math0401116/hyp.tex
1: %% Academic Press `C' Style Template File.
2: %% Current version: May 12, 1999
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: % LaTeX2e:
5: \documentclass{cjour}
6: \usepackage{amssymb,psfig,url}
7: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
8: \def\Frac#1#2{\frac{\displaystyle{#1}}{\displaystyle{#2}}}
9: \topmargin 2cm
10: \begin{document}
11: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
12: %%%%% To be entered at Academic Press: =====>>           %
13: % \journame{}                                            %
14: % \articlenumber{}                                       %
15: % \yearofpublication{}                                   %
16: % \volume{}                                              %
17: % \cccline{}                                             %
18: % \received{}                                            %
19: % \revised{}                                             %
20: % \accepted{}                                            %
21:  \authorrunninghead{A. Gil, W. Koepf $\&$ J. Segura}    %
22:  \titlerunninghead{Computation of zeros of hypergeometric functions}    %
23: % communication line, use: \commline{Communicated by...} %
24: % \commline{ }                                           %
25: %\setcounter{page}{261} %% This command is optional.     %
26: %% <<== End of commands to be entered at Academic Press  %
27: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
28: \title{Numerical algorithms for the real zeros of hypergeometric functions}
29: \author{Amparo Gil$^{a}$, Wolfram Koepf$^{b}$ and Javier Segura$^{c}$}
30: \affil{$^a$Depto. de Matem\'aticas, Universidad Aut\'onoma de Madrid,
31: 28049-Madrid, Spain \\
32:  E-mail: amparo.gil@uam.es\\
33: $^b$ Universit\"at Kassel, FB 17 Mathematik-Informatik, 34132-Kassel, Germany\\
34:   E-mail: koepf@mathematik.uni-kassel.de\\
35: $^c$ Depto. de Matem\'aticas, Estad\'{\i}stica y Computaci\'on.
36: Universidad de Cantabria. 39005-Santander,
37: Spain\\
38:   E-mail: javier.segura@unican.es   \\
39: }
40: 
41: \abstract{ Algorithms for the computation of the real zeros of
42: hypergeometric functions which are solutions of second order ODEs
43: are described. The algorithms are based on global fixed point
44: iterations which apply to families of functions satisfying first
45: order linear difference differential equations with continuous
46: coefficients. In order to compute the zeros of arbitrary solutions
47: of the hypergeometric equations, we have at our disposal several
48: different sets of difference differential equations (DDE). We
49: analyze the behavior of these different sets regarding the rate of
50: convergence of the associated fixed point iteration. It is shown
51: how combinations of different sets of DDEs, depending on the range
52: of parameters and the dependent variable, is able to produce
53: efficient methods for the computation of zeros with a fairly
54: uniform convergence rate for each zero. 
55: 
56: }
57: 
58: 
59: 
60: {\it AMS subject classification: } 33Cxx, 65H05  
61: 
62: \keywords{Zeros; hypergeometric functions; fixed point iterations; numerical algorithms}
63: 
64: 
65: 
66: \begin{article}
67: 
68: \section{Introduction}
69: 
70: The zeros of hypergeometric functions are quantities which appear
71: in a vast number of physical and mathematical applications. For
72: example, the zeros of classical orthogonal polynomials (OP) are
73: the nodes of Gaussian quadrature; classical OP (Hermite, Laguerre
74: and Jacobi polynomials) are particular cases of hypergeometric
75: functions. Also, the zeros of Bessel functions and their
76: derivatives appear in many physical applications and there exists
77: a variety of methods of software for computing these zeros.
78: 
79: However, an efficient algorithm which can be applied to the computation
80: of all the zeros of any hypergeometric function in any real interval (not
81: containing a singular point of the defining ODE) is still missing.
82: 
83: In \cite{Seg2, Seg3} methods were introduced which are
84: capable of performing this task for hypergeometric functions which are solutions
85: of a second order ODE; an explicit Maple algorithm was presented in \cite{Gil}.
86: The starting point of the methods is the construction of a first order
87: system of differential equations
88: \begin{equation}
89: \begin{array}{ll}
90: y' (x)=\alpha (x) y(x) +\delta (x) w (x)\\
91: w' (x)=\beta (x) w(x) +\gamma (x) y (x),
92: \end{array}
93: \end{equation}
94: \noindent with continuous coefficients $\alpha (x)$, $\beta (x)$,
95: $\gamma (x)$ and $\delta (x)$ in the interval of interest,
96: relating our problem function $y(x)$ with a contrast function
97: $w(x)$, whose zeros are interlaced with those of $y(x)$.
98: Typically, the contrast function $w(x)$ satisfies a second order
99: ODE similar to the second order ODE satisfied by the problem
100: function.
101: 
102: Given a hypergeometric function $y(x)$ there are several known options to choose as contrast
103: functions $w(x)$. As an example, considering a Jacobi polynomial
104: \begin{equation}
105: y(x)=P_n^{(\alpha ,\beta)}(x)=\Frac{(\alpha+1)_{n}}{n!}\,_2\mbox{F}_{1} (-n,n+\alpha+\beta+1;
106: \alpha+1;(1-x)/2)
107: \end{equation}
108: \noindent we could take as contrast function $w_{\mbox{\tiny{OP}}} (x)=P_{n-1}^{(\alpha ,\beta)}(x)$
109: but also $w_{\mbox{\tiny{D}}}(x)=\Frac{d}{dx}P^{(\alpha ,\beta)}_n(x)$ is a possible choice. Both contrast
110: functions are $_2\mbox{F}_1 (a,b;c;x)$ hypergeometric functions with parameters $a,b,c$ differing
111:  by integer numbers from the parameters of the problem function $y(x)$.
112: 
113: When the contrast function in the previous example is $w_{\mbox{\tiny{OP}}}
114: (x)=P_{n-1}^{(\alpha ,\beta)}(x)$
115: the first order differential system is related to the three term recurrence relation for Jacobi
116: polynomials. It may seem that this is a natural differential system to consider. However, it was
117: numerically observed that the fixed point method which can be obtained from this differential
118: system becomes relatively slow for the zeros of $P_n^{(\alpha,\beta)}$ close to $\pm 1$ \cite{Gil}. Because
119: the extreme zeros approach to $\pm 1$ as $n\rightarrow +\infty$, the efficiency for the computation
120: of such zeros decreases as the order increases. Similar problems arise, for example, when
121: $\alpha \rightarrow -1^+$ or $\beta \rightarrow -1^+$. In fact, the number of iterations required
122:  to compute the extreme zeros tend to infinity in these limits. Similar problems take place for
123:  Laguerre polynomials $L_{n}^{\alpha }(x)$ for the smallest (positive) zero. Fortunately we will
124:  later show how the selection of $w_{\mbox{\tiny{D}}}(x)$ as contrast function
125: gives a much better asymptotic behavior for the resulting fixed
126: point iteration for the extreme Jacobi
127:  zeros. For the Laguerre case, a similar solution is possible.
128: 
129: These two examples illustrate the need to analyze the convergence
130: of the resulting fixed point iteration for the different available
131: contrast functions. Although for any adequate contrast function
132: (satisfying the necessary conditions \cite{Seg2,Seg3}) the
133: resulting fixed point method is quadratically convergent, the
134: non-local behavior of the method and the corresponding estimation
135: of first guess values for the zeros may result in disaster for
136: certain contrast
137:  functions in some limits. As a result of this study, we will
138: obtain explicit methods for the computation of the real zeros of
139: hypergeometric functions with a good asymptotic behavior and a
140: fairly uniform convergence rate in the whole range of parameters.
141: 
142: \section{Theoretical background}
143: \label{repaso}
144: 
145: Let us now briefly outline the main ingredients of the numerical method. For more details
146: we refer to \cite{Seg3,Seg2}.
147: It was shown in \cite{Seg3,Seg2} that, given a family of functions $\{y_k^{(1)}, y_k^{(2)}\}$,
148: depending on one parameter $k$, which are independent solutions of second order ODEs
149: \begin{equation}
150: y''_k+B_k(x)y'_k+A_k(x)y_k=0\,,k=n,n-1
151: \label{ODE}
152: \end{equation}
153: and satisfy relations of the type:
154: \begin{equation}
155: \begin{array}{rl}
156: y'_n=& a_n(x)y_n+d_n(x)y_{n-1}\\
157: y'_{n-1}= &b_n(x)y_{n-1}+e_n(x)y_n
158: \label{ddr}
159: \end{array}
160: \end{equation}
161: the coefficients $a_n(x)$, $b_n(x)$, $d_n(x)$, $e_n(x)$, $B_k (x)$
162: and $A_k(x)$ being continuous and $d_n e_n<0$ in a given interval
163: $[x_1, x_2]$, fixed point methods (Eq.\ (\ref{iter})) can be built
164: to compute all the zeros of the solutions of (\ref{ODE}) inside
165: this interval. These difference-differential equations (\ref{ddr})
166: are called general because they are satisfied by a basis of
167: solutions $\{y_k^{(1)}, y_k^{(2)}\}$. The fact that the DDEs are
168: general and with continuous coefficients in an interval $I$
169: implies \cite{Seg2} that $d_n e_n \neq 0$ in this interval.
170: Conversely, given $\{y_n^{(i)}, y_{n-1}^{(i)}\}$, $i=1,2$
171: independent solutions of the system (\ref{ddr}) and $d_n e_n\neq
172: 0$, then $\{y_k^{(1)}, y_k^{(2)}\}$, $k=n, n-1$, are independent
173: solutions of the ODEs (\ref{ODE}). The method can then be applied
174: to compute the zeros of any solution of such ODEs.
175: 
176: It was shown that the ratios $H_i (z)$ ($i=\pm 1$):
177: \begin{equation}
178: \begin{array}{l}
179: H_i(z)=-i\,{\rm sign\:}(d_{n_i})K_{n_i}\Frac{y_n(x(z))}{y_{n_{i}}(x(z))}\\
180: K_{n_i}=\left(-\Frac{d_{n_i}}{e_{n_i}}\right)^{i/2}\,,\,
181: z(x)=\int \sqrt{-d_{n_i} e_{n_i}} dx
182: \end{array}
183: \label{FixP}
184: \end{equation}
185: where $n_{+1}=n+1\,,\,n_{-1}=n$, satisfy the first order equations
186: \begin{equation}
187: \dot{H}_i (z)=1+H_i (z)^2-2\eta_i (x(z)) H_i (z)
188: \label{ratH}
189: \end{equation}
190: \noindent where
191: $$
192: \eta_i (x)=i\Frac{1}{\sqrt{-d_{n_i}e_{n_i}}}\left(a_{n_i}-b_{n_i}+\frac12
193: \left(\Frac{e'_{n_i}}{e_{n_i}}-\Frac{d'_{n_i}}{d_{n_i}}\right)\right)
194: $$
195: and the dot means derivative with respect to $z$ while the prime
196: is the derivative with respect to $x$. \noindent Using Eq.\
197: (\ref{ratH}), one can show that
198: \begin{equation}
199:    T_i(z)=z - \arctan (H_i (z))
200: \label{iter}
201: \end{equation}
202: are globally convergent fixed point iterations (FPI): given a
203: value $z_0$ between two con\-se\-cu\-tive zeros
204: ($z_{n_i}^{1},z_{n_i}^{2}$) of $y_{n_i}(x(z))$ (consecutive
205: singularities of $H_i (z)$),
206:  the iteration of (\ref{iter}) converges to $z_n\in (z_{n_i}^{1},z_{n_i}^{2})$, where $x_n=x(z_n)$
207: is a zero of $y_n (x)$.
208: 
209: Global bounds for the distance were provided which lead to iteration steps
210: that can be used to compute new starting values for obtaining all the zeros inside a given
211: interval.
212: 
213: It was shown that, in intervals where $\eta_i$ does not change sign,
214:  either
215: \begin{equation}
216: |z(x_{n_i}^{1})-z_n|<\pi /2 \mbox{ and  } |z(x_{n_i}^{2})-z_n|>\pi /2\,\,\,(\eta_i>0)
217: \label{etamas}
218: \end{equation}
219: \noindent
220: or
221: \begin{equation}
222: |z(x_{n_i}^{1})-z_n|>\pi /2 \mbox{ and  } |z(x_{n_i}^{2})-z_n|<\pi /2\,\,\, (\eta_i<0).
223: \label{etamenos}
224: \end{equation}
225: \noindent In this
226: way $\pi /2$ is the choice for the iteration step when the second situation takes place ($\eta_i<0$); this means
227: that if $y_{n_i}(x(z))$ has at least a zero larger than $z_n$, then
228: \begin{equation}
229: \lim_{j\rightarrow\infty}T^{(j)}(z_n+\pi/2 )
230: \label{nonimp}
231: \end{equation}
232: is the smallest zero larger than $z_n$. Similarly, when $\eta_i>0$, the iteration step will be $-\pi/2$ instead
233: of $\pi/2$ (backward sweep). When $\eta_i$ changes sign, forward and backward schemes can be combined \cite{Seg3}.
234: 
235: The functions $H_i (z)$ can be written as a ratio of functions $H_i (z)=\tilde{y}_{n}(z)/\tilde{y}_{n_i}(z)$, where
236: $\tilde{y}_k(z)=\lambda_k (z) y_k (x(z))$ and $\lambda_k$ have no zeros,
237: in such a way that the functions $\tilde{y}_n (z)$ and
238: $\tilde{y}_{n_i}(z)$ (with the same
239: zeros as $y_n (x(z))$ and $y_{n_i} (x(z))$ respectively) satisfy
240:  second order ODEs in normal form:
241: \begin{equation}
242: \begin{array}{ll}
243: \Frac{d^2 \tilde{y}_n}{dz^2}+\tilde{\mbox{A}}_n \tilde{y}_n=0\,, &
244: \Frac{d^2 \tilde{y}_{n_i}}{dz^2}+\tilde{\mbox{A}}_{n_i} \tilde{y}_{n_i}=0\,,  \\
245: & \\
246: \tilde{\mbox{A}}_n (z)=1+\dot{\eta}_i -\eta_i^2\,, & \tilde{\mbox{A}}_{n_i} (z)=1-\dot{\eta}_i -\eta_i^2\,.
247: \end{array}
248: \label{ODEsnor}
249: \end{equation}
250: 
251: Finally, we recall that using monotony conditions of $A_n (z)$ the
252: iteration steps $\pm \pi /2$ (Eq.\ (\ref{nonimp})) can be improved
253: according to Theorem 2.4 of \cite{Seg3}:
254: 
255: \begin{theorem}
256: If $z_{-1}<z_0<z_1$ are three consecutive zeros of $y_n (x(z))$ and $\eta_i (z)\dot{\tilde{\mbox{A}}}_n (z)>0$ in
257: $(z_{-1}, z_{1})$ then $z_{j}=\lim_{n\rightarrow\infty}T^{(n)}(z_0+\Delta z_0)$ where $\Delta z_0=z_0-z_{-j}$,
258: $j=\mbox{sign}(\eta)$. The convergence is monotonic.
259: \label{T2.4}
260: \end{theorem}
261: 
262: \subsection{Oscillatory conditions}
263: 
264: We are interested in computing zeros of oscillatory solutions of second order ODEs and, in particular, on building
265: algorithms for the computation of the zeros of the hypergeometric functions.
266: If a second order differential equation has a given number of singular regular points,
267:  we divide the real axis in subintervals determined by the singularities and search for
268: the zeros in each of these subintervals. We only apply the algorithms if it is not disregarded
269:  that the function can have two zeros at least
270: in the subinterval under consideration.
271: 
272:  We consider that an ODE has
273: oscillatory solutions in one of these subintervals if it has
274: solutions with at least two zeros in this subinterval; otherwise,
275: if all the solutions have one zero at most we will call these
276: zeros isolated zeros. The fixed point methods (FPMs) before
277: described deal with the zeros of any function satisfying a given
278: differential equation, no matter what the initial conditions are
279: on this function. Isolated zeros for a given solution depend on
280: initial conditions or boundary conditions for this solution and
281: are, in any case, easy to locate and compute.
282: 
283: There are several ways to ensure that a solution $y_n(x)$ has at most one zero in an interval; among them:
284: 
285: \begin{theorem} If one of the following conditions is satisfied in an interval $I$ (where all the coefficients
286: of the DDEs are continuous)
287: then $y_n$ and ($y_{n_i}$) have at most one
288: zero in the interval $I$ (trivial solutions excluded):
289: \begin{enumerate}
290: \item{$d_{n_i}(x) e_{n_i}(x)\ge 0$ in} $I$ \cite{Seg2}.
291: \item{$|\eta_i (x)|\ge 1$} in $I$ \cite{Seg3}.
292: \item{$\tilde{\mbox{A}}_n<0$} ($\tilde{\mbox{A}}_{n_i}<0$) \cite{Seg3}.
293: \end{enumerate}
294: \label{osccond}
295: \end{theorem}
296: 
297: The condition $e_{n_i}d_{n_i}<0$ is required for the method to apply. Furthermore, it is known that
298: when the DDEs (\ref{ddr}) are general, $d_{n_i} e_{n_i}$ can not change sign. Therefore $d_{n_i} e_{n_i}<0$
299: is a clear signature for the oscillatory character of the differential equation.
300: 
301: 
302: \subsection{Hypergeometric functions; selection of the optimal DDEs}
303: 
304: For hypergeometric functions several DDEs are available
305: for the construction of fixed point iterations (FPIs), depending on the
306: selection of contrast function.
307: 
308: Let us start by considering, for example, the case of the confluent
309: hypergeometric
310:  equation
311: \begin{equation}
312: xy^{\prime\prime}+(c-x)y^{\prime}-a y=0
313: \label{11}
314: \end{equation}
315: 
316: One of the solutions of this differential equation
317: are Kummer's series
318: $$M(a,c,x)\equiv  _1\!\!\mbox{F}_1(a;c;x)=\sum_{k=0}^{\infty}\Frac{(a)_k}{(c)_k k!}x ^k\,,
319: $$
320: for which different difference-differential relations are
321: available. Indeed, denoting $\alpha_n=\alpha+k\,n$,
322: $\gamma_n=\gamma+m\, n$ and $y_n\equiv M(\alpha_n ,\gamma_n ,x)$
323: we will have different sets of DDES (Eq.\ (\ref{ddr})) for
324: different selections of $(k,m)$.
325: 
326: For Gauss hypergeometric functions $_2\mbox{F}_1 (a,b;c;x)$, which are solutions
327: of the ODE
328: \begin{equation}
329: x(1-x)y''+[c-(a+b+1)x]y'-aby=0\,,
330: \label{GHF}
331: \end{equation}
332: \noindent the possible DDEs are determined by three-vectors with
333: integer components, that is, we will consider $y_n\equiv \:
334: _2\mbox{F}_1 (\alpha+k\,n,\beta+l\,n; \gamma+m\,n;x)$ and the
335: associated DDEs will be named $(k,l,m)$-DDEs. Finally, for the
336: case of the hypergeometric functions $_0\mbox{F}_1(;c;x)$, we can
337: only consider families $y_n= \:_0\mbox{F}_1 (;\gamma+k\,n ;x)$ and
338: the different relations are described by the integer numbers $k$.
339: 
340: Our FPMs can only be applied to solutions of second order ODEs.
341: This restricts our study to the hypergeometric functions
342: $_0\mbox{F}_1(;c;x)$,
343:  $_2\mbox{F}_0(a,b;;x)$,
344:  $_1\mbox{F}_1 (a;c;x)$ and
345: $_2 F_1 (a,b;c;x)$.
346: 
347: Regarding the selection of the different DDEs available, we will restrict ourselves to:
348: \begin{enumerate}
349: \item{DDEs} with continuous coefficients except at the singular points of the defining differential
350: equations.
351: \item{The} most simple DDEs in a given recurrence direction which allow the use of improved iteration steps.
352: Taking as example the case of confluent hypergeometric
353: functions, this means that the $(1,0)$-DDE will be described, and the
354: analysis of the $(-1,0)$, $(2,0)$,... DDEs will be skipped.
355: \end{enumerate}
356: 
357: The first restriction is convenient for simplicity and it means that the problem function and the contrast function
358: have zeros interlaced in each subdivision of the real interval defined by the singular points of
359: the differential equation; this is a convenient property for a simple application
360: of the FPMs and enables the application of each DDE to compute all the zeros in the different subintervals
361: of continuity of the solutions of the differential equation.
362: 
363: Regarding the second restriction, and considering the case of
364: confluent hypergeometric functions as example, it should be noted
365: that for the $(k,m)$-DDE,
366: generally two FPIs are available, one of them based on the ratio
367: $H_{-1}=y_n(x)/ y_{n-1}(x)$ and a second one based on
368: $H_{+1}(x)=y_n (x)/y_{n+1}(x)$. As described in \cite{Seg3},
369: generally one of these two iterations is preferable because
370: improved iteration steps can be considered according to Theorem
371: \ref{T2.4} (Theorem 2.4 in \cite{Seg3}). If we considered the
372: $(-k,-m)$-DDE, the two associated ratios $H_i$, $i=\pm 1$, would
373: be the same as before (replacing $i$ by $-i$). Because both
374: selections of DDEs are equivalent, only one of them will be
375: discussed.
376: By convention, we will consider pairs $(k,m)$ for which
377: the iteration  on $H_{-1}$ can take advantage of the monotony
378: property of $\hat{\mbox{A}}_n (x(z))$, as described in
379: \cite{Seg3}, for classical orthogonal polynomial cases (Jacobi, Hermite,
380: Laguerre).
381:  Once we have fixed this criterion the index $i$ in
382: Equations (\ref{ratH})-(\ref{ODEsnor}) can be dropped.
383: We consider the following additional notation: given a vector
384: $\vec{u}$ with integer components, we will denote by
385: DDE$(\vec{u})$ (FP$(\vec{u})$) the corresponding DDE (FPI) based
386: on the ratio $y_n/y_{n-1}$ for $x>0$.
387: 
388: On the other hand, and considering the confluent case as illustration, we will not analyze DDE$(2,2)$ nor any
389: successive multiples of the DDE$(1,1)$. This is so because the first restriction is generally
390: violated if successive multiples of a DDE are considered (there are exceptions to this; see the case of $_0\mbox{F}_1$
391: hypergeometric functions).
392: 
393: 
394: \subsubsection{Selection of the optimal DDEs}
395: 
396: There are several (and related) criteria to select among the available DDEs to compute the zeros
397: of a given function $y$. The associated FPMs tend to be more efficient as we are
398: closer to any of the following two situations:
399: 
400: \begin{enumerate}
401: \item{$\eta (x)=0$,}
402: \item{the} coefficient $\tilde{\mbox{A}}_n$ is constant.
403: \end{enumerate}
404: 
405: Of course, the first condition implies the second one (see Eq.\
406: (\ref{ODEsnor})). The first condition makes the FPI converge with
407: one iteration for any starting value. The second condition makes
408: the method an exact one using improved iteration steps (there is
409: even no need to iterate the FPM).
410: 
411: Let us recall that the FPIs associated to a given system of DDEs
412: are quadratically convergent to a zero $z_0$ (in the transformed
413: variable $z$) with asymptotic error constant $\eta(x(z_0))$.
414: Therefore, the smaller $|\eta (x)|$ is the fastest the convergence
415: is expected to be, at least for starting values close enough to
416: $z_0$.  On the other hand, the smaller the absolute value of
417: variation of $\tilde{\mbox{A}}_n$ is, the better the improved
418: iteration (Theorem \ref{T2.4}) will work because this implies the
419: exactness of the iteration criteria to estimate starting values
420: from previously computed zeros. This second criterion (on the
421: variation of $\tilde{\mbox{A}}$) is more difficult to apply, as we
422: will later see. It is, however, more relevant to improve the
423: iterative steps for obtaining starting values to compute zeros
424: than to improve the local convergence properties of the fixed
425: point methods, which are quadratically convergent anyway.
426: 
427: Indeed, as was described in \cite{Gil}, the natural FPIs for
428: orthogonal polynomials of confluent hypergeometric type
429: (FP$(-1,0)$) tend to converge slowly for the computation of the
430: first positive zeros when they become very small. This, for
431: instance, is the case  for Laguerre polynomials $L_{n}^{\alpha}
432: (x)$ when $\alpha\rightarrow -1^+$. The reason for this behavior
433: lies in the fact that the associated change of variables is
434: singular at $x=0$:
435: $$
436: z=\displaystyle\sqrt{(b-a)(1-a)}\ln x\,,
437: $$
438: In this way, the interval of orthogonality
439:  for the Laguerre polynomials $(0,+\infty)$
440: is transformed into $(-\infty,\infty)$ in the $z$ variable. This
441: means that the zeros which are very small in the $x$ variable,
442: tend to go to $-\infty$ in the $z$ variable. Therefore after
443: computing the second smallest zero, $x_2$, the next initial guess
444: for the FPI, $z(x_2)-\pi /2$, may lie well far apart for the value
445: $z(x_1)$, $x_1$ being the smallest zero.
446:  Although it is guaranteed
447: that the FPI will converge to $z(x_1)$, it could take a considerable
448: number of iterations to approach this value.
449: 
450: Fortunately, we will see that the rest of FPMs (different to
451: FP$(k,0)$) do not show such a singularity; therefore, we expect
452: better behavior near $x=0$ for these iterations.
453: 
454: This suggests that, given two FPIs with associated change of
455: variables $z_1(x)$, $z_2 (x)$ respectively, one should choose that
456: one which gives the largest displacement in the $x$ variable for
457: the same step in the corresponding $z$ (the typical value being
458: $\pi /2$). Let us stress that the possibility of passing the next
459: zero is ruled out: in the algorithms the sequence of all $z$
460: values calculated in a backward (forward) sweep form monotonically
461: decreasing (increasing) sequences.
462: 
463: We will therefore say that the change of variables $z_1$ behaves better than $z_2$
464: if $x(z_1+\Delta z)>x(z_2+\Delta z)$, for a typical value of $\Delta z$ ($\approx \pi /2$). Given the
465: definition of the changes of variables $z(x)$:
466: $$
467: x(z+\Delta z)-x (z)=\int_{z}^{z+\Delta z}\Frac{1}{\tilde{d}_{n}(x(z))}dz
468: $$
469: \noindent where $\tilde{d}_n=\sqrt{-d_n e_n}$, we can say that the
470: change of variable $z_1 (x)$ (and its associated FPI) is more
471: appropriate than the change $z_2 (x)$ when its coefficient
472: $\tilde{d}_{n}=\sqrt{-d_n e_n}$ is smaller than the corresponding
473: coefficient for $z_2(x)$.
474: 
475: Therefore, an alternative non-local prescription to that one
476: dealing with $\tilde{\mbox{A}}_n$ is the following: among the
477: possible DDEs and associated fixed point iterations, choose that
478: one for which $|d_n e_n|$ is smallest. As we will see, this is an
479: easy to apply criterion which correctly predicts the more
480: appropriate DDEs and FPI depending on the range of the parameters
481: and the dependent variable.
482: 
483: \section{Analysis of Hypergeometric functions}
484: 
485: We will use the DDEs satisfied by hypergeometric series as generated by
486: the Maple package {\it hsum.mpl} \cite{Koe}.
487: The results for each of the family of functions (change of
488: variable $z(x)$, function $\eta(x)$, etc.) considered can be
489: automatically generated using the package {\it zeros.mpl}
490: \cite{Gil}.
491: 
492: \subsection{Hypergeometric function $_0\mbox{F}_1 (;c;x)$}
493: 
494: The ODE satisfied by the function $y(x)= \:_0\mbox{F}_1(;c;x)$ is
495: \begin{equation}
496: x^2 y'' +c x y' -xy=0
497: \label{01}
498: \end{equation}
499: The solutions of these differential equations have an infinite
500: number of zeros for negative $x$ and are related to Bessel
501: functions:
502: $$
503: _0\mbox{F}_1 (;c;z)=\Gamma (c) (-z)^{(1-c)/2}J_{c-1}(2\sqrt{-z})
504: $$
505: 
506: \subsubsection{First DDE}
507: 
508: Let us consider the DDEs for the family of functions $y_n = \:
509: _0\mbox{F}_1(;\gamma +n;-x)$
510: 
511: The DDEs for this family
512: reads:
513: \begin{equation}
514: \begin{array}{l}
515: y'_{n}=-\Frac{c-1}{x}y_{n-1}+\Frac{c-1}{x}y_n\\[2mm]
516: y'_{n-1}=-\Frac{1}{c-1}y_{n} \label{DDE01}
517: \end{array}
518: \end{equation}
519: \noindent
520: where $c=\gamma+n$.
521: 
522: The relation with Bessel functions can be expressed saying that, if $y (x)$ is a solution of
523:  $x^2 y'' +(\nu+1) x y' +xy=0$ then
524: \begin{equation}
525: w(x)=x^{\nu} y(x^2 /4)
526: \label{relac}
527: \end{equation}
528: \noindent is a solution of the Bessel equation $x^2 w'' +x w' +(x^2-\nu^2 )w=0$.
529: 
530: Transforming the DDEs (\ref{DDE01}) as described in \cite{Seg3},
531: and in Section \ref{repaso}, the relevant functions are:
532: \begin{equation}
533: \eta (z(x)) =\Frac{\nu -1/2}{2\sqrt{x}},\\
534: \tilde{\mbox{A}}_n=1-\Frac{\nu^2 -1/4}{4 x},\\
535: z(x)=2\sqrt{x} \;
536: \end{equation}
537: 
538: The fixed point method deriving from this set of DDEs will have
539: identical performance as the system considered in \cite{Seg3},
540: which holds for Ricatti-Bessel functions
541: $j_{\nu}(x)=\sqrt{x}J_{\nu} (x)$, that are solutions of the
542:  second
543: order ODEs $y''+A(x)y=0$, with $A(x)=1-(\nu^2-1/4)/x^2$. The
544: identification of both methods with the replacement $x\rightarrow
545: x^2/4$ (according to Eq.\ (\ref{relac}) and to the change of
546: variable $z(x)$) is evident by comparing the $A(x)$
547:  and $\tilde{\mbox{A}}_n (x)$ coefficients. This is not surprising given that both methods compare the same problem
548: function, $J_{\nu}(x)$, with the same contrast function, $J_{\nu -1}(x)$, up to factors which do not vanish
549: (for example, the factor $\sqrt{x}$ for Ricatti-Bessel functions) and up to changes of variable.
550: 
551: \subsubsection{Second DDE}
552: 
553: For this type of hypergeometric functions, the only alternative
554: DDEs that can be built are those based on the family of functions
555: $y_n=\:_0\mbox{F}_1(;\gamma+m\,n;-x)$; $m=1$ corresponds to the
556: DDEs (\ref{DDE01}). We will only consider the case $m=2$
557: (equivalent, in the sense described before, to $m=-2$). For
558: $|m|>2$, the DDEs violate the first imposed condition on the
559: continuity of the coefficients. For the functions
560: $y_n=\:_0\mbox{F}_1 (;\gamma+2n;-x)$, the associated DDEs read:
561: \begin{equation}
562: \begin{array}{l}
563: y'_{n}=-\Frac{(c-2)^2+(c-2)-x}{(c-2) x}y_{n}+\Frac{c-1}{x}y_{n-1}\\
564: \\
565: y'_{n-1}=-\Frac{1}{c-2}y_{n-1}-\Frac{x}{(c-1)(c-2)^2}y_{n}
566: \label{DDE012}
567: \end{array}
568: \end{equation}
569: \noindent
570: where $c=\gamma+2n$.
571:  The relevant functions in this case are (again, writing $\nu=c-1$):
572: \begin{equation}
573: \begin{array}{l}
574: \eta(z(x))=\Frac{(\nu-1)^2}{2x}-1\\
575: \\
576: \tilde{\mbox{A}}_n(z(x))=\left(\Frac{\nu-1}{2x}\right)^2 (4x-(\nu^2-1))\\
577: \\
578: z(x)=\Frac{x}{\nu -1}
579: \end{array}
580: \end{equation}
581: 
582: \subsubsection{Comparison between DDEs}
583: 
584: The second DDE is no longer equivalent to the first one; in fact,
585: it has quite different characteristics. Let us compare the
586: expected performance of these two DDEs, according to the different
587: criteria described above.
588: 
589: To begin with, the $\eta(x)$ parameter never vanishes for the
590: first DDE (DDE1), except when $\nu=1/2$, in which case
591:  the method
592: with improved steps is exact without the need to iterate the FPI
593: even once (forward or backward sweeps \cite{Seg2} are used
594: depending on the sign of $\eta (x)$). In contrast, DDE2 has an
595: $\eta$-function which changes sign at $x_{\eta}=(\nu -1)^2/2$ and
596: the zeros are computed by an expansive sweep \cite{Seg2}. Close to
597: $x_{\eta}$ we can expect that DDE2 tends to behave better in
598: relation to local convergence, because the asymptotic error
599: constant tends to be small.
600: 
601: We observe that, as $x\rightarrow +\infty$,  $\eta(x)$ goes to zero for DDE1 but it tends to $-1$ for DDE2. This suggests
602: that DDE1 will have faster local convergence than DDE2 for large $x$. On the other hand, as $\nu$ increases, $\eta(x)$
603: becomes larger, however, it is difficult to quantify the impact on local converge because as $\nu$ increases also
604: the smallest zero becomes larger. Let us also take into account that DDE2 will have small $\eta (x)$ for
605: $x$ close to $x_{\eta}$, which becomes large for large $\nu$.
606: 
607: Regarding the behavior of $\tilde{\mbox{A}}_n (z)$, it is
608: monotonic for DDE1 and has a maximum for DDE2, which allows the
609: use of improved iteration steps. For DDE1, $\tilde{\mbox{A}}_n
610: (z)$  is constant when $\nu=1/2$, which means that sweep with
611: improved iteration steps is exact, as commented  before. The
612: maximum for DDE2 is at $x_m=(\nu^2-1)/2$, where
613: $\tilde{\mbox{A}}_n(x_m)=(\nu-1)/(\nu+1)$. Around this extremum,
614: the improved iteration steps tend to work better because
615: $\tilde{\mbox{A}}_n (x)$ will be approximately constant; how
616: constant $\tilde{\mbox{A}}_n (x)$ is around $x_m$ can be measured
617: by the convexity at this point. We find:
618: $$
619: \ddot{\tilde{\mbox{A}}}_n (z_m)=-8\Frac{\nu -1}{(\nu +1)^3}
620: $$
621: \noindent where $z_m=z(x_m)$. As $\nu$ becomes larger, $\ddot{\tilde{\mbox{A}}}_n (z)$  
622: becomes smaller around the maximum of $\tilde{\mbox{A}}_n (z)$ and 
623: the improved iteration will work better. This fact again, favors DDE2 for large $\nu$.
624: 
625: Finally, considering the criterion of smaller $D_n=|d_n e_n|$, we find that, for DDE1
626: \begin{equation}
627: D_n=\Frac{1}{x}
628: \end{equation}
629: \noindent while for DDE2
630: \begin{equation}
631: D_n=\Frac{1}{(\nu -1)^2}
632: \end{equation}
633: 
634: This again shows that DDE1 will improve as $x$ increases while DDE2 will be better for large $\nu$. Numerical
635: experiments show that for $\nu>100$ the second DDE is preferable over the first, particularly for computing
636: the smallest zeros.
637: 
638: The different criteria yield basically the same information. However the prescription on $D_n=|d_n e_n|$ is the simplest
639: one to apply. From now on, we will not repeat the analysis for the different criteria. Instead, we adopt this last
640: criterion to analyze the rest of cases.
641: 
642: \vspace*{0.5cm}
643: \begin{minipage}{6cm}
644: \centerline{\protect\hbox{\psfig{file=f01iter10.eps,width=5.cm}}}
645: \end{minipage}
646: \hfill
647: \begin{minipage}{6cm}
648: \centerline{\protect\hbox{\psfig{file=f0110m1.eps,width=5.cm}}}
649: \end{minipage}
650: { {\bf Figure 1.} {\bf Left}: Ratio between the number of
651: iterations needed for the second and first DDEs for the
652: computation of the zeros of $_0\mbox{F}_1(;11,-x)$ (the zeros of
653: $J_{10}(2\sqrt{x})$). {\bf Right}: Number of iterations needed for
654: the first DDE. }
655: 
656: \begin{minipage}{6cm}
657: \centerline{\protect\hbox{\psfig{file=f01iter200.eps,width=5.cm}}}
658: \end{minipage}
659: \hfill
660: \begin{minipage}{6cm}
661: \centerline{\protect\hbox{\psfig{file=f01200m1.eps,width=5.cm}}}
662: \end{minipage}
663: { {\bf Figure 2.}
664: {\bf Left}: Ratio between the number of iterations
665: needed for the second and first DDEs for the computation of the zeros
666: of $_0\mbox{F}_1(;201,-x)$ (the zeros of $J_{200}(2\sqrt{x})$).
667: {\bf Right}: Number of iterations needed for the first DDE.
668: }
669: 
670: \subsection{Confluent hypergeometric function}
671: 
672: For the confluent hypergeometric case we have a larger variety of
673: DDEs to choose, because we can choose families of functions
674: $y_n=\:_1\mbox{F}_1(a+k\, n; c+m\, n;x)$, or, more generally,
675: $y_n=\phi (a+k\, n; c+m\, n;x)$, being $\phi$ any solution of Eq.\
676: (\ref{11}). The families which give rise to DDEs satisfying all
677: our requirements are three, corresponding to the following
678: selections of $(k,m)$: $(1,0)$, $(1,1)$ and $(0,-1)$.
679: 
680: We will give the corresponding DDEs and the associated functions. At the same time, we will
681: restrict the range of parameters for which the functions are oscillatory (considering the
682: first oscillatory condition in Theorem \ref{osccond}).
683: 
684: We can restrict the study to $x>0$ because, if $\phi (a;c;x)$ is a
685: solution of Eq.\ (\ref{11}), then $e^{x}\phi (c-a;c;-x)$ is also a
686: solution of Eq.\ (\ref{11}).
687: 
688:  \subsubsection{$(k,m)=(1,0) \rightarrow y_n=\phi$($\alpha +n$; $\gamma$ ; x)}
689: 
690: Let us write, for shortness and in order to
691: compare with other recurrences $a=\alpha+n$, $c=\gamma$. As before commented,
692: we consider simultaneously the equivalent directions $(k,m)=(1,0)$ and $(k,m)=(-1,0)$ but
693: we present only the DDEs and related functions for the recurrence direction for
694: which the FPI based on the ratio $H_{-1}=y_n /y_{n-1}$ can be used with improved
695: iteration steps (Theorem \ref{T2.4}); this is the direction $(1,0)$.
696: 
697: The $(-1,0)$ direction is the natural one for orthogonal polynomials of hypergeometric type
698: (Laguerre, Hermite), which
699: are related to confluent hypergeometric series of the
700: type $_1\mbox{F}_1(-n;\gamma ;x)$ (see, for instance, \cite{Leb}, Eqs. (9.13.8-10)).
701: 
702: The DDEs for $(k,m)=(1,0)$ read
703: \begin{equation}
704: \begin{array}{ll}
705: y'_n=&\Frac{a-c+x}{x}y_n-\Frac{a-c}{x}  y_{n-1}\\
706: \\
707: y'_{n-1}= &-\Frac{a-1}{x}y_{n-1}+\Frac{a-1}{x}y_n
708: \label{10}
709: \end{array}
710: \end{equation}
711: \noindent and therefore, applying the first oscillatory condition of Theorem \ref{osccond},
712:  the parameters are restricted to:
713: $$
714: (a-c)(a-1) > 0
715: $$
716: \noindent otherwise the functions $y_{n}$, $y_{n-1}$ will be non-oscillatory. If we repeat the same
717: argument for $(k,m)=(-1,0)$ or, equivalently, apply the same criteria for the DDEs for the functions
718: $y_n$, $y_{n+1}$ the following restriction is obtained:
719: $$
720: (a-c+1)a >0\,.
721: $$
722: 
723: The associated functions for this DDE are
724: \begin{equation}
725: \begin{array}{l}
726: \eta (z(x))=-\Frac{2a+c+x}{2\displaystyle\sqrt{(c-a)(1-a)}}\\
727: \\
728: \tilde{\mbox{A}}_n(z(x))= \Frac{-x^2+2(c-2a)x-(c-1)^2}{4(c-a)(1-a)}\\
729: \\
730: z(x)=\displaystyle\sqrt{(c-a)(1-a)}\ln x
731: \end{array}
732: \end{equation}
733: 
734: Let us notice that $\eta (x)$ becomes negative for large $x$,
735: which gives the most appropriate sweep for large $x$ (forward)
736: since $\tilde{\mbox{A}}_n (z(x))$ decreases for large $x$.
737: 
738: 
739: Observe that the lack of a singularity in  $\tilde{\mbox{A}}_n (z(x))$ is only apparent because the function $z(x)$ is singular
740: at $x=0$.
741: 
742: 
743: \subsubsection{$(k,m)=(0,-1)\rightarrow y_n=\phi$($\alpha$; $\gamma -n$ ; x)}
744: 
745: As before, we denote $a=\alpha$, $c=\gamma -n$.  The DDEs for the system read:
746: \begin{equation}
747: \begin{array}{ll}
748: y'_n=& y_n  + \Frac{a-c}{c} y_{n-1}\\
749: \\
750: y'_{n-1}= & -\Frac{c}{x}  y_{n-1}  +\Frac{c}{x}y_n
751: \label{0-1}
752: \end{array}
753: \end{equation}
754: \noindent which implies the restriction $(c-a)x>0$ for the
755: solutions to have oscillatory nature. Repeating the same for
756: $(k,m)=(0,1)$, we arrive at the condition $(c-a-1)x>0$, which for
757: $x>0$ gives a more restrictive condition $c-a>1$ (for $x<0$, the
758: first condition is more restrictive and gives $a-c<0$).
759: 
760: Let us recall that the condition $c-a>1$ for $x>0$ means that, if this condition is not met,
761: neither $y_n$ nor $y_{n-1}$ can have two zeros in $x>0$. Given that we are interested in computing
762: zeros of oscillatory functions, we consider these restrictions. Let us however notice that the
763: possible isolated zero of $y_n$ for $x>0$ and $0<c-a<1$, could be also computed by means
764: of the FPM associated to the DDEs (\ref{0-1}).
765: 
766: Considering also the restrictions imposed in the previous selection of
767: DDEs, we obtain the following:
768: \begin{theorem}
769: Let $y$ be a solution of the confluent hypergeometric equation (\ref{11}) for $x>0$. If $y$
770: has at least two zeros then $c-a>1$ and $a<0$.
771: 
772: Let $y$ be a solution of the confluent hypergeometric equation (\ref{11}) for $x<0$. If $y$
773: has at least two zeros then $c-a<0$ and $a>1$.
774: \label{condiex}
775: \end{theorem}
776: For more detailed results on the number of zeros of confluent
777: hypergeometric functions, we refer the reader to \cite{Bat},
778: Volume 1, Section 6.16.
779: 
780: 
781: The associated functions are
782: \begin{equation}
783: \begin{array}{l}
784: \eta (z(x))=-\Frac{2c-1+2x}{4\displaystyle\sqrt{(c-a)x}}\\
785: \\
786: \tilde{\mbox{A}}_n(z(x))= \Frac{8c x-16x a-3+8c-4c^2-4x^2}{16(c-a)x}\\
787: \\
788: z=2\displaystyle\sqrt{(c-a) x}
789: \end{array}
790: \end{equation}
791: 
792: Let us notice that $\eta(z(x))$ becomes negative for large $x$,
793: which gives the most appropriate sweep for large $x$ (forward)
794: since, for positive $x$, $\tilde{\mbox{A}}_n (z(x))$ decreases for
795: large $x$.
796: 
797: 
798: 
799: \subsubsection{$(k,m)=(1,1)\rightarrow y_n=\phi$($\alpha +n$; $\gamma +n$ ; x)}
800: 
801: The DDEs are the following (writing $a=\alpha+n$, $c=\gamma +n$).
802: \begin{equation}
803: \begin{array}{ll}
804: y'_n=&   \Frac{x+1-c}{x}y_n +\Frac{c-1}{x}  y_{n-1}\\
805: \\
806: y'_{n-1}= & \Frac{a-1}{c-1} y_n
807: \end{array}
808: \label{11F11}
809: \end{equation}
810: \noindent which implies the oscillatory condition $(a-1)x>0$;
811: considering $(k,m)=(-1,-1)$ or, equivalently, the DDEs relating
812: $y_{n}$, $y_{n+1}$ and their derivatives, the condition obtained
813: is $ax>0$. These conditions are consistent with Theorem
814: \ref{condiex}.
815: 
816: The functions associated to these DDEs for $x>0$ are
817: \begin{equation}
818: \begin{array}{l}
819: \eta= -\Frac{2x+3-2c }{ 4 \displaystyle\sqrt{(1-a)x}}\\
820: \\
821: z=2\displaystyle\sqrt{(1-a) x}\\
822: \\
823:  \tilde{\mbox{A}}_n (z(x))= \Frac{16x a+4x^2-8x c+3-8 c+4 c^2}{16(-1+a)x}
824: \end{array}
825: \end{equation}
826: 
827: This iteration cannot be used for $c=1$ (see Eq.\ (\ref{11F11}))
828: unless the FPI stemming from the ratio $y_n/y_{n+1}$ is used, in
829: which case the improved iteration steps cannot be applied.
830: 
831: 
832: \subsubsection{Comparing fixed point iterations}
833: 
834: As commented before, we will consider the prescription consisting in choosing the DDEs for
835: which the product $D_n=-d_n e_n$ is smaller. Applying literally this criterion, we find the
836: following preferred regions of application:
837: \begin{enumerate}
838: \item{FP(1,1)} should be applied for $x<c-a$ and FP(1,0) for $x>c-a$.
839: \item{FP(1,1)} is always better than FP(0,-1).
840: \item{FP(0,-1)} is better than FP(1,0) for $x<1-a$, but $1-a<c-a$ and FP(1,1)
841: is better there.
842: \end{enumerate}
843: 
844: Therefore, the best combination of the considered FPIs is FP(1,1)
845: for $x<c-a$ and FP(1,0) for $x>c-a$. The iteration FP(0,-1) has
846: the same behavior as FP(1,1) and can be used as a replacement when
847: $c=1$ (in this case FP(1,1) can not be used). In fact, the
848: FP(0,-1) and FP(1,1) are not independent because, as it is well
849: known, if $\psi (\alpha ;\gamma ;x)$ are solutions of the
850: confluent hypergeometric equation
851: $x\psi''+(\gamma-x)\psi'-\alpha\psi =0$, then $y(a;c;x)=x^{1-c}
852: \psi (1+a+c;2-c;x)$ is a solution of $x y'' + (c-x)y' -a y=0$.
853: 
854: Let us illustrate this behavior with numerical examples:
855: 
856: In Fig.\ 3 we compare FP(1,1) with FP(1,0) for the case of
857: Laguerre polynomials $L_{n}^{\alpha}(x)$ for the quite extreme
858: case $n=50$, $\alpha=-0.9999$. On the left, the ratio between the
859: number of iterations employed by FP(1,1) and FP(1,0) is shown as a
860: function of the location of the zeros. The first two zeros are
861: skipped in the left figure to show in more detail the behavior for
862: large $x$ (for the first zero the ratio was 40 while for the
863: second it was 5). The improvement for small $x$ when considering
864: FP(1,1) is quite noticeable. For larger $x$ ($x>c-a\simeq 50$),
865: FP(1,0) works generally better than FP(1,1) but the improvement is
866: not so noticeable. On the right figure, the number of iterations
867: used to compute each zero when considering FP(1,1) is shown.
868: 
869: \begin{minipage}{6cm}
870: \centerline{\protect\hbox{\psfig{file=ratio50.eps,width=6.cm}}}
871: \end{minipage}
872: \hfill
873: \begin{minipage}{6cm}
874: \centerline{\protect\hbox{\psfig{file=iter1150.eps,width=6.cm}}}
875: \end{minipage}
876: 
877: { {\bf Figure 3.} {\bf Left:} Ratio between the number of iterations
878: needed by FP(1,0) and FP(1,1) for the calculation of the zeros of the generalized Laguerre polynomial
879: $L_{50}^{-.9999}(x)$, as a function of the location of the zeros.
880: {\bf Right:} Number of iterations needed by FP(1,1).}
881: 
882: In conclusion, FP(1,1) is a more appropriate choice than the natural recurrence for
883: orthogonal polynomials (FP(1,0)), although FP(1,0) slightly improves the convergence
884: of FP(1,1) for $x>c-a$.
885: 
886: In Fig.\ 4 we compare FP(1,0) with FP(0,-1) for the case of
887: generalized Laguerre polynomials but now for a choice of the
888: parameters $n=50$, $\alpha=0$. This situation corresponds to the
889: case $c=1$ where FP(1,1) can not be applied. As in Fig.\ 3, the
890: ratio of the number of iterations for the first two zeros is not
891: plotted (the ratio of the first zero was 8 and for the second it
892: was 5). As expected from our previous analysis, the iteration
893: FP(0,-1) behaves quite better than FP(1,0) for small $x$.
894: 
895: \begin{minipage}{6cm}
896: \centerline{\protect\hbox{\psfig{file=ratio2.eps,width=6.cm}}}
897: \end{minipage}
898: \hfill
899: \begin{minipage}{6cm}
900: \centerline{\protect\hbox{\psfig{file=iter0m1.eps,width=6.cm}}}
901: \end{minipage}
902: 
903: {{\bf Figure 4.} {\bf Left:} Ratio between the number of iterations
904: needed by FP(1,0) and FP(0,-1) for the calculation of the zeros of the generalized Laguerre polynomial
905: $L_{50}^{0}(x)$, as a function of the location of the zeros.
906: {\bf Right:} Number of iterations needed by FP(0,-1).}
907: 
908: \subsection{Mysterious hypergeometric function}
909: 
910: This is the name given to the hypergeometric series $_2\mbox{F}_0 (a,b;;x)$,
911:  which diverges for all $x\neq 0$ (except in the terminating cases)
912: and can only be interpreted
913: in an asymptotic sense. It is well know that, for negative $x$ we have:
914: \begin{equation}
915: _2\mbox{F}_0(a,b;;x) = (-1/x)^a U(a,1+a-b,-1/x)
916: \label{myst}
917: \end{equation}
918: \noindent where $U(a,c,x)$ is a solution of the confluent hypergeometric
919: equation (\ref{11}). The functions $_2\mbox{F}_0$ can
920:  be analytically continued
921:  to the whole complex
922: plane cut along the line $\Re(z)>1$, $\Im(z)=0$. The function (\ref{myst}) is
923: a solution
924: of the 2-0 hypergeometric differential equation:
925: \begin{equation}
926: x^2 y''+[-1+x(a+b+1)]y'+ab\,y=0\,.
927: \label{mysE}
928: \end{equation}
929: 
930: In general terms, without referring to any particular solution of
931: the corresponding differential equations, the problem of computing
932: the real zeros of the mysterious hypergeometric function for
933: negative $x$ can be transformed into a problem of computation of
934: the zeros of confluent hypergeometric functions Eq.\ (\ref{myst}).
935: This is so because one can check that, if we denote by
936: $y(\alpha,\beta,x)$ a set of solutions of the confluent
937: hypergeometric equation, then $w(x)=|x|^{-a} y(a,1+a-b,-1/x)$, for
938: $x>0$ or $x<0$, are solutions of Eq.\ (\ref{mysE}).
939: 
940: For this reason and for brevity we omit further details.
941: 
942: \subsection{Gauss Hypergeometric Functions}
943: 
944:    Let us consider the hypergeometric function $_2\mbox{F}_1(a,b;c;x)$. We will consider the
945:  DDEs for families of functions of the type $\psi (\alpha + kn, \beta +m n; \gamma + l n; x)$, with
946:  $\psi (a,b;c;x)$ solutions of the hypergeometric equation (\ref{GHF}); we use the DDEs for $_2\mbox{F}_1 (a,b;c;x)$
947: series as generated by ${\it hsum.mpl}$.
948: 
949: Similarly as we did for the confluent case, we can obtain oscillatory conditions
950: for the coefficients $a$, $b$ and $c$, depending on the range of $x$.
951:  If these conditions are not satisfied by the parameters,
952: then we can assure that if there exists one zero of the function, this is an isolated zero.  As before, these conditions
953: are obtained by requiring that $d_n e_n<0$; combining the restrictions imposed by this condition for the DDEs that
954: we will later show we obtain the following:
955: 
956: \begin{theorem} Let $\psi (a,b;c;x)$ be a solution of Eq.\ (\ref{GHF}) defined in $(-\infty, 0)$, then if this function
957: is oscillatory in this interval (it has at least two zeros) then one of the following sets of conditions must be verified:
958:   $$
959:    \left\{
960:    \begin{array}{l}
961:    \mbox{(C1):  } a<0\,,\,b<0\,,\,c-a>1\,,\,c-b>1 \\
962:     \mbox{(C2):  }  a>1\,,\,b>1\,,\,c-a<0\,,\,c-b<0
963:    \end{array}
964:    \right.
965:   $$
966: 
967:  Similarly,if $\psi (a,b;c;x)$ is a solution of Eq.\ (\ref{GHF}) defined in $(0, 1)$, the oscillatory conditions are
968:    $$
969:    \left\{
970:    \begin{array}{l}
971:     \mbox{(C3):  } a<0\,,\,b>1\,,\,c-a>1\,,\,c-b<0 \\
972:      \mbox{(C4):  }  a>1\,,\,b<0\,,\,c-a<0\,,\,c-b>1
973:    \end{array}
974:    \right.
975:   $$
976: 
977:  Finally, if $\psi (a,b;c;x)$ is a solution of Eq.\ (\ref{GHF}) defined in $(1, +\infty)$ and is oscillatory in this
978: interval,
979:  one of the following sets of conditions must be verified:
980:    $$
981:    \left\{
982:    \begin{array}{l}
983:     \mbox{(C5):  } a<0\,,\,b<0\,,\,c-a<0\,,\,c-b<0 \\
984:     \mbox{(C6):  } a>1\,,\,b>1\,,\,c-a>1\,,\,c-b>1
985:    \end{array}
986:    \right.
987:   $$
988: \label{oscG}
989: \end{theorem}
990: 
991:   The different sets of conditions for the three subintervals in Theorem \ref{oscG} can be obtained by combining
992: the restrictions obtained
993: from the following values of $(k,l,m)$: $(\pm 1,0,0)$, $(\pm 1,\pm 1,0)$, $(\pm 1, 0,\pm 1)$, $(\pm 1,\mp 1,0)$,
994:  $(0,0,\pm 1)$ and $(\pm 1,\pm 1, \pm 1)$. In any case, the analysis for the three different subintervals are
995: not independent because, as it is well know (see \cite{Bat}, Vol.\
996: I, Chap.\ II), if we denote by $\psi (\alpha,\beta;\gamma,x)$ the
997: solutions of the hypergeometric equation
998: $x(1-x)y''+(\gamma-(\alpha +\beta +1)x)y'-\alpha\beta y=0$ in the
999: interval $(0,1)$ one can write solutions in the other two
1000: intervals by using that both
1001: \begin{equation}
1002: y(a,b;c;x)=(1-x)^{-a} \psi (a,c-b;c;x/(x-1))\,,x<0
1003: \label{x<0}
1004: \end{equation}
1005: \noindent and
1006: \begin{equation}
1007: y(a,b;c;x)=x^{-a} \psi (a,a+1-c;a+b+1-c;1-1/x)\,,x>1
1008: \label{x>1}
1009: \end{equation}
1010: are solutions of the hypergeometric differential equation
1011: $x(1-x)y''+(c-(a+b+1)x)y'-aby=0$. With this, it is easy to see
1012: that the conditions C1 and C2 can be obtained from C3 and C4
1013: respectively by using Eq.\ (\ref{x<0}), while C5 and C6 derive
1014: from C3, C4 and Eq.\ (\ref{x>1}).
1015: 
1016:  Notice, in addition, that the 6 conditions in Theorem \ref{oscG} are mutually exclusive, which means that
1017: 
1018: \begin{theorem}
1019: Given three values of the parameters $a$, $b$ and $c$, at most one
1020: of the subintervals $(-\infty,0)$, $(0,1)$, $(1,+\infty)$
1021: possesses oscillatory solutions (solutions with at least two
1022: zeros).
1023: \end{theorem}
1024: 
1025:  The oscillatory conditions C6 is of no use for non-terminating hypergeometric series,
1026:  because they diverge for $x>1$; however, this is a possible case for
1027:  other solutions of the differential equation. The conditions C3 correspond
1028:  to Jacobi polynomials $P_{n}^{(\alpha ,\beta)}(x)=\Frac{(\alpha +1)_n}{n!}$ $_
1029: 2\mbox{F}_1(-n,1+\alpha+\beta+n;\alpha +1;(1-x)/2)$ ($\alpha
1030: ,\beta >-1$) of order $n\ge 2$ (for order $n=1$ the conditions are
1031: not satisfied, not surprisingly because according to our criteria
1032: a polynomial of degree 1 is non-oscillating). Particular cases of
1033: Jacobi polynomials are Gegenbauer ($\alpha=\beta$), Legendre
1034: ($\alpha=\beta=0$) and Chebyshev ($\alpha=\beta =-1/2$)
1035: polynomials.
1036: 
1037: \subsubsection{DDEs and change of variables}
1038: 
1039:  In this section, we compile the expressions for the different DDEs as well as the associated
1040: change of variable. For brevity, the associated functions $\eta
1041: (x)$ and $\tilde{\mbox{A}}_n (x)$ are not shown. 
1042: 
1043: It is understood that $a=\alpha + k\,n$, $b=\beta +l\,n$, $c=\gamma +m\,n$ for DDE$(k,l,m)$.
1044: 
1045: \begin{enumerate}
1046: \item{DDE$(1,0,0)$}
1047: $$
1048: \begin{array}{ll}
1049: y'_n=&\Frac{-a-bx+c}{x(x-1)}y_n+\Frac{a-c}{x(x-1)}  y_{n-1}\\
1050: \\
1051: y'_{n-1}= & \Frac{1-a}{x}y_{n-1}+\Frac{a-1}{x}y_n\\
1052: \end{array}
1053: $$
1054: 
1055: Change of variable:
1056: $z(x)=-2\sqrt{(c-a)(1-a)}\tanh^{-1}(\sqrt{1-x})$
1057: 
1058: \item{$DDE(1,1,0)$ :}
1059: $$
1060: \begin{array}{ll}
1061: y'_n=&\left(\Frac{-(a+b)(a+b-c-1)+ab-c}{
1062: x(a+b-c-1)}+\Frac{a+b-c}{x(1-x)}\right)y_n+
1063: \Frac{(a-c)(c-b)}{x(1-x)(a+b-c-1)}
1064:   y_{n-1}\\
1065: \\
1066: y'_{n-1}= & \Frac{a+b-ab-1}{(a+b-c-1)x} y_{n-1}
1067: -\Frac{(1-x)(a-1)(1-b)}{(a+b-c-1)x}y_n\\
1068: \end{array}
1069: $$
1070: 
1071: Change of variable: $z(x)=\Frac{\sqrt{(b-c)(c-a)(b-1)(1-a)}}{|
1072: a+b-c-1|}\ln x$,  $x>0$.
1073: 
1074: \item{$DDE(1,1,2)$ :}
1075: $$
1076: \begin{array}{ll}
1077: y'_n=& \Frac{(1-x)\left[(1-a-b)(c-1)+ab\right]+(c-1)(1+a+b-c)-ab}{x(1-x)(c-2)}y_n-\Frac{(1-c)}{x(1-x)} y_{n-1}\\
1078: \\
1079: y'_{n-1}= &\Frac{1-a-b+ab}{(1-x)(c-2)} y_{n-1}
1080: -\Frac{x(a-c+1)(1-a)(c-b-1)(b-1)}{(1-x)(c-1)(c-2)^2}y_n\\
1081: \end{array}
1082: $$
1083: 
1084: Change of variable: $z(x)=-\displaystyle\sqrt{\Frac{(c-a-1)(1-a)(1+b-c)(b-1)}{(c-2)^2}}\ln (1-x)$.
1085: 
1086: \item{$DDE(1,0,1)$ :}
1087: $$
1088: \begin{array}{ll}
1089: y'_n=& \Frac{1+xb-c}{x(1-x)}y_n-
1090: \Frac{(1-c)}{x(1-x)}
1091:   y_{n-1}\\
1092: \\
1093: y'_{n-1}= & -\Frac{(1-a) }{(1-x) } y_{n-1}
1094: +\Frac{(1-a)(b+1-c)}{(1-x)(1-c)}y_n\\
1095: \end{array}
1096: $$
1097: 
1098: Change of variable: $z(x)=2\sqrt{(1-a)(b+1-c)}\tanh^{-1} \sqrt{x}$.
1099: 
1100: \item{$DDE(1,-1,0)$ :}
1101: $$
1102: \begin{array}{ll}
1103: y'_n=&b\Frac{x(b-a+1)+a-c}{x(1-x)(b-a+1)}y_n+
1104: \Frac{b(c-a)}{x(1-x)(b-a+1)}y_{n-1}\\
1105: \\
1106: y'_{n-1}= & (1-a)\Frac{(1-x)(b-a+1)+a-c}{x(1-x)(b-a+1)} y_{n-1}
1107: -\Frac{(1-a)(1+b-c)}{x(1-x)(b-a+1)}y_n\\
1108: \end{array}
1109: $$
1110: 
1111: Change of variable: $z(x)=
1112: \Frac{\sqrt{b(c-a)(1-a)(1+b-c)}}{b-a+1}\ln\left(\Frac{x}{1-x}\right)$.
1113: 
1114: \item{$DDE(0,0,-1)$ :}
1115: $$
1116: \begin{array}{ll}
1117: y'_n=&\Frac{b+a-c}{1-x}y_n-\Frac{(b-c)(c-a)}{(1-x)c }y_{n-1}\\
1118: \\
1119: y'_{n-1}= & -\Frac{c}{x} y_{n-1}+\Frac{c}{x}y_n\\
1120: \end{array}
1121: $$
1122: 
1123: Change of variable: $z(x)=\sqrt{(b-c)(c-a)} \arcsin (2x-1)$.
1124: 
1125: \item{$DDE(1,1,1)$ :}
1126: $$
1127: \begin{array}{ll}
1128: y'_n=&\Frac{x(a+b-1)-c+1}{x(1-x)}y_n-\Frac{1-c}{x(1-x)}y_{n-1}\\
1129: \\
1130: y'_{n-1}= & \Frac{(b-1)(1-a)}{(1-c)}y_n\\
1131: \end{array}
1132: $$
1133: 
1134: Change of variable: $z(x)=\sqrt{(b-1)(1-a)} \arcsin (2x-1)$.
1135: 
1136: \end{enumerate}
1137: 
1138: \subsubsection{Comparison of FPIs}
1139: 
1140:    The following table shows the $|D_{n}|=|-d_n e_n|$ coefficient for the FPIs
1141: $(1,0,0)$, $(1,1,0)$, $(1,1,2)$, $(1,0,1)$, $(1,-1,0)$, $(0,0,-1)$ and $(1,1,1)$.
1142: $$
1143: \begin{array}{l|c}
1144: \mbox{Iteration} & |D_n|  \\
1145: \hline
1146:  (1,0,0) &   \displaystyle\left|\Frac{(a-c)(a-1)}{x^2(1-x)}\right|\\
1147: \hline
1148: (1,1,0) &    \displaystyle\left|\Frac{(b-c)(a-c)(b-1)(a-1)}{x^2(a+b-c-1)^2}\right|        \\
1149: \hline
1150: (1,1,2) &    \displaystyle\left|\Frac{(c-a-1)(1-a)(1+b-c)(b-1)}{(1-x)^2(c-2)^2}\right|        \\
1151: \hline
1152: (1,0,1) &    \displaystyle\left|\Frac{(a-1)(1+b-c)}{x(1-x)^2}\right|     \\
1153: \hline
1154: (1,-1,0) &    \displaystyle\left|\Frac{b(c-a)(a-1)(c-b-1)}{x^2(1-x)^2(a-b-1)^2}\right|        \\
1155: \hline
1156: (0,0,-1) &     \displaystyle\left|\Frac{(b-c)(a-c)}{x(1-x)}\right|      \\
1157: \hline
1158: (1,1,1)  &      \displaystyle\left|\Frac{(b-1)(a-1)}{x(1-x)}\right|
1159: \end{array}
1160: $$
1161: 
1162: As can be inferred from the table, the most appropriate FPIs in
1163: the interval $(0,1)$ are the $(0,0,-1)$ and $(1,1,1)$ iterations.
1164: As commented, some hypergeometric functions in this interval with
1165: particular values of their parameters are orthogonal polynomials
1166: (Jacobi and derived polynomials).
1167: In the case of orthogonal polynomials, the ``natural'' iteration to be considered is $(1,-1,0)$
1168: (or $(-1,1,0)$ equivalently) which is not the optimal iteration. In order to illustrate this fact,
1169: let us consider the evaluation of the zeros of the hypergeometric function
1170: $_2\mbox{F}_1(-50,54;5/2;x)$ in the interval $(0,1)$. The zeros of this function
1171: correspond to the zeros of the Jacobi polynomial $P^{(3/2,3/2)}_{50}(1-2x)$.
1172:  In Figure 5, we show the ratio between the number of iterations
1173: needed by FP$(1,-1,0)$ and FP$(1,1,1)$ as a function of the location of the zeros.
1174: 
1175: On the contrary, the most appropriate iteration in the $(1,
1176: \infty)$ interval is the $(1,0,0)$ iteration. This could also be
1177: understood taking into account that the best iterations in the
1178: interval $(0,1)$ are FPI(0,0,-1) and FPI(1,1,1) and using Eq.\
1179: (\ref{x>1}). In Figure 6  we show the ratio between the number of
1180: iterations needed by FP$(1,1,1)$ and FP$(1,0,0)$.
1181: 
1182: The main conclusion for $_2\mbox{F}_1$ hypergeometric functions is
1183: that the iteration $(1,1,1)$ is the preferred one and that
1184: $(0,0,-1)$ can be considered as a replacement, with similar
1185: performance. For the other two intervals, the relations
1186: (\ref{x<0}) and (\ref{x>1}) indicate that the most appropriate
1187: iterations will be $(1,0,1)$ for $(-\infty, 0)$ and $(1,0,0)$ for
1188: $(1,+\infty)$; in any case, the solutions in these intervals can
1189: be related to solutions in $(0,1)$.
1190: 
1191: \vspace*{0.2cm}
1192: 
1193: \centerline{\protect\hbox{\psfig{file=hyp1.eps,width=7.cm,height=6.cm}}}
1194: 
1195: {{\bf Figure 5}. Ratio between the number of iterations
1196: needed by FP(1,-1,0) and FP(1,1,1) for the calculation of the zeros of the hypergeometric
1197: function $_2\mbox{F}_1(-50,54;5/2;x)$, as a function of the location of the zeros.}
1198: 
1199: 
1200: \centerline{\protect\hbox{\psfig{file=hyp2.eps,width=7.cm,height=6.cm}}}
1201: 
1202: {{\bf Figure 6}. Ratio between the number of iterations
1203: needed by FP(1,1,1) and FP(1,0,0) for the calculation of the zeros of the hypergeometric
1204: function $_2\mbox{F}_1(-30,-32;-70;x)$, as a function of the location of the zeros.}
1205: 
1206: \section{Conclusions}
1207: 
1208: We have developed a detailed study of the performance of the available fixed point methods for hypergeometric
1209:  functions. This will allow the construction of efficient algorithms for the computation of the real
1210: zeros of hypergeometric functions with a good asymptotic behavior.
1211: The next table summarized the main results.
1212: 
1213: $$
1214: \begin{array}{cccc}
1215: \mbox{Function} & \mbox{Contrast function} & \mbox{Range of application} &
1216: \mbox{Change of variables } z(x)\\
1217: \hline
1218: _0\mbox{F}_1 (;c;-x) & _0\mbox{F}_1 (;c_{-};-x) &
1219: \left\{\begin{array}{l}
1220: c<100\\
1221: c>100\,,\,x>c^2/2
1222: \end{array}
1223: \right. & 2\sqrt{x}\\
1224: `` & _0\mbox{F}_1 (;c_{-};-x) & c>100\,,\,x<c^2 /2 & x/(\nu-1)\\
1225: \hline
1226: _1\mbox{F}_1(a;c;x) & _1\mbox{F}_1 (a_{-};c_{-};x) & x<c-a & 2\sqrt{(1-a)}x\\
1227: `` & _1\mbox{F}_1 (a_{-};c;x) & x>c-a & N_{ac} \ln x \\
1228: \hline
1229: _2\mbox{F}_1 (a,b;c;x) & _2\mbox{F}_1 (a_{-},b_{-};c_{-};x) & & N_{bc} \arcsin(2x-1)\\
1230: \hline
1231: \end{array}
1232: $$
1233: \noindent where $N_{ac}=\sqrt{(c-a)(1-a)}$ and $N_{bc}=\sqrt{(b-1)(1-a)}$ and
1234: $a_{-}=a-1$, $b_{-}=b-1$, $c_{-}=c-1$.
1235: 
1236: In the table we restrict ourselves to $x>0$ in all cases, with the
1237: additional restriction $x<1$ for the $_2\mbox{F}_1$ functions. For
1238: the rest of the intervals, as commented before, relations are
1239: available which map these other regions into the intervals
1240: considered in the table.
1241: 
1242: It should be noted that, when a priori approximations to the roots
1243: are available (for instance, asymptotic approximations like in
1244: \cite{Tem}) the performance of the algorithms can be improved.
1245: However, the methods presented here have the advantage of being
1246: efficient methods that do not require specific approximations for
1247: specific functions (which, on the other hand, are difficult to
1248: obtain for three parameter functions like the $_2\mbox{F}_1$
1249: hypergeometric functions). In addition, even in the simple cases
1250: of one parameter functions, the methods are very efficient by
1251: themselves.
1252: 
1253: To conclude, it is worth mentioning that one on the main reasons
1254: for the good performance of the algorithms if that the analytical
1255: transformations of the DDEs, and in particular, the associated
1256: change of variable $z(x)=\int\sqrt{-d_n e_n}dx $, tend to
1257: uniformize the distance between zeros. Generally speaking, the
1258: most successful  methods are those which produce smaller
1259: variations of the distances between zeros, because the first
1260: guesses for the zeros become more accurate. In connection to this,
1261: these changes of variable lead to interesting analytical
1262: information about these zeros \cite{Dea}.
1263: 
1264: \begin{appendix}{Maple Code}
1265: In this appendix we would like to  explain how we received the DDEs
1266: automatically. Our Maple code uses a Zeilberger type approach
1267: \cite{Zei} and is completely on the lines of \cite{Koe}. The
1268: authors provide a Maple program {\it rules.mpl} which can be used
1269: in combination with {\it hsum.mpl} \cite{Koe} to get equations
1270: (15), (18), (22), (24), (26), (28), and the DDEs in \S~3.4.1.
1271: These computations are collected in the Maple worksheet {\it
1272: rules.mpl}. All these files can be obtained from the web site
1273: \url{http://www.mathematik.uni-kassel.de/~koepf/Publikationen}.
1274: \end{appendix}
1275: 
1276: \begin{acknowledgment}
1277:   A. Gil acknowledges support from A. von Humboldt foundation. J. Segura acknowledges support from DAAD.
1278: \end{acknowledgment}
1279: 
1280: \pagebreak
1281: \begin{references}
1282: 
1283: \bibitem{Seg3} Gil, A. and J. Segura, ``Computing zeros and turning points of linear homogeneous second order ODEs'',
1284: SIAM J. Numer. Anal.  41 (2003) 827-855.
1285: 
1286: \bibitem{Bat} Erd\'elyi, A., W. Magnus, F. Oberhettinger, F.G. Tricomi. Higher Transcendental Functions, McGraw-Hill
1287: (1953).
1288: 
1289: \bibitem{Dea} Dea\~no, A., A. Gil, J. Segura. ``New inequalities from the classical Sturm comparison theorem'', in
1290: preparation.
1291: 
1292: \bibitem{Gil} Gil, A. and J. Segura, ``A combined symbolic and numerical algorithm for the computation
1293: of zeros of orthogonal polynomials and special functions'', J. Symbol. Comp. 35 (2003) 465-485.
1294: 
1295: \bibitem{Koe} Koepf, Wolfram. ``Hypergeometric Summation. An Algorithmic Approach to Summation and Special Function
1296: Identities''. Vieweg, Braunschweig/Wiesbaden, 1998
1297: 
1298: \bibitem{Leb} Lebedev, N.N. ``Special Functions and Their Applications''. Dover Publications (1972).
1299: 
1300: \bibitem{Seg2} Segura, J., ``The zeros of special functions from a fixed point method'', SIAM J. Numer.
1301: Anal. 40 (2002) 114-133.
1302: 
1303: \bibitem{Tem} Temme, N.M., ``An algorithm with Algol 60 program for the computation of the zeros of ordinary Bessel
1304: functions and those of their derivatives'', J. Comput. Phys. 32 (1979) 270-270.
1305: 
1306: \bibitem{Zei} Zeilberger, D., ``A fast algorithm for proving terminating hypergeometric
1307: identities.'' Discrete Math.\ 80 (1990) 207-211.
1308: 
1309: \end{references}
1310: 
1311: \end{article}
1312: \end{document}
1313: