quant-ph0412073/hi.tex
1: \documentclass[twocolumn, prl]{revtex4}
2: \usepackage{graphicx}
3: 
4: \begin{document}
5: %\draft
6: \title{Spectral analysis of short time signals}
7: \author{Zbyszek P. Karkuszewski}
8: 
9: \affiliation{
10: Institute of Physics, Jagiellonian University, Cracow, Poland\\
11: and\\
12: Los Alamos National Laboratory, Los Alamos, NM 87545.
13: }
14: 
15: \date{\today}
16: 
17: \begin{abstract}
18: The very old problem of extracting frequencies from time signals is addressed
19:   in the case of signals that are very short as compared to their 
20:   intrinsic time scales.
21: The solution of the problem is not only important to the classic signal processing but also helps to disqualify several common formulations of the quantum
22: mechanical time-energy uncertainty principle.
23: \\ \\
24: LAUR-04-5290
25: \end{abstract}
26: 
27: \maketitle
28: 
29: \section{Introduction}
30: The goal of many scientific efforts is to predict future evolution of
31: physical systems on the basis of their known past behavior.
32: One example of the very limited success of such an activity is
33: weather forecasts. Even if the history of all important parameters 
34: like temperature, pressure, humidity, wind velocity, etc. is known for many 
35: years back at almost every point on Earth, the reasonably accurate prediction 
36: of the coming weather conditions can be made only for several days.
37: One may argue that Earth's atmosphere is especially tough system to consider
38: due to its intrinsic instabilities: Even a tiny perturbation of air in one 
39: place can lead to huge changes of weather on a distant continent. In this
40: work we will not be able to deal with such instable systems either.
41: 
42: The other extreme is represented by very stable systems such as 
43: celestial objects. Centuries-long observations of the Moon and the Sun allowed 
44: ancient astronomers to predict accurately, Moon's phases, risings, settings 
45: and eclipses for coming millennia without any knowledge of
46: gravitational forces or Kepler's laws. Such precision was possible 
47: because the 
48: observations of the system had been made over much longer period than the 
49: system's characteristic time scales: days, (sidereal) months and years. 
50: 
51: Would the same quality predictions be possible if we observed the Moon 
52: just for fifteen minutes, i.e. for time much shorter than the shortest 
53: characteristic timescale?
54: 
55: The situation is even more interesting in quantum world where according
56: to some formulations of the time-energy uncertainty principle it would be
57: fundamentally impossible to accurately predict evolution of a quantum system
58: that has been observed only for a very short period of time.
59: 
60: In this work we show a practical way of achieving exact predictions of a future
61: based on a very short history of a system. Our predictions will be limited
62: only to the quantities evolution of which can be well described by finite 
63: Fourier series. This restriction is crucial. Still, there are many important
64: quantities that fall into this category.
65: 
66: 
67: Before introducing a method designed to perform such a task let us restate the
68: problem in a more formal way.
69: Suppose that a continuous quantity (a signal) $c(t)$ is given only in a finite 
70: time interval $t\in [0,T]$ and that it can be expressed in the following way 
71: \begin{equation}
72: c(t) = \sum_{k=1}^K d_k e^{-i\omega_k t},
73: \label{sigc}
74: \end{equation}
75: where $K$ is an integer, $d_k$ is a real positive \cite{ssp} amplitude 
76: and $\omega_k$ is a real (characteristic) frequency. The quantity $c(t)$ is, 
77: for simplicity, a complex function, but it can be made real by appropriate 
78: addition of terms with opposite sign frequencies $\omega_{k'}=-\omega_k$.
79: 
80: Our goal is to find unknown amplitudes $d_k$ and frequencies $\omega_k$ of 
81: given $c(t)$ in the case when length $T$ of the time interval is much smaller
82: than the smallest characteristic timescale $T\ll 1/\omega_{max}$. The method
83: outlined bellow does not require the prior knowledge of the number of the
84: Fourier components $K$ in (\ref{sigc}).
85: 
86: Mathematically, one can see that $c(t)$ is an analytic function of
87: time $t$ and, as such, can be uniquely extended beyond the interval $[0,T]$.
88: This means that, in principle, even for tiny $T$ it is possible to get all
89: amplitudes and frequencies from (\ref{sigc}) to any desired precision.  
90: Unfortunately it is difficult to solve this nonlinear problem analytically
91: and numerical methods cannot handle continuous signals due to infinite number
92: of data points.  One way of getting around this problem is to take only finite 
93: number of points at the cost of loss of the uniqueness of the extension.
94: From now on we will assume that the signal $c(t)$ is
95: known only at $N+1$ equidistant time points $t_n=n\delta t$ for $n=0,...,N$
96: and $t_N=T$. (\ref{sigc}) can be rewritten as a set of $N+1$ equations
97: \begin{equation}
98: \sum_{k=1}^K d_k e^{-i\omega_k t_n} = c_n,
99: \label{sigd}
100: \end{equation}
101: where $c_n\equiv c(n\delta t)$. This set has $2K$ real unknowns 
102: ($K$ amplitudes and $K$ frequencies) so the number of complex equations 
103: $N+1$ has to be equal or greater than $K$. This condition would have opened the possibility
104: of existence of a unique solution if the equations were linear in $\omega_k$ 
105: and $d_k$. It is not the case here. There will always be infinite number of solutions
106: to (\ref{sigd}) if the set is self-consistent and no solutions otherwise.
107: If one has an additional information about the range of frequencies $\omega_k$
108: in the problem (for instance Moon trajectory on the sky should not involve 
109: frequencies higher than 1/Hour) then the solution will be unique for small 
110: enough $\delta t$.
111: 
112: Similar problems, but for large $T$, are usually treated by a
113: discrete Fourier transform method (DFT).
114: That method {\it assumes} a grid of $N$ equidistant frequencies and solves 
115: (\ref{sigd}) only for $d_k$ as a linear set of equations. The spacing
116: of the assumed frequencies is proportional to $1/T$ and thus the method
117: is useless for estimating values of $c$ outside very short interval $T$.
118: 
119: The more challenging task of solving (\ref{sigd}) for both
120: amplitudes and frequencies is performed by, so called, harmonic inversion 
121: method.
122: 
123: \section{Harmonic inversion}
124: The historical roots of the method can be traced back to the two 
125: centuries old work by Gaspard Riche (Baron de Prony) \cite{Prony}.
126: A numerical implementation of the original approach can be found 
127: in \cite{Marple}. Here we present yet another derivation of the algorithm.
128: 
129: The key idea behind the harmonic inversion method is to replace the 
130: nonlinear problem (\ref{sigd}) with an eigenvalue problem for some
131: fictitious operator $\hat U$.
132: 
133: It is not necessary but very convenient to use basic formalism of quantum 
134: mechanics to introduce mathematical structure of the method. We will also 
135: benefit from this formalism when we turn our attention to quantum time-energy 
136: uncertainty principles.
137: 
138: Suppose that an evolution of a normalized quantum state $|\Phi_0\rangle$ 
139: is generated by an unitary operator $\hat U(\delta t)$ and 
140: $|\Phi_n\rangle \equiv \hat U^n(\delta t)|\Phi_0\rangle$. 
141: For every signal $c$ in (\ref{sigd}) there exists such an 
142: evolution operator $\hat U$ that the signal can be presented as 
143: the following autocorrelation function
144: \begin{equation}
145: c_n = \langle \Phi_0| \Phi_n\rangle.
146: \label{auto}
147: \end{equation}
148: It is enough to find eigenvalues $u_k=\exp(-i\omega_k\delta t)$ of the operator 
149: $U(\delta t)$ to find all characteristic frequencies. 
150: The matrix elements of $\hat U$ in the basis of states
151: $|\Phi_n \rangle$ can be expressed in terms of $c_n$ alone 
152: \begin{equation}
153: U_{ij}\equiv \langle \Phi_i | U| \Phi_j\rangle = 
154: \langle \Phi_0| U^{j-i+1} | \Phi_0\rangle = c_{j-i+1} 
155: \label{me}
156: \end{equation}
157: where $i,j=0,..., N-1$. 
158: The negative indices of $c$ in the equation above introduce no complication 
159: since $c_{-n}=c^*_n$. To obtain $K$ eigenvalues the dimension $N$ of the matrix 
160: $U$ must be equal or greater than $K$. This means that the number of
161: complex signal points $c_n$ required by the method exceeds the half of the 
162: number of unknowns. Vectors $|\Phi_n\rangle$ are not orthogonal,
163: thus the eigenequation for matrix $U$ takes the form
164: \begin{equation}
165: U |u_k\rangle = u_k S |u_k\rangle,
166: \label{gep}
167: \end{equation}
168: where $S$ is a matrix of scalar products $S_{ij}\equiv\langle
169: \Phi_i|\Phi_j\rangle=c_{j-i}$ with $i,j=0,...,N-1$. The rank of the matrix
170: $S$ gives the number of Fourier components of the signal $K$ for $N\ge K$.
171: 
172: The harmonic inversion method consists of two stages. First, one numerically 
173: solves the generalized eigenvalue problem (\ref{gep}), in order to get all 
174: frequencies. Second, when all frequencies in (\ref{sigd}) are known, a 
175: linear set of equations for the amplitudes $d_k$ is solved.
176: 
177: This brilliant algorithm has been developed and used by physicists and 
178: chemists for several years now, \cite{Neuhauser,Taylor1,Taylor2}. The 
179: only problem is that it fails when applied to short (small $T$) 
180: signals, \cite{Taylor2}. This limitation has been phrased in a form of 
181: Fourier-like uncertainty relation stating that the local density of 
182: frequencies that can be resolved by harmonic inversion must be smaller than the length $T$ of the time span of the signal \cite{Taylor2}.
183: 
184: Here we claim that the applicability of the method is not limited by the length
185: $T$ of the time interval but rather by noise affecting the signal $c(t)$.
186: 
187: In short, solving (\ref{gep}) requires calculating an inverse of the Hermitian 
188: scalar product matrix $S$. $S$ has $K$ positive and $N-K$ zero eigenvalues.
189: There are algebraic techniques to cope with singular matrices, see \cite{Golub}.
190: However, the smallest positive eigenvalue $\lambda_{min}$ becomes very small
191: for small $T$ or large $K$.
192: \begin{equation}
193: \frac{\lambda_{min}}{KN} \approx [f(\omega_k)T  \Omega]^{2(K-1)}
194: \label{mineig}
195: \end{equation}
196: where $f(\omega_k)$ is a function of order of unity,
197: and $\Omega$ stands for the greatest frequency magnitude present in the signal.
198: The formula above is valid only for short signals, $\Omega T \ll 1$, and the
199: expression in rectangular brackets is less than 1. 
200: 
201: Even small addition of noise to the signal may result in such a perturbation
202: of $\lambda_{min}$ that it will be impossible to distinguish it from, also 
203: perturbed, zero eigenvalues of the matrix $S$. When that happens harmonic 
204: inversion fails. The optimistic approach to (\ref{mineig}) notices that
205: $\lambda_{min}$ increases very fast with increasing $T$.
206: 
207: Assuming that the signal $c(t)$ is corrupted by noise $\eta(t), 
208: \quad\eta(t)\in [-\eta_{max}, \eta_{max}]$, the new 
209: signal $\tilde c_n = c_n + \eta_n$ has to be used to build matrices in 
210: (\ref{gep}). Harmonic inversion will extract all frequencies if
211: \begin{equation}
212: \lambda_{min} \ge 4N\eta_{max}
213: \label{nc}
214: \end{equation}
215: which is a necessary condition assuring that all $K$ positive eigenvalues 
216: of $S$ can be found.
217: Presence of noise leads to a set of perturbed frequencies 
218: $\tilde \omega_k$, which differ from $\omega_k$ 
219: \begin{equation}
220: |\tilde \omega_k - \omega_k|T \le \frac{2KN^2}{\lambda_{min}}\eta_{max}.
221: \label{ina}
222: \end{equation}
223: Notice that this is a "certainty" rather than uncertainty relation. 
224: The accuracy of frequencies found with the harmonic inversion can be made 
225: as high as needed by reducing the amplitude of noise $\eta_{max}$. 
226: This is the central result of this work.
227: 
228: In numerical studies where the method failed for small $T$ the role of noise was
229: played by roundoff errors. Figure \ref{Fig1} shows an exemplary application
230: of the harmonic inversion to a very short signal. The values of 
231: parameters in the example were deliberately chosen to expose the importance 
232: of the precision of the signal.
233: 
234: \begin{figure}[htb]
235: \includegraphics*[width=8.6cm]{fig1a.eps}
236: \caption{As an example, the harmonic inversion method was applied
237: to a signal with $K=10$ frequencies drawn from the interval $(0.5,1.0)$,
238: sampled at $N+1=14$ points with $T=0.01$ (upper plot). 
239: 85 digits precision, i.e. $\eta_{max}=10^{-84}$, was used.
240: Based on the points of the upper plot the signal is reconstructed (red line)
241: with the help of harmonic inversion.
242: Exact (black line) and reconstructed (red line) signals are indistinguishable
243: (middle plot).
244: In fact, the two lines start to differ by 1\% for $t>1,000,000$.
245: The discrepancies, that are due to roundoff errors (noise) of the initial 
246: points, are visible in the bottom plot.
247: In the example above $\lambda_{min}=4.07\times 10^{-78}$ and this justifies
248: the used 85 digits precision.
249: As a curiosity, the DFT algorithm applied to this signal would give just one 
250: frequency $\omega=0$, so that the prediction would be a horizontal
251: line at $|c(t)|=1$.}
252: \label{Fig1}
253: \end{figure}
254: 
255: Coming back to the problem of determining the future positions of the Moon in 
256: the sky just from a short observation we see that it is possible under
257: one condition: The observation has to be very accurate. The major obstacle
258: in achieving required accuracy would be refraction of the incoming moonbeams 
259: in the Earth's atmosphere, which results in a significant shift of the
260: apparent Moon's position with respect to the actual position \cite{Fantz}.
261: 
262: The method described above can be applied to any signal of classical
263: or quantum origin as long as it has the form of (\ref{sigd}).
264: When applied to quantum systems, it gives important insight to the problem
265: of validity of several formulations of the time-energy uncertainty relations.
266: 
267: \section{Quantum uncertainty relations}
268: Uncertainty principles play a central role in quantum mechanics.
269: They impose constraints on the states allowed by the theory. For example,
270: no quantum state can yield a product of momentum and position standard 
271: deviations smaller than $\hbar/2$ i.e. $\Delta p\Delta x\ge \hbar/2$,
272: where momentum $p$ and position $x$ are a pair of canonically conjugate 
273: operators $[x,p]=i\hbar$. This also means that if we prepare a large number of
274: quantum systems, all in the same state, and perform an {\it exact} 
275: measurement of position on half of them and an {\it exact} measurement of 
276: momentum on the other half then the spread of measured values must satisfy the 
277: uncertainty principle above.
278: 
279: Similar in form is the relation often provided for time $t$ and energy $E$
280: \begin{equation}
281: \Delta E\Delta t \ge \hbar/2,
282: \label{dedt}
283: \end{equation}
284: which is interpreted in various ways in literature. Unlike energy, 
285: time is just a parameter in quantum mechanics and the analogy between 
286: (\ref{dedt}) and other uncertainty relations cannot be carried too far.
287: One of the formulations of the time-energy uncertainty relation is presented
288: in textbooks on quantum mechanics \cite{Schiff,Messiah} as follows:
289: The measurement of energy of a quantum system performed over time $\Delta t$ 
290: inevitably results in inaccuracy $\Delta E$, so that (\ref{dedt}) is
291: satisfied. For more examples of absurd interpretations of the uncertainty
292: see \cite{Asher}.
293: 
294: It has been pointed out by Aharonov and Bohm \cite{Aharonov1} that the above
295: interpretation is wrong and one can measure accurate energy
296: of a quantum system in as short time interval as one pleases.
297: With the use of harmonic inversion we can provide a simple argument supporting
298: that claim. Moreover, the argument outlined bellow is much simpler and more
299: general than the one used in \cite{Aharonov1}.
300: 
301: Suppose that the state of an isolated quantum system spans over a finite
302: number $K$ of its energy eigenstates $|k\rangle$, so that at any time $t$
303: it assumes the form
304: \begin{equation}
305: |\Phi(t)\rangle = \sum_{k=1}^K a_k e^{-i \omega_k t} |k\rangle,
306: \end{equation}
307: where $a_k$ is a complex number and eigenenergy $E_k=\hbar\omega_k$.
308: 
309: We will show that, in principle, not only an expectation value of the 
310: energy but all energies $E_k$ that contribute to the evolution of the system
311: can be measured exactly no matter how short the measurement time is.
312: 
313: First, it has been experimentally proved that one can actually measure
314: a wave function $\Phi(x)$ in position representation. The technique used in
315: the measurement is known as quantum tomography. The relevant theory and 
316: applications are reviewed in \cite{Raymer}.
317: 
318: Second, being able to measure $\Phi(x)$ at different times implies that the time
319: autocorrelation function itself $c(t)=\langle\Phi(0)|\Phi(t)\rangle$ can be measured.
320: Moreover, accuracy of the estimates of $c(t)$ improves systematically with 
321: the increasing number of copies of the system on which such measurements are
322: performed. 
323: 
324: The third and the last step is to use harmonic inversion described above
325: to extract all eigenenergies and respective amplitudes from $c(t)$.
326: 
327: In short, the accuracy of obtained eigenenergies $E_k$ is only limited by
328: the precision of the measured autocorrelation function $c(t)$ and this precision
329: can be, in principle, as high as one needs.
330: 
331: \section{Unbreakable relation}
332: Many formulations of the time-energy uncertainty principle were invented using
333: intuition or dimensional analysis. And, as in the case discussed above, 
334: they are of limited applicability or simply wrong. There is, however,
335: one formulation that is rigorously derived from the quantum theory
336: \cite{Mandelshtam}.  
337: 
338: The Heisenberg uncertainty relations are manifestations of the following
339: theorem:
340: If $\hat A$ and $\hat B$ are two self-adjoint operators and a state
341: $|\Psi\rangle$ belongs simultaneously to the domains of $\hat A$, $\hat B$,
342: $\hat A\hat B$, $\hat B\hat A$, $\hat A^2$ and $\hat B^2$, then
343: \begin{equation}
344: \Delta A\Delta B \ge \frac{1}{2} |\langle [\hat A, \hat B]\rangle|,
345: \label{Hut}
346: \end{equation}
347: where $(\Delta A)^2\equiv \langle \hat A^2\rangle - \langle \hat A\rangle^2$
348: and $\langle ...\rangle\equiv \langle \Psi| ...|\Psi\rangle $. 
349: The uncertainty above is an intrinsic feature of the state $|\Psi\rangle$ and
350: has nothing to do with measurement inaccuracies.
351: 
352: This theorem applied to the position and momentum operators yields familiar
353: relation for standard deviations of position and momentum. In the case of time
354: and energy, however, (\ref{Hut}) results in $0\ge 0$ since time enters the
355: Schr\"odinger equation as a parameter rather than an operator. On the other
356: hand if $\hat A$ in (\ref{Hut}) is replaced with a Hamiltonian $\hat H$ and 
357: $|\Phi\rangle$ is not a stationary state, then using Heisenberg equation of
358: motion for an incompatible operator $\hat B$, 
359: $\mbox{d}\langle\hat B\rangle/\mbox{d}t = i/\hbar \langle [\hat H, \hat
360: B]\rangle$, one arrives at
361: \begin{equation}
362: \Delta H\frac{\Delta B}{\left | \frac{\mbox{d}\langle \hat B\rangle}
363: {\mbox{d}t}\right |} \ge \frac{\hbar}{2}
364: \end{equation}
365: the uncertainty relation of energy and something that has dimension of time --
366: the lifetime of the state $|\Psi\rangle$ with respect to the observable $B$.
367: This uncertainty relation cannot be broken or circumvented, it holds as
368: long as quantum mechanics is valid.
369: 
370: There is also a time-energy relation introduced for the case where only one
371: copy of a quantum system is used \cite{Aharonov2}.
372: 
373: \section{Summary}
374: 
375: In this work we show a practical way of extracting accurate frequencies and 
376: amplitudes from a signal that is available only over very short period of time.
377: The price is that the signal itself must be known to a very high
378: precision. The required precision is not achievable in real world
379: experiments where observed signals are both short and involve many frequencies.
380: The situation is somewhat better in numerical simulations where one has
381: more control over generated data. 
382: The role of this letter is to identify the source of the difficulty with
383: a spectral decomposition of short signals and present a tool to perform
384: such a decomposition when possible.
385: 
386: The method equips us also with a powerful argument against some
387: interpretations of time-energy uncertainty relations in quantum mechanics.
388: 
389: Harmonic inversion can also be applied to signals with continuous spectra.
390: In this case it provides lowest moments of the relevant frequency
391: distribution.
392: 
393: Harmonic inversion is superior to DFT even when long time signals are 
394: considered. Readers interested in testing the method on long signals and 
395: where complex frequencies might be involved should read \cite{Taylor2}.
396: 
397: 
398: \section{Acknowledgments}
399: I am grateful to Jacek Dziarmaga, Krzysztof Sacha, Jakub Zakrzewski and
400: George Zweig for many stimulating discussions.
401: Work supported by KBN grant 5 P03B 088 21.
402: 
403: 
404: 
405: \begin{thebibliography}{}
406: \bibitem{ssp} Negative amplitudes
407: would only require replacing an ordinary scalar product with a symmetric 
408: scalar product in this Letter, see \cite{Taylor2} for details.
409: \bibitem{Prony} Baron de Prony, J. E. Polytech., {\bf 1} 24 (1795).
410: \bibitem{Marple} S. L. Marple, Jr., Digital Spectral Analysis, Prentice-Hall,
411: Inc. (1987).
412: \bibitem{Neuhauser} M. R. Wall and D. Neuhauser, J. Chem. Phys. {\bf 102}, 8011
413: (1995)
414: \bibitem{Taylor1} V. A. Mandelshtam and H. S. Taylor, J. Chem. Phys. 
415: {\bf 106}, 5085 (1997).
416: \bibitem{Taylor2} V. A. Mandelshtam and H. S. Taylor, J. Chem. Phys.
417: {\bf 107}, 6756 (1997).
418: \bibitem{Golub} G. H. Golub and C. F. van Loan, "Matrix computations", The
419: John Hopkins University Press, Baltimore 1996.
420: \bibitem{Fantz} U. Fantz, Contemp. Phys. {\bf 45} 93 (2004).
421: \bibitem{Schiff} L. I. Schiff, Quantum Mechanics, McGraw-Hill Companies, 3rd ed. (1968).
422: \bibitem{Messiah} A. Messiah, Quantum Mechanics, Dover Publications (2000).
423: \bibitem{Asher} A. Peres, "Quantum Theory: Concepts and Methods", (Kluwer
424: Academic Publishers, 1995).
425: \bibitem{Aharonov1} Y. Aharonov and D. Bohm, Phys. Rev. {\bf 122}, 1649 (1961).
426: \bibitem{Raymer} M. G. Raymer, Contemp. Phys. {\bf 38} 343 (1997).
427: \bibitem{Mandelshtam} L. Madelshtam and I. Tamm, J. Phys. (USSR) {\bf 9} 249
428: (1945).
429: \bibitem{Aharonov2} Y. Aharonov, S. Massar, S. Popescu, Phys. Rev. A {\bf 66},
430: 052107 (2002).
431: \end{thebibliography}
432: 
433: 
434: \end{document}
435: