math-ph0611008/S.tex
1: \documentstyle[12pt,epsfig]{article}
2: \textheight 640pt \textwidth 480pt \oddsidemargin 0pt \voffset
3: -1.5cm
4: \title{Variationally Improved Spectral Method as an extremely accurate
5: technique for solving time-independent Schr\"{o}dinger equation}
6: \author{P. Pedram\thanks{Email: pedram@sbu.ac.ir}, M. Mirzaei, and S. S. Gousheh
7: \\ {\small Department of Physics, Shahid Beheshti University,
8: Evin, Tehran 19839, Iran}}
9: \begin{document}
10: \maketitle \baselineskip 24pt
11: % ----------------------------------------------------------------
12: \begin{abstract}
13: We introduce three distinct, yet equivalent, optimization
14: procedures for the Fourier Spectral Method which increase its
15: accuracy. This optimization procedure also allows us to uniquely
16: define the error for the cases which are not exactly solvable, and
17: this error matches closely its counterpart for the cases which are
18: exactly solvable. Moreover, this method is very simple to program,
19: fast, extremely accurate ({\it e.g.} an error of order $10^{-130}$
20: is usually obtainable as compared to the exact results), very
21: robust and stable. Most importantly, one can obtain the energies
22: and the wave functions of as many of the bound states as desired
23: with a single run of the algorithm. We first  thoroughly test this
24: method against an exactly solvable problem and then apply it to
25: two problems which are not exactly solvable, all for finding the
26: bound states of the time-independent Schr\"{o}dinger equation. We
27: present a detailed comparison between the results obtained by this
28: method and some of the more routine methods.
29: 
30: \vspace{5mm} 
31: \end{abstract}
32: \maketitle
33: % ----------------------------------------------------------------
34: \section{Introduction}
35: Eighty years after the birth of quantum mechanics
36: \cite{Schrodinger}, the Schr\"{o}dinger's famous equation still
37: remains a subject for numerous studies, aiming at extending its
38: field of applications and at developing more efficient analytic and
39: approximation methods for obtaining its solutions. There has always
40: been a remarkable interest in studying exactly solvable
41: Schr\"{o}dinger equations. At this point, we have to state that
42: traditionally the term ``exactly solvable" has been used in a
43: well-defined mathematical sense, meaning that the eigenvalues and
44: eigenfunctions of the Hamiltonian under consideration may be
45: expressed in an explicit and closed form \cite{CloseForm}. In this
46: sense, the exact solubility has been found for only a very limited
47: number of potentials, most of them being classified already by
48: Infeld and Hull \cite{Infeld} on the basis of the Schr\"{o}dinger
49: factorization method \cite{factorization}, which in turn appeared to
50: be a rediscovery of the formalism stated nearly 120 years ago by
51: Darboux \cite{Darboux}. However, a vast majority of the problems of
52: physical interest do not fall in the above category when we
53: formulate a more or less realistic model for them. Then we have to
54: resort to approximation techniques which can be analytic or numeric.
55: Examples of approximate analytic techniques would be the usual
56: perturbation method, the semiclassical or WKB approximation, and the
57: variational method \cite{bohm,landau}. In the perturbation method
58: the problem is divided into two segments. The main segment is
59: supposed to be exactly solvable. The second segment is supposed to
60: modify the solution obtained in the main segment only very slightly.
61: This modification can be obtained analytically for any desired
62: degree of accuracy. On the other hand, the numeric solutions could
63: be either perturbative in nature or completely numeric. Indeed, the
64: relevant Schr\"{o}dinger equation can always be solved numerically,
65: and in view of the immensely increased computational power, this
66: numerical solution could be extremely accurate. Such solutions could
67: serve the dual purpose of being a substitute for the exact solution
68: and being used as a comparison with experimental results, when
69: applicable, whose accuracy are also rapidly on the rise. The need
70: for such methods have stimulated development of more sophisticated
71: integration approaches, {\it e.g.} embedded exponentially-fitted
72: Runge-Kutta \cite{Runge-Kutta} and dissipative Numerov-type
73: \cite{Numerov} methods, as well as interesting techniques, such as a
74: relaxational approach \cite{Relax} based on the Henyey algorithm
75: \cite{Henyey}, an adaptive basis set using a hierarchical finite
76: element method \cite{hierarchical}, and an approach based on
77: microgenetic algorithm \cite{microgenetic}, which is a variation of
78: a global optimization strategy proposed by Holland \cite{Holland}.
79: To be more specific, Amore \cite{Amore} has used the variational
80: sinc collocation method to obtain the bound state solutions for a
81: Schr\"{o}dinger equation and has obtained accuracies of order
82: $10^{-42}$. Also Jentschura \cite{Ulrich} has used the Instanton
83: method for similar kind of problems and obtained accuracies of order
84: $10^{-180}$. Problems which consist of systems of coupled ODE's with
85: multiple delta function potentials can be solved by, for example,
86: the method of shooting to multiple fitting points and implementing
87: the delta functions as boundary conditions at those points
88: \cite{Gousheh}. Moreover, problems involving moving singularities
89: are always are difficult to handle. However a few algorithm have
90: been recently introduced to solve these problems (see for example
91: \cite{siamak2}). However, even in this simplest case, the success of
92: applying any direct numerical integration method depends on the
93: quality of initial guesses for the boundary conditions and energy
94: eigenvalues. Moreover, one usually encounters difficulties with the
95: intrinsic instabilities of typical problems, and rarely with the
96: existence of actual solutions which posses rapid oscillation.
97: 
98: Here we discuss an alternative technique for finding the bound
99: states of the time-independent Schr\"{o}dinger equation,
100: applicable for any potential which supports such states. This
101: method has the following seven distinct advantages: It is very
102: simple, fast, can be extremely accurate, does not have the
103: aforementioned difficulties with the choice of boundary
104: conditions. It is very robust and stable, {\it i.e.} it does not
105: have the instability problems due to the usual existence of
106: divergent solutions of most physical problems. These problems
107: usually produce difficulties for the spatial integration routines
108: such as Finite Difference Method (FDM). Moreover, we have
109: encountered problems in quantum cosmology whose exact solutions
110: posses very rapid oscillations \cite{wavepacket} which prevent any
111: successful application of more routines such as FDM, and we were
112: able to solve this problem using our alternative technique with
113: ease \cite{SM}. Here, our main goal is to improve this method by
114: an optimization procedure which first of all allows us to define
115: the errors uniquely and secondly decrease them drastically. This
116: method can also easily handle cases with mild moving
117: singularities, which also occurred in the aforementioned problem.
118: Finally, and perhaps most importantly, we can obtain the wave
119: functions and energies of as many of the bound states as desired
120: with a single run of the algorithm. This method, which was first
121: introduced by Galerkin over 90 years ago, consists of first
122: choosing a complete orthonormal set of eigenstates of a,
123: preferably relevant, hermitian operator to be used as a suitable
124: basis for our solution. For this numerical method we obviously can
125: not choose the whole set of the complete basis, as these are
126: usually infinite. Therefore we make the approximation of
127: representing the solution by a superposition of only a finite
128: number of the basis functions. By substituting this approximate
129: solution into the differential equation, a matrix equation is
130: obtained. The energies and expansion coefficients of these
131: approximate solutions could be determined by the eigenvalues and
132: eigenfunctions of this matrix, respectively. This method has been
133: called the Galerkin Method, and is a subset of the more general
134: Spectral Method (SM) \cite{SP1,SP2,SP3}. Spectral methods
135: generally fall into two broad categories. The “interpolating”, and
136: the “non–interpolating” method. In the first category, which
137: includes the Pseudospectral and the Spectral Element Methods, one
138: divides the configuration space into a set of grid points. Then
139: one demands that the differential equation be satisfied exactly at
140: a set of points known as the “collocation” or “interpolation”
141: points. Presumably, as residual function is forced to vanish at an
142: increasingly larger number of discrete points, it will be smaller
143: and smaller in the gaps between the collocation points. The
144: “non–interpolating” category includes the Lanczos tau-method and
145: the Galerkin method, mentioned above. The latter is the method
146: that we use and, in conformity with the usual nomenclature, we
147: shall simply refer to it as the Spectral Method. The interesting
148: characteristic of this method is that it is completely distinct
149: from the usual spatial integration routines, such as FDM, which
150: concentrate on spatial points. In SM the concentration is on the
151: basis functions and we expect the final numerical solution to be
152: approximately independent of the actual basis used. Moreover in
153: this method, the refinement of the solution is accomplished by
154: choosing a larger set of basis functions, rather than choosing
155: more grid points, as in the numerical integration methods. We
156: should note that we are implicitly assuming that the true solution
157: is expandable in any complete orthonormal basis such as the
158: Fourier basis. However, first of all this requirement is usually
159: satisfied for cases of physical applications, secondly we have
160: found that when this requirement is not satisfied the method does
161: not fail, but we loose overall accuracy.
162: 
163: In this paper we concentrate on the Fourier basis and refine SM by
164: introducing three distinct optimization methods. Then we show that
165: these three procedures, perhaps surprisingly, yield identical
166: results which we shall henceforth call the Variationally Improved
167: Spectral Method (VISM). This refinement adds two particularly
168: important advantages to SM: first we find a unique method to
169: define the error for the cases which are not exactly solvable, and
170: when tested on the exactly solvable cases matches closely with the
171: standard definition of the error. Second, This refinement, as we
172: shall show, usually improves the accuracy drastically. Then we
173: show an application of this amazingly powerful method to the
174: problem of finding the bound states of the time independent
175: Schr\"{o}dinger equation.
176: 
177: The remainder of this paper is organized as follows. In Section 2,
178: we present the underlying theoretical bases for the formulation of
179: the SM in connection with the problems of quantum nature. Then we
180: introduce our optimization procedures to obtain VISM. In Section
181: 3, we first use this method for the Simple Harmonic Oscillator
182: (SHO), which is an exactly solvable problem for three important
183: reasons: first to illustrate the method, second to show that the
184: three distinct optimization procedures yield the same results, and
185: third to test the method thoroughly. We then apply this method to
186: two perturbed harmonic oscillators, the first with quartic
187: anharmonic term, and second with a rapidly oscillating
188: trigonometric anharmonic term. Neither problem is exactly
189: solvable, and are particularly chosen to illustrate some of the
190: powerful features of this method. We compare the results for the
191: earlier case with those obtained by the usual first order
192: perturbation theory method, the conventional and a variationally
193: improved Sturmian approximation method \cite{mostafazadeh}, and a
194: highly accurate method \cite{exactanhrmonic}. The latter, though
195: in principle an approximate method, can determine rather precisely
196: the energy levels for this problem from the quantization of an
197: angle variable with an accuracy of order $10^{-90}$. The
198: trigonometric anharmonic problem has some very interesting
199: physical and mathematical aspects which we shall point out. In
200: Section 4, we state our conclusions and some final remarks.
201: 
202: \section{The Spectral Method}
203: Let us consider the time-independent one-dimensional
204: Schr\"{o}dinger equation,
205: \begin{equation}\label{Schrodinger}
206: -\frac{\hbar^2}{2m}\frac{d^2\psi(x)}{dx^2}+U(x)\psi(x)=E\psi(x),
207: \end{equation}
208: where $m$, $U(x)$, and $E$ stand for the reduced mass, potential
209: energy, and energy, respectively. Throughout this paper, we only
210: examine the bound states of this problem, i.e. the states which
211: are the square integrable. Therefore the general ODE that we want
212: to solve is a linear one that can be written in the form,
213: \begin{equation}\label{ODE}
214: -\frac{d^2\psi(x)}{dx^2}+\hat{f}(x)\psi(x)=\varepsilon\,\psi(x),
215: \end{equation}
216: where,
217: \begin{eqnarray}
218: \hat f(x)=\frac{2m}{\hbar^2}\, U(x),\label{f}
219: \hspace{2cm}\varepsilon=\frac{2m}{\hbar^2}\, E.
220: \end{eqnarray}
221: As mentioned before, any complete orthonormal set can be used for
222: the SM. We use the Fourier series basis as an example. That is,
223: since we need to choose a finite subspace of a countably infinite
224: basis, we restrict ourselves to the finite region $-L<x<L$. This
225: means that we can expand the solution as,
226: \begin{eqnarray}
227: \psi(x)=\sum_{i=1}^{2} \sum_{m=0}^{\infty} A_{m,i} \,\,\,
228: g_i\left(\frac{m \pi x}{L}\right), \label{eqpsitrigonometric}
229: \end{eqnarray}
230: where,
231: \begin{eqnarray}
232: \left\{
233:   \begin{array}{ll}
234:   g_1\left(\frac{m \pi
235: x}{L}\right)=\frac{1}{\sqrt{LR_m}} \sin\left(\frac{m \pi
236: x}{L}\right), &  \\
237:   g_2\left(\frac{m \pi
238: x}{L}\right)=\frac{1}{\sqrt{LR_m}}\cos\left(\frac{m \pi
239: x}{L}\right), & \\
240:   \end{array}\hspace{1cm}R_{m}=\left\{
241:   \begin{array}{ll}
242:     2 & \hspace{1cm}\hbox{m=0,} \\
243:     1 & \hspace {1cm}\hbox{otherwise.}
244:   \end{array}\right.
245: \right.
246: \end{eqnarray}
247: In the above choice of the basis we are implicitly assuming
248: periodic boundary condition. It is interesting to note that since
249: we are interested only in the bound states, we could choose only
250: the sine series but with the shifted problem $x\rightarrow x-L$.
251: Our shifted domain will be $0<x<2L$ and the potential function is
252: shifted accordingly. That is our new basis functions and the
253: differential equation are
254: \begin{eqnarray}
255:   g\left(\frac{m \pi
256: x}{2L}\right)=\frac{1}{\sqrt{L}} \sin\left(\frac{m \pi
257: x}{2L}\right),\hspace{1cm} \mbox{and} \hspace{1cm}
258: -\frac{d^2\psi(x)}{dx^2}+\hat{f}(x-L)\psi(x)=\varepsilon\,\psi(x),
259: \end{eqnarray}
260: For simplicity we shall refer to this alternate approach the
261: confinement boundary condition. Now going back to our original
262: formulation we can also make the following expansion,
263: \begin{eqnarray}
264: \hat f(x) \psi(x)=\sum_i \sum_{m} B_{m,i} \,\,\, g_i\left(\frac{m
265: \pi x}{L}\right),\label{eqV}
266: \end{eqnarray}
267: where $B_{m,i}$ are coefficients that can be determined once $\hat
268: f(x)$ is specified. By substituting Eqs.\
269: (\ref{eqpsitrigonometric},\ref{eqV}) into Eq.\ (\ref{ODE}) and
270: using the differential equation of the Fourier basis we obtain,
271: \begin{eqnarray}
272: \sum_{m,i}\left[\left(\frac{m \pi}{L}\right)^2
273: A_{m,i}+B_{m,i}\right]g_i\left(\frac{m \pi
274: x}{L}\right)=\varepsilon\sum_{m,i}A_{m,i}\,g_i\left(\frac{m \pi
275: x}{L}\right).\label{eqAB1}
276: \end{eqnarray}
277: Because of the linear independence of $g_i(\frac{m \pi x}{L})$,
278: every term in the summation must satisfy,
279: \begin{eqnarray}
280: \left(\frac{m \pi}{L}\right)^2 A_{m,i}+B_{m,i}=\varepsilon\,
281: A_{m,i}.\label{eqAB2}
282: \end{eqnarray}
283: It only remains to determine the matrix $B$. Using Eq.\
284: (\ref{eqV}) and Eq.\ (\ref{eqpsitrigonometric}) we have,
285: \begin{eqnarray}
286: \sum_{m,i} B_{m,i} g_i\left(\frac{m \pi
287: x}{L}\right)\,\,\,=\sum_{m,i} A_{m,i} \hat f(x) g_i\left(\frac{m
288: \pi x}{L}\right).
289: \end{eqnarray}
290: By multiplying both sides of the above equation by
291: $g_{i'}(\frac{m' \pi x}{L})$ and integrating over the $x$-space
292: and using the orthonormality condition of the basis functions, one
293: finds,
294: \begin{eqnarray}\label{eqbmi}
295: B_{m,i}=\sum_{m',i'} A_{m',i'} \int_{-L}^{L} g_i\left(\frac{m \pi
296: x}{L}\right)\,\,\,\hat f(x)\,\,\, g_{i'}\left(\frac{m' \pi
297: x}{L}\right)\,\,\, dx= \sum_{m',i'} A_{m',i'} C_{m,m',i,i'}\,\,,
298: \end{eqnarray}
299: Therefore we can rewrite Eq.\ (\ref{eqAB2}) as,
300: \begin{eqnarray}
301: \left(\frac{m \pi}{L}\right)^2 A_{m,i}+ \sum_{m',i'}
302: C_{m,m',i,i'}\,\, A_{m',i'}=\varepsilon\, A_{m,i}.\label{eqAC}
303: \end{eqnarray}
304: Where the coefficients $C_{m,m',i,i'}$ are defined by Eq.
305: (\ref{eqbmi}). It is obvious that the presence of the operator
306: $\hat f(x)$ in Eq.\ (\ref{ODE}), leads to nonzero coefficients
307: $C_{m,m',i,i'}$ in Eq.\ (\ref{eqAC}), which in principle could
308: couple all of the matrix elements of $A$. Therefore we have to
309: resort to a numerical solution. In general the number of basis
310: elements are at least countably infinite. The aforementioned
311: coupling of terms in the main matrix Eq.\ (\ref{eqAC}) forces us
312: to make the approximation of using a finite basis. It is easy to
313: see that the more basis functions we include, the closer our
314: solution will be to the exact one. By selecting a finite subset of
315: the basis functions, {\it e.g.} choosing the first $2N$ which
316: could be accomplished by letting the index $m$ run from 1 to $N$
317: in the summations, equation (\ref{eqAC}) can be written as,
318: \begin{eqnarray}
319: D\, A=\varepsilon\, A, \label{eqmatrix}
320: \end{eqnarray}
321: where $D$ is a square matrix with $(2N) \times (2N)$ elements. Its
322: elements can be obtained from Eq.\ (\ref{eqAC}). The eigenvalues
323: and eigenfunctions of the Schr\"{o}dinger equation are
324: approximately equal to the corresponding quantities of the matrix
325: $D$. That is the solution to this matrix equation simultaneously
326: yields $2N$ sought after eigenstates and eigenvalues. The only
327: problem which remains is to solve the eigenvalue problem Eq.\
328: (\ref{eqmatrix}), and to control the round-off errors. This is
329: often a serious issue for the usual spatial integration method
330: using double precision. However, we can easily overcome this
331: problem and obtain an extremely high precision. This is possible
332: if the following conditions are satisfied. First, the potential
333: energy and its derivatives should be smooth. Second, the
334: programming language (such as MATHEMATICA, MAPLE, etc) should be
335: capable of keeping the desired number of significant digits.
336: Third, when the integrations involved in the calculations (Eq.
337: \ref{eqbmi}) can be done analytically the coefficients in the
338: matrix equation (Eq. \ref{eqAC}) become exact and the  accuracy
339: increases. This also reduces the computational time drastically.
340: At this point it is worth mentioning that when the Hamiltonian
341: commutes with the parity operator, we can simplify our task of
342: solving Eq.\ (\ref{eqmatrix}) by separating the search for the
343: positive and negative parity solutions.
344: 
345: Now we can introduce our optimization procedure. We are free to
346: adjust two parameters: $2N$, the number of basis elements used and
347: the length of the spatial region, $2L$. This length should be
348: preferably larger than spatial spreading of all the sought after
349: wave functions. However, if $2L$ is chosen to be too large we
350: loose overall accuracy. As we shall show, the error decreases
351: extremely rapidly as the number of basis elements is increased.
352: However, it is important to note that for each $N$, $L$ has to be
353: properly adjusted. This is in fact an optimization problem and is
354: not a trivial task and requires some further analysis. We shall
355: denote this optimal quantity by $\hat{L}(N)$. We have come up with
356: three methods to determine this quantity. These are based on the
357: minimization of any of the eigenvalues (usually chosen to be those
358: of the ground state), its error, or the error of the corresponding
359: eigenfunction. For the minimization of the energy we proceed as
360: follows. For a few fixed values of $N$ we compute $E(N,L)$ which
361: invariably has an inflection point in the periodic boundary
362: condition and a minimum in the confinement boundary condition.
363: Interestingly enough this points exactly coincide. Therefore, all
364: we have to do is to compute the position of these inflection or
365: minimum points and compute an interpolating function for obtaining
366: $\hat{L}(N)$. Obviously the more points we choose the better our
367: results will be. We shall explain the other two optimization
368: methods in section 3. Once $\hat{L}(N)$ is computed, we can use
369: the method described above to compute the eigenvalues and
370: eigenfunctions. The necessary program is usually about 15 lines.
371: As we shall see, the addition of this refinement to SM can have
372: dramatic consequences. Throughout this paper we use VISM.
373: 
374: Computation of the relative error in the exactly solvable cases is
375: straightforward. For example for computing the relative error of
376: the eigenvalue, denoted by $\delta _E$, we only need to find the
377: absolute value of the difference between the result and the exact
378: one and divide by the latter. For cases which are not exactly
379: solvable, we compute the difference between the eigenvalues for a
380: given $N$ and those obtained with $N+1$, both lying on the
381: $\hat{L}(N)$ curve. We shall denote the error computed by this
382: procedure $\hat{\delta}_E$. We have computed $\hat{L}(N)$ for all
383: cases, and subsequently computed the eigenfunctions, eigenvalues
384: and their errors using this method, and  checked their validity in
385: the exactly solvable case of SHO. Obviously to obtain consistent
386: results we have to keep the same precision throughout the
387: calculations.
388: 
389: At this point we should mention that the only weakness of this
390: method that we have found is that, like most other routines, the
391: more discontinuous the potential or its derivatives, the less
392: accurate our solution will be. This is due to the fact that these
393: discontinuities would induce associated discontinuities in the
394: wave functions or their derivatives via the Schr\"{o}dinger
395: equation. In these cases we would need more basis functions, in
396: particular high frequency ones, to reproduce these features of the
397: wave functions.
398: 
399: \section{Some applications of the Spectral Method}
400: In this section, for illustrative purposes, we first apply VISM to
401: find the bound states of a Simple Harmonic Oscillator (SHO) which
402: is an exactly solvable case, for the reasons stated at the end of
403: the Introduction section. Briefly, the purpose is to readily check
404: the validity of our three distinct optimization procedures, which
405: includes our prescriptions for finding $\hat{L}(N)$, and compare
406: them. We can also check the overall accuracy of our results. We
407: then apply this method to two perturbed harmonic oscillators, the
408: first with quartic anharmonic term, and second with a rapidly
409: oscillating trigonometric anharmonic term. Neither problem is
410: exactly solvable. We compare our results for the earlier case with
411: some other reported results.
412: 
413: \textit{a. Simple Harmonic Oscillator}\\
414: The Schr\"{o}dinger equation for a SHO is,
415: \begin{equation}\label{SHO}
416: -\frac{\hbar^2}{2m}\frac{d^2\psi(x)}{dx^2}+\frac{1}{2}m \omega^2
417: x^2 \psi(x)=E\psi(x),
418: \end{equation}
419: where $\omega$ is the natural frequency of the oscillator.
420: Dividing both sides by $\hbar\omega/2$, we convert this
421: differential equation into the following dimensionless form,
422: \begin{equation}\label{SHO2}
423: -\frac{d^2\psi(x')}{dx'^2}+x'^2 \psi(x')=E'\psi(x'),\hspace{0.5cm}
424: \mbox{where} \hspace{0.5cm}x'=\sqrt{\frac{m\omega}{\hbar}}x,
425: \hspace{0.5cm} E'=\frac{2}{\hbar\omega}E.
426: \end{equation}
427: This differential equation is exactly solvable and its eigenvalues
428: and eigenfunctions, which are all bound states, can be easily
429: found analytically and are well known,
430: \begin{eqnarray}
431: E_n=(n+\frac{1}{2})\hbar\omega,\hspace{1cm}\psi_n(x)=\left(\frac{\omega}{\pi}\right)^{1/4}\frac{H_n(
432: \sqrt{\omega}x)} {\sqrt{2^n n!}}e^{-\omega
433: x^2/2},\hspace{1cm}n=\{0,1,2,...\},\label{SHO1}
434: \end{eqnarray}
435: where $H_n(x)$ denote the Hermite polynomials. Using VISM we can
436: calculate approximately the energy levels and the corresponding
437: eigenfunctions of this Hamiltonian. The computation of the errors
438: of the wave functions are analogous to that of the energy. We
439: divide the configuration space into $M$ grid points. Then, we
440: average the square of the absolute value of the difference between
441: the exact solution and that obtained by the VISM on the grid
442: points,
443: \begin{equation}
444: \delta_{E_n}=\frac{|E_n^{\mbox{\footnotesize{exact}}}-E_n^{SM}|}{E_n^{\mbox{\footnotesize{exact}}}},\hspace{2cm}
445: \delta^2_{\psi_n}=\frac{\sum_{i=1}^M
446: |\psi^{\mbox{\footnotesize{exact}}}_n(i)-\psi^{(N)}(i)|^2}{\sum_{i=1}^M
447: |\psi^{\mbox{\footnotesize{exact}}}_n(i)|^2}.\label{eqerror}
448: \end{equation}
449: In the left part of figure \ref{figNL} we show the ground state
450: energy computed using SM for the fixed value of the $N=5$ as a
451: function of $L$ using periodic boundary condition. In the right
452: part of figure \ref{figNL} we show the corresponding quantity
453: computed using confinement boundary condition. Note the existence
454: of the inflection and minimum points which exactly correspond to
455: each other in the figure. These points determine $\hat{L}(5)$. We
456: repeat this procedure for a few other values of $N$. After
457: plotting these values we can obtain an interpolating function
458: $\hat{L}(N)$. Our second optimization method is based on the
459: minimization of the error of the eigenvalue ($\delta_{E_0}$), and
460: to proceed we first exhibit a semi-log plot of the square of the
461: exact error for the ground state energy using SM in terms of $N$
462: and $L$ in Fig. \ref{fig3D}. Note the existence of a valley in
463: this figure indicating the optimal quantity $\hat{L}(N)$, which
464: can gives us the best values for the eigenvalues and
465: eigenfunctions using VISM. Our third optimization method is based
466: on the minimization of the error of the wave function
467: ($\delta_{\psi_0}$), and proceeds analogously to the previous
468: method. In Fig. \ref{figLN2} we show our results for $\hat{L}(N)$
469: obtained using the three aforementioned methods and their
470: interpolating function. Having determined $\hat{L}(N)$, we can
471: proceed to compute the bound states. We have checked the validity
472: of our results for the eigenvalues and their errors, and the
473: eigenfunctions using this method as compared to the corresponding
474: exact values. Table \ref{Table Oscillator} shows the results for
475: the first 10 eigenfunctions for $N=100$. Note the outstanding
476: accuracy of $\delta_E\approx 10^{-130}$ and
477: $\delta_{\psi_0}\approx 10^{-70}$. The values for $\hat{\delta}_E$
478: are also shown in Table \ref{Table Oscillator}. The values for
479: $\hat{\delta}_{\psi_0}$ could similarly be calculated, but are not
480: shown here. Now we want to exhibit explicitly the errors of the
481: ground state wave function, whose value is shown in
482: Table~\ref{Table Oscillator} for $N=100$, for some other values of
483: $N$. The left part of Figure \ref{fig SHO} shows the exact and
484: approximate ground state wave functions for $N=\{3,5,7\}$ using SM
485: with fixed un-optimized $L=10$. The right part of the same figure
486: shows the exact and approximate ground state wave functions for
487: $N=\{1,2\}$ using VISM with optimized
488: $\hat{L}=\{2.52479,3.04635\}$, respectively. Note that in the
489: un-optimized case (SM) for $N \geq 7$ the exact and approximate
490: wave functions are practically indistinguishable on the graph,
491: while in the optimized case (VISM) this occurs for $N \geq 2$.
492: This clearly shows the significance of our variational improvement
493: to SM. Here, we can report that, using VISM,
494: $\hat\delta_{E_0}=\{2\times10^{-4},4\times10^{-6},9\times10^{-9},2\times10^{-13}\}$
495: for $N=\{2,3,5,7\}$, respectively. Extrapolating from this, our
496: minuscule errors for $N=100$ are easily justified. In the left
497: part of Fig. \ref{fig time} we show a semi-log plot of the error
498: for the ground state energy, obtained using VISM, in terms of $N$.
499: As can be extrapolated from the figure significantly smaller
500: errors are easily obtainable. The interesting feature of this
501: graph is that the error decreases exactly exponentially and,
502: interestingly enough, the error of the simple Fourier expansion of
503: smooth functions has exactly the same behavior. We interpret this
504: as a strong signal that our procedure works well.
505: \begin{figure}
506:   % Requires \usepackage{graphicx}
507:  \centerline{
508: \begin{tabular}{cc}
509: \includegraphics[width=8cm]{1.eps}
510: \includegraphics[width=8cm]{1b.eps}
511: \end{tabular}}
512: \caption{Ground state energy  for SHO versus $L$ for $N=5$, using SM
513: in units where $\hbar\omega=2$. In the left part the results using
514: periodic boundary condition are shown and on the right using
515: confinement boundary condition.}\label{figNL}
516: \end{figure}
517: \begin{figure}
518:   % Requires \usepackage{graphicx}
519:   \centering
520:   \includegraphics[width=13cm]{2.eps}\\
521:   \caption{Semi-log plot of the square of the error of the ground state
522:   energy of SHO
523:   (in units where $\hbar\omega=2$) versus $N$ and $L$ using SM. Superimposed on top of the figure
524:   is the $\hat{L}(N)$ obtained from this figure, which indicates the direction of the valley. }\label{fig3D}
525: \end{figure}
526: \begin{figure}
527:   % Requires \usepackage{graphicx}
528:   \centering
529:   \includegraphics[width=10cm]{3.eps}\\
530:   \caption{$\hat{L}$ versus $N$ computed by our optimization methods involving minimization of $E_0$ in confinement boundary condition
531:   (boxes), minimization of $\delta E_0$ (circles), as shown in Fig. \ref{fig3D}, and minimization of $\delta_{\psi_0}$ (diamonds). Also shown in the figure
532:   is the interpolating function $\hat{L}(N)$.}\label{figLN2}
533: \end{figure}
534: \begin{table}
535:   \centering
536: \begin{tabular}{|c|c|c|c|c|c|} \hline
537:   $n$   & $E_n^{exact}$ & $E_n^{SM}$       & $\delta_E$  &$\hat{\delta}_E$&$\delta_{\psi}$ \\ \hline
538:    0    &1.      &    1.     & $2.6\times10^{-139}$ & $2.5\times10^{-139}$     &   $7.3\times10^{-70}$                \\ \hline
539:    1    &3.      &    3.     & $1.1\times10^{-133}$ & $1.1\times10^{-133}$     &   $3.5\times10^{-68}$                \\ \hline
540:    2    &5.      &    5.     & $5.9\times10^{-134}$ & $5.7\times10^{-134}$     &   $3.1\times10^{-67}$                \\ \hline
541:    3    &7.      &    7.     & $7.5\times10^{-129}$ & $7.2\times10^{-129}$     &   $7.9\times10^{-66}$                \\ \hline
542:    4    &9.      &    9.     & $2.2\times10^{-129}$ & $2.1\times10^{-129}$     &   $6.2\times10^{-65}$                \\ \hline
543:    5    &11.     &    11.    & $1.5\times10^{-124}$ & $1.4\times10^{-124}$     &   $1.0\times10^{-63}$                \\ \hline
544:    6    &13.     &    13.    & $3.1\times10^{-125}$ & $3.1\times10^{-125}$     &   $6.8\times10^{-63}$                \\ \hline
545:    7    &15.     &    15.    & $1.3\times10^{-120}$ & $1.3\times10^{-120}$     &   $9.5\times10^{-62}$                \\ \hline
546:    8    &17.     &    17.    & $2.4\times10^{-121}$ & $2.3\times10^{-121}$     &   $6.1\times10^{-61}$                \\ \hline
547:    9    &19.     &    19.    & $7.1\times10^{-117}$ & $6.8\times10^{-117}$     &   $6.8\times10^{-50}$                \\ \hline
548: \end{tabular}
549:   \caption{The results for the first 10 eigenvalues and eigenfunctions of the SHO in units where
550:    $\hbar\omega=2$. $E^{SM}_n$ are values obtained using VISM with $N=100$. The values of $\delta_E$
551:    and $\delta_{\psi}$ refer to the difference between the quantities computed using VISM and the corresponding
552:    exact quantities as given by Eq. (\ref{eqerror}). Note the good correspondence between $\hat{\delta}_E$ and
553:    $\delta_E$ shows the consistency of our method. Obviously we could not display all the digits of $E^{SM}_n$
554:   due to insignificant errors. The total computation time was about $50$ seconds on a typical Pentium 2.4 GHz
555:   machine for obtaining the first 200 eigenvalues and eigenfunctions, which were all obtained with a same single
556:   run of the algorithm. It is worth mentioning that choosing N=30 we obtain $\delta_E\approx10^{-15}$ with
557:   computational time of 0.5 Sec. } \label{Table Oscillator}
558: \end{table}
559: \begin{figure}
560: \centerline{
561: \begin{tabular}{cc}
562: \includegraphics[width=7.5cm]{5.eps}
563: \includegraphics[width=7.5cm]{4.eps}
564: \end{tabular}}
565: \caption{Left, the exact and approximate ground state wave
566: functions of SHO for $N=\{3,5,7\}$ using SM with fixed
567: un-optimized $L=10$. Note that for $N \geq 7$ the exact and
568: approximate wave functions are almost indistinguishable on the
569: graph. Right, the exact and approximate ground state wave
570: functions of SHO for $N=\{1,2\}$ using VISM with optimized
571: $\hat{L}=\{2.52479,3.04635\}$, respectively. Note that for $N \geq
572: 2$ the exact and approximate wave functions are practically
573: indistinguishable on the graph. Also note the drastic effect of
574: our optimization procedures.} \label{fig SHO}
575: \end{figure}
576: 
577: In Fig. \ref{fig time} we show a semi-log plot of the corrected
578: computation time versus $N$. Note that the curvature of this plot is
579: negative and considering this together with the plot of $\ln(
580: \delta_E)$ shows how the effective efficiency of our program
581: increases with $N$.
582: \begin{figure}
583:   % Requires \usepackage{graphicx}
584:  \centerline{
585: \begin{tabular}{cc}
586: \includegraphics[width=7.5cm]{6.eps}
587: \includegraphics[width=7cm]{7.eps}
588: \end{tabular}}
589: \caption{Left, semi-log plot of the error for the first eigenvalue
590: of the SHO obtained by VISM using various number of basis
591: functions ($2N$ equals the number of basis functions used). Right,
592: semi-log plot of the computation time versus $N$ for the ground
593: state of SHO.}\label{fig time}
594: \end{figure}
595: %\pagebreak
596: 
597: %*************************************Anharmonic Oscillator******************
598: \textit{b. Anharmonic Oscillator with a quartic term}\\
599: Now we apply this method to an anharmonic oscillator which has a
600: quartic term. This is probably one of the most famous problem in
601: quantum mechanics which is not exactly solvable and is used as at
602: least as a toy model. The Schr\"{o}dinger equation for this model
603: is,
604: \begin{equation}\label{ASHO}
605: -\frac{\hbar^2}{2m}\frac{d^2\psi(x)}{dx^2}+\left(\frac{1}{2}m\omega^2
606: x^2+\epsilon x^4\right) \psi(x)=E\psi(x).
607: \end{equation}
608: Using VISM we can find the bound states energy spectrum and the
609: corresponding eigenfunctions of this Hamiltonian. The results that
610: we have obtained using $N=100$ are extremely accurate
611: ($\delta_{E_0}\approx 10^{-120}$) (Table \ref{Table anharmonic}).
612: We have also compared our method and results with some other more
613: or less routine methods such as the zero and first order
614: perturbation theory (Table \ref{Table anharmonic}), the
615: conventional Sturmian approximation of Ref. \cite{CSA}, the zero,
616: first and second order variational sturmian approximation of Ref
617: \cite{mostafazadeh}, and the highly accurate values of Ref.
618: \cite{exactanhrmonic}. The results of these comparisons is that
619: the seven advantages of our method mentioned in the introduction
620: are clearly justified. We just have to mention that the reported
621: accuracy in Ref. \cite{exactanhrmonic} was only 90 significant
622: digits. The accuracy of the other methods were even lower, that is
623: not better than ${\mathcal{O}}(10^{-4})$.
624:  {\footnotesize
625: \begin{table}
626: \begin{tabular}{|c|c|c|c|c|c|c|} \hline
627: $n$&$E_n^{SM}$                                 & $SD$
628: &$E_n^{(0)}$& $\frac{|E_n^{(0)}-E_n^{SM}|}{E_n^{SM}}$ &$E_n^{(1)}$
629: & $\frac{|E_n^{(1)}-E_n^{SM}|}{E_n^{SM}}$  \\ \hline
630: 0&1.0652855095437176888570916287890930843044864178189& 124  &1
631: &0.061     &1.075          &0.009      \\ \hline
632: 1&3.3068720131529135071281216846928690495946552097516& 121  &3
633: &0.032     &3.450          &0.043 \\ \hline
634: 2&5.7479592688335633047335031184771312788809760663913& 120  &5
635: &0.13      &5.975          &0.039         \\ \hline
636: 3&8.3526778257857547121552577346436977053951052605059& 118  &7
637: &0.16      &8.875          &0.063 \\ \hline
638: 4&11.098595622633043011086458749297403250621831282348& 118  &9
639: &0.19      &12.08          &0.088\\ \hline
640: 5&13.969926197742799300973433956842133961140713634295& 116  &11
641: &0.21      &15.58          &0.11\\ \hline
642: 6&16.954794686144151337692616508817134375549987258361& 114  &13
643: &0.23      &19.38          &0.14\\ \hline
644: 7&20.043863604188461233641421107385111570572266905826& 115  &15
645: &0.25      &23.48          &0.17 \\ \hline
646: 8&23.229552179939289070647087434323318243534938599487& 112  &17
647: &0.27      &27.88          &0.20      \\ \hline
648: 9&26.505554752536617417469503006738723676057932189542& 110  &19
649: &0.28      &32.58          &0.23\\ \hline
650: \end{tabular}
651:   \caption{The first 10 energy levels of the Schr\"{o}dinger equation
652:   (Eq. (\ref{ASHO})) whose dimensionless form is $(-d^2/dx'^2+x'^2+\epsilon' x'^4)\psi(x')=E'\psi(x')$.
653:   We have chosen the parameter $\epsilon'=4 \epsilon/(m \omega^4)=1/10$ and exhibited the results for the energies in
654:   units where $\hbar\omega=2$. $E_n^{SM}$ are the values obtained using VISM with $N=100$, and $SD$
655:   denotes the significant digits. For space limitations only
656:   the first 50 significant digits are displayed.
657:   $E_n^{(0)}$ and $E_n^{(1)}$ denote the energy eigenvalues obtained
658: using zero and first order perturbation theory, respectively.
659: Obviously the accuracy of these methods are far inferior compared
660: to VISM.}\label{Table anharmonic}
661: \end{table}}
662: %*************************************Oscillatory****************************
663: 
664: \textit{c. Harmonic Oscillator perturbed by rapid oscillations}\\
665: Now we want to solve an example which exhibits another less
666: explored and powerful feature of VISM. As mentioned in the
667: introduction this method can easily handle problems whose
668: solutions exhibit rapid oscillations, in sharp contrast to the
669: spatial integration methods. A rather interesting example that we
670: have constructed for this purpose is a harmonic oscillator
671: perturbed by rapid oscillations whose precise  Schr\"{o}dinger
672: equation is given by,
673: \begin{equation}\label{SHOR}
674: -\frac{\hbar^2}{2m}\frac{d^2\psi(x)}{dx^2}+\left(\frac{1}{2}m
675: \omega^2 x^2+\alpha \cos(\beta \pi x)\right) \psi(x)=E\psi(x),
676: \end{equation}
677: where $\omega$ is the natural frequency of the oscillator and
678: $\alpha$ and $\beta$ are arbitrary constants.  This differential
679: equation is not exactly solvable and for high frequency
680: perturbations $\beta$ the spatial integration routines like FDM
681: would have serious difficulties in these situations. They should
682: consider many spatial points for overcoming the rapid oscillations
683: of the potential which increases the time and round off errors of
684: the routine and decreases the efficiency and stability of the
685: method. However, this is not a case for the VISM which can also
686: handle these type of potentials very easily. For large $\beta$ the
687: behavior of the potential is very oscillatory and centered around
688: the curve $\frac{1}{2} m\omega x^2$. As expected all the
689: eigenstates of this problem are bound states. The results for the
690: ground state are shown in the Fig.\ \ref{PHO}. In the left part of
691: the figure we show the full potential $U(x)$, the ground state
692: wave function ($\psi_0(x)$), and a zoomed box highlighting the
693: fine structural behavior of the wave function. In the right part
694: of the figure we show the ground state energy $E_0$ versus $N$.
695: Note for $N$ smaller than $\beta \hat{L}$ (100 here) VISM is not
696: sensitive enough to respond to the rapidly oscillating part of the
697: potential and the results are very close to those of the
698: (unperturbed) SHO. As is apparent from the figure, for $N$ even
699: slightly larger than $\beta \hat{L}$, the VISM easily incorporates
700: the rapidly oscillating part of the potential and rapidly
701: approaches the exact energy eigenvalue as $N$ increases. This
702: explains the sharp drop in the graph of $E_0(N)$. In particular
703: for $N=150$ we obtain $\hat{\delta}_{E_0}\approx 10^{-50}$. The
704: physics of this phenomenon is very clear: When the set of basis
705: functions is large enough to include those whose frequencies are
706: at least as large as the ones induced in the wave function by the
707: oscillations or discontinuities in the potential, via the
708: Schr\"{o}dinger equation, VISM can easily finds the accurate
709: eigenfunctions and the corresponding eigenvalues. the fine
710: structure of the ground state energy exhibited in the left part of
711: the figure is also very interesting. The slight ripples in the
712: wave function are produced by the conflicting tendencies of the
713: wave function to accumulate in the rapidly occurring wells of the
714: perturbed part of the potential and that of the second derivative
715: operator to smooth out the wave function.
716: \begin{figure}
717: \centering
718: \begin{tabular}{ccc} \epsfig{figure=9.eps,width=9cm}
719: &\hspace{1.cm}& \epsfig{figure=10.eps,width=7cm}
720: \end{tabular}
721: \caption{Left, the potential of the harmonic oscillator perturbed
722: by rapid oscillations, whose Schr\"{o}dinger equation (Eq.
723: (\ref{SHOR})) in dimensionless form is $(-d^2/dx'^2+x'^2+\alpha'
724: \cos(\beta'\pi x'))\psi(x')=E'\psi(x')$. We have chosen the
725: parameters $\beta'=\sqrt{2/m} \beta/\omega=10$, and
726: $\alpha'=(2/\hbar\omega)\alpha=10$ and exhibited the results for
727: the energies in  units where $\hbar\omega=2$. Superimposed on the
728: same graph is the ground state wave function calculated using VISM
729: with $N=150$. The zoomed box exhibits the behavior of the wave
730: function and the potential for $-0.7<x<0.7$. Right, the ground
731: state energy versus $N$ for the same parameters as in the left
732: figure.} \label{PHO}
733: \end{figure}
734: 
735: 
736: \section{Conclusions}
737: We have introduced three distinct yet equivalent variational
738: improvements for the Fourier Spectral Method and used them as an
739: extremely accurate method for obtaining the energies and wave
740: functions of the bound states of the time-independent
741: Schr\"{o}dinger equation. In this method a finite basis is used for
742: approximating the solutions. Our variational improvement of the
743: method is to calculate an optimized spatial domain for a given
744: number of basis elements, denoted by $\hat{L}(N)$. The three
745: optimization procedures are based on minimizing the eigenvalues, the
746: error of the eigenvalues, and on minimizing the error of the
747: eigenstates. Our refinement schemes usually improve the accuracy of
748: SM drastically, as can be seen in Fig. \ref{fig SHO}, for example.
749: This improvement increases rapidly with $N$. In particular, when the
750: problem is not exactly solvable there does not seem to be any other
751: canonical way to fix $L$ and the errors. We applied this method to
752: an exactly solvable problem and easily found an extraordinarily good
753: agreement with the exact solutions (errors of order $10^{-130}$). In
754: the quartic anharmonic oscillator case which is not exactly
755: solvable, the accuracy of this method is much higher than some of
756: the more conventional methods such as the perturbation method, the
757: conventional sturmian approximation, and the variational sturmian
758: approximation. In the problem of SHO perturbed by trigonometric
759: anharmonic term, we observed how easily and accurately VISM handles
760: problems involving potentials with very rapid oscillations. To
761: summarize, this method is very simple, fast, extremely accurate in
762: most cases, very robust and stable, can easily handle solutions with
763: rapid oscillations and moving singularities, and there is no need to
764: specify the boundary conditions on the slopes. Most importantly, one
765: can obtain the energies and the wave functions of as many of the
766: bound states as desired with a single run of the algorithm. The main
767: sources of error are using too few number of basis elements and
768: having potentials with major discontinuities. When the latter dose
769: not exist we can obtain extraordinary good results, {\it e.g.} 130
770: significant digits, by choosing appropriate parameters. This method
771: can be easily extended to $D$ dimensional Schr\"{o}dinger equation.
772: For example, We have solved 2-D Harmonic Oscillator $V(x,y)=x^2+y^2$
773: and QCD potential $V(x,y)=x^2 y^2$ using VISM and found extremely
774: highly accurate results \cite{2D}. In 2-D SHO case the accuracy with
775: 15 significant digits can be easily obtained with only 22 basis
776: functions, and this shows the efficiency of VISM.
777: %\pagebreak
778: \vskip20pt\noindent {\large {\bf Acknowledgement}}\vskip5pt\noindent
779: The Authors thank F. M. Fernandez and A. Turbiner for their useful
780: comments. \vskip10pt
781: % ----------------------------------------------------------------
782: \begin{thebibliography}{99}
783: \bibitem{Schrodinger}E. Schr\"{o}dinger, Quantisierung als Eigenwertproblem. (Erste Mitteilung.),
784: Ann. Phys. (Leipzig) 79 (1926), 361–376; Quantisierung als Eigenwertproblem. (Zweite Mitteilung.),
785: Ann. Phys. (Leipzig) 79 (1926), 489–527; Quantisierung als Eigenwertproblem. (Dritte Mitteilung.),
786: Ann. Phys. (Leipzig) 80 (1926), 437–490; \"{U}ber das Verh\"{a}ltnis der Heisenberg-Born-
787: Jordan'schen Quantenmechanik zu der meinen. Ann. Phys. (Leipzig), 79 (1926),
788: 734-756; An Undulatory Theory of the Mechanics of Atoms and Molecules, Phys. Rev.
789: 28, 1049 (1926).
790: \bibitem{CloseForm}G. Junker and P. Roy, Conditionally exactly solvable potentials: A supersymmetric construction method,
791: Ann. Phys. (Leipzig) 270, 155 (1998).
792: \bibitem{Infeld}L. Infeld and T.D. Hull, The Factorization Method, Rev. Mod. Phys. 23, 21 (1951).
793: \bibitem{factorization}E. Schr\"{o}dinger, A method of determining quantum mechanical eigenvalues and eigenfunctions,
794: Proc. R. Ir. Acad. Sect. A, Math. Astron. Phys. Sci. 46, 9–16 (1940);
795: Further studies on solving eigenvalue problems byfactorization, 47A, 183–206 (1941).
796: \bibitem{Darboux}G. Darboux, Sur une proposition relative aux \'{e}quations lin\'{e}arires, C R. Acad. Sci. III 94, 1456 (1882).
797: \bibitem{bohm}D. Bohm, Quantum Theory (Prentice-Hall, New York, 1951);
798: J. J. Sakurai, Modern Quantum Mechanics (Addison-Wesley, Reading,
799: Massachusetts, 1994).
800: \bibitem{landau}L. D. Landau and E. M. Lifshitz, Quantum Mechanics (Butterworth Heinemann,oxford,1977).
801: \bibitem{Runge-Kutta}G. Avdelas, T.E. Simos, and J. VigoAguiar, An embedded exponentially-fitted Runge-Kutta method for the numerical solution of
802: the Schr\"{o}dinger equation and related periodic initial-value,
803: problems Comput. Phys. Commun. 131, 52 (2000).
804: \bibitem{Numerov}G. Avdelas and T.E. Simos, Dissipative high phase-lag order Numerov-type methods for the numerical solution of the
805: Schr\"{o}dinger equation, Phys. Rev. E 62, 1375 (2000).
806: \bibitem{Relax}J.D. Praeger, Relaxational approach to solving the Schr\"{o}dinger
807: equation, Phys. Rev. A 63, 022115 (2001).
808: \bibitem{Henyey}L.G. Henyey, L. Wilets, K.H. B\"{o}hm, R. Lelevier, and R.D.
809: Lev\'{e}e, A method for automatic computation of stellar
810: evolution, Astrophys. J. 129, 628 (1959).
811: \bibitem{hierarchical}M. Sugawara, Adaptive basis set for quantum mechanical calculation based on
812: hierarchical finite element method,
813: Chem. Phys. Lett. 295, 423 (1998).
814: \bibitem{microgenetic}H. Nakanishi and M. Sugawara, Numerical solution of the Schr\"{o}dinger equation by a microgenetic algorithm,
815: Chem. Phys. Lett. 327, 429 (2000).
816: \bibitem{Holland}J. H. Holland, Adaptation in Natural and Artificial Systems
817: (University of Michigan Press, Ann Arbor, 1975, 1992).
818: \bibitem{Amore}Paolo Amore, J. Phys. A: Math. Gen. 39 (2006) L349-L355.
819: \bibitem{Ulrich} U. D. Jentschura, J. Zinn-Justin, J. Phys. A 34 (2001) L253-L258.
820: \bibitem{Gousheh}Siamak S. Gousheh, An Efficient Algorithm for Solving Coupled Schr\"{o}dinger Type
821: ODE's, Whose Potentials Include $\delta$-Functions, J. Comput.
822: Phys. 123, 162 (1996).
823: \bibitem{siamak2} S. S. Gousheh and H. R.
824: Sepangi, K. Ghafoori-Tabrizi, Computer Physics Communications, 149
825: (2003), 135-141.
826: \bibitem{wavepacket}S. S. Gousheh and H. R. Sepangi, Phys.Lett. A272
827: (2000), 304-312.
828: \bibitem{SM}P. Pedram, M. Mirzaei, S.S. Gousheh, Computer Physics
829: Communications (2007), doi: 10.1016/j.cpc.2007.01.004.
830: \bibitem{SP1}J. P. Boyd,  Chebyshev \& Fourier Spectral Methods, Springer-Verlag, BerlinHeidelberg, (1989).
831: \bibitem{SP2}D. Gottleib and S. Ortega, Numerical analysis of spectral methods:
832: theory and applications, SIAM, Philadelphia (1977).
833: \bibitem{SP3}C. Canuto, M. Y. Hussaini, A. Quateroni and T. Zang, Spectral
834: Methods in Fluid Dynamics Springer, Berlin (1988).
835: \bibitem{mostafazadeh}A. Mostafazadeh, Variational Sturmian Approximations: A nonperturbative method of
836: solving time-independent Schr\"{o}dinger equation, J. Math. Phys.
837: 42, 3372-3389 (2001); quant-ph/0105047.
838: \bibitem{exactanhrmonic}B. Bacus, Y. Meurice, and A. Soemadi, Precise determination of the energy levels
839: of the anharmonic oscillator from the quantization of the angle
840: variable, J. Phys. A: Math. Gen. 28, L381 (1995).
841: \bibitem{CSA}F. Antonsen, Sturmian basis functions for the harmonic oscillator, Phys. Rev. A 60, 812 (1999).
842: \bibitem{2D}P. Pedram, M. Mirzaei, S. S. Gousheh. math-ph/0611063.
843: \end{thebibliography}
844: \end{document}
845: % ----------------------------------------------------------------
846: