1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %%% (December 2003)
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4:
5: \documentclass{iopart}
6: \usepackage{epsfig}
7: \usepackage{latexsym}
8: \begin{document}
9:
10: \newcommand{\tbox}[1]{\mbox{\tiny #1}}
11: \newcommand{\half}{\mbox{\small $\frac{1}{2}$}}
12: \newcommand{\mbf}[1]{{\mathbf #1}}
13: \newcommand{\mpg}[3][b]{\begin{minipage}[#1]{#2}{#3}\end{minipage}}
14:
15: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
16:
17: \title{Consolidating boundary methods
18: for finding the eigenstates of billiards}
19:
20: \author{Doron Cohen$^{1}$, Natasha Lepore$^{2}$ and Eric J. Heller$^{2,3}$}
21:
22: \address{
23: \mbox{$^1$ Department of Physics, Ben-Gurion University,
24: Beer-Sheva, Israel.} \\
25: \mbox{$^2$ Department of Physics, Harvard University,
26: Cambridge, Massachusetts.} \\
27: \mbox{$^3$ Department of Chemistry and Chemical Biology,
28: Harvard University, Cambridge, Massachusetts.} \\
29: }
30:
31:
32: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
33:
34: \begin{abstract}
35: The plane-wave decomposition method (PWDM), a widely used
36: means of numerically finding eigenstates of the Helmholtz equation in
37: billiard systems is described as a variant of the mathematically
38: well-established boundary integral method (BIM). A new unified
39: framework encompassing the two methods is discussed. Furthermore,
40: a third numerical method,
41: which we call the Gauge Freedom Method (GFM) is derived from the BIM
42: equations. This opens the way to further improvements in eigenstate
43: search techniques.
44: \end{abstract}
45:
46: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
47:
48: %\maketitle
49: \vspace*{-0.5cm}
50:
51: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
52: \section{Introduction}
53:
54: Solving the Helmholtz equation within a domain
55: given Dirichlet boundary conditions
56: is of great interest to both physicists \cite{sto} and
57: engineers. Firstly, the Helmholtz equation is the simplest
58: example of a \mbox{\em wave equation}. Furthermore, this
59: equation may be used to describe acoustics waves, microwave
60: systems, and in particular the wavefunction
61: of a quantal particle inside nano-scale devices \cite{datta}
62: such as quantum-dots, where the motion of the electrons
63: can be regarded as a free~motion within a box. For this reason
64: it has become a prototype problem in studies of quantum chaos.
65:
66: Of particular interest are the wavefunctions $\Psi(x)$ of a
67: stationary particle in a two-dimensional box (a so-called
68: billiard system). These wavefunctions are
69: solutions of the homogeneous Helmholtz equation
70: \mbox{${\cal H}\Psi(x) = 0$}, where
71: the differential operator ${\cal H}$ is defined as
72: %
73: \begin{eqnarray} \label{e1}
74: {\cal H} = -\nabla^2 - k^2.
75: \end{eqnarray}
76: %
77: Note that for the special case $k=0$, the Helmholtz equation
78: reduces to Laplace's equation. Given a closed boundary
79: we can ask whether this equation has a non-trivial solution
80: that satisfies Dirichlet boundary conditions $\Psi(x)=0$.
81:
82: Two main numerical strategies have been suggested to date in the
83: literature in order to find the eigenstates of the Helmholtz equation
84: (for more comprehensive reviews and references, see
85: for example \cite{kuttler,backer,barnett}).
86: The first strategy can be described as a `Laplacian diagonalization'. A basis
87: is selected such that the functions it contains satisfy the Dirichlet
88: boundary conditions. For example, in some cases one can use conformal
89: mapping to determine a basis \cite{conformal} (and see also
90: \cite{reichl}). The Laplacian operator is then written in this
91: basis and diagonalized. Numerically, some truncation
92: is required, and the diagonalization only determines all the
93: eigenstates up to some maximum wavenumber $k_{\tbox{max}}$.
94: Thus, the Laplacian diagonalization strategy is inherently limited,
95: and can not be used for the purpose of finding high-lying eigenstates.
96:
97:
98: The second numerical strategy, which is the object of this paper, can be
99: described as a `boundary approach'. This strategy is based on the
100: observation that the eigenfunctions are completely determined by their
101: behavior at the boundary.
102: The boundary methods use basis functions that satisfy the Helmholtz equation
103: inside the billiard at fixed $k$. A linear combination of
104: the basis functions is then
105: selected such that the boundary conditions are satisfied.
106: Thus, in order to find the eigenstates, one only needs to study the
107: small $k$~window that contains the energy range of interest.
108: Therefore the method is naturally suitable for the purpose of
109: finding high-lying eigenstates.
110: For 2D~billiards, the Laplacian diagonalization
111: requires 2D~grid calculations. This is a heavy numerical task. The
112: boundary approach, on the other hand reduces the calculations to a
113: 1D~boundary grid.
114:
115:
116: In the quantum chaos community, two boundary methods are commonly
117: employed. The first one is referred to as the boundary integral method
118: (BIM) \cite{berry1}, while the other is what we call here
119: the decomposition method (DEM), of which the
120: plane-wave decomposition method (PWDM) \cite{heller2}
121: is a special case. Extensions of the standard PWDM have
122: been used in \cite{vergini,VS} and in \cite{barnett}.
123:
124: Usually, the BIM and the PWDM are considered to be two
125: independent self-contained procedures. Several studies
126: have been done in order to compare their capabilities \cite{li}.
127: While the BIM equation is exact, its convergence is very slow
128: (power law in the number $b$ of discretization points per half-wavelength).
129: On the other hand while the PWDM is mathematically
130: limited (e.g. the maximal $b$ is semiclassically determined),
131: it is still found to be extremely efficient in practice.
132: Hence there is definitely a need to develop hybrid boundary methods.
133:
134:
135: % However, Li and coworkers plotted the
136: % PWDM tension versus shape for several billiards with a shape parameters
137: % that varied the dynamics from integrable to chaotic. They
138: % confirmed that the accuracy of PWDM is reduced according to
139: % the amount of chaos in the system. Vega {\it et al.} \cite{vega}
140: % computed the number of states generated by the PWDM in polygonal
141: % billiards, and found that a significant fraction of these were missing
142: % when the results were compared to expected numbers given by Weyl's law.
143: % Finally, for non-convex shapes, the BIM is plagued with
144: % exterior chords while the PWDM fails completely (see\cite{li, boris}).
145:
146: In the present paper, we adopt a new point of view through
147: which we regard the BIM and the DEM as sequences of
148: four independent steps. By doing so, we are going to make the observation
149: that the DEM and the BIM are strongly related: {\em The two procedures
150: are based on the diagonalization of literally the same matrix!}
151: As a bridge between them, we will highlight an intermediate strategy
152: which we call the gauge-freedom method (GFM). In a follow-up paper,
153: this framework will lead the way to improved eigenstate
154: search techniques combining the strengths of the two boundary
155: methods \cite{bmi}.
156:
157: Our unified description of the different boundary methods
158: can be summarized by the following set of four steps that are
159: common to the BIM and the DEM, and as we will show later, to the GFM:
160: %
161: %\begin{minipage}{10cm}
162: \begin{itemize}
163: \item Choice of a set of basis functions $F_j(x;k)$.
164: \item Definition of the Fredholm matrix $\mbf{A}_{js}(k)$.
165: \item Procedure for construction of the wavefunction $\Psi_r$.
166: \item Definition of the quantization measure $S(k)$.
167: \end{itemize}
168: %\end{minipage}
169:
170:
171: The first step consists of selecting a set of basis
172: functions $F_j(x;k)$ labeled $j=1..N$.
173: All boundary methods rely on basis functions that
174: satisfy the Helmholtz equation {\em inside} the billiard.
175: Thus, a superposition of such basis functions is an
176: eigenfunction if it vanishes along the boundary.
177: The choices of bases that correspond to the PWDM,
178: to the primitive version of the BIM and to the simplest
179: variation of the GFM are as follows:
180: %
181: \begin{eqnarray} \label{e2}
182: F_j(x;k) \ \ =& \ \ \cos(\phi_j + k {n}_j \cdot {x})
183: \ \ & \ \ \ \ \mbox{PWDM}
184: \\ \label{e3}
185: F_j(x;k) \ \ =& \ \ Y_0(k|{x}-{x}_j|)
186: \ \ & \ \ \ \ \mbox{Y0-BIM}
187: \\ \label{e4}
188: F_j(x;k) \ \ =& \ \ J_0(k|{x}-{x}_j|)
189: \ \ & \ \ \ \ \mbox{J0-GFM}
190: \end{eqnarray}
191:
192:
193: For the purpose of the numerical treatment we represent
194: the boundary by a set of points $x_s$ with $s = 1\cdots M$.
195: In practice, we choose a set of $M$ equally spaced points,
196: so that the spacing is $\Delta s = L/M$ where $L$ is the
197: perimeter of the billiard. Depending on details of
198: the numerical strategy,
199: the number of points along the boundary is either taken to
200: be equal to the number of basis functions ($M=N$),
201: or it may be larger ($M>N$). The Fredholm matrix is defined as
202: %
203: \begin{eqnarray} \label{e5}
204: \mbf{A}_{js}(k) \ \ \equiv \ \ F_j(x_s;k)
205: \end{eqnarray}
206: %
207: Given $k$, one may perform the singular value decomposition (SVD)
208: of the matrix $\mbf{A}$.
209: The smallest singular value is the
210: one which we care about. If it is a minimum at a given $k$,
211: then the billiard system is likely to have an eigenvalue at that energy.
212:
213:
214: In the third step, the left and right
215: eigenvectors of the smallest singular value
216: ($\Phi_s$ and ${\mathsf C}_j$, resp.)
217: are used to construct a wavefunction $\Psi_r$ through a linear
218: transformation. We select a grid of points $X_r$ on which
219: the wavefunction $\Psi_r \equiv \Psi(X_r)$ is calculated.
220: In the DEM, the left eigenvector ${\mathsf C}$ is used for
221: the purpose of this construction, and the linear transformation
222: which is applied is:
223: %
224: \begin{eqnarray} \label{e6}
225: \Psi_r = \sum_j {\mathsf C}_j \mbf{F}_{jr}
226: \end{eqnarray}
227: %
228: where $\mbf{F}_{jr} \equiv F_j(X_r;k)$.
229: Note that ${\mathsf C}$ contains the expansion coefficients
230: of $\Psi(x)$ in the chosen basis $F_j(x;k)$.
231: %
232: For the BIM, the right eigenvector $\Phi$
233: is used in order to build the wavefunction,
234: and the linear transformation in this case is
235: %
236: \begin{eqnarray} \label{e7}
237: \Psi_r = \sum_s \mbf{G}_{rs} \Phi_s,
238: \end{eqnarray}
239: %
240: where $\mbf{G}_{rs}$ is the discretized version of
241: the Green function. Thus, the vector $\Phi_s$
242: represents a `charge' that is distributed along the boundary.
243:
244: In the final step, a measure $S(k)$ is defined such that
245: $S(k)=0$ if $k$ is an eigenvalue and $S(k)>0$ otherwise.
246: In practice, the eigenvalues are determined by searching for the local
247: minima of $S(k)$. By construction, the wavefunction which was built
248: in step three satisfies the Helmholtz equation inside the boundary.
249: Therefore, the most natural choice of $S(k)$ is the tension,
250: the sum of the square of the wavefunction along the boundary.
251: The tension is thus a
252: measure for the roughness of the constructed $\Psi(x)$
253: along the boundary. This definition of $S(k)$ is
254: traditionally used with the PWDM.
255: Other possibilities for the measure include
256: the smallest singular value, and the Fredholm
257: determinant of $\mbf{A}$. These two latter choices
258: of $S(k)$ are the ones that are usually associated with the BIM.
259: In Sec.~\ref{IV} we discuss the mathematical equivalence
260: of the three possible measures, and compare
261: their respective numerical effectiveness.
262:
263:
264:
265: %%%%%%%%%%%%%%%%%%%%%%
266: \subsection{Outline}
267:
268: In Sec.~\ref{II}, we give a concise presentation
269: of the BIM and the related GFM. Our derivation of
270: the BIM equation contains some significant
271: improvements over previous ones. Most
272: importantly, it naturally leads to the
273: existence the GFM.
274: Furthermore, we have succeeded in avoiding the use
275: of the complicated "regularized" method of images,
276: which was the major ingredient in the derivation
277: of Ref.~\cite{li}.
278:
279: Strategies for constructing the wavefunction
280: are discussed in Sec.~\ref{III}. An explanation of
281: the Green function method is given,
282: as well as a critical discussion of the DEM and
283: its numerical variants.
284:
285: Sec.~\ref{IV} explores the practicality of using different
286: choices for the quantization measure. In particular,
287: it is demonstrated that a tension measure can be defined
288: not only for the PWDM, but for the case of the BIM as well.
289: An important issue emerges as to whether the
290: quantization measures can be used as to determine
291: the error in the bulk wavefunction.
292: We address this issue, and also make a comparison
293: between the numerical accuracies of the BIM and of the PWDM.
294:
295: Sec.\ref{V} explains how the GFM bridges between
296: the BIM and the DEM. It is found that for
297: any DEM, an associate GFM exists, whereas the
298: inverse statement is not true.
299:
300:
301: The shape that we have studied numerically is presented in Fig.1.
302: We have used the cornerless, generic 'Pond' shape in order to avoid
303: the range of problems that arise with more complicated geometries.
304: These problems are the subject of a follow-up study \cite{bmi},
305: where we suggest mixed BIM/DEM methods for finding eigenfunctions.
306: This is done using the above theoretical framework,
307: while regarding the 'Pond' shape as a reference against which to judge
308: the effectiveness of our efforts. Another direction of
309: research is related to billiards in magnetic fields \cite{klaus}.
310:
311: For the convenience of the reader, our
312: numerical notations are concentrated in Table~1.
313: Further information about Fig.~1, Table~1,
314: and the numerical analysis is integrated within
315: the main text.
316:
317:
318:
319: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
320: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
321: \begin{table}
322: \begin{tabular}{|lll|}
323: %
324: \hline
325: \ & \ & \\
326: $x_s$ &=& vector of boundary points \hspace*{3cm} \\
327: $x_{{\mathsf s}}$ &=& vector of outer-boundary points \\
328: $X_r$ &=& vector of interior grid points \\
329: $X_0$ &=& randomly selected interior point \\
330: \ & \ & \\
331: $\Psi_r$ &=& wavefunction on the grid points \\
332: $\Psi_s$ &=& wavefunction on the boundary points \\
333: $\Phi_s$ &=& `charge' along the boundary \\
334: \ & \ & \\
335: $\| \Psi_r \|$ &=& norm of the wavefunction (see Sec.III) \\
336: $\| \Psi_s \|$ &=& tension of the wavefunction (see Sec.III) \\
337: \ & \ & \\
338: $n(s)$ &=& unit normal at the boundary point $x_s$ \\
339: \ & \ & \\
340: $\mbf{w}_{s}$ &=& $(1/(2k^2)) \ n(s) \cdot x_s $ \\
341: \ & \ & \\
342: $\mbf{G}_{rs}$ &=& $G(X_r,x_s)$ \\
343: \ & \ & \\
344: $\mbf{A}_{j0}$ &=& $F_j(X_0;k)$ \\
345: \ & \ & \\
346: $\mbf{A}_{js}$ &=& $F_j(x_s;k)$ \\
347: \ & \ & \\
348: $\mbf{D}_{js}$ &=& $\partial F_j(x_s;k)$ \\
349: \ & \ & \\
350: $\mbf{B}_{ij}$ &=& $\Delta s \ \sum_s \mbf{w}_s \mbf{D}_{is}\mbf{D}_{js}
351: \ \ = \ \
352: \Delta s \ (\mbf{D}\mbf{w}\mbf{D}^{\dag})_{ij}$ \\
353: \ & \ & \\
354: $\mbf{T}_{ij}$ &=& $\Delta s \ \sum_s \mbf{A}_{is} \mbf{A}_{js}
355: \ \ \ \ = \ \
356: \Delta s \ (\mbf{A} \mbf{A}^{\dag})_{ij}$ \\
357: \ & \ & \\
358: \hline
359: %
360: \end{tabular}
361: %
362: \caption{Notations}
363: \end{table}
364: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
365: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
366:
367:
368:
369:
370: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
371: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
372: \section{The BIM and the GFM}
373: \label{II}
374:
375:
376: The gist of the BIM is that,
377: from the knowledge of the gradient of the wavefunction
378: on the boundary and from Green's theorem, it is possible
379: to find the value of the wavefunction everywhere
380: inside the billiard. We give a derivation of this method
381: in this section. This procedure will lead us naturally
382: to the existence of the GFM.
383:
384: The free space Green function $G(x,x')$ is defined
385: by the equation ${\cal H}G(x,x')=\delta(x-x')$.
386: The most general solution can be written as
387: %
388: \begin{eqnarray} \label{e8}
389: G(x,x') \ \ = \ \ - \frac{1}{4} Y_0(k|x-x'|)
390: \ \ + \ \ {\cal C}(x,x')
391: \end{eqnarray}
392: %
393: where ${\cal C}(x,x')$ is any solution of
394: the homogeneous equation ${\cal H}{\cal C}(x,x')=0$.
395: Note that in the electrostatic limit $k\rightarrow0$
396: we have $G(x,x')=-(1/(2\pi))\ln(r) + {\cal C}$,
397: where ${\cal C}$ is a constant~\cite{rmrk2}.
398: We shall refer to the choice of ${\cal C}(x,x')$
399: as gauge freedom. This term is
400: at the core of the GFM.
401:
402: By the definition of the Green function, it follows
403: that a solution of the generalized Poisson-Laplace (GPL)
404: equation ${\cal H}\Psi(x) = \rho(x)$ is~\cite{rmrk3}
405: %
406: \begin{eqnarray} \label{e9}
407: \Psi(x) \ \ = \ \ \int G(x,x')\rho(x') dx'
408: \end{eqnarray}
409: %
410: We refer to $\rho(x)$ as the `charge density',
411: by analogy to its electrostatic equivalent.
412:
413: We shall use the notation $\Phi(s)$ in order to refer
414: to the (surface) charge density upon the boundary.
415: In the latter case, the equation above reduces to
416: %
417: \begin{eqnarray} \label{e10}
418: \Psi(x) \ \ = \ \ \oint G(x,x(s)) \Phi(s) ds
419: \end{eqnarray}
420: %
421: where $s$ parameterizes the boundary.
422:
423: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
424: \subsection{The BIM}
425:
426: Let us assume that $k$ is an eigenvalue of the billiard.
427: In such a case, there exists a non-vanishing $\Psi(x)$
428: inside the boundary that satisfies $\Psi(x)=0$ on the boundary.
429: It can be shown from Green's theorem that the interior
430: wavefunction satisfies Eq.(\ref{e10}) with
431: %
432: \begin{eqnarray} \label{e11}
433: \Phi(s) \ = \ \partial_{-} \Psi(x(s)) \ \equiv \
434: \lim_{x\uparrow x(s)} n(s)\cdot\nabla \Psi(x)
435: \end{eqnarray}
436: %
437: where $n(s)$ is the outward pointing normal at point $s$, and
438: $\partial_{-}$ is used for the normal derivative evaluated inside
439: of the billiard walls.
440:
441: In electrostatics, it is known that forcing the
442: scalar potential to be zero on the boundary induces
443: a boundary charge. From Green's theorem,
444: the induced charge is proportional to the normal
445: component of the electric field. Here the wavefunction
446: acts as the equivalent of the scalar potential. Similarly
447: to the electrostatic case, there exists an induced
448: `boundary charge', that is in this case proportional
449: to the normal derivative of the wavefunction.
450:
451: The BIM is based on the fact that if an eigenstate exists,
452: then there also exists a charge density $\Phi(s)$ given by
453: Eq.(\ref{e11}), such that Eq.(\ref{e10})
454: is satisfied. On the boundary, Eq.(\ref{e10}) yields
455: %
456: \begin{eqnarray} \label{e12}
457: \int G(x(j),x(s)) \Phi(s) ds \ \ = \ \ 0
458: \ \ \ \mbox{[BIM equation]}
459: \end{eqnarray}
460: %
461: Thus, having an eigenstate $\Psi(x)$ implies
462: that the kernel $G(x(j),x(s))$ has an eigenvector
463: $\Phi(s)$ that corresponds to a zero eigenvalue.
464: Fig.2 shows an example
465: of a boundary charge density $\Phi(s)$ that was found
466: via the BIM equation (for more details, see next section).
467: The converse
468: is also true: Once a non-trivial charge density is
469: found that satisfies Eq.(\ref{e12}), the associated
470: eigenstate can be constructed using Eq.(\ref{e10}).
471: We discuss this construction issue in more details
472: in the next section.
473:
474: For numerical purposes, it is convenient to use the
475: discretized version Eq.(\ref{e7}) of the above formula.
476: The BIM equation can then be
477: written as the matrix equation $\mbf{A}\Phi=0$,
478: where $\mbf{A}_{js} = G(x(j),x(s))$. The gauge
479: term ${\cal C}(x,x')$ allows some freedom in the
480: determination of the Green function.
481: Using the Neumann Bessel function $Y_0(k|x-x'|)$
482: for the Green function, one obtains the
483: matrix $\mbf{A}_{js}$ as defined
484: by Eq.(\ref{e5}) with~(\ref{e3}).
485: Another possibility is to use the Hankel Bessel
486: function $H_0(k|x-x'|)$. Accordingly, we
487: will distinguish between the Y0-BIM version
488: and the H0-BIM version. We shall later discuss
489: the numerical implication of using the
490: complex $H(k|x-x'|)$ rather than the real $Y(k|x-x'|)$.
491:
492: The primitive BIM uses Eq.(\ref{e12}) literally.
493: However, this version of the BIM is not the one that
494: is generally favored because $G(x(j),x(s))$
495: is singular for $x(j)\rightarrow x(s)$,
496: leading to some difficulty in determining the diagonal
497: matrix elements of $\mbf{A}_{js}$.
498: Therefore, other versions of the BIM
499: have become popular (see Appendices B,C).
500:
501:
502:
503:
504: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
505: \subsection{The GFM}
506:
507: The GFM is a different strategy to obtain the
508: charge density $\Phi(s)$. Rather than
509: using the BIM equation Eq.~(\ref{e12})
510: or one of its variants, a gauge freedom argument
511: is invoked in order to introduce
512: a new type of equation (Eq.(\ref{e13}) below).
513: %
514: It is clear that Eq.(\ref{e12}) should be
515: valid for {\em any} choice of gauge.
516: In other words, Eq.(\ref{e12}) should be
517: satisfied for any Green function (Eq.(\ref{e8})),
518: whatever the choice of ${\cal C}(x,x')$.
519: Thus, for a given ${\cal C}(x,x')$,
520: the charge density $\Phi(s)$ must satisfy
521: the equation
522: %
523: \begin{eqnarray} \label{e13}
524: \int {\cal C}(x(j),x(s)) \Phi(s) ds \ \ = \ \ 0
525: \ \ \ \mbox{[GFM equation]}
526: \end{eqnarray}
527: %
528: For example, we may take ${\cal C}(x,x') = J_0(k|x-x'|)$,
529: and we shall refer to this version of GFM as J0-GFM.
530: For numerical purposes, it is once again convenient to
531: discretize the integral expression, which can then be
532: written as the matrix equation $\mbf{A}\Phi=0$.
533:
534:
535: The kernel $\mbf{A}_{js} = J_0(k|x-x'|)$
536: of the J0-GFM is non-singular, and very well-behaved.
537: Thus, the J0-GFM method, unlike the Y0-BIM, provides
538: an extremely convenient way of obtaining the eigenvalues
539: of the Helmholtz equation. Fig.2 shows an example
540: of a charge density that was found via the J0-GFM equation
541: (for more details see next section).
542: The result is indistinguishable from the charge density
543: generated by the traditional H1-BIM.
544: [We note however that the J0-GFM method has certain
545: numerical limitations that we are going to discuss later].
546: %
547: Once the eigenvector $\Phi(s)$ is found via the GFM equation,
548: we can proceed as with the traditional BIM, and construct
549: the wavefunction $\Psi(x)$ using Eq.(\ref{e7}).
550:
551:
552:
553: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
554: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
555: \section{Constructing the wavefunction}
556: \label{III}
557:
558: In this section we explain how a wavefunction $\Psi(x)$
559: is constructed for a given $k$. It is assumed that $k$ is an
560: eigenvalue. The (numerical) question how to determine the
561: eigenvalues $k=k_n$ is differed to Section~4.
562:
563:
564: \subsection{Green function method (Eq.~(\ref{e7}))}
565: \label{III_A}
566:
567: Both the BIM and GFM make use of Eq.(\ref{e7}) in order
568: to construct the wavefunction. In order to find
569: the charge vector $\Phi_s$ the BIM equation (Eq.(\ref{e12}))
570: and the GFM equation (Eq.(\ref{e13}))
571: are written as the matrix equation \mbox{$\mbf{A}\Phi=0$}.
572: The only difference between the two
573: is in the expression for $\mbf{A}$.
574: Given $k$, one performs the SVD
575: of the matrix~$\mbf{A}$. Fig.4 displays an example
576: of the output of such a SVD procedure.
577: One then finds the right eigenvector $\Phi$ that
578: corresponds to the {\em smallest} singular value.
579:
580:
581: Once the charge vector $\Phi_s$ has been determined,
582: as in the example of Fig.2, one can construct
583: the wavefunction using Eq.(\ref{e7}).
584: For the Green function Eq.(\ref{e8}), it is
585: most natural to use the simplest gauge (${\cal C}=0$).
586: If $k$ is known to be an eigenvalue,
587: then any gauge should give the same result,
588: and in particular, the wavefunction
589: associated with any complex part of the Green
590: function (such as that of the Hankel function)
591: should vanish. The outcome of the Green function method
592: is illustrated in Fig.3.
593:
594:
595: It is natural to ask how the constructed wavefunction $\Psi(x)$
596: look like outside of the boundary. The answer turns out to be
597: $\Psi(x)=0$. For completeness, we give a proof
598: of this statement. Let us define an extended function
599: $\Psi_{\tbox{ex}}(x)$ such that
600: $\Psi_{\tbox{ex}}(x)=\Psi(x)$ inside
601: and $\Psi_{\tbox{ex}}(x)=0$ outside of the boundary.
602: We would like to show that $\Psi(x)$ as defined
603: by Eq.(\ref{e10}) is also equal to $\Psi_{\tbox{ex}}(x)$
604: outside of the boundary.
605: It is clear that $\Psi(x)$ is a solution of the GPL equation
606: by construction [see discussion following Eq.(\ref{e9})].
607: In the next paragraph, we are going to argue
608: that $\Psi_{\tbox{ex}}(x)$ is a solution of
609: the {\em same} GPL equation. It follows that
610: the difference $R(x)=\Psi(x)-\Psi_{\tbox{ex}}(x)$
611: is a solution of Helmholtz equation in free space.
612: From the definition of $\Psi_{\tbox{ex}}(x)$,
613: we have $R(x)=0$ in the interior region,
614: which implies by the unique continuation property
615: that $R(x)=0$ over all space.
616:
617: The proof that $\Psi_{\tbox{ex}}(x)$ is a solution of
618: the GPL equation with a charge density given by
619: Eq.(\ref{e11}) goes as follows: By construction,
620: $\Psi_{\tbox{ex}}(x)$ satisfies the GPL equation inside
621: as well as outside of the boundary. All we have to show
622: is that it also satisfies the GPL equation across
623: the boundary. The latter statement is most easily
624: established by invoking Gauss' law. This approach
625: is valid because at short distances, $G(x,x')$ coincides
626: with the electrostatic Green function.
627: Thus, the gradient of $\Psi_{\tbox{ex}}(x)$ corresponds,
628: up to a sign, to the electric field. Gauss' law implies
629: that the electric field should have a discontinuity
630: equal to the charge density $\Phi(s)$.
631: Indeed, $\Psi_{\tbox{ex}}(x)$ is consistent with
632: this requirement.
633:
634:
635: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
636: \subsection{Decomposition method (Eq.~(\ref{e6}))}
637: \label{IIIB}
638:
639:
640: The other procedure to construct the wavefunction
641: is to use the DEM Eq.(\ref{e6}). The idea is to regard $F_j(x;k)$
642: as a basis for the expansion:
643: %
644: \begin{eqnarray} \label{e16_0}
645: \Psi(x) \approx \sum_j {\mathsf C}_j F_j(x;k)
646: \end{eqnarray}
647: %
648: Any such superposition at fixed $k$ is a solution of
649: the Helmholtz equation within the interior region. Thus, in order
650: to satisfy the Dirichlet boundary conditions, one looks for a vector
651: ${\mathsf C}$ of expansion coefficients that
652: satisfy ${\mathsf C} \mbf{A} = 0$.
653: It turns out that the direct numerical implementation of
654: this simple idea is a complicated issue (see discussion of
655: the null space problem later in this section).
656:
657:
658: Any set of $F_j(x;k)$ which are solutions of the Helmholtz equation
659: may be used for the DEM. However, it should be remembered
660: that computationally not all bases are equivalent.
661: For instance, the $Y_0$ basis defined by Eq.(\ref{e3})), which might
662: appear to be the best choice as a DEM basis due to its association
663: with the BIM, does not give the best numerical results
664: when compared against other options.
665: In particular, it turns out that the PWDM is typically much more
666: effective (recall that the PWDM is a special case of
667: the DEM, corresponding to the choice Eq.(\ref{e2}) of basis functions.)
668: Finally we note that the set of $J_0$s of Eq.(\ref{e4})
669: can not be regarded as a mathematically legitimate basis
670: for a DEM. This latter point will be explained in Section~5.
671:
672:
673: For the $Y_0$ basis the BIM and the DEM lead to the same equation.
674: It is only the mathematical interpretations that is different.
675: Within the DEM, one regards the $Y_0$ as basis functions to be used
676: in an expansion, while the same $Y_0$ in the BIM context serves
677: as the Green function. In the context of DEM, one may be bothered
678: by the singular nature of the $Y_0$ functions: The constructed wavefunction
679: should be zero on the boundary. Mathematically this is achieved
680: in the $N\rightarrow\infty$ limit, so the $Y_0$ basis is a valid choice.
681: But in an actual numerical implementation, the wavefunction so constructed
682: will always have singularities on the boundary.
683: One possible remedy consists of enforcing the boundary conditions
684: on intermediate boundary points, or on points that lie
685: on an outer boundary. Alternatively, one may replace the
686: bare $Y_0$ basis by smooth superpositions of $Y_0$ functions (see Appendix C).
687:
688:
689:
690: In Fig.2, we display an example of a numerically determined
691: ${\mathsf C}$ (using the PWDM basis) for one of the Pond eigenstates,
692: while in Fig.~3 we illustrate the constructed wavefunction.
693: Unlike the Green function construction,
694: the DEM/PWDM constructed wavefunction does not
695: vanish outside of the boundary. Actually, it is
696: quite the opposite: Typically the DEM/PWDM wavefunction
697: becomes exponentially large as we go further away from
698: the boundary. Whenever this behaviors occurs,
699: it constitutes an indication
700: of the {\em evanescent} nature of the wavefunction.
701: Namely, in such cases, the wavefunction acquires
702: sub-wavelength features in order to
703: accommodate the boundary. This requires
704: exponential behavior (negative kinetic energy)
705: in the transverse space direction, in order to keep
706: the total energy fixed.
707:
708:
709:
710: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
711: \subsection{Normalization of the wavefunction}
712:
713:
714: A standard SVD procedure generates vectors $\Phi_s$ and ${\mathsf C}_j$
715: that are normalized in the sense $\sum_s |\Phi_s|^2 = 1$ and
716: $\sum_j |{\mathsf C}_j|^2 = 1$. Therefore, the constructed $\Psi_r$
717: is not properly normalized within the interior region.
718: Adopting the usual philosophy of boundary methods,
719: the problem of calculating the $\Psi_r$ normalization
720: is reduced to that of evaluating a boundary integral,
721: namely~\cite{berry1,boasman}
722: %
723: \begin{eqnarray} \label{e15}
724: \int\int |\Psi(x)|^2 dx =
725: \frac{1}{2k^2} \oint |\Phi(s)|^2 (n(s){\cdot}x(s))ds
726: \end{eqnarray}
727: %
728: %
729: %
730: For the BIM, by discretizing of Eq.~(\ref{e15})
731: we obtain the following numerical
732: expression for the normalization factor:
733: %
734: \begin{eqnarray} \label{e15a}
735: \| \Psi_r \| \ \ = \ \ \frac{1}{\Delta s}
736: \sum_s \mbf{w}_s (\Phi_s)^2
737: \ \ = \ \ \Phi^{\dag} \mbf{W} \Phi.
738: \end{eqnarray}
739: %
740: Here $\mbf{W}=(1/\Delta s)\mbox{diag}(\mbf{w}_s)$
741: is a diagonal matrix, and the weight factor
742: $\mbf{w}_s$ is defined in Table~1.
743: %
744: %
745: %
746: As for the DEM, by using the derivative of Eq.(\ref{e16_0})
747: in Eq.(\ref{e11}) and substituting into Eq.(\ref{e15}), we get:
748: %
749: \begin{eqnarray} \label{e16a}
750: \| \Psi_r \| \ &=& \ \Delta s \sum_s \mbf{w}_s
751: \left(\sum_j {\mathsf C}_j \mbf{D}_{js}\right)^2
752: \\ \label{e16aa}
753: \ &=& \ \sum_{ij} {\mathsf C}_i \mbf{B}_{ij} {\mathsf C}_j
754: \ = \ {\mathsf C}\mbf{B}{\mathsf C}^{\dag}
755: \end{eqnarray}
756: %
757: The definitions of $D_{js}$ and of the metric $B_{ij}$
758: can be found in Table 1.
759:
760: The normalization $\| \Psi_r \|$ can be calculated
761: using the metric $\mbf{B}_{ij}$. This method looks quite
762: elegant, but it turns out not to be very effective
763: numerically. Consider Eq.(\ref{e16a}). This
764: equation is quite safe computationally for two reasons:
765: (i) all its terms are non-negative;
766: (ii) standard summation routines order these terms
767: in descending order. Now, let us look instead at
768: Eq.(\ref{e16aa}). In this case the numerical calculation
769: can give {\em any} result (if we go
770: to large $k$). Sometimes, the answer even
771: comes out to be negative! This occurs because the calculation
772: involves many arbitrarily ordered terms that
773: each have a different algebraic sign.
774:
775:
776:
777:
778:
779:
780: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
781: \subsection{The tension along the boundary}
782:
783:
784: The numerical wavefunction $\Psi_r$ satisfies Helmholtz
785: equation in the interior region {\em by construction}.
786: Thus, whether $\Psi_r$ is an actual eigenstate depends
787: on its behavior along the boundary.
788: In this subsection we would like to discuss
789: the definition of a `tension' measure
790: that estimates whether, and to what accuracy,
791: the numerical $\Psi_r$ satisfies the boundary conditions.
792:
793:
794:
795: For the case of the DEM, following \cite{heller2},
796: the tension is defined as the boundary integral
797: %
798: \begin{eqnarray} \label{e16b}
799: \| \Psi_s \| \ &=& \ \Delta s \sum_s
800: \left(\sum_j {\mathsf C}_j \mbf{A}_{js}\right)^2
801: \ = \ \sum_{ij} {\mathsf C}_i \mbf{T}_{ij} {\mathsf C}_j
802: \ = \ {\mathsf C}\mbf{T}{\mathsf C}^{\dag}
803: \end{eqnarray}
804: %
805: The standard practice to date for the
806: tension calculation has been to
807: use a denser set of boundary points
808: located between the $x_s$ points.
809: However, our experience (see also \cite{barnett})
810: is that the tension estimate obtained from
811: the initial set of points is just as effective.
812: This is demonstrated in Fig.3d.
813: Therefore, we routinely rely on the same set
814: of boundary points to determine the tension.
815:
816:
817: For the BIM on the other hand, the above definition
818: is not practical due to the singular nature of the basis functions.
819: For any finite $M$, the numerical wavefunction
820: blows up at each boundary point.
821: However, since the BIM wavefunction vanishes everywhere
822: outside of the billiard, a numerically unambiguous definition
823: of tension arises as an integral of $|\Psi(x)|^2$ along
824: an outer boundary:
825: %
826: \begin{eqnarray} \label{e15b}
827: \| \Psi_ {{\mathsf s}}\| \ = \
828: \Delta s \ \sum_{{\mathsf s}} (\Psi_{{\mathsf s}})^2
829: \ = \ \Delta s \ \sum_{{\mathsf s}}
830: \left(\sum_s\mbf{G}_{{\mathsf s}s}\Phi_s \right)^2
831: \end{eqnarray}
832: %
833: By outer boundary (see Fig.1), we mean the set of
834: external points (${\mathsf s}$~points, as opposed to $s$~points
835: for the true boundary) that have a fixed
836: transverse distance $\Delta L$ from the the true boundary.
837: The distance $\Delta L$ between the boundary and the outer boundary
838: should be small on any classical scale but large compared with $ds$,
839: in order for the tension to be independent of the
840: choice of $\Delta L$. See Fig.5c for a numerical demonstration.
841:
842:
843:
844: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
845: \subsection{The PWDM and the null space problem}
846:
847: One may think that ${\mathsf C}$ could be found
848: simply by computing the (left) eigenvector
849: that corresponds to the smallest singular
850: value of $\mbf{A}_{js}$. Numerically this definition
851: is hard to implement. This difficulty can be explained
852: by looking at the behavior of the singular values
853: of $\mbf{A}_{js}$ for the PWDM basis.
854: Fig.4 gives some examples of singular values
855: resulting from the SVD of the $\mbf{A}_{js}$ matrix.
856: In the case of the PWDM, as $k$ become large, one observes
857: that the the singular values separate into two groups:
858: rather than having one distinctly smaller singular
859: value, we obtain a whole set of them.
860: Accordingly, we can define a numerical `null space'
861: of the $\mbf{A}_{js}$ matrix.
862:
863:
864: The interpretation of this null space is quite clear.
865: It is well known \cite{dietz,uzy} that it is not efficient
866: to include much more than $N_{sc}$
867: plane waves in the basis set $F_j(x;k)$, where
868: %
869: \begin{eqnarray} \label{e16c}
870: N_{sc} \ = \ \frac{1}{\pi} kL
871: \end{eqnarray}
872: %
873: and $L$ is the perimeter of the billiard.
874: The reason for this ineffectiveness is
875: that $k_i$ and $k_j$ cannot
876: be distinguished numerically within the interior
877: region unless \mbox{$|k_i-k_j| L > 1$}.
878: In order to obtain the semiclassical result (\ref{e16c}),
879: the total phase space area ($L\times(2mv)$)
880: of the boundary Poincar\'e section is divided by the size
881: of Planck cell ($2\pi\hbar$). If we use
882: $N>N_{sc}$ plane waves, then we can create
883: {\em wavefunctions that are nearly zero in the interior},
884: and become large only as we go far away from the center\cite{berry2}.
885: It is clear that the SVD can be used to determine
886: the $N-N_{sc}$ null space of these evanescent states.
887: Whenever $k$ is an eigenvalue, this null space includes,
888: in addition to the evanescent waves, the single eigenvector
889: that constitutes an eigenstate of the Helmholtz equation.
890: The problem is to distinguish this eigenvector from the other
891: vectors in the null-space.
892:
893:
894: The basic difference between the eigenvector that leads
895: to an eigenstate (which exists if $k=k_n$), and the
896: other vectors in the null space is related to the normalization.
897: As discussed before, a standard SVD procedure generates
898: vectors ${\mathsf C}_j$ that are normalized
899: in the sense $\sum_j |{\mathsf C}_j|^2 = 1$.
900: Therefore, the $\Psi_r$ of Eq.(\ref{e6}) is not properly normalized
901: within the interior region. Normalizing the wavefunction
902: has the effect of magnifying the evanescent solutions
903: in the interior as well as on the boundary,
904: while the eigenfunction (if it exists) remains
905: small on the boundary. In appendix D, we give a detailed
906: explanation of the numerical procedure for
907: finding ${\mathsf C}_j$ that can be derived from the above observation.
908:
909:
910:
911:
912: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
913: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
914: \section{The quantization measure}
915: \label{IV}
916:
917:
918: Once we have constructed the wavefunction at a given $k$, the next step
919: is to determine whether $\Psi$ is an eigenstate. As we will explain below,
920: our choice of measures reduces to finding the minima of one of:
921: %
922: \begin{eqnarray} \label{e28a}
923: S(k) \ &=& \ \mbox{tension}
924: \\ \label{e28b}
925: S(k) \ &=& \ \mbox{smallest singular value}
926: \\ \label{e28c}
927: S(k) \ &=& \ \mbox{determinant}
928: \end{eqnarray}
929: %
930: We call $S(k)$ the quantization measure.
931: Below we give further explanation of the above definitions.
932:
933:
934:
935: It is clear that the most natural quantization measure is the tension.
936: If a properly normalized wavefunction has ``zero tension" on the boundary,
937: it means that the corresponding $k$ is an eigenvalue.
938: The normalization issue is further discussed in Appendix~D.
939: The question that arises is whether we can use a numerically simpler
940: measure, and what price we pay for doing so.
941:
942:
943: The BIM Eq.~(\ref{e12}) and the GFM Eq.~(\ref{e13}) can both be written as
944: $\mbf{A}\Phi = 0$, with the appropriate choice of $\mbf{A}$.
945: Thus, if $k$ is an eigenvalue, $\mbf{A}$ should have a singular value
946: that tends to $0$ as $N$ increases. The determinant of $\mbf{A}$
947: is defined as the product of all the singular values, and therefore
948: it should vanish whenever one of the singular values does.
949: Using the {\em GFM-DEM duality} which is discussed in Section~5,
950: it is clear that for the DEM (and for the PWDM in particular)
951: the determinant of $\mbf{A}$ vanishes whenever $k$ is an eigenvalue.
952: It is important to realize that in the latter argumentation,
953: {\em we do not rely on inside-outside duality} \cite{uzy} considerations,
954: but rather on the much simpler {\em GFM-DEM duality}.
955:
956:
957: Thus, a low tension must be correlated with
958: having a vanishingly small singular value or determinant.
959: The converse is not true: It is well known that
960: SVD based quantization measures may lead to spurious
961: minima (see \cite{backer} and references therein).
962: Therefore SVD based procedure for finding eigenvalues
963: requires a post-selection procedure whose aim
964: is to distinguish true zeros from fake ones.
965:
966:
967: It is important to realize that neither the traditional implementation
968: of PWDM, nor that of the BIM should be considered to be 'package deals'.
969: For example, the BIM could be used with the
970: tension as a measure (defined in the next section), rather than looking
971: for minima of the the singular values.
972: Similarly, the usual Heller method of PWDM implementation (see Appendix~D)
973: could be replaced by a search over determinant values.
974:
975:
976:
977: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
978: \subsection{The tension as a quantization measure}
979:
980:
981: The tension is a robust measure of quantization.
982: Fig.5 displays some examples of the corresponding
983: $S(k)$ plots. The PWDM minima
984: are typically much sharper than their BIM equivalents.
985: Zooming over a PWDM minimum (Fig.5d) reveals
986: some amount of roughness.
987: This feature is actually helpful, because it
988: gives an indication of and control over the accuracy
989: of the numerics. We interpret the roughness of the
990: PWDM minimum as a reflection for the
991: existence of a null-space. In the same spirit,
992: the smoothness of the BIM minima can be regarded
993: as an indication that better accuracy can be
994: obtained by making $N$ larger. We discuss these issues
995: below.
996:
997: The tension provides a common measure that
998: may be used to monitor improvements,
999: as well as to compare the success of the different methods.
1000: Naturally, the first issue to discuss is
1001: the dependence of the tension on the
1002: size $N$ of the basis set (see Fig.6).
1003: For the BIM, the tension becomes better as $N$ grows,
1004: and disregarding the computer hardware,
1005: there is no reason to suspect that there is an inherent
1006: limitation on the accuracy. The situation is different for the PWDM.
1007: Here, taking $N$ much larger than $N_{sc}$ is not effective.
1008: In practice, the method reaches a limiting accuracy,
1009: which, taking into account present hardware limitations,
1010: is still very good compared with that of the BIM.
1011:
1012: From Fig.6, it is also clear that the tension of the PWDM becomes
1013: much better as $k$ becomes larger. This is expected on the
1014: basis of the following semiclassical reasoning: larger uncertainties
1015: in $k$ result from confining a particle
1016: to a smaller box (taking a smaller box for a given $k$
1017: is equivalent to making $k$ smaller for a given box size).
1018: Thus, it is more difficult to build a wavefunction with a precise
1019: value of $|k_j|=k$ for low lying eigenstates.
1020: On the other hand, the BIM does not seem to be sensitive
1021: to the value of $k$.
1022:
1023:
1024:
1025:
1026: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1027: \subsection{The tension as an indication for the global error}
1028:
1029:
1030: The tension can be regarded as a measure
1031: of the {\em local error} in the
1032: determination of the eigenfunction.
1033: The tension is local in the sense that it pertains only
1034: to points along boundary. We can also define
1035: a measure for the {\em global error},
1036: that is the error which is associated with all
1037: the interior points:
1038: %
1039: \begin{eqnarray} \label{e_error}
1040: (\Delta \Psi)^2 \ \ = \ \
1041: \langle |\Psi_r - \Psi_{\tbox{exct}}(X_r)|^2 \rangle
1042: \end{eqnarray}
1043: %
1044: Here $\Psi_{\tbox{exct}}(x)$ is the numerically exact
1045: wavefunction. The average is taken over the
1046: set $X_r$ of selected points inside of the boundary.
1047: Fig.7 gives an example for the variation of the
1048: error along the cross section line of Fig.1.
1049: In order to eliminate a possible bias due to a
1050: global normalization error, we renormalized
1051: the inexact wavefunction so that
1052: $\Psi_r = \Psi_{\tbox{exct}}(X_r)$ at a randomly selected
1053: point $X=X_0$. In retrospect, we realized that
1054: such an error did not significantly affect the result.
1055: However, we still chose to be on the
1056: safe side, and we adopted this procedure routinely.
1057:
1058: It is natural to expect the average error
1059: $(\Delta \Psi)^2$ to be correlated with
1060: the tension. In other words, if
1061: $|\Psi_r - \Psi_{\tbox{exct}}(X_r)|$ is small on the boundary,
1062: then one may expect it to be small in the interior.
1063: The degree of such correlation is important
1064: for practical reasons. Moreover, we have introduced two different versions
1065: of tension definitions, one for each of the PWDM and the BIM.
1066: It is not a-priori clear that the above correlation is independent
1067: of the choice of numerical method.
1068: In Fig.8, we study this issue by plotting $(\Delta \Psi)^2$
1069: against the tension for the BIM and the PWDM.
1070: In the case of the PWDM, the error saturates below a critical
1071: tension. After this point, further improvements on the boundary
1072: do not seem to affect the bulk of the eigenstate.
1073: It is not clear from the numerics whether or not the BIM
1074: saturates. What is clear however is that
1075: the BIM does a poorer job at reproducing
1076: the wavefunction inside of the boundary than the PWDM with
1077: the same tension.
1078:
1079: The saturation of the error well inside the billiard
1080: can be explained as a manifestation of the fact that
1081: the wavefunction there is not very sensitive to
1082: sub-wavelength roughness of the boundary:
1083: If $N$ is reasonably large, the numerical wavefunction
1084: vanishes on a nodal line that almost coincides with the
1085: true (pre-defined) boundary. Increasing $N$ further
1086: affects the sub-wavelength features of the
1087: (distance) difference between that nodal-line and
1088: the true boundary. This distance difference is
1089: important for the tension, but barely
1090: affects the wavefunction well inside the billiard.
1091:
1092:
1093: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1094:
1095: \subsection{The determinant as a quantization measure}
1096:
1097:
1098: The tension is the natural choice for a quantization measure.
1099: However, from a numerical point of view, it is much more convenient
1100: and time effective to compute the singular values of $\mbf{A}$,
1101: without having to find the eigenvectors for each $k$ value,
1102: and without having to compute the wavefunction
1103: along the boundary (for tension calculation).
1104:
1105:
1106: The smallest singular value is traditionally used
1107: as a quantization measure for the BIM.
1108: From Fig.7, it is quite clear that for the BIM
1109: one of the singular values is significantly smaller
1110: than the others, so that the eigenstate
1111: is unambiguously determined by this method.
1112: Is it possible to use the same approach, with comparable success,
1113: for the PWDM? We have already determined that looking at the
1114: smallest singular values is not very meaningful numerically.
1115: For $N > N_{sc}$, there exists a large null-space of evanescent
1116: states for any $k$. The metric method (Appendix D)
1117: is not a practical solution to this problem
1118: since we want to gain numerical efficiency (if efficiency is not
1119: the issue then it is better to use the tension as a quantization measure).
1120:
1121:
1122: One simple way to improve the numerical stability
1123: is to use the determinant rather than the smallest
1124: singular value as a quantization measure:
1125: Each time that $k=k_n$, the null space should include one more
1126: `dimension'. Therefore, the determinant, rather than the
1127: smallest singular value, becomes the reasonable quantity
1128: to look at. Thus, from numerical point of view (\ref{e28b})
1129: should be superior compared with (\ref{e28a}).
1130:
1131:
1132: Fig.9 illustrates how the determinant can be used in practice
1133: as a quantization measure. As a general rule, as is the case
1134: for the tension, the PWDM/GFM minima are sharper than the BIM
1135: ones. On the one hand, this extra sharpness can be regarded as
1136: an advantage, because it leads to a better resolution of the
1137: eigenvalue spectrum. However, more computer time is needed in order
1138: to find these minima. The BIM minima are broader,
1139: and therefore digging algorithms that search for local minima
1140: become extremely effective.
1141:
1142:
1143: In the case of the traditional BIM, using a larger $N$ leads to
1144: a better resolution of the local minima, as expected.
1145: The traditional H-BIM uses the complex Hankel Bessel
1146: function as its Green function. One may wonder why
1147: the real Neumann function could not be
1148: used instead. A-priori, there is no reason to
1149: insist on Hankel choice. However, it seems that
1150: with Neumann choice the numerics are not very stable:
1151: The locations of the local minima vary on a $k$ range which is
1152: large compared to their $k$ width.
1153: Because of this problem, search routines
1154: relying on the Y-BIM may yield misleading values for the
1155: error in the $k_n$ determination.
1156: Thus, the numerical stability of the H-BIM can be attributed
1157: to the fact that the BIM equation $\mbf{A}\Phi=0$ becomes
1158: complex. Its real part is just the Y-BIM equation,
1159: while its imaginary part is the J-GFM equation.
1160: Thus one may say that the H-BIM benefits from combining
1161: the Y-BIM with the J-GFM.
1162:
1163:
1164:
1165: Is it practical to use the Fredholm determinant
1166: as a quantization measure also in the GFM/PWDM case?
1167: Here we observe that the null-space problem is reflected
1168: in the stability of the determinant calculation.
1169: It is useful to characterize the numerics
1170: using the discretization parameter~$b$:
1171: %
1172: \begin{eqnarray} \label{e_bdef}
1173: b \ \ = \ \ \frac{N}{N_{sc}} \ \
1174: =\Big|_{\tbox{$M$=$N$}} \ \ \frac{\lambda/2}{\Delta s}
1175: \end{eqnarray}
1176: %
1177: The last equality holds if we take $M=N$,
1178: leading to the interpretation of $b$ as the
1179: number of boundary points per half De-Broglie wavelength.
1180: If $b<1$ the null space problem does not exist,
1181: and we can define ${\mathsf C}$ as the
1182: (left) eigenvector that corresponds to the smallest
1183: singular value of $\mbf{A}_{js}$.
1184: Of course, we want to push PWDM to the limit,
1185: and therefore in practice we always take $b>1$.
1186:
1187:
1188:
1189: The natural question is whether choosing a
1190: very large $b$ is numerically useful.
1191: Our numerical experience is that for $1<b<1.8$
1192: we get nice minima, which actually look much
1193: sharper than the BIM ones (see Fig.~9).
1194: As we try to increase $b$ in order to improve accuracy,
1195: the numerics loose stability (what we mean by
1196: instability is demonstrated in Fig.~9e).
1197: The same phenomenon occurs with J0-GFM,
1198: which has somewhat larger tendency for instability.
1199: This is apparently because the J0-GFM
1200: is involved with a larger null-space (see Fig.~4).
1201:
1202:
1203:
1204: Thus we conclude that taking $b>1$ does improve the
1205: numerics, while $b \gg 1$ generally leads to
1206: instabilities that should be avoided.
1207: The optimal choice of $b$ depends
1208: on the details of the implementation
1209: and on the computer hardware. It should be clear that if the
1210: numerical accuracy were unlimited,
1211: then the $b\rightarrow\infty$ limit would lead
1212: to a numerically exact solution in cases
1213: where the wavefunctions may be written
1214: as superpositions of plane waves.
1215: This is not always possible \cite{boris}.
1216: Note however that evanescent features
1217: of the wavefunction can be reconstructed
1218: by a suitable superposition of plane waves \cite{berry2}.
1219:
1220:
1221:
1222:
1223: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1224: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1225: \section{The duality of the GFM and the DEM}
1226: \label{V}
1227:
1228: In addition to providing a boundary method of its own,
1229: the GFM also serves to bridge the gap between
1230: the BIM and the DEM. Consider the version of the GFM
1231: that is based on the choice ${\cal C}(x,x') = F_j(x;k)$,
1232: where the $F_j$ are solutions of the Helmholtz equation
1233: in free space (with neither singularities nor cuts).
1234: With this choice, we immediately realize that
1235: the DEM and the GFM are dual methods:
1236: %
1237: \begin{eqnarray}
1238: \label{e14a}
1239: \mbf{A} \Phi=0 \hspace*{1.5cm} & \mbox{[GFM equation]} \\
1240: \label{e14b}
1241: {\mathsf C} \mbf{A}=0 \hspace*{1.5cm} & \mbox{[DEM equation]}
1242: \end{eqnarray}
1243: %
1244: The only difference lies in whether one looks
1245: for the {\em left} or the {\em right} eigenvector.
1246: This point is numerically demonstrated in Fig.2.
1247:
1248: The PWDM version of the DEM also satisfies this duality.
1249: In this special case, a somewhat more elegant version
1250: of the above argument is as follows:
1251: Consider the version of the GFM that is based on the choice
1252: \mbox{${\cal C}(x,x') = \exp(ik n_j\cdot(x-x'))$},
1253: where $n_j$ is a unit vector in a given direction.
1254: We can take $N$ different choices of $n_j$,
1255: thus obtaining the matrix equation $\mbf{A} \Phi = 0$ with
1256: %
1257: \begin{eqnarray} \label{e14}
1258: \mbf{A}_{js} \ \ = \ \
1259: \exp(ik n_j \cdot x_s)
1260: \end{eqnarray}
1261: %
1262: An equivalent matrix equation is found by
1263: multiplying each equation by $\exp(ik\phi_j)$,
1264: where $\phi_j$ are random phases. We can then
1265: take the real part of these equations, thus
1266: obtaining a set of equations that involves the same
1267: matrix $\mbf{A}$ as that of PWDM,
1268: namely (\ref{e5}) with the basis defined by (\ref{e2}).
1269:
1270:
1271: The duality of the PWDM and the GFM is very important from
1272: a mathematical point of view. The mathematical
1273: foundations of the PWDM are quite shaky.
1274: It is clear that PWDM is well-established mathematically
1275: if a strict `inside-outside duality' (IOD) \cite{uzy} is satisfied.
1276: %
1277: % That is, the interior (billiard) region has an
1278: %eigenvalue at a given $k$ if and only if the
1279: %scattering matrix does for the outside scattering problem.
1280: %
1281: In this case, the wavefunction can be extended to the
1282: whole plane so that the boundary of the billiard can be
1283: regarded as a {\em nodal line} of some plane-wave superposition.
1284: Obviously, this is rarely possible \cite{berry2}.
1285: Therefore, one may wonder whether
1286: we indeed have $\det(\mbf{A})=0$ whenever $k=k_n$.
1287: Using the duality Eq.(\ref{e14a}), it becomes obvious
1288: that the Fredholm determinant indeed vanishes at the eigenstates,
1289: even in the absence of an exact IOD.
1290:
1291: It is quite clear from the first paragraph
1292: of this section that any expansion method can
1293: be associated with a corresponding GFM.
1294: Whenever the left eigenvector
1295: is used with the expansion method,
1296: the right eigenvector can be used with the GFM.
1297: We have already demonstrated this point in Fig.2.
1298: Is it possible to make the inverse statement?
1299: Do we have a well defined expansion method
1300: associated with any GFM? The answer is negative.
1301: We discuss this issue in the rest of
1302: this section, and it can be skipped at first reading.
1303:
1304: For the following, it is convenient to
1305: consider Eq.(\ref{e6}) as $N\rightarrow\infty$.
1306: Subsequently, we are going to talk about whether this
1307: limit is meaningful.
1308: In the case of the usual PWDM,
1309: the $N\rightarrow\infty$ limit of Eq.(\ref{e6})
1310: can be written as
1311: %
1312: \begin{eqnarray} \label{e16_1}
1313: \Psi(x) = \int_0^{2\pi}
1314: C(\theta)d\theta
1315: \ \exp(ik n(\theta)\cdot x)
1316: \end{eqnarray}
1317: %
1318: Similarly, in the case of the $J_0$ decomposition,
1319: using the basis functions of Eq.(\ref{e4}),
1320: we can write in complete analogy:
1321: %
1322: \begin{eqnarray} \label{e16_2}
1323: \Psi(x) = \oint
1324: \Phi(s)ds \ J_0(k|x-x(s)|)
1325: \end{eqnarray}
1326: %
1327: In writing Eq.(\ref{e16_2}) we have used the fact
1328: that $J_0(x(j)-x(s))$ is a symmetric kernel,
1329: and therefore we could make the substitution ${\mathsf C}=\Phi$.
1330:
1331:
1332: Eq.(\ref{e16_2}) looks at first sight
1333: like an innocent variation of Eq.(\ref{e16_1}).
1334: The Bessel function $J_0$ is just a superposition
1335: of plane waves, and therefore one may
1336: possess the (incorrect) idea that there is a simple
1337: way to go from Eq.(\ref{e16_2}) to Eq.(\ref{e16_1}).
1338: If we expand each $J_0$ in Eq.(\ref{e16_2})
1339: in plane waves, and re-arrange the expression
1340: in order to identify the PWDM coefficients,
1341: we end up with the relation
1342: %
1343: \begin{eqnarray} \label{e18}
1344: {\mathsf C}(\theta) =
1345: \int \mbox{e}^{- ik n(\theta){\cdot}x(s)}
1346: \ \Phi(s) ds
1347: \end{eqnarray}
1348: %
1349: This relation implies the trivial result
1350: ${\mathsf C}(\theta)=0$ due to gauge freedom
1351: [see discussion of Eq.(\ref{e14})].
1352: Hence we conclude that the constructed
1353: wavefunction is $\Psi(x) \equiv 0$ in the
1354: $N\rightarrow\infty$ limit!
1355:
1356: Having $\Psi(x) \equiv 0$ from Eq.(\ref{e16_2})
1357: could have been anticipated using a simpler
1358: argument: We know that Eq.(\ref{e10}) should hold
1359: for {\em any} gauge choice.
1360: This gauge freedom implies that we have
1361: %
1362: \begin{eqnarray} \label{e17}
1363: \oint {\cal C}(x,x(s)) \Phi(s) ds \ \ = \ \ 0
1364: \end{eqnarray}
1365: %
1366: The above equation should hold for
1367: any $x$ inside
1368: as well as on the boundary.
1369: Furthermore, the left hand side
1370: of (\ref{e17}) is manifestly a solution
1371: of Helmholtz equation in free space,
1372: and it follows from the unique continuation
1373: property that Eq.(\ref{e17}) holds also
1374: for points outside of the boundary.
1375: %
1376: Having $\Psi(x)=0$ as a result of the
1377: integration in Eq.(\ref{e16_2})
1378: is just a particular case of Eq.(\ref{e17}).
1379:
1380: In spite of the observation that Eq.(\ref{e16_2})
1381: yields $\Psi(x) \equiv 0$ in the $N\rightarrow\infty$
1382: limit, the vector $\Psi_r$ is non-zero numerically
1383: for any finite $N$. In fact,
1384: after proper re-normalization, $\Psi_r$ becomes a quite good
1385: approximation to the wavefunction (see Fig.3c and Fig.3e).
1386: Sometimes the result so obtained is
1387: even better than the one which is found
1388: via the traditional BIM Eq.~(\ref{e7}).
1389: As strange as it sounds, this success is entirely
1390: due to the fact that we are using finite $N$.
1391:
1392: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1393:
1394: \section{Conclusion}
1395:
1396: The BIM and the DEM were written as different faces of a unified boundary
1397: procedure comprising four steps. In the process of doing so, yet a third
1398: boundary method was derived as part of the same framework, the GFM.
1399: The DEM and the GFM are strongly related, as they are respectively the
1400: left and right eigenvectors of the same Fredholm matrix. We think that
1401: the presented approach opens the way towards a controlled fusion
1402: of the BIM and the DEM into a more powerful numerical procedure.
1403:
1404: The unified treatment of quantization measures allowed us to compare
1405: the efficiency of the various methods, and to analyze both the local
1406: and the global errors in the numerically determined wavefunction.
1407: In particular, a numerically valid definition of the BIM tension was given,
1408: and was found to possess smooth minima at the eigenstates.
1409: Using the tension as a quantization measure is one possible way
1410: to avoid some problems \cite{backer} that are encountered in the traditional
1411: implementation of the BIM.
1412:
1413: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1414: \appendix
1415:
1416:
1417: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1418: \section{The BIM for the scattering problem}
1419:
1420: The solution of the Helmholtz equation for the scattering problem
1421: is just another variation of the BIM. Consider a boundary, one that in
1422: general may be composed of several disconnected pieces. The incident
1423: wave $\Psi_{\tbox{incident}}(x)$ is a solution of Helmholtz
1424: equation in free space. Formally, $\Psi_{\tbox{incident}}(x)$ includes both
1425: the ingoing and the outgoing wave components. We look for a solution
1426: $\Psi(x)$ that has the same ingoing
1427: component as $\Psi_{\tbox{incident}}(x)$,
1428: and that satisfies $\Psi(x)=0$ on the boundary.
1429: Such solution can be written as a sum
1430: of the incident wave and a scattered wave,
1431: and hence must be of the form
1432: %
1433: \begin{eqnarray} \label{eb1}
1434: \Psi(x) = \Psi_{\tbox{incident}}(x)
1435: + \oint G(x,x(s')) \Phi(s') ds
1436: \end{eqnarray}
1437: %
1438: Equation (\ref{eb1}) is a variation of Eq.(\ref{e10}).
1439: Note that the Green function should
1440: satisfy outgoing boundary conditions in order to
1441: yield the desired solution.
1442: The charge density $\Phi(s)$ is fixed by
1443: the requirement that $\Psi(x)=0$, which leads to the boundary equation
1444: %
1445: \begin{eqnarray} \label{eb2}
1446: \oint G(x(s),x(s')) \Phi(s') ds = - \Psi_{\tbox{incident}}(s)
1447: \end{eqnarray}
1448: %
1449: This inhomogeneous equation is a straightforward
1450: generalization of Eq.(\ref{e12}). A discretized
1451: version of it was used in Ref.\cite{lupo} in order
1452: to obtain numerical solutions of the Helmholtz equation
1453: for some scattering problems.
1454:
1455:
1456: The derivation of Eq.(\ref{eb2}) in Ref.\cite{lupo}
1457: is much more complicated than ours, and involves the use
1458: of Lippmann-Schwinger equation.
1459: The boundary is represented by a large delta-potential $V$,
1460: and the limit $V\rightarrow\infty$ is taken.
1461: Using this procedure, the charge density $\Phi(s)$ can
1462: be obtained as the $V\rightarrow\infty$ limit of $V\Psi(x(s))$.
1463: Note the correctness of the physical units. Namely,
1464: $[{\cal H}][\Psi]=[\rho]$ and therefore $[V][\Psi]=[\Phi]$.
1465: Note also that the wavefunction is in general non-zero
1466: in both sides of the boundary. Therefore the charge density
1467: $\Phi(s)$ is equal to the {\em difference} between the
1468: normal derivatives on both sides of the boundary.
1469: The simplest way to derive the relation between $\Phi(s)$
1470: and $V\Psi(x(s))$ is to integrate the Helmholtz equation
1471: over an infinitesimal range across the boundary,
1472: as in the treatment of the 1D Shchrodinger equation
1473: with delta potential.
1474:
1475:
1476: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1477: \section{Traditional BIM equation}
1478:
1479: The primitive BIM equation (Eq.(\ref{e12}))
1480: is based on Eq.(\ref{e10}).
1481: The traditional BIM is a variation of the same idea
1482: which avoids the boundary singularities that
1483: plague the more simplistic version.
1484: Rather than using Eq.(\ref{e10}) directly,
1485: one considers its gradient, leading to
1486: %
1487: \begin{eqnarray} \label{ea1}
1488: \partial_{\pm} \Psi(x(s)) \ = \
1489: \oint \partial_{\pm} G(x(s),x(s')) \Phi(s') ds'
1490: \end{eqnarray}
1491: %
1492: This equation is analogous to Eq.(\ref{e12}).
1493: We use the notation $\partial_{+}$ and $\partial_{-}$
1494: in order to refer to the normal derivative
1495:
1496: on the interior and exterior sides of the boundary
1497: By definition, \mbox{$\partial_{-} \Psi(x(s)) = \Phi(s)$},
1498: and from the discussion in section III
1499: we have \mbox{$\partial_{+} \Psi(x(s)) = 0$}.
1500: Adding the two equations of (\ref{ea1}), we obtain
1501: %
1502: \begin{eqnarray} \label{ea2}
1503: \Phi(x(s)) \ = \
1504: \oint 2\partial G(x(s),x(s')) \Phi(s') ds'
1505: \end{eqnarray}
1506: %
1507: where $\partial \equiv (\partial_{+}+\partial_{-})/2$
1508: is just the derivative on the boundary in the principal sense.
1509: Thus, in the traditional BIM, the definition
1510: of the Fredholm matrix is
1511: %
1512: \begin{eqnarray} \label{ea3}
1513: \mbf{A}_{ss'} \ = \ \frac{1}{\Delta s}\delta_{ss'} \
1514: - \ 2\partial G(x(s),x(s'))
1515: \end{eqnarray}
1516: %
1517: and the BIM equation (\ref{ea2}) is $\mbf{A}\Phi=0$.
1518: The kernel $\partial G(x(s),x(s'))$ is well-
1519: behaved, and its diagonal elements are finite
1520: thanks to the presence of a geometrical factor.
1521: Namely, if $G(x(s),x(s'))=g(k|x(s)-x(s')|)$ then
1522: %
1523: \begin{eqnarray}
1524: \partial G = k\frac{n(s)\cdot(x(s)-x(s'))}{|x(s)-x(s')|}
1525: g'(k|x(s)-x(s')|)
1526: \end{eqnarray}
1527: %
1528: If either one of the Bessel functions $H_0$ or $Y_0$
1529: is used for the Green function, then the definition
1530: of $\mbf{A}_{ss'}$ above involves
1531: either $H_1$ or $Y_1$, respectively.
1532:
1533:
1534:
1535: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1536: \section{Transformed BIM equations}
1537:
1538: There exists another elegant version of the BIM
1539: \cite{klaus,klaus-long,KKR} which does not exhibit the singularities associated
1540: with the Y0-BIM. Namely, the BIM equation is rewritten as
1541: %
1542: \begin{eqnarray}
1543: \int \tilde{G}(x(j),\kappa)
1544: \tilde{\Phi}(\kappa) d\kappa \ \ = \ \ 0,
1545: \end{eqnarray}
1546: %
1547: where the vector $\tilde{\Phi}(\kappa)$ is related by a linear transformation
1548: to the vector $\Phi(s)$. This transformation corresponds to the selection of a
1549: complete basis set of boundary wavefunctions. The KKR method \cite{KKR}
1550: is a particular implementation which uses the spherical harmonics.
1551: Another (more general) choice \cite{klaus,klaus-long} consists of taking
1552: $\tilde{\Phi}(\kappa)$ as the Fourier components of $\Phi(s)$.
1553: In the latter case $\tilde{G}$ is related to $G$ by the Fourier
1554: transform $s\mapsto\kappa$. However, obtaining $\tilde{G}$ from $G$ is not
1555: a simple matter for most boundary shapes.
1556:
1557:
1558:
1559: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1560: \section{The PWDM, in practice}
1561:
1562: The mathematically clean solution for the
1563: null-space problem is to adopt a method based on a metric.
1564: As we explain in the next paragraph,
1565: this procedure is sensitive to cumulative numerical errors.
1566: A modified implementation of the metric method,
1567: that avoids some of the numerical problems,
1568: has been introduced by Barnett~\cite{barnett}.
1569: The other possibility is to use a very simple
1570: procedure which is known as Heller's method~\cite{heller2}.
1571: Below we discuss the latter as well.
1572:
1573:
1574: The metric method works as follows:
1575: First, one finds the basis in which the
1576: normalization metric $\mbf{B}_{ij}$ becomes $\delta_{ij}$.
1577: The tension metric $\mbf{T}_{ij}$ should then be written
1578: in that same basis. The SVD is done on the transformed
1579: tension metric. In this case, the null space becomes at most
1580: one-dimensional (whenever $k=k_n$). Unfortunately, this
1581: elegant and straightforward metric scheme does not work very well,
1582: due to the finite precision problems
1583: discussed in connection with Eq.(\ref{e16aa}).
1584: Furthermore, $\mbf{B}$ and $\mbf{T}$ are "squares" of $\mbf{A}$,
1585: which leads to a loss of numerical precision
1586: compared with an $\mbf{A}$-based strategy.
1587:
1588:
1589: The most widely used $\mbf{A}$-based
1590: strategy is referred to as Heller's method~\cite{heller2}.
1591: The idea is to find ${\mathsf C}_j$ as the solution
1592: of the $M \geq N$ set of equations
1593: $\sum_j {\mathsf C}_j \mbf{A}_{js} = 0$,
1594: with the additional constraint
1595: $\sum_j {\mathsf C}_j \mbf{A}_{j0} = 1$.
1596: Table 1 gives the definition of $\mbf{A}_{j0}$.
1597:
1598: By constraining the wavefunction to be $\Psi(X_0)=1$
1599: at a selected point $X_0$ in the interior of the billiard,
1600: we eliminate the problems associated with evanescent states,
1601: as the associated (evanescent) wavefunctions no longer vanish
1602: on the boundary and thus, there is no longer
1603: a null-space problem. As a result, quite large~$b$
1604: can be used without encountering numerical instabilities.
1605: The only worry with this method is that $X_0$ may happen
1606: to be very close to a nodal line.
1607: In such cases, the tension will be large due to an improper
1608: normalization, so we will miss these eigenstates.
1609:
1610:
1611:
1612: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1613:
1614: \ \\
1615:
1616: \noindent
1617: {\bf Acknowledgments:}
1618:
1619: \noindent
1620: It is our pleasure to thank Alex Barnett, Areez Mody,
1621: Lev Kaplan and Michael Haggerty for many useful discussions,
1622: and Klaus Hornberger and Uzy Smilansky for their comments.
1623: This work was funded by ITAMP and the National Science Foundation.
1624:
1625: \ \\
1626:
1627: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1628:
1629: \begin{thebibliography}{99}
1630:
1631: \bibitem{sto}
1632: H.J. Stockmann, "Quantum Chaos : An Introduction"
1633: (Cambridge Univ Pr 1999).
1634:
1635: \bibitem{datta}
1636: S. Datta, "Electronic Transport in Mesoscopic Systems"
1637: (Cambridge Univ Pr 1997).
1638:
1639: \bibitem{kuttler}
1640: J.R.Kuttler and V.G.Sigillito, "Eigenvalues of the
1641: Laplacian", SIAM Review 26 (1984) 163-193.
1642:
1643: \bibitem{backer} A. Backer {\em in}
1644: "The mathematical aspects of quantum maps",
1645: M.Degli Esposti and S.Graffi (Eds), Springer (2003)
1646: [nlin.CD/0204061].
1647:
1648: \bibitem{barnett}
1649: A.~Barnett, PhD thesis (Harvard, Sept. 2000).
1650:
1651: % \bibitem{heller1}
1652: % E.J. Heller, Phys. Rev. Lett. {\bf 53}, 1515 (1984).
1653:
1654: \bibitem{conformal}
1655: M. Robnik, J. Phys. A {\bf 17}, 1049 (1983).
1656:
1657: \bibitem{reichl}
1658: G.A. Luna-Acosta, Kyungsun Na, and L.E. Reichl,
1659: Phys. Rev. E {\bf 53}, 3271 (1996).
1660:
1661: \bibitem{berry1}
1662: M.V.~Berry and M. Wilkinson,
1663: Proc. R. Soc. London A {\bf 392}, 15 (1984).
1664: R.~E. Kleinman and G.~F. Roach,
1665: SIAM Review {\bf 16}(2), 214--236 (1974).
1666: R.~J. Riddell, J. Comp. Phys. {\bf 31}, 21 (1979).
1667: S.~W. McDonald and A.~N. Kaufman,
1668: Phys. Rev. A {\bf 37}(8), 3067--3086 (1988).
1669:
1670: \bibitem{heller2}
1671: E.J Heller, {\it Chaos and Quantum Systems}, ed. M.-J.~Giannoni, A.~Voros,
1672: J.~Zinn-Justin (Elsevier, Amsterdam, 1991), p. 548.
1673:
1674: \bibitem{vergini}
1675: E.~Vergini, PhD thesis (Universidad de Buenos Aires, 1995).
1676:
1677: \bibitem{VS}
1678: E. Vergini and M. Saraceno, Phys. Rev. E, {\bf 52}, 2204 (1995).
1679:
1680: \bibitem{li}
1681: B.~Li, M.~Robnik, B.~Hu, Phys. Rev. E {\bf 57}, 4095 (1998).
1682:
1683: \bibitem{bmi} N. Lepore, D. Cohen and E.J. Heller, in preparation.
1684:
1685: \bibitem{klaus}
1686: K. Hornberger and U. Smilansky,
1687: J. Phys. A {\bf 33}, 2829 (2000).
1688:
1689: \bibitem{klaus-long}
1690: K. Hornberger, Spectral Properties of Magnetic Edge States,
1691: Dissertation, Ludwig-Maximilians-Universitat Munchen (2001).
1692: [http://www.mpipks-dresden.mpg.de/~klh/pubs/].
1693:
1694: \bibitem{KKR}
1695: M.V. Berry, Annals of Physics {\bf 131}, 163-216 (1981).
1696:
1697: \bibitem{boris}
1698: B. Gutkin, archive preprint nlin.CD/0301031.
1699:
1700: \bibitem{boasman}
1701: P.A.~Boasman, Nonlinearity {\bf 7}, 485.
1702:
1703: \bibitem{dietz}
1704: B.~Dietz and U.~Smilansky, Chaos {\bf 3}, 581 (1993).
1705:
1706: \bibitem{uzy}
1707: B. Dietz et al., Phys. Rev. E {\bf 51}, 4222 (1995).
1708:
1709: \bibitem{berry2}
1710: M.V.~Berry, J. Phys. A, {\bf 27}, L391 (1994).
1711:
1712: \bibitem{lupo}
1713: M.G.E. da Luz, A. Lupu Sax, and E.J. Heller,
1714: Phys. Rev. E {\bf 56}, 2496 (1997).
1715:
1716: \bibitem{vega}
1717: J.L. Vega, T. Uzer, J. Ford, Phys. Rev. E {\bf 52}, 1490 (1995).
1718:
1719: \bibitem{rmrk} The polar equation of the Pond shape is
1720: \mbox{$r=1+0.2*\cos(\theta+0.9*\cos(\theta))$}.
1721:
1722: \bibitem{rmrk2} We assume here the usual free space
1723: boundary conditions used in electrostatics.
1724:
1725: \bibitem{rmrk3} Note that Eq.~(\ref{e9}) always gives "a solution"
1726: of the GPL equation. This is true irrespective
1727: of the choice of boundary conditions and Green function.
1728:
1729: \end{thebibliography}
1730: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1731: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1732:
1733:
1734:
1735:
1736: %FIGURES:
1737:
1738:
1739: \newpage
1740: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1741: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1742: \centerline{\epsfig{figure=shapes,width=5cm}}
1743: \vspace{.1in}
1744: {\footnotesize {\bf FIG.1:}
1745: Full line: The 'Pond' shape~\cite{rmrk}.
1746: Its radius is roughly $1$, and its perimeter is $L=6.364$.
1747: Dashed line: The cross section line that is used e.g. in Fig.3c.
1748: Dots: The 'outer boundary' which is used for the
1749: BIM tension calculation (see Sec.III-A).
1750: The transverse distance between
1751: the actual boundary and the outer boundary
1752: was chosen in most cases to be $\Delta L \sim 10\Delta s$.
1753: Star: The selected point which is used
1754: in Heller's implementation of the PWDM method. }
1755: \\ \ \\
1756: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1757: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1758:
1759:
1760: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1761: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1762: \centerline{\epsfig{figure=PhiChi,width=0.8\hsize}}
1763: \vspace{.1in}
1764: {\footnotesize {\bf FIG.2:}
1765: The eigenvectors of the Fredholm matrix (Eq.(\ref{e5}))
1766: are found for $k = k_n = 6.82754592867694$.
1767: {\bf Right plot:} The right-eigenvector $\Phi$.
1768: The symbols (x) and (+) and (o) correspond respectively
1769: to the PWDM choice of Eq.(\ref{e2}),
1770: to the H1-BIM choice of Eq.(\ref{ea3}),
1771: and to the J0-GFM choice of Eq.(\ref{e4}).
1772: {\bf Left plot:} The left-eigenvector ${\mathsf C}$
1773: for the PWDM Fredholm matrix.}
1774: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1775: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1776:
1777:
1778:
1779: \newpage
1780: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1781: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1782: \mpg{2.6cm}{(a) \\ \vspace*{5cm} \\ (b) \\ \vspace*{3cm} }
1783: {\epsfig{figure=estate,width=0.5\hsize}}
1784:
1785: \mpg{1.4cm}{(c) \\ \vspace*{3cm}}
1786: \epsfig{figure=Xsections,width=0.6\hsize}
1787:
1788: \vspace{.1in}
1789: {\footnotesize FIG.3abc: See captions in the next page.}
1790:
1791: \mpg{0.5cm}{(d) \\ \vspace*{1.7cm} \\ (e) \vspace*{1.7cm} \\ (f) \vspace*{1.5cm}}
1792: \epsfig{figure=tensions,width=0.8\hsize}
1793: \vspace{.1in}
1794:
1795: {\footnotesize {\bf FIG.3:}
1796: The eigenfunction at $k=k_n$ is found using
1797: the eigenvectors of Fig.2.
1798: (a) An image of $\Psi(x)$ using Eq.(\ref{e7}).
1799: (b) The same using PWDM and Eq.(\ref{e6}).
1800: (c) Plot of $\Psi(x)$ along the crossection
1801: line of Fig.1. The vertical lines indicate
1802: the location of the boundary. The wavefunctions
1803: that correspond to images a-b are shown with (+)
1804: and (x) respectively. We also show with (o)
1805: the wavefunction that is obtained using J0-GFM
1806: and Eq.(\ref{e6}).
1807: Panels d-e-f display plots of $\log(|\Psi(x)|^2)$
1808: along the boundary. The symbols are as in c. }
1809: \\ \ \\
1810: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1811: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1812:
1813:
1814:
1815: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1816: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1817: \centerline{\epsfig{figure=sinvals,width=0.8\hsize}}
1818:
1819: \vspace{.1in}
1820:
1821: {\footnotesize {\bf FIG.4:}
1822: Singular values of the Fredholm matrix
1823: for $k_n=6.82754592867694$ ({\bf left panel})
1824: and for $k_n = 50.05474912004408$ ({\bf right panel}).
1825: The various symbols are as in Fig.2. }
1826: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1827: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1828:
1829:
1830: \newpage
1831: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1832: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1833: \centerline{\epsfig{figure=Sk,width=0.65\hsize}}
1834:
1835: \mpg{1cm}{(c) \\ \vspace*{4cm}}
1836: {\epsfig{figure=Sk_BIM,width=0.65\hsize}}
1837:
1838: \mpg{1cm}{(d) \\ \vspace*{2cm}}
1839: {\epsfig{figure=Sk_PWDM,width=0.65\hsize}}
1840:
1841: \vspace{.1in}
1842:
1843: {\footnotesize {\bf FIG.5:}
1844: The tension as a function of $k$ in the vicinity
1845: of $k_n=10.14707971517264$ (panels (a),(c),(d)),
1846: and of $k_n=50.05474912004408$ (panel(b)).
1847: In the upper panels (a-b) the full line is for the
1848: PWDM constructed wavefunction, while the dashed line
1849: is for the BIM constructed wavefunction.
1850: For the low $k$ state we chose $b=4$,
1851: while for the high $k$ we used $b=2$.
1852: Panel (c) demonstrates the dependence
1853: of the BIM tension on the choice of the distance $\Delta L$.
1854: The different curves (from the upper to lower)
1855: correspond to $\Delta L/\Delta s = 1,8,16,12,4$.
1856: Panel (d) is a zoom over the PWDM minimum. }
1857: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1858: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1859:
1860:
1861: \newpage
1862: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1863: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1864: \centerline{
1865: \epsfig{figure=TvsN1,width=0.5\hsize}
1866: \epsfig{figure=TvsN2,width=0.5\hsize}
1867: }
1868:
1869: \vspace{.1in}
1870:
1871: {\footnotesize {\bf FIG.6:}
1872: The tension for the constructed eigenstate
1873: versus the number $N$ of basis functions.
1874: The {\bf left panel} is for the
1875: $k_n = 2.40425657792391$ eigenstate,
1876: and the {\bf right panel} is for the
1877: $k_n = 6.82754592867694$ eigenstate.
1878: The symbols (o) and (+) are for
1879: the BIM and for the PWDM, respectively.}
1880: \\ \ \\
1881: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1882: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1883:
1884:
1885: \mpg{0.45\hsize}{
1886: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1887: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1888: \centerline{\epsfig{figure=Xtensions,width=\hsize}}
1889:
1890: \vspace{.1in}
1891:
1892: {\footnotesize {\bf FIG.7:}
1893: Plot of the error $|\Psi_r - \Psi_{exct}(X_r)|^2$,
1894: along the cross-section line of Fig.1.
1895: We refer here to the $k_n = 6.82754592867694$ eigenfunction.
1896: The numerically `exact' wavefunction
1897: is our best PWDM-constructed wavefunction ($N=69$)
1898: with tension$=10^{-13}$.
1899: The `non-exact' wavefunction is either
1900: BIM-constructed (+) or PWDM-constructed (x),
1901: with tension $\approx 10^{-8}$.
1902: In the middle point the error is zero
1903: by construction (see explanation in the text).}
1904: \\ \ \\ \ \\
1905: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1906: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1907: }
1908: %
1909: \ \ \ \ \ \ \
1910: %
1911: \mpg{0.45\hsize}{
1912: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1913: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1914: \centerline{\epsfig{figure=XtensvsN,width=\hsize}}
1915:
1916: \vspace{.1in}
1917:
1918: {\footnotesize {\bf FIG.8:}
1919: The averaged error in the
1920: determination of the wavefunction,
1921: versus the tension for the
1922: $k_n = 6.82754592867694$ eigenfunction.
1923: For the BIM-constructed wavefunction
1924: we use squares and (o),
1925: while for the PWDM one
1926: we use (+) and (x).
1927: The error has been determined with respect to
1928: the `exact' wavefunction $\Psi_{exct}$.
1929: The latter is numerically defined as either
1930: the best BIM eigenstate (squares and (+))
1931: or as the best PWDM one ((o) and (x)).
1932: For both choices $\Psi_{exct}$
1933: had a tension better than $10^{-10}$. }
1934: \\ \ \\
1935: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1936: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1937: }
1938:
1939:
1940:
1941:
1942: \newpage
1943: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1944: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1945:
1946: \mpg{0.7cm}{(a) \\ \vspace*{1.7cm} \\ (b) \vspace*{1.7cm} \\ (c) \vspace*{1.5cm}}
1947: {\epsfig{figure=det_k,width=0.72\hsize}}
1948:
1949: \mpg{1.5cm}{(d) \\ \vspace*{1.7cm}}
1950: {\epsfig{figure=det_BIM,width=0.65\hsize}}
1951:
1952: \mpg{1.5cm}{(e) \\ \vspace*{1.7cm}}
1953: {\epsfig{figure=det_PDM,width=0.65\hsize}}
1954:
1955: \vspace{.1in}
1956:
1957: {\footnotesize {\bf FIG.9:}
1958: The Fredholm determinant ($S(k)$)
1959: versus $k$ in the vicinity
1960: of $k_n = 10.14707971517264$.
1961: Note that $S(k)$ is normalized such
1962: that $S(k)=1$ away from the minima.
1963: Panels a-b-c show $S(k)$
1964: in the cases of the H1-BIM, the Y1-BIM and the PWDM, respectively.
1965: The lines plotted,
1966: in order of decreasing $S(k)$ minimum,
1967: correspond to $b=2,3,4,8$ in panel (a),
1968: $b=4,8,13,12$ in panel (b)
1969: and $b=2$ in panel (c).
1970: The PWDM run in panel
1971: (c) was repeated 3 times with different values of the
1972: randomly chosen plane wave phases.
1973: Panels (d) and (e) give a zoom over
1974: the minima of panels (a) and (c), respectively.
1975: We witness some numerical instabilities for both
1976: the Y1-BIM and the PWDM, though in the latter
1977: case it is much much weaker, and can be resolved
1978: only in the zoomed plot (panel~(e)).
1979: For larger $b$ values, the PWDM instability
1980: is enhanced, and the results are reduced to
1981: numerical garbage (not shown). }
1982: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1983: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1984:
1985:
1986:
1987: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1988: \end{document}
1989: