1: \documentclass[makeidx,reqno]{amsart}
2: %\usepackage[T1]{fontenc}
3: %\usepackage[latin9]{inputenc}
4: %\usepackage{geometry}
5: \usepackage[active]{srcltx}
6: \usepackage{epsfig}
7: \usepackage{hyperref}
8: \usepackage{amsmath}
9: \usepackage{amssymb}
10: %\usepackage{esint}
11: \newtheorem{Proposition}{Proposition}
12: \newtheorem{Remark}{Remark}
13: \newtheorem{Corollary}[Proposition]{Corollary}
14: \newtheorem{Lemma}[Proposition]{Lemma}
15: \newtheorem{Theo}[Proposition]{Theorem}
16: \newtheorem{Theorem}{Theorem}
17: \newtheorem{Consequence}[Proposition]{Consequence}
18: \newtheorem{Definition}[Proposition]{Definition}
19: \newtheorem{Note}[Remark]{Note}
20: \newtheorem{Convention}[Proposition]{Convention}
21: \newtheorem{Assumptions}{Assumption}
22: \newtheorem{Normalization}{Normalization}
23: \newcommand {\z}{{\noindent}}
24: \def\HH{\mathbb{H}}
25: \def\CC{\mathbb{C}}
26: \def\RR{\mathbb{R}}
27: \def\NN{\mathbb{N}}
28: \def\ZZ{\mathbb{Z}}
29: \def\QQ{\mathbb{Q}}
30: \def\Re{\mathrm{Re}}
31: \def\Im{\mathrm{Im}}
32: \def\ds{\displaystyle}
33: \renewcommand{\labelenumi}{$ $\hskip -0.4cm (\roman{enumi})}
34: \newcommand{\theitemi}{(\roman{itemi})}
35: \def\bb{b} \def\({\left(} \def\){\right)} \makeindex
36: \author{O. Costin, M. Huang} \address{Mathematics Department//The Ohio State University//Columbus, OH 43220} \title{Behavior of lacunary
37: series at the natural boundary} \gdef\shorttitle{Natural boundaries}
38: \begin{document}
39: \begin{abstract}
40: We develop a local theory of lacunary Dirichlet series of the form
41: $\sum\limits_{k=1}^{\infty}c_k\exp(-zg(k)), \Re(z)>0$ as $z$
42: approaches the boundary $i\RR$, under the assumption $g'\to\infty$
43: and further assumptions on $c_k$. These series occur in many
44: applications in Fourier analysis, infinite order differential
45: operators, number theory and holomorphic dynamics among others. For
46: relatively general series with $c_k=1$, the case we primarily focus
47: on, we obtain blow up rates in measure along the imaginary line and
48: asymptotic information at $z=0$.
49:
50: When sufficient analyticity information on $g$
51: exists, we obtain Borel summable expansions at points on
52: the boundary, giving exact local
53: description. Borel
54: summability of the expansions provides property-preserving
55: extensions beyond the barrier.
56:
57: The singular
58: behavior has remarkable universality and self-similarity features. If
59: $g(k)=k^b$, $c_k=1$, $b=n$ or $b=(n+1)/n$, $n\in\NN$,
60: behavior near the boundary is roughly of the standard form $\Re(z)^{-b'}Q(x)$ where
61: $Q(x)=1/q$ if $x=p/q\in\QQ$ and zero otherwise.
62:
63: The B\"otcher map at infinity of polynomial iterations of the form
64: $x_{n+1}=\lambda P(x_n)$, $|\lambda|<\lambda_0(P)$, turns out to have
65: uniformly convergent Fourier expansions in terms of simple lacunary
66: series. For the quadratic
67: map $P(x) =x-x^2$, $\lambda_0=1$, and the Julia set is the graph of this Fourier expansion in the main cardioid of the Mandelbrot set.
68:
69:
70:
71: \end{abstract}
72: \maketitle
73: \tableofcontents
74: \section{Introduction}
75:
76: Natural boundaries (NBs) occur frequently in many
77: applications of analysis, in the theory of Fourier series, in
78: holomorphic dynamics (see \cite{Douady}, \cite{Devaney}, \cite{CPAM}
79: and references there), in analytic number theory, see
80: \cite{Titchmarsh}, physics, see \cite{Kawai} and even in relatively simple ODEs such as the
81: Chazy equation \cite{Ince}, an equation arising in conformal mapping
82: theory, or the Jacobi equation.
83:
84:
85: The intimate structure of NBs turns out to be particularly rich,
86: bridging analysis, number theory and complex dynamics.
87:
88:
89:
90: Nonetheless (cf. also \cite{Kawai}), the study of NBs of concrete functions is yet to
91: be completed from a pure analytic point of view. The aim of the present paper is a detailed study near the analyticity
92: boundary of a prototypical functions exhibiting this
93: singularity structure, classes of {\em lacunary series}. For such functions, we
94: develop a theory of generalized local asymptotic expansions at NBs, and explore their consequences and applications. The expansions are asymptotic in the sense
95: that they become increasingly accurate as the singular curve is approach, and in many cases exact, in that the function can be recovered from these expansions.
96:
97:
98: Lacunary series, sums of the form $h(s)=\sum_{j\ge 1}{c_j}s^{g_j}$, or written
99: as Dirichlet series $f(z)=\sum_{j\ge 1}{c_j}e^{-z{g(j)}}$), where
100: $j/g_j=o(1)$ for large $j$, often occur in applications and have deep
101: connections with infinite order differential operators, as found and studied by
102: Kawai \cite{Kawai1}, \cite{Kawai}. Under the lacunarity assumption
103: above, if the unit disk is the maximal disk of analyticity of $h$,
104: then the unit circle is its NB (\cite{Mandelbrojt}). For
105: instance, the series
106: \begin{equation}
107: \label{eq03}
108: h(s)=\sum_{j=1}^{\infty}s^{2^j}\:(|s|<1)
109: \end{equation}
110: studied by Jacobi \cite{Jacobi}
111: before the advent of modern Complex Analysis, clearly has the
112: unit disk as a singular curve: $h(s)\to+\infty$ as $|s|\uparrow
113: 1$ along any ray of angle $2^{-n}m\pi $ with $m,n\in\NN$.
114:
115:
116: We show that if $c_j=1$ and $g'\to\infty$,
117: then, in a measure theoretic sense, $f(x+iy)$ blows up
118: as $x\to 0^+$ uniformly in $y$ at a calculable rate. We find
119: interesting {\em universality properties in the blow-up profile}.
120:
121:
122: In special cases of interest, Borel summable power series, in powers
123: of the distance to the boundary, and more generally convergent
124: expansions
125: as series of small exponentials multiplying Borel summed series power series\footnote{These are simple instances of Borel (or Ecalle-Borel) summed transseries. A brief summary of the definitions and properties
126: of transseries and generalized Borel summability, and references to
127: the literature are given in \S\ref{A4}. \label{f1}}
128: representations
129: can be determined on a dense set on the singularity
130: barrier. Examples are
131: \begin{itemize}
132:
133: \item $\sum_{j\ge 1}e^{-z j^b}$ where $b>1$, or its dual $q>1$, where
134: $ b^{-1}+q^{-1}=1$, is integer (relating to exponential sums and van
135: der Corput dualities \cite{Montgomery}; the special self-dual case
136: $\sum_{j\ge 1}e^{-z j^2}$, is related to the Jacobi theta function);
137:
138: \item $\sum_{j\ge 1}e^{-z a^j}$, $1<a\in\NN$;
139: \end{itemize}
140: More generally, if $c_j=c(j)$ and $g_j=g(j)$ have suitable analyticity
141: properties in $j$, then the behavior at $z=0^+$, and possibly at other
142: points, is described in terms of Ecalle-Borel summed expansions (see
143: footnote \ref{f1}). Then the analysis leads to a natural,
144: properties-preserving, continuation formulas across the boundary.
145:
146:
147:
148: In general, the blow-up profile along the barrier is closely related to
149: {\em exponential sums,} expressions of the form
150: \begin{equation}
151: \label{eq:es1}
152: S_N=\sum_{k=1}^N c_n\exp(2\pi i g(n)),\ \ g(n)\in\RR
153: \end{equation}
154: where for us $g'\to\infty$ as $n\to\infty$. The corresponding lacunary
155: series are in a sense the continuation of (\ref{eq:es1}) in the
156: complex domain, replacing $2\pi i $ by $-z$, $\Re(z)>0$, and letting
157: $N\to \infty$. The asymptotic behavior of lacunary series as the
158: imaginary line is approached in nearly-tangential directions is
159: described by dual, van der Corput-like, expansions. The method we use
160: extend to exponential sums, which will be the subject of a different
161: paper.
162:
163: \section{Results}
164: \subsection{Results under general assumptions}
165: \subsubsection{Blow-up on a full measure set}
166: We consider lacunary Dirichlet series of the form
167: \begin{equation}
168: \label{lacn1}
169: f(z)= \sum_{k=0}^\infty e^{-zg(k)}
170: \end{equation}
171: but, as it can be seen from the proofs, the analysis extends easily to series of the form
172: $$f(z)= \sum_{j=0}^\infty c_j e^{-zg(j)}$$
173: under suitable smoothness and growth conditions, see
174: \S\ref{genblowup}. The results in the paper apply under the further restriction,
175: \begin{Assumptions}\label{A1}
176: The function $g$ is differentiable and $g'(j)\to \infty$ as $j\to\infty$.
177: \end{Assumptions}
178:
179:
180: In particular, $g$ is eventually increasing. By subtracting a finite
181: sum of terms from $f$ (a finite sum is clearly entire), we arrange that $g$ is
182: {\bf increasing}. If $g(0)=a$, we can multiply $f$ by $e^{-za}$ to
183: arrange that $g(0)=0$.
184:
185: \begin{Normalization}
186:
187:
188: (i) $g$ is differentiable on $[0,\infty)$, $g'>0$ and $g'\to\infty$
189: along $\RR^+$.
190:
191: (ii) $g(0)=0$.
192: \end{Normalization}
193:
194: \bigskip
195:
196: \z {\bf Notation.} We write $\ds |H(\cdot)|\mathop{=}^\mu 1+o(1)$ if $|H(y)|dy$ converges to
197: the Lebesgue measure $d\mu(y)$.
198:
199: \bigskip
200:
201: Under Assumption \ref{A1}, after normalization, we have the following result,
202: giving exact blow-up rates in measure, as well as sharp pointwise blow-up upper bounds.
203: \begin{Theorem}\label{P1}
204: (i) We have the uniform
205: blow-up rate in measure\footnote{It turns out that in general, $L^1$ or a.e. convergence of $|f|$ do
206: not hold.}.
207: \begin{equation}
208: \label{eq:asbh}
209: |f(x+i\cdot)|^2\mathop=^\mu\int_{0}^{\infty}e^{-2g(s)x}ds\left(1+o(1)\right)
210: \end{equation}
211: It can be checked that $ \int_{0}^{\infty}e^{-2g(s)x}ds\ \ge g^{-1}(1/x)\to\infty {\text{ as }} x\to 0^+$; see also Note~\ref{n1}.
212:
213:
214:
215: (ii) The following pointwise estimate holds:
216: \begin{equation}
217: \label{eq:locb}
218: \|f(x+i\cdot)\|_{\infty}\le \int_{0}^{\infty}e^{-g(s)x}ds(1+o(1))\ \ \text{as } x\to 0^+
219: \end{equation}
220: This is sharp at $z=0$, cf. Proposition \ref{Prop2}, and
221: in many cases it is only reached at $z=0$; see Proposition~\ref{Pp4}.
222: \end{Theorem}
223:
224:
225: \subsubsection{General behavior near $z=0$}
226: At $z=0^+$ a more detailed asymptotic description is possible.
227: \begin{Theorem}\label{Prop2}
228: (i) As $z\to 0^+$ we have
229: \begin{equation}
230: \label{eq:asympt1}
231: f(z)=\int_0^{\infty}e^{-z g(s)}ds-\frac{1}{2}+o(1)\ \to\infty\ \ \ \text{as } z\to 0^+
232: \end{equation}
233: In fact,
234: \begin{equation}
235: \label{eq:asympt1}
236: f(z)-\int_0^{\infty}e^{-z g(s)}ds=-z\int_{0}^{\infty} e^{-zu}
237: \{g^{-1}(u)\}du
238: \end{equation}
239: (where $\{\cdot\}$ denotes the fractional part, and we used $g(0)=0$)\footnote{As mentioned, often this maximal growth
240: is achieved at zero but in special cases it occurs, up to a
241: bounded function, on a dense set of measure zero.}.
242:
243:
244: (ii) If $g(s)$ has a
245: differentiable asymptotic expansion as $s\to\infty$ in terms of (integer or noninteger)
246: powers of $s$ and $\log s$, and $g(s)\sim const. s^b $, $b>1$, then after
247: subtracting the blowing up term, $f$
248: has a Taylor series at $z=0$ (generally divergent, even when $g$ is analytic,
249: which can be calculated explicitly),
250: $$
251: f(z)=
252: \int_0^{\infty}e^{-z g(s)}ds-\frac{1}{2}+zs(z), \ \ s\in C^\infty[0,\infty)$$
253: (as an example, see (\ref{eq:srjtob}).
254: \end{Theorem}
255:
256: \begin{Note}\label{n1}
257: Often $g$ has
258: an asymptotic expansion starting with a combination of powers, exponentials and
259: logs. Let $\phi=g^{-1}$. Then $\phi(\nu x)/\phi(\nu)\to \phi_1(x)$ as $\nu\to\infty$ and $x>0$ is
260: fixed and
261: $$\int_0^{\infty}e^{-x g(s)}ds=C_gg^{-1}(1/x)(1+o(1));\ \ \ C_g=\int_0^{\infty}e^{-u}\phi_1(u)du$$
262: \end{Note}
263: For instance, if $g(k)=k^b,b>1$ we have, as $\rho\to 1$,
264: \begin{eqnarray}
265: \label{nb}
266: |f(z)|\mathop=^\mu x^{-\frac{1}{2b}} \,\,\Gamma(1+1/b)^{-1/2}2^{-\frac{1}{2b}}(1+o(1)) \ \ \ (x\to 0^+)\nonumber \\
267: f(x)=x^{-\frac{1}{b}} \,\, \Gamma(1+1/b)(1+o(1))\ \ \ \ \ \ \ (x\to 0^+)
268: \end{eqnarray}
269:
270: \subsubsection{Blow-up profile along barrier}
271:
272: \begin{Theorem}\label{Cexp}
273: (i) Assume that for some $y\in\RR$ there is a smooth increasing
274: function $\rho(N;y)=:\rho(N)\in(0,N]$ such that the following
275: weighted exponential sum (see \cite{Montgomery}) has a limit:
276: \begin{equation}
277: \label{eq:expos}
278: S_{\rho,N}:= \rho(N)^{-1}\sum_{j=1}^N e^{i y g(j)}\to L(y)\ \ \text{as}\ \ N\to \infty
279: \end{equation}
280: where $\rho''$ is uniformly bounded and nonpositive for sufficiently
281: large $k$. (Without loss of generality, we may assume $\rho''(k)\le
282: 0$ for all $k$.) Let
283: \begin{equation}
284: \label{eq:H1}
285: \Phi(x)=\int_0^{\infty}e^{-xg(u)}\rho'(u)du
286: \end{equation}
287: Then, we have the asymptotic behavior
288: \begin{equation}
289: \label{eq:lim2}
290: f(x+iy)=L(y)\Phi(x)+o(\Phi(x))\ \ \ (x\to 0^+)
291: \end{equation}
292: (ii) As a pointwise upper bound we have:
293: \begin{equation}
294: \label{eq:lsup}
295: \limsup_{N\ge 0} |S_{\rho,N}|=L<\infty\ \Rightarrow \ f(x+iy)=O(\Phi(x))\ \ \ (x\to 0^+)
296: \end{equation}
297: \end{Theorem}\label{P4}
298:
299: \subsection{Results in specific cases}
300:
301:
302:
303:
304: \subsubsection{Settings leading to convergent expansions}
305: The cases $g(j)=j^2$ and $g(j)=a^j$, $a>1$ are distinguished, since
306: the expansions at some points near the boundary converge.
307:
308: \begin{Proposition}\label{b=2}
309: \z If $b=2$, then we have the identity \[
310: f(z)=\frac{1}{2}\sqrt{\frac{\pi}{z}}-\dfrac{1}{2}+\sqrt{\frac{\pi}{z}}\sum_{k=1}^{\infty}e^{-\frac{k^{2}\pi^{2}}{z}}\]
311:
312: \end{Proposition}
313: Clearly, this is most useful when $z\to 0$. It also shows the identity
314: associated to the Jacobi theta function
315: \begin{equation}
316: \label{eq:ide2}
317: f(z)=\frac{1}{2}\sqrt{\frac{\pi}{z}}-\dfrac{1}{2}
318: +\sqrt{\frac{\pi}{z}}f\left(\frac{k^2\pi^2}{z}\right)
319: \end{equation}
320: \begin{Proposition}\label{Prop4}
321: If $g(j)=a^j$, $a>1$, then, as $z\to 0^+$, $f(z)$
322: is convergently given by
323: \begin{multline}\label{convt}
324: f(z)=-\dfrac{\log \zeta}{\log a}+\sum_{n=1}^{\infty}\frac{(-\zeta )^{n}}{n!(1-a^{n})}+c_{0}+\frac{1}{\log a}\sum_{k\neq0}\Gamma\left(-\frac{2k\pi i}{\log a}\right)\zeta ^{\frac{2k\pi i}{\log a}}\\=-\dfrac{\log \zeta }{\log
325: a}+\sum_{n=1}^{\infty}\frac{(-\zeta )^{n}}{n!(1-a^{n})}
326: +c_{0}-
327: \frac{1}{2\pi i}\int_{0}^{\infty}\log_{\RR}[-(s/\zeta )^{\frac{2\pi i}{\log a}}]e^{-s}ds
328: \end{multline}
329: where $\zeta=1-e^{-z}$.
330: (Here $\log_{\RR}$ is the usual branch of the log with a cut along
331: $\RR^-$ and {\bf not} the log on the universal covering of
332: $\CC\setminus \{0\}$.)
333:
334: It is clear that for $a\in\NN$ the transseries
335: can be easily calculated for any $z=\rho\exp(2\pi i m/a^j)$, $(m,j)\in\NN, 0\le \rho<1$ since
336: $$f(\rho \exp(2\pi i m/a^j))=\sum_{n=1}^{j} \rho^{a^j}\exp(2\pi i m a^{n-j})+f(\rho)$$
337: where the sum is a polynomial, thus analytic.
338: \end{Proposition}
339:
340:
341:
342: \subsubsection{Borel summable transseries representations; resurgence (cf. \S\ref{A4})}
343:
344:
345:
346:
347: When $c_j\equiv 1$ it is clear that the growth rate as $z\to 0^+$
348: majorizes the rate at any point on $i\RR$. There may be {\bf no} other
349: point with this growth, as is the case when $g(j)=j^b$, $b\in (1,2)$
350: as seen in
351: see Proposition \ref{Pp4} below, or densely many if, for instance,
352: $g(j)=j^b$, $b\in\NN$, or when $g(j)=a^j$, $a\in\NN$. The behavior
353: near points of maximal growth merits special attention.
354:
355:
356:
357: For $g=j^b$, $1<b\ne 2$, $f$ has asymptotic expansions which do not,
358: in general converge. They are however generalized Borel summable.
359:
360:
361: Define $d$ by \begin{equation}
362: \label{eq:eqdu}
363: \frac{1}{b}+\frac{1}{d}=1
364: \end{equation}
365:
366:
367:
368: \begin{Theorem}\label{PP3}{ Let $g=j^b;\,b>1$. Then,
369:
370:
371: (i) The asymptotic series
372: of $f(z)$ for small $z$,
373: \begin{equation}
374: \label{eq:tr1}
375: \tilde{f}_0=\Gamma\left(1+\dfrac{1}{{b}}\right)z^{-\frac{1}{{b}}}-\dfrac{1}{2}+\frac{i}{2\pi}\sum_{j=1}^{\infty}(1-(-1)^{j{b}})\zeta(j{b}+1)b_{j}z^{j}
376: \end{equation}
377: is Borel summable in $X=z^{-1/(b-1)}$,
378: along any ray $\arg(X)=c$ if $c\ne-\arg k^qs_{\pm}, k\in\NN $, where
379: $s_{\pm}=t_{\pm}\mp 2\pi i t_{\pm}^{1/b}$ and $t_{\pm}=( \pm 2\pi i/b)^{b/(b-1)}$. More precisely, (a)
380: \begin{equation}
381: \label{eq:eqfz}
382: f(z)=\Gamma(1+1/b)z^{-1/b}-\frac{1}{2}+ z^{-b/(b-1)}\int_0^{\infty}e^{-z^{-1/(b-1)s}}H(s)ds
383: \end{equation}
384: where $H(s)=H_{b}(s^{b-1})$, where $H_{b}$ is analytic at zero and $H_{b}(0)=0$; (b) $H$ is
385: analytic on the Riemann surface of the log, with square root branch points
386: at all points of the form $k^qs_{\pm}, k\in\NN$ and (c) making appropriate cuts (or working on Riemann surfaces), $u^{-(b-1)^2/b}H$ is bounded at infinity.
387:
388: If $\arg(X)\in(\theta_-,\theta_+)$, then $\mathcal{LB}\tilde{f}_0=f$ \footnote{Note that
389: the variable of Borel summation, or {\em critical time}, is not
390: $1/z$ but $z^{-1/(b-1)}$.} In a general complex direction, $f$ has a nontrivial transseries, see (iii).
391:
392:
393: (iii) For a given direction $\varphi$,
394: $\sigma$ be $\pm 1$ if $\pm \varphi>0$ and $0$ otherwise. If $\sigma \arg z\in (\theta_{\sigma }, \pi/2)$, then the transseries
395: of $f$ is
396:
397: \begin{equation}
398: \tilde{f_0}(z)+\sigma \frac{i}{2\pi}\sum\limits _{k=1}^{\infty}
399: e^{-{s_{{\sigma}}k^{{q}}}{x^{-q+1}}}\sum\limits _{j=0}^{\infty}c_{j}(\sigma k)^{\frac{-2j{b}+2-{b}}{2({b}-1)}}z^{\frac{2j-1}{2({b}-1)}}
400: \end{equation}
401: and it is Borel summable as well.}
402: \end{Theorem}
403:
404:
405:
406:
407:
408: \begin{Note}
409: {\rm The duality $k^b\leftrightarrow k^d$ is the same as in van der Corput
410: formulas; see \cite{Montgomery}.}
411: \end{Note}
412: \subsubsection{Examples} (i)
413: For $b=3$, $f$ the transseries is given by
414: \begin{equation}
415: \tilde{f}_0(z)+\sum_{k\in\ZZ}\sigma e^{-\kappa_3k^{3/2}z^{-1/2}}\left[\left(\dfrac{\pi
416: i k}{6}\right)^{\frac{1}{4}}{z}^{-\frac{1}{4}}+\dfrac{5}{32}\dfrac{(ik)^{\frac{7}{4}}}{6^{\frac{3}{4}}\pi^{\frac{5}{4}}}z^{\frac{1}{4}}+...\right]
417: \end{equation}
418: where with $\kappa_{3}=\pi^{3/2}\sqrt{\frac{32}{27}}(-1)^{\frac{1}{4}}$ and
419: $$ \tilde{f}_0(z)=\frac{\Gamma(4/3)}{z^{1/3}}-\dfrac{1}{2}-\dfrac{z}{120}+\frac{z^{3}}{792}+...$$
420:
421:
422: \z (ii) For $b=3/2$, with $\kappa=32 i\pi^3/27$, $f$ has the Borel summable transseries
423: \begin{equation}
424: \label{eq:3/2f}
425: \tilde{f}_0(z)+ \sigma \sum\limits _{k=1}^{\infty}e^{\kappa \sigma k^{3}z^{-2}}
426: \left[\dfrac{4\sqrt{2}(\sigma i)^{-\frac{1}{2}}\pi}{3}\dfrac{k^{\frac{1}{2}}}{z}+
427: \dfrac{(\sigma i)^{\frac{1}{2}}}{16\sqrt{2}\pi^{\frac{5}{4}}}\dfrac{z^2}{k^{\frac{5}{2}}}+...
428: \right]
429: \end{equation}
430: where $\theta_{\pm}=\pi/4$ and
431: $$\tilde{f}_0(z)=\Gamma\left(\text{\small $\frac{5}{3}$}\right)z^{-\frac{2}{3}}-\dfrac{1}{2}-\dfrac{3\zeta(\frac{5}{2})}{16\pi^{2}}z+\frac{1}{240}z^{2}+\frac{315\zeta(\frac{11}{2})}{2048\pi^{5}}z^{3}+...$$
432:
433:
434:
435: \subsubsection{Properties-preserving extensions beyond the barrier}
436: It is natural to require for an extension beyond the barrier that it
437: has the following properties:
438: \begin{enumerate}
439: \item It reduces to usual analytic continuation when the latter exists.
440:
441: \item It commutes with all properties with which analytic continuation
442: is compatible (principle of preservation of relations, a vaguely
443: stated concept; this requirement is rather open-ended).
444:
445:
446: \end{enumerate}
447: Borel summable series (more generally transseries)
448: or suitable convergent representation representations allow for
449: extension beyond the barrier, as follows. In the case $g(j)=j^b$
450: (and, in fact in others in which $g$ has a convergent or summable
451: expansion at infinity), $f(z)$ can be written, after Borel
452: summation in the form (see
453: \S\ref{AC1}) \begin{equation} \label{eq:eqac} z^{-{d}}\int_0^{
454: \infty}e^{-z^{-1/(b-1)}
455: s}H_1(s)ds=z^{\frac{1}{b-1}-d}\int_0^{ \infty}e^{-t}H_1(t
456: z^{1/(b-1)})dt=:\int_0^{\infty}e^{-s}F(s,z)ds \end{equation} where $H_1$ is analytic near the
457: origin and in $\CC$ except for arrays of isolated singularities
458: along finitely many rays. Furthermore, $H_1$ is polynomially
459: bounded at infinity. This means that the formal series is
460: summable in all but finitely many directions in
461: $z$.
462: \begin{Definition}
463: We define the Borel sum continuation of $f$ through point $z$ on the
464: natural boundary, in the direction $d$, to be
465: the Borel sum of the formal series of $f$ at $z$ in the direction $d$,
466: if the Borel sum exists.
467:
468: This extension simply amounts to analytically
469: continuing $F(s,z)$ in $z$, in (\ref{eq:eqac}), for
470: small $s$ and then analytically continuing
471: $F(s,z)$ in $s$ for fixed $z$.
472: along $\RR^+$.
473: \end{Definition}
474:
475: \subsubsection{Notes} (cf. Appendix \S\ref{Fi}.) \begin{enumerate} \item The Borel sum provides a natural, properties-preserving,
476: extension \cite{OCtransseries}. Borel summation commutes with all common operations
477: such as addition, multiplication, differentiation. Thus,
478: the function and its extensions will have the same
479: properties.
480:
481:
482: \item
483: Also, when a series converges, the Borel sum coincides with the usual
484: sum. Thus, when analytic continuation exists, it coincides with the
485: extension.
486:
487: \item
488:
489: With or without a boundary, the Borel sum of a
490: divergent series changes as the direction of summation crosses the
491: {\em Stokes directions} in $\CC$. Yet, the properties of the family of
492: functions thus obtained are preserved.
493: There may then exist extensions {\em along} the barrier as well. Of course,
494: all this cannot mean that there is analytic continuation across/along
495: the boundary.
496: \end{enumerate}
497:
498:
499:
500: See also Eq. (\ref{convt}), a convergent expansion, where $z$ can also
501: be replaced by $-z$. In this case, due to strong lacunarity, the extension
502: {\em changes} along {\em every} direction, as though there existed densely
503: many Stokes lines.
504:
505: \subsection{Universal behavior near boundary in specific cases}
506: \begin{figure}
507: \centering
508: \includegraphics[scale=0.3]{Riemann1}
509: \caption{The standard function $Q(x)$. The points above $Q=0.05$ are
510: the only ones present in the actual graph.}
511: \label{fig:1}
512:
513: \includegraphics[scale=0.3]{barrn40}
514: \caption{Point-plot graph of $Q_{4,4}$, normalized to
515: one.}
516: \label{fig:1}
517: \end{figure}
518: \begin{figure}
519: \includegraphics[scale=0.4]{fig030}
520: \caption{Point-plot graph of $Q_{3,3}$; $Q_{3,2}$ follows from it
521: through the transformation (\ref{eq:dual}).
522: }
523: \label{fig:1}
524:
525:
526: \end{figure}
527:
528:
529: In many cases, $\phi(g^{-1}(u/x))$ has an asymptotic expansion, and then
530: $\Phi$ in turn has an asymptotic expansion in $x$.
531: Detailed behavior along the boundary can be obtained in special cases such as $g(j)=j^b;\,b\in\NN$,
532: or $g(j)=j^{(b+1)/b};\,b\in\NN$. Properly scaled sums converge to everywhere discontinuous
533: functions.
534: \begin{Theorem}\label{Pp4}
535: (i) If $g(j)=j^b$, $b\in\NN$, we have $\displaystyle \limsup
536: x^{1/b} |f(x +iy)|<M<\infty$. We let
537: $Q_{b,d}(s)=\displaystyle \lim_{x\to
538: 0^+}x^{1/d}f(x+2\pi i s)$ (whenever it exists); then,
539:
540: \begin{multline}
541: \label{eq:bd2}
542: \frac{Q_{b,b}(s)}{\Gamma(1+b^{-1})}=\begin{cases} \displaystyle \,\,\frac{1}{n}\sum_{l=1}^{n}e^{-2\pi
543: i\dfrac{m}{n}l^{b}}, \ s=\frac{m}{n}, \ m,n\in \NN\ \text{relatively prime} \\ \\
544: 0\ \ \ a.e. \end{cases}
545: \end{multline}
546:
547: (ii) For $b=3/2$ and $z\to 0^+$, $f$ grows like $z^{-2/3}$. For any other point
548: on $i\RR$ the growth is slower, at most $(\Re\, z)^{-1/2}$. Furthermore,
549:
550:
551: \begin{equation}\label{case3/2}
552: \frac{Q_{3/2,1/2}(\sqrt{s})}{\sqrt{6\pi i}s^{1/4}}=\begin{cases}\displaystyle\frac{1}{n}\sum_{l=1}^{n}e^{-2\pi
553: i\dfrac{m}{n}l^{3}},\ \ s =\dfrac{16}{27}\frac{n}{m}, \ n,m\text{ as in (i)}\\ \\
554: 0\ \ a.e.\end{cases}
555: \end{equation}
556: In particular, we have the profile duality relation
557: \begin{equation}
558: \label{eq:dual}
559: Q_{3/2,1/2}(s)=\frac{\sqrt{6\pi i}s^{1/2}}{\Gamma(4/3)}Q_{3,3}\(2^83^{-6}s^{-2}\)
560: \end{equation}
561: for any $s\in\RR$ for which $Q_{3/2}(s)$ and/or
562: $Q_{3,3}(s)$ is well defined, for instance, in the cases given in
563: (\ref{eq:bd2}) and (\ref{case3/2}).
564: \end{Theorem}
565: For large $n$, the sum over $l$ in Eq. (\ref{eq:bd2}) is,
566: statistically, expected to be of order one. After $x^{\frac{1}{d}}$
567: rescaling, the template behavior ``in the bulk'' is given by the
568: familiar function $Q(x)=n^{-1}$ if $x=m/n, (m,n)\in\NN^2$ and zero
569: otherwise, shown in Fig. \ref{fig:1}.
570: \subsection{Fourier series of the B\"otcher map and structure of Julia sets}\label{holom}
571:
572: We show how lacunary series are building blocks for fractal structures
573: appearing in holomorphic dynamics (of the vast literature on holomorphic
574: dynamics we refer here in particular to \cite{Beardon}, \cite{Devaney} and \cite{Milnor}).
575: In \S\ref{S5} we mention a few known facts about polynomial
576: maps. Consider for simplicity the
577: quadratic map
578: \begin{equation}
579: \label{eq:logist1}
580: x_{n+1}=\lambda x_n(1-x_n)
581: \end{equation}
582: It will be apparent from the proof that the results and method extend easily
583: to polynomial iterations of the form
584: \begin{equation}
585: \label{eq:eqpol1}
586: x_{n+1}=\lambda P_k(x_n)
587: \end{equation}
588: with $\lambda$ relatively small.
589: The substitution $x=-(\lambda y)^{-1}$ transforms (\ref{eq:logist1}) into
590: \begin{equation}
591: \label{eq:infty1}
592: y_{n+1}=\frac{y_n^2}{\lambda(1+ y_n)}=f(y_n)
593: \end{equation}
594:
595: By B\"otcher's theorem (we give a self contained proof in
596: \S\ref{S5}, for (\ref{eq:logist1}), which extends in fact to the
597: general case), there exists a map $\phi$,\footnote{The notations
598: $\phi,\psi, H$, designate different objects than those in previous
599: section.} analytic near zero, with $\phi(0)=0$,
600: $\phi'(0)=\lambda^{-1}$ so that $(\phi f \phi^{-1})(z)=z^2$. Its
601: inverse, $\psi$, conjugates (\ref{eq:infty1}) to the canonical map
602: $z_{n+1}=z_n^2$, and it can be checked that
603: \begin{equation}
604: \label{eq:eqG}
605: \psi(z)^2=\lambda \psi(z^2)(1+\psi(z));\ \ \psi(0)=0,\ \psi'(0)=\lambda
606: \end{equation}
607: Let $\mathcal A(\mathbb{D})$ denote the Banach space of functions
608: analytic in the unit disk $\mathbb{D}$ and continuous in $\overline{\mathbb{D}}$, with the sup norm. We define the linear operator $\mathfrak T=\mathfrak T_2$, on $\mathcal A(\mathbb{D})$ by
609: \begin{equation}
610: \label{eq:defT}
611: (\mathfrak T f)(z)=\frac{1}{2}\sum_{k=0}^{\infty}2^{-k}f(z^{2^k})
612: \end{equation}
613: This is the inverse of the operator $f\mapsto 2f-f^{\vee 2}$,
614: where $f^{\vee p}(z)=f(z^p)$.
615: Clearly, $\mathfrak T f$ is an isometry on $\mathcal A(\mathbb{D})$
616: and it maps simple functions, such as generic polynomials, to
617: functions having $\partial \mathbb{D}$ as a natural boundary; it reproduces $f$
618: across vanishingly small scales.
619:
620:
621: \bigskip
622:
623: \begin{Theorem}\label{T1}
624:
625: (i) $\psi$ and $H=1/\psi$ are analytic in ($\lambda,z$) in $\mathbb{D}\times\mathbb{D}$ ($\lambda\in \mathbb{D}$ corresponds to the main cardioid
626: in the Mandelbrot set, see \S\ref{S5}), and continuous in $\mathbb{D}\times\overline{\mathbb{D}}$. The series
627: \begin{equation}
628: \label{eq:formG}
629: \psi(\lambda,z)=\lambda z+\sum_{k=1}^{\infty}z\lambda^{k+1} \psi_k(z)
630: \end{equation}
631: converges in $\mathbb{D}\times\overline{\mathbb{D}}$ (and so does the series
632: of $H$), but not in $\overline{\mathbb{D}}\times\overline{\mathbb{D}}$. Here
633: \begin{equation}
634: \label{eq:formPhi1}
635: \psi_1(z)=\mathfrak T z=\sum_{k=1}^{\infty}2^{-k+1}z^{2^k}
636: \end{equation}
637: (note that $\psi(z)=\frac{1}{2}z+\frac{1}{2}\int_0^zs^{-1}h(s)ds$ with $h$ given in (\ref{eq03})), and in general
638: \begin{equation}
639: \label{eq:frakt}
640: \psi_k =\mathfrak{T}\left(z\sum_{j=0}^{k-1}\psi_j^{\vee 2}\psi_{k-1-j}-\sum_{j=1}^{k-1}\psi_j\psi_{k-j}\right)
641: \end{equation}
642:
643: (ii) All $\psi_k, k\ge 1$ have binarily lacunary series: In $\psi_k$,
644: the coefficient of $z^p$ is nonzero only if $p$ has at most $k$ binary-digits equal to $1$, i.e.,
645: \begin{equation}
646: \label{eq:pk}
647: p=2^{j_1}+\cdots+2^{j_k},\ \ j_i=0,1,2,... \ (\text{``$k$''is the same as in $\psi_k$})
648: \end{equation}
649: (iii) For $|\lambda|<1$, the Julia curve of (\ref{eq:logist1}) is given
650: by a uniformly convergent Fourier series, by (i),
651: \begin{equation}
652: \label{eq:eqJ}
653: J=\left\{-\left(\Re\, H(e^{it}),\Im\, H(e^{it})\right): t\in [0,2\pi)\right\}
654: \end{equation}
655: \end{Theorem}
656: \begin{Remark}{\rm
657: The effective lacunarity of the Fourier series makes calculations of the
658: Julia set numerically effective if $\lambda$ is not too large.}
659: \end{Remark}
660:
661: \begin{figure}
662: \centering
663: \includegraphics[scale=0.3]{lambda3}\includegraphics[scale=0.3]{lambda3i}
664: \caption{The Julia set of $x_{n+1}=\lambda x_n(1-x_n)$, for $\lambda=0.3$ and
665: $\lambda=0.3i$,
666: calculated from the Fourier series (\ref{eq:formG}) discarding all $o(\lambda^2)$ terms. They coincide, within
667: plot precision, with numerically calculated ones using standard iteration of maps
668: algorithms.
669: }
670: \label{fig:3}
671:
672: \end{figure}
673:
674: \begin{Note} \label{N4} $ $ \hskip -1cm {\rm
675: \begin{enumerate}
676: \item Lacunarity of $\psi_k$ is a strong
677: indication that $\psi$ has a natural boundary (in fact, it does have one),
678: but not a proof. Transseriation at {\em singular} points and
679: summation of convergent series (here in the parameter $\lambda$)
680: {\bf do not commute}. The transseries of $H$ on the barrier can
681: be calculated by transasymptotic matching, see \cite{Inventiones},
682: but in this case is more simply found directly from the functional
683: relation; see Note \ref{Explc}.
684: \item In assessing the fine structure
685: of the fractal using the Fourier expansion truncated to $o(\lambda^n)$, the scale of analysis
686: cannot evidently go below $O(\lambda^n)$.
687: \end{enumerate}
688: }
689: \end{Note}
690:
691:
692: \begin{Remark}{\rm
693:
694: \begin{enumerate}
695:
696:
697: \item For $|\lambda|$ sufficiently small,
698: Theorem~\ref{T1} provides a convenient way to determine
699: the Julia set as well as the discrete evolution on the boundary.
700:
701:
702:
703: \item For small $\lambda$, the self similar structure
704: is seen in
705: $$\psi_1(\rho \exp(2\pi i m/2^n))=
706: \sum_{k=1}^{n-1}\frac{ \rho^{2^j}
707: \exp(2\pi i m 2^{k-n})}{2^j}+\frac{ \rho^{2^n-2}}{2^{n-1}}\psi_1(\rho)$$
708: where the sum is a polynomial, thus analytic. Up to a scale factor
709: of $2^{-n+1}$, if $|\lambda|\ll 2^{-n+1}$, the nontrivial structure of $\psi$
710: at $\exp(2\pi i m/2^n)$ and at $1$ are the same, see Note \ref{N4}; that is
711: $$\psi(\exp(2\pi i m/2^n))={2^{-n+1}}\psi(1)+\text{regular}+o(\lambda)$$
712: Exact transseries can be obtained for $\psi$; see also Note~\ref {Explc}.
713:
714: \item For iterations of the form
715: $x_{n+1}=\lambda^{k-1}P_{k}(x)$ where $P_k$ is a polynomial of degree $k>2$, the calculations and the results, for small
716: $\lambda$, are essentially the same. The lacunary series
717: would involve the powers $z^{k^j}$. For instance if the
718: recurrence is $x_{n+1}=\lambda^2x_n(1-3x_n+x_n^2)$, then $\psi$
719: is to be replaced by the solution of
720: $$\psi=\lambda^2\psi^{\vee 3}(1+3\psi+\psi^2)$$
721: and the small $\lambda$ series will now have
722: $\psi_1=\sum_{k=1}^{\infty}3^{-k-1}{{z^{3^k}}}$ and so on.
723: \end{enumerate}
724: }
725: \end{Remark}
726:
727:
728:
729: \section{Proofs}
730: \begin{Lemma}\label{linv}
731: Under the assumptions of Theorem~\ref{P1}, (iii), $h(y):=g^{-1}(y)$ also has a
732: differentiable asymptotic power series as $y\to\infty$.
733: \end{Lemma}
734: \begin{proof}
735: Straightforward inversion of power series asymptotics, cf \S\ref{appendix}.
736: \end{proof}
737:
738: \subsection{Proof of Theorem \ref{P1}}
739: The proof of (\ref{eq:asbh}) (i) essentially amounts to showing that
740: $|f|^2$ is diagonally dominant, in that terms containing $g(j)$ and $g(k)$
741: with $j\ne k$ are comparatively small, as shown in \S\ref{genblowup}.
742:
743: (ii) Since $g$ is increasing on
744: $(0,\infty)$, the result follows from the usual
745: integral upper and lower bounds for a sum. Equation (\ref{eq:asympt1})
746: follows from simple calculations, cf. \S\ref{appendix}.
747:
748: \subsection{Proof of Theorem \ref{Prop2}}
749: We prove part (ii); part (i) is similar, and simpler. By standard Fourier analysis we get
750: \begin{equation}
751: \label{eq:eq21}
752: \{u\}=\frac{1}{2}-\sum_{k=1}^{\infty}\frac{\sin2k\pi u}{k\pi}, \ \ \ \ u\notin\mathbb{Z}\end{equation}
753: where
754: \begin{equation}
755: \label{eq:eqM}
756: \left\|\{u\}-\frac{1}{2}+\sum_{k=1}^{M}\frac{\sin2k\pi u}{k\pi}\right\|_{\infty}\le C_M \to {\text{\rm Si}}(\pi) \text{ as } M\to \infty
757: \end{equation}
758: where Si is the sine integral, and Si($\pi$) is the constant in the Gibbs phenomenon.
759: Let $g_m$ be an analytic function such that
760: $g(s)-g_m(s)=o(s^{-m})$ for large $s$.
761: \begin{Lemma}
762: The analysis reduces to the case where $h$ is a finite sum of powers. Indeed,
763: \begin{equation}
764: \label{eq:eqf}
765: f=e^{-xg(0)}+\int_0^{\infty} e^{-xu}g^{-1}(u)du+\int_{g(0)}^{\infty}e^{-xu} \{h_m(u)\}du+R_{m-1}(x)
766: \end{equation}
767: where $R_{m-1}$ is $C^{m-1}$ and $h_m$ is a truncation of the asymptotic
768: expansion of $h$, such that $h(u)-h_m(u)=o(u^{-m})$.
769: \end{Lemma}
770: \begin{proof}
771: We have
772:
773: \begin{multline}
774: \label{eq:sum2}
775: \int_0^{\infty} e^{-xu} \{h(u)\} du=
776: \sum_{N=0}^{\infty}\int_{g(N)}^{g(N+1)} e^{-xu}(h(u)-N) du\\=
777: \sum_{N=0}^{\infty}\int_{g_m(N)}^{g_m(N+1)} e^{-xu}(h_m(u)-N)du
778: +R_{m-1}(x)
779: \end{multline}
780: where
781: \begin{multline}
782: \label{eq:eqR}
783: R_{m-1}(x)=\sum_{N=0}^{\infty}\left(\int_{g(N)}^{g_m(N)}+\int_{g_m(N+1)}^{g(N+1)}\right)
784: e^{-xu} \{h(u)\} du\\+\sum_{N=0}^{\infty}\int_{g_m(N)}^{g_m(N+1)}e^{-xu}(h(u)-h_m(u))du
785: \end{multline}
786: Using (\ref{eq:eq21}) and (\ref{eq:eqM}) the proof follows and the
787: fact that $g(N)-g_m(N)=o(N^{-m})$ and $h(u)-h_m(u)=o(u^{-m})$, the sum is rapidly convergent, and the result follows. \end{proof}
788: \begin{Lemma}
789: If $h$ is a finite sum of powers, then $f-e^{-xg(0)}-\int_0^{\infty} e^{-xu}g^{-1}(u)du\in C^\infty$.
790: \end{Lemma}
791: \begin{proof}
792:
793: Using (\ref{eq:eq21}) and (\ref{eq:eqM}) we have
794: \begin{multline}
795: \label{eq:eqhm}
796: f_1(x)= x\int_0^{\infty}e^{-xu} \{h_m(u)\}du=\frac{1}{2}-
797: \sum_{k=1}^{\infty}\int_0^{\infty}e^{-xu}\frac{\sin 2k\pi h_m(u)}{k\pi}du\\
798: =\frac{1}{2}-
799: \sum_{k=1}^{\infty}\frac{1}{2k\pi i}\sum_{\sigma=\pm 1}\sigma \int_0^{\infty}e^{-xu+\sigma 2k\pi i h_m(u)}du
800: \end{multline}
801: We deform the contours of integration along the directions
802: $\sigma\alpha$ respectively, say for $\alpha=\pi/2$. The integral
803: \begin{equation}
804: \label{eq:vert}
805: f_1= \int_0^{i\infty}u^n e^{2k\pi i h_m(u)}du
806: \end{equation}
807: exists for any $n$ and it is estimated by
808: \begin{equation}
809: \label{eq:estm3}
810: \left|\int_0^{i\infty}u^n e^{2k\pi i h_m(u)}du\right|<const. k^{-b(n+1)}\Gamma(n(b+1))
811: \end{equation}
812: The termwise nth derivatives at $0^+$ of the series of $f_1(x)$ converge rapidly,
813: and the result follows.
814: \end{proof}
815:
816: \subsection{Proof of Theorem \ref{Cexp}}
817: \begin{proof}
818: Let $\epsilon>0$ be arbitrary, let
819: $$S_N=\sum_{j=1}^N e^{i y g(j)};\ \ \ S_0:=0$$
820: and let $N_1$ be large enough so that $|S_N/\rho(N)-L|\le \epsilon$ for $N>N_1$. Then, by looking at $e^{i\phi} f$ if necessary, we can assume that $L\ge 0$. We have
821: \begin{multline}
822: \label{eq:eqf3}
823: f(x+iy)=\sum_{k=1}^{\infty}e^{-xg(k)}(S_{k}-S_{k-1})=\sum_{k=1}^{\infty}\frac{S_k}{\rho(k)}(e^{-xg(k)}-e^{-xg(k+1)})\rho(k)\\=L(y)\sum_{k=1}^{\infty}(e^{-xg(k-1)}-e^{-xg(k)}){\rho(k)}+\sum_{k=1}^{\infty}(e^{-xg(k)}-e^{-xg(k+1)})d_k{\rho(k)}
824: \end{multline}
825: where $d_k=L(y)-S_k/\rho(k) \to0$ as $k\to\infty$. Now,
826: \begin{equation}
827: \label{eq:67}
828: \sum_{k=1}^{\infty}e^{-xg(k)}(\rho(k+1)-\rho(k)) \sim \int_0^{\infty}e^{-xg(k)}\rho'(k)dk =:\Phi_{\rho}(x).
829: \end{equation}
830: and under the given assumptions $\Phi_{\rho}\to \infty$ as $x\to 0$.
831: Note that $\Phi_{\rho}\to \infty$ as $x\to 0$ and the facts that $e^{-xg(k)}-e^{-xg(k+1)}>0$, $\rho(k)>0$ and $d_k\to 0$, readily imply that the last sum in (\ref{eq:eqf3}) is $o(\Phi_{\rho}(x))$; (\ref{eq:lsup}) is follows in a similar way.
832: \end{proof}
833:
834: \subsection{Proof of Theorem\ref{Pp4}}
835: We rely on Theorem \ref{Cexp}, and analyze the case $b=3/2$; the
836: case $b\in\NN$ is simpler. We have $\beta=t/(2\pi)$. Let
837: $\check{S}_j=\sum \limits_{k=1}^{j}e^{-2\pi i \dfrac{m}{n} k^3}\!\!.$ It
838: is clear that for $m,n\in\NN$ we have $j^{-1}\check{S}_j\to \sum
839: \limits_{k=1}^{n}e^{-2\pi i \dfrac{m}{n} k^3}=L_{mn}$. On the other
840: hand, by summation by parts we get
841:
842: \begin{multline}
843: \label{eq:eq01}
844: S_N= \frac{2\sqrt{2}e^{\pi i/4}}{3\beta}\sum_{k=1}^{k_N-1} k^{1/2}
845: \exp\left(-\frac{8\pi i}{27\beta^2}k^3\right)+O(N^{1/4})=O(N^{1/4})\\+
846: \frac{2\sqrt{2}e^{\pi i/4}}{3\beta}(k_N-1)^{3/2}\frac{\check{S}_{k_N-1}}{k_N-1}
847: +\frac{2\sqrt{2}e^{\pi i/4}}{3\beta}\sum_{k=1}^{k_N-1}k^{-1}\check{S}_k k(k^{1/2}-(k-1)^{1/2})
848: \end{multline}
849: and it is easy to check that, as $N\to\infty$ we have
850: \begin{equation}
851: \label{eq:sum}
852: S_N\sim \left(\frac{2\sqrt{2}e^{\pi i/4}}{3\beta}(1+1/3)\right) k_N^{3/2} L_{mn}=\frac{8\sqrt{2}e^{\pi i/4}}{9\beta}\sqrt{\frac{3\beta}{2}}N^{3/4}
853: \end{equation}
854: where we used the definition of $k_N$ following
855: eq. (\ref{eq:eq01}), which implies $k_N\sim \frac{3}{2}N^{1/2}\beta$. The result follows by changes of variables, using
856: (\ref{eq:lim2}) and noting that
857: \begin{equation}
858: \label{eq:int4}
859: \frac{3}{4} \int_0^{\infty}e^{-x k^{3/2}}k^{-1/4}dk=\frac{\sqrt{\pi}}{2\sqrt{x}}
860: \end{equation}
861:
862:
863: It is also clear that $|j^{-1}\check{S}_j|\le 1$ and a similar calculation
864: provides an overall upper bound.
865:
866:
867:
868: Section \ref{Direct,3} provides an independent way to calculate the behavior
869: along the boundary.
870: \subsection{Proof of Proposition \ref{b=2}}
871: For $b=2$ using (\ref{eq:eqhm}) and
872: $\int_{0}^{\infty}e^{-px}\sin2k\pi
873: p^{\frac{1}{2}}dp=\pi^{3/2}ke^{-\frac{k^{2}\pi^{2}}{x}}x^{-3/2}$
874: we immediately obtain
875:
876: \[
877: f(x)=\frac{\sqrt{\pi}}{2\sqrt{x}}-\dfrac{1}{2}+\frac{\sqrt{\pi}}{\sqrt{x}}\sum_{k=1}^{\infty}e^{-\frac{k^{2}\pi^{2}}{x}}\]
878:
879: \subsection{The case $g(j)=a^j$}
880:
881:
882:
883:
884: It is easy to see that, for this choice of $g$, we have the functional
885: relation $$f(z)-f(az)=e^{-z}$$
886:
887: Since $e^{-z}\to 1$ as $z\rightarrow0+$, the leading behavior formally
888: satisfies $f(z)-f(az)\sim1$ i.e. $f(z)\sim-\dfrac{\log z}{\log a}$
889:
890: In view of this we let $f(z)=-\dfrac{\log z}{\log a}+G(z)$ which
891: gives $G(z)-G(az)=e^{-z}-1$ {\em i.e.}
892:
893: \begin{equation}
894: \label{eq:eqg}
895: G(z)=G(z/a)+(1-e^{-z/a})
896: \end{equation}
897: We first obtain a solution a solution $\check{G}$ of the homogeneous equation and then
898: write $G=\check{G}+h(a^z)$, where $h$ now satisfies
899: \begin{equation}
900: \label{eq:eqhy}
901: {h}(y)={h}(y+1)
902: \end{equation}
903: Iterating (\ref{eq:eqg}) we obtain
904: \[
905: \check{G}(z)=\sum_{n=0}^{\infty}(1-e^{-z/a^n})=-\sum_{n=0}^{\infty}\sum_{k=1}^{\infty}\frac{(-z)^{k}}{k!a^{nk}}=\sum_{n=1}^{\infty}\frac{(-z)^{n}}{n!(1-a^{n})}\]
906: which is indeed an entire function and satisfies the functional
907: relation. Now we return to (\ref{eq:eqhy}). Since by its connection to
908: $f(z)$, ${h}$ is obviously smooth in $y$, it can be expressed in
909: terms of its Fourier series
910:
911: \[
912: {h}(y)=\sum_{k=-\infty}^{\infty}c_{k}e^{2k\pi iy}\] where the
913: coefficients $c_{k}=\bar{c}_{-k}$ can be found by using the original
914: function $f$. Recall that we have $$f(z)=-\dfrac{\log z}{\log
915: a}+\check{G}(z)+{h}\(\dfrac{\log z}{\log a}\)$$ which
916: implies
917:
918: \[
919: {h}(y)=f(a^{y})+y-\check{G}(a^{y})\]
920: \begin{Lemma}\label{Lgeom}
921: The Fourier coefficients of the periodic function ${h}$ are given
922: by
923: \begin{equation}
924: \label{eq:ck}
925: c_{k}=\frac{1}{\log a}\Gamma\(-\frac{2k\pi i}{\log a}\)\:\(k\neq0\)
926: \end{equation}
927: and
928: $$c_{0}=\int_{0}^{1}f(a^{y})dy+\int_{0}^{1}ydy-\int_{0}^{1}\check{G}(a^{y})dy$$
929: Since $c_{k}\sim\sqrt{\dfrac{1}{\pi\log a}}e^{\frac{-|k|\pi^{2}}{\log a}}$,
930: (implying that the Fourier expansion for $h$ is valid exactly
931: for $\Re\(z\)>0$) we have
932: \[
933: f\(z\)=-\dfrac{\log z}{\log a}+\sum_{n=1}^{\infty}\frac{\(-z\)^{n}}{n!\(1-a^{n}\)}+c_{0}+\frac{1}{\log a}\sum_{k\neq0}\Gamma\(-\frac{2k\pi i}{\log a}\)z^{\frac{2k\pi i}{\log a}}\]
934: The series further resums to
935: \begin{equation}
936: \label{eq:resum}
937: f\(z\)=-\dfrac{\log z}{\log
938: a}+\sum_{n=1}^{\infty}\frac{\(-z\)^{n}}{n!\(1-a^{n}\)}+c_{0}-\frac{z}{2\pi
939: i}\int_{0}^{\infty}\log_{\RR}\(-s^{\frac{2\pi i}{\log a}}\)e^{-sz}ds
940: \end{equation}
941: valid for $z>0$.
942: \end{Lemma}
943: The proof is given in \S\ref{pfLgeom}.
944:
945:
946: Similar results hold for other rational angles if $\bb\in\NN$ (by
947: grouping terms with the same phase). For example,
948: \begin{multline}
949: \sum_{n=0}^{\infty}e^{-2^{n}(z+\frac{2}{5}\pi i)}\\=e^{-\frac{2}{5}\pi
950: i}\sum_{n=0}^{\infty}e^{-2^{4n}z}+e^{-\frac{4}{5}\pi
951: i}\sum_{n=0}^{\infty}e^{-2^{4n+1}z}+e^{-\frac{8}{5}\pi
952: i}\sum_{n=0}^{\infty}e^{-2^{4n+2}z}+e^{-\frac{6}{5}\pi
953: i}\sum_{n=0}^{\infty}e^{-2^{4n+3}z}\\
954: =e^{-\frac{2}{5}\pi
955: i}\sum_{n=0}^{\infty}e^{-16^{n}z}+e^{-\frac{4}{5}\pi
956: i}\sum_{n=0}^{\infty}e^{-16^{n}(2z)}+e^{-\frac{8}{5}\pi
957: i}\sum_{n=0}^{\infty}e^{-16^{n}(4z)}+e^{-\frac{6}{5}\pi
958: i}\sum_{n=0}^{\infty}e^{-16^{n}(8z)}
959: \end{multline}
960: \subsection{Proof of Theorem~\ref{PP3}}\label{AC1}
961: (i) and (ii):
962: By an argument similar to the one leading to (\ref{eq:eqhm}) we have
963: \begin{equation}
964: \label{eq:dec3}
965: f(z)=\int_0^{\infty} e^{-zg(s)} ds-\frac{1}{2}+
966: \sum_{k=1}^{\infty}\frac{1}{2k\pi i}\sum_{\sigma= \pm 1}\int_0^{\sigma i\infty}e^{-zu+\sigma 2k\pi i u^{1/b}}du
967: \end{equation}
968: The exponential term ensures convergence in $k$.
969: Taking $s^b=t$ we see that
970: \begin{equation}
971: \label{eq:eqt1}
972: \int_0^{\infty} e^{-z s^b}dp=\Gamma(1+1/b)z^{-1/b}
973: \end{equation}
974: We now analyze the case $\sigma=-1$, the other case being similar. A term in the sum is
975: \begin{multline}
976: \label{eq:trm1}
977: \frac{1}{2k\pi i}\int_0^{ i\infty}e^{-zu-2k\pi i u^{1/b}}du=
978: \left(\frac{k}{z}\right)^{{d}}\int_0^{
979: \infty}e^{-\nu_k(t+2\pi i t^{1/b})}dt\\=\left(\frac{k}{z}\right)^{{d}}
980: \int_0^{
981: \infty}e^{-\nu_k s}{\check{H}_{-}}(s)ds=z^{-{d}}\int_0^{
982: \infty}e^{-z^{-1/(b-1)} s}{\check{H}_{-}}(s/k^{{d}})ds
983: \end{multline}
984: where $\nu_k={k^{{d}}}{z^{-1/(b-1)}}$, ${\check{H}_{-}}(s)=\phi'(s)$ and
985: \begin{equation}
986: \label{eq:eqphi}
987: \Phi(\phi(u)):= \phi(u)+2\pi i \phi(u)^{1/b}=u
988: \end{equation}
989: (a) Near the origin, we write $\phi=u^b H_{b}$, and we get
990: $$H_{b}=\frac{1}{2\pi i}(1-u^{b-1}H_{b})^b$$
991: and analyticity of $H_{b}$ in $u^{b-1}$ follows, for instance, from the contractive
992: mapping principle.
993:
994: (b) The only singularities of $\phi$, thus of ${\check{H}_{-}}$, are the points implicitly given by $\Phi'=0$.
995:
996:
997: (b) It is also easy to check that for some $C>0$ we have
998: \begin{equation}
999: \label{eq:eqH1}
1000: |{\check{H}_{-}}(u)|\le C\frac{|u^{b-1}|}{1+|u^{b-1}|}
1001: \end{equation}
1002: uniformly in $\CC$, with a
1003: cut at the singularity, or on a corresponding Riemann surface.
1004: From (\ref{eq:eqH1}) we see that the sum
1005: \begin{equation}
1006: \label{eq:eqsum}
1007: H_{-}=\sum_{k=1}^{\infty} {\check{H}_{-}}(s/k^{{q}})
1008: \end{equation}
1009: converges, on compact sets in $s$, at least as fast as $const\sum k^{-b}$,
1010: thus it is an analytic function wherever {\em all} ${\check{H}_{-}}(s/k^{{d}})$ are
1011: analytic, that is, in $\CC$ except for the
1012: points $k^d s_0$. Using an integral estimate, we get the
1013: global bound
1014: \begin{equation}
1015: \label{eq:eqinfty}
1016: |H_{-}|\le const \sum_{k=1}^{\infty} \frac{|u|^{b-1}}{k^q+|u|^{b-1}}\le const |u|^{(b-1)^{2}/b}
1017: \end{equation}
1018: as $|u|\to \infty$.
1019:
1020: The function $H$ in the lemma is simply $H_{-}+H_{-}$ where
1021: $H_+$ is obtained from $H_{-}$ by replacing $i$ by $-i$. The calculation
1022: of the explicit power series is straightforward, from (\ref{eq:eqphi}),
1023: (\ref{eq:eqsum}) and the similar formulae for $H_+$, using dominated
1024: convergence based on (\ref{eq:eqinfty}). We provide the details for convenience.
1025:
1026: We write the last term in (\ref{eq:trm1}) in the form
1027: \begin{equation}
1028: \label{eq:deffk}
1029: f_{k}(z)=\(\frac{k}{z}\)^{\frac{b}{b-1}}\intop_{C}e^{-s(\frac{k^{b}}{z})^{\frac{1}{b-1}}}\check{H}_{-}ds
1030: \end{equation}
1031: where the contour $C$ is a curve from the origin to $\infty$ in the
1032: first quadrant.
1033:
1034:
1035:
1036: Watson's lemma implies \[
1037: f_{k}(z)=\sum_{j=1}^{N}b_{j}k^{-jb}z^{j-1}+R_{N}(k,z)\] where
1038: $b_{j}$can be calculated explicitly from Lagrange-B\"urmann inversion
1039: formula used for the inverse function $\phi^{-1}$ and $R_{N}(k,z)\leqq
1040: Ck^{-(N+1)b}z^{N}$, for arbitrary $N\in\mathbb{N}$. It follows that
1041: \begin{equation}
1042: \label{eq:srjtob}
1043: f(z)\sim\Gamma(1+\dfrac{1}{b})z^{-\frac{1}{b}}-\dfrac{1}{2}+\frac{i}{2\pi}\sum_{j=1}^{\infty}(1-(-1)^{jb})\zeta(jb+1)b_{j}z^{j}
1044: \end{equation}
1045: which holds for $z\rightarrow0$ in the right half plane in any
1046: direction not tangential to the imaginary axis.
1047:
1048: (iii) We obtain the transseries (which gives us information
1049: near the imaginary axis) by using the global properties of
1050: $\check{H}_{-}$, and standard deformation of the Laplace contour.
1051:
1052: As $z$ goes around the complex plane, as usual in Laplace-like
1053: integrals, we rotate $s$ in (\ref{eq:deffk}) simultaneously,
1054: to keep the exponent real and positive. In the process,
1055: as we cross singularities, we collect a
1056: contribution to the integral from the point $s_{-}$ above; the contribution
1057: is an integral around a cut originating at $s_{-}$. The singularity is integrable, and
1058: collapsing the contour to the cut itself, we get a
1059: contribution again in the form of a Laplace transform.
1060: This is the Borel sum (in the same variable, $z^{-1/(b-1)}$.
1061:
1062: Generically $s_{-}$ is a
1063: square root branch point and we have
1064: \[
1065: \frac{ds_{-}}{ds}=\sum_{j=0}^{\infty}\tilde{c}_{j}\left[s-\left(\dfrac{b
1066: i}{2\pi}\right)^{\frac{b}{1-b}}\right]^{\frac{j-1}{2}}\]
1067: The asymptotic expansion of the cut contribution, Borel summable
1068: as we mentioned, is
1069: \[
1070: \exp \left({-(2\pi)^{\frac{b}{b-1}}(b^{\frac{b}{1-b}}-b^{\frac{1}{1-b}})i^{\frac{b}{1-b}}\(\frac{k^{b}}{z}\)^{\frac{1}{b-1}}}\right)\sum_{j=0}^{\infty}c_{j}\(\frac{z}{k^{b}}\)^{\frac{2j+1}{2(b-1)}}\]
1071: The exponential term ensures convergence in $k$ of the Borel summed transseries. A similar result
1072: can be obtained for $-k.$
1073:
1074: Thus, the transseries of $f$ is of the form
1075:
1076: \[
1077: \tilde{f}(z)=\Gamma(1+\dfrac{1}{b})z^{-\frac{1}{b}}-\dfrac{1}{2}+\frac{i}{2\pi}\sum_{j=1}^{\infty}(1-(-1)^{jb})\zeta(jb+1)b_{j}z^{j}\]
1078:
1079:
1080: \[
1081: +\begin{cases}
1082: \frac{i}{2\pi}\sum\limits _{k=1}^{\infty}e^{(b-1)(\frac{2\pi}{b i})^{\frac{b}{b-1}}(\frac{k^{b}}{z})^{\frac{1}{b-1}}}\sum\limits _{j=0}^{\infty}c_{j}k^{\frac{-2jb+2-b}{2(b-1)}}z^{\frac{2j-1}{2(b-1)}} & \,-\frac{\pi}{2}\leqslant\arg z\leqslant\theta_{1}\\
1083: -\frac{i}{2\pi}\sum\limits _{k=1}^{\infty}e^{(b-1)(\frac{2\pi}{b
1084: i})^{\frac{b}{b-1}}(\frac{(-k)^{b}}{z})^{\frac{1}{b-1}}}\sum\limits
1085: _{j=0}^{\infty}c_{j}(-k)^{\frac{-2jb+2-b}{2(b-1)}}z^{\frac{2j-1}{2(b-1)}}
1086: & \,\theta_{2}\leqslant\arg z\leqslant\frac{\pi}{2}\end{cases}\]
1087:
1088: \begin{Remark}
1089: The analysis can be extended to the case
1090:
1091: \[
1092: f(z)=\sum_{k=1}^{\infty}F(k)e^{-k^{\bb}z}\:(\lambda\neq0)\]
1093: by noticing that
1094:
1095: \[
1096: f(z)=\int_{0}^{\infty}(zF(p)-F'(p))e^{-pz}[p^{\frac{1}{\bb}}]dp\] If
1097: the expression of $F(k)$ is simple, for example $F(k)$ is a finite
1098: combination of terms of the form
1099: $\mu^{k^{\frac{1}{\bb}}}k^{\lambda}(\log k)^{m}\:(m=0,1,2,3...)$,
1100: the method in this section applies with little change to calculate the
1101: transseries of $f(z)$.
1102: \end{Remark}
1103: For special values of $\bb$, asymptotic information as $z$ approaches
1104: the imaginary line can be obtained in the following way. Let
1105: $z=\delta+2\pi i\beta$ and
1106: \[
1107: f(z)=\sum_{k=1}^{\infty}e^{-k^{\bb}(\delta+2\pi i\beta)}\:(\Re(\delta)>0)\]
1108: If $1<\bb\in\NN$, we
1109: may obtain the asymptotic behavior for all rational $\beta=\dfrac{m}{n}$,
1110: by noting that $e^{-k^{\bb}(2\pi i\beta)}=e^{-(k+n)^{\bb}(2\pi
1111: i\beta)}$ and splitting the sum into
1112: \[
1113: f(z)=\sum_{j=0}^{\infty}\sum_{l=1}^{n}e^{-(nj+l)^{\bb}\(\delta+2\pi
1114: i\dfrac{m}{n}\)}=\sum_{l=1}^{n}e^{-2\pi
1115: i\dfrac{m}{n}l^{\bb}}\sum_{j=0}^{\infty}e^{-(nj+l)^{\bb}\delta}\]
1116: It follows that
1117: \[
1118: \sum_{j=0}^{\infty}e^{-(nj+l)^{\bb}\delta}=n^{\bb}\delta\int_{l^b/n^b}^{\infty}e^{-n^{\bb}\delta s}\left(s^{\frac{1}{\bb}}-l/n\right)ds\]
1119: \[
1120: =n^{\bb}\delta \int_{l^\bb/n^\bb}^{\infty}e^{-n^{\bb}\delta s}
1121: \(s^{\frac{1}{\bb}}-l/n\)ds-n^{\bb}\delta
1122: \int_{l^\bb/n^\bb}^{\infty}e^{-n^{\bb}\delta s}\left\{s^{\frac{1}{\bb}}-l/n\right\}ds\]
1123: by the argument above. Since the absolute value of the
1124: fractional part does not exceed one, we have the
1125: estimate
1126: \[
1127: \sum_{j=0}^{\infty}e^{-(nj+l)^{\bb}\delta }=n^{\bb}\delta \int_{(\frac{l}{n})^{\bb}}^{\infty}e^{-n^{\bb}\delta s}s^{\frac{1}{\bb}}ds+O(1)=\frac{1}{n}{\Gamma\(1+\dfrac{1}{\bb}\)}\delta ^{-\frac{1}{\bb}}+O(1)\:(z\rightarrow0)\]
1128: This implies
1129: \begin{equation}
1130: \label{eq:deffz}
1131: f(z)=\frac{1}{n}\sum_{l=1}^{n}e^{-2\pi
1132: i\dfrac{m}{n}l^{\bb}}\Gamma\(1+\dfrac{1}{\bb}\)z^{-\frac{1}{\bb}}+O(1)
1133: \end{equation}
1134: Therefore $f(z)$ either blows up like $z^{-\frac{1}{\bb}}$(when
1135: $\sum_{l=1}^{n}e^{-2\pi i\dfrac{m}{n}l^{\bb}}\neq0$) or it is bounded
1136: (when $\sum_{l=1}^{n}e^{-2\pi i\dfrac{m}{n}l^{\bb}}=0$).
1137:
1138: The Fourier expansion of the fractional part can be used to calculate
1139: the transseries as we did for $\beta=0$, but we shall omit the
1140: calculation here.
1141:
1142: For special values of $b$, asymptotic information is relatively easy
1143: to obtain on a dense set along the barrier. This is the case when
1144: $\bb=\dfrac{r+1}{r}$ where $r\in\NN$; then, the transseries contains
1145: exponential sums in terms of integer powers, $k^{r}$, a
1146: consequence of the duality relation $\dfrac{1}{b}+\dfrac{1}{d}=1$,
1147: which at the transseries level is of the form $\sum
1148: e^{-k^{b}z}\rightarrow g_{1}+\sum e^{ck^{d}z^{-d/b}}g_{2}$ where
1149: $g_{1},g_{2}$ are power series. We illustrate this for
1150: $\bb=\dfrac{3}{2}$.
1151:
1152: Without loss of generality, we assume $\beta<0$. The transseries of
1153: $f$ is given in (\ref{Tr3/2}). To estimate the asymptotic behavior of
1154: $f(z)$ as $z$ approaches the imaginary line, we rewrite (\ref{Tr3/2}) as
1155: \begin{multline}
1156: f(z)=\Gamma\(\frac{5}{3}\)z^{-\frac{2}{3}}-\dfrac{1}{2}\\+\frac{i}{2\pi}\sum_{k=1}^{\infty}\(\frac{k}{z}\)^{2}\int_{0}^{\infty}e^{-s\frac{k^{3}}{z^{2}}}\(\dfrac{3}{4\sqrt{2}(\pi
1157: i)^{\frac{3}{2}}}s^{\frac{1}{2}}-\dfrac{3i}{8\pi^{3}}s-\dfrac{105i^{\frac{3}{2}}}{256\sqrt{2}\pi^{\frac{9}{2}}}s^{\frac{3}{2}}+\cdots\)ds\\
1158: +\sum_{k=1}^{\infty}e^{\dfrac{32i\pi^{3}k^{3}}{27z^{2}}}\dfrac{4\sqrt{2}i^{-\frac{1}{2}}\pi}{3}k^{\frac{1}{2}}z^{-1}\\\frac{i}{\pi}\sum_{k=1}^{\infty}\(\frac{k}{z}\)^{2}e^{-\dfrac{32i\pi^{3}k^{3}}{27z^{2}}}\int_{0}^{\infty}e^{-s\frac{k^{3}}{z^{2}}}\(\dfrac{i^{-\frac{1}{2}}}{8\sqrt{2}\pi^{\frac{3}{4}}}(s-s_{0})^{\frac{1}{2}}+\cdots\)ds
1159: \end{multline}
1160: Watson's Lemma implies that
1161: \begin{equation}
1162: \label{eq:eqtest0}
1163: f(z)=\dfrac{4\sqrt{2}i^{-\frac{1}{2}}\pi}{3z}\sum_{k=1}^{\infty}k^{\frac{1}{2}}e^{\dfrac{32i\pi^{3}k^{3}}{27z^{2}}}+O(1)
1164: \end{equation}
1165: The is sum in (\ref{eq:eqtest0}) is similar to the sum with
1166: $\bb=3$, and can be estimated in a similar way:
1167:
1168: \[
1169: \sum_{k=1}^{\infty}k^{\frac{1}{2}}e^{-(y+2\pi i\frac{m}{n})}=\(\sum_{l=1}^{n}e^{-2\pi i\dfrac{m}{n}l^{3}}\)\frac{\sqrt{\pi}}{3n\sqrt{y}}+o\(\frac{1}{\sqrt{y}}\)\]
1170:
1171:
1172: Setting $\dfrac{32i\pi^{3}k^{3}}{27z^{2}}=y+2\pi i\dfrac{m}{n}$ , we
1173: have $z=-\dfrac{4\pi
1174: i}{3\sqrt{3}}\sqrt{\dfrac{n}{m}}$+$\(\dfrac{n}{m}\)^{\frac{3}{2}}y+o(y)$.
1175: The asymptotic behavior can be obtained for $\beta=-\dfrac{4\pi
1176: i}{3\sqrt{3}}\sqrt{\dfrac{n}{m}}$ (this includes all rationals) by
1177: substituting $y=\dfrac{32i\pi^{3}k^{3}}{27z^{2}}-2\pi i\dfrac{m}{n}$
1178: in the above estimates. Setting $z =-\dfrac{4\pi
1179: i}{3\sqrt{3}}\sqrt{\dfrac{n}{m}}$+$\delta$ ($\delta>0$), a direct
1180: calculation shows that
1181: \begin{equation}
1182: \label{eq:eqtest}
1183: \sqrt{\Re(z )}f(z )=\frac{\sqrt{6\pi
1184: i}}{3^{\frac{7}{4}}}\(\frac{n}{m}\)^{\frac{1}{4}}\frac{1}{n}\sum_{l=1}^{n}e^{-2\pi
1185: i\dfrac{m}{n}l^{3}}+o(1)
1186: \end{equation}
1187: \subsection{Details of the proof of
1188: Theorem~\ref{P1}}\label{genblowup}
1189: Consider, more generally,
1190: \[
1191: f(\delta,\beta)=\sum_{k=0}^{\infty}a(k)e^{-g(k)(\delta+2\pi i\beta)}\:, \ \ \delta>0\]
1192: where $g(k)>0$ is a real function
1193: and$\int_{0}^{\infty}|a(t)|^{2}dt=\infty$.
1194:
1195: We find the behavior of
1196: \begin{multline}
1197: \int_{\beta_{0}}^{\beta_{1}}|f(\delta,\beta)|^{2}d\beta\\
1198: =\int_{\beta_{0}}^{\beta_{1}}\sum_{k=0}^{\infty}|a(k)|^{2}e^{-2g(k)\delta}d\beta+\int_{\beta_{0}}^{\beta_{1}}\sum_{k\neq
1199: j}a(k)\bar{a}(j)e^{-(g(k)+g(j))\delta+(g(j)-g(k))2\pi i\beta}d\beta
1200: \\
1201: =(\beta_{1}-\beta_{0})\sum_{k=0}^{\infty}|a(k)|^{2}e^{-2g(k)\delta}\\
1202: +\frac{1}{2\pi
1203: i}\sum_{k\neq
1204: j}\frac{a(k)\bar{a}(j)}{g(j)-g(k)}e^{-(g(k)+g(j))\delta}e^{(g(j)-g(k))2\pi
1205: i\beta_{0}}\left(e^{(g(j)-g(k))2\pi i(\beta_{1}-\beta_{0})}-1\right)
1206: \end{multline}
1207: where
1208: $\beta_{0,1}\in\RR$ are arbitrary, or after $m$ integrations,
1209: \begin{multline}
1210: F(\delta)=\int_{\beta_{m-1}}^{\beta_{m-1}+c_{m-1}}\cdots\int_{\beta_{1}}^{\beta_{1}+c_1}\int_{\beta_{0}}^{\beta_{0}+c_0}|f(\delta,\beta)|^{2}d\beta d\beta_{0}\cdots
1211: d\beta_{m-2}\\
1212: =c_0c_1\cdots c_m\sum_{k=0}^{\infty}|a(k)|^{2}e^{-2g(k)\delta}+O\left(\sum_{k\neq j}\left|\frac{a(k)\bar{a}(j)}{(g(j)-g(k))^{m}}\right|e^{-(g(k)+g(j))\delta}\right)
1213: \end{multline}
1214:
1215: Note that
1216: \begin{multline}
1217: \sum_{k\neq
1218: j}\left|\frac{a(k)\bar{a}(j)}{(g(j)-g(k))^{m}}\right|e^{-(g(k)+g(j))\delta}\\=2\sum_{k=0}^{\infty}\sum_{n=1}^{\infty}\left|\frac{a(k)\bar{a}(k+n)}{(g(k+n)-g(k))^{m}}\right|e^{-(g(k)+g(k+n))\delta}\\
1219: =2\sum_{k=0}^{\infty}|a(k)|e^{-2g(k)\delta}\sum_{n=1}^{\infty}\left|\frac{a(k+n)}{(g(k+n)-g(k))^{m}}\right|e^{-(g(k+n)-g(k))\delta}\\
1220: =O\left(\sum_{k=0}^{\infty}|a(k)|e^{-2g(k)\delta}\sum_{n=1}^{\infty}\left|\frac{a(k+n)}{(g(k+n)-g(k))^{m}}\right|\right)
1221: \end{multline}
1222: under our assumption.
1223:
1224: If furthermore we have
1225: \[
1226: \sum_{n=1}^{\infty}\left|\frac{a(k+n)}{(g(k+n)-g(k))^{m}}\right|=o(a(k))\]
1227: then we obtain
1228: \begin{equation}
1229: F(\delta)=c_0c_1\cdots c_m\sum_{k=0}^{\infty}|a(k)|^{2}e^{-2g(k)\delta}+O\left(\sum_{k=0}^{\infty}\frac{|a(k)|^{2}}{G(k)}e^{-2g(k)\delta}\right)
1230: \end{equation}
1231: where $G(k)>1,\, G(k)\rightarrow\infty$ as $k\rightarrow\infty$,
1232: or
1233: \begin{equation}
1234: \label{eq:eqB}
1235: F(\delta)\left(\sum_{k=0}^{\infty}|a(k)|^{2}e^{-2g(k)\delta}\right)^{-1} =c_0\cdots c_m+o(1)\ \ as\ \delta\to 0^+
1236: \end{equation}
1237: The result now follows from the following lemma.
1238: \begin{Proposition}
1239: Assume
1240:
1241: (i) $h_n:\RR\to [0,\infty)$ are locally $L^1$, and
1242:
1243: (ii) $\displaystyle \lim_{n\to\infty}\int_{B} h_n(x_1+\cdots +x_N)dx_1\cdots dx_N= meas(B)$ for any box $B\displaystyle=\prod_{i=1}^N[a_i,b_i]$.
1244:
1245: Then
1246: $h_n\to 1$ in the dual of $C[\alpha,\beta]$ for any $[\alpha,\beta]$.
1247:
1248: \end{Proposition}
1249:
1250: \begin{proof}
1251: We first take $N=2$, the general case will follow by induction on $N$.
1252:
1253: Consider the rectangle $B_{c}=(a,b-c)\times (0,c)$, $0<c<b-a$. By changing coordinates
1254: to $x+y=s,y=y'$, we get that
1255: \begin{equation}
1256: \label{eq:Tc}
1257: c^{-1}\int_{B} h_ndydx= \int_a^bh_n(s)T_{a,b;c}(s)ds\to meas(T_{a,b;c});\ \text{as } n\to\infty
1258: \end{equation}
1259: where $T_{a,b;c}(\cdot)$ is the function having $T_{a,b;c}$ as a
1260: graph, $T_{a,b;c}$ being an isosceles trapezoid with lower base the
1261: interval $(a,b)$ and upper base of length $b-a-2c$ at height 1.
1262:
1263:
1264: We also note that the indicator function of $[a,b]$, $\mathbf{1}_{ab}$,
1265: satisfies the inequalities $T_{a-c,b+c;c}\ge \mathbf{1}_{ab}\ge
1266: T_{a,b;c}$. Thus, since $c$ is arbitrary and $h_n\ge 0$, and both
1267: $meas (T_{a-c,b+c;c})$, and $meas (T_{a,b;c})$ tend to $(b-a)$ as
1268: $c\to 0$, we have
1269: \begin{equation}
1270: \label{eq:(b-a)}
1271: \lim_{n\to\infty}\int_a^bh_n(s)ds =(b-a) =meas([a,b])
1272: \end{equation}
1273: In particular, given $\alpha<\beta$, $\|h_n\|_{L^1[\alpha,\beta]}$ are uniformly
1274: bounded, that is, for some $C\ge 1$ we have
1275: \begin{equation}
1276: \label{eq:L1}
1277: \sup_{n\ge 1}\|h_n\|_{L^1[\alpha,\beta]}\le C(\beta-\alpha)
1278: \end{equation}
1279: Since a continuous function on $[\alpha,\beta]$ is approximated
1280: arbitrarily well in sup norm by finite linear combinations of
1281: indicator functions of intervals, it follows from (\ref{eq:(b-a)}),
1282: (\ref{eq:L1}) \footnote{Alternatively, and somewhat more compactly,
1283: one can prove the result without the intermediate steps (\ref{eq:(b-a)})
1284: and (\ref{eq:L1}) by upper and lower bounding continuous functions
1285: by sums of trapezoids.}
1286: and the triangle inequality that
1287: \begin{equation}
1288: \label{eq:concl}
1289: \int_\alpha^\beta h_n(s)f(s)ds\to\int_\alpha^\beta f(s)ds, \ \forall\,f\in C[\alpha,\beta]
1290: \end{equation}
1291: For general $N$ we use $h^-_n(s)=\int_{B'}h_n(s+x_1+...+x_{N-1})dx_1\cdots dx_{N-1}$ and
1292: (\ref{eq:(b-a)}) to reduce the problem to $N-1$.
1293: \end{proof}
1294:
1295:
1296: The condition
1297:
1298: \[
1299: \sum_{n=1}^{\infty}\left|\frac{a(k+n)}{(g(k+n)-g(k))^{m}}\right|=o(a(k))\]
1300: is satisfied, for instance, if
1301:
1302: (1) $\exists c>0$ so that $c<|a(k)|<c^{-1}$, or $|a(k)|$
1303: decreases to 0. (Note that $g(k+n)-g(k)=g'(k+tn)n$, where
1304: $0\leqslant t\leqslant1$, and $g'(k)\rightarrow\infty$ as
1305: $k\rightarrow\infty$.)
1306:
1307: (2) $\exists c,r$ so that $c<a(k)<k^{r}$. $g'(k)\geqslant
1308: k^{\varepsilon}$ for some small $\varepsilon>0$.
1309: \section{Proof of Theorem~\ref{T1}}\label{S5}
1310: In Appendix \S\ref{Fi} we list some known facts about iterations of maps.
1311:
1312: \begin{proof}[Proof of B\"otcher's theorem, for (\ref{eq:eqG})]
1313: ({\bf Note:} this line of proof extends to general analytic maps.)
1314:
1315: We
1316: write $\psi=\lambda z+\lambda^2zg(z)$ and obtain
1317: \begin{equation}
1318: \label{eq:H}
1319: g(z)-\frac{1}{2}g(z^2)=\frac{1}{2}z+\frac{1}{2}\lambda\left[g(z)(z-g(z))+g(z^2)\right]+\frac{\lambda^2 z}{2}g(z)g(z^2)=N(g)
1320: \end{equation}
1321: Let $\mathcal{A_{\lambda}}$ denote the functions analytic
1322: in the polydisk $\mathbb{P}_{1,\epsilon}=\mathbb{D}\times
1323: \{\lambda:|\lambda|<\epsilon\}$. We write (\ref{eq:H}) in the form (see \ref{eq:defT}))
1324: \begin{equation}
1325: \label{eq:eqH4}
1326: g=2\mathfrak TN(g)
1327: \end{equation}
1328: This equation is manifestly contractive in the sup norm, in a ball of
1329: radius slightly larger than $1/2$ in $\mathcal{A_{\lambda}}$, if
1330: $\epsilon$ is small enough. For $\lambda\ne 0$, evidently
1331: $\psi=\phi^{-1}$ is also analytic at zero. \end{proof}
1332: \begin{Lemma}
1333: $\psi$ is analytic in
1334: $\mathbb{D}_1$ for all $\lambda$ with $|\lambda|<1$.
1335: \end{Lemma}
1336: \begin{proof} We have
1337: \begin{equation}
1338: \label{eq:eqGM}
1339: \psi(z)=\frac{\lambda}{2}\left( X+\sqrt{X^2+4X/\lambda}\right)=:F(X);\ X=\psi(z^2)
1340: \end{equation}
1341: For small $z\ne 0$, $\psi(z)=O(z)$ and thus $F(\psi(z^2))$ is well defined and analytic. Note that (\ref{eq:eqGM}) provides analytic
1342: continuation
1343: of $\psi$ from $\mathbb{D}_{\rho^2}$ to $\mathbb{D}_{{\rho}}$, provided nowhere
1344: in $\mathbb{D}_{\rho^2}$ do we have $\psi=-4/\lambda$ (certainly the case if $\rho$ is small). We assume, to get a contradiction, that there is a $z_0$,
1345: $|z_0|=\lambda_0<1$ so that $\psi(z_0)=-4/\lambda$, and we choose the least
1346: $\lambda_0$ with this property. By the previous discussion, $\psi$
1347: is analytic in the open disk $\mathbb{D}_{\sqrt{\lambda_0}}$.
1348: Then we use the ``backward'' iteration $\psi(z^2)=
1349: \lambda^{-1}\psi(z)^2/(1+\psi(z))$ to calculate $\psi(z^2)$
1350: from $\psi(z)$, starting with $z=z_0$. This is in fact
1351: equivalent to (\ref{eq:infty1}); after the substitution $x=(-\lambda y)^{-1}$
1352: we return to (\ref{eq:logist1}), with $x_0=\lambda/4$. Using (vi) and
1353: (vii) of \S\ref{Fi}, it follows that $1/x_n\not\to 0$, that is, $\psi(z_0^{2^n})\not \to 0$. This impossible, since $\psi$ is analytic and $\psi(0)=0$.
1354: \end{proof}
1355:
1356: \begin{proof}[Proof of Theorem~\ref{T1}, (i)]
1357: We return to
1358: (\ref{eq:eqG}). Taking $a\in (0,1)$ $m_n=\sup \{|\psi(z)|:
1359: |z|<a^{1/2^n}$ we note that
1360: \begin{equation}
1361: \label{eq:eqm}
1362: m_{n+1}\le \frac{1}{2} |\lambda|(m_n+\sqrt{m_n^2+4 m_n/|\lambda|})
1363: \end{equation}
1364: The sequence of $m_n$ is bounded by the sequence of $M_n$, defined by replacing
1365: ``$\le$'' with ``$=$'' in (\ref{eq:eqm}). Since $\frac{1}{2}
1366: |\lambda|(x+\sqrt{x^2+4 x/|\lambda|})<x$ if
1367: $x>A:=|\lambda|/(1-|\lambda|)$, we have $\limsup_n M_n\le A$. By the
1368: maximum principle, $|\psi(\lambda,z)|< A$ in
1369: $\mathbb{D}\times\mathbb{D}$. Thus, by Cauchy's formula in $\lambda$
1370: we have $|\psi_n(z)|\le A$ for all $n$ and $z\in \mathbb{D}$. The
1371: radius of convergence of (\ref{eq:formG}) in $\lambda$ is at least
1372: one. By \S\ref{Fi}, (vii), the radius of convergence is exactly
1373: one.
1374:
1375: Indeed, note first that (a) if $\psi$ is analytic in
1376: $\mathbb{D}$ then $\psi'\ne 0 $ in $\mathbb{D}$, otherwise
1377: $\psi'(z_1)=0$ would imply $\psi'(z_1^{2^n})=0$ in contradiction
1378: with $\psi'(0)=\lambda$. This means that if there is a $z_0$, $\psi(z_0)=-4/\lambda$, then $\sqrt{z_0}$ is a singular point of $\psi$.
1379:
1380: Secondly, any $\lambda$ of the form $1+i\epsilon$ with small
1381: $\epsilon$ correspond to $c=1/4+1/4 \epsilon^2,$ outside the
1382: Mandelbrot set. Thus, in the iteration (\ref{eq:infty1}), the initial
1383: condition $y_0=-4/\lambda$ implies $y_n\to 0$. We can now use the
1384: implicit function theorem to suitably match $y_n$, once it is small
1385: enough, to some value of $\psi$ near zero. Indeed, the equation
1386: $y_n=\lambda z_{0}^{2^n}+O(z_0^{2^{n+1}})$ has $2^n$ solutions. This
1387: means that for such a $z_0$, using (\ref{eq:eqGM}) to iterate
1388: backwards and to determine $\psi(z_0)$ (noting the parallel to
1389: (\ref{eq:infty1})), we have $\psi(z_0)=-4/\lambda$, and by (a) above,
1390: $\psi$ cannot be analytic in $z$ in $\mathbb{D}$.
1391:
1392: Formula (\ref{eq:frakt}) follows by straightforward expansion of (\ref{eq:eqG})
1393: and identification of powers of $\lambda$.\end{proof}
1394: \begin{proof}[Proof of Theorem~\ref{T1}, (ii)]
1395: The stated type of lacunarity of $\psi_k$ follows from (\ref{eq:frakt}) by induction,
1396: noting the discrete convolution structure in $k$.\end{proof}
1397: \begin{proof}[Proof of Theorem~\ref{T1}, (iii)]
1398: Continuity of $\psi_k$ in $\overline{\mathbb{D}}$ also follows by induction from (\ref{eq:frakt}) and the
1399: properties of $\mathfrak{T}$. By dominated convergence (applied to the
1400: discrete measure $|\lambda|^n$), for $\lambda<1$, $\psi$ is
1401: continuous in $\overline{\mathbb{D}}$ and the Fourier series converges
1402: pointwise in $\partial{\mathbb{D}}$.
1403:
1404: To show convergence of the Fourier series of $H$ we only need to show
1405: $\inf_{\mathbb{D}}|\psi|>0$. Now, $\psi$ clearly cannot
1406: vanish for any $z_0=\mathbb{D}$, otherwise $\psi(z_0^{2^n})=0$, would imply
1407: by analyticity $\psi\equiv 0$. If
1408: $\min_{\mathbb D_{\rho}}|\psi(\rho)|=\epsilon$ would be small enough, then
1409: $\min_{\mathbb D_{\rho^2}}|\psi(\rho)|\le O(\epsilon^2)\ll
1410: \epsilon$, contradicting the maximum principle for $z/\psi(z)$.
1411:
1412:
1413: The rest of the proof is straightforward
1414: calculation, using the analyticity of $\psi$. \end{proof} The extension of the small
1415: $\lambda$ analysis to higher
1416: order polynomials is also straightforward.
1417: \begin{Note}\label{Explc} The transseries of the B\"otcher map at binary rational
1418: numbers can be calculated rather explicitly. This is beyond
1419: the scope here, and will be the subject of a different paper. A less explicit
1420: expression has been obtained in \cite{CPAM}. We note that the constant
1421: $\log_2(2\pi)$ in (1.17) of \cite{CPAM} should be $(2\pi)/\log
1422: 2$.
1423: \end{Note}
1424: \section{Appendix}\label{appendix}
1425: \subsection{Proof of Lemma~\ref{linv}}
1426: For every $m$, we write $g(s)=g_m+o(s^{-m})$ where $g_m$
1427: is a finite sum, an initial sum in the asymptotic series
1428: of $g$. It is straightforward to show that $g_m^{-1}(y)$ has
1429: an asymptotic power series as $y\to\infty$. Then, in the equation
1430: $g(s)=y$ we write $s=s_m+\epsilon$ where $g_m(s_m)=y$. Then,
1431: $y=g(s)=g(s_m+\epsilon)=g_m(s_m)+g'(\xi)\epsilon+(g(s_m)-g_m(s_m))$
1432: implies $\epsilon=(g_m(s_m)-g(s_m))/g'(\xi)=o(s^{-m-\theta})$ where $g'(\xi)\sim a x^{\theta}$. Now $\theta$ is fixed and $m$ is arbitrary, and then the result
1433: follows.
1434: \subsection{Proof of Lemma~\ref{eq:asympt1}}
1435: We have
1436: \begin{multline}
1437: \label{eq:34}
1438: f-e^{-xg(0)}=-\sum_{k=1}^{\infty}k(e^{-xg(k+1)}-e^{-xg(k)})=x\int_0^{\infty} e^{-xg(s)}g'(s)\lfloor s \rfloor ds\\=x\int_{g(0)}^{\infty} e^{-xu}\lfloor g^{-1}(u)\rfloor du=x\int_{g(0)}^{\infty} e^{-xu} g^{-1}(u)du-
1439: x\int_{g(0)}^{\infty} e^{-xu} \{g^{-1}(u)\} du
1440: \end{multline}
1441: and we also have
1442: \begin{equation}
1443: \label{eq:ginv}
1444: x\int_{g(0)}^{\infty} e^{-xu} g^{-1}(u)du=-\int_{0}^{\infty} (e^{-xg(u)})'u du=
1445: \int_{0}^{\infty} e^{-xg(u)}du
1446: \end{equation}
1447: whereas
1448: $$0<\int_0^{\infty} e^{-xu} xg'(u)\{u\}du\le \int_0^{\infty} e^{-xu} xg'(u)du=e^{xg(0)}$$
1449: \subsection{Proof of Lemma \ref{Lgeom}}\label{pfLgeom}
1450: By the Fourier coefficients formula we have
1451: \begin{multline}
1452: c_{0}=\int_{0}^{1}f(a^{y})dy+\int_{0}^{1}ydy-\int_{0}^{1}\check{G}(a^{y})dy\\
1453: c_{k}=\int_{0}^{1}f(a^{y})e^{-2k\pi iy}dy+\int_{0}^{1}ye^{-2k\pi iy}dy-\int_{0}^{1}\check{G}(a^{y})e^{-2k\pi iy}dy\\
1454: =\frac{i}{2k\pi}+\sum_{n=0}^{\infty}\int_{0}^{1}e^{-a^{n+y}-2k\pi iy}dy+\sum_{n=1}^{\infty}\int_{0}^{1}\frac{(-1)^{n}a^{ny}}{n!(a^{n}-1)}e^{-2k\pi iy}dy\\
1455: =\frac{i}{2k\pi}+\frac{1}{\log a}\sum_{n=0}^{\infty}a^{\frac{2kn\pi i}{\log a}}\(\Gamma\(-\frac{2k\pi i}{\log a},a^{n}\)-\Gamma\(-\frac{2k\pi i}{\log a},a^{1+n}\)\)\\+\sum_{n=1}^{\infty}\frac{(-1)^{n+1}}{n!(2k\pi i-n\log a)} =\sum_{n=0}^{\infty}\frac{(-1)^{n+1}}{n!(2k\pi i-n\log a)}+\frac{1}{\log a}\Gamma\(-\frac{2k\pi i}{\log a},1\)\:(k\neq0)
1456: \end{multline}
1457:
1458:
1459: Note that since $\mathcal{L}^{-1}(\frac{1}{2k\pi i-n\log a})=\dfrac{1}{\log a}e^{\frac{2k\pi ip}{\log a}}\:(n\rightarrow p)$
1460: we have
1461: \begin{multline*}
1462: \sum_{n=0}^{\infty}\frac{(-1)^{n+1}}{n!(2k\pi i-n\log
1463: a)}=\int_{0}^{\infty}\sum_{n=0}^{\infty}\frac{(-1)^{n+1}e^{-np}}{n!}\dfrac{1}{\log
1464: a}e^{\frac{2k\pi ip}{\log a}}dp\\=\dfrac{1}{\log
1465: a}\int_{0}^{\infty}e^{-e^{-p}}e^{\frac{2k\pi ip}{\log
1466: a}}dp\\=\dfrac{1}{\log a}\int_{0}^{1}e^{-t}t^{\frac{-2k\pi i}{\log
1467: a}-1}dt=\frac{1}{\log a}\(\Gamma(-\frac{2k\pi i}{\log
1468: a})-\Gamma(-\frac{2k\pi i}{\log a},1)\)
1469: \end{multline*}
1470:
1471:
1472: The above procedure is justified for $k$ in the upper half plane.
1473: By analytic continuation the expression holds for $k$ real as well.
1474: Eq. (\ref{eq:ck}) follows.
1475:
1476:
1477:
1478:
1479: We can further resum the series in the above expression by noting
1480: that
1481: \begin{multline}
1482: \sum_{k\neq0}\Gamma\(-\frac{2k\pi i}{\log a}\)
1483: z^{\frac{2k\pi i}{\log a}}=\sum_{k=1}^{\infty}
1484: \frac{-\log a}{2k\pi i}\int_{0}^{\infty}
1485: t^{\frac{-2k\pi i}{\log a}}e^{-t}z^{\frac{2k\pi i}{\log a}}dt
1486: \\+\sum_{k=-\infty}^{-1}\frac{-\log a}{2k\pi i}=\int_{0}^{\infty}\sum_{k=1}^{\infty}\frac{-\log a}{2k\pi i}
1487: \(\frac{z}{t}\)^{\frac{2k\pi i}{\log a}}e^{-t}dt
1488: \int_{0}^{\infty}t^{\frac{-2k\pi i}{\log a}}e^{-t}z^{\frac{2k\pi i}{\log a}}dt
1489: \\+
1490: \int_{0}^{\infty}\sum_{k=-\infty}^{-1}\frac{-\log a}{2k\pi i}
1491: \(\frac{z}{t}\)^{\frac{2k\pi i}{\log a}}e^{-t}dt=\frac{\log a}{2\pi i}\int_{0}^{\infty}
1492: \log_{\RR}\(1-\(\frac{z}{t}\)^{\frac{2\pi i}{\log a}}\)e^{-t}dt
1493: \\-
1494: \frac{\log a}{2\pi i}\int_{0}^{\infty}
1495: \log_{\RR}\(1-\(\frac{z}{t}\)^{-\frac{2\pi i}{\log a}}\)e^{-t}dt=\frac{\log a}{2\pi i}\int_{0}^{\infty}
1496: \log_{\RR}\(-\(\frac{z}{t}\)^{\frac{2\pi i}{\log a}}\)e^{-t}dt
1497: \\
1498: =-\frac{\log a z}{2\pi i}\int_{0}^{\infty}\log_{\RR}\(-s^{\frac{2\pi
1499: i}{\log a}}\)e^{-sz}ds
1500: \end{multline}
1501: which can be justified by analytic continuation, for the last
1502: expression and the sum are both analytic, and equal to each other on
1503: the real line. The logarithm $\log_{\RR}$ is defined with a branch cut along $\RR^-$.
1504:
1505: We finally obtain the integral representation (\ref{eq:resum}),
1506: valid for $z$ in the right half plane.
1507:
1508: \begin{Remark}
1509: Since $\log_{\RR}\(-s^{\frac{2\pi i}{\log a}}\)=i\arg\(\dfrac{2\pi \log
1510: s}{\log a}-\pi\)=2\pi i \(\left\{\dfrac{\log s}{\log a}\right\}-\dfrac{1}{2}\)$,
1511: we actually recover the last term of (\ref{eq:asympt1}).
1512: \end{Remark}
1513:
1514:
1515: \subsection{Direct calculations for $b\in\NN$ integer; the cases $b=3,b=3/2$}\label{Direct,3}
1516: \begin{Proposition}
1517: If $\bb$ is an integer, the behavior of \[
1518: f(\delta+2\pi i \beta)=\sum_{k=1}^{\infty}e^{-k^{\bb}(\delta+2\pi i\beta)}\:(\Re(\delta)>0)\]
1519: where $\beta=2\pi i m/n$, m and n being integers, as $\delta$ approaches
1520: 0 is
1521: \[
1522: f(\delta+2\pi i \beta)=\left[\frac{1}{n}\sum_{l=1}^{n}e^{-2\pi
1523: i\dfrac{m}{n}l^{\bb}}\Gamma\(1+\dfrac{1}{\bb}\)\right]\delta^{-\frac{1}{\bb}}+O(1)\]
1524: Therefore $f(\delta)$ either blows up like $\delta^{-\frac{1}{\bb}}$ or is
1525: bounded.
1526:
1527: The more general case $\bb=\dfrac{r+1}{r}$ where $r$ is an integer
1528: can be treated similarly. In particular, if
1529: $\bb=\dfrac{3}{2}$, for $\beta=-\dfrac{4\pi
1530: i}{3\sqrt{3}}\sqrt{\dfrac{n}{m}}$(this includes all rational
1531: numbers), we have
1532: \[
1533: \sum_{k=1}^{\infty}e^{-k^{3/2}z}=\frac{\sqrt{6\pi
1534: i}}{3^{\frac{7}{4}}m^{\frac{1}{4}}n^{\frac{3}{4}}}\sum_{l=1}^{n}e^{-2\pi
1535: i\dfrac{m}{n}l^{3}}\frac{1}{\sqrt{\delta}}+o\left(\frac{1}{\sqrt{\delta}}\right)\]
1536: with $z=-\dfrac{4\pi
1537: i}{3\sqrt{3}}\sqrt{\dfrac{n}{m}}$+$\delta$($\delta\to 0^+$)
1538: \end{Proposition}
1539:
1540: \begin{proof}
1541:
1542:
1543: In general, to find the asymptotic behavior of $f(z)$ we analyze
1544: the functions
1545: \begin{equation}
1546: \label{eq:deffk}
1547: f_{k}(z)=\int_{0}^{\infty}e^{-pz-2k\pi
1548: ip^{\frac{1}{\bb}}}dp,\ k\in\ZZ
1549: \end{equation}
1550: Letting $p=q\(\dfrac{z}{k}\)^{\frac{\bb}{1-\bb}}$ we have
1551:
1552: \[
1553: \int_{0}^{\infty}e^{-pz-2k\pi
1554: ip^{\frac{1}{\bb}}}dp=\(\frac{k}{z}\)^{\frac{\bb}{\bb-1}}\int_{0}^{\infty}e^{-(q+2\pi
1555: iq^{\frac{1}{\bb}})(\frac{k^{\bb}}{z})^{\frac{1}{\bb-1}}}dq\]
1556:
1557:
1558: Next we let $s=h(q)=q+2\pi iq^{\frac{1}{\bb}}$and
1559: $f_{k}(z)=(\frac{k}{z})^{\frac{\bb}{\bb-1}}\intop_{C_1}e^{-s(\frac{k^{\bb}}{z})^{\frac{1}{\bb-1}}}\ds\frac{1}{h'(h^{-1}(s))}ds$
1560: where the contour $C_1$ is a curve from the origin to $\infty$ in
1561: the first quadrant.
1562:
1563: We can find the asymptotic behavior of $f_{k}(z)$ using Watson's
1564: Lemma \cite{benderorszag}. By iterating the contractive map
1565: $q\rightarrow\(\dfrac{s-q}{2\pi i}\)^{\bb}$ near 0, we can easily
1566: see that $\dfrac{h^{-1}(s)}{s^{\bb}}$ is analytic in $s^{\bb-1}$,
1567: which implies $\ds\frac{1}{h'(h^{-1}(s))}$ is analytic in
1568: $s^{\bb-1}$ near 0 with no constant term.
1569:
1570:
1571:
1572:
1573:
1574:
1575: Now let's consider the examples $\bb=3$ and $b=3/2$; for $\bb=3$ we
1576: have $s=h(q)=q+2\pi iq^{\frac{1}{3}}$ and
1577: $\dfrac{ds}{dq}=$$\dfrac{3i}{8\pi^{3}}s^{2}+\dfrac{15}{64\pi^{6}}s^{4}-\dfrac{21i}{128\pi^{9}}s^{6}\cdots$. Thus, the asymptotic power series is
1578:
1579: \[
1580: f(z)\sim\Gamma\(\frac{4}{3}\)z^{-\frac{1}{3}}-\dfrac{1}{2}-\dfrac{z}{120}+\frac{z^{3}}{792}+\cdots\]
1581:
1582:
1583: The branch point of $h^{-1}$ is located at
1584: $s_{0}=\pi^{\frac{3}{2}}\dfrac{4\sqrt{2}}{3\sqrt{3}}(-1)^{\frac{1}{4}}$,
1585: which is between the contour $C_1$ defined above and the $x$-axis. As we start
1586: rotating $z$ from $z>0$, we have, cf (\ref{eq:deffk}),
1587: \begin{multline}
1588: f_{k}(z)=\(\frac{k}{z}\)^{\frac{3}{2}}\intop_{C_1}\frac{e^{-s(\frac{k^{3}}{z})^{\frac{1}{2}}}}{h'(h^{-1}(s))}ds\\
1589: =\(\frac{k}{z}\)^{\frac{3}{2}}\intop_{0}^{\infty}\frac{e^{-s(\frac{k^{3}}{z})^{\frac{1}{2}}}}{h'(h^{-1}(s))}ds+\(\frac{k}{z}\)^{\frac{3}{2}}e^{-s_{0}(\frac{k^{3}}{z})^{\frac{1}{2}}}\intop_{C_2}\frac{e^{-s(\frac{k^{3}}{z})^{\frac{1}{2}}}}{h'(h^{-1}(s+s_{0}))}ds\\
1590: =\(\frac{k}{z}\)^{\frac{3}{2}}\int_{0}^{\infty}\frac{e^{-s(\frac{k^{3}}{z})^{\frac{1}{2}}}}{h'(h^{-1}(s))}ds+2\(\frac{k}{z}\)^{\frac{3}{2}}e^{-s_{0}(\frac{k^{3}}{z})^{\frac{1}{2}}}\int_{0}^{\infty}\frac{e^{-s(\frac{k^{3}}{z})^{\frac{1}{2}}}}{h'(h^{-1}(s+s_{0}))}ds
1591: \end{multline}
1592: where the contour $C_2$ starts at $\infty$, goes clockwise around
1593: the origin, then ends at $\infty.$ Since now
1594:
1595: $$\dfrac{ds}{dq}=\dfrac{(-\pi
1596: i)^{\frac{3}{4}}}{6^{\frac{1}{4}}}(s-s_{0})^{-\frac{1}{2}}+\dfrac{5}{6}+\dfrac{5}{16}\(\dfrac{i}{6\pi}\)^{\frac{3}{4}}(s-s_{0})^{\frac{1}{2}}+\cdots$$
1597:
1598: We have
1599: \begin{multline}
1600: \(\frac{k}{z}\)^{\frac{3}{2}}e^{-s_{0}z}\intop_{C_2}\frac{e^{-s(\frac{k^{3}}{z})^{\frac{1}{2}}}}{h'(h^{-1}(s+s_{0}))}ds\\=e^{-\pi^{\frac{3}{2}}\frac{4\sqrt{2}}{3\sqrt{3}}(-1)^{\frac{1}{4}}z}\(\(\dfrac{\pi
1601: i}{6}\)^{\frac{1}{4}}k^{-\frac{1}{4}}z^{-\frac{1}{4}}+\dfrac{5}{32}\dfrac{i^{\frac{7}{4}}}{6^{\frac{3}{4}}\pi^{\frac{5}{4}}}k^{-\frac{7}{4}}z^{\frac{1}{4}}+\cdots\)
1602: \end{multline}
1603:
1604:
1605: Therefore the transseries is
1606: \begin{multline}
1607: \label{eq:trans1}
1608: \tilde{f}(z)=\Gamma(\frac{4}{3})z^{-\frac{1}{3}}-\dfrac{1}{2}-\dfrac{z}{120}+\frac{z^{3}}{792}+\cdots\\
1609: +\sum_{k=1}^{\infty}e^{-\pi^{\frac{3}{2}}\frac{4\sqrt{2}}{3\sqrt{3}}(-1)^{\frac{1}{4}}(\frac{k^{3}}{z})^{\frac{1}{2}}}\left[\(\dfrac{\pi
1610: i}{6}\)^{\frac{1}{4}}k^{-\frac{1}{4}}z^{-\frac{1}{4}}+\dfrac{5}{32}\dfrac{i^{\frac{7}{4}}}{6^{\frac{3}{4}}\pi^{\frac{5}{4}}}k^{-\frac{7}{4}}z^{\frac{1}{4}}+\cdots\right]\\
1611: -\sum_{k=1}^{\infty}e^{-\pi^{\frac{3}{2}}\frac{4\sqrt{2}}{3\sqrt{3}}(-1)^{\frac{1}{4}}(\frac{-k^{3}}{z})^{\frac{1}{2}}}\left[\(\dfrac{\pi
1612: i}{6}\)^{\frac{1}{4}}(-k)^{-\frac{1}{4}}z^{-\frac{1}{4}}+\dfrac{5}{32}\dfrac{i^{\frac{7}{4}}}{6^{\frac{3}{4}}\pi^{\frac{5}{4}}}(-k)^{\frac{7}{4}}z^{\frac{1}{4}}+\cdots\right]
1613: \end{multline}
1614: The calculation for $\bb=\dfrac{3}{2}$ is similar: in this case
1615: $\dfrac{ds}{dq}=\dfrac{3}{4\sqrt{2}(\pi
1616: i)^{\frac{3}{2}}}s^{\frac{1}{2}}-\dfrac{3i}{8\pi^{3}}s-\dfrac{105i^{\frac{3}{2}}}{256\sqrt{2}\pi^{\frac{9}{2}}}s^{\frac{3}{2}}+\cdots$
1617: and the asymptotic power series is
1618:
1619: \[
1620: f(z)\sim\Gamma\(\frac{5}{3}\)z^{-\frac{2}{3}}-\dfrac{1}{2}-\dfrac{3\zeta(\frac{5}{2})}{16\pi^{2}}z+\frac{1}{240}z^{2}+\frac{315\zeta(\frac{11}{2})}{2048\pi^{5}}z^{3}+\cdots\]
1621:
1622:
1623: The exponential sum is slightly different than in the previous case, for
1624: now the branch point $s_{0}=-\dfrac{32}{27}i\pi^{3}$ lies in the
1625: lower half plane, which means the contour $C_2$ can be deformed
1626: to $[0,$$+\infty)$ without passing through any singularity.
1627:
1628: We collect the contribution from the branch point only when $\arg z$
1629: decreases to $-\dfrac{\pi}{4}$from 0. Since
1630: $$\dfrac{ds}{dq}=\dfrac{-4\sqrt{2}i^{\frac{1}{2}}\pi^{\frac{3}{2}}}{3}(s-s_{0})^{-\frac{1}{2}}+\dfrac{4}{3}+\dfrac{i^{-\frac{1}{2}}}{8\sqrt{2}\pi^{\frac{3}{4}}}(s-s_{0})^{\frac{1}{2}}+\cdots$$
1631: we have for the exponential part of the sum
1632: $$\(\dfrac{4\sqrt{2}i^{-\frac{1}{2}}\pi}{3}k^{\frac{1}{2}}z^{-1}+\dfrac{i^{\frac{1}{2}}}{16\sqrt{2}\pi^{\frac{5}{4}}}k^{-\frac{5}{2}}z^{2}+\cdots\)\exp\({\dfrac{32i\pi^{3}k^{3}}{27z^{2}}}\)$$
1633: Therefore, for small $z$ in the right half plane, the transseries is given by
1634: \begin{multline}
1635: \label{Tr3/2}
1636: \tilde{f}(z)=\Gamma\(\frac{5}{3}\)z^{-\frac{2}{3}}-\dfrac{1}{2}-\dfrac{3\zeta(\frac{5}{2})}{16\pi^{2}}z+\frac{1}{240}z^{2}+\frac{315\zeta(\frac{11}{2})}{2048\pi^{5}}z^{3}+\cdots\\+\begin{cases}
1637: \sum\limits _{k=1}^{\infty}e^{\frac{32i\pi^{3}k^{3}}{27z^{2}}}\(\frac{4\sqrt{2}i^{-\frac{1}{2}}\pi}{3}k^{\frac{1}{2}}z^{-1}+\frac{i^{\frac{1}{2}}}{16\sqrt{2}\pi^{\frac{5}{4}}}k^{-\frac{5}{2}}z^{2}+\cdots\); \,\,-\frac{\pi}{2}\leqslant\arg z\leqslant-\frac{\pi}{4}\\
1638: -\sum\limits
1639: _{k=1}^{\infty}e^{-\frac{32i\pi^{3}k^{3}}{27z^{2}}}\(\frac{4\sqrt{2}i^{-\frac{1}{2}}\pi}{3}(-k)^{\frac{1}{2}}z^{-1}+\frac{i^{\frac{1}{2}}}{16\sqrt{2}\pi^{\frac{5}{4}}}(-k)^{-\frac{5}{2}}z^{2}+\cdots\);\,\frac{\pi}{4}\leqslant\arg
1640: z\leqslant\frac{\pi}{2}\end{cases}
1641: \end{multline}
1642: The effect of the exponential part of the transseries affects the leading
1643: order when
1644: $z\rightarrow0$ nearly tangentially to the imaginary line.
1645:
1646: For example, to see the effect of the first exponential term for
1647: $\bb=\dfrac{3}{2}$, we let $z^{-\frac{2}{3}}\sim
1648: e^{\frac{32i\pi^{3}}{27z^{2}}}z^{-1}$ and $z=-ire^{i\theta}$ near
1649: the negative imaginary line.
1650:
1651: The critical curve along which the power term and the exponential
1652: term are of equal order is
1653: $\theta\sim\dfrac{9}{64\pi^{3}}r^{2}\log\dfrac{1}{r}$.
1654: \end{proof}
1655: \begin{figure}[h!]
1656: $ $ \hskip -3cm
1657: \includegraphics[scale=0.38]{M}
1658: \caption{The Madelbrot set (drawn with xaos 3.1 \cite{xaos}).} \label{M1}
1659: \label{fig:12}
1660: \end{figure}
1661: \subsection{Notes about iterations of maps}\label{Fi}
1662:
1663: For the following, see e.g., \cite{Beardon,Devaney,Milnor}.
1664: \begin{enumerate}
1665: \item For $|\lambda|<1$, three types of behavior are possible for
1666: the solution of (\ref{eq:logist1}): if the initial condition $x_0\in \mathcal{F}$, the connected component
1667: of the origin in {\em the Fatou set}, then $x_n\to 0$ as $n\to \infty$.
1668: Clearly, $x_0\in \mathcal{F}$ if $x_0$ is small enough. If $x_0\in \overline{\mathcal{F}}^c$, the connected
1669: component of infinity in the Fatou set, then
1670: $|x_n|\to\infty$. Clearly, $x_0\in \overline{\mathcal{F}}^c$ if it is large enough. Finally, for $x_0\in \partial{F}=J$, the Julia set, a (connected) curve
1671: of nontrivial Hausdorff dimension invariant under
1672: the map, $\{x_n\}_n$ are dense in $J$ and the evolution is chaotic.
1673:
1674:
1675: \item $J$ is the closure of the set of repelling periodic points.
1676:
1677: \item For polynomial maps, and more generally, for entire maps, $J$ is the boundary of the set of points which converge to infinity under iteration.
1678:
1679: \item If the maximal disk of analyticity of $\psi$ is the unit disk $\mathbb{D}_1$, then $\psi$ maps $\mathbb{D}_1$ biholomorphically onto the immediate basin $\mathcal{A}_0$ of zero. If on the contrary the maximal disk is $\mathbb{D}_r,\,r<1$, then there is at least one other critical point in $\mathcal{A}_0$, lying in $\psi(\partial\mathbb{D}_r)=J_y$, the Julia set of \ref{eq:infty1}.
1680:
1681: \item If $r=1$, it follows that $\psi(\partial\mathbb{D}_1)=J_y$.
1682:
1683:
1684: \item By the change
1685: of variable $x_n=-(1/\lambda)t_n+1/2$, (\ref{eq:logist1}) is brought to the ``$c$ form''
1686: $t_{n+1}=t_n^2+c$, $c=\lambda/2-\lambda^2/4$. The Mandelbrot set is defined
1687: as (see e.g. \cite{Devaney})
1688: \begin{equation}
1689: \label{eq:defM}
1690: \mathcal{M}=\{c: t_n \text{ bounded }\text{ if } t_0=0\}
1691: \end{equation}
1692: If $c\in\mathcal{M}$, then clearly $y_n$ in (\ref{eq:infty1}) are bounded away from zero. Note that $t_0=0$ corresponds to $x_0=1/2$ implying $x_1=-\lambda/4$.
1693: \item \label{(iii)} $\mathcal{M}$ is a compact set; it coincides
1694: with the set of $c$ for which $J$ is connected. The
1695: cardioid $\mathcal{H}=\{(2e^{it}-e^{2it})/4:t\in [0,2\pi)\}$ is contained in
1696: $\mathcal{M}$; see \cite{Devaney}. This means $\{\lambda:|\lambda|<1\}$
1697: corresponds to the interior of $\mathcal{M}$. We have
1698: $|\lambda|=1\Rightarrow c\in \partial \mathcal{M}\subset\mathcal{M}$.
1699:
1700: \end{enumerate}
1701:
1702:
1703: \subsection{Overview of Borel summability and transseries}\label{A4}
1704:
1705: There is a vast literature on transseries, Borel summability, and
1706: resurgence, see, for example \cite{Sauzin}. Most of the modern theory
1707: originates in Ecalle's work \cite{Ecalle}.
1708: \begin{Definition}\label{def1}
1709: We say that $f$ is given by a Borel summable transseries for
1710: $x>\nu$, if there exists a $\beta \in\CC$, a sequence $c_k$, with
1711: $\Re\,c_k\ge Ck$ for some $C>0$, and a sequence of functions $Y_k$,
1712: analytic in a neighborhood of $(0,\infty)$, having convergent Puiseux
1713: series at zero, and $|Y_k(p)|\le |B^ke^{\nu p}|$ (where $B$ and
1714: $\nu$ are independent of $k$) such that
1715: \begin{equation}
1716: \label{eq:deftr}
1717: f(x)=\sum_{k=0}^{\infty}e^{-c_k x}x^{(k+1)\beta}\mathcal{L} Y_k
1718: \end{equation}
1719: where $\mathcal{L}$ is the usual Laplace transform:
1720: \begin{equation}
1721: \label{eq:lapl}
1722: (\mathcal{L}Y)(x)=\int_0^{\infty}e^{-px}Y(p)dp
1723: \end{equation}
1724: The definition for other directions $\theta$ in the $x$ complex domain is obtained
1725: by changing the variable to $x'=xe^{-i\theta}$.
1726:
1727: The Borel-Laplace summation operator is denoted by $\mathcal{LB}$.
1728: \end{Definition}
1729: \begin{Definition}
1730: A formal power series in powers of $1/x$ is Borel summable as
1731: $x\to\infty$ if it is the asymptotic series of $\mathcal{L} Y$,
1732: where $Y$ is as in Definition \ref{def1}. (We note that by Watson's
1733: Lemma \cite{benderorszag}, $(\mathcal{L}Y)(x)$ has an asymptotic
1734: series as $x\to\infty$, which is the termwise Laplace transform of
1735: the Taylor series of $Y$ at zero.)
1736: \end{Definition}
1737: Transseries representations contain therefore manifest asymptotic information.
1738: \begin{Definition}
1739: A function $Y(p)$ is resurgent in $p$ in the sense of Ecalle \cite{Ecalle}, if it
1740: is analytic on the Riemann surface of $\CC\setminus J$, where $J$ is
1741: a discrete set, and has uniform exponential bounds along any
1742: direction towards infinity cf. \cite{Sauzin}.\footnote{It is also required
1743: that certain relations, called bridge equations, hold; in our
1744: case it would be easy to derive them from the
1745: explicit form of $H$, but we omit this calculation.} By abuse of language,
1746: $f(x)$ is called resurgent if it satisfies the requirements in
1747: Definition \ref{def1} and {\bf all} $Y_k$ are resurgent.
1748: \end{Definition}
1749: This is especially useful when global information about $f$ for
1750: $x\in\CC$ is needed: deformation of contours in $p$, and collecting
1751: residues when/if singularities are crossed, provides a straightforward
1752: way to obtain this information.
1753:
1754:
1755:
1756:
1757: \begin{thebibliography}{99}
1758: \bibitem{Abramowitz} M Abramowitz and I A Stegun, Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, New York: Dover (1965).
1759: \bibitem{benderorszag} C Bender and S Orszag, {\em Advanced
1760: Mathematical Methods for scientists and engineers}, McGraw-Hill, 1978, Springer-Verlag 1999.
1761:
1762: \bibitem{Beardon} A F Beardon, {\em Iteration of Rational Functions} Springer Verlag, New York (1991).
1763: \bibitem{OCtransseries}O. Costin, Topological construction of transseries and introduction to generalized Borel summability (AMS CONM series vol. 373, 137--177)
1764: \bibitem{CPAM} O. Costin and M. D. Kruskal, {\em Analytic methods
1765: for obstruction to integrability in discrete dynamical systems}
1766: Comm. Pure Appl. Math. 58 no. 6, 723--749) (2005).
1767: \bibitem{Inventiones} O. {}Costin and R D Costin, Invent. Math, 145, 3, pp 425-485 (2001).
1768:
1769:
1770: \bibitem{Devaney} R L Devaney, An Introduction to Chaotic Dynamical Systems, 2nd Edition, Westview Press (2003).
1771: \bibitem{Douady} A Douady and J Hubbard {\em On the dynamics of polynomial-like mappings}, Ann. Sci. Ec. Norm. Sup. 18 pp. 287–344 (1985).
1772:
1773:
1774:
1775:
1776: \bibitem{Ecalle} J Ecalle, Les fonctions resurgentes, vol. I, II and III, Publ. Math. Orsay, 1985.
1777:
1778: \bibitem{Jacobi} D C G Iacobo. Fundamenta Nova Theoriae Functionum Ellipticarum.
1779: Konigsberg:, Sumtibus Fratrum Borntraeger, 1829.
1780: \bibitem{Ince} E L Ince, Ordinary Differential Equations, Dover Publications (1956).
1781: \bibitem{Kawai1} T. Kawai, {Amer. J. Math.}, 109, 57--64 (1989).
1782: \bibitem{Kawai} T. Kawai, in {\em Towards the Exact WKB Analysis of Differential
1783: Equations, Linear or Nonlinear}, C.J. Howls, T. Kawai and Y. Takei, eds., Kyoto
1784: Univerisity Press, Kyoto, pp. 231--244, (2000).
1785:
1786: \bibitem{Mahler} K
1787: Mahler,
1788: {\em An unsolved problem on the powers of $3/2$}.
1789: J. Austral. Math. Soc. 8 1968 313--321.
1790: \bibitem{Mandelbrojt} S. Mandelbrojt {\em S\'eries lacunaires},
1791: Actualit\'es scientifiques et industrielles, Paris, 305 (1936).
1792: \bibitem{Milnor} J. Milnor {\em Dynamics in one complex variable}, Annals of Mathematics
1793: studies Nr. 160, Princeton (2006).
1794: \bibitem{Montgomery} H. L. Montgomery, {\em Ten lectures on the interface between
1795: Analytic Number Theory and Harmonic Analysis}, Providence, R.I. : Published for the Conference Board of the Mathematical Sciences by the American Mathematical Society, (1994).
1796: \bibitem{xaos} Project Xaos, Gnu freeware, http://sourceforge.net/projects/xaos.
1797: \bibitem{Sauzin} D. {}Sauzin {\em Resurgent functions and splitting problems}
1798: RIMS Kokyuroku 1493 (31/05/2006) 48-117, \begin{verbatim}{http://arxiv.org/abs/0706.0137v1.}\end{verbatim}
1799: \bibitem{Titchmarsh} E C Titchmarsh, {\em The Theory of the Riemann Zeta-Function}, Oxford University Press, (1987).
1800:
1801:
1802: \end{thebibliography}
1803: \end{document}
1804: -(1/2)*E(Pi^(1/2)*gamma^(1/2)*(N-xN)*(-i)^(1/2))/(gamma^(1/2)*(-i)^(1/2))
1805:
1806:
1807:
1808: 30B50, 30B10, 30B30,40G10,34E05, 37F10,37F50