1: \documentclass[twocolumn, prl]{revtex4}
2: \usepackage{graphicx}
3:
4: \begin{document}
5: %\draft
6: \title{Harmonic inversion helps to beat time-energy uncertainty relations.}
7: \author{Zbyszek P. Karkuszewski}
8:
9: \affiliation{
10: Institute of Physics, Jagiellonian University, ul. Reymonta 4, 30-059 Cracow,
11: Poland
12: }
13:
14: \date{\today}
15:
16: \begin{abstract}
17: It is impossible to obtain accurate frequencies from time signals of a very
18: short duration. This is a common believe among contemporary physicists.
19: Here I present a practical way of extracting energies to a high precision from
20: very short time signals produced by a quantum system. The product of time span
21: of the signal and the precision of found energies is well bellow the limit
22: imposed by the time-energy uncertainty relation.
23: \end{abstract}
24:
25: \maketitle
26:
27: \section{Introduction}
28: Many methods of obtaining frequencies or energies from time signals have limited
29: resolution. This limitation is often expressed in the form of a time-energy
30: uncertainty relation. For instance, the width $\Delta E$ of spectral lines
31: in atomic spectroscopy is determined by life time $\tau$ of the excited
32: atoms
33: $$
34: \Delta E \tau = \hbar.
35: $$
36: Another example is a Fourier transform applied to a discrete time signal in
37: the interval of length $T$. The grid step in a frequency domain and thus the
38: resolution of the method is given by $\Delta \omega = 2\pi/T$, which in turn
39: yields
40: $$
41: \Delta \omega T = 2\pi.
42: $$
43: All relations of this kind may give a wrong impression that something more
44: fundamental is the source of the uncertainties -- a principle that one
45: cannot measure energies to an arbitrary precision in a very short time.
46: Such an intuition is common if the system subjected to energy measurement is
47: quantum.
48:
49: Uncertainty principles in quantum mechanics are a mathematical
50: consequence of the following theorem.
51: Two self-adjoint operators $\hat A$ and $\hat B$ defined on the same
52: Hilbert space necessarily obey the relation
53: \begin{equation}
54: \Delta A\Delta B \ge \frac{1}{2}|\langle[\hat A,\hat B]\rangle|,
55: \label{qur}
56: \end{equation}
57: where $[...,...]$ is a commutator, $\langle ...\rangle$ stands for quantum
58: average in a given state from the domain of $\hat A$ and $\hat B$,
59: and $(\Delta Z)^2 \equiv \langle \hat Z^2\rangle-\langle \hat Z \rangle ^2$.
60: In the case of position $\hat x$ and momentum $\hat p$ operators (\ref{qur})
61: gives
62: \begin{equation}
63: \Delta x \Delta p \ge \frac{\hbar}{2}.
64: \label{xpr}
65: \end{equation}
66: Notice that $\Delta x$ and $\Delta p$ are completely determined by the state
67: of the system being measured and have nothing to do with an accuracy of
68: the measuring apparatus. (\ref{xpr}) holds even if this accuracy is infinite.
69: The correct approach to the quantum time-energy uncertainties can be found
70: in~\cite{Asher}.
71: The theorem (\ref{qur}) does not apply to the case of time and energy,
72: simply because time is not an operator, it is just a parameter in quantum
73: mechanics. One gains nothing forcing the idea that a parameter is a special
74: kind operator because (\ref{qur}) gives zeros on both sides and does not
75: provide grounds for existence of any time-energy uncertainty relation.
76: Indeed, it has been shown how to precisely measure energy in an
77: arbitrarily short time \cite{Aharonov1}. Some methods of the energy measurement
78: have their own limitations \cite{Aharonov2} reflected in loss of accuracy
79: when applied for very short time signals. Here I use the method free of such
80: inconveniences.
81:
82: Let us setup some general issues before getting to the details of the
83: method.
84: Suppose that a continuous signal $c(t)$ is given in a finite time interval
85: \begin{equation}
86: c(t) = \sum_{k=1}^K d_k e^{-i\omega_k t}, \qquad \mbox{for}\quad 0\le t \le T,
87: \label{sigc}
88: \end{equation}
89: where $K$ is finite, $d_k$ is a real amplitude and $\omega_k$ is a real
90: frequency. Complex frequencies will be considered in future.
91: The task is to find unknown amplitudes and frequencies of
92: $c(t)$. What one can see is that $c(t)$ is an analytic function of
93: time $t$ and, as such, can be uniquely extended beyond the interval $(0,T)$.
94: This means that, in principle, even for tiny $T$ it is possible to get all
95: amplitudes and frequencies to a high precision from (\ref{sigc}).
96: Unfortunately it seems difficult to solve this nonlinear problem analytically
97: and numerical methods cannot handle continuous signals due to infinite number
98: of data points. One way of getting around this problem is to take only finite
99: number of points at cost of loss of uniqueness of the extension.
100: However, as will be shown later, this drawback is usually not severe for
101: practical purposes. From now on I shall assume that the signal $c(t)$ is
102: known only at $N+1$ equidistant time points $t_n=n\delta t$ for $n=0,...,N$
103: and $t_N=T$.
104: (\ref{sigc}) can be rewritten as a set of $N+1$ equations
105: \begin{equation}
106: \sum_{k=1}^K d_k e^{-i\omega_k t_n} = c_n,
107: \label{sigd}
108: \end{equation}
109: where $c_n\equiv c(n\delta t)$. This set has $2K$ real unknowns
110: ($K$ amplitudes and $K$ frequencies) so the number of equations $N+1$ has to
111: be equal or greater than $K$. This condition would have opened the possibility
112: of existence of a unique solution if the equations were linear in $\omega_k$
113: and $d_k$. It is not the case here. There will always be infinite number of solutions
114: to (\ref{sigd}) if the set is self-consistent and no solutions otherwise.
115: For reasons explained in the next sections it will be required that $N\ge K$.
116: Notice that discrete Fourier transform method assumes the grid of $K$
117: equidistant frequencies and solves (\ref{sigd}) only for $d_k$ as a linear
118: set of
119: equations. The more challenging task of solving (\ref{sigd}) for
120: amplitudes and frequencies is performed by, so called, harmonic inversion
121: method.
122:
123: \section{Intervals of unique solutions}
124: The two functions $e^{i\omega n \delta t}$ and $e^{i\omega' n \delta t}$ have
125: the same values at equidistant time points $n \delta t$, $n=0,...,N$ if
126: $$
127: \omega' = \omega+ l\frac{2\pi}{n\delta t}.
128: $$
129: The integer number $l$ has to be chosen in such a way that the fraction
130: $l/n$ is integer as well. This implies that $l=mN!$, where $m$ is integer
131: and
132: \begin{equation}
133: \omega' = \omega + m\frac{2\pi N!}{\delta t}.
134: \label{uniq}
135: \end{equation}
136: The time discretization of the signal $c(t)$ , the transition from
137: (\ref{sigc}) to (\ref{sigd}), happens at cost of the loss of uniqueness of
138: solutions. However, as can be seen from (\ref{uniq}) solutions are unique
139: in finite intervals in frequency domain. Frequency $\omega$ is unique
140: in the interval
141: $$
142: \left( \omega -\frac{2\pi N N!}{T}, \omega + \frac{2\pi N N!}{T}\right),
143: $$
144: which can be made large by increasing $N$ and/or decreasing $T$.
145: Conversely, if one posses prior knowledge about the range of frequencies
146: in a signal it is straightforward to design a sampling step $\delta t$ to
147: extract all frequencies in a unique way.
148:
149:
150: \section{Harmonic inversion}
151: The key idea behind the harmonic inversion method is to replace the original
152: nonlinear problem (\ref{sigd}) with an eigenvalue problem of an operator, as
153: in (\ref{gep}). The presentation of the method in this section follows that
154: in \cite{Taylor2}.
155:
156: Lets start with a normalized quantum state $|\Phi_0\rangle$.
157: The evolution of this state is generated by unitary evolution operator
158: $\hat U(\delta t)$ and $|\Phi_n\rangle = \hat U^n(\delta t)|\Phi_0\rangle$.
159: The states $|\Phi_n\rangle$ are so important that they were given a name of
160: Krylov states. For every time signal in (\ref{sigd}) there exists such an
161: evolution by time $\delta t$ operator $\hat U$ that the signal can be
162: viewed as an autocorrelation function
163: \begin{equation}
164: c_n = \langle \Phi_0| \Phi_n\rangle.
165: \label{auto}
166: \end{equation}
167: Krylov states in (\ref{auto}) can be written in the orthonormal basis of
168: eigenvectors of $\hat U$ defined by $\hat U|u_k\rangle = u_k|u_k\rangle$.
169: Namely $|\Phi_n\rangle = \sum_{k=1}^K \alpha_k u_k^n|u_k\rangle$, where
170: $\alpha_k$ stands for a time independent complex number. (\ref{auto}) in the
171: new form appears as
172: \begin{equation}
173: c_n = \sum_{k=1}^K |\alpha_k|^2 u_k^n.
174: \label{nauto}
175: \end{equation}
176: Comparing (\ref{nauto}) and (\ref{gep}) one can see that eigenvalues
177: $u_k = e^{-i\omega_k \delta t}$ and $|\alpha_k|^2 = d_k$. It is enough
178: to find eigenvalues of $\hat U$ in order to obtain all frequencies $\omega_k$
179: in the signal $c_n$. These frequencies multiplied by $\hbar$ can be viewed as
180: eigenenergies of a hypothetical Hamiltonian governing the evolution of a
181: system in the initial state $|\Phi_0\rangle$. Of course, one can start
182: from the autocorrelation function (\ref{auto}) and find the eigenenergies
183: of the real system.
184:
185: It turns out that the matrix elements of $\hat U$ in the basis of Krylov states
186: can be expressed in terms of $c_n$ alone, i.e. without explicit reference to
187: the states $|\Phi_n\rangle$
188: \begin{equation}
189: U_{ij}\equiv \langle \Phi_i | U| \Phi_j\rangle =
190: \langle \Phi_0| U^{j-i+1} | \Phi_0\rangle = c_{j-i+1}
191: \label{me}
192: \end{equation}
193: where $i,j=0,..., N-1$. The negative
194: indices of $c$ in the equation above introduce no complication since
195: $c_{-n}=c^*_n$. To obtain $K$ eigenvalues the dimension $N$ of the matrix
196: $U$ must be equal or greater than $K$. It means that the number of
197: complex signal points $c_n$ required by the method exceeds the half of the
198: number of unknowns. The Krylov vectors $|\Phi_n\rangle$ are normalized but not
199: orthogonal.
200: Thus the eigenequation for matrix $U$ in the Krylov representation
201: (\ref{me}) takes the form
202: \begin{equation}
203: U |u_k\rangle = u_k S |u_k\rangle,
204: \label{gep}
205: \end{equation}
206: where $S$ is a matrix of scalar products $S_{ij}\equiv\langle
207: \Phi_i|\Phi_j\rangle=c_{j-i}$ with $i,j=0,...,N-1$.
208:
209: The harmonic inversion method consists of two stages. First is to solve
210: generalized eigenvalue problem (\ref{gep}), where all matrix elements
211: are expressed in terms of $c_n$, in order to get all frequencies.
212: Second, when all frequencies in (\ref{sigd}) are known, it is enough to
213: solve linear set of equations for the amplitudes $d_k$.
214: There are some practical difficulties in proceeding with the first stage.
215: For instance, numerical algorithms fail in finding the eigenvalues $u_k$
216: if the signal duration $T$ is small and the number of frequencies $K>4$.
217: The second stage is straightforward and will be skipped in this work. It
218: will be assumed that all real amplitudes $d_k$ are equal and the signal
219: is normalized to unity, i.e. $c_0 = 1$.
220:
221: In the next section we will see that the crucial role in the numerical
222: approach to (\ref{gep}) is played by the smallest positive eigenvalue of $S$.
223:
224: \section{Properties of matrix S}
225: One way of dealing with the generalized eigenvalue problem is to
226: reduce it to the ordinary eigenvalue problem by multiplying both sides of
227: (\ref{gep}) by the inverse of the Hermitian Toeplitz matrix $S$. However,
228: matrix $S$ is singular unless $N=K$ and the inverse does not exist.
229: Indeed, recall that every vector $|\Phi_n\rangle$ can be decomposed into a
230: superposition of $K$ linearly independent eigenvectors of $\hat U$.
231: Therefore the rank of the matrix $S$ is $K$. Additionally, those nonzero
232: eigenvalues are positive because a scalar product
233: $\langle \varphi|\varphi \rangle > 0$ for any nonzero state
234: $|\varphi \rangle$ in a Hilbert space.
235: Suppose that columns of a $N\times K$ matrix $G$ are represented by
236: eigenvectors of matrix $S$ to the positive eigenvalues. In order to
237: restrict (\ref{gep}) to the range of $S$ we define $U'\equiv G^\dagger U G$,
238: $S'\equiv G^\dagger S G$ and $|u_k'\rangle \equiv G^{-1}|u_k\rangle$.
239: $S'$ is diagonal and positive defined thus the ordinary eigenvalue problem
240: converts to
241: \begin{equation}
242: S'^{-1} U' |u_k'\rangle = u_k |u_k'\rangle .
243: \label{ep}
244: \end{equation}
245: Construction of matrix $G$ is possible only if one can extract all positive
246: eigenvalues of $S$. This task becomes hopeless if the smallest eigenvalue
247: of $S$ cannot be distinguished from zero due to limited accuracy.
248: The exact analytical expression for the smallest eigenvalue of a Toeplitz
249: matrices is not known yet.
250: Here is the approximate formula for the magnitude of the smallest
251: eigenvalue for the short time span $T$ of the signal
252: \begin{equation}
253: \frac{\lambda_{min}}{KN} \approx [a(\omega_1,...,\omega_K)\Omega T]^{2(K-1)}.
254: \label{mineig}
255: \end{equation}
256: $\Omega$ is the magnitude of the greatest (to the absolute value)
257: frequency in the signal. For the derivation of (\ref{mineig}) see the
258: directions in the Appendix.
259: The formula is valid only if the shortness
260: condition is fulfilled, $T\Omega \ll 1$. The expression in square
261: brackets in (\ref{mineig}) is smaller than 1 and the $\lambda_{min}$ decreases
262: exponentially fast with increasing $K$. For the estimated magnitude of the
263: function $a(\omega_1,...,\omega_K)$ see Fig.\ref{Fig1}.
264: \begin{figure}[htb]
265: \includegraphics*[width=8.6cm]{fig1b.eps}
266: \caption{Statistical properties of $a(\omega_1,...,\omega_K)$. For every $K$
267: 100 sequences of $K$ frequencies were randomly generated. Each frequency has
268: been uniformly drawn from the interval $(0.5,1.0)$. For each
269: sequence $a(\omega_1,...,\omega_K)$ was computed. Circles denote the most
270: probable value of $a(\omega_1,...,\omega_K)$ and vertical bars embrace 90\%
271: of sequences.}
272: \label{Fig1}
273: \end{figure}
274:
275: The eigenvalue $\lambda_{min}$ is additionally
276: corrupted by inaccuracies of the signal. The accuracy analysis is presented
277: in the next section.
278:
279:
280: \section{Impact of noise}
281: From now on it will be assumed that the accurate signal $c_n$ is perturbed
282: by noise $\eta_n$ and the new signal
283: $$
284: \tilde c_n = c_n + \eta_n.
285: $$
286: The noise is limited, $\eta_n \in (-\eta_{max}, \eta_{max})$ for all $n$
287: and $\eta_{max}\ge 0$.
288: Matrix $\tilde S$, which is constructed using $\tilde c_n$ instead
289: of $c_n$ (see the text under (\ref{gep})), remains Hermitian.
290: It can be shown \cite{golub} that the smallest eigenvalue of $\tilde S$
291: must obey the inequality
292: $$
293: |\tilde \lambda_{min} - \lambda_{min}|\le 2N\eta_{max}+ {\cal O}(\eta_{max}^2).
294: $$
295: If $\tilde \lambda_{min}$ is to be distinguished from perturbed zero
296: eigenvalues the following lower bound on $\lambda_{min}$ arises
297: \begin{equation}
298: \lambda_{min}\ge 4N\eta_{max}.
299: \label{bound}
300: \end{equation}
301: This diagnostic condition combined with (\ref{mineig}) provides the limits of
302: applicability of the harmonic inversion method for short signals.
303:
304: Tracking the inaccuracy propagation when solving (\ref{ep}) one arrives
305: at the surprising at first sight "certainty relation" on perturbed
306: frequencies $\tilde \omega_k$
307: \begin{equation}
308: |\tilde \omega_k - \omega_k|T \le \frac{2KN^2}{\lambda_{min}}\eta_{max}.
309: \label{ina}
310: \end{equation}
311: The error concerning frequencies of short signals can be made arbitrarily
312: small by reducing the noise amplitude!
313: Higher order terms in $\eta_{max}$ were dropped in (\ref{ina}).
314: For time signals of short duration the presence of $\lambda_{min}$ in the
315: denominator in RHS of (\ref{ina}) is unfavorable. As stated in
316: (\ref{mineig}) $\lambda_{min}$ rapidly goes to zero as $K$ increases.
317: Therefore, to extract many accurate frequencies from a short signal the very
318: low level of the noise $\eta_{max}$ will be required.
319:
320: In numerical experiments the noise is caused by finite precision of floating
321: point numbers. As an example, the harmonic inversion method was used
322: on a signal with $K=10$ frequencies drawn from an interval $(0.5,1.0)$,
323: sampled at $N=14$ points with $T=0.01$ and using 85 digits precision i.e.
324: $\eta_{max}=10^{-84}$. The results are shown in Table~\ref{tab:res}.
325: \begin{table}[htb]
326: \caption{\label{tab:res}Numerical example. All amplitudes $d_k$ were equal.}
327: \begin{ruledtabular}
328: \begin{tabular}{c | c | c}
329: $\omega_k$ & $\tilde \omega_k$& $|\tilde\omega_k - \omega_k|$\\
330: \hline
331: 0.50415486481506&0.50415486481507&0.00000000000001\\
332: 0.51315149664879&0.51315149664880&0.00000000000002\\
333: 0.66526505816728&0.66526505819813&0.00000000003085\\
334: 0.71068158128764&0.71068158894354&0.00000000765589\\
335: 0.73253390251193&0.73253391404702&0.00000001153508\\
336: 0.75819833122659&0.75819832230088&0.00000000892572\\
337: 0.79694000270683&0.79694000183578&0.00000000087105\\
338: 0.85043252643663&0.85043252642270&0.00000000001394\\
339: 0.88220469909720&0.88220469909823&0.00000000000103\\
340: 0.93358750757761&0.93358750757763&0.00000000000001\\
341: \end{tabular}
342: \end{ruledtabular}
343: \end{table}
344: Notice that the Fourier algorithm applied to this signal would give
345: the grid step in frequency domain $200\pi$ which is several orders
346: of magnitude greater than the error in Table~\ref{tab:res}.
347: In the example above $\lambda_{min}=4.07\times 10^{-78}$.
348:
349:
350: \section{Disadvantages of the method}
351: Harmonic inversion method does pretty well when applied to short time signals
352: provided the level of inaccuracies in the input data is very small.
353:
354: In principle this level can be kept very low in measurement of quantum
355: systems. For instance, the measurement of quantum autocorrelation
356: function squared (probability of finding a system in the initial state at
357: different times of its evolution) can be arbitrarily precise if done
358: simultaneously on many copies of the system. The inaccuracy decreases
359: as an inverse square root of the number of the copies. This function involves
360: $(K-1)K/2$ frequencies rather than $K$ as in (\ref{auto}).
361:
362: In experimental practice, however, inaccuracies are big and the method might
363: be useful only for signals with small number of frequencies.
364:
365: %Another area, where the harmonic inversion might be useful is a numerical
366: %stochastic evolution of interacting quantum many body systems
367: %~\cite{Carusotto}. No matter what precision during the calculation is
368: %used the accurate quantum averages, the "signal", can be obtained only on
369: %a short time interval. Harmonic inversion might help extend the signal further
370: %in time.
371:
372: Even if desired accuracy is provided another problem may show up.
373: The harmonic inversion method as described here involves diagonalization
374: of $N\times N$ matrix $\tilde S$ and $K\times K$ matrix
375: ${\tilde S}^{-1}\tilde U$.
376: $\tilde S$ matrix is Hermitian Toeplitz, the matrix elements
377: are constant along diagonals, and it is enough to store only one row of
378: $\tilde S$. Solving a linear set of equations with Toeplitz matrix requires
379: $\propto N\log^2_2(N)$ operations using superfast algorithms \cite{superfast}.
380: Perhaps diagonalization of Toeplitz matrices can be also speed up bellow
381: $\propto N^3$ operations.
382: The complexity of the second diagonalization is $\propto K^3$ operations
383: but here no eigenvectors need to be computed.
384: Remember that all these operations has to be done with high precision and are,
385: therefore, relatively slow.
386:
387: The diagonalization related difficulties has been overcome for long signals
388: \cite{Neuhauser,Taylor1,Taylor2}.
389: For such problems the frequency domain was effectively divided into smaller
390: intervals containing fewer frequencies and
391: the harmonic inversion, known there as filter diagonalization, was applied to
392: one interval at the time. Splitting the frequency domain introduces however
393: its own inaccuracies into the computed frequencies of the signal and filter
394: diagonalization methods are said to obey their own time-frequency uncertainty
395: relations. The total duration of the signal is limited from bellow by local
396: density of frequencies \cite{Taylor2} or by minimal and average frequency
397: distance \cite{Neuhauser}. It is not clear yet whether these restrictions
398: can be invalidated by improved precision of calculations.
399:
400: \section{Summary}
401: In this work I presented the application of the harmonic inversion
402: method to short time signals. I have demonstrated that the method has no
403: fundamental
404: limitations regarding the length of the signals. It works for arbitrarily short
405: signals if sufficient accuracy of the input data is provided.
406: In particular, it has been shown that it is possible
407: to extract energies from the short autocorrelation function generated by a
408: quantum system. The method can be also applied to short autocorrelation
409: functions squared which have a clear experimental interpretation. Its
410: measurement can be, in principle, carried out in arbitrarily short
411: interval of time and still all energies involved in its evolution can be
412: extracted to a desired accuracy.
413:
414:
415: \section{Acknowledgments}
416: I am grateful to Jacek Dziarmaga, Krzysztof Sacha, Jakub Zakrzewski and
417: George Zweig for many stimulating discussions.
418: Work supported by KBN grant 5 P03B 088 21.
419:
420: \appendix
421: \section{\label{app}Appendix: Estimate of $\lambda_{min}$}
422: The direct calculation of $\lambda_{min}$ is hard. What can be done
423: without much effort is the prediction of the general form of the expression
424: for this quantity.
425: The first step is the observation that since each matrix element of $S$
426: is a sum of $K$ exponents as in (\ref{sigd}) then $S$ can be written as
427: a sum of $K$ matrices $S= S_1 + S_2 + ... + S_K$, where each matrix $S_k$
428: depends only on one frequency $\omega_k$. Matrix $S_k$ is Hermitian and has
429: only one nonzero eigenvalue $N$ and corresponding eigenvector
430: $|s_k\rangle = [1, e^{i\omega_k\delta t}, e^{i\omega_k2\delta t},...
431: ,e^{i\omega_k(N-1)\delta t}]/\sqrt{N}$. Eigenvectors $|s_k\rangle$
432: form a nonorthogonal basis. Lets construct a $K\times N$ matrix $R$ by
433: filing its rows with vectors $|s_k\rangle$. Eigenvalues of $RR^\dagger$
434: multiplied by $N$ give the all nonzero eigenvalues of $S$.
435: The characteristic equation for eigenvalues of $RR^\dagger$
436: $$
437: a_K(\lambda/N)^K +...+a_1(\lambda/N) + \mbox{det}(RR^\dagger) = 0
438: $$
439: can be expanded at $\lambda = 0$ to the first order to give an
440: estimate for the smallest eigenvalue
441: $$
442: \frac{\lambda_{min}}{NK} \approx -\frac{\mbox{det}(RR^\dagger/K)}{a_1}.
443: $$
444: The matrix $RR^\dagger$ above is conveniently divided by $K$ to normalize its
445: trace to unity.
446: For $N=K$ matrix $R$ becomes square and
447: $\mbox{det}(R)\propto (\Omega T)^{K(K-1)/2}$ for $\Omega T \ll 1$.
448: Therefore, $\mbox{det}(RR^\dagger/K)\approx (c_1\Omega T)^{K(K-1)}$.
449: The parameter $a_1$ is given by sum of minors of
450: $RR^\dagger/K$ and $a_1 \approx (c_2\Omega T)^{(K-1)(K-2)}$.
451: $c_1$ and $c_2$ are functions of normalized frequencies $\omega_k/\Omega$.
452: Similar approximations are obtained if $N>K$.
453: All calculations are summarized by the estimate (\ref{mineig}).
454:
455:
456:
457:
458: \begin{thebibliography}{}
459: \bibitem{Asher} A. Peres, "Quantum Theory: Concepts and Methods", (Kluwer
460: Academic Publishers, 1995).
461: \bibitem{Aharonov1} Y. Aharonov and D. Bohm, Phys. Rev. {\bf 122}, 1649 (1961).
462: \bibitem{Aharonov2} Y. Aharonov, S. Massar, S. Popescu, Phys. Rev. A {\bf 66},
463: 052107 (2002).
464: \bibitem{Taylor2} V. A. Mandelshtam and H. S. Taylor. J. Chem. Phys.
465: {\bf 107}, 6756 (1997)
466: %\bibitem{FUTURE} Future publication.
467: \bibitem{golub} G. H. Golub and C. F. van Loan, "Matrix computations", The
468: John Hopkins University Press, Baltimore 1996.
469: %\bibitem{Carusotto} I. Carusotto, Y. Castin and J. Dalibard, Phys. Rev. A
470: %{\bf 63} 023610 (2001).
471: \bibitem{superfast} G. S. Ammar and W. B. Gragg, SIAM J. Matrix. Anal. Appl.,
472: {\bf 9}, 61 (1988).
473: \bibitem{Neuhauser} M. R. Wall and D. Neuhauser, J. Chem. Phys. {\bf 102}, 8011
474: (1995)
475: \bibitem{Taylor1} V. A. Mandelshtam and H. S. Taylor, J. Chem. Phys.
476: {\bf 106}, 5085 (1997).
477: \end{thebibliography}
478:
479:
480: \end{document}
481: