math0211386/g2.tex
1: 
2: \documentclass[12pt]{article}
3: \usepackage{latexsym,amssymb,psfig}
4: 
5: %\documentstyle[12pt]{article}
6: \textheight 9in
7: \textwidth 6in
8: \font\of=msbm10 scaled 1200
9: \def\R{\mbox{\of R}}
10: \def\C{\mbox{\of C}}
11: \def\Z{\mbox{\of Z}}
12: \def\N{\mbox{\of N}}
13: \renewcommand\arraystretch{1.3}
14: \hoffset =-1cm
15: \voffset =-2cm
16: 
17: \newtheorem{definition}{Definition}
18: \newtheorem{remark}{Remark}
19: \newtheorem{theorem}{Theorem}%[section]
20: \newtheorem{lemma}{Lemma}
21: \newtheorem{proposition}{Proposition}
22: \newtheorem{corollary}{Corollary}
23: 
24: \title{ Complete hyperelliptic integrals of the first kind and their
25: non-oscillation }
26: 
27: \author{Lubomir Gavrilov \\
28: \normalsize \it Laboratoire Emile Picard , CNRS UMR 5580,  Universit\'e
29: Paul Sabatier\\
30: \normalsize \it 118, route de Narbonne, 31062 Toulouse Cedex, France \\
31: \\ Iliya D. Iliev\\
32: \normalsize \it Institute of Mathematics, Bulgarian Academy of Sciences\\
33: \normalsize \it P.O. Box
34: 373, 1090 Sofia, Bulgaria } \date{}
35: 
36: \begin{document}
37: \maketitle
38: \begin{abstract}
39: Let $P(x)$ be a real polynomial of degree $2g+1$, $H=y^2+P(x)$ and
40: $\delta(h)$ be an oval contained in the level set $\{H=h\}$.
41: We study complete Abelian integrals of the form
42: $$I(h)=\int_{\delta(h)}
43: \frac{(\alpha_0+\alpha_1 x+\ldots + \alpha_{g-1}x^{g-1})dx}{y},
44: \;\;h\in \Sigma,$$
45: where $\alpha_i$ are real and $\Sigma\subset \R$ is a maximal open interval
46: on which a continuous family of ovals $\{\delta(h)\}$ exists. We show that
47: the $g$-dimensional real vector space of these integrals is not Chebyshev
48: in general: for any $g>1$, there are hyperelliptic Hamiltonians $H$ and
49: continuous families of ovals $\delta(h)\subset\{H=h\}$, $h\in\Sigma$,
50: such that the Abelian integral $I(h)$ can
51: have at least $[\frac32g]-1$ zeros in $\Sigma$. Our main result is Theorem
52: \ref{main} in which we show that when $g=2$, exceptional families of ovals
53: $\{\delta(h)\}$ exist, such that the corresponding vector space is still
54: Chebyshev.
55: \end{abstract}
56: 
57: \section{Introduction}
58: 
59: 
60: 
61: 
62: 
63: 
64: Take real polynomials  $H,f,g\in \R[x,y]$ and let $\delta (h) \subset
65: \{(x,y)\in \R^2: H(x,y)=h\}$, $h\in\Sigma$,
66: be a continuous family of ovals.
67: For  sufficiently small $\varepsilon$ and generic $H, f, g,$
68: the limit cycles of the perturbed plane Hamiltonian system
69: $$
70: dH + \varepsilon (f dx + g dy) = 0
71: $$
72: which tend to certain ovals from the continuous
73: family when $\varepsilon\to 0$, are in one-to-one
74: correspondence with the zeros of the complete Abelian integral
75: $$
76: I(h) = \int_{\delta (h)} f(x,y) dx + g(x,y) dy, \quad h\in\Sigma.
77: $$
78: For this reason the problem of finding the zeros of $I(h)$ in terms of the
79: degrees of $H,f,g$
80: was called  by Arnold \cite[p. 313]{Arnold1} the
81: ``{\it weakened 16th Hilbert problem}"  (compare to Hilbert \cite{hil}, see
82: also Arnold \cite{Arnold2,Arnold1+,ten}).
83: Note that the level sets $ \{H=h\} $ will contain in general several
84: continuous families of ovals which need be considered separately.
85: 
86: It follows from the Varchenko-Khovanskii theorem that the number of the
87: zeros of $I(h)$ is bounded
88: by a constant $N(n,d)<\infty$, uniformly in all $H,f,g$, such that
89: $deg\,f\leq d$,  $deg\,g \leq d$,  $deg\,H\leq n$.
90: 
91: In the so called ``hyperelliptic case" ($H=y^2+P(x)$) it was proved by
92: Novikov and Yakovenko\cite{nov-yak} that there exists an algorithm producing
93: a function $C(n,d)$ such that $\infty > C(n,d)> N(n,d)$,
94: which is given by a tower function (an iterated exponent) of
95: height at least five. Their
96: proof is based on the analytic properties of a suitable Picard-Fuchs system
97: satisfied by Abelian
98: integrals (including the magnitude of the coefficients of the system and some
99: restrictions on its monodromy group).
100: 
101: The progress in solving the weakened 16th Hilbert problem (finding the
102: {\it exact} number of the zeros of
103: $I(h)$) concerned so far the ``elliptic case" only (the complex algebraic
104: curve $\{H=h\}$
105: is of genus at most one). It was proved for instance
106:  that in several cases,
107: the vector space ${\cal A}_{H,d}$
108: of Abelian integrals of degree $d$ polynomials along the ovals of $H$,
109: obeys the so called {\it Chebyshev property} (the number
110: of the zeros of each integral is smaller than the dimension of the vector
111: space ${\cal A}_{H,d}$), see \cite{Petrov86,gav98,gav99,gas-li}.
112: In this relation Arnold asked in \cite[the 7th problem]{ten} whether the
113: $g$-dimensional vector
114: space of Abelian integrals
115: \begin{equation}
116: \label{question}
117: I(h)=\int_{\delta(h)}
118: \frac{(\alpha_0+\alpha_1 x+\ldots + \alpha_{g-1}x^{g-1})dx}{y},
119: \;\;h\in \Sigma
120: \end{equation}
121:  where
122: $
123: \delta (h)\subset \{(x,y)\in \R^2:\; y^2+P(x)=h\},\;\;g=[\frac12(deg\, P-1)]>1,
124: $
125: is Chebyshev.
126: 
127: In an attempt to solve this problem Givental \cite{giv} obtained a
128: non-oscillation theorem for Lagrangian
129: planes of the Picard-Fuchs system satisfied by the Abelian integrals
130: $\int_{\delta(h)}x^idx/y$. He
131: used the fact that every Picard-Fuchs system has a Hamiltonian (or Poisson)
132: structure. He failed,
133: however, to produce bounds for the zeros of the Abelian integrals.
134: 
135: In the present paper we try to explore the algebro-geometrical properties
136: of the Abelian integrals,
137: without using the Picard-Fuchs system. Our main result is Theorem
138: \ref{main} in which we show that
139: when $g=2$ and $deg\,P=5$, exceptional families of ovals
140: $\{\delta(h)\}$, $ h\in \Sigma$, exist such that every Abelian integral of
141: the form
142: $$
143: I(h)=\int_{\delta(h)}
144: \frac{(\alpha_0+\alpha_1 x)dx}{y},\quad \alpha _0^2+\alpha _1^2\neq 0
145: $$
146: has at most one simple zero on the  interval $\Sigma$. At a first sight
147: this seems to be an easy
148: observation (equivalent to the claim that $(\int_{\delta(h)}
149: xdx/y)/(\int_{\delta(h)}dx/y)$ is
150: monotonous on $\Sigma$). This is, however, the first result of such a kind
151: for non-elliptic curves
152: $\{H=h\}$. The proof uses the Riemann bilinear relations on differentials
153: of the first kind together
154: with  the fact that a Jacobian variety with its polarization cannot be a
155: direct product of
156: principally polarized Abelian varieties. Our arguments can be adapted to
157: other situations
158: (for instance when the degree of $P(x)$ is $6$), but we shall not do
159: this here.
160: 
161: We give also a negative answer (Proposition \ref{nocheb}) to the initial
162: question posed by Arnold:
163: it turns out that there exist Abelian integrals of the form
164: (\ref{question}) with exactly
165: $[\frac32g]-1$ zeros in a neighborhood of the origin.
166: 
167: 
168: 
169: \section{The cyclicity at a center}
170: In this section we prove that in general, the integral (\ref{question})
171: does not belong to a Chebyshev space. Consider first the particular case
172: when $n=5$. Let $H= y^2+ P(x)$ where $P(x)= x^2(1+a_1 x +a_2 x^2+a_3 x^3)$,
173: $a_1,a_2, a_3 \in \R$, $a_3\neq 0$, and denote by
174: $\delta (t) \subset \{y^2+P(x)=t\}$  a continuous family of cycles vanishing
175: at the origin as $t$ tends to zero.
176: 
177: 
178: \begin{proposition}
179: \label{cyclicity}
180: The Abelian integral
181: $$
182: I(t) = \int_{\delta (t)} \frac{(\alpha_0+\alpha_1 x) dx}{y}=
183: \alpha_0 I_0(t)+\alpha_1 I_1(t)\not\equiv 0
184: $$
185: can have a zero at the origin of multiplicity at most two. Moreover, for
186: fixed $\alpha_1, a_2, a_3$, $\alpha_1 a_3\neq 0$, there exist a sufficiently
187: small
188: $\varepsilon >0$  and $\alpha_0$, $a_1$ in a small neighborhood of zero,
189: such that $I(t)$ has exactly two simple zeros in the interval
190: $(0,\varepsilon )$.
191: \end{proposition}
192: The above shows that the real vector space generated by the functions
193: $I_0(t)$, $I_1(t)$ is not Chebyshev. Recall that a real vector space $V$ of
194: functions defined on some interval $\Sigma $ is said to be Chebyshev,
195: provided that each $f\in V$ has at most $\dim V - 1$ zeros (counted with
196: multiplicity) on $\Sigma $, and Chebyshev with accuracy $m$, if
197: each $f\in V$ has at most $\dim V+m - 1$ zeros there.
198: 
199: 
200: \vspace{2ex}
201: \noindent
202: {\bf Proof of Proposition \ref{cyclicity}.}
203: For a small $x$, denote $X=x(1+a_1 x +a_2 x^2+a_3 x^3)^{1/2}$.
204: Then an easy calculation yields the inverse transformation
205: $$\textstyle
206: x=\varphi(X)\sim X-\frac12a_1X^2+(\frac58a_1^2-\frac12a_2)X^3
207: -(a_1^3-\frac32a_1a_2+\frac12a_3)X^4+\ldots \quad \mbox{\rm as}\quad
208: X\to 0.$$
209: Therefore, for small positive $t$,
210: $$\begin{array}{l}{\displaystyle
211: I(t)=\int_{y^2+X^2=t}\frac{[\alpha_0+\alpha_1\varphi(X)]
212: \varphi'(X)}{y}dX}\\[5mm]
213: {\displaystyle
214: = \int_{y^2+X^2=t} \frac{\alpha_0[1+O(X^2)] - \alpha_1[\frac32a_1X^2
215: +\frac52(\frac{21}{8}a_1^3-\frac72a_1a_2+a_3)X^4 +O(X^6)]}{y}dX}\\[5mm]
216: =2\pi\alpha_0[1+O(t)]-2\pi\alpha_1[\frac34a_1t+
217: \frac{15}{16}(\frac{21}{8}a_1^3- \frac72a_1a_2+a_3)t^2 +O(t^3)].
218: \end{array}$$
219: When $\alpha_0=a_1=0$, the function $I(t)$ has a double zero at the
220: origin since $\alpha_1 a_3\neq 0$.
221: Taking $\alpha_1=1$, $|\alpha_0|<\!\!<|a_1|<\!\!<|a_3|$ and
222: $\alpha_0 a_1>0>a_1a_3$, one finishes the proof of the proposition. $\Box$
223: 
224: Clearly, the above approach could be applied to hyperelliptic Hamiltonians
225: of any degree that have a center.
226: 
227: Denote by ${\cal H}_n$ the set of Hamiltonians
228: $H=y^2+x^2+a_1 x^3+\ldots +a_{n-2}x^n$, where $a_k\in \R$, $a_{n-2}\neq 0$.
229: Denote by $g=[\frac12(n-1)]\geq 2$ the genus of the hyperelliptic curve.
230: Given $H\in {\cal H}_n$, let $\delta(t)$, $t\in (0, T_H)$, be an oval from
231: the continuous family of ovals surrounding the center at the origin.
232: Consider the vector space ${\cal A}_H$ formed by the Abelian integrals
233: 
234: 
235: \begin{equation}
236: \label{g}
237: I(t) = \int_{\delta (t)} \frac{G(x) dx}{y},\quad t\in(0,T_H),
238: \quad G\in \R[x],\quad deg\, G<g.
239: \end{equation}
240: Let us point out that $dim\,{\cal H}_H=g$ in general, but
241: $dim\,{\cal H}_H=[\frac12(g+1)]$ provided that $H$ contains no odd degree
242: monomials. In the proposition below, we give a partial answer to the question
243: raised by V. Arnold'd in \cite{ten}.
244: 
245: 
246: 
247: \begin{proposition}
248: \label{nocheb}
249: There exist $H\in{\cal H}_n$ and a polynomial $G$ such that the
250: corresponding integral $(\ref{g})$ has exactly $[\frac32g]-1$ small positive
251: simple zeroes. Thus, such a space ${\cal A}_H$ could be Chebyshev with
252: accuracy at least $[\frac12g]$ in $(0,T_H)$.
253: \end{proposition}
254: 
255: 
256: 
257: Given an integer $k\geq 0$ and $H\in{\cal H}_n$, denote
258: $$I_k(t)=\int_{\delta(t)}\frac{x^kdx}{y}, \quad 0<t<T_H.$$
259: The proof of the above proposition is based on the following lemma.
260: 
261: 
262: 
263: \begin{lemma}
264: \label{int} The following asymptotic expansions hold for
265: small positive $t$:
266: 
267: \vspace{1ex}
268: \noindent
269: {\em(i)}$\quad I_{2k}(t)= t^k[c_k+O(t)],\;\;\mbox{\it where} \;\;
270: c_k=2\pi\frac{(2k-1)!!}{(2k)!!},$
271: 
272: \vspace{1ex}
273: \noindent
274: {\em(ii)}$\quad\displaystyle I_{2k+1}(t)=-{\textstyle\frac12}
275: \sum\limits_{j=1}^g a_{2j-1}(2k+2j+1)t^{k+j}[c_{k+j}+O(t)],$
276: 
277: \vspace{1ex}
278: \noindent
279: therefore the coefficients in the last expansion belong to the ideal
280: generated by the odd-numbered parameters $\{a_1, a_3,\ldots, a_{2g-1}\}$.
281: \end{lemma}
282: 
283: 
284: 
285: \noindent
286: {\bf Proof.} Take as above $X=x(1+a_1 x +\ldots+ a_{n-2} x^{n-2})^{1/2}$,
287: then the inverse transformation $x=\varphi(X,a_1,a_2,\ldots,a_{n-2})$
288: for small $X$ reads $ x=X+A_1X^2+A_2X^3 +A_3X^4+\ldots$,
289: where the coefficients $A_j=A_j(a_1,a_2,\ldots,a_{n-2})$ are polynomials.
290: It is easy to see that if one attaches a weight $k$ to the coefficient $a_k$,
291: then $A_j$ is a polynomial of weight $j$. Indeed, directly from the definition
292: of $\varphi$ it follows that
293: $$\varphi(\omega X,a_1,a_2,\ldots,a_{n-2})=\omega\varphi(X, \omega a_1,
294: \omega^2 a_2,\ldots, \omega^{n-2}a_{n-2}).$$
295: Applying this identity to
296: $$\varphi(X, a_1, \ldots, a_{n-2})=X(1+\sum_{k=1}^\infty A_k(a_1,\ldots,
297: a_{n-2})X^k),$$
298: we get $A_k(\omega a_1, \ldots, \omega^{n-2} a_{n-2})=\omega^k
299: A_k(a_1, \ldots, a_{n-2})$.
300: To calculate the coefficient at $a_k$ in $A_k$, we take $a_1=\ldots=a_{k-1}=0$,
301: and easily obtain that it equals $-\frac12$. Therefore, one deduces that
302: $A_k=-\frac12a_k+A_k^*$ where $A_k^*$ belongs to the polynomial ideal
303: generated by $a_1,\ldots, a_{k-1}$.
304: 
305: In case (i), one immediately obtains that for small positive $t$,
306: $$I_{2k}(t)=\frac{1}{2k+1}\int_{y^2+X^2=t}\frac{d\varphi^{2k+1}(X)}{y}=
307: \int_{y^2+X^2=t}\frac{(X^{2k}+\ldots)dX}{y}=c_kt^k+O(t^{k+1}).
308: $$
309: To calculate the integral in case (ii), we write $\varphi^{2k+2}(X)$
310: in the form
311: $$\varphi^{2k+2}=X^{2k+2}\left(1+\sum\limits_{j=1}^\infty
312: (-{\textstyle\frac12}a_j+A_j^*)X^j\right)^{2k+2}$$
313: $$=X^{2k+2}\left(1+\sum\limits_{j=1}^\infty
314: [(2k+2)(-{\textstyle\frac12}a_j+A_j^*)X^j+O(X^{j+1})]\right)$$
315: $$=X^{2k+2}\left(1-\sum\limits_{j=1}^{n-2} (k+1)a_jX^j[1+O(X)]\right)$$
316: $$=-(k+1)\sum\limits_{j=1}^g a_{2j-1}X^{2k+2j+1}[1+O(X^2)]+ \Phi(X)$$
317: where $\Phi(X)$ is a series in $X$ containing only even order powers.
318: Therefore,
319: $$I_{2k+1}(t)=\frac{1}{2k+2}\int_{y^2+X^2=t}\frac{d\varphi^{2k+2}(X)}{y}$$
320: $$=-{\textstyle\frac12}\sum\limits_{j=1}^g a_{2j-1}(2k+2j+1)t^{k+j}
321: [c_{k+j}+O(t)].\;\;\Box$$
322: 
323: \vspace{2ex}
324: \noindent
325: {\bf Proof of Proposition \ref{nocheb}.} Denote $m=[\frac12g]$ and take
326: in (\ref{g})
327: $$G(x)=\sum_{k=0}^{m-1}\gamma_{2k}x^{2k}-x^{2m-1}.$$
328: Then by Lemma \ref{int},
329: $$I(t)=\sum_{k=0}^{m-1}\gamma_{2k}t^k[c_k+O(t)]+{\textstyle\frac12}
330: \sum_{j=1}^g a_{2j-1}(2m+2j-1)t^{m+j-1}[c_{m+j-1}+O(t)].$$
331: Now, freeze the even-numbered coefficients in $H$ and then choose $\gamma_{2j}$
332: and $a_{2j-1}$ so that $\gamma_{2j}\gamma_{2j+2}<0$,
333: $a_{2j-1}a_{2j+1}<0$, $\gamma_{2m-2}a_1<0$ and
334: $$|\gamma_0|<\!\!<|\gamma_2|<\!\!<\ldots<\!\!<|\gamma_{2m-2}|
335: <\!\!<|a_1|<\!\!<\ldots <\!\!<|a_{2g-1}|<\!\!<1.$$
336: Such a choice guarantees that $I(t)$ will have exactly $m+g-1=[\frac32g]-1$
337: zeroes in a certain small interval $(0,\varepsilon)$. $\Box$
338: 
339: This result raises the problem to describe the hyperelliptic Hamiltonians
340: $H$ and the continuous families of ovals $\{\delta(h)\}$ for which the
341: space of integrals (\ref{question}) obeys the Chebyshev property.
342: In the rest of the paper, we concentrate our efforts on the simplest case
343: when $deg\,P=5$. We begin with some preliminaries.
344: 
345: \section{The normal form, the bifurcation diagram and the Dynkin diagram
346: for $n=5$}
347: We consider the hyperelliptic Hamiltonian $H=\frac12y^2+P(x)$
348: where $P\in \R[x]$ and $deg\,P=5$. Assume that a certain level set
349: $\{H=h\}$ contains an oval (compact closed smooth curve without critical
350: points). As the Poincar\'e index of the respective Hamiltonian vector
351: field $X_H$ is zero, then $X_H$ has at least two different real critical
352: points. Without any loss of generality, we can place all the four critical
353: points at $(0,0)$, $(\mu, 0)$, $(\lambda,0)$ and $(1,0)$, where either
354: $0\leq\mu\leq\lambda\leq 1$ (the real case)
355: or $\bar\mu=\lambda\in\C\setminus\R$ (the complex case).
356: Finally, we can assume that the coefficient at $x^5$ is positive.
357: Then after rescaling of the $y$ variable, we come to the normal form
358: \begin{equation}
359: H(x,y)=\frac12y^2-\frac{\lambda\mu}{2}x^2+
360: \frac{\lambda+\mu+\lambda\mu}{3}x^3-\frac{1+\lambda+\mu}{4}x^4+\frac15x^5
361: \equiv \frac12y^2+P(x)
362: \label{normal}
363: \end{equation}
364: and the Hamiltonian flow $X_H$ takes the form
365: \begin{equation}
366: \label{system}
367: \begin{array}{l}
368: \dot{x}=y,\\
369: \dot{y}=-x(x-\mu)(x-\lambda)(x-1).
370: \end{array}
371: \end{equation}
372: Until the end of the paper, we will use this normal form of $H$ only.
373: 
374: Clearly, the origin is a hyperbolic saddle if $\mu\neq 0$, and $(1,0)$ is
375: a nondegenerate center provided that $\lambda\neq 1$. In addition,
376: when $0<\mu<\lambda<1$ in the real case, $(\mu, 0)$ is a center
377: and $(\lambda, 0)$ is a saddle. Denote by $h_0, h_\mu, h_\lambda$ and $h_1$
378: the corresponding critical levels of $H$. One has
379: $$\begin{array}{l}
380: h_0=0,\quad h_1=-\frac{1}{60}(3-5\lambda-5\mu+10\lambda\mu),\\[2mm]
381: h_\lambda=-\frac{\lambda^3}{60}(3\lambda^2-5\lambda\mu-5\lambda+10\mu),\\[2mm]
382: h_\mu=-\frac{\mu^3}{60}(3\mu^2-5\lambda\mu-5\mu+10\lambda). \end{array}$$
383: If $0<\mu<\lambda<1$, we have $h_0>h_\mu$, $h_\lambda>h_\mu$, $h_\lambda>h_1$.
384: The bifurcation diagram $D$ in the real case consists of the boundary
385: of the triangle $T=\{0\leq\mu\leq\lambda\leq 1\}$ and the curve $\gamma:\;
386: h_\lambda=h_0$ having an equation
387: $$\mu=\frac{3\lambda^2-5\lambda}{5(\lambda-2)},$$
388: see Fig. \ref{bdrc}.
389: 
390: \begin{figure}[htbp]
391: \begin{center}
392:   \psfig{file=save1.eps}
393: \end{center}
394: \caption{ Bifurcation diagram in the $(\lambda,\mu)$-plane in the real case}
395: \label{bdrc}
396: \end{figure}
397: 
398: There are two open components in $T\setminus D$ giving the
399: two generic cases. Namely, for parameters $\lambda, \mu$ above the
400: curve $\gamma$, $X_H$ has three period annuli, and below this curve $X_H$
401: has just two period annuli. There are also five codimension-one and four
402: codimension-two degenerate (nongeneric) cases corresponding
403: to the parameters on the bifurcation diagram $D$. In the complex case,
404: there is a unique period annulus around $(1,0)$ which terminates
405: at the saddle-loop through $(0,0)$ and the phase portraits for all
406: $\lambda\in\C\setminus\R$ are topologically equivalent. We will denote
407: the period annuli (continuous families of ovals) around $(1,0)$, $(\mu,0)$
408: and the eight loop by ${\cal O}_1$, ${\cal O}_\mu$ and ${\cal O}_e$,
409: respectively.
410: 
411: Below we describe the Dynkin diagram of the real polynomial $H(x,y)$
412: when the critical levels are distinct. Recall that it is a graph
413: defined in the same way as the Dynkin diagram (or D-diagram)
414: of a germ of an analytical function with an isolated singularity, see
415: \cite{Arnold,dim}. The vertices of the diagram are in one-to-one
416: correspondence with the vanishing cycles of the polynomial. Two
417: vertices are connected by an edge if the intersection number of the
418: cycles is not zero. To an edge one associates also a label equal
419: to the intersection number, but it will be omitted here because is of no
420: importance for us. The vanishing cycles ``vanish" along suitable paths,
421: and hence the Dynkin diagram depends also on the family of paths. The
422: precise definition in our situation is as follows.
423: 
424: Consider first the real case $0<\mu<\lambda<1$. There are five possible
425: distributions of the critical values as follows:
426: $$\begin{array}{lll}
427: h_1 < h_\mu < h_\lambda < h_0, &
428: h_1 < h_\mu < h_0 < h_\lambda, &
429: h_\mu < h_1 < h_0 < h_\lambda, \\
430: h_\mu < h_0 < h_1 < h_\lambda, &
431: h_\mu < h_1 < h_\lambda < h_0. &
432: \end{array}$$
433: For definiteness, assume that $h_1<h_\mu < h_\lambda < h_0$.
434: Denote ${\cal D} = \C \backslash [h_\lambda , \infty)$ and let
435: $$
436: l_0,l_1,l_\lambda,l_\mu:\; [0,1] \rightarrow \{h\in \C: Im\,h \geq 0\}
437: $$
438: be continuous paths connecting some fixed regular value
439: $\tilde{h} \in {\cal D}$ to
440: $h_0$, $h_1$, $h_\lambda$, $h_\mu$ respectively, as shown on
441: Fig. \ref{realdynk}.
442: 
443: % \begin{figure}
444: % \setlength{\unitlength}{.03\textwidth}
445: %
446: % \begin{picture}(30,15)(0,0)
447: % \put(2,8){
448: % \put(5,0){\vector(1,0){15}}
449: % \put(12,-7){\vector(0,1){15}}
450: % \qbezier[200](9,-6,)(14.96,0)(9,6)
451: % \put(2,0){\qbezier[200](15,-6,)(9.04,0)(15,6)}
452: % \put(20.5,-.3){$Re \lambda$}
453: % \put(11,8.5){$Im \lambda$} }
454: % \end{picture}
455: % \caption{The hyperbola  $\Gamma:\; h_\lambda=h_\mu$ in the complex case}
456: % \label{fhyp}
457: % \end{figure}
458: 
459: The Dynkin diagram of $H$ with respect to the above paths
460: is a graph with four vertices corresponding to the
461: four families of cycles vanishing along the paths
462: $l_0$, $l_1$, $l_\lambda$, $l_\mu$, which we denote by
463: $\delta_0(h)$, $\delta_1(h)$, $\delta_\mu (h)$, $\delta_\lambda (h)$
464: respectively. Two vertices are connected by an
465: edge if the intersection number of the cycles for $h=\tilde{h}$
466: is not zero (in which case it is $\pm 1$ and may be supposed equal
467: to one). The computation of the Dynkin diagram of a real polynomial
468: with real critical points only is well known, see for instance
469: \cite{cam,Arnold}. It yields the diagram shown on Fig. \ref{realdynk}.
470: We note that the computation depends only on the distribution of the
471: four critical points $0<\mu<\lambda<1$ and not on the distribution of
472: the respective critical values (hence, the Dynkin diagram is the same
473: for all possible arrangements of the critical values
474: with corresponding paths $l_0$, $l_1$, $l_\lambda$, $l_\mu$).
475: 
476: Consider now the complex case when $\lambda = \bar{\mu}\not\in\R $.
477: The method we use is a deformation of $H$ to a polynomial having real
478: critical points. Denote ${\cal D}=\C\backslash [h_0,\infty)$ and let
479: $$
480: l_0, l_1, l_\lambda, l_\mu:\; [0,1] \rightarrow {\cal D}
481: $$
482: be continuous paths connecting some fixed regular value
483: $\tilde{h} \in {\cal D}$ to $h_0$, $h_1$, $h_\lambda$, $h_\mu$
484: respectively, as shown on Fig. \ref{dynkin1}. The Dynkin diagram
485: related to these paths is defined as above. Consider in
486: the complex plane the hyperbola $\Gamma$ defined by the equation
487: \begin{equation}
488: \label{hyperbola}
489: \Gamma:\;\;
490: h_\lambda=h_\mu\quad \Leftrightarrow\quad (Im\,\lambda)^2=
491: 5Re\,\lambda(Re\,\lambda-1)
492: \end{equation}
493: and drawn on Fig. \ref{imconj}.
494: 
495: 
496: \begin{figure}
497: \setlength{\unitlength}{.03\textwidth}
498: \begin{picture}(30,15)(0,0)
499: %critical values
500: \put(0,5){
501:  \begin{picture}(15,15)(0,0)
502:  \put(0,0){\circle*{.3}} \put(0,0.5){$h_1$}
503:  \put(3,0){\circle*{.3}}  \put(3,0.5){$h_\mu$}
504:  \put(6,0){\circle*{.3}} \put(6,0.5){$h_\lambda$}
505:  \put(9,0){\circle*{.3}} \put(9,0.5){$h_0$}
506:  \put(6,0){\thicklines{\vector(1,0){6}}}
507:  %\put(9,0){\thicklines{\vector(1,0){3}}}
508: 
509:  \qbezier[20](0,0)(0,5)(6,6)
510:  \qbezier[20](3,0)(3,2)(6,6)
511:  \qbezier[20](6,0)(5,3)(6,6)
512:  \qbezier[20](9,0)(7.5,2.5)(6,6)
513:  \put(6,6){\circle*{.3}} \put(6.5,6){$\tilde{h}$}
514:  \end{picture}}
515: %Dynkin diagram
516: \put(20,5){
517:  \begin{picture}(15,15)(0,0)
518:  \put(0,0){\line(1,0){3}} \put(0,0){\circle*{.3}}
519:  \put(3,0){\line(1,0){3}} \put(3,0){\circle*{.3}}
520:  \put(6,0){\line(1,0){3}} \put(6,0){\circle*{.3}}
521:  \put(9,0){\circle*{.3}}
522: 
523:  \put(0,.5){$\delta _1$}
524:  \put(2.5,.5){$\delta _\lambda$}
525:  \put(5.5,.5){$\delta _\mu$}
526:  \put(8.5,.5){$\delta _0$}
527: 
528:  \end{picture}
529:               }
530: \put(7,0){ (i)}
531: \put(22,0){ (ii)}
532: \end{picture}
533: \caption{
534: (i) The paths $l_0$, $l_1$, $l_\lambda$, $l_\mu$ in the real case
535: $0<\mu<\lambda<1$ when $h_1<h_\mu<h_\lambda<h_0$;
536: (ii) Dynkin diagram in the real case $0<\mu <\lambda <1$ }
537: \label{realdynk}
538: \end{figure}
539: 
540: 
541: Suppose for definiteness that $\lambda$ lies inside its left branch.
542: First, by continuous deformation of $\lambda = \bar{\mu }$ such that
543: in the course of the deformation $\lambda $ remains in the left branch
544: of (\ref{hyperbola}), we may achieve that $\lambda =\mu<0<1$.
545: This, combined with $h_\lambda =h_\mu <0$ already implies that
546: \begin{itemize}
547: \item
548: the intersection number of $\delta _1(\tilde{h})$ and $\delta _0(\tilde{h})$
549: is equal to $\pm 1$;
550: \item
551: the intersection number of $\delta _\lambda (\tilde{h})$,
552: $\delta _\mu  (\tilde{h})$ with $\delta _1 (\tilde{h})$ is zero;
553: \item
554: the intersection number of $\delta _\lambda (\tilde{h})$ and
555: $\delta _\mu  (\tilde{h})$ is equal to $\pm 1$;
556: \item
557: The intersection number of either $\delta _\lambda (\tilde{h})$ or
558: $\delta _\mu(\tilde{h})$ with $\delta _0(\tilde{h})$ is equal
559: to $\pm 1$.
560: \end{itemize}
561: Finally, we may use the real structure of the complex algebraic  curve
562: $$\Gamma _h = \{(x,y)\in \C^2 :\; H(x,y) =h\}$$
563: for $h\in \R$.
564: This structure is defined by an antiholomorphic involution $i$ on the
565: curve which is the complex conjugation. It induces on its hand an
566: involution $i_*$ of $H_1(\Gamma _h, \Z)$, $(i_* )^2=id$.
567: It is easy to see that for $h<0$ we have
568: $$
569: i_* \delta _\lambda(h) = \delta _\mu(h) ,\quad
570: i_* \delta _0(h) = - \delta _0(h) .
571: $$
572: This combined with the identity
573: $$
574: \overline{\langle \delta (h) , \gamma (h)\rangle} =
575: - \langle i_*\delta (h) , i_*\gamma (h)\rangle
576: $$
577: shows that
578: $$\langle \delta_\lambda  (h) , \delta _0 (h) \rangle=
579: \langle \delta_\mu   (h) , \delta _0 (h)\rangle.
580: $$
581: This implies the Dynkin diagram as shown on Fig. \ref{dynkin1}(ii)).
582: The remaining cases shown there are studied in a similar way.
583: 
584: \begin{figure}
585: \setlength{\unitlength}{.03\textwidth}
586: 
587: \begin{picture}(30,15)(0,0)
588: %critical values
589: \put(0,5){
590:    \begin{picture}(15,15)(0,0)
591: \put(0,0){\circle*{.3}} \put(0,0.5){$h_1$}
592: \put(3,0){\circle*{.3}}  \put(3,0.5){$h_0$}
593: \put(1.5,1.5){\circle*{.3}} \put(1.5,2){$h_\lambda$}
594: \put(1.5,-1.5){\circle*{.3}} \put(2,-1.5){$h_\mu$}
595: \put(3,0){\thicklines{\vector(1,0){9}}}
596: 
597: \qbezier[20](0,0)(0,5)(6,6)
598: \qbezier[20](3,0)(3,2)(6,6)
599: \qbezier[20](1.5,1.5)(4,5)(6,6)
600: \qbezier[25](1.5,-1.5)(3,3.5)(6,6)
601: \put(6,6){\circle*{.3}} \put(6.5,6){$\tilde{h}$}
602:    \end{picture}}
603: %Dynkin diagram
604: \put(20,5){
605:    \begin{picture}(15,15)(0,0)
606: \put(0,0){\line(1,0){3}} \put(0,0){\circle*{.3}}
607: \put(3,0){\line(1,1){2.1213}} \put(3,0){\circle*{.3}}
608: \put(3,0){\line(1,-1){2.1213}}
609: \put(5.1213,2.1213){\circle*{.3}}
610: \put(5.1213,-2.1213){\circle*{.3}}
611: 
612: \put(2.5,0.5){$\delta _0$}
613: \put(0,0.5){$\delta _1$}
614: \put(5.6213,2.1213){$\delta _\lambda$}
615: \put(5.6213,-2.1213){$\delta _\mu$}
616: 
617: \put(5.1213,-2.1213){\line(0,1){4.2426}}
618:    \end{picture}
619:               }
620: \put(7,0){ (i)}
621: \put(22,0){ (ii)}
622: 
623: \end{picture}
624: \begin{picture}(30,15)(0,0)
625: %Dynkin diagram
626: \put(3,7){
627:  \begin{picture}(15,15)(0,0)
628:  \put(0,0){\line(1,0){3}} \put(0,0){\circle*{.3}}
629:  \put(3,0){\line(1,1){2.1213}} \put(3,0){\circle*{.3}}
630:  \put(3,0){\line(1,-1){2.1213}}
631:  \put(5.1213,2.1213){\circle*{.3}}
632:  \put(5.1213,-2.1213){\circle*{.3}}
633: 
634:  \put(2.5,0.5){$\delta _1$}
635:  \put(0,0.5){$\delta _0$}
636:  \put(5.6213,2.1213){$\delta _\lambda$}
637:  \put(5.6213,-2.1213){$\delta _\mu$}
638: 
639:  \put(5.1213,-2.1213){\line(0,1){4.2426}}
640:     \end{picture}
641:           }
642: %Dynkin diagram
643: \put(20,5){
644:  \begin{picture}(15,15)(0,0)
645:  \put(0,0){\line(1,0){4}} \put(0,0){\circle*{.3}}
646:  \put(4,0){\line(0,1){4}} \put(4,0){\circle*{.3}}
647:  \put(0,4){\line(1,0){4}}
648:  \put(0,0){\line(0,1){4}}
649:  \put(4,4){\circle*{.3}}
650:  \put(0,4){\circle*{.3}}
651: 
652:  \put(-.5,- 0.8){$\delta _0$}
653:  \put(4.3,-.8){$\delta _\lambda $}
654:  \put(-.5,4.5){$\delta _\mu$}
655:  \put(4.3,4.5){$\delta _1$}
656: 
657:  \put(0,0){\line(1,1){4}}
658:  \end{picture}
659:               }
660: \put(7,0){ (iii)}
661: \put(22,0){ (iv)}
662: 
663: \end{picture}
664: \caption{(i) The paths $l_0$, $l_1$, $l_\lambda$, $l_\mu$ in the case when
665: $\lambda \not \in \R$ and $h_\lambda \neq h_\mu $;
666: (ii) Dynkin diagram in the case when $\lambda \not \in \R$ lies inside the
667: left branch of the hyperbola $\Gamma$;
668: (iii) Dynkin diagram in the case when $\lambda \not \in \R$ lies inside the
669: right branch of the hyperbola $\Gamma$;
670: (iv) Dynkin diagram in the case when $\lambda \not \in \R$ lies between the
671: two branches of the hyperbola $\Gamma$.}
672: \label{dynkin1}
673: \end{figure}
674: 
675: We finish this section by introducing a property of the continuous
676: families of ovals which is dependent on the structure of the related
677: Dynkin diagram. Suppose that the real polynomial $H= y^2+ P(x)$ has
678: distinct critical values. Let $\delta (h) \subset \{H=h\}$ be a
679: continuous family of ovals defined on a maximal open interval
680: $\Sigma =(h_c,h_s)$, where for $h=h_c$ the oval degenerates to a point
681: $\delta (h_c)$ which is a center and for $h=h_s$ the oval becomes a
682: homoclinic loop of the Hamiltonian system $dH=0$. The family
683: $\{ \delta (h)\}$ represents a continuous family of cycles vanishing
684: at the center $\delta (h_c)$.
685: To formulate our main result we shall need the following
686: \begin{definition}
687: \label{excd}
688: {\em It is said that $\{ \delta (h)\}$ is {\em an exceptional family of ovals},
689: provided that for every polynomial one-form $\omega $ the Abelian integral
690: $$
691: I(h)= \int_{\delta (h)} \omega , \quad h\in \Sigma
692: $$
693: allows an analytic continuation in the sector $S_\varepsilon (h_s)= \{h\in \C:
694: Arg (h-h_s) \in (-\varepsilon, 2\pi + \varepsilon )\}$ for some
695: strictly  positive
696: $\varepsilon$. The corresponding family of vanishing cycles represented by
697: $\delta (h)$ is called {\em an exceptional family of vanishing cycles.}}
698: \end{definition}
699: The above definition has in fact a geometric nature.
700: Indeed, $I(h)$ has an analytic continuation in
701: the sector  $S_\varepsilon (h_s)$ if and only if $\delta (h)$ has an
702: appropriate intersection number with each of the other families of
703: vanishing cycles. For instance, if $\lambda$, $\mu$ are real and
704: $h_1 < h_\mu < h_\lambda < h_0$, then the family of cycles $\delta_1(h)$ is
705: exceptional, because it has zero intersection number with $\delta _0(h)$.
706: On the contrary, the family $\delta _\mu (h)$ is not exceptional, because it
707: has a non-zero intersection number with  $\delta _0(h)$.
708: The remaining possible distributions of the critical
709: values, as well as the case $\lambda, \mu \not \in \R$, are studied in the
710: same way. We summarize all this in the following
711: \begin{proposition}
712: \label{exceptional}
713: Suppose that the Hamiltonian $H(x,y)$ is taken in a normal form
714: $(\ref{normal})$ and has four distinct critical values. Then:
715: 
716: \vspace{1ex}
717: \noindent
718: $(i)$
719: In the real case $(0 <\mu< \lambda <1)$, the continuous family of ovals
720: ${\cal O}_1$ surrounding the  center
721: $(1,0)$ is exceptional, and the continuous family of ovals ${\cal O}_\mu$
722: surrounding the center $(\mu,0)$ is not exceptional.
723: 
724: \vspace{1ex}
725: \noindent
726: $(ii)$
727: In the complex case $(\lambda = \bar{\mu }\not \in \R)$,
728: the continuous family of  ovals ${\cal O}_1$
729: is exceptional if and only if $\lambda$ lies inside the left
730: branch of the hyperbola $\Gamma$ $($see
731: $(\ref{hyperbola})$ and {\em Fig. \ref{imconj})}.
732: \end{proposition}
733: 
734: \section{The period integral in a complex domain}
735: Assume that the level curve $\{H=h\}$ contains an oval $\delta(h)$ and
736: consider the integral (oriented along with the vector field (\ref{system}))
737: $$I_0(h)=\oint_{\delta(h)}\frac{dx}{y}$$
738: which is the derivative of the area inside $\delta(h)$ and hence
739: determines the period of the periodic orbit lying on $\delta(h)$.
740: Denote by $\alpha(h)$ and by $\beta(h)$ the minimal and the maximal
741: values of $x$ for $(x,y)\in\delta(h)$. These are among the real solutions
742: of the equation $P(x)=h$. Then one can express $I_0(h)$ in the
743: form
744: $$I_0(h)=\sqrt{2}\int_{\alpha(h)}^{\beta(h)}\frac{dx}{\sqrt{h-P(x)}}.$$
745: Consider first the case $0<\mu<\lambda<1$ and assume that $\delta(h)$ belongs
746: to either of the continuous families of
747: ovals ${\cal O}_\mu$, ${\cal O}_1$ surrounding centers $(\mu,0)$ and
748: $(1,0)$, respectively. The corresponding families of cycles vanish
749: respectively at $h=h_\mu$ and $h=h_1$. Denote this value in both cases
750: by $h_c$ and assume that the real ovals are defined in  $\Sigma=(h_c, h_s)$.
751: Finally, denote by $x_c$ the $x$-coordinate of the corresponding center.
752: We keep the same notation in the complex case and in the
753: cases when either $\mu=0$ or $\mu=\lambda$  or $\lambda=1$.
754: Then $x_c$ is the $x$-coordinate of the unique nondegenerate center
755: (as long as $(\lambda,\mu)\not\in\{(1,0), (1,1)\}$).
756: 
757: The integral $I_0(h)$ has an analytic continuation for $h<h_s$. We wish to
758: prove that $I_0(h)\neq 0$ in  $(-\infty, h_c)$. The precise statement
759: is as follows.
760: 
761: \begin{proposition}
762: \label{nozero}
763: Assume that the parameters $\lambda,\mu$ in $(\ref{normal})$ are either
764: real or lie inside the hyperbola $\Gamma$. Then the integral $I_0(h)$
765: related to any period annulus around a nondegenerate center of
766: $(\ref{system})$ has an analytic continuation in $(-\infty,h_s)$
767: and takes positive values there.
768: \end{proposition}
769: {\bf Proof.}
770: The claim that $I_0(h)$ has an analytic continuation in the interval
771: $(-\infty,h_s)$ is obvious in the complex case ($\lambda=
772: \bar{\mu}\not\in\R$). In the real case this follows from the Dynkin
773: diagram shown on Fig. \ref{realdynk}. For instance, if
774: $h_1 < h_\mu < h_\lambda < h_0$ as on Fig. \ref{realdynk}, and if
775: $h_s=h_\lambda$, $h_c= h_1$, then the claim is obvious. If $h_s=h_\lambda$,
776: $h_c= h_\mu$, then the claim is true again because the intersection number
777: of $\delta_\mu$ and $\delta_1$ is equal to zero. The remaining four
778: configurations of critical values and period annuli are examined in the
779: same way (with the help of Fig. \ref{realdynk}). Note, however, that the
780: integral $I_0(h)$ corresponding to the period annulus ${\cal O}_e$
781: (in this case $\Sigma =(h_\lambda, 0)$) has no analytic continuation on
782: the real axis outside $\Sigma$. This is so because the singularities in
783: both $h=h_0$ and $h=h_\lambda$ are unavoidable. The same conclusion holds
784: for some of the degenerate cases as well.
785: 
786: 
787: Next we shall prove the positivity of the period integral $I_0(h)$
788: on $(-\infty,h_s)$. We first choose a proper formula to present $I_0(h)$
789: when $h<h_c$. Take $z=x+iy\in\C$, $h<h_s$ and consider the polynomial
790: $h-P(z)$. By Taylor's formula,
791: $$\begin{array}{rl}
792: h-P(z)=&\!\!\!h-P(x)+\frac12y^2P''(x)-\frac{1}{24}y^4P''''(x)
793: -iy[P'(x)-\frac16y^2P'''(x)+\frac15y^4]\\[2mm]
794: \equiv&\!\!\!h-Q(x,y)-iyR(x,y)\end{array}$$
795: Given $h<h_s$, consider the equation  $h-P(z)=0$ and denote by
796: $\zeta_k(h)$, $k=1,2$, the two branches of the algebraic function defined by
797: $h-P(\zeta(h))\equiv 0$ which satisfy $\zeta(h_c)=(x_c,0)$.
798: These functions are unique since $P'(z)\neq 0$ for $z\neq 0,\mu,\lambda,1$.
799: For $h_c<h<h_s$, one has $\zeta_1(h)=(\alpha(h),0)=\alpha(h)$ and
800: $\zeta_2(h)=(\beta(h),0)=\beta(h)$ where $\alpha, \beta$ are as above.
801: For $h<h_c$, one has $\zeta_1(h)=\bar\zeta_2(h)\in\C$.
802: For a similarity in notation, we put $\zeta_1(h)=\alpha(h)$ and
803: $\zeta_2(h)=\beta(h)=\bar\alpha(h)$ where $Im\,\alpha<0$.
804: Denote by $C_h$ the curve in the complex plane connecting the
805: points $\alpha(h)$, $\beta(h)$ along $R(x,y)=0$, if $h<h_c$,
806: and along the real line $y=0$, if $h_c<h<h_s$. Then the integral
807: $I_0(h)$ is expressed as
808: \begin{equation}
809: \label{cmplx}
810: I_0(h)=\sqrt{2}\int_{C_h}\frac{dz}{\sqrt{h-Q(x,y)}},\quad h<h_s.
811: \end{equation}
812: Let $h<h_c$. We will establish below that $y=Im\,z$ can be used as a local
813: coordinate on $C_h$. Thus, $dz=[x'(y)+i]dy$ where $x(-y)=x(y)=x$. Denote for
814: short $y_h=Im\,\beta(h)$ and write $\int_{C_h}=\int_{-y_h}^0+\int_0^{y_h}$.
815: Replacing $y$ with $-y$ in the first integral, we obtain
816: \begin{equation}
817: \label{y}
818: I_0(h)=2\sqrt{2}\int_0^{y_h}\frac{dy}{\sqrt{Q(x(y),y)-h}},\quad h<h_c.
819: \end{equation}
820: 
821: Below we study the curve ${\cal R}:\;R(x,y)=0$ and establish that on this
822: curve the real-valued function $Q(x,y)-h$ is positive for all $h<h_c$.
823: 
824: Recall that $\cal R$ has an equation
825: $\frac15y^4-\frac16y^2P'''(x)+P'(x)=0$. We will study how this algebraic
826: curve changes when varying the parameters $\lambda,\mu$. It turns out that
827: a local analysis can be used to determine the global behavior of $\cal R$.
828: To begin with, we observe that the Poincar\'e index of any sufficiently big
829: circle $x^2+y^2=r^2$ subject to the vector field $dR(x,y)=0$ is $-3$.
830: The critical points of $R$ are determined from $R_x=R_y=0$, namely
831: 
832: \vspace{1ex}
833: (i) $y=0$, $P''(x)=0$;
834: 
835: \vspace{1ex}
836: (ii) $y^2=\frac{5}{12}P'''(x)$, $P''(x)-\frac{5}{72}P'''(x)P''''(x)=0$.
837: 
838: \vspace{1ex}
839: \noindent
840: It is easy to verify that the Hessian $R_{xx}R_{yy}-R_{xy}^2$ is negative
841: at any (real) nondegenerate critical point $(\xi,\eta)$. Indeed, it equals
842: $-\frac13(P'''(\xi))^2$ and
843: $-\frac{5}{108}P'''(\xi)(P''''(\xi))^2-\frac49(P'''(\xi))^2$, respectively
844: in cases (i) and (ii). Hence, $R$ has in general three nondegenerate saddles
845: and when two of the saddles collide, the resulting degenerate
846: critical point has only hyperbolic sectors around. In particular,
847: (ii) has no solutions for $0<\mu<\lambda<1$, since all the
848: three critical points are then given by (i). Also, in this case the curve
849: $\cal R$ is free of critical points because the value of $R$ at the saddle
850: $(\xi,0)$ is $P'(\xi)\neq 0$. In general, one can prove that $\cal R$
851: has no critical points outside the real line. Indeed, taking a critical point
852: $z=\xi+i\eta$ with $\eta\neq 0$, one has
853: $P'(z)=R(\xi,\eta)+\eta(R_y(\xi,\eta)+iR_x(\xi,\eta))=R(\xi,\eta)\neq 0$
854: unless $z=\lambda$ or $z=\mu$ (which may happen in the complex case).
855: However, by (ii), $(Re\,\lambda,\pm Im\,\lambda)$ is a critical point
856: if and only if $\lambda$ lies on the hyperbola $\Gamma$,
857: which is not our case. We have proved that the only critical points
858: $\cal R$  may have are $(0,0)$, $(\lambda,0)$ and $(1,0)$
859: provided that $\mu=0$, $\lambda=\mu$, $\lambda=1$, respectively.
860: 
861: As a consequence, the branches of the level curve $\cal R$ do not intersect
862: outside the real line and all they do escape to infinity (in the directions
863: of $x$ and $y$ altogether, as shown by the asymptotics at infinity). This
864: is because any closed compact curve necessarily surrounds a critical point
865: having an index $+1$ and $R(x,y)=const$ has no such points.
866: 
867: Assume first that $0<\mu<\lambda<1$. Then the four branches of $\cal R$ have
868: no common points at all. This implies that the branches through
869: $(0,0)$ and $(\mu,0)$ go to $-\infty$ in the $x$-direction while the
870: branches through $(\lambda,0)$ and $(1,0)$ go to $+\infty$, see Fig.
871: \ref{curve}. When running any of the branches, the $y$ variable changes
872: in a monotone way. (Otherwise, there would exist a horizontal line $y=const$
873: having at least 6 intersections with $\cal R$ which is impossible.)
874: Finally, denote by $\xi$ any of $0,1,\lambda,\mu$. The local equation
875: of $\cal R$ near the point $(\xi,0)$ is
876: $y^2=[6P''(\xi)/P'''(\xi)](x-\xi)+O((x-\xi)^2)$ where the coefficient
877: is negative for $\xi=0$, positive for $\xi=1$ and changes sign if
878: $\xi=\lambda,\mu$. This latter is because $P'''(\lambda)$ and
879: $P'''(\mu)$ change sign in $T$.
880: On the $(\lambda,\mu)$ curve where $P'''$ vanishes, the local equation
881: of $\cal R$
882: becomes $y^4\sim-5P''(\xi)(x-\xi)$ for $\xi=\lambda,\mu$.
883: Summing up, we obtain the form of the curve $\cal R$ as given in
884: Fig. \ref{curve}.
885: The arrows indicate the direction of the Hamiltonian vector field $dR=0$.
886: 
887: \begin{figure}[htbp]
888: \begin{center}
889:   \psfig{file=save2.eps}
890: \end{center}
891: \caption{The curve $R(x,y)=0$ for the  case $0<\mu<\lambda<1$.}
892: \label{curve}
893: \end{figure}
894: 
895: It is easy to see that $q(y)=Q(x(y),y)-h$ does not change sign when
896: $y\in(0,y_h)$. As $q(y_h)=0$, it suffices to show that $q'(y)\neq 0$.
897: We first note that $\dot{y}=-R_x<0$ on the branch through $(h_c,0)$
898: we consider. Then, by Cauchy-Riemann equations,
899: $$Q_x=(yR)_y,\quad Q_y=-(yR)_x,$$
900: and we obtain on $R=0$ that
901: $$q'=Q_y+x'Q_x=-\frac{y}{R_x}(R_x^2+R_y^2)<0.$$
902: This implies that $Q(x(y),y)-h>0$ for $y\in(0,y_h)$
903: and hence by (\ref{y}), $I_0(h)>0$ for $h<h_s$.
904: 
905: The same analysis is applicable when $\mu=0$ or $\mu=\lambda$ or $\lambda=1$,
906: the unique difference is that on Fig. \ref{curve} the corresponding
907: branches touch each other at $y=0$.
908: 
909: Let us turn now to the complex case. Apart of the real case, the branches
910: through $(0,0)$ and $(1,0)$ (let us denote them by $R_0$ and $R_1$) may
911: bifurcate significantly. This is because they are not necessarily monotone
912: with respect to $y$ and may escape to either $-\infty$ or $+\infty$, in the
913: $x$-direction. The complex plane is divided by the hyperbola $\Gamma$
914: into three parts. In the outer domain
915: $(Im\,\lambda)^2>5Re\,\lambda(Re\,\lambda-1)$, $R_0$ and $R_1$ behave as
916: in Fig. \ref{curve},
917: but are not necessarily monotone in $y$. Inside the left branch of the
918: hyperbola, $R_0$ and $R_1$ look like the branches through $\lambda$ and $1$
919: on Fig. \ref{curve}, respectively. And inside the right branch, $R_0$ and
920: $R_1$ behave as the branches through $0$ and $\mu$ on Fig. \ref{curve},
921: respectively. All this is easily verified by a simple deformation argument,
922: with a starting point on the appropriate side of triangle $T$.
923: 
924: To avoid complications, we regard only the parameters $\lambda$,
925: $\mu$ in the complex plane which belong to the interior of $\Gamma$
926: (which is the case to be considered in our main theorem). Taking
927: $\lambda=a+ib$, then $b^2<5a(a-1)$.
928: We are going to prove that $R_1$ is monotone with respect to $y$
929: in this domain. If so, one obtains just as in the real case that
930: $q(y)=Q(x(y),y)-h>0$ and the proof is complete.
931: 
932: Consider first the interior of the left branch of $\Gamma$. Then $R_1$ lies
933: in the half-plane $x\geq 1$. To establish monotonicity, we rewrite $R(x,y)$
934: as a polynomial in $t=x-1$. One obtains
935: \begin{equation}
936: \label{descartes}
937: R=t^4+(3-2a)t^3+(3-4a+a^2+b^2-2y^2)t^2+(1-2a+a^2+b^2+(2a-3)y^2)t+R(1,y).
938: \end{equation}
939: Assume that $a\leq\frac12$. As the first two coefficients of the polynomial
940: are positive, then according to Descartes rule of signs, (\ref{descartes})
941: can have three positive roots only if the coefficient at $t^2$ is negative
942: and the coefficient at $t$ is positive. However, these two conditions
943: contradict each other. Therefore each line $y=const$ can intersect $R_1$
944: in at most two points (counting multiplicity) and hence in a single point
945: because the number of intersections should be an odd number.
946: 
947: Now take $\lambda$ in the interior of the right branch of $\Gamma$ and denote
948: by $(\xi,\eta)$ the right-most point of $R_1$ in the half-plane $y\geq 0$.
949: Thus, $\xi\geq 1$ and $\eta\geq 0$. As above, one obtains with
950: $t=\xi-x \geq 0$ that
951: \begin{equation}
952: \label{descart1}
953: {\textstyle
954: R=t^4-\frac16P''''(\xi)t^3+(\frac12P'''(\xi)-2y^2)t^2-(P''(\xi)
955: -\frac16P''''(\xi)y^2)t+\frac15(y^2-\frac{5}{12}P'''(\xi))^2.}
956: \end{equation}
957: As $\eta=0$ if $P'''(\xi)\leq 0$ and $\eta^2=\frac{5}{12}P'''(\xi)$
958: otherwise, the coefficient at $t^2$ is negative for $y>\eta$.
959: Then by Descartes rule,
960: (\ref{descart1}) has only two positive roots and therefore the part of
961: $R_1$ above the line $y=\eta$ is monotone. It remains to consider the
962: part below it (if $\eta>0)$. In this domain (\ref{descart1}) can
963: have two complex roots and Descartes rule does not yield the required result.
964: We can proceed as follows. Assume that the part of $R_1$ in the upper
965: half-plane lying below $y=\eta$
966: is not monotone. Denote by $(\xi_1,\eta_1)$ the first local maximum on $R_1$
967: (starting from $(1,0)$), thus $1<\xi_1<\xi$ and $0<\eta_1<\eta$. Consider
968: also the branch $R_+$ of $\cal R$ which is placed in the upper half-plane
969: to the right of the line $x=\xi$ (and above the line $y=\eta_1$).
970: An easy calculation yields the equations
971: $y=\pm\sqrt{5\pm2\sqrt5}(x-\frac{1+2a}{4})$
972: for the tangential lines of $\cal R$ at infinity. As their common point
973: is different from $(\xi_1,\eta_1)$, there would always exist a tangential
974: line $l$ to $R_+$ at a finite point, which goes through $(\xi_1,\eta_1)$.
975: However, then $l$ would intersect $R_+$ in at least two points and $R_1$
976: in at least three points, which is impossible because $\cal R$ is a quartic
977: curve. The monotonicity of $R_1$ is established. $\Box$
978: 
979: \vspace{2ex}
980: \noindent
981: \begin{proposition}
982: \label{nozero1}
983: Assume that $\delta(h)$ is an exceptional family of ovals according to
984: Definition $\ref{excd}$. Then the period integral
985: $I_0(h)= \int_{\delta(h)} dx/y$
986: has an analytic continuation in the complex domain ${\cal D}= \C \setminus
987: [h_s,\infty)$ and does not vanish there. \end{proposition}
988: {\bf Proof.}
989: The proof that $I_0(h)$ has an analytic continuation in $\cal D$
990: follows immediately from the definition of an exceptional family of ovals.
991: To count the zeros of the analytic function $I_0(h)$ in $\cal D$,
992: we apply the argument principle to the complex domain
993: ${\cal D}_r$  obtained from $\cal D$ by removing the discs
994: $\{h\in \C: |h-h_{s}|< r \}$, $\{h\in \C: |h| > 1/r \}$, where $ r>0$
995: is a small enough constant.  For this purpose we
996: evaluate the increase of the argument of $I_{0}(h)$ when $h$ runs
997: the boundary of ${\cal D}_r$ in a positive direction.
998: 
999: We consider first the real case ($0< \mu < \lambda < 1$).
1000: The Picard-Lefschetz formula implies that in a suitable complex
1001: neighborhood of $h_s$ it holds
1002: $$I_{0}(h) = \varphi(h) \log(h-h_{s}) + \psi(h), \quad
1003: \varphi(h)= \frac{1}{2\pi} \int_{\delta_{s}(h)} \frac{dx}{y}
1004: $$
1005: where $\varphi,\psi$ are locally holomorphic and, as it is easily
1006: checked, $\varphi(h_s)\neq 0$.
1007: 
1008: 
1009: The behavior of $I_{0}(h)$ for $|h|$ sufficiently big follows from
1010: the Picard-Lefschetz formula too, but the argument is more delicate.
1011: Let $l$ be a simple closed path which makes one turn in a positive
1012: direction around all the critical values of $H$. It induces a
1013: monodromy map
1014: $$
1015: l_* : H_1(\Gamma_{\tilde{h}},\Z) \rightarrow H_1(\Gamma_{\tilde{h}},\Z),
1016: \quad \Gamma_{\tilde{h}}= \{(x,y)\in \C:\, H(x,y)=\tilde{h} \} \; .
1017: $$
1018: It is shown for instance in \cite{gav97} that $l_{*}$ coincides
1019: with the operator of classical monodromy of the singularity
1020: $\frac{1}{2}y^2 + \frac{1}{5}x^5 $ (which is the highest
1021: weighted homogeneous part of $H$). The characteristic polynomial $P(z)$
1022: of $l_*$ is now easily computed to be equal to
1023: $P(z)= (z^5+1)/(z+1)$ (see \cite{bri}). This shows that in a neighborhood
1024: of $h=\infty$ on the projective sphere the integral
1025: $I_0(h)$ is a meromorphic function of $h^{1/10}$. The substitution
1026: $$
1027: x \rightarrow x h^{1/5},\quad y \rightarrow y h^{1/2},\quad h \rightarrow
1028: \infty
1029: $$
1030: defines an isotopy of the regular fibers of $H$ and the regular fibers
1031: of its weighted homogeneous part
1032: $\frac{1}{2}y^2 + \frac{1}{5}x^5 $ \cite{gav97}.
1033: From this fact we deduce that
1034: $$I_{0}(h)=c_{\infty}h^{1/5-1/2} (1+O(h^{-1/10})).$$
1035: For a further use we note that
1036: $c_{\infty} \neq 0$. Indeed, the one-form $dx/y$ is not cohomologous to zero
1037: on the regular fibers of the polynomial $\frac{1}{2}y^2 + \frac{1}{5}x^5 $
1038: (this follows from \cite[Theorem 1.1]{gav98a}).
1039: On the other hand, the form of the characteristic polynomial of $l_{*}$ shows
1040: that it is irreducible over the field of rational numbers. Therefore the span
1041: of $l_*^k \delta$, $k=0, 1, 2, 3$ generates the whole
1042: homology group of the fiber. This implies that if $c_\infty=0$, then the
1043: restriction of  $dx/y$ on the regular fibers of
1044: $\frac{1}{2}y^2 + \frac{1}{5}x^5$  defines the zero
1045: cohomology class which is a contradiction.
1046: 
1047: 
1048: The above consideration shows that the decrease of the argument of $I_{0}(h)$
1049: along $\{h\in \C: |h| = 1/r \}$ is close to $3 \pi/5$ and the increase of the
1050: argument of $I_{0}(h)$  along $\{h\in \C: |h-h_{s}| = r \}$ is close to zero.
1051: 
1052: 
1053: We claim further that the imaginary part of $I_{0}(h)$ along
1054: $(r+h_s,1/r)$ does not vanish. Namely, denote the two determinations of
1055: $I_{0}(h)$ on $(r+h_s,1/r)$ by $I_{0}^{\pm}(h)$. Similarly, denote by
1056: $\delta^{\pm}(h)$ the two determinations of the continuous family
1057: $\delta(h)$ on $(r+h_s,1/r)$.
1058: Since $I_{0}$ is a real analytic function  along $(-\infty, h_{s})$, we have
1059: $$I_{0}^{+}(h)= \overline{I_{0}^{-}(h)},\quad
1060: \delta^{+}(h) =\overline{\delta^{-}(h)}, \quad h>h_{s}
1061: $$
1062: and hence $2\,\sqrt{-1} \, Im\, I_{0}^{+}(h) =
1063: I_{0}^{+}(h) - I_{0}^{-}(h)$. On the other hand $\delta(h)$ is an exceptional
1064: family of ovals. Therefore the Picard-Lefschetz formula implies that for
1065: $h> h_s$,
1066: $$
1067: \delta^{+}(h) - \delta^{-}(h) = \gamma(h),\qquad
1068: I_{0}^{+}(h) - I_{0}^{-}(h) = \int_{\gamma(h)} \frac{dx}{y}
1069: $$
1070: where $\gamma(h)$ is a continuous family of cycles vanishing at
1071: $h=h_{s}$, with intersection number $\langle\gamma,\delta^-\rangle=1$.
1072: 
1073: Finally, we note that after substituting $y$ by $\sqrt{-1} y$,
1074: the family of cycles $\gamma(h)$ corresponds to a family
1075: of ovals surrounding a center. Moreover, in the real case the full
1076: bifurcation diagram (consisting of the boundary of $T$ and the parts of
1077: the curves $h_\lambda=h_0$, $h_\mu=h_1$, $h_0=h_1$ inside) images onto
1078: itself. For later use we also mention that in the complex case, the
1079: interior of the hyperbola $\Gamma$ images onto itself. All this holds
1080: because the corresponding Hamiltonians when taken in a normal form
1081: (\ref{normal}) are related by the formula
1082: $$H(x, \sqrt{-1}y,\lambda,\mu)=h_1-H(1-x,y,1-\mu,1-\lambda).$$
1083: Hence, it follows from Proposition \ref{nozero}  that
1084: $\int_{\gamma(h)} dx/y$ does not vanish along $(h_{s},\infty)$. Therefore
1085: if  $h\in (r+h_s,1/r)$, then $Im\, I_{0}^{\pm}(h) \neq 0$ and hence
1086: the increase of the argument of $Im\, I_{0}^{\pm}(h)$
1087: along $(r+h_{s},1/r)$ is at most $\pi$.
1088: 
1089: Summing up the above information we conclude that the increase of the
1090: argument of $I_{0}(h)$ along the boundary of ${\cal D}_r$ is strictly
1091: less than $2\pi$. By the argument principle, $I_{0}(h)$ does not vanish
1092: in ${\cal D}_r$, and hence in $\cal D$.
1093: 
1094: The complex case when $\lambda = \bar{\mu }\not \in \R$ is studied in the
1095: same way. $\Box$
1096: 
1097: \section{The Chebyshev property}
1098: Suppose that the real polynomial $H= y^2+ P(x)$ has distinct critical values.
1099: Let $\delta (h) \subset \{H=h\}$ be a continuous family of ovals defined on
1100: a maximal open interval
1101: $\Sigma =(h_c,h_s)$, where for $h=h_c$ the oval degenerates to a point
1102: $\delta (h_c)$ which is a
1103: center and for $h=h_s$ the oval becomes a homoclinic loop of the
1104: Hamiltonian system $dH=0$. Denote ${\cal D}=\C\setminus[h_s,\infty)$.
1105: The following theorem is our main result.
1106: 
1107: 
1108: \begin{theorem}
1109: \label{main}
1110: Let $\delta (h)$, $h\in \Sigma $ be an exceptional family of ovals
1111: of $(\ref{normal})$. Then the real vector space
1112: ${\cal A}=\{\alpha_0 I_0 + \alpha_1 I_1: \alpha_0, \alpha_1 \in \R \}$
1113: of Abelian integrals
1114: $$I_0(h) =  \int_{\delta (h)} \frac{ dx}{y}, \qquad I_1(h) =
1115: \int_{\delta (h)} \frac{x dx}{y}, \qquad h\in \Sigma $$
1116: is Chebyshev in $\cal D$ and hence in $\Sigma$.
1117: \end{theorem}
1118: {\bf Proof.}
1119: Let $\delta(h)$ where $h \in \Sigma = (h_{c},h_{s})$ be an exceptional
1120: family of ovals. According to Proposition \ref{nozero1}, it suffices to
1121: show that the analytic function $F(h)= \alpha_0 + I_{1}(h)/I_{0}(h)$ has
1122: at most one zero in the domain $\cal D$. To prove this claim we apply the
1123: argument principle to $F(h)$ in the complex domain ${\cal D}_r$ obtained
1124: from $\cal D$ by removing the discs
1125: $\{h\in \C: |h-h_{s}|< r \}$, $\{h\in \C: |h| > 1/r \}$,
1126: where $ r>0$ is a small enough constant.
1127: 
1128:  As in the proof of Proposition \ref{nozero1}, we may check that:
1129: \begin{itemize}
1130:      \item  if $h\sim \infty$, then $I_{1}(h)$ is meromorphic
1131:       with respect to $h^{-1/10}$ and
1132:       $$I_{1}(h) = c_{\infty} h^{2/5-1/2}(1+O(h^{-1/10}))
1133:       \;\;\mbox{\rm where}\;\; c_{\infty} \neq 0;$$
1134: 
1135:       \item if $h \sim h_{s}$, then
1136:       $$I_{1}(h) = \varphi_1(h) \log(h-h_{s}) + \psi_1(h),
1137:  \quad \varphi_1(h)= \frac{1}{2\pi} \int_{\delta_{s}(h)} \frac{x\,dx}{y},$$
1138:   where $\varphi_1$ and $\psi_1$ are locally holomorphic. Moreover,
1139:   it is easily verified that
1140:   $\lim\limits_{h \rightarrow h_{s}} I_{1}(h)/I_{0}(h)=
1141:   \varphi_1(h_s)/\varphi(h_s) = x_{s} $
1142:   where $(x_{s},0)$ is the saddle point corresponding to the
1143:   critical value $h_{s}$.
1144: \end{itemize}
1145: 
1146: \noindent
1147: Therefore the increase of the argument of $F(h)$ along
1148: $\{h\in \C: |h| = 1/r \}$
1149: is close to $2 \pi/5$ and the increase of the argument of $F(h)$
1150: along $\{h\in \C: |h| = r \}$ is close to zero or strictly negative.
1151: 
1152: We claim further that
1153: \begin{itemize}
1154:         \item  the imaginary part of $F(h)$ along $(r+h_{s},1/r)$
1155:         does not vanish.
1156: \end{itemize}
1157: 
1158: 
1159: \noindent
1160: If this claim were proved then summing up the above information we would get
1161: that the argument of $F(h)$ increases along the boundary of ${\cal D}_r$
1162: by at most $2\pi + 2 \pi/5$. Thus the
1163: function $F(h)$ can have at most one zero in  ${\cal D}_r$  and hence
1164: in $\C \backslash [h_{s}, \infty)$.
1165: 
1166: It remains to prove the above claim. Recall that, since $\delta (h)$ is an
1167: exceptional family of ovals, then the two determinations of $F(h)$
1168: over $(h_s,\infty)$, namely $F^\pm(h)$, are analytic
1169: along $(h_s,\infty)$. The function $F(h)$ is real analytic along
1170: $(-\infty,h_s)$ and hence
1171: $$
1172: 2\,\sqrt{-1}\,Im\, F(h) = F^+(h) - F^-(h) \;\; \forall h\in (h_s,\infty).
1173: $$
1174: Denote by $\delta ^\pm(h)$ the two determinations of
1175: $\delta (h)$, $h\in {\cal D}$ on $(h_s,\infty)$, and let
1176: $$\Gamma_{h}~=~\{(x,y)\in \C^2: H(x,y) =h\}.$$
1177: It was shown in the proof of Proposition~\ref{nozero1} that
1178: $$
1179: \int_{\delta^{\pm} (h)} \frac{ dx}{y} \neq 0\quad\forall h \in (h_s,\infty).
1180: $$
1181: Then we obtain for each $h\in (h_s,\infty)$
1182: $$
1183:    2\,\sqrt{-1}\,Im\, F(h)  =  \frac{\int_{\delta^+(h)} \frac{x dx}{y}}
1184:    {\int_{\delta^{+} (h)} \frac{dx}{y}}
1185:                   - \frac{\int_{\delta^{-} (h)} \frac{x dx}{y}}
1186:    {\int_{\delta^{-} (h)} \frac{ dx}{y}}
1187:  =  \frac{ det \left( \begin{array}{cc}
1188:                 \int_{\delta^{+} (h)} \frac{x dx}{y} & \int_{\delta^{+}
1189: (h)} \frac{ dx}{y}  \\[1mm]
1190:                 \int_{\delta^{-} (h)} \frac{x dx}{y} & \int_{\delta^{-}
1191: (h)} \frac{ dx}{y} \end{array} \right) }{
1192:         |\int_{\delta^{\pm} (h)} \frac{ dx}{y}|^{2}}.
1193: $$
1194: Therefore we have to prove that the analytic function
1195: $$
1196: \Delta (h) =  det \left( \begin{array}{cc}
1197:                 \int_{\delta^{+} (h)} \frac{x dx}{y} & \int_{\delta^{+}
1198: (h)} \frac{ dx}{y}  \\[1mm]
1199:                 \int_{\delta^{-} (h)} \frac{x dx}{y} & \int_{\delta^{-}
1200: (h)} \frac{ dx}{y} \end{array} \right),\quad  h \in (h_s,\infty)
1201: $$
1202: does not vanish.
1203: 
1204: Suppose first that $h=h_0 \in (h_s,\infty)$ is a critical value of $H$.
1205: The affine curve $\Gamma_{h_0}$ is singular of arithmetic genus two. Let
1206: $\bar{\Gamma} _{h_0}$ be the corresponding completed and normalized Riemann
1207: surface. As the geometric genus of $\bar{\Gamma} _{h_0}$ is one and the intersection
1208: number $\langle \delta^{+},\delta^- \rangle$ equals $+1$, then
1209: $\delta^{+}(h_0)$ and $\delta^-(h_0)$ form a canonical homology basis of
1210: $\bar{\Gamma} _{h_0}$. We have $P(x)=(x-x_0)^2(x-x_1)(x-x_2)(x-x_3)$
1211: where $x_i\neq x_j$ for $i\neq j$. Therefore, $(x-x_0)dx/y$ induces a
1212: holomorphic one-form on $\bar{\Gamma} _{h_0}$ and $dx/y$ induces a
1213: meromorphic one-form of the third kind on $\bar{\Gamma} _{h_0}$ with simple
1214: poles at $P^\pm$, where $P^\pm$ are the two distinct pre-images
1215: of the singular point $(x_0,0)\in \Gamma _{h_0}$ under the canonical
1216: projection $ \bar{\Gamma} _{h_0}\rightarrow \Gamma_{h_0}$. The reciprocity
1217: law for differentials of the first and third kind \cite{gh} applied to
1218: $(x-x_0)dx/y$, $dx/y$ on the elliptic curve $\bar{\Gamma} _{h_0}$ gives
1219: \begin{eqnarray*}
1220: \Delta (h) & = & 2\pi \sqrt{-1}\left[
1221: Res|_{P^+}\left(\frac{dx}{y} \int_{P^0}^{P} \frac{(x-x_0)dx}{y}\right)
1222: + Res|_{P^-}\left(\frac{dx}{y} \int_{P^0}^{P} \frac{(x-x_0)dx}{y}\right)
1223: \right]\\
1224: &=& \frac{\pm 2 \pi \sqrt{-1}}{\sqrt{(x_0-x_1)(x_0-x_2)(x_0-x_3)}}
1225: \int_{P^-}^{P^+}\frac{dx}{\sqrt{(x-x_1)(x-x_2)(x-x_3)}
1226: }
1227: \end{eqnarray*}
1228: where the path of integration is on the Riemann surface
1229: $ \bar{\Gamma} _{h_0}$ of the affine elliptic curve
1230: $$\{y^2= (x-x_1)(x-x_2)(x-x_3) \}$$
1231: and $P^\pm = (x_0,\pm \sqrt{(x_0-x_1)(x_0-x_2)(x_0-x_3)}) $. The function
1232: $$
1233: P \rightarrow \int^P \frac{dx}{\sqrt{(x-x_1)(x-x_2)(x-x_3)}}
1234: $$
1235: is an uniformizing variable on the elliptic curve $ \bar{\Gamma} _{h_0}$
1236: which shows that
1237: $$
1238: \int_{P^-}^{P^+}\frac{dx}{\sqrt{(x-x_1)(x-x_2)(x-x_3)}} = 0
1239: \;\;\Rightarrow\;\; P^+=P^- .
1240: $$
1241: As $P^+ \neq P^-$, then $\Delta (h_0) \neq 0$.
1242: 
1243: 
1244: Suppose now that $h\in (h_s,\infty)$ is a regular value of $H$. The
1245: intersection number $\langle \delta^{+},\delta^- \rangle$ equals $+1$ and
1246: it is easy to check that we may choose a canonical homology basis
1247: $\delta_{1}(h),\ldots,\delta_{4}(h)$ of the lattice
1248: $H_{1}(\bar{\Gamma}_{h},\Z)$ such that $\delta_{1}(h) = \delta^{+}$,
1249: $\delta_{3}(h) = \delta^{-}$. Further, let $\omega_{1}, \omega_{2}$ be a
1250: normalized basis of holomorphic differentials on $\bar{\Gamma}_{h}$. The
1251: corresponding period lattice $\Pi$ reads
1252: $$
1253: \Pi = \left(
1254: \begin{array}{cccc}
1255:         \int_{\delta_{1} (h)} \omega_{1} & \int_{\delta_{2} (h)} \omega_{1} &
1256:         \int_{\delta_{3} (h)} \omega_{1} & \int_{\delta_{4} (h)} \omega_{1}\\
1257: \int_{\delta_{1} (h)} \omega_{2} & \int_{\delta_{2} (h)} \omega_{2} &
1258:         \int_{\delta_{3} (h)} \omega_{2} & \int_{\delta_{4} (h)} \omega_{2}
1259: \end{array}
1260: \right) =
1261: \left(
1262: \begin{array}{cccc}
1263:         1 & 0 &
1264:         a & b  \\
1265: 0 & 1 &
1266:         c & d
1267: \end{array}
1268: \right)
1269: $$
1270: where
1271: $$
1272: \left(
1273: \begin{array}{c}
1274:         \omega _1 \\
1275:         \omega _2
1276: \end{array}
1277: \right) =
1278: \left(
1279: \begin{array}{cc}
1280:    \int_{\delta_{1} (h)} \frac{xdx}{y} & \int_{\delta_{2} (h)}\frac{xdx}{y}\\
1281:    \int_{\delta_{1} (h)} \frac{dx}{y} & \int_{\delta_{2} (h)}\frac{dx}{y}
1282: \end{array}
1283: \right)^{-1}
1284: \left(
1285: \begin{array}{c}
1286:         \frac{xdx}{y} \\
1287:       \frac{ dx}{y}
1288: \end{array}
1289: \right) .
1290: $$
1291: We get that
1292: $$
1293: \Delta (h) = \det \left(
1294: \begin{array}{cc}
1295:     \int_{\delta_{1} (h)} \frac{xdx}{y} & \int_{\delta_{2} (h)}\frac{xdx}{y}\\
1296:     \int_{\delta_{1} (h)} \frac{dx}{y} & \int_{\delta_{2} (h)}\frac{dx}{y}
1297: \end{array}
1298: \right) \det
1299: \left(
1300: \begin{array}{cc}
1301:         1 & a \\
1302:      0 & c
1303: \end{array}
1304: \right)
1305: $$
1306: and hence
1307: $$
1308: \Delta (h)=0 \;\;\Leftrightarrow \;\; c=0.
1309: $$
1310: Let $\Lambda $ be the lattice generated by the columns of $\Pi $.
1311: The Riemann bilinear relations \cite{gh} on $\Pi $ imply
1312: $$
1313: b=c.
1314: $$
1315: Therefore if $\Delta (h)=0$, then the Jacobian variety
1316: $J(\bar{\Gamma} _{h}) = \C^2/\Lambda $ is
1317: a direct product of the two elliptic curves
1318: \begin{equation}
1319: \label{curves}
1320: \C / \{\Z \oplus a \Z  \},\quad \C / \{\Z \oplus d \Z  \}.
1321: \end{equation}
1322: It is well known that this is impossible (e.g. the Remark on page 49 in
1323: \cite{arbarello}) which completes the proof of Theorem \ref{main}.
1324: 
1325: For convenience of the reader, we present a proof of the last claim.
1326: Consider the Riemann theta function
1327: $$
1328: \theta(z)=  \sum_{m\in \Z^{2}} e^{\pi \sqrt{-1} \langle m,Bm\rangle + 2 \pi
1329: \sqrt{-1}\langle m,z\rangle}
1330: $$
1331: and the corresponding theta divisor
1332: $\Theta = \{z \in  J(\bar{\Gamma}_{h}): \;\theta (z)=0 \}$ where
1333: $$
1334: B= \pmatrix{ a & b  \cr c & d},\quad z= \pmatrix{z_{1}\cr z_{2}},\quad
1335:  \langle m,z\rangle = m_{1}z_{1}+ m_{2}z_{2}.
1336: $$
1337: By the Riemann theorem, the divisor $\Theta$ is isomorphic to the curve
1338: $\bar{\Gamma} _{h}$ and in particular is irreducible. If $\Delta(h)=0$,
1339: then $b=c=0$ and $\theta(z)$ factorizes into a product of two elliptic
1340: theta functions
1341: $$
1342: \theta(z)=
1343: \sum_{m_{1}\in \Z} e^{\pi \sqrt{-1} a m_{1}^{2} + 2 \pi
1344: \sqrt{-1} m_{1} z_{1}}
1345: \sum_{m_{2}\in \Z} e^{\pi \sqrt{-1} d m_{2}^{2} + 2 \pi
1346: \sqrt{-1} m_{2} z_{2}}  .
1347: $$
1348: Therefore the Riemann theta divisor $\Theta$ is a product of the two
1349: elliptic curves (\ref{curves}) which is a contradiction.  $\Box$
1350: 
1351: \section{Concluding remarks}
1352: We finish this paper with some general hypotheses concerning the zeros of
1353: the function $\alpha_0I_0+\alpha_1I_1$. In the previous section we
1354: demonstrated that this linear combination belongs to a Chebyshev space,
1355: provided that the continuous family of ovals $\{\delta(h)\}$ is exceptional.
1356: In the non-exceptional case a big part of our proof still holds,
1357: provided that $\lambda$ and $\mu$ are real. This fact combined
1358: with the exact local result obtained in Proposition \ref{cyclicity}
1359: suggests the following more general statement.
1360: 
1361: \vspace{2ex}
1362: \noindent
1363: {\bf Conjecture 1.}
1364: Assume that the polynomial $P(x)$ has distinct real critical values.
1365:  Then the corresponding
1366: vector space of Abelian integrals $\alpha_0 I_0+\alpha_1I_1$
1367: is Chebyshev with accuracy at most one in ${\cal D}$.
1368: 
1369: %It turns out that, at least locally in a neighborhood of a center, these two
1370: %bounds do not coincide. The bifurcation curves in the $\lambda,\mu$-space
1371: %which separate domains in which the centers have different cyclicity turn
1372: %out to be algebraic, and hence globally defined. This suggests immediately
1373: %a global conjecture which we formulate below.
1374: 
1375: \vspace{2ex}
1376: \noindent
1377: One may also ask whether the upper bound (two) for the number of
1378: the zeros near the center always holds in the whole interval $\Sigma$ and
1379: next, whether the bounds for the real zeros when $h\in\Sigma$ and for the
1380: complex zeros when $h$ belongs to a certain complex extension of $\Sigma$
1381: do coincide. It turns out that the answers of all these questions are
1382: negative. In particular, we establish below that when $P(x)$ has
1383: complex critical values, then there are three zeros in $\Sigma$
1384: for appropriate $\lambda\in\C$.
1385: 
1386: 
1387: One can try to predict the possible number of real zeros by studying
1388: the behavior of the ratio $F(h)=I_1(h)/I_0(h)$ near the ends of
1389: the interval $\Sigma$. For reader's convenience, we first recall some
1390: formulas.
1391: 
1392: Let us take a continuous function $f(x)$ and consider the integral
1393: $$I(h)=\int_{\delta(h)}\frac{f(x)dx}{y},\quad h\in\Sigma$$
1394: where $\Sigma=(h_c,h_s)$ for the families ${\cal O}_1$, ${\cal O}_\mu$
1395: and $\Sigma=(h_s,h_0)$ for the family ${\cal O}_e$.
1396: Using Picard-Lefschetz formula or direct calculations of integrals,
1397: one can obtain the following asymptotical expansions in a neighborhood
1398: of the critical levels $h_s$ corresponding to a saddle or a cuspidal point
1399: $(x_s,0)$:
1400: 
1401: \vspace{1ex}
1402: \noindent
1403: (i) For the saddle-loop and eight-loop cases,
1404: $$I(h)=-c_sf(x_s)\log|h-h_s|+\varphi(h)$$
1405: where $c_s$ is a specific positive constant independent on $f$ and
1406: $\varphi(h)$ is a continuous function in a neighborhood of $h_s$.
1407: We are not going to write up the exact expression of $c_s$ since do
1408: not need it here.
1409: 
1410: \vspace{1ex}
1411: \noindent
1412: (ii) For the heteroclinic loop cases,
1413: $$I(h)=-(c_0f(0)+c_s f(x_s))\log|h-h_s|+\varphi(h)$$
1414: where $c_0$, $c_s$ and $\varphi(h)$ are as above.
1415: 
1416: \vspace{1ex}
1417: \noindent
1418: (iii) For the cuspidal loop cases,
1419: $$I(h)=c_sf(x_s)|h-h_s|^{-1/6}+\varphi(h),$$
1420: where $c_s$ and $\varphi(h)$ are as above.
1421: 
1422: \vspace{1ex}
1423: \noindent
1424: (In fact, these formulas hold in a much more general context.)
1425: Applying the above expansions to $F(h)$, we get $F(h_s)= x_s/(1+c_0/c_s)$
1426: for the heteroclinic loop case and $F(h_s)=x_s$ for all the remaining cases.
1427: Moreover, what concerns the saddle-loops, it is evident that
1428: $F(h)-x_s\neq 0$ in $\Sigma$ and the sign of the expression depends on
1429: the position of the saddle with respect to the family of ovals.
1430: Consider now the case of ${\cal O}_e$ where the situation is more
1431: complicated. For $h=h_\lambda$, we obtain from (\ref{normal}) that
1432: $$\textstyle y^2=\frac25(x-\lambda)^2p(x),\quad p(x)=(x-x_1)(x-x_2)(x_3-x)$$
1433: where $x_1<0<x_2\leq\lambda\leq x_3$ and $x_2$, $x_3$ are the left-most and
1434: the right-most points of the eight-loop (or the cuspidal loop, as a
1435: limit case). By (i)--(iii), the function $\varphi(h)=I_1(h)-\lambda I_0(h)$
1436: is continuous in a neighborhood of $h_\lambda$, which yields immediately that
1437: $$\varphi(h_\lambda)=
1438: \sqrt{10}\left(\int_\lambda^{x_3}-\int_{x_2}^\lambda\right)\frac{dx}
1439: {\sqrt{p(x)}}.$$
1440: Clearly, $\varphi(h_\lambda)\to-\infty$ when $x_2\to x_1$,
1441: $\varphi(h_\lambda)\to\pm \sqrt{10}\int_{x_2}^{x_3}dx/\sqrt{p(x)}$
1442: when $\lambda\to x_2$ and $\lambda\to x_3$, respectively.
1443: As $x_2=x_1$ is equivalent to $h_\lambda=0$,
1444: $\lambda=x_2$ is equivalent to $\lambda=\mu$ and $\lambda=x_3$
1445: is equivalent to $\lambda=1$, this means that the curve
1446: $$\gamma_s=\{(\lambda,\mu)\in T:\;I_1(h_\lambda)-\lambda I_0(h_\lambda)=0\}$$
1447: connects the points $(0,0)$ and $(1,1)$ and lies in the part of $T$ above
1448: $\gamma$. Surprisingly, although expressed in terms of incomplete elliptic
1449: integrals, it turns out that $\gamma_s$ is a part of an algebraic curve of
1450: degree 6. This can be verified by transforming the integrals to a standard
1451: form. From the explicit algebraic equation thus obtained it is easily seen
1452: that $\gamma_s$ is a simple and connected curve (we omit the details).
1453: The curve $\gamma_s$ is an element of the bifurcation diagram because for
1454: $h$ close to $h_\lambda$, the function $F(h)-\lambda$ will change the sign
1455: when crossing $\gamma_s$ in the parameter space.
1456: 
1457: Among other, the above analysis implies that for $\lambda=\mu$, the function
1458: $\alpha_0I_0+\alpha_1I_1=I_0(\alpha_0+\alpha_1 F)$ can have three zeros
1459: in $(h_1,h_\lambda)\cup(h_\lambda,h_0)$. Hence, the same fact remains true
1460: for ${\cal O}_1$ and the respective $\Sigma=(h_1,h_0)$ in the complex case,
1461: at least for parameters $\lambda=a+ib$ with $a\in(0,1)$ and $|b|$ small.
1462: Based on the behavior of $F$ near the other endpoint which corresponds
1463: to a center (see below), we conjecture that the domain in the parameter space
1464: $\lambda\in\C$ where the equation $F(h)=const$ can have three real zeros,
1465: is surrounded by a curve on which they join into a triple zero.
1466: 
1467: Looking at the bifurcation diagram again, let us denote by $\gamma_c$
1468: the curve
1469: $$\textstyle\gamma_c=\{(\lambda,\mu)\in T:\;
1470: \lambda=\frac{3\mu^2-2\mu}{2\mu-1}\}
1471: \quad\mbox{\rm (in the real case),}$$
1472: $$\gamma_c=\{\lambda\in \C:\; |\lambda-2|=1\}
1473: \quad\mbox{\rm (in the complex case).}$$
1474: Further, let us denote by $\gamma^*$ the (hypothetical) curve
1475: $$\gamma^*=\{\lambda\in \C:\;
1476: \exists h^*\in\Sigma\;\; \mbox{\rm with}\;\; F'(h^*)=F''(h^*)=0\}.$$
1477: We believe that $\gamma^*$ is a simple closed curve in the complex plane
1478: going through $(0,0)$ and $(1,0)$. On $\gamma^*$, a bifurcation
1479: of a triple real zero into a simple one occurs.
1480: 
1481: Next, in the real case, denote by $\Omega_\mu$ the part of $T$ placed
1482: between the curves $\gamma_c$ and $\gamma$, by $\Omega_e$ the part of
1483: $T$ placed between the curve $\gamma_s$ and the line $\lambda=\mu$
1484: and finally, in the complex case, denote by $\Omega_1$
1485: the unit disk inside $\gamma_c$ and by $\Omega_1^*$
1486: the interior of $\gamma^*$ (see figures \ref{imconj} and  \ref{reconj}).
1487: %\begin{figure}[htbp]
1488: %\begin{center}
1489: %  \psfig{file=save3.eps}
1490: %\end{center}
1491: %\caption{ Bifurcation diagram}
1492: %\label{reconj}
1493: %\end{figure}
1494: 
1495:         \begin{figure}
1496: \setlength{\unitlength}{.04\textwidth}
1497: 
1498: \begin{picture}(25,15)(5,0)
1499: \put(2,8){
1500:  \put(5,0){\vector(1,0){20}}
1501: \put(12,-7){\vector(0,1){15}}
1502: \qbezier[200](9,-6,)(14.96,0)(9,6)
1503: \put(4,0){\qbezier[200](15,-6,)(9.04,0)(15,6)}
1504: \put(12,0){
1505: \qbezier[50](0,0)(0,6)(2,6)
1506: \qbezier[50](0,0)(0,-6)(2,-6)
1507: \qbezier[50](2,6)(4,6)(4,0)
1508: \qbezier[50](2,-6)(4,-6)(4,0)
1509: }
1510: \put(16,0){
1511: \qbezier[100](0,0)(0,2)(2,2)
1512: \qbezier[100](0,0)(0,-2)(2,-2)
1513: \qbezier[100](2,2)(4,2)(4,0)
1514: \qbezier[100](2,-2)(4,-2)(4,0)}
1515: \put(25.5,-.3){$Re \lambda$}
1516: \put(13.5,.3){$\Omega_1^*$}
1517: \put(17,.3){$\Omega_1$}
1518: \put(11,-.8){$0$}
1519: \put(16.5,-.8){$1$}
1520: \put(20.5,-.8){$3$}
1521: \put(15,-6){$\gamma^*$}
1522: \put(19.5,-2){$\gamma_c$}
1523: \put(19.7,6){$\Gamma$}
1524: \put(8,6){$\Gamma$}
1525: \put(11,8.5){$Im \lambda$} }
1526: \end{picture}
1527: \caption{Partition of the parameter space in the complex case 
1528: $\lambda = \bar{\mu} \in \C$.}
1529: \label{imconj}
1530: \end{figure}
1531: \begin{figure}
1532: \setlength{\unitlength}{.04\textwidth}
1533: 
1534: \begin{picture}(25,25)(0,0)
1535: \put(0,0){\vector(1,0){22}}
1536: \put(0,0){\vector(0,1){20}}
1537: \put(20,0){\line(0,1){20}}
1538: 
1539: \qbezier[100](0,0)(10,4.5)(20,6)
1540: \qbezier[100](0,0)(9,6)(20,10)
1541: \qbezier[100](0,0)(15,10)(20,20)
1542: \put(0,0){\line(1,1){20}}
1543: \put(0,-.6){$0$}
1544: \put(20,-.6){$1$}
1545: \put(22,.6){$\lambda$}
1546: \put(0.6,19.6){$\mu$}
1547: \put(10,3.2){$\gamma_c$}
1548: \put(16,7){$\Omega_\mu$}
1549: \put(14,8.1){$\gamma$}
1550: \put(17,14.6){$\gamma_s$}
1551: \put(10,9.1){$\Omega_e$}
1552: 
1553: \end{picture}
1554: \caption{Partition of the parameter space in the real case $\lambda,\mu \in \R$.
1555: }
1556: \label{reconj}
1557: \end{figure}
1558: \begin{proposition}
1559: \label{twocycles}
1560: The vector spaces ${\cal A}_\mu$, ${\cal A}_e$, ${\cal A}_1$ of Abelian
1561: integrals $\alpha_0I_0+\alpha_1I_1$ corresponding to the continuous
1562: families of ovals ${\cal O}_\mu$ for $(\lambda,\mu)\in\Omega_\mu$,
1563: ${\cal O}_e$ for $(\lambda,\mu)\in\Omega_e$, and
1564: ${\cal O}_1$ for $\lambda\in\Omega_1\cup\Omega_1^*$, are not Chebyshev
1565: in the respective intervals $\Sigma$.
1566: \end{proposition}
1567: {\bf Proof.} The assertion follows from local analysis.
1568: Consider first ${\cal O}_\mu$. Then $\Sigma=(h_\mu, 0)$ and
1569: $F(h_\mu)=\mu>0$, $F(0)=0$. By the proof of Proposition \ref{cyclicity},
1570: we obtain that $F'(h_\mu)>0$ is equivalent to $a_1<0$. This coefficient
1571: is easily calculated from the normal form (\ref{normal}) and the
1572: result is
1573: $$a_1=\frac{P'''(\mu)}{3P''(\mu)}=
1574: \frac{6\mu^2-4\lambda\mu-4\mu+2\lambda}{3\mu(\lambda-\mu)(1-\mu)}$$
1575: which proves the claim. The proof in the case of ${\cal O}_1$
1576: is similar. The assertion for ${\cal O}_e$ follows from the
1577: definition of $\gamma_s$. $\Box$
1578: 
1579: \vspace{1ex}
1580: Making use of the whole information about the local behavior of $F$ near
1581: the endpoints of $\Sigma$, one can formulate the following hypothesis.
1582: 
1583: \vspace{2ex}
1584: \noindent
1585: {\bf Conjecture 2.} (i).
1586: The vector spaces ${\cal A}_\mu$, ${\cal A}_e$ are Chebyshev with accuracy
1587: one in $\Sigma$ for the cases considered in Proposition \ref{twocycles},
1588: and they are Chebyshev in $\Sigma$ for all the remaining cases.
1589: 
1590: (ii). The vector space ${\cal A}_1$ is Chebyshev with accuracy one in
1591: $\Sigma$ for $\lambda\in\Omega_1$, it is Chebyshev with accuracy two in
1592: $\Sigma$ for $\lambda\in \Omega_1^*$ and it is Chebyshev in $\Sigma$ for
1593: the remaining cases.
1594: 
1595: (iii). Given $\alpha_0$ and $\alpha_1$, the total number of zeros of
1596: the function $\alpha_0I_0+\alpha_1I_1$ relative to all the period annuli
1597: is three and this bound is attained in $\Omega_e\cup\Omega_1^*$.
1598: 
1599: \vspace{2ex}
1600: \noindent
1601: It is plausible that Conjecture 2 is a consequence of Conjecture 1, except in
1602: the case when the polynomial $P(x)$ has complex critical points.
1603: 
1604: The next problem in difficulty should be to study the space of Abelian
1605: integrals of the first kind related to genus three hyperelliptic curves. We
1606: hope that this problem can be settled by combining the methods of the present
1607: paper with deformation arguments as in \cite{Gav4}. Properly understood,
1608: this might lead to the solution of the problem for arbitrary genus $g$.
1609: 
1610: \vspace{2ex}
1611: \noindent
1612: {\bf Acknowledgments.} This research is partially supported by the
1613: Ministry of Science and Education of Bulgaria under Grant no. 1003/2000.
1614: 
1615: \begin{thebibliography}{99}
1616: \bibitem{cam}
1617: %
1618: N. A'Campo,
1619: Le groupe de monodromie du d\'eploiement des singularit\'es isol\'ees de
1620: courbes planes. I. {\it Math. Ann.} {\bf 213} (1975), 1--32.
1621: %
1622: \bibitem{arbarello} E. Arbarello, M. Cornalba, P.A. Griffiths, J. Harris,
1623: {\it Geometry of Algebraic Curves}, vol. 1, Springer, New York, 1985.
1624: %
1625: \bibitem{Arnold} V.I. Arnold, S.M. Gusein-Zade, A.N. Varchenko,
1626: {\it Singularities of Differentiable Maps}, vols. 1 and 2,
1627: Monographs in Mathematics, Birkh\"auser, Boston, 1985 and 1988.
1628: %
1629: \bibitem{Arnold0} V.I. Arnold, Yu.S. Il'yashenko, Ordinary Differential
1630: Equations, in: {\it Dynamical Systems} I, Encyclopaedia of Math. Sci.,
1631: vol. 1, Springer, Berlin, 1988.
1632: %
1633: \bibitem{Arnold1} V.I. Arnold, {\it Geometrical Methods in the Theory of
1634: Ordinary Differential Equations}, Springer, New York, 1988.
1635: %
1636: \bibitem{Arnold2} V.I. Arnold, Sur quelques probl\`emes de la
1637: th\'eorie des syst\`emes dynamiques, {\it Topological Methods in Nonlinear
1638: Analysis}, {\bf  4} (1994), 209--225.
1639: %
1640: \bibitem{Arnold1+} V.I. Arnold, Some unsolved problems in the theory of
1641: differential equations and mathematical physics,
1642: {\it Russian Math. Surveys} {\bf 44} (1989), no. 4, 157--171.
1643: %
1644: \bibitem{ten} V.I. Arnold, Ten problems, in: {\it Theory of
1645: singularities and its applications}, {\it Adv. Soviet Math.} {\bf 1},
1646: pp. 1--8, Amer. Math. Soc., Providence, 1990.
1647: %
1648: \bibitem{bri} E. Brieskorn, Die Monodromie der isolierten Singularit\"{a}ten
1649: von Hyperfl\"{a}chen, {\it  Manuscripta Math.} {\bf 2} (1970), 103--161.
1650: %
1651: \bibitem{dim} A. Dimca, {\it Singularities and Topology of Hypersurfaces},
1652: Springer, Berlin, 1992.
1653: %
1654: \bibitem{gas-li} A. Gasull, W. Li, J. Llibre, Zh. Zhang, Chebyshev property
1655: of complete elliptic integrals and its application to Abelian integrals,
1656: {\it Pacific J. Math.}  {\bf 202} (2002), no. 2, 341--362.
1657: %
1658: \bibitem{gav97} L. Gavrilov, Isochronicity of plane polynomial
1659: Hamiltonian systems, {\it Nonlinearity} {\bf 10} (1997), no. 2, 433--448.
1660: %
1661: \bibitem{gav98a} L. Gavrilov, Petrov modules and zeros of Abelian
1662: integrals, {\it Bull. Sci. Math.} {\bf 122} (1998), no. 8, 571--584.
1663: %
1664: \bibitem{gav99} L. Gavrilov, Abelian integrals related to Morse polynomials
1665: and perturbations of plane Hamiltonian vector fields, {\it Ann. Inst. Fourier}
1666: {\bf 49} (1999), no. 2, 611--652.
1667: %
1668: \bibitem{gav98} L. Gavrilov, Nonoscillation of elliptic integrals related
1669: to cubic polynomials with symmetry of order three, {\it Bull. London Math.
1670: Soc.} {\bf 30} (1998), no. 2,  267--273.
1671: %
1672: \bibitem{Gav4} L. Gavrilov, The infinitesimal 16th Hilbert problem in the
1673: quadratic case, {\it Invent. Math.} {\bf 143} (2001), 449--497.
1674: %
1675: \bibitem{giv} A.B. Givental, Sturm's theorem for hyperelliptic integrals,
1676: {\it Leningrad Math. J.} {\bf 1} (1990), 1157--1163.
1677: %
1678: \bibitem{gh} P.A. Griffiths, J. Harris, {\it Principles of algebraic
1679: geometry}, Pure and Appl. Math., John Wiley and Sons, New York, 1978.
1680: %
1681: \bibitem{hil} D. Hilbert, Mathematische probleme,
1682: {\it Gesammelte Abhandlungen} III, Springer-Verlag, Berlin (1935),
1683: pp. 403--479.
1684: %
1685: \bibitem{nov-yak} D. Novikov and S. Yakovenko, Tangential Hilbert problem
1686: for perturbations of hyperelliptic Hamiltonian systems, {\it Electron. Res.
1687: Announc. Amer. Math. Soc.} {\bf 5} (1999), 55--65 (electronic).
1688: %
1689: % \bibitem{zade} S.M. Husein-Zade, Dynkin diagrams of singularities of
1690: % functions of two variables, {\it Functional Anal. Appl.} {\bf 8}  (1974),
1691: % 10--13.  % 295-300.
1692: %
1693: \bibitem{Petrov86} G.S. Petrov, Number of zeros of complete elliptic integrals,
1694: {\it Funct.Anal.Appl.} {\bf 18} (1984), 72--73; {\bf 20} (1986), 37--40;
1695: {\bf 21 } (1987), 87--88; {\bf 22} (1988), 37--40;
1696: {\bf 23} (1989), 88--89; {\bf 24} (1990), 45--50.
1697: %
1698: \bibitem{pont} L.S. Pontryagin, On dynamic systems close to Hamiltonian
1699: systems, {\it Zh. Eksp. Teor. Fiz.} {\bf 4} (1934), 234--238 (Russian).
1700: %
1701: \bibitem{Roussarie1} R. Roussarie, {\it Bifurcation of Planar Vector Fields
1702: and Hilbert's sixteenth Problem}, Progress in Mathematics, vol. 164,
1703: Birkh{\"a}user, Basel, 1998.
1704: %
1705: \end{thebibliography}
1706: 
1707: \end{document}
1708: %=======================================================================
1709: 
1710: 
1711: 
1712: 
1713: 
1714: 
1715: 
1716: 
1717: 
1718: 
1719: 
1720: 
1721: 
1722: 
1723: 
1724: 
1725: 
1726: 
1727: 
1728: 
1729: 
1730: 
1731: 
1732: 
1733: 
1734: 
1735: 
1736: 
1737: 
1738: 
1739: 
1740: 
1741: 
1742: 
1743: 
1744: 
1745: 
1746: 
1747: 
1748: 
1749: 
1750: 
1751: 
1752: