quant-ph0504229/wp2.tex
1: %documentstyle[preprint,prl,aps]{revtex}
2: \documentclass[12pt]{article}
3: \usepackage{graphicx,amssymb,amsmath,amsthm}
4: %\usepackage{time} %for displaying of the time
5: \usepackage{cite}
6: \newtheorem{theorem}{Theorem}
7: \newtheorem{definition}{Definition}
8: \newtheorem{proposition}{Proposition}
9: \newtheorem{property}{Property}
10: \newtheorem{corollary}{Corollary}
11: \newcommand{\Li}{[L_1, L_2]}
12: \newcommand{\Lp}{[L_1^{'}, L_2^{'}]}
13: \newcommand{\be}{\begin{equation}}
14: \newcommand{\ee}{\end{equation}}
15: \newcommand{\bea}{\begin{align}}
16: \newcommand{\eea}{\end{align}}
17: \newcommand{\ben}{\begin{eqnarray}}
18: \newcommand{\een}{\end{eqnarray}}
19: 
20: \newcommand{\spann}{{\mbox{\rm{span}}}}
21: 
22: \newcommand{\til}{\tilde}
23: \newcommand{\ov}{\overline}
24: \newcommand{\ra}{\rangle}
25: \newcommand{\la}{\langle}
26: \newcommand{\ii}{\hat{I}}
27: 
28: \newcommand{\sumn}{\sum_{n=1}^N}
29: 
30: 
31: \newcommand{\spa}{,\,}
32: \newcommand{\arexna}{\imath\frac{a n\pi x }{L}}
33: 
34: \newcommand{\subz}{n \in \N}
35: \newcommand{\sumz}{\sum_{\subz}}
36: \def\N{\mathbb{N}}
37: \newcommand{\h}{{\cal{H}}}
38: \newcommand{\W}{{\cal{W}}}
39: 
40: 
41: \setlength\unitlength{1cm}
42: \setlength\topmargin{-2cm}
43: \setlength\oddsidemargin{-0.1in}
44: \setlength\textwidth{17cm}
45: \setlength\textheight{23cm}
46: 
47: \title{On the truncation of the harmonic oscillator wavepacket } 
48: 
49: \author{L. Rebollo-Neira and S. Jain\\
50: Aston University,\\
51: Birmingham B4 7ET, United Kingdom}
52: 
53: \date{}
54: \begin{document}
55: \maketitle
56: \baselineskip = 1.7 \baselineskip
57: 
58: \begin{abstract}
59: We present an interesting result regarding the 
60: implication of truncating the wavepacket of the 
61: harmonic oscillator. We show that
62:  disregarding the  non-significant tails of 
63: a function which is the superposition of 
64: eigenfunctions of the harmonic oscillator 
65: has a remarkable consequence. Namely, there exit 
66: infinitely many different superpositions giving rise to 
67:  the same function on the interval. Uniqueness, in the 
68: case of a wavepacket, is restored by a postulate of  
69: quantum mechanics. 
70: 
71: PACS: 03.65.-w, 03.65.Ca
72: \end{abstract}
73: \section{Introduction}
74: We analyse the effect of truncating the wavepacket of the 
75: harmonic oscillator in the light of the {\em frame theory}. 
76: Such a theory, developed in 1952 by  Duffin and Shaffer 
77: in the context of harmonic analysis \cite{da} has been 
78: applied, for over fifteen years, to construct 
79: coherent states. More specifically, we 
80: should mention affine coherent states, 
81: also called wavelets, 
82: and  Weyl-Heisenberg coherent states, 
83: also  known as Gabor frames 
84: \cite{do,wa,ka1,ali3,ali1,do2,han,alib,rewf,rs}.
85: We recall in the 
86: next paragraph the general definition of frames and a
87: few properties, which is all what we need for the 
88: purpose of the present effort. For a complete treatment 
89: of frames we refer to \cite{yo,ftf,c2}.  
90: 
91: Given a Hilbert space $\h$, a family 
92: $\{\phi_n\}_{n\in \N}$ in $\h$ 
93: is called a {\em{frame}} for $\h$ if 
94: for every $f \in \h$ 
95: there exists a pair of constants
96: $0 < A \le B <\infty $ such that
97: \be
98: A \la  f , f \ra  \le  \sum_{\subz} | \la \phi_n , f \ra|^2 \le
99: B \la  f , f \ra.
100: \label{fc}
101: \ee
102: The constants $A$ and $B$ are called the
103: {\em {frame bounds}} and (\ref{fc}) the 
104: {\em frame condition}. 
105: From its definition it is clear that a frame
106: is a complete set,  since
107: the relations $ \la \phi_n, f\ra=0 \spa  \subz$ imply
108: that $f\equiv 0$. The removal of an element
109: from a frame leaves either a frame or an incomplete set.
110: A frame that ceases to be complete if
111: an arbitrary element $\phi_n$ is removed, is called
112: {\em{exact}}. This last property implies that only 
113: exact frames are bases;
114: in the general case
115: the  family $\{\phi_n\}_{n\in \N}$ may be
116:  over complete.
117: When the condition $A=B$ holds, the frame is said to be
118: a {\em{tight}} frame. Assuming that all the elements 
119: $\phi_n,\,  \subz $ are normalised to unity,
120: a tight frame is an orthogonal basis if and only if $A=B=1$.
121: Notice that this implies that a tight frame with
122: frame bounds $A=B=1$ is an orthogonal basis only if
123: the elements are normalised to unity, otherwise it
124: is a redundant frame. 
125: 
126: In this paper we introduce a class of redundant tight frames 
127: which are trivially obtained 
128: by redefining the functions of an orthonormal basis  
129: to be zero outside an interval. Such a restriction allows 
130: us to use the truncated functions to represent a given function  
131: vanishing outside the identical interval. This has 
132: a remarkable consequence, namely, the 
133: coefficients of the corresponding linear span are {\em{not unique}}. 
134: We discuss this important consequence in relation to the truncation of 
135: the wavepacket of an harmonic oscillator. We show that, 
136: by disregarding the
137: non-significant tails of a function which is  the
138: superposition of the harmonic oscillator eigenfunctions 
139: one creates a {\em{null space}}. As a consequence, the 
140: restriction of the wavepacket to the reduced interval can be 
141: realized by infinitely many different coefficients 
142: giving rise to the same function on the interval. This 
143: is, certainly, a striking result. 
144: Nevertheless, uniqueness can be 
145: restored by means of one of the postulates of 
146:  quantum mechanics. 
147: 
148: The paper is organised as follows: In Section II we 
149:  introduce the construction of tight frames for 
150: the Hilbert space of functions vanishing 
151: outside an interval 
152: by simple truncation of orthonormal basis functions. 
153: In Section III we apply these results to analyse the effect of 
154: truncating the wavepacket of the harmonic oscillator. The 
155: conclusions are drawn in Section IV. 
156: 
157: \section{Building tight frames from orthonormal bases}
158: The next proposition shows that, for the space of
159: square integrable functions vanishing outside an interval, 
160: one can construct tight frames by simple restriction of 
161: orthonormal functions to the corresponding interval. 
162: \begin{proposition}
163: \label{p1}
164: Let $\{\psi_n\}_{\subz}$ be an orthonormal basis for $L^2 \Li $, i.e.,
165: \be
166: \la \psi_m, \psi_n \ra =\int_{L_1}^{L_2} \psi_m^\ast(x)\psi_n(x)\, dx =
167: \delta_{m,n}
168: \ee 
169: and ${\chi}_{\Lp}$ the characteristic function for the 
170: interval ${\Lp}$, i.e, 
171: $${\chi}_{\Lp}(x)= \left\{\begin{array}{cll} 1
172:  \, \, & {\mbox{if}} \,\, x \in \Lp \\
173:                        0 \,\,& {\mbox{otherwise.}}
174: \end{array} \right.$$ 
175: Functions $\{\psi'_n\}_{\subz}$, obtained as
176: $\psi'_n(x) ={\chi}_{\Lp}(x)\psi_n(x) \spa \subz$, with
177:  $\Lp \subset \Li$, constitute a tight frame for the
178: subspace ${\cal{W}}$ of square integrable functions
179: vanishing outside $\Lp$.
180: \end{proposition}
181: \begin{proof}
182: On the one hand, since $\{\psi_n\}_{\subz}$ is an
183: orthonormal basis of $L^2\Li$, for all $f \in {L^2\Li}$ we have
184: \be
185: ||f||^2= \la f,f \ra=
186: \sumz \la f , \psi_n \ra \la  \psi_n,  f \ra.
187: \ee
188: On the other hand, for all $f \in {\cal{W}}$ it is true
189: that ${\chi}_{\Lp} f= f$ and we further have:
190: $$\sumz \la f , \psi_n' \ra \la  \psi_n', f \ra =
191:  \sumz \la f {\chi}_{\Lp}, \psi_n \ra \la \psi_n,
192: {\chi}_{\Lp} f \ra
193:   = ||f||^2,$$
194: which proves that $\{\psi_n\}_{\subz}$ is a redundant tight
195: frame for ${\cal{W}}$, since the elements $\psi_n'$ are not normalised 
196: to unity.
197: \end{proof}
198: %\begin{corollary}
199: %\label{fou}
200: %The set ${\chi}_{[-L\,\,L]} \frac{{1}}{\sqrt{2L}}e^{\arexna}$
201: %with $a\in (0\,,\,1)$
202: %is a redundant tight frame for the subspace of square
203: %integrable functions vanishing outside $[-L\,,\,L]$.
204: %The corresponding frame bound is $A=a^{-1}$.
205: %\end{corollary}
206: %\begin{proof}It is a direct consequence of the previous
207: %proposition,
208: %as $\frac{ \sqrt{a}}{\sqrt{2L}}e^{\arexna},\, \subz$ is an orthonormal
209: %basis for $\ii_{L^2[-\frac{L}{a}\,,\,\frac{L}{a}]}.$
210: %\end{proof}
211: We prove next that, through the finite subset of 
212: frames elements $\psi_1',\ldots,\psi_N'$
213: constructed as indicated in Proposition~\ref{p1} 
214: we can construct the orthogonal 
215: projection of $f$ onto 
216: ${\cal{S}}=\spann\{\psi_1,\ldots,\psi_N\}$, restricted to 
217: $\Lp$.
218: \begin{proposition}
219: Let $\psi_1',\ldots,\psi_N'$ be as defined in 
220: Proposition~\ref{p1}, and for each $f \in {\cal{W}}$ let us define  
221: a function $f^N$ as:
222: \be
223: \label{opp}
224: f^N= \sum_{i=1}^N \psi_i' \la \psi'_i, f \ra.
225: \ee
226: The function given in \eqref{opp} satisfies: 
227: $$f^N= \left\{\begin{array}{cll} \hat{P}_{\cal{S}}f(x)
228:  \, \, & {\mbox{if}} \,\, x \in \Lp \\
229:                    0 \,\,& {\mbox{otherwise,}}
230: \end{array} \right.$$
231: where $\hat{P}_{\cal{S}}$ stands for the orthogonal projector 
232:  operator onto ${\cal{S}}$.
233: \end{proposition}
234: \begin{proof}
235: Since $\{\psi_1,\ldots,\psi_N\}$ is an orthonormal set for
236: $L^2 \Li$, the orthogonal projection of $f \in L^2 \Li$  onto 
237: ${\cal{S}}$ is given as
238: \be
239: \hat{P}_{\cal{S}}f = \sum_{i=1}^N \psi_i \la \psi_i, f\ra. 
240: \ee
241: For $f \in {\cal{W}}$ it holds  that
242: $$\sum_{i=1}^N \psi_i \la \psi_i, f\ra= \sum_{i=1}^N \psi_i \la \psi_i', f\ra.$$Then 
243: \be
244: \hat{P}_{\cal{S}}f = \sum_{i=1}^N \psi_i \la \psi_i', f\ra,
245: \ee
246: and the proof follows by multiplication of both sides of the equation 
247: by ${\chi}_{\Lp}$. 
248: \end{proof}
249: A convenient property of tight frames is that 
250:  the coefficients of a linear span in a tight frame superposition
251: are obtained by inner products with the frame functions.
252: In our context this implies that  
253: the coefficients of an orthonormal 
254: expansion and the ones to span a function
255: in $\W$ by means of the frame in Proposition~\ref{p1}
256:  are computed in an equivalent manner.
257: The essential difference is that, as discussed below,
258: the coefficients in the frame superposition are 
259: {\em{not unique}}.
260:  
261: Let us consider the frame constructed in Proposition \ref{p1} and 
262: let us define ${\cal{S}'}=\spann\{\psi'_1,\ldots,\psi'_N\}$. For every 
263: $f\in {\cal{S}'} \subset {\cal{W}}$ we have: 
264: \be
265: \label{fd}
266: f=\sum_{n=1}^{N} 
267: c_n \psi'_n, \,\,\,\,\,\text{with}\,\,\,\,\,c_n=\la \psi_n', f \ra.
268: \ee
269: Since $f \in \W$ it is true that $c_n=\la \psi_n', f \ra= \la \psi_n, f\ra$ 
270: and the equivalence with the orthonormal case follows. However, the 
271: redundancy of the frame implies that the coefficients in \eqref{fd} are 
272: not unique. Indeed, since a redundant frame is linearly dependent,  
273: the following situation can occur:
274: \be
275: \label{ld}
276: 0=\sum_{n=1}^{N} c'_n \psi'_n,\,\,\,\,\,\text{for}\,\,\,\,\,\sum_{n=1}^N |c'_n|^2 \ne 0.
277: \ee
278: Taking inner product both sides of the equation on the left 
279: with each 
280:  $\psi'_m$   we have
281: \be 
282: 0=\sum_{n=1}^{N} c'_n \la \psi'_m , \psi'_n \ra,
283: \ee
284: that we can recast as 
285: \be
286: \label{ld2}
287: 0= G \vec{c'}
288: \ee
289: where $G$ is a matrix of elements 
290: $g_{m,n}= \la \psi'_m , \psi'_n \ra \,,n,m= 1,\ldots,N$ and
291: $\vec{c'}$ a vector, the component of which are 
292: the coefficients $c'_n,\, n= 1,\ldots,N$. 
293: Equations \eqref{ld} and \eqref{ld2} imply that the 
294: general form for \eqref{fd}  is
295: \be
296: \label{eq}
297: f=\sumn c_n \psi'_n + \sumn c'_n \psi'_n,
298: \ee
299: with $c_n, \,n=1,\ldots,N$ given in \eqref{fd} and 
300: $c'_n,\, n=1,\ldots,N$ the components of a vector 
301: $\vec{c'} \in$ null($G$). 
302: 
303: It is appropriate to stress the significance of 
304: Proposition~\ref{p1} when the interval $\Li$ is 
305: actually the whole real line. Then, for numerical 
306: calculations one is obliged 
307: to work on a finite domain. An important consequence of this 
308: fact will be discussed in the next section.
309: 
310: \section{On the truncation of the harmonic Oscillator 
311: wavepacket}
312: We show here that the results of the previous section 
313: are relevant to the analysis of the truncation of the
314: wavepacket of the harmonic oscillator. To this end 
315: we simulate two different situations, which are specially 
316: devised to illustrate the phenomenon we wish to 
317: discuss. 
318: 
319: Let us consider that a normalised to unity function 
320: $\Psi(x) \in \h$ is generated as linear 
321: superposition of the harmonic oscillator eigenfunctions, 
322: i.e., 
323: \be
324: \Psi(x)=
325: \sum_{n=1}^{N} c_n\frac{e^{-0.5x^2} H_n(x)}
326: {\sqrt{2^{n-1}(n-1)!\sqrt{\pi}}}, \,\,\,\,\,\,n,m=1,\ldots,N,
327: \label{wp}
328: \ee
329: where we have written 
330: $\psi_n(x) =\frac{e^{-0.5x^2}H_n(x)}{\sqrt{2^{n-1}(n-1)!\sqrt{\pi}}}$ in terms of the 
331:  Hermite polynomial $H_n(x)$ . The 
332: coefficients $c_n,\, n=1,\ldots,N$ in (\ref{wp}) 
333: are simulated according to the equation:
334: \be
335: \label{cn}
336: c_n=\frac{e^{-0.0032(n-80)^2}}{\sqrt{\sum_{n=1}^N e^{-0.0064(n-80)^2}}}.  
337: \ee
338: We consider $N=160$. 
339: The left graph of Figure~\ref{wp1} depicts 
340: the corresponding function for $x\in[-1,30]$. 
341: Since the values $\Psi(x)$ are 
342: very small outside this interval ($\Psi(-1)=8.7 \times 10^{-11}$ 
343: and $\Psi(30)= 2.7 \times 10^{-92}$) one does not commit a significant 
344: error by calculating expectation values using this domain. In order to 
345: illustrate this fact let us calculate $\ov{x}$ and $\ov{x^2}$ 
346: as
347: \ben
348: \ov{x}&=& \int_{-40}^{40} |\Psi(x)|^2 x\,dx = 12.56967570231863 
349: \nonumber \\
350: \ov{x}&= &\int_{-1}^{30} |\Psi(x)|^2 x\,dx = 12.56967570231863
351: \een
352: \ben
353: \ov{x^2}&= &\int_{-40}^{40} |\Psi(x)|^2 x^2\,dx = 158.4912423027778
354: \nonumber \\
355: \ov{x^2}&= &\int_{-1}^{30} |\Psi(x)|^2 x^2\,dx = 158.4912423027777.
356: \een
357: The difference in the values of  $\ov{x}$ 
358: cannot be observed in the given format and the values of $\ov{x^2}$
359: differ only in the last of the 16 digits.
360: Thus one could conclude that, 
361: for the purpose of computing expectation values, neglecting the   
362: tails of the distribution outside the interval $x\in[-1,30]$  is not harmful. 
363: However, the assumption that $|\Psi(x)|=0$ for 
364: $x \notin [-1,30]$ has a tremendous consequence: 
365: the coefficients of the  superposition (\ref{wp}) 
366: are thereby not unique.
367: In order to illustrate this we compute the vectors 
368: in the null space of 
369: matrix $G$ of elements 
370: \be
371: g_{m,n}= \int_{-1}^{30} H_m(x)H_n(x)\frac{e^{-x^2}}
372: {\sqrt{2^{m-1}(m-1)! 2^{n-1}(n-1)!}\pi}\,dx.
373: \ee
374: We use just one of the  eigenvectors spanning null$(G)$, 
375: say the vector $\vec{c'}$,  to construct the coefficients 
376: $c''_n=c_n + c'_n, \,n=1,\ldots,160$, with $c_n$ as in \eqref{cn} 
377: and $c'_n, n=1,\ldots,160$ the components of $\vec{c'} \in$ null$(G)$.
378: The right graph in Figure~\ref{wp1} plots the coefficients $c''_n$. 
379: We use now these coefficients to construct the function $\Psi''(x)$.  
380: By calculating 
381: \be
382: ||\Psi'' - \Psi|| =\sqrt{\int_{-1}^{30}|\Psi''(x) - \Psi(x)|^2\,dx}=
383:  1.4198\times 10^{-18}
384: \ee
385: we do not see any significant difference in the functions. 
386: However $||\vec{c''}-\vec{c}||=||\vec{c'}||=1$, which clearly shows that 
387: the  function are considering can be generated in
388: many different ways on the interval $[-1,30]$. Even considering the 
389: interval $[-10,40]$ null($G$) is still not empty. It would be empty for 
390: $[-15,40]$, though, but only  if the maximum 
391: number of states $N=160$ is maintained fixed.  
392: 
393: It should be noted that, although the `critical' interval 
394: depends on the number $N$ and a smaller value of $N$ would 
395: decrease the length of the critical interval, 
396: there is room for adjustments. 
397: Indeed, by considering zero the coefficients which do not 
398: intervene in the original superposition,
399: but allowing the transformation to have larger dimension, the critical interval is enlarged. 
400: Notice that this opens the possibility of producing the 
401: identical function by means of states that were not 
402:  present in the original superposition. The next example illustrates 
403: this situation. 
404: 
405: Consider that the coefficients of the  superposition 
406: (\ref{wp}) are now  simulated as 
407: \be
408: \label{cn2}
409: c_n=\frac{e^{-n}}{\sqrt{\sum_{n=1}^{20}e^{-2n}}}, \,\,\,\,\,\,\, n=1,\ldots,20
410: \ee
411: The corresponding function $\Psi(x)$ is plotted in the left graph of 
412:  Figure~\ref{wp2}. In this case the 
413: critical interval yielding lack of uniqueness is included in the 
414: main support of $\Psi(x)$. Hence, the restriction to such an interval 
415: is not possible. Now, increasing the value of $N$ to 130 for instance, 
416: and considering 
417: $c_n=0,\,n=31,\ldots,130$, the  function $\Psi(x)$ does not change. 
418: Nevertheless,  
419: we have a $130 \times 130$ matrix  $G$  constructed by  
420: extending the interval to one containing the most significant support of 
421:  $\Psi(x)$ (in this case $[-7\,,\,10]$). 
422: Taking one of the vectors in null$(G)$ we
423: construct the coefficients $\vec{c''}$ 
424: by an equivalent process as in the 
425: previous example. With this coefficients, 
426: plotted in  the right 
427: graph of Figure~\ref{wp2}, we 
428: construct the  graph on the left of Figure~\ref{wp2}. 
429: Notice that, the fact that coefficients $c''_n$ have significant values
430: for $n=31,\ldots,130$, implies that  $\Psi(x)$  
431: can be realized by states which were not  present in the  
432: original superposition. Nevertheless, the average energy is the same. 
433: Indeed,  
434: \be
435: \ov{E}= \la \psi'', \hat{H} \psi'' \ra=\sum_{n=1}^{N} \sum_{m=1}^{N} 
436: (c_n^{\ast} +{c'_n}^{\ast})\la \psi'_n, \hat{H} (c_m +c'_m) \psi'_m \ra
437: \ee
438: and, since by hypothesis $\sum_{m=1}^{N} c'_m \psi'_m,=0$, 
439: it follows that
440: \be
441: \ov{E}= \sum_{n=1}^{N} \sum_{m=1}^{N} c_n^{\ast} \la \psi'_n, \hat{H}
442: c_m \psi'_m \ra =  \la \psi',  \hat{H} \psi' \ra.
443: \ee
444: This result stresses the point that was 
445:  made initially. Namely, that without normalising the 
446: coefficients, {\em there  could be infinitely  many 
447: different ways of realizing the harmonic
448: oscillator wavepacket on a finite interval}. 
449: However, according to quantum mechanics
450:  each $|c''_n|^2$ represents the probability of 
451: finding the harmonic oscillator in the state $n$. Hence, to 
452:  be able to maintain this interpretation we must 
453:  impose the normalisation 
454: condition on the coefficients $c''_n$. As a consequence, we 
455: cannot use a vector $\vec{c'}$ of arbitrary norm. In fact, 
456: in order to construct a normalised vector 
457:  we need to consider $\vec{c''}= \vec{c}+D \vec{c'}$,
458:  with $\vec{c'}\in$ null$(G)$
459: and $D$ a constant to be determined by the condition 
460: $||\vec{c''}||^2=1$. Writing this condition explicitly 
461: we have
462: \be
463: ||\vec{c''}||^2= ||\vec{c}||^2  + D\la \vec{c},\vec{c'} \ra + 
464: D^\ast\la \vec{c'},\vec{c} \ra 
465: +|D|^2||\vec{c'}||^2 =1,
466: \ee
467: and, since $\vec{c'}$ and $\vec{c}$ are orthogonal to
468: each other we further have
469: \be
470: ||\vec{c''}||^2= ||\vec{c}||^2 + |D|^2||\vec{c'}||^2=1.
471: \ee
472: The value of $||\vec{c}||^2$ is fixed by the function 
473: $\Psi(x)$, in our case $||\vec{c}||^2=1$. 
474: Thus, the only possible solution to the above
475: equation is $D=0$.  This leads to  the conclusion that 
476:  amongst all the possible superposition 
477: giving rise to the  same 
478: function $\Psi(x)$ on the interval,
479: there is only {\em one} which is  consistent
480: with the physical 
481: significance of quantum mechanics. 
482: 
483: \section{Conclusions}
484: An interesting result, arising by limiting the domain of a
485: normalised function which is the superposition
486: of the harmonic oscillator eigenfunctions,
487: has been discussed. It was shown that,
488: when restricted to a finite interval,
489: such a function can be realized in many different ways.
490: Although the mean values of the
491: physical quantities are not affected in any significant
492: manner,  they can be the result of infinitely many
493: different combinations of eigenfunctions. Uniqueness is
494: restored by endowing the function with the significance of a
495: wavepacket. It was proved that, in that case, 
496: there is only {\em one} set of coefficients that can 
497: fulfil the normalisation condition.
498: 
499: \begin{figure}[h!]
500: \begin{center}
501: \mbox{\includegraphics[angle=0,width=6.5cm]{psi1.eps}
502: \includegraphics[angle=0,width=6.5cm]{coep1.eps}}
503: \caption{\small{The graph on the left 
504: represents the function $\Psi$ 
505: constructed by using the coefficients given in \eqref{cn} 
506: and also the coefficients $c''_n$ plotted
507: in the right graph.}}
508: \label{wp1}
509: \end{center}
510: \end{figure}
511: 
512: \begin{figure}[h!]
513: \begin{center}
514: \mbox{\includegraphics[angle=0,width=6.5cm]{psi2.eps}
515: \includegraphics[angle=0,width=6.5cm]{coep2.eps}}
516: \caption{\small{The graph on the left represents the function $\Psi$
517: constructed by using the coefficients given in \eqref{cn2}
518: and also the coefficients $c''_n$ plotted in the right graph.}}
519: \label{wp2}
520: \end{center}
521: \end{figure}
522: 
523: \begin{thebibliography}{99}
524: %\begin{references}
525: \bibitem{da} R. J. Duffin, A. C. Shaffer, ``A Class of Nonharmonic Fourier
526:  Series,''  Trans. Amer. Math. Soc. {\bf {72}}, 341-366 (1952).
527: \bibitem{do} I. Daubechies, A. Grossmann, Y. Meyer, ``Painless nonorthogonal
528: expansions'', Journal of Mathematical Physics, {\bf{27}}, 1271-1283 (1986).
529: \bibitem{wa} C. Heil, D. Walnut, ``Continuous and discrete wavelet transforms,''
530:  SIAM Rev. {\bf{31,}} 628-666 (1989).
531: \bibitem{ka1} G. Kaiser,
532:  {\em Quantum Physics, Relativity and Complex Space time:
533: Towards a New Synthesis }  (North-Holland, Amsterdam, 1990).
534: \bibitem{ali3}S. T. Ali, J. P. Antoine, J. P. Gazeau,
535: ``Square integrability of group representation on homogeneous spaces.
536: II. Coherent and quasi-coherent states. The case of the Poincar\'e group,''
537:   Ann. Inst. Henri Poincar\'e {\bf {55,}} 857-890 (1991).
538: \bibitem{ali1} S. T. Ali, J. P. Antoine, J. P. Gazeau, ``Continuous
539: Frames in Hilbert Space,''  Annals of Physics {\bf {222,}} 1-37 (1993).
540: \bibitem{do2} I. Daubechies, ``The Wavelets Transform, Time Frequency
541: Localization  and Signal Analysis,'' IEEE Trans. Inform. Theory
542: {\bf 36}, 961-1005 (1990).
543: \bibitem{han} D. Han, D. R. Larson,
544: {\em Frames, Bases and Group Representations},
545: (Memoirs of the  American Mathematical Society, Number 697, 2000)
546: \bibitem{alib} S. T. Ali, J. P. Antoine, J. P. Gazeau,
547: ``Coherent States, Wavelets and their generalizations'', 
548: (Springer, London, 2000).
549: \bibitem{rewf}  L. Rebollo-Neira, A. Plastino,
550:  ``The cross Wigner distribution as a generator of frames on the
551: the Euclidean plane from frames on the real line'',
552: Journal of Physics A: Mathematical and General,
553:  Vol(33, 15), 3053-3061, (2000).
554: \bibitem{rs}L. Rebollo-Neira,``Frames in two dimensions arising
555: from wavelet transforms'', Proceedings of the Royal
556: Society, series A, {\bf{457}}, 2013, 2079-2091. (2001).
557: \bibitem{yo} R. M. Young, {\em An introduction to Nonharmonic Fourier
558:  Series} (Academic Press, New York, 1980).
559: \bibitem{ftf} K. Gr\"ochening,
560: {\em Foundations of time-frequency analysis},
561: (Birkh\"auser, Boston, 2000)
562: \bibitem{c2}O. Christensen,  {\em An Introduction to
563: Frames and Riesz Bases},  (Birkh\"auser, Berlin, 2002)
564: \end{thebibliography}
565: %\end{references}
566: \end{document}
567: