrices of $N\times 7$ to
350: use the sparseness of matrices. In order to retain the original positions of
351: each element we construct an index matrix. The second part solves the
352: generalized system of equations using the inverse power iteration or a
353: standard diagonalization freeware routine \cite{http1}, if the
354: dimensions of the matrices are small (less than $\approx 5000\times 5000$).
355: We obtain $\approx 500$ eigenvalues with error less than $5\%$ of the
356: average spacing between levels.
357:
358: The input for the first routine is a set of points localised on the border
359: of the region of interest and a list of intervals of this
360: enumeration in which we want Dirichlet boundary conditions. We assume
361: Neumann boundary conditions in the rest of the border.
362:
363: For the second routine the inputs are the
364: number of eigenvectors, the tolerance for the Gauss Seidel and inverse power
365: iterations, the constant $C$, and the superconvergence factor $\alpha $. The
366: output consists of two files with the corresponding eigenvalues and
367: eigenvectors.
368:
369: \begin{center}
370: {\bf IV Application to the acoustic stadium}
371: \end{center}
372: By way of example we apply the program to some particular two-dimensional
373: region $R$; such as the Bunimovich stadium which is completely chaotic in
374: classical mechanics. This geometry consists of two
375: semicircles of radius $r$, joined by two straight lines with length $2l$
376: \cite{Bunimovich}. We will apply pure Neumann boundary conditions. On
377: a grid of $\sim 3000$ points we obtain, by the inverse
378: power iteration method described above, $200$
379: eigenvectors and eigenvalues with a CPU time of approximately
380: $200$ $\sec $. on an ALPHA Work-Station.
381:
382: To analised a discrete sequence of eigenvalues we first define the cumulative
383: level density or staircase function
384: \begin{equation}
385: N(E)=\sum_{i=1}^N\Theta (E-E_i) \label{nde}
386: \end{equation}
387: and its derivative $\rho (E)$ (the level density),
388:
389: \begin{equation}
390: \rho (E)=\sum_{i=1}^N\delta (E-E_i).
391: \end{equation}
392: Here $\Theta $ and $\delta $ are the Heaviside and Dirac delta functions
393: respectively. The staircase function is usually divided in a smooth part
394: plus a fluctuating part:
395:
396: \begin{equation}
397: N(E)=\overline{N}(E)+N_{fluct}(E).
398: \end{equation}
399:
400: \begin{figure}[tbp]
401: {\hspace*{-0.5cm}\psfig{figure=figure1ab.eps,height=6cm}} \vspace*{.12in}
402: \caption{a) Cummulative level density for a quarter of
403: Bunimovich stadium with Neumann boundary conditions and discretised with
404: 2791 elements. The dashed curve correspond to the theoretical result given
405: by the Weyl formula and the dotted curve is the Weyl formula corrected by
406: the polynomial from the equations in finite elements for the square. b)
407: Cummulative level density for a $50\times 50$ square with periodic boundary
408: conditions. As reference the theoretical prediction for the square is also
409: plotted.} \label {fig:1}
410: \end{figure}
411:
412: The cumulative level density obtained by the finite element method for the
413: stadium with Neumann boundary conditions is plotted in Fig. 1a. For
414: comparison the average level density
415:
416: \begin{equation}
417: \overline{N}_{Weyl}(E)=(AE+P\sqrt{E}+\kappa ) \,
418: \end{equation}
419: obtained from the Weyl formula \cite{Gutzwiller} is depicted. Here $A$ is the
420: area of the billiard, $P$ is the length of its perimeter and $\kappa $ is a
421: constant that contains information on the topological nature of the billiard
422: and the curvature of its boundary. From this figure we can see that the finite
423: element method gives a systematic deviation of the eigenvalues\cite{BaezG}.
424: This systematic deviation in the cumulative level density is due to the
425: discretisation of the finite element. To show this we calculate in the
426: appendix the eigenvalues $\lambda _{n,m}$ for the equations in finite
427: elements for an $a\times a$ square with periodic boundary conditions. The
428: final equation is
429:
430: \begin{equation}
431: \lambda _{n,m}=\frac{4-2\left( \cos \left( k_x\right) +\cos \left(
432: k_y\right) \right) }{\frac 12+\frac 16\left( \cos \left( k_x\right) +\cos
433: \left( k_y\right) +\cos \left( k_x+k_y\right) \right) }\, \label{eqsinfe}
434: \end{equation}
435: where $k_x=\frac{n\pi }a$ and $k_y=\frac{m\pi }a$ are the wave number on the
436: $x$ and $y$ directions, $a$ is the size of the side and $n,m=0,\pm 1,\pm
437: 2,\ldots $
438:
439: In Fig. 1b we show the cumulative level density obtained from Eq. (\ref
440: {eqsinfe}) for a $50\times 50$ square. We also plotted the cumulative level
441: density for the exact square ($\propto k_x^2+k_y^2$). The one coming from
442: the diagonalisation is indistinguishable from the obtained by the Eq. (\ref
443: {eqsinfe}). The systematic deviation observed in the square with periodic
444: boundary conditions will be used to correct the Weyl formula for
445: arbirary-shaped billiards and arbitray boundary conditions. To do this we
446: calculate the difference between the fits for both cumulative level
447: densities --Eq. (\ref{eqsinfe})-- and the exact square. The polynomial for the
448: difference is $A(4.60055-13.75335E-10.44418E^2+0.45601E^3)/2500$, where $A$
449: is the area of the billiard.
450: If this polynomial is added to the Weyl prediction,
451: the resulting curve yields good agreement
452: up to $\sim 400$ eigenvalues in the stadium (See Fig. 1a).
453:
454: We shall study the fluctuation properties of the spectrum. In order to do
455: this for a typical sequence of eigenvalues, it is necessary to suppress the
456: secular variation. This ``unfolding'' of the levels is done through the
457: mapping
458:
459: \begin{equation}
460: E_i\longmapsto E_i^{\prime }=\overline{N}(E_i).
461: \end{equation}
462:
463: \begin{figure}[tbp]
464: {\hspace*{-.5cm}\psfig{figure=figure2.eps,height=6cm}} \vspace*{.12in}
465: \caption{Nearest-Neighbour spacing distribution for the
466: GOE (solid line), and for the Poisson sequence (dotted line). The one
467: obtained for the stadium (200 eigenvalues for each symmetry--NN,DD,ND,DN--
468: of a quarter of the Bunimovich stadium with Neumann boundary conditions) is
469: depicted by the histogram.} \label {fig:2}
470: \end{figure}
471:
472: The effect of such mapping on the original sequence is that the new sequence
473: has on the average a constant spacing equal to one. Although we should use the
474: Weyl formula corrected by the polynomial, we can use a polynomial fit
475: $\overline{N}(E)$ for $N(E)$ that takes into account the systematic deviation
476: due to the discretisation. We can then calculate different statistical
477: measures of the fluctuating part of the spectrum. The first statistic we
478: shall use is the nearest-neighbour spacing distribution $p(s)$ where
479: $s=E_{i+1}^{\prime }-E_i^{\prime },$ which gives information on the short
480: range correlations. The $p(s)$ for the stadium is given in Fig. 2. Notice
481: that it agrees with the values predicted by the gaussian orthogonal ensemble
482: (GOE) typical for chaotic systems. For completeness, the Poisson
483: case, typical for integrable systems, is also depicted. The spectrum of the
484: acoustic Bunimovich stadium shows $p(s\rightarrow 0)\rightarrow 0$. This
485: behaviour is called level repulsion. In fact the nearest-neighbour spacing
486: distribution agrees with
487:
488: \begin{equation}
489: p_{Wigner}(s)=\frac \pi 2s\exp (-\frac \pi 4s^2),\qquad s\geq 0 \, \label{pds}
490: \end{equation}
491: know in spectral statistics as the Wigner surmise, which is very close to
492: the GOE\ prediction.
493:
494: We can also define the $k$th-neighbour spacing distribution $p(k;s_k).$ Now
495: $s_k=E_{i+k+1}^{\prime }-E_i^{\prime }$, so that $p(s)=p(0;s_0\equiv s)$. It
496: is well known \cite{Brodyetal} that these distributions tend to a normal
497: distribution as $k$ grows. Since $\left\langle s_k\right\rangle =k+1$, the
498: only relevant parameter left is the width $\sigma (k)$ of the distribution.
499: In Fig. 3 we show $\sigma (k)$ for the stadium, GOE and Poisson cases.
500:
501: \begin{figure}[tbp]
502: {\hspace*{-.5cm}\psfig{figure=figure3.eps,height=6cm}} \vspace*{.12in}
503: \caption{The width $\sigma (k)$ of the $k$th-neighbour
504: spacing distribution $p(k;s_k)$ as a function of $k$ for the GOE (solid
505: line), the Poisson (dotted line) and the acoustic stadium (squares).} \label {fig:3}
506: \end{figure}
507:
508: The correlation coefficient between adjacent spacings is another short range
509: fluctuation measure. For the acoustic stadium we obtained $C=-0.26$ near to
510: the GOE value ($C_{GOE}=-0.27$) and far from the Poisson value
511: ($C_{Poisson}=0$).
512:
513: Another commonly used statistic is the number variance $\Sigma^2(L)$, defined
514: as the second moment of the number of levels $\nu (L)$ within an interval of
515: length $L$, and given for the stadium, GOE and the Poisson cases in Fig. 4.
516: For the stadium and GOE cases and for large $L$, $\Sigma^2(L)\approx \ln (L)$
517: indicating a very rigid sequence.
518:
519: \begin{figure}[tbp]
520: {\hspace*{-.5cm}\psfig{figure=figure4.eps,height=6cm}} \vspace*{.12in}
521: \caption{The number variance $\Sigma^2(L)$ as a function of the length $L$ of
522: the interval for the same cases as in Fig. 3.} \label {fig:4}
523: \end{figure}
524:
525: The number variance depends exclusively on the two-point function. We
526: consider further moments
527: \begin{equation}
528: \Sigma^k(L)=\left\langle \left[ \nu (L)-\left\langle \nu
529: (L)\right\rangle \right] ^k\right\rangle
530: \end{equation}
531: that depend on higher correlations. To emphasise the three- or four-point
532: properties\cite{BHP}, it is useful to consider the skewness
533:
534: \begin{equation}
535: \gamma _1(L)=\Sigma^3(L)\times \left[ \Sigma^2(L)\right]
536: ^{-3/2} \label{gamma1}
537: \end{equation}
538: and the excess
539: \begin{equation}
540: \gamma _2(rL)=\Sigma^4(L)\times \left[ \Sigma^2(L)\right]
541: ^{-2}-3. \label{gamma2}
542: \end{equation}
543:
544: The numerical values obtained for the stadium are shown in Figs. 5 and 6.
545:
546: \begin{figure}[tbp]
547: {\hspace*{-.5cm}\psfig{figure=figure5.eps,height=6cm}} \vspace*{.12in}
548: \caption{The skewness $\gamma _1(L)$ --Eq. (20)-- as a function of
549: the length $L$ of the interval for the same cases as in Fig. 3.} \label{fig:5}
550: \end{figure}
551:
552: \begin{figure}[tbp]
553: {\hspace*{-.5cm}\psfig{figure=figure6.eps,height=6cm}} \vspace*{.12in}
554: \caption{The excess $\gamma _2(L)$ --Eq. (21)-- as a function of the
555: length $L$ of the interval for the same cases as in Fig. 3.}
556: \label {fig:6}
557: \end{figure}
558:
559: In many instances the Fourier transform of the spectra has also proven
560: useful. It contains the same information as the spectrum itself. On the
561: other hand the power spectrum:
562:
563: \begin{equation}
564: \left| C(t)\right| ^2=\frac 1{2\pi }\left| \int\limits_{-\infty }^\infty
565: e^{-2i\pi Et}\rho (E)dE\right| ^2, \label{pow}
566: \end{equation}
567: depends exclusively on the two-point function\cite{Lombardietal}. The
568: short-range part of $P(t)$ gives specific information concerning the
569: long-range stiffness. Numerical values are given in Fig. 7.
570:
571: \begin{figure}[tbp]
572: {\hspace*{-.5cm}\psfig{figure=figure7.eps,height=6cm}} \vspace*{.12in}
573: \caption{The power spectrum $\left| C(t)\right| ^2$ --Eq. (22)-- as a
574: function of the dimensionless variable $t$, for the same cases as in Fig. 3.
575: We eliminated the divergence at the origin due to the finite range of the
576: spectrum.} \label {fig:7}
577: \end{figure}
578:
579: \begin{figure}[tbp]
580: {\hspace*{-.5cm}\psfig{figure=figure8ab.eps,height=9cm}} \vspace*{.12in}
581: \caption{Eigenfunctions of Bunimovich stadium calculated by finite element
582: method for a quadrant and reflecting with respect to both axis: a) Dirichlet
583: boundary conditions with a relation $l/r = 1/2$. This figure show a
584: whispering gallery state; b) Neumann boundary conditions with
585: $l/r=1$. Typical state scarred by a near ``bouncing-ball'' orbit.}
586: \label{fig:8}
587: \end{figure}
588:
589: The eigenfunctions for the rectangle were also successfully compared with
590: the exact solution. In Fig. 8 we show two eigenfunctions of the stadium: one
591: of them with Dirichlet boundary conditions which shows typical feature of the
592: whispering gallery states\cite{McDonald&Kauffman}, and the last one, with
593: Neumann boundary conditions shows a near bouncing ball state. There are other
594: kinds of features in different eigenfunctions, like scars
595: \cite{Heller1,Heller2}, and others that resemble noise \cite{Berry1,Berry2}.
596: Some which have been reproduced by the autors with this algorithm can be seen
597: in ref. \cite{Arcosetal}.
598:
599: Finally, we performed all the calculation on a quarter stadium and
600: used Neumann (N) or Dirichlet (D) boundary conditions on the symmetry
601: lines. This implies that we studied the symmetric or antisymmetric solutions
602: with respect to both reflection symmetries of the stadium. The full solution,
603: shown in the figures, is recovered making the corresponding reflections respect
604: each symmetry axis.
605:
606:
607: \begin{center}
608: {\bf V Conclusions}
609: \end{center}
610:
611: We have presented an algorithm based on the finite element method that
612: computes eigenfunctions and eigenvalues of the two-dimensional Helmholtz
613: equation with mixed Neumann and Dirichlet boundary conditions. The algorithm
614: is divided in two parts: one that computes the matrix elements and another
615: that diagonalizes the generalised eigenvalue equation using the inverse
616: power and Gauss-Seidel Methods. The Gauss-Seidel method runs efficiently if
617: we use over-relaxation where the gain in computing time peaks at a factor of
618: $\approx 3$ around the value 1.75 for the superconvergence
619: (over-relaxation). A systematic error in frequencies was found. This error
620: is due to the discretisation and can be estimated by using the eigenvalues
621: of the equations in finite elements. The programs were used to calculate the
622: eigenvalues of the acoustic stadium. The fluctuation measures of the
623: eigenvalues were compared with the random matrix predictions.
624:
625: The algorithm is very useful in diverse branches of wave physics. The
626: program can obtain eigenfunctions and eigenvalues of two-dimensional
627: acoustic cavities, two-dimensional microwave cavities, membranes, quantum
628: billiards, elastic valleys in certain approximations, etc., and can be
629: readily generalised to other problems.
630:
631: \begin{center}
632: {\bf Acknowledgements}
633: \end{center}
634:
635: This work was supported by the UNAM---CRAY Research Inc. project SC101094,
636: by UNAM-DGAPA project IN106894. G. B\'{a}ez and R. A. M\'{e}ndez received
637: fellowships by UNAM-DGAPA. We want to thank to the IF-UNAM in which, the main
638: part of this work was developped.
639:
640: \newpage
641: \begin{center}
642: {\bf Appendix: Eigenvalues for the equations in finite elements.}
643: \end{center}
644:
645: In this appendix we deduce the equations in finite elements for a square
646: with periodic boundary conditions. If we denote by
647:
648: \begin{equation}
649: \text{\bf x}=e^{i(k_xn+k_ym)} \label{periodic}
650: \end{equation}
651: the amplitude in the grid point $(n,m)$, the Eq. (\ref{generalised}) for the
652: bulk element is given by
653: \begin{eqnarray}
654: &&(4 -e^{ik_x} -e^{ik_y} -e^{-ik_x}
655: -e^{-ik_y})\text{\bf x} \nonumber \\
656: &&= \lambda _{n,m} \Big( \frac 12 +\frac 1{12}e^{ik_x} +
657: \frac 1{12}e^{ik_y} +\frac 1{12}e^{-ik_x} +
658: \frac 1{12}e^{-ik_y} \\
659: && + \frac 1{12}e^{i(k_x+k_y)}
660: +\frac 1{12}e^{-i(k_x+k_y)} \Big) \text{\bf x.} \nonumber
661: \end{eqnarray}
662: Here $k_x=\frac{n\pi }a$ and $k_y=\frac{m\pi }a$. We also assumed that the
663: size of the grid is one and that for the bulk $A_{i,i}=4,$ $A_{i,j\pm
664: 1}=A_{i\pm 1,j}=-1$, $B_{i,j\pm 1}=B_{i\pm1,j}=B_{i\pm 1,j\pm 1}=\frac 1{12}$
665: and $B_{i,i}=\frac 12$. The final form for the eigenvalue
666: equation in finite elements is given by
667:
668: \begin{eqnarray}
669: &&4-2\left( \cos (k_x)+\cos (k_y)\right) \nonumber \\
670: &&=\lambda _{n,m}\left( \frac 12+\frac 16\left( \cos (k_x)+\cos (k_y)+
671: \cos (k_x+k_y)\right) \right) .
672: \end{eqnarray}
673:
674: \begin{references}
675:
676:
677: \bibitem{Neu1} J. M. Neuberger and D. W. Noid.(1984). {\sl Chem. Phys.
678: Lett.} {\bf 104} 1.
679:
680: \bibitem{Neu2} J. M. Neuberger and D. W. Noid.(1984). {\sl Chem. Phys.
681: Lett.} {\bf 112} 393.
682:
683: \bibitem{Neu3} B. Neuberger. (private communication).
684:
685: \bibitem{Novaroetal} O. Novaro, T. H. Seligman, J. M. \'{A}lvarez-Tostado,
686: J. L. Mateos, \& J. Flores. (1990). {\sl Bull. Seism. Soc. Am., {\bf 80%
687: }}, 239.
688:
689: \bibitem{Mateosetal} J. L. Mateos, J. Flores, O. Novaro, T. H. Seligman and
690: J. M. \'{A}lvarez-Tostado (1992). {\sl Geophys. J. Int. {\bf 113}},
691: 449.
692:
693: \bibitem{BaezG} G. B\'{a}ez, ``Caos Ac\'{u}stico'', Thesis Facultad de
694: Ciencias, University of Mexico (1993).
695:
696: \bibitem{MendezR} R. A. M\'{e}ndez-S\'{a}nchez, ``C\'{a}lculo de modos
697: resonantes en cavidades arbitrarias bidimensionales'', Thesis, Facultad de
698: Ciencias, University of Mexico (1992).
699:
700: \bibitem{Baezetal} J. B\'{a}ez, Flores, J., \& Seligman, T.\ H.''A comment
701: on the spectral rigidity of chaotic systems''. Proceedings of the IV Wigner
702: Symposium. Word Scientific 1996.
703:
704: \bibitem{BGS1} O. Bohigas, M. J. Giannoni and C. Schmit, {\sl Phys. Rev.
705: Lett.} {\bf 52}, 1-4 (1984). See also O. Bohigas, M. J. Giannoni and C.
706: Schmit, ``Spectral fluctuations of classically chaotic quantum systems'' in
707: T.H. Seligman and H. Nishioka (eds.) {\it Quantum Chaos and Statistical
708: Nuclear Physics. }(Springer, Berlin-Heidelberg-New York, 1986). pp. 18-40.
709:
710: \bibitem{Berry} M. V. Berry. ``Semiclassical mechanics of regular and
711: irregular motion'' in {\it Chaotic Behavior of Deterministic Systems},
712: Les Huoches Lectures XXXVI; R. H. Helleman and G. Iooss Eds. North Holland
713: Amsterdam (1985), pp.171-271.
714:
715: \bibitem{Arcosetal} E. Arcos, G. B\'{a}ez, P. Cuatl\'{a}yol, H.
716: Hern\'{a}ndez-Salda\~{n}a, M.L. Hern\'{a}ndez-Prian and R. A.
717: M\'{e}ndez-S\'{a}nchez. {\sl Am. J. Phys.}, {\bf 66}(7), 601, 1998.
718:
719: \bibitem{Sridhar} S. Sridhar. {\sl Phys. Rev. Lett.} {\bf 67},
720: 785-788 (1991).
721:
722: \bibitem{Richteretal} H. -D. Gr\"{a}f, H. L. Harney, H. Lengeler, C. H.
723: Lewenkopf, C. Rangacharyulu, A. Richter, P. Schard and H. A.
724: Weindem\"{u}ller. {\sl Phys. Rev. Lett.} {\bf 69}, 1296 (1992).
725:
726:
727: \bibitem{Stockmann&Stein} H. -J. St\"{o}ckmann, J. Stein, and M. Kollmann%
728: {\it ``Microwave studies in irregularly shaped billiards''} in: Quantum
729: Chaos, edited by G. Casati and B. Chirikov Cambridge University Press (1995).
730:
731: \bibitem{Metha} M. L. Mehta. {\it Random Matrices}. Revised and Enlarged
732: 2nd Ed. ({\sl Academic Press. San Diego, Cal., 1991}).
733:
734: \bibitem{Brodyetal} T. A. Brody, J. Flores, J. B. French, P.A. Mello, A.
735: Pandey and S. S. M. Wong, {\sl Rev. Mod. Phys.} {\bf 53} (1981) 385.
736:
737: \bibitem{Stein&Stockmann} J. Stein, and H.-J. St\"{o}ckmann. {\sl Phys.
738: Rev. Lett.} {\bf 68}, 2867 (1992).
739:
740: \bibitem{Legrandetal} O. Legrand, C. Schmit and D. Sornette. {\sl %
741: Europhys. Lett.}, {\bf 18} (2) 101-106(1992).
742:
743: \bibitem{Guhr} C. Ellegaard, T. Guhr, K. Lindemann, J. Nyg\aa rd and M.
744: Oxborrow {\it ``Symmetry breaking and Acoustic Chaos}'' in: Proceedings of
745: The IV Wigner Symposium, edited by N. Atakishiyev, T. H. Seligman and K. B.
746: Wolf, 1996.
747:
748: \bibitem{Guhr2} C. Ellegaard, T. Guhr, K. Lindemann, H. Q. Lorensen, J.
749: Nyg\aa rd and M. Oxborrow. {\sl Phys. Rev. Lett. }{\bf 75} (1995)1546.
750:
751: \bibitem{Weaver} R. L. Weaver. {\sl J. Acoust. Soc. Am.} {\bf 85}
752: (1989)1005.
753:
754: \bibitem{Marcusetal} L. P. Kouwenhoven, C. M. Marcus, P. L. McEuen, S.
755: Tarucha, R. M. Westervelt and N. S. Wingreen, {\it ``Electron Transport
756: in Quantum Dots'' Nato ASI conference proceedings, }ed. by L. P.
757: Kouwenhoven, G. Sch\"{o}n and L.L. Sohn (Klewer, 1997).
758:
759: \bibitem{Crommieetal} M. F. Crommie, C. P. Lutz, D. M. Eigler. {\sl %
760: Phys. Today} {\bf 46}, 17, Nov. 1993.
761:
762: \bibitem{Crommieetal&Heller} M. F. Crommie, C. P. Lutz, D. M. Eigler and E.
763: J. Heller. {\sl Surf. Rev. Lett.} {\bf 2(1)}, 127 (1995).
764:
765: \bibitem{Blumeletal} R. Bl\"{u}mel, I. H. Davidson, W. P. Reinhart, H. Lin
766: and M. Sharnoff. {\sl Phys. Rev. A} {\bf 45}, 2641-2644 (1992).
767:
768: \bibitem{Ottetal} L. Couchman, E. Ott and T. M. Antonsen Jr., {\sl Phys.
769: Rev. A.} {\bf 46}, 6193(1992).
770:
771: \bibitem{Kochetal} L. Sirko, P. M.Koch and R. Bl\"{u}mel. {\sl Phys.
772: Rev. Lett.}{\bf 78}, 2940(1997)
773:
774: \bibitem{Strang&Fix} G. Strang and G. Fix (1973). {\it An Analysis of
775: the Finite Element Method}. Prentice-Hall, Englewood Cliffs, New Jersey.
776:
777: \bibitem{Schwarz} H. R. Schwarz, (1988). {\it Finite Element Methods},
778: Academic Press, San Diego, CA.
779:
780: \bibitem{Bethe&Wilson} K. J. Bethe and E. L. Wilson (1976). {\it Numerical
781: Methods in Finite Element Analysis}, Prentice Hall, Inc. Englewood
782: Cliffs, New Jersey.
783:
784: \bibitem{Mori} M. Mori. {\it The Finite Element Method and Its
785: Applications}. Ed. by McMillan. Publishing Company, 1986.
786:
787:
788: \bibitem{BHP} O. Bohigas, R. U. Haq and A. Pandey, {\sl Phys. Rev. Lett.}
789: {\bf 54} (1984) 1645.
790:
791: \bibitem{Lombardietal} R. Jost, and M. Lombardi (1986) in {\it Quantum
792: Chaos and Statistical Nuclear Physics}. T. H. Seligman and H. Nishioka, eds.
793: Lecture Notes in Physics, 263. (Springer-Verlag, Berlin Heidelberg, 1986). pp.
794: 72.
795:
796: \bibitem{McDonald&Kauffman} S. W. McDonald and A. N. Kaufman. (1988).
797: {\sl Phys. Rev. A {\bf 37}}, 3067.
798:
799: \bibitem{Heller1} E. J. Heller. (1984). {\sl Phy. Rev. Lett., {\bf 53}}, 1515.
800:
801: \bibitem{Heller2} E. J. Heller. (1986), in: T. H. Seligman and H. Nishioka
802: (Eds.) ``{\it Quantum Chaos and Statistical Nuclear Physics}''. Springer,
803: Berlin-Heidelberg-New York.
804:
805: \bibitem{Berry1} M. V. Berry. (1977). {\sl J. Phys. A {\bf 10}}, 2083.
806:
807: \bibitem{Berry2} M. V. Berry. (1977). {\sl Philos. Trans. R. Soc. A
808: {\bf 287}}, 237.
809:
810: \bibitem{http1} http://math.nist.gov
811:
812: \bibitem{Bunimovich} L. A. Bunimovich. (1979). {\sl Commun. Math. Phys. }
813: {\bf 65}, 295.
814:
815: \bibitem{Gutzwiller} Gutzwiller Martin, C. (1990).{\it Chaos in Classical and
816: Quantum Mechanics}. (Springer-Verlag, New York Inc. 1990). pp. 259.
817: \end{references}
818:
819:
820: \end{document}