1: \documentclass{article}
2: \usepackage{amssymb}
3: \usepackage{amsmath}
4: \begin{document}
5: \input epsf.sty
6:
7:
8:
9:
10:
11: \title{On polynomials orthogonal to all powers of a given
12: polynomial on a segment}
13: \author{F. Pakovich}
14: \date{}
15:
16:
17: \maketitle
18:
19:
20: %\tableofcontents
21: %\pagebreak
22:
23:
24: \def\be{\begin{equation}}
25: \def\ee{\end{equation}}
26: \def\bs{$\square$ \vskip 0.2cm}
27: \def\d{{\rm d}}
28: \def\D{{\rm D}}
29: \def\I{{\rm I}}
30: \def\C{{\mathbb C}}
31: \def\N{{\mathbb N}}
32: \def\P{{\mathbb P}}
33: \def\Z{{\mathbb Z}}
34: \def\R{{\mathbb R}}
35: \def\ord{{\rm ord}}
36:
37: \def\e{\eqref}
38: \def\phi{{\varphi}}
39: \def\v{{\varepsilon}}
40: \def\deg{{\rm deg\,}}
41: \def\Det{{\rm Det}}
42: \def\dim{{\rm dim\,}}
43: \def\Ker{{\rm Ker\,}}
44: \def\Gal{{\rm Gal\,}}
45: \def\St{{\rm St\,}}
46: \def\exp{{\rm exp\,}}
47: \def\cos{{\rm cos\,}}
48: \def\circ{{\rm circ\,}}
49: \def\diag{{\rm diag\,}}
50: \def\GCD{{\rm GCD }}
51: \def\LCM{{\rm LCM }}
52: \def\mod{{\rm mod\ }}
53:
54: \def\bp{\begin{proposition}}
55: \def\ep{\end{proposition}}
56: \def\bt{\begin{theorem}}
57: \def\et{\end{theorem}}
58: \def\be{\begin{equation}}
59: \def\l{\label}
60: \def\ee{\end{equation}}
61: \def\bl{\begin{lemma}}
62: \def\el{\end{lemma}}
63: \def\bc{\begin{corollary}}
64: \def\ec{\end{corollary}}
65: \def\pr{\noindent{\it Proof. }}
66: \def\note{\noindent{\bf Note. }}
67: \def\bd{\begin{definition}}
68: \def\ed{\end{definition}}
69:
70: \newtheorem{theorem}{Theorem}[section]
71: \newtheorem{lemma}{Lemma}[section]
72: \newtheorem{definition}{Definition}[section]
73: \newtheorem{corollary}{Corollary}[section]
74: \newtheorem{proposition}{Proposition}[section]
75:
76:
77:
78: \section{Introduction}
79:
80: In this paper we investigate the following
81: ``polynomial moment problem":
82: {\it for a complex polynomial $P(z)$ and distinct complex numbers $a,b$
83: to describe polynomials $q(z)$
84: such that}
85: \be \l{1}
86: \int^b_a P^i(z)q(z)\d z=0
87: \ee
88: {\it for all integer non-negative $i$.}
89: This problem attracted
90: attention recently in the
91: series of papers \cite{bby}-\cite{by}, \cite{y1}, where (1) arose
92: as an infinitesimal version of the
93: center problem for the Abel differential equation with polynomial
94: coefficients in the complex domain.
95:
96: The example of $P(z)=z$ shows that $q(z)=0$ can be the only
97: polynomial solution to
98: \eqref{1} since the Stone-Weierstrass
99: theorem implies that the only continuous complex-valued function
100: which is orthogonal to all powers of $z$ is zero. On the other
101: hand, if $P(a)=P(b),$ then non-trivial polynomial solutions to (1) always
102: exist. Indeed, it is enough to set
103: $q(z)=R^{\prime}(P(z))P^{\prime}(z),$ where $R(z)$ is any
104: complex polynomial. Then for any $i\geq 0$ $$\int^b_a
105: P^i(z)q(z)\d z=\int^{P(b)}_{P(a)}y^i R^{\prime}(y)\d y=0.$$
106:
107:
108: More generally, the following ``composition condition" imposed
109: on $P(z)$ and $Q(z)=\int q(z)\d z$ is sufficient for polynomials
110: $P(z),q(z)$ to satisfy (1): {\it there exist polynomials $\tilde
111: P(z),$ $\tilde Q(z),$ $W(z)$ such that}
112: \be \l{2}
113: P(z)=\tilde P(W(z)), \
114: \ Q(z)=\tilde Q(W(z)), \ {\it and} \ W(a)=W(b).
115: \ee
116: The
117: sufficiency of condition (2) follows from $W(a)=W(b)$ after the
118: change of variable $z \rightarrow W(z)$. It was suggested in the
119: papers \cite{bfy1}-\cite{bfy5} (``the composition conjecture") that,
120: under the additional assumption $P(a)=P(b),$
121: condition (1) is actually equivalent to
122: condition (2). This conjecture is true in several special cases.
123: In particular, when $a,b$ are not critical points of $P(z)$
124: (\cite{c}), when $P(z)$ is indecomposable (\cite{pa2}),
125: and in some other cases (see \cite{pry}, \cite{ro}, and the papers
126: cited above).
127:
128: Nevertheless, in general the composition conjecture fails to be
129: true; a class of counterexamples to the composition conjecture
130: was constructed in \cite{pa1}. These counterexamples exploit
131: polynomials
132: which admit
133: ``double decompositions'' of the form
134: $P(z)=A(B(z))=C(D(z)),$ where $A(z),$ $B(z),$ $C(z),$ $D(z)$
135: are non-linear polynomials. If $P(z)$ is such a polynomial and,
136: in addition, the equalities $B(a)=B(b),$ $D(a)=D(b)$ hold, then
137: for any polynomial
138: $Q(z),$ which can be represented as $Q(z)=E(B(z))+F(D(z))$ for some
139: polynomials $E(z),F(z),$ condition (1) is satisfied with
140: $q(z)=Q^{\prime}(z)$ by linearity.
141: On the other hand, it can be shown (see \cite{pa1}) that if $\deg B(z)$ and
142: $\deg D(z)$ are coprime then condition (2) is not satisfied already for
143: $Q(z)=B(z)+D(z)$. The simplest explicit counterexample
144: to the composition conjecture has
145: the following form:
146: $$ P(z)=T_6(z), \ \ \ \ q(z)=T_2^{\prime}(z)+T_3^{\prime}(z),\ \ \ \
147: a=-\sqrt{3}/2. \ \ \ \ b=\sqrt{3}/2,$$
148: where $T_n(z)$ denotes $n$-th Chebyshev polynomial.
149: %Note that, by the second Ritt theorem,
150: %double decompositions with $\deg A(z)=\deg D(z),$
151: %$\deg B(z)=\deg C(z)$ and $\deg B(z),\deg D(z)$
152: %coprime
153: %are equivalent either to decompositions with $$A(z)=z^nR^m(z),\ \ \
154: %B(z)=z^m, \ \ \ C(z)=z^m,\ \ \ D(z)= z^nR(z^m)$$ for a polynomial $R(z)$
155: %and ${\rm GCD}(n,m)=1$ or to decompositions with $$A(z)=T_m(z), \ \ \ B(z)=
156: %T_n(z),\ \ \
157: %C(z)= T_n(z), \ \ \ D(z)=T_m(z)$$
158: %for Chebyshev polynomials $T_n(z),$ $T_m(z)$ and ${\rm GCD}(n,m)=1$ (see
159: %\cite{ri},
160: %\cite{sch}).
161:
162: The counterexamples above
163: suggest to transform the composition conjecture
164: as follows (\cite{pa4}): {\it non-zero polynomials $P(z),$ $q(z)$ satisfy
165: condition (1)
166: if and only if $\int q(z)\d z$ can be represented as a {\it sum} of
167: polynomials $Q_j$ such that
168: \be \l{cc}
169: P(z)=\tilde P_j(W_j(z)), \ \ \
170: Q_j(z)=\tilde Q_j(W_j(z)), \ \ \ {\it and} \ \ \ W_j(a)=W_j(b)
171: \ee
172: for some $\tilde P_j(z), \tilde Q_j(z), W_j(z)\in \C[z]$.}
173: Note that we do not make any additional assumptions on the
174: values of $P(z)$ at the points $a,b$ any more.
175: In particular, the conjecture
176: implies that non-zero polynomials orthogonal to
177: all powers of a given polynomial $P(z)$ on $[a,b]$ exist if and only
178: if $P(a)=P(b).$
179: For $P(z)=T_n(z)$
180: conjecture \eqref{cc} was verified in \cite{pa3}.
181:
182: This paper is organized as follows. We start from the description of
183: necessary and sufficient
184: conditions for polynomials $P(z),q(z)$ to satisfy \eqref{1} in terms
185: of linear relations between
186: branches of the algebraic function $Q(P^{-1}(z)),$ where $P^{-1}(z)$ denotes
187: the algebraic function which is inverse to $P(z).$ More precisely,
188: let $P^{-1}_{i}(z),$ $1\leq i \leq n,$ be single-valued branches of $P^{-1}(z)$
189: in a simply-connected domain $U\subset \C$
190: containing no critical values of $P(z).$
191: We show that there exists a system of
192: equations
193: \be \l{su}
194: \sum_{i=1}^nf_{s,i}Q(P^{-1}_{i}(z))=0, \ \ \ \ \ \ \ \ 1\leq s \leq \deg P(z),
195: \ee
196: where $f_{s,i}\in\{0,-1,1\},$
197: such that \eqref{1} holds if and only if \eqref{su} holds.
198: This system depends on $P(z),a,b$ and can be calculated explicitly
199: via a special graph $\lambda_P,$ called the ``cactus'' of $P(z),$
200: which contains
201: all the information about the monodromy of $P(z).$
202:
203:
204: The criterion
205: \eqref{su}
206: has a number of applications.
207: For example, it allows us
208: to reduce an infinite set of
209: equations \eqref{1} to a finite set of equations $v_k=0,$ ${0\leq k \leq M,}$
210: where $v_k$ are initial coefficients of
211: the Puiseux expansions of the combinations of branches in \eqref{su} and
212: $M$ depends only on degrees of $P(z)$ and $Q(z)$.
213: Furthermore, using the equivalence of \eqref{1} and \eqref{su},
214: we provide a variety of different conditions on a
215: collection $P(z),a,b$
216: under which conditions \eqref{1} and
217: \eqref{2} are equivalent;
218: such conditions are of interest because
219: of applications to the
220: Abel equation (see \cite{bby}, \cite{bry}, \cite{by}).
221: Essentially the finding of such conditions reduces to the finding conditions
222: under which system \eqref{su} implies that
223: \be \l{eb}
224: Q(P^{-1}_{i_1}(z))=Q(P^{-1}_{i_2}(z))
225: \ee
226: for some $i_1\neq i_2.$
227: In its turn these conditions can be naturally given in terms
228: of combinatorics of the graph $\lambda_P.$
229:
230:
231: While an explicit form of system \eqref{su} depends on
232: the collection $P(z),a,b,$
233: there exists a necessary condition for \e{1} to be satisfied the form
234: of which is invariant with respect to $P(z),a,b.$
235: Namely, it is known (\cite{pa2}, \cite{pry}) that \e{1}
236: implies the equality
237: \be \l{e1}
238: \frac{1}{d_{a}}\sum_{s=1}^{d_{a}}
239: Q(P_{a_s}^{-1}(z))=\frac{1}{d_{b}}\sum_{s=1}^{d_{b}}
240: Q(P_{b_s}^{-1}(z)),
241: \ee if $P(a)=P(b)$, or
242: the system
243: \be \l{e2}
244: \frac{1}{d_{a}}\sum_{s=1}^{d_{a}}
245: Q(P_{a_s}^{-1}(z))=0, \ \ \ \ \ \ \ \ \ \ \frac{1}{d_{b}}\sum_{s=1}^{d_{b}}
246: Q(P_{b_s}^{-1}(z))=0,
247: \ee
248: if $P(a)\neq P(b),$ where
249: $P_{a_1}^{-1}(z),$
250: $P_{a_2}^{-1}(z), ... , P_{a_{f_a}}^{-1}(z)$
251: (resp. $P_{b_1}^{-1}(z),$ $P_{b_2}^{-1}(z),$ ... , \linebreak
252: $P_{b_{f_b}}^{-1}(z)$) denote
253: the branches
254: of $P^{-1}(z)$
255: which map points close
256: to $P(a)$ (resp. $P(b)$)
257: to points close to $a$ (resp. $b$).
258:
259:
260: In the fourth section of this paper we investigate equation \eqref{e1} and
261: system \eqref{e2}. In particular, we establish a specific geometric
262: property of the monodromy groups of polynomials
263: from which we deduce that if \e{e1} or \e{e2} is satisfied
264: for $P(z),Q(z)\in \C[z],$ $\deg P(z)=n, \deg Q(z)=m,$
265: then for coefficients of the Puiseux expansions near infinity
266: \be \l{ps2}
267: Q(P^{-1}_j(z))=\sum_{k=-m}^{\infty}
268: u_k\varepsilon_n^{jk}z^{-\frac{k}{n}}
269: \ee
270: the equality $u_k=0$ holds whenever $\GCD(k,n)=1.$
271: This fact agrees with conjecture \eqref{cc} and, in particular,
272: implies that for $P(z),q(z)$ satisfying \eqref{1} the numbers
273: $n$ and $m$ can not be coprime.
274: As an application of our analysis of \eqref{e1},\eqref{e2}
275: we show in the fifth section that conditions \eqref{1} and \eqref{2}
276: are equivalent if at least one from points $a,b$ is not a critical
277: point of $P(z)$ or if $\deg P(z)=p^r$ for a prime number $p.$
278:
279: Finally,
280: on the base of obtained results,
281: in the sixth section we show that for any
282: collection $P(z),a,b$ with $\deg P(z)<10$
283: conditions \eqref{1}
284: and \eqref{2} are equivalent except the case when
285: $P(z),a,b$ is linearly equivalent to the collection
286: $T_6(z),- \sqrt{3}/2,\sqrt{3}/2.$
287: Since for $P(z)=T_n(z)$ all solutions to \eqref{1} were obtained
288: in \cite{pa3}, this provides
289: the complete solution of the polynomial moment problem for
290: $P(z),a,b$ with $\deg P(z)<10.$
291:
292:
293: \section{Criterion for a polynomial to be orthogonal to all
294: powers of a given polynomial}
295:
296: \subsection{Cauchy type integrals of algebraic functions}
297: A quite general approach to the polynomial moment problem
298: was proposed in the paper \cite{pry} concerning Cauchy type integrals of
299: algebraic functions
300: \be \l{ci}
301: I(t)=I(\gamma,g,t)=\frac{1} {2\pi i}\int_{\gamma} \frac{g(z)dz}
302: {z-t}\,.
303: \ee
304: In this subsection we briefly recall it (see \cite{pry} for
305: details) and outline in this context the approach of this paper.
306:
307:
308: First of all notice that condition \eqref{1} is equivalent
309: to the condition
310: \be \l{1+}
311: \int^b_a P^i(z)Q(z)P^{\prime}(z)\d z=0
312: \ee
313: for $i\geq 0,$ where $Q(z)=\int q(z) \d z$ is normalized
314: by the condition ${Q(a)=Q(b)=0}$
315: ($Q(a)$ always equals $Q(b)$
316: by \e{1} taken for $i=0$).
317: Furthermore, vice versa, condition \e{1+}
318: with $Q(a)=Q(b)=0$ implies that \e{1} holds with $q(z)=Q^{\prime}(z).$
319: (Actually, it was the condition
320: \eqref{1+} that appeared initially in the papers
321: on differential equations cited above).
322:
323: Indeed, condition \e{1+} is equivalent to the condition that
324: the function
325: $$
326: H(t)= \int_a^b \frac{Q(z)P^{\prime}(z)dz}{P(z)-t} \ \ \ \ \
327: $$ vanishes identically near infinity, since near infinity
328: $$H(t)=-\sum_{i=0}^{\infty}m_it^{-(i+1)}, \ \ \ \ {\rm where}
329: \ \ \ \ m_i=
330: \int^b_a P^i(z)Q(z)P^{\prime}(z)\d z.$$
331: On the other hand,
332: we have:
333: \begin{multline} \l{mul}
334: \frac{dH(t)}{dt}=\int_a^b
335: \frac{Q(z)P^{\prime}(z)dz}{(P(z)-t)^2}=
336: -\int_a^b Q(z)d\left(\frac{1}{P(z)-t}\right)=\\
337: =\frac{Q(a)}{P(a)-t}-\frac{Q(b)}{P(b)-t}+\tilde H(t),
338: \end{multline}
339: where
340: $$\tilde H(t)= \int_a^b \frac{q(z)dz}{P(z)-t}.$$ Since
341: near infinity
342: $$\tilde H(t)=-\sum_{i=0}^{\infty}\tilde m_it^{-(i+1)}, \ \ \ \ {\rm where}
343: \ \ \ \ \tilde m_i=
344: \int^b_a P^i(z)q(z)\d z,$$
345: it follows from \e{mul} that conditions \e{1} and \e{1+} are equivalent
346: whenever \linebreak
347: ${Q(a)=Q(b)=0.}$
348:
349: Furthermore, performing the change of variable
350: $z\rightarrow P(z),$
351: %the vanishing of $H(t)$
352: %near infinity becomes equivalent to
353: %the vanishing near infinity of Cauchy type integral
354: %\e{ci}, where $\gamma=P([a,b])$ and
355: %$g(z)$ is an algebraic function obtained by the analytic continuation
356: %of a germ of the algebraic function
357: %$g(z)=Q(P^{-1}(z))$ along $\gamma.$
358: we see that $H(t)$ coincides with integral \e{ci} where $\gamma=P([a,b])$ and
359: $g(z)$ is an algebraic function obtained by the analytic continuation
360: of a germ of the algebraic function
361: $g(z)=Q(P^{-1}(z))$ along $\gamma.$
362: Integral representation \eqref{ci}
363: defines a collection of univalent regular functions
364: $I_i(t);$ each $I_i(t)$ is defined in a domain $D_i$ of the complement
365: of $\gamma$ in $\C\P^1$. Denote by $I_{\infty}(t)$
366: the function defined in
367: the domain $D_{\infty}$ containing infinity. Then the vanishing of $H(t)$
368: near infinity becomes equivalent to the equality $I_{\infty}(t)\equiv 0.$
369:
370:
371: More generally,
372: consider integral \e{ci}, where $\gamma$ is a curve in the complex
373: plane $\C$ and $g(z)$ is any
374: ``piecewise-algebraic'' function on $\gamma$. More precisely, we assume that
375: after removing from $\gamma$ a finite
376: set of points $\Sigma_{\gamma},$
377: the set $\gamma\setminus \Sigma_{\gamma}$ is an union
378: of topological segments $\cup \gamma_s$ such that
379: for each $\gamma_s$ there exists a
380: domain $V_s\supset \gamma_s$ and an analytic in $V_s$
381: algebraic function $g_s(z)$ such that $g(z)$
382: on $\gamma_s$ coincides with
383: $g_s(z).$ Furthermore, we
384: assume that at the points of $\Sigma_{\gamma}$, complete analytic
385: continuations $\hat g_s(z)$ of $g_s(z)$ can ramify
386: but do not have
387: poles.
388: %Integral representation \eqref{ci}
389: %defines a collection of univalent regular functions
390: %$I_i(t);$ each $I_i(t)$ is defined in a domain $D_i$ of the complement
391: %of $\gamma$ in $\C\P^1$. Denote by $I_{\infty}(t)$
392: %the function defined in
393: %the domain $D_{\infty}$ containing infinity.
394: Below we sketch conditions for $I_{\infty}(t)$
395: to be a rational function; if these conditions are satisfied, then in
396: order to verify the equality
397: $I_{\infty}(t)\equiv 0$
398: it is enough to examine possible poles.
399:
400:
401:
402: Denote by $\Sigma_g$ the set
403: of all singularities of $\hat g_s(z)$
404: in $\C\P^1.$
405: Show that any element $(I_i(t),U_i)$ can be analytically continued
406: along any curve
407: $S=S_{t_1,t_2}$ connecting points $t_1,t_2 \in \C\P^1$
408: and avoiding points from the sets $\Sigma_g$
409: and $\Sigma_{\gamma}.$
410: First of all notice that if $t_2\in \partial U_i$ then
411: an analytical extension of $(I_i(t),U_i)$
412: to a domain containing $t_2$
413: is given simply by the integral
414: $I(\tilde\gamma,g,t),$ where $\tilde \gamma$ is a small
415: deformation of the $\gamma$
416: such that $t_2\in \tilde U_i.$
417: Furthermore, if $S=S_{t_1,t_2}$ is a simple curve connecting points
418: $t_1\in U_i,$ $t_2\in U_j,$ where
419: $U_i,$ $U_j$ are domains with a common segment of the boundary
420: $\gamma_{s}$ and $(g_s,V_s)$ is
421: the corresponding algebraic function,
422: then the well-known
423: boundary property of Cauchy type integrals (see e.g. \cite{mus}) implies that
424: $$(I_i(t),U_i\cap V_s)=(I_j(t),U_j\cap V_s)+(g_s,V_s).$$
425: Therefore, the analytic continuation of
426: $(I_i(t),U_i)$ along $S$ can be defined via
427: the analytic continuation of the right side of this formula.
428:
429: Finally,
430: for arbitrary domains $U_i,$ $U_j$ and
431: a curve $S=S_{t_1,t_2}$ connecting points $t_1\in U_i$ and $t_2\in U_j,$
432: the analytic continuation $(I_i(t),U_i)_S$ of
433: the element $(I_i(t),U_i)$ along $S$ can be defined inductively as follows.
434: Let $S\cap \gamma =\{c_1,c_2, ... ,c_r\},$
435: $c_s\in V_s,$ $1\leq s \leq r,$
436: and
437: let $(g_1,V_1), (g_2,V_2), ... , (g_r,V_r)$ be
438: the corresponding algebraic functions.
439: Define a germ $g_{\gamma,S}$ of an algebraic function near the point
440: $t_2$ by the formula:
441: $$g_{\gamma,S}=\sum_{i=1}^r (g_i,V_i)_{S_{c_i}}, $$
442: where $(g_i,V_i)_{S_{c_i}}$, $1\leq i \leq r,$ denotes the analytic
443: continuation
444: of the element $(g_i,V_i)$ (taken with the sign corresponding
445: to the orientation of the crossing of $S$ and $\gamma$)
446: along a part of $S$ from $c_i$ to $t_2.$
447: Then by induction we have:
448: \be \l{skl}
449: (I_i(t),U_i)_{S}=(I_j(t),U_j)+g_{\gamma,S}.
450: \ee
451: In particular, a complete analytic continuation $\hat I_i(t)$ of the element
452: $(I_i(t),U_i)$ is
453: a multi-valued analytic function with a finite set of
454: singularities $\Sigma_{\hat I_i} \subset \Sigma_g \cup \Sigma_{\gamma}.$
455:
456: From formula \e{skl} one deduces the following criterion
457: (\cite{pry}):
458: {\it $\hat I_i(t)$ is a rational function
459: if and only if the equality
460: \be \l{us}
461: g_{\gamma,S}=0
462: \ee
463: holds for any curve
464: $S=S_{t_1,t_2}$ as above with $t_1=t_2\in U_i.$}
465: Indeed, the necessity of
466: \e{us} is obvious. To establish the sufficiency
467: observe that \e{us} implies, in particular, that $\hat I_i(t)$ has
468: no ramification in its singularities.
469: Since
470: a Cauchy type integral near the end of the line of integration $u$
471: (and, similarly, near points of $\Sigma_{\gamma}$) is
472: a sum of a part which has a logarithmic branching at $u$
473: with a part that bounded at
474: $u$ (see e.g. \cite{mus}) and
475: $\hat g_s(z)$ do not have
476: poles at $\Sigma_{\gamma},$
477: this fact implies that the singularities of $\hat I_i(t)$
478: located on $\gamma$ are removable.
479: On the other hand, if a singularity
480: $t_0$ of $\hat I_i(t)$ is contained in $\C\P^1\setminus \gamma,$ then
481: formula \e{skl} implies that
482: $t_0\in \Sigma_g.$ Since $\hat I_i(t)$ has no ramification
483: at $t_0$ it follows that in this case $t_0$
484: is a pole the worst.
485:
486: Although the method above in principle is constructive
487: its practical application is rather
488: difficult since the calculation of sums $g_{\gamma,S}$
489: is complicated.
490: In this paper we propose a modification
491: of the method above
492: designed
493: specially for the polynomial moment problem.
494: This modification
495: permits to avoid any analytic continuations
496: and allows us to obtain a necessary and sufficient conditions
497: for equality \e{1} to be satisfied in
498: a closed and convenient form.
499: The idea is
500: to choose a very special way of integration $\Gamma$ connecting points
501: $a,b$ (we can use any of them since integrals \e{1}
502: do not depend on $\Gamma$). It turns out that
503: $\Gamma$ can by chosen
504: so that $\C\P^1\setminus P(\Gamma)$ consists of
505: a {\it unique} domain. Then condition $I_{\infty}(t)\equiv 0$
506: simply reduces to the condition that
507: the corresponding algebraic functions $g_s(z)$ vanish on $P(\Gamma).$
508: Furthermore, we choose $\Gamma$ as a subset of a special tree $\lambda_P,$
509: called the cactus of $P(z),$ which contains
510: all the information about the monodromy $P(z).$
511: This allows us to relate properties of the collection
512: $P(z),a,b$ which are connected to the polynomial moment
513: problem to combinatorial properties of the pair consisting of
514: the tree $\lambda_P$ and the
515: path
516: connecting points $a,b$ on $\lambda_P.$
517:
518:
519: \subsection{Cacti} \l{cacti}
520: To visualize the monodromy group of a polynomial $P(z)$
521: it is convenient to consider a graphical object $\lambda_P$ called the
522: {\it cactus} of $P(z)$ (see e. g. \cite{cac1}, \cite{cac2}).
523:
524: Let $c_1,c_2, ... ,c_k$ be all
525: finite critical values of $P(z)$ and let $c$ be a not critical value. Draw
526: a star $S$ joining $c$ with $c_1,c_2, ... ,c_k$ by non intersecting arcs
527: $\gamma_1, \gamma_2, ... ,\gamma_k.$ We will suppose that $c_1,c_2, ...
528: ,c_k$ are numerated in such a way that in a counterclockwise rotation
529: around $c$ the arc $\gamma_s$, $1 \leq s \leq k-1,$ is followed by the arc
530: $\gamma_{s+1}.$ By definition, the cactus $\lambda_P$ is the preimage of $S$
531: under the map $P(z)\,:\, \C \rightarrow \C.$ More precisely, we consider
532: $\lambda_P$ as a $k+1$-colored graph embedded into the Riemann sphere:
533: vertices of $\lambda_P$ colored by the $s$-th color, where $1\leq s \leq
534: k,$ are preimages of the point $c_s,$ vertices colored by the $k+1$-th
535: color (to
536: be definite we will suppose that it is the white color) are
537: preimages
538: of the point $c,$ and edges of $\lambda_P$ are preimages of the arcs
539: $\gamma_s,$ $1 \leq s \leq k.$ It is not difficult to show that the graph
540: $\lambda_P$ is connected and has no cycles. Therefore,
541: $\lambda_P$ is a plane tree.
542:
543: The valency of a non-white vertex $z$ of $\lambda_P$
544: coincides with the multiplicity of $z$ with respect to $P(z)$ while
545: all white vertices of $\lambda_P$ are of the same valency
546: $n=\deg P(z).$ The set of all edges of $\lambda_P$ adjacent to a
547: white vertex $w$ is called by a {\it star} of $\lambda_P$ centered at $w.$
548: Clearly, $\lambda_P$ has $nk$ edges and $n$ stars. The set of stars
549: of $\lambda_P$ is naturally identified with the set of branches of $P^{-1}(z)$
550: as follows.
551: Let $U$ be a
552: simply connected domain containing no critical values of $P(z)$ such
553: that $S\setminus\{c_1,c_2,
554: ... ,c_k\}\subset U.$ By the monodromy theorem in $U$ there exist
555: $n$ single valued branches
556: of $P^{-1}(z).$
557: Any such a branch $P_i^{-1}(z),$ $1\leq i \leq n,$
558: maps $S\setminus\{c_1,c_2, ... ,c_k\}$ into a star of
559: $\lambda_P$ and we will label the corresponding star by the symbol $S_i$
560: (see Fig. 1).
561:
562:
563: \begin{figure}
564: \medskip
565: \epsfxsize=12truecm
566: \centerline{\epsffile{1.eps}}
567: \smallskip
568: \centerline{Figure 1}
569: \medskip
570: \end{figure}
571:
572: The cactus $\lambda_P$ permits to reconstruct the monodromy group
573: $G_P$ of $P(z).$ Indeed, $G_P$ is generated by
574: the permutations $g_{s}\in S_n,$ $1\leq s \leq k,$ where
575: $g_{s}$ is defined by the condition
576: that the analytic continuation of
577: the element $(P_i^{-1}(z),U),$ $1\leq i \leq n,$ along a counterclockwise
578: oriented loop $l_s$ around $c_s$
579: is the element $(P_{g_s(i)}^{-1}(z),U).$ Having in mind the identification of
580: the set of stars of $\lambda_P$ with the set of branches of $P^{-1}(z),$
581: the permutation $g_s,$ $1\leq s \leq k,$ can be
582: identified with the permutation $\hat g_s,$
583: $1\leq s \leq k,$
584: acting
585: on the set of starts of $\lambda_P$ in the following way: $\hat g_s$ sends the
586: star $S_i,$ $1\leq i \leq n,$ to the ``next'' one in a counterclockwise
587: direction around its vertex of color $s.$ For example,
588: for the cactus shown on Fig. 1 we have: $g_1=(1)(2)(37)(4)(5)(6)(8),$
589: $g_2=(1)(2)(3)(47)(56)(8),$ $g_3=(1238)(4)(57)(6).$
590:
591: Note that since $P(z)$ is a polynomial, the permutation
592: $g_{\infty}=g_1g_2 ... g_k$ is a cycle of
593: length $n.$ Usually, we will choose the numeration of $S_i,$ $1\leq i \leq
594: n,$ in such a way that this cycle coincides with the cycle $(12...n).$
595:
596:
597: \subsection{Criterion}\l{criter} In this subsection
598: we give explicit necessary and sufficient conditions for $P(z),$ $q(z)\in
599: \C[z]$ and $a,b\in \C,$ $a\neq b,$ to satisfy \e{1}, \e{1+}.
600: For this propose we choose the way of integration $\Gamma_{a,b}$ connecting
601: $a,b$ so
602: that $\Gamma_{a,b}$ would be a subset of $\lambda_P.$
603:
604:
605: More precisely, for any $P(z)\in \C[z]$ and $a,b\in \C$ let us define
606: an {\it extended} cactus
607: $\tilde \lambda_P=\tilde \lambda_P(c_1,c_2,\, ... \,,c_{\tilde k})$
608: as follows. Let $c_1,c_2, ... ,c_{\tilde k}$ be all finite critical values
609: of $P(z)$ complemented by $P(a)$ or $P(b)$ (or by both of them) if $P(a)$
610: or $P(b)$ is not a critical value of $P(z).$ Consider an extended star
611: $\tilde S$ connecting $c$ with $c_1,c_2, ... ,c_{\tilde k}$ and set
612: $\tilde \lambda_P=P^{-1}\{\tilde S\}$ (we suppose that $c$ is chosen
613: distinct from $P(a),P(b)$). Clearly, $\tilde \lambda_P$ considered as a
614: $\tilde k+1$ colored graph is still connected and has no cycles.
615: Furthermore,
616: by construction the points $a,b$ are vertices of $\tilde \lambda_P.$ Since
617: $\tilde \lambda_P$ is connected there exists an oriented path $\Gamma_{a,b}
618: \subset \tilde \lambda_P$
619: with the starting point $a$ and the ending point $b.$ Moreover, since
620: $\tilde \lambda_P$ has no cycles there exists exactly one such a path. We
621: choose $\Gamma_{a,b}$ as a new way of integration.
622:
623: Let $U$ be a domain as above and let $Q(z)=\int q(z) \d z$ be normalized
624: by the condition ${Q(a)=Q(b)=0}.$
625: For each
626: $s,$ $1\leq s \leq \tilde k,$ define a linear combination
627: $\phi_s(z)$ of branches $Q(P^{-1}_{i}(z)),$ $1\leq i \leq n,$ in $U$
628: as follows. Set
629: \be \l{piz}
630: \phi_s(z)=\sum_{i=1}^nf_{s,i}Q(P^{-1}_{i}(z)),
631: \ee
632: where $f_{s,i}\neq 0$ if and only if the path
633: $\Gamma_{a,b}$ passes through a vertex $v$ of the star
634: $S_i$
635: colored by the
636: $s$-th color
637: (we do not take into account the stars $S_i$ for which $\Gamma_{a,b}\cap S_i$
638: contains only the point $v$). Furthermore, if under a moving along $\Gamma_{a,b}$ the vertex $v$
639: is followed by the center of $S_i$ then $f_{s,i}=-1$ otherwise
640: $f_{s,i}=1$. As an example consider the cactus shown on Fig. 1. Then
641: for the path
642: $\Gamma_{a,b}$ pictured by the fat
643: line we
644: have: $$\phi_1(z)=
645: -Q(P^{-1}_{2}(z))+Q(P^{-1}_{3}(z))- Q(P^{-1}_{7}(z)),$$ $$\phi_2(z)=
646: Q(P^{-1}_{7}(z))-Q(P^{-1}_{4}(z)),$$ $$\phi_3(z)=
647: Q(P^{-1}_{2}(z))-Q(P^{-1}_{3}(z))+ Q(P^{-1}_{4}(z)).$$
648: \vskip 0.1cm
649: \bt \l{t1}
650: Let $P(z),q(z)\in \C[z]$, $a,b\in \C,$ $a\neq b,$
651: and let $\tilde \lambda_P(c_1,c_2,\, ... \,,c_{\tilde k})$ be an extended
652: cactus corresponding to the collection $P(z),a,b.$
653: Then
654: \eqref{1} holds if and only if
655: the equality $\phi_s(z)\equiv 0$ holds
656: in $U$ for any $s,$ $1\leq s \leq \tilde k.$
657: \et
658:
659: \pr Indeed,
660: condition \eqref{1+} is equivalent to the condition that the function
661: $$
662: H(t)= \int_{\Gamma_{a,b}} \frac{Q(z)P^{\prime}(z)dz}{P(z)-t} \ \ \ \ \
663: $$ vanishes identically near infinity.
664: On the other hand, using
665: the change of variable $z\rightarrow P(z),$ we can express the function
666: $H(t)$ as a sum of Cauchy type integrals of algebraic functions as follows:
667: \be \l{f}
668: H(t)=\sum_{s=1}^{\tilde k}\int_{\gamma_s}\frac{\phi_s(z)}{z-t}\, \d
669: z.
670: \ee
671: Since this formula implies that $H(t)$ is analytic in a domain
672: $V=\C\P^1\setminus S$ we see that the vanishing of $H(t)$
673: near infinity is equivalent to the
674: condition that $H(t)\equiv 0$ in $V.$
675:
676: Let $z_0$ be an interior point of $\gamma_s,$ $1\leq s \leq \tilde k.$
677: Then by the well-known boundary property of
678: Cauchy type integrals (see e.g. \cite{mus}) we have:
679:
680: $$
681: \lim_{t \to z_0}\!\!\!\!\,^+H(t)-\lim_{t \to z_0}
682: \!\!\!\!\,^-H(t)=\phi_s(t_0),
683: $$
684: where the limits are taken respectively for $t$
685: tending to $z_0$ from the ``left'' and
686: from the ``right'' parts of $V$ with respect to $\gamma_s.$
687: If $H(t)\equiv 0$ in $V,$ then
688: $$
689: \lim_{t \to z_0}\!\!\!\!\,^+H(t)=\lim_{t \to z_0}
690: \!\!\!\!\,^-H(t)=0,
691: $$
692: and, therefore, $\phi_s(z_0)= 0.$
693: Since this equality holds for any
694: interior point $z_0$ of any arc $\gamma_s,$ $1\leq s \leq \tilde k,$
695: we conclude that
696: $\phi_s(z)\equiv 0,$ $1\leq s \leq \tilde k,$ in $U.$ On the other hand,
697: if $\phi_s(z)\equiv 0,$ $1\leq s \leq \tilde k,$ in $U,$ then
698: it follows directly from formula \e{f} that
699: $H(t)\equiv 0$ in $V.$
700:
701:
702: \vskip 0.2cm
703: Note that some of equations $\phi_s(z)\equiv 0,$ $1\leq s \leq \tilde k,$
704: could be trivial.
705: This happens exactly for the values $s$ such that
706: the path $\Gamma_{a,b}$ does not pass through vertices
707: colored by the $s$-th color.
708: Note also that
709: equations \e{piz}
710: are linearly dependent. Indeed,
711: for each $i$ such that there exists an index $s,$
712: $1\leq s
713: \leq \tilde k,$ with $f_{s,i}\neq 0$ there exist exactly two such
714: indices $s_1,s_2$ and $c_{s_1,i}=-c_{s_2,i}.$ Therefore, the equality
715: $$\sum_{s=1}^{\tilde k}\phi_s(t)=0$$
716: holds in $U$.
717:
718:
719: \subsection{Checking the criterion}
720: Let $P(z),Q(z)\in \C[z],$ $\deg P(z)=n,$ $\deg Q(z)=m.$
721: Let $U$ be a simply connected domain containing no critical values of $P(z)$
722: and let $P_i^{-1}(z),$ $1\leq i \leq n,$ be branches of $P^{-1}(z)$
723: in $U.$
724: In this subsection we provide a simple estimation
725: for the order of a zero in $U$ of a function of the form
726: $$\psi(z)=\sum_{i=1}^nf_iQ(P^{-1}_i(z)), \ \ \ \ \ \ \ \ \ \ \ \ f_i\in \C,$$
727: via the degrees of $P(z)$ and $Q(z).$ This reduces
728: the verification of the criterion
729: to the calculation of a finite set of initial coefficients of
730: Puiseux expansions of functions \e{piz}
731: and, as a corollary, provides a practical method for checking an infinite
732: set of equation \e{1} in a finite number of steps.
733:
734:
735: \bl\l{ur} If $\psi(z)\neq 0$ then $\psi(z)$
736: satisfies an equation
737: \be \l{urr}
738: y^N(z)+a_1(z)y^{N-1}(z)+\ ... \ + a_N(z)=0,
739: \ee
740: where $a_j(z)\in \C[z],$
741: $a_N(z)\neq 0,$ and $N\leq n!.$
742: Furthermore,
743: $\deg a_j(z)\leq \left(\frac{m}{n}\right)^{j},$ $1\leq j \leq N.$
744: \el
745:
746: \pr Indeed, if $\psi(z)\neq 0$ then, since $\psi(z)$ is a sum of algebraic
747: functions, $\psi(z)$ itself is an algebraic function and
748: therefore satisfies an algebraic equation \e{urr}
749: with $a_i(z)\in \C(z),$ $1\leq j \leq N.$ Furthermore, we can
750: suppose that this equation
751: is irreducible. Then $a_N(z)\neq 0$
752: and the number $N$
753: coincides with the number of different analytic continuations
754: $\psi_j(z)$ of
755: $\psi(z)$ along closed curves. Clearly,
756: $N$ can be bounded by the number $N_1$ of different elements of
757: the monodromy group of $P(z).$ In its turn, $N_1$ is bounded by the
758: number of elements of the full symmetric group $S_n.$ Hence, $N\leq n!.$
759:
760: Furthermore, since $P(z),Q(z)$ are polynomials, the rational functions
761: $a_j(z),$ $1\leq j \leq N,$ as the symmetric functions
762: of $\psi_j(z),$ $1\leq j \leq N,$ have no poles in $\C$ and therefore
763: are polynomials. Finally, since near infinity branches $P^{-1}_i(z),$
764: $1\leq i \leq n,$
765: of $P^{-1}(z)$
766: are represented
767: by the Puiseux series
768: \be
769: P^{-1}_i(z)=\sum_{k=-1}^{\infty}v_k\varepsilon_n^{ik}z^{-\frac{k}{n}},
770: \ \ \ \ \ \ \ \ v_k\in \C, \ \ \ \ \ \ \ \ \varepsilon_n=exp(2\pi i/n),
771: \ee
772: the first non-zero exponent of the Puiseux series
773: at infinity for the functions
774: $\psi_j(z),$ $1\leq j \leq N,$
775: is less or equal than $m/n.$ It follows that
776: $\deg a_j(z)\leq \left(\frac{m}{n}\right)^{j},$ $1\leq j \leq N.$
777:
778: \bc Let $z_0 \in U.$ To verify that $\psi(z) \equiv 0$ it is enough
779: to check that the first $\left(\frac{m}{n}\right)^{n!}+1$ coefficients of
780: the series
781: $\psi(z)=\sum_{k=0}^{\infty} v_k (z-z_0)^k$ vanish.
782:
783: \ec
784:
785: \pr Indeed, suppose that $\ord_{z_0}\psi(z) > \left(\frac{m}{n}\right)^{n!}$
786: but $\psi(z)\neq 0.$ Then,
787: by lemma \ref{ur}, $\psi(z)$ satisfies \e{urr}, where
788: $\deg a_j(z)\leq \left(\frac{m}{n}\right)^{j}\leq \left(\frac{m}{n}
789: \right)^{n!},$ $1\leq j \leq N,$ and $a_N\neq 0.$
790: It follows that
791: $$\ord_{z_0}\{\psi^N(z)\} >
792: \ord_{z_0}\{a_{i_1}(z)\psi^{N-i_1}(z)\} >\ ... \ >\ord_{z_0}\{a_{i_k}(z)
793: \psi^{N-i_k}(z)\},$$ where $0 < i_1 < i_2 <\ ... \ <i_k=N$ are all indices
794: for which $a_i(z) \neq 0.$
795: Therefore,
796: $$\ord_{z_0}\{\psi^N(z)+
797: a_1(z)\psi^{N-1}(z) +\ ... \ +a_{N}(z)\}
798: =\ord_{z_0}\{a_{N}(z)\}<\infty $$ in contradiction with equality
799: \e{urr}.
800:
801:
802:
803: \section{Application to a description of definite polynomials}
804: In this section, as a first
805: application of theorem \ref{t1}, we provide
806: a number of conditions
807: on a collection $P(z),a,b,$ where $P(z)\in \C[z],$ $a,b\in \C,$
808: $a\neq b,$ under which conditions \e{1} and \e{2}
809: are equivalent; such collections are called {\it definite} and
810: are of interest because of applications to the Abel equation (see \cite{bby},
811: \cite{bry}, \cite{by}).
812:
813: \subsection{A combinatorial condition for a change of variable}
814:
815: The simplest form of the equality
816: $\phi_s(z)=0$
817: is equality \e{eb}.
818: Furthermore, \e{eb} has a clear compositional meaning.
819:
820:
821: \bl \l{lcomp} The equality \eqref{eb} holds if and only if
822: \be \l{comp}
823: P(z)=\tilde P(W(z)), \ \ \ \ \ \ \ \ Q(z)=\tilde Q(W(z))
824: \ee
825: for
826: some polynomials $\tilde P(z),$ $\tilde Q(z),$ $W(z)$ with $\deg W(z)>1.$
827: \el
828:
829: The proof of this lemma easily follows from the L\"{u}roth theorem
830: (see e.g. \cite{pa2}, \cite{ro}).
831: If condition \e{comp} is satisfied we say that
832: {\it polynomials $P(z),$ $Q(z)$ have a (non-trivial)
833: common right divisor in the composition algebra of polynomials.}
834:
835: Below we give a convenient combinatorial
836: condition on a collection $P(z),a,b$ which implies that for any
837: $q(z)$ satisfying \e{1} polynomials $P(z),$ $Q(z)=\int q(z)\d z$ have a
838: common right divisor in the composition algebra of polynomials.
839: The use of this condition permits, after the change of variable
840: $z\rightarrow W(z),$
841: to reduce the solution of the polynomial
842: moment problem for a polynomial $P(z)$ to that for a polynomial
843: of lesser degree $\tilde P(z).$
844:
845: Let $\tilde \lambda_P$ be a $\tilde k+1$ colored extended cactus
846: corresponding to a collection $P(z),a,b$ and let
847: $\Gamma_{a,b}\subset \lambda_P$
848: be the path connecting points $a,b$ on $\tilde \lambda_P.$
849: For each $s,$ $1\leq s \leq \tilde k,$
850: define the weight $w(s)$ of the $s$-th color
851: on $\Gamma_{a,b}$ as a number of vertices $v\in \Gamma_{a,b}$
852: colored by the $s$-th color with the convention that vertices $a,b$ are
853: counted with the coefficient $1/2.$ For example, for $\Gamma_{a,b}$
854: shown on Fig. 1 we have $w(1)=w(3)=3/2,$ $w(2)=1.$
855:
856: \bt \l{rt} Let $P(z),q(z)\in \C[z]$, $q(z)\neq 0,$
857: $a,b\in \C,$ $a\neq b$ satisfy \e{1}. Suppose that there exists
858: $s,$ $1\leq s \leq \tilde k,$ such that $w(s)=1$ on $\Gamma_{a,b}.$
859: Then $P(z),Q(z)$ have a common right divisor in the composition algebra.
860: \et
861: \pr Indeed, the construction of $\Gamma_{a,b}$ implies that if
862: $w(s)=1,$ then $f_{s,i}\neq 0$
863: exactly for two values $i_1,i_2,$ $1\leq i_1,i_2 \leq n.$ Moreover, for these values we have
864: $c_{s,{i_1}}=-c_{s,{i_2}}$ and
865: hence the equality $\phi_s(z)=0$
866: reduces to \e{eb}. Therefore,
867: $P(z)$ and $Q(z)$ have a common right divisor in the composition algebra
868: by lemma \ref{lcomp}.
869:
870: \subsection{Reduction}
871:
872:
873:
874: Although condition \eqref{comp} in general is weaker than condition \eqref{2}
875: it turns out that in order to prove
876: that for any collection
877: $P(z),a,b,$ $a\neq b,$
878: satisfying some condition $\mathcal R$
879: conditions \eqref{1} and \eqref{2} are equivalent
880: it is often enough to show
881: that for any such a collection
882: condition \e{1} implies condition \e{comp}.
883: Say that a condition $\mathcal R$
884: is {\it compositionally stable} if for any
885: collection $P(z),a,b,$ $a\neq b,$ satisfying $\mathcal R$
886: such that $P(z)=\tilde P(W(z))$ for
887: some $\tilde P(z), W(z)\in \C[z],$ ${\deg W(z) >1,}$ $W(a)\neq W(b),$
888: the collection $\tilde P(z),W(a),$ $W(b)$ also
889: satisfies $\mathcal R.$ For instance, the following condition is
890: compositional stable: at least one point from $a,b$ is
891: not a critical point of $P(z).$ An other example of a
892: compositional stable condition
893: is the following one: $\deg P(z)=p^r,$ where $p$ is a prime.
894:
895:
896:
897: \vskip 0.2cm
898:
899: \bl \l{rl}
900: Let $\mathcal R$ be
901: a compositionally stable condition.
902: Suppose that for any collection $P(z),a,b,$ $a\neq b,$
903: satisfying $\mathcal R,$
904: condition \e{1} implies condition \e{comp}.
905: Then for any collection $P(z),a,b,$ $a\neq b,$
906: satisfying $\mathcal R$ conditions \e{1} and \e{2} are equivalent.
907: \el
908:
909: \vskip 0.2cm
910:
911: \pr
912: Let $P(z),a,b$ be a collection satisfying $\mathcal R.$
913: Suppose that \e{1} holds for some $q(z)\in \C[z].$ Then it follows from
914: \e{comp} that $\C(P,Q)$ is a proper
915: subfield of $\C(z).$ Therefore, by the L\"{u}roth theorem
916: \be\l{lu}
917: \C(P,Q)=\C(W_1)
918: \ee for some rational function $W_1(z),$ $\deg W_1(z) > 1,$
919: and without loss of generality we can assume that
920: $W_1(z)$ is a polynomial. To prove the lemma it is enough to show
921: that the equality $W_1(a)=W_1(b)$ holds.
922:
923: Observe that equality \e{lu} is equivalent to the statement
924: that
925: \be \l{comp+}
926: P(z)=P_1(W_1(z)), \ \ \ \ \ \ \ \ Q(z)=Q_1(W_1(z))
927: \ee
928: for
929: some polynomials $P_1(z),$ $Q_1(z)$ such that
930: $P_1(z)$ and $Q_1(z)$ have no a
931: common right divisor in the composition algebra.
932: On the other hand,
933: performing the change of variable $z\rightarrow W_1(z)$
934: we see that condition \e{1} is satisfied also for
935: $P_1(z), Q_1^{\prime}(z), W_1(a),W_1(b).$
936: Therefore, if $W_1(a)\neq W_1(b),$ then,
937: taking into account that $\mathcal{R}$ is compositionally stable,
938: we find that $$P_1(z)=P_2
939: (W_2(z)), \ \ \ \ \ \ \ Q_1(z)=
940: Q_2(W_2(z))$$ for some $P_2(z), Q_2(z), W_2(z)\in \C[z]$ with
941: $\deg W_2(z) >1$ that contradicts the fact that
942: $P_1(z),$ $Q_1(z)$ have no a
943: common right divisor in the composition algebra.
944:
945:
946: \vskip 0.2cm
947:
948:
949:
950:
951:
952: \vskip 0.2cm
953:
954:
955:
956:
957:
958: \subsection{Description of some classes of definite
959: polynomials}
960:
961: As a first application of theorem
962: \ref{rt} and lemma \ref{rl} we give a simple proof
963: of the following assertion conjectured in \cite{pry}.
964:
965: \bc \l{c1} Let $P(z),q(z)\in \C[z]$, $q(z)\neq 0,$
966: $a,b\in \C,$ $a\neq b.$ Suppose that
967: $P(a)=P(b)=c_1$ and that all the
968: points of the preimage
969: $P^{-1}(c_1)$ except possibly $a,b$ are not critical points
970: of $P(z).$ Then conditions \e{1} and \e{2} are equivalent.
971: \ec
972:
973: \pr
974: Since the chain rule implies that
975: the condition of the corollary is compositionally
976: stable it is enough to show that
977: $P(z),$ $Q(z)$ have a
978: common right divisor in the composition algebra.
979: To establish it observe that $\Gamma_{a,b}$ can not pass through
980: vertices of $\lambda_P$ of the valency $1$ distinct from $a,b.$
981: Therefore, the
982: condition of the corollary implies
983: that $w(1)=1.$ It follows now from theorem \ref{rt} that
984: $P(z),$ $Q(z)$ have a
985: common right divisor in the composition algebra.
986:
987: \vskip 0.2cm A slight modification of the idea used in the proof
988: of corollary \e{c1}
989: leads to the following
990: statement.
991:
992:
993: \bc Let $P(z),q(z)\in \C[z]$, $q(z)\neq 0,$
994: $a,b\in \C,$ $a\neq b.$ Suppose that
995: $P(a)=P(b)=c_1$ and that for any critical
996: value $c$ of $P(z)$ except possibly $c_1$ the preimage $P^{-1}(c)$
997: contains only one critical point. Then conditions \e{1} and \e{2} are
998: equivalent. \ec
999:
1000: \pr
1001: Again, it follows from the chain rule that
1002: the condition of the corollary is compositionally stable.
1003: Furthermore, observe that
1004: the path $\Gamma_{a,b}$ contains at least one vertex $v$
1005: of a color $s\neq 1.$
1006: Since $\Gamma_{a,b}$ can not pass through vertices of the
1007: valency
1008: $1$ distinct from $a,b,$ it follows from the condition
1009: of the corollary that
1010: the equality
1011: $w(s)=1$ holds and, therefore, by theorem \ref{rt},
1012: $P(z),$ $Q(z)$ have a
1013: common right divisor in the composition algebra.
1014:
1015: \vskip 0.2cm
1016:
1017: Finally, we give a new proof of an
1018: assertion from the paper \cite{pry}
1019: which provides some geometric condition for a collection $P(z),a,b$
1020: to be definite. It turns out that
1021: this assertion actually also can be regarded as a particular
1022: case of theorem \ref{rt}.
1023: For a curve $M$
1024: denote by $V_{M,\infty}$
1025: the domain from the collection of domains $\C\P^1\setminus M$ which contains infinity.
1026: For an oriented curve $L$ and points
1027: $d_1,d_2\in L$
1028: denote by $L_{d_1,d_2}$ the part of
1029: $L$ between $d_1$ and $d_2.$
1030:
1031: \bc Let $P(z),q(z)\in \C[z]$, $q(z)\neq 0,$
1032: $a,b\in \C,$ $a\neq b.$ Suppose that
1033: $P(a)=P(b)=c_1$ and that
1034: there exists a curve $L$
1035: connecting points $a,b$
1036: such that $c_1$
1037: is a simple point of $P(L)$ and $c_1\in \partial V_{P(L),\infty}.$
1038: Then conditions \e{1} and \e{2} are
1039: equivalent.
1040: \ec
1041:
1042:
1043: \pr
1044: We will keep the notation introduced in subsection \ref{cacti}
1045: and \ref{criter}.
1046: Let $a^+$ (resp. $b^-$) be a point on $L$ near the point $a$
1047: (resp. $b$) and
1048: let $U$ be a
1049: simply connected domain containing no critical values of $P(z)$ such
1050: that the sets $S\setminus\{c_1,c_2,... ,c_{\tilde k}\},$
1051: $P(L_{a,a^+})\setminus c_1,$
1052: and $P(L_{b^-,b})\setminus c_1$ are subsets of
1053: $U.$ Recall that there is a natural correspondence between branches
1054: $P_i^{-1}(z),$ $1\leq i \leq n,$ of
1055: $P^{-1}(z)$ in $U$ and stars of the cactus $\lambda_P:$
1056: branch $P_i^{-1}(z)$ maps $U$ on a domain $U_i$ containing $S_i.$
1057: Denote by $U_{j_1}$ (resp. $U_{j_2}$) the domain containing the point $a^+$
1058: (resp. $b^-$). Then by construction the result of the analytic continuation
1059: of the element $(P^{-1}_{j_1}(z),U)$ along the curve
1060: $P(L_{a^+,b^-})$ is the element $(P^{-1}_{j_2}(z),U).$
1061: \begin{figure}
1062: \medskip
1063: \epsfxsize=4.5truecm
1064: \centerline{\epsffile{gg.eps}}
1065: \smallskip
1066: \centerline{Figure 2}
1067: \medskip
1068: \end{figure}
1069: Let $c_0$ be an interior point of $U$ close to $c_1.$
1070: Consider a small deformation $M$ of the curve $P(L_{a,b})$ obtained
1071: as follows: change the part of $P(L_{a,b})$ connecting $c_1$
1072: and $P(a^+)$
1073: to an arc $\gamma^+\subset U$ connecting $c_0$ with $P(a^+)$
1074: and, similarly, change the part of $P(L_{a,b})$ connecting
1075: $P(b^-)$ and $c_1$
1076: to an arc $\gamma^-\subset U$ connecting $P(b^-)$ and $c_0$ (see Fig. 2).
1077:
1078: Let now $l_M=l_{i_1}^{j_1}l_{i_2}^{j_2}\, ...\, l_{i_r}^{j_r}$
1079: be the image of the curve $M$ in
1080: $\pi_1(X, c_0),$ where $X=\C\P^1\setminus \{c_1,c_2,c_3, \, ... \,
1081: c_{k}\}.$
1082: Since the result of the analytic continuation
1083: of the element $(P^{-1}_{j_1}(z),U)$ along the curve
1084: $M$ is still the element $(P^{-1}_{j_2}(z),U),$
1085: the final element of the chain of stars
1086: $$\Omega=<S_{j_1}\,,\;S_{g_{i_1}^{j_1}(j)}\,,\;
1087: S_{g_{i_1}^{j_1}g_{i_2}^{j_2}(j)}
1088: \,,\;
1089: S_{g_{i_1}^{j_1}g_{i_2}^{j_2}g_{i_3}^{j_3}(j)}\,,\;...\;,\;
1090: S_{g_{i_1}^{j_1}g_{i_2}^{j_2}\,...\; g_{i_r}^{j_r}(j)}>$$
1091: is the star $S_{j_2}.$
1092: In particular, the path $\Gamma_{a,b}$ is contained in
1093: $\Omega.$
1094: Since $c_1\in V_{M,\infty},$
1095: the loop $l_1$ does not appear
1096: among the loops $l_{i_1},l_{i_2},\,...\,l_{i_r}.$
1097: Therefore, the common vertex of any two successive stars in the chain
1098: $\Omega$ is not contained in the set $P^{-1}(c_1).$
1099: In particular, among of vertices of $\Gamma_{a,b}$ there are no preimages
1100: of $c_1$ distinct from $a,b$ and hence
1101: $w(1)=1$ on $\Gamma_{a,b}.$
1102:
1103: To finish the proof notice that the condition of the
1104: corollary is compositionally stable. Indeed, if
1105: $L$ is a curve connecting points $a,b$
1106: such that $c_1=P(a)=P(b)$
1107: is a simple point of $P(L)$ and $c_1\in \partial V_{P(L),\infty},$
1108: then $W(L)$ is a curve connecting points $W(a),W(b)$
1109: such that $c_1=\tilde P(W(a))=\tilde P(W(b))$
1110: is a simple point of $\tilde P(W(L))$ and $c_1\in \partial V_{\tilde
1111: P(W(L)),\infty}.$
1112:
1113:
1114: \vskip 0.2cm
1115:
1116:
1117:
1118:
1119:
1120: \section{On functional equations
1121: \e{e1} and \e{e2}}
1122:
1123: \subsection{Derivation of \e{e1} and \e{e2}}
1124: Let $U$ be a simply connected domain containing no critical values of $P(z)$
1125: such that $S\setminus\{c_1,c_2, ... ,c_{\tilde k}\}\subset U.$
1126: Denote by
1127: $P_{a_1}^{-1}(z),$
1128: $P_{a_2}^{-1}(z), ... , P_{a_{f_a}}^{-1}(z)$
1129: (resp. $P_{b_1}^{-1}(z),$ $P_{b_2}^{-1}(z),$ ... ,
1130: $P_{b_{f_b}}^{-1}(z)$)
1131: the branches
1132: of $P^{-1}(z)$ in $U$
1133: which map points close
1134: to $P(a)$ (resp. $P(b)$)
1135: to points close to $a$ (resp. $b$).
1136: In particular, $f_a$ (resp. $f_b$)
1137: equals the multiplicity of the point $a$ (resp. $b$)
1138: with respect to $P(z).$ It was shown in \cite{pa2} for $P(a)=P(b)$
1139: and in \cite{pry} in general case that condition \e{1} implies
1140: equality \e{e1} or system \e{e2}, where
1141: as above $Q(z)=\int q(z) \d z$ is normalized
1142: by the condition ${Q(a)=Q(b)=0}.$
1143:
1144: For the sake of self-containedness of this paper below
1145: we provide a short derivation of \e{e1}, \e{e2} from
1146: theorem \ref{t1}.
1147:
1148:
1149: \bp \l{brc} Suppose that condition \eqref{1} holds. Then, if $P(a)=P(b),$
1150: equation \e{e1}
1151: holds in $U$.
1152: Furthermore, if $P(a)\neq P(b),$ then system \e{e2} holds
1153: in $U.$
1154: \ep
1155:
1156: \pr Suppose first that $P(a)= P(b)=c_1.$ Examine the relation
1157: $$\phi_1(z)=\sum_{i=1}^nf_{1,i}Q(P^{-1}_{i}(z))=0.$$
1158: Let $i,$ $1\leq i \leq n,$ be an index such that $f_{1,i}\neq 0$
1159: and let $x$ be a vertex of the star $S_i$ such that $P(x)=c_1.$
1160: Observe that if $x\neq a,$ $x\neq b,$ then there exists an index
1161: $\tilde i$ such
1162: that $x$ is also a vertex of the star $S_{\tilde i}$ and
1163: $f_{1,\tilde i}=-f_{1,i}.$ Furthermore, we have $\tilde i=g_1^j(i)$
1164: for some natural number $j.$ Therefore, $\phi_1(z)$ has the form
1165: $$\phi_1(z)=-Q(P^{-1}_{i_a}(z))+ $$
1166: $$
1167: Q(P^{-1}_{i_1}(z))-Q(P^{-1}_{g_1^{j_1}(i_1)}(z))+
1168: \, ... \, +
1169: Q(P^{-1}_{i_r}(z))-Q(P^{-1}_{g_1^{j_r}(i_r)}(z))$$
1170: $$+Q(P^{-1}_{i_b}(z))=0,$$
1171: where $i_a$ (resp. $i_b$) is the index such that
1172: $a\subset S_{i_a}$
1173: (resp. $b\subset S_{i_b}$), $i_1,i_2,\, ... \, i_r$
1174: are some other indices and $j_1,j_2,\, ... \, j_r$
1175: are natural numbers.
1176:
1177: Let $n_1$ be the order of the element $g_1$ in the group $G_P.$
1178: For each $s,$ \linebreak ${0\leq s \leq n_1-1,}$ the equality
1179: $$-Q(P^{-1}_{g^s_1(i_a)}(z))+ $$
1180: $$
1181: Q(P^{-1}_{g^s_1(i_1)}(z))-Q(P^{-1}_{g_1^{j_1+s}(i_1)}(z))+
1182: \, ... \, +
1183: Q(P^{-1}_{g^s_1(i_r)}(z))-Q(P^{-1}_{g_1^{j_r+s}(i_r)}(z))$$
1184: $$+Q(P^{-1}_{g^s_1(i_b)}(z))=0$$ holds by the analytic continuation
1185: of the equality
1186: $\phi_1(z)=0.$ Summing these equalities and taking into account that
1187: for any $i,$ $1\leq i \leq n,$ and any natural number $j$ we have:
1188: $$\sum_{s=0}^{n_1-1}
1189: Q(P^{-1}_{g^s_1(i)}(z))=\sum_{s=0}^{n_1-1} Q(P^{-1}_{g_1^{j+s}(i)}(z)),$$
1190: we obtain equality \e{e1}.
1191:
1192: In the case when $P(a)\neq P(b)$ the proof is similar: if $P(a)=c_1,$
1193: $P(b)=c_2,$ then one must examine relations $\phi_1(z)=0$ and $\phi_2(z)=0.$
1194:
1195: \vskip 0.2cm
1196: Note that if points $a,b$ are not critical
1197: points of $P(z),$ then \eqref{e1} reduces to \eqref{eb}
1198: while \eqref{e2} leads to the equality $q(z)\equiv 0.$
1199: In view of lemmas \ref{lcomp} and \ref{rl} this implies
1200: immediately the following result from \cite{c}
1201: (see also \cite{pa2},\cite{pry}).
1202:
1203: \bc \l{cris}
1204: Let $P(z),q(z)\in \C[z]$, $q(z)\neq 0,$
1205: $a,b\in \C,$ $a\neq b.$ Suppose that
1206: $a,$ $b$ are not critical points of $P(z).$
1207: Then conditions \e{1} and \e{2} are
1208: equivalent. \ec
1209:
1210:
1211: \subsection{Relations between branches of $Q(P^{-1}(z))$}
1212: In this subsection we examine how linear relations between branches of \linebreak $Q(P^{-1}(z))$ over $\C$ reflect on the structure of coefficients of Puiseux
1213: expansions of $Q(P^{-1}(z))$ near
1214: infinity.
1215:
1216: Let $P(z)$ be a non-constant polynomial of degree $n$
1217: and
1218: let $z_0\in \C$ be a non-critical
1219: value of $P(z).$
1220: If
1221: $\vert z_0\vert $ is sufficiently large then in a neighborhood $U_{z_0}$
1222: of $z_0$ each
1223: branch of $P^{-1}(z)$ can be
1224: represented
1225: by a Puiseux series centered at infinity. More precisely,
1226: if $P_0^{-1}(z)$ is a fixed branch of $P^{-1}(z)$ near $z_0$ then
1227: in $U_{z_0}$ we have: $$P_0^{-1}(z)=
1228: \sum_{k=-1}^{\infty}v_kz^{-\frac{k}{n}}, \ \ \ \ \ \ \ \ v_k\in \C, \ \ \
1229: \ \ \ \ \ \varepsilon_n=exp(2\pi i/n),$$
1230: where $z^{\frac{1}{n}}$ is a branch of
1231: the algebraic function which is inverse to $z^n$ in $U_{z_0}.$
1232: If $l$ is a loop around infinity then
1233: the result of the analytic continuation of the branch
1234: $P^{-1}_0(z)$ along $l^j,$ $0\leq j \leq n-1,$ is represented
1235: by the series
1236: \be \l{cn}
1237: P^{-1}_j(z)=\sum_{k=-1}^{\infty}v_k\varepsilon_n^{jk}z^{-\frac{k}{n}}.
1238: \ee
1239: The numeration of branches of $P^{-1}(z)$ near $z_0$ defined by equation
1240: \eqref{cn}
1241: is called canonical. Clearly, such a numeration depends on
1242: the choice of $P^{-1}_0(z).$ Nevertheless, any canonical numeration
1243: induces
1244: the same cyclic ordering of branches of $P^{-1}(z)$ in $U_{z_0}.$
1245: This cyclic ordering also will be called canonical.
1246: For any non-zero polynomial $Q(z),$ $\deg Q(z)=m,$
1247: the composition $Q(P^{-1}_j(z)),$ $0\leq j \leq n-1,$
1248: is represented near $z_0$ by the series \e{ps2}
1249: obtained by the substitution of series \eqref{cn} in $Q(z).$
1250:
1251: Let $U$ be a simply-connected domain containing no critical values of
1252: $P(z)$ such that some linear combination of branches
1253: of $Q(P^{-1}(z))$ over $\C$ identically vanishes in $U$.
1254: Considering in case of necessity a bigger domain we can
1255: suppose without loss of generality that $\infty \in \partial U.$
1256: Then series \e{cn} converge in a domain
1257: $V\subset U.$
1258: Furthermore, we can assume that
1259: the numeration of branches of $P^{-1}(z)$ in $U$
1260: is induced by a canonical numeration
1261: of branches of $P^{-1}(z)$ in $V$.
1262: If equality
1263: \be \l{x}
1264: \sum_{i=0}^{n-1}f_jQ(P^{-1}_j(z))=0, \ \ \ \ \ \ f_j\in \C,
1265: \ee
1266: holds in $U,$ then substituting in \e{x} expansions \e{ps2}
1267: we see that \e{x} reduces to the system
1268: $$\sum_{j=0}^{n-1}f_ju_k\varepsilon^{kj}_n=0, \ \ \ \ \ \ \ k\geq -m.$$
1269: Introducing the notation $F(z)=\sum_{j=0}^{n-1}f_jz^j$
1270: and summing up we get:
1271:
1272: \bl \l{la} The equality \e{x}
1273: holds in $U$ if and only if
1274: for any $k\geq -m$
1275: either
1276: $u_k=0$
1277: or $F(\varepsilon^k_n)=0.$
1278: \el
1279:
1280: In particular, since all $u_k$ can not
1281: vanish
1282: and $\deg F(z)<n,$ the following statement is true.
1283:
1284: \bc \l{la+}
1285: If equality \e{x} holds in $U$
1286: %for non-constant polynomials $P(z),Q(z),$
1287: then $F(\varepsilon^r_n)=0$ for at least one $r,$ ${0\leq r \leq n-1.}$ On the other hand, for at least one $r,$ $0\leq r \leq n-1,$ the equality $u_k=0$ holds whenever $k\equiv r \ \mod n.$
1288: \ec
1289:
1290: \subsection{Lemma about monodromy groups of polynomials}
1291: In order to relate \e{e1}, \e{e2}
1292: with coefficients of Puiseux
1293: expansions of $Q(P^{-1}(z))$ near
1294: infinity
1295: we are going to examine
1296: which roots of unity can be roots of the corresponding polynomial
1297: \be \l{e11}
1298: r(z)=\frac{1}{d_{a}}\sum_{s=1}^{d_{a}}
1299: z^{a_s}-\frac{1}{d_{b}}\sum_{s=1}^{d_{b}}
1300: z^{b_s},
1301: \ee or common roots of the corresponding pair of polynomials
1302: \be \l{e22}
1303: r_1(z)=\frac{1}{d_{a}}\sum_{s=1}^{d_{a}}
1304: z^{a_s}, \ \ \ \ \ \ \ \ \ \ r_2(z)=\frac{1}{d_{b}}\sum_{s=1}^{d_{b}}
1305: z^{b_s}.
1306: \ee
1307: For this propose we establish now a
1308: geometric property of monodromy groups of polynomials
1309: which concerns the mutual arrangement
1310: of indices $a_1,$
1311: $a_2, ... , a_{f_a}$
1312: and $b_1,$ $b_2,$ ... ,
1313: $b_{f_b}$ under assumption that the numeration of branches is canonical.
1314:
1315: Let $P(z)\in \C[z],$ $\deg P(z)=n,$ $a,b\in \C,$ $a\neq b.$
1316: Let $U$
1317: be a simply-connected domain containing no critical values of $P(z)$
1318: such that $P(a),P(b), \infty \in \partial U.$
1319: Fix a canonical numeration of branches of $P^{-1}(z)$ in $U$
1320: and let $P_{u_1}^{-1}(z),$
1321: $P_{u_2}^{-1}(z), ... ,$ $P_{u_{f_a}}^{-1}(z)$
1322: (resp. $P_{v_1}^{-1}(z),$ $P_{v_2}^{-1}(z), ... ,
1323: P_{v_{f_b}}^{-1}(z)$) be
1324: the branches of $P^{-1}(z)$ in $U$ which map points close
1325: to $P(a)$ (resp. $P(b)$)
1326: to points close to the point $a$ (resp. $b$) numbered
1327: by means of this numeration.
1328: The lemma below describes the
1329: mutual position on the unit circle of the sets
1330: $V(a)=\{ \v_n^{a_1},\v_n^{a_2}, ... , \v^{a_{f_a}}_n \}$ and
1331: $V(b)=\{\v_n^{b_1},\v_n^{b_2}, ... ,\v_n^{b_{f_b}}\},$
1332: where $\varepsilon_n=exp(2\pi i/n).$
1333: Let us introduce the following
1334: definitions.
1335: Say that two sets of points $X,Y$ on the unit circle $S_1$ are
1336: {\it disjointed} if there exist $s_1, s_2 \in S_1$
1337: such that all points from $X$ are on
1338: the one of two connected components of $S_1\setminus \{s_1,
1339: s_2\}$
1340: while all points from $Y$ are on
1341: the other one. Say that $X,Y$ are
1342: {\it almost disjointed} if $X\cap Y$ consists of a single point $s_1$
1343: and there exists a point $s_2\in S_1$ such that
1344: all points from $X\setminus s_1$ are on
1345: the one of two connected components of $S_1\setminus \{s_1,
1346: s_2\}$ while all points from $Y\setminus s_1$ are on
1347: the other one.
1348:
1349: \vskip 0.2cm
1350:
1351:
1352: \noindent{\bf Monodromy Lemma.} {\it
1353: The sets $V(a)$ and $V(b)$ are almost disjointed. Furthermore,
1354: if $P(a)= P(b)$ then $V(a)$ and $V(b)$ are disjointed.
1355: }
1356:
1357:
1358: \vskip 0.2cm
1359:
1360: \noindent{\it Proof.}
1361: Consider first the case when $P(a)=P(b).$
1362: \begin{figure}
1363: \medskip
1364: \epsfxsize=9truecm
1365: \centerline{\epsffile{G1.eps}}
1366: \smallskip
1367: \centerline{Figure 3}
1368: \medskip
1369: \end{figure}
1370: Let $M\subset U$ be a simple curve connecting points $P(a)=P(b)$
1371: and $\infty.$
1372: Consider the preimage
1373: $P^{-1}\{M\}$ of $M$ under the map $P(z)\,:\,\C\P^1\rightarrow
1374: \C\P^1.$
1375: It is convenient to consider $P^{-1}\{M\}$ as a bicolored graph
1376: $\Omega$ embedded into the Riemann sphere:
1377: the black vertices of $\Omega$ are preimages of
1378: $P(a)=P(b),$ the
1379: unique white vertex is the preimage of $\infty,$ and
1380: the edges of $\Omega$ are preimages of $M$ (see Fig. 3).
1381: Since the multiplicity of the vertex $\infty$ equals $n$ and $\Omega$
1382: has $n$ edges, $\Omega$ is connected.
1383: The edges of $\Omega$ are identified with branches of $P^{-1}(z)$ in $U$
1384: as follows:
1385: to a branch $P^{-1}_k(z),$ $1\leq k \leq n,$
1386: corresponds the edge $e_k$
1387: such that $P^{-1}_k(z)$ maps $M\setminus \{P(a),\infty\}$ into $e_k.$
1388: For any vertex $v$ of $\Omega$ the orientation of $\C\P^1$
1389: induces a natural cyclic ordering on edges of $\Omega$ adjacent to
1390: $v.$ In particular, taking $v=\infty,$ we obtain a
1391: cyclic ordering on edges of $\Omega.$
1392: Clearly, this cyclic ordering coincides with that induced by
1393: the canonical cyclic ordering of branches of $P^{-1}(z)$ in $U.$
1394: Let $E_a=\{e_{a_1},e_{a_2},\, ... \,, e_{a_{f_a}}\}$
1395: (resp. $E_b=\{e_{b_1},e_{b_2},\, ... \,, e_{b_{f_b}}\}$) be the union
1396: of edges of
1397: $\Omega$ which are adjacent to the vertex $a$ (resp. $b$).
1398: Let $D$
1399: be the domain from the collection of domains
1400: $\C\P^{1}\setminus E_a$ which contains point $b$
1401: and let $e_s,e_t\in E_a$ be the edges which bound $D.$
1402: Clearly, all the edges from $E_a$ are contained in
1403: $\C\P^1\setminus D.$
1404: Therefore, the lemma is equivalent to the following statement:
1405: the domain $D$ contains $e_h\setminus {\infty}$
1406: for all $e_h\in E_b.$ But the last statement is a corollary
1407: of the Jordan theorem since an edge $e_h\in E_b$ can intersect $e_s$ or
1408: $e_t$ only at infinity.
1409:
1410:
1411: In the case when $P(a)\neq P(b)$ the proof is modified as follows.
1412: Divide the boundary of $U$ into three
1413: parts $M_1,M_2,M_3,$ where $M_1$ connects the point $\infty$ with the
1414: point
1415: $P(a),$ $M_2$ connects the point $\infty$ with the point
1416: $P(b),$ and $M_3$ connects the point $P(a)$ with the point
1417: $P(b).$
1418: Consider now $P^{-1}\{\partial U\}$ as a graph
1419: $\Omega$ embedded into the Riemann sphere. The vertices of $\Omega$
1420: are divided into three groups: the first one consists of vertices
1421: that are preimages of $\infty,$ the second one consists of vertices
1422: that are preimages of $P(b),$ and the third one consists of vertices
1423: that are preimages of $P(a).$ Similarly, the edges of $\Omega$
1424: also are divided into three groups: the first one consists of edges
1425: that are preimages of $M_1,$ the second one consists of edges
1426: that are preimages of $M_2,$ and the third one consists of edges
1427: that are preimages of $M_3.$ Finally, the faces of $\Omega$ are divided
1428: into two groups: the first one consists of faces
1429: that are preimages of $U$ and the second one consists of faces
1430: that are preimages of $\C\P^1\setminus \overline{U}$
1431: (see Fig. 4).
1432:
1433:
1434: \begin{figure}
1435: \medskip
1436: \epsfxsize=11truecm
1437: \centerline{\epsffile{G2.eps}}
1438: \smallskip
1439: \centerline{Figure 4}
1440: \medskip
1441: \end{figure}
1442:
1443:
1444: The faces from the first group are identified with branches of $P^{-1}(z)$
1445: in $U$
1446: as follows:
1447: %if $N_3$ is a numeration of branches of $P^{-1}(z)$ in $V$
1448: to a branch $P^{-1}_k(z),$ $1\leq k \leq n,$
1449: corresponds the face $f_k$
1450: such that $P^{-1}_k(z)$ maps bijectively $U$ on $f_k\setminus \partial f_k.$
1451: The edges from the corresponding groups which bound $f_k$ will be denoted
1452: by
1453: $e^1_k, e^2_k, e^3_k$ correspondingly.
1454: Note that in a counterclockwise direction around infinity the edge
1455: $e^1_k,$ $1\leq k \leq n,$ is
1456: followed by the edge $e^2_k.$
1457: The canonical cyclic ordering of
1458: branches of $P^{-1}(z)$ in $U$ induces a cyclic ordering of faces of the
1459: first group of $\Omega$ and this
1460: ordering coincides with that induced by the
1461: orientation of $\C\P^1.$
1462: Let $E_a^1=\{e^1_{a_1},e^1_{a_2},\, ... \,, e^1_{a_{f_a}}\}$
1463: (resp. $E_b^2=\{e^2_{b_1},e^2_{b_2},\, ... \,, e^2_{b_{f_b}}\}$)
1464: be the union of edges from the first (resp.
1465: the second) group
1466: $\Omega$ which are adjacent to the vertex $a$ (resp. $b$).
1467: Let $D$
1468: be the domain from the collection of domains
1469: $\C\P^{1}\setminus E_a^1$ which contains point $b.$
1470: %and let $e_s^1,e_t^1\in E_a$ be the edges which bound $D.$
1471: Once again the Jordan theorem implies that
1472: all the edges from $E_a^1$ are contained in
1473: $\C\P^1\setminus D$ while
1474: $D$ contains $e_h^2\setminus {\infty}$
1475: for all $e_h^2\in E_b^2.$ Taking into account that
1476: for any $k,$ $1\leq k \leq n,$
1477: the edge $e^1_k$ is
1478: followed by $e^2_k$ this fact implies
1479: that $V(a)$ and $V(b)$ are almost disjointed.
1480: Note that, in contrast to the case when
1481: $P(a)=P(b),$
1482: now the sets $V(a)$ and $V(b)$ can have a non-empty intersection
1483: consisting of a single element.
1484:
1485: \subsection{On coefficients of Puiseux expansion of $Q(P^{-1}(z))$}
1486:
1487: In this subsection we deduce from the monodromy lemma the following
1488: important
1489: property of
1490: the Puiseux expansions \e{ps2}
1491: for pairs $P(z),Q(z)$
1492: satisfying \e{e1}, \e{e2}.
1493:
1494:
1495: \bt \l{pu}
1496: Let $P(z),Q(z)\in \C[z]$, $\deg P(z)=n,$
1497: $a,b\in \C,$ $a\neq b.$
1498: Suppose that \e{e1} or \e{e2} holds.
1499: Then $u_k=0$ for any $k$ such that $\GCD(k,n)=1.$
1500:
1501: \et
1502:
1503: \pr
1504: Suppose first that $P(a)=P(b).$
1505: Then lemma \ref{la} implies that
1506: $u_k = 0$ whenever the number $\varepsilon_n^{k}$ is not
1507: a root of the polynomial \e{e11}.
1508: Let us show that if $\GCD(k,n)=1$
1509: then
1510: the equality $r(\varepsilon_n^{k})=0$ is impossible.
1511: Indeed, if $(k,n)=1,$
1512: then $\varepsilon_n^{k}$
1513: is a primitive $n$-th root of unity.
1514: Since
1515: the $n$-th cyclotomic polynomial $\Phi_n(z)$ is irreducible over $\Z$,
1516: the equality $r(\varepsilon_n^{k})=0$
1517: implies that $\Phi_n(z)$ divides $r(z)$ in the ring $\Z[z].$
1518: Therefore, the primitive $n$-th root of unity
1519: $\varepsilon_n=exp(2\pi i/n)$ also is
1520: a root of $r(z)$ and hence the equality
1521: $$\sum_{s=1}^{d_a}\varepsilon_n^{a_s}/d_a =
1522: \sum_{s=1}^{d_b}\varepsilon_n^{b_s}/d_b$$ holds.
1523: The last equality is equivalent to the statement
1524: that the mass centers of the sets $V(a)$ and $V(b)$
1525: coincide. But this contradicts to the monodromy lemma. Indeed,
1526: the mass center of a system of points in $\C$
1527: is inside of the convex envelope of this system and therefore
1528: the mass centers of disjointed sets must be distinct.
1529:
1530: If $P(a)\neq P(b)$ then, similarly, the inequality $u_k\neq 0$ for
1531: $\GCD(k,n)=1$ implies that
1532: $$\sum_{s=1}^{d_a}\varepsilon_n^{a_s}/d_a =0, \ \ \ \ \
1533: \sum_{s=1}^{d_b}\varepsilon_n^{b_s}/d_b=0.$$
1534: But this again contradicts the monodromy lemma. Indeed, the fact that
1535: the sets $V(a)$ and $V(b)$ are almost disjoint implies that at least
1536: one
1537: from these sets is contained in an open half plane bounded by a line
1538: passing through the origin and therefore the mass center of this set
1539: is distinct from zero.
1540:
1541:
1542:
1543: \bc
1544: Let $P(z),q(z)\in \C[z]$, $q(z)\neq 0,$ $\deg P(z)=n,$
1545: $\deg Q(z)=m,$
1546: $a,b\in \C,$ $a\neq b.$ Suppose that \e{1} holds.
1547: Then $\GCD(m,n)>1.$
1548: \ec
1549:
1550: \pr Since in expansions \e{cn}
1551: the coefficient $v_{-1}$ is distinct from zero,
1552: the coefficient $u_{-m}=v_{-1}^m$ in expansions \e{ps2}
1553: is distinct from zero.
1554: Since \e{1} implies \e{e1} or \e{e2}
1555: by proposition \ref{brc} it follows from
1556: theorem \ref{pu} that $\GCD(m,n)>1.$
1557:
1558: \vskip 0.2cm
1559: Notice that theorem \ref{pu} agrees with conjecture
1560: \e{cc}. Indeed, if
1561: \be \l{a}
1562: Q(z)=\tilde Q_1(W_1(z))+\tilde Q_1(W_1(z))+ \ ... \ + \tilde Q_r(W_r(z)),
1563: \ee
1564: where $W_1(z),W_2(z),...,W_r(z)$ are (non-trivial) right divisors of
1565: $P(z)$ in the composition algebra,
1566: $$P(z)=\tilde P_1(W_1(z))=\tilde P_2(W_2(z))=\ ... \ =
1567: \tilde P_r(W_r(z)),$$
1568: %for non-linear
1569: %$W_1(z),W_2(z),...,W_r(z)$$W_1(z),W_2(z),...,W_r(z)$
1570: then
1571: the expansion \e{ps2} has the form
1572: $$Q(P^{-1}(z))=\tilde Q_1(\tilde P_1^{-1}(z))+\tilde Q_2(\tilde
1573: P_2^{-1}(z))+\
1574: ... \ +\tilde Q_r(\tilde P_r^{-1}(z)).$$
1575: Since $\deg \tilde P_j(z) < n,$ $1\leq j \leq r,$ it follows
1576: easily that
1577: $u_k=0$ for any $k$ such that $\GCD(k,n)=1.$
1578: Conjecturally, vice versa, equalities
1579: $u_k=0$ for all $k$ with $\GCD(k,n)=1$ imply
1580: that $Q(z)$ has form \e{a} at least under some additional
1581: assumptions. We plan to discuss this topic in another paper.
1582:
1583:
1584:
1585:
1586:
1587: \section{Further description of definite polynomials}
1588:
1589:
1590: \subsection{Case when $a$ or $b$ is not a critical
1591: point of $P(z)$}
1592: As a first application of the Puiseux expansions technique
1593: we provide in this section the following generalization of
1594: corollary \ref{cris}.
1595:
1596: \bt \l{alo} Let $P(z),q(z)\in \C[z]$, $q(z)\neq 0,$
1597: $a,b\in \C,$ $a\neq b.$ Suppose that at least one from
1598: points $a$ and $b$ is not a critical point of the polynomial $P(z).$
1599: Then
1600: conditions \e{1} and \e{2} are equivalent.
1601: \et
1602: \pr Since the condition of the theorem is compositionally
1603: stable it follows from lemmas \ref{rl}, \ref{lcomp}
1604: that we only must show that equality \e{eb} holds.
1605: To be definite suppose that the point $a$ is not a critical point of
1606: $P(z).$
1607: By proposition \ref{brc} either the system
1608: \be \l{pk}
1609: Q(p_{a_1}^{-1}(z))=0, \ \ \ \ \ \ \ \ \sum_{s=1}^{d_{b}}
1610: Q(p_{b_s}^{-1}(z))=0
1611: \ee
1612: or the equality
1613: \be \l{pk2}
1614: Q(p_{a_1}^{-1}(z))=\frac{1}{d_{b}}\sum_{s=1}^{d_{b}}
1615: Q(p_{b_s}^{-1}(z))
1616: \ee
1617: holds. Nevertheless, since the first equation of system \e{pk}
1618: leads to the equality ${q(z)\equiv 0,}$ we must only consider
1619: equation \e{pk2}.
1620:
1621: Applying lemma \ref{la} we see
1622: that for any $k$ such that $u_k\neq 0$
1623: the equality
1624: $$d_b(\varepsilon_n^k)^{a_1}=\sum_{s=1}^{d_{b}} (\varepsilon_n^k)^{b_s} $$
1625: holds.
1626: The triangle inequality implies that
1627: this is possible only if
1628: $$(\varepsilon_n^k)^{a_1} =(\varepsilon_n^k)^{b_1}=(\varepsilon_n^k)^{b_2}=
1629: ... =(\varepsilon_n^k)^{b_{d_b}}.$$
1630: Therefore,
1631: $$Q(p_{a_1}^{-1}(z))=
1632: Q(p_{b_1}^{-1}(z))=Q(p_{b_2}^{-1}(z))= ... =
1633: Q(p_{b_{d_s}}^{-1}(z))
1634: .$$
1635:
1636:
1637:
1638:
1639: \subsection{Case when
1640: $\deg P(z)=p^n$}
1641: In this subsection we deduce from theorem \ref{pu} the solution
1642: of the polynomial moment problem in the case
1643: when $\deg P(z)=p^r$ for $p$ prime.
1644:
1645: \bt \l{prime} Let $P(z),q(z)\in \C[z],$ $q(z)\neq 0,$
1646: $a,b\in \C,$ $a\neq b.$
1647: Suppose that $\deg P(z)=p^r,$
1648: where $p$ is a prime number. Then conditions \e{1} and \e{2} are equivalent.
1649: \et
1650:
1651: \pr Again, since the condition of the theorem is compositionally
1652: stable, it is enough to show that \e{eb} holds. Consider
1653: expansion \e{ps2}. By theorem \ref{pu} the equality
1654: $u_k=0$ holds for any $k$ with $\GCD(k,p^r)=1.$
1655: Show that this fact implies the equality
1656: $$
1657: Q(P^{-1}_j(z))=Q(P^{-1}_{j+p^{r-1}}(z))
1658: $$
1659: for any $j,$ $0\leq j\leq n-1.$
1660: Indeed, we have:
1661: $$
1662: Q(P^{-1}_j(z))-Q(P^{-1}_{j+p^{r-1}}(z))
1663: =\sum_{k=-m}^{\infty}
1664: w_kz^{-\frac{k}{n}},
1665: $$
1666: where
1667: $$
1668: w_k=u_k(\varepsilon_{p^r}^{jk}-\varepsilon_{p^r}^{(j+p^{r-1})k}).
1669: $$
1670: If $\GCD(k,p^r)=1$ then $u_k=0$ and hence $w_k=0.$
1671: Otherwise, $k=p\tilde k$ for some $\tilde k\in \Z.$
1672: Therefore,
1673: $$\varepsilon_{p^r}^{(j+p^{r-1})k}=
1674: \varepsilon_{p^r}^{jk}\varepsilon_{p^r}^{p^{r}\tilde k}=
1675: \varepsilon_{p^r}^{jk}$$ and hence again
1676: $w_k=0.$
1677:
1678:
1679: \subsection{Case when $P(z)$ is indecomposable}
1680: %In this subsection we give, using the obtained results,
1681: Theorems \ref{t1}, \ref{prime} allow us to give
1682: a short proof of the theorem proved in \cite{pa2}, \cite{pa4}
1683: which describes solutions to \e{1} in case when $P(z)$
1684: is indecomposable that is
1685: can not be represented as a composition
1686: $P(z)=P_1(P_2(z))$ with non-linear polynomials $P_1(z),$
1687: $P_2(z).$
1688: \bt \l{nepr}
1689: Let $P(z),q(z)\in \C[z],$ $q(z)\neq 0,$
1690: $a,b\in \C,$ $a\neq b.$
1691: Suppose that $P(z)$ is indecomposable.
1692: Then conditions \e{1} and \e{2} are equivalent. In more details,
1693: $Q(z)$ is a polynomial in $P(z)$
1694: and $P(a)=P(b).$
1695: \et
1696:
1697: \pr
1698: Once again we only must prove that \e{eb} holds.
1699: Suppose the contrary that is that
1700: all $Q(P^{-1}_i(z)),$ $1\leq i \leq n,$ where $n=\deg P(z)$ are
1701: different; then the monodromy group $G$ of the algebraic
1702: function $Q(P^{-1}(z))$ obtained by the complete analytic continuation
1703: of $Q(P^{-1}_i(z)),$ $1\leq i \leq n,$
1704: coincides with that of $P^{-1}(z)$.
1705: Since $P(z)$ is indecomposable, $G$ is primitive
1706: by the Ritt theorem \cite{ri}.
1707: Since for the case when $n=\deg P(z)$ is a prime number the statement follows
1708: from theorem \ref{prime} we can suppose that $n$ is a composite number.
1709: By the Schur theorem (see e.g. \cite{wi}, Th. 25.3)
1710: a primitive permutation group of composite degree $n$
1711: which contains an $n$-cycle is doubly transitive.
1712:
1713: Recall now the following fact:
1714: roots
1715: $\alpha_i,$ $1\leq i \leq n,$ of an
1716: irreducible algebraic equation over a field $k$ of characteristic zero
1717: with doubly transitive Galois group can not satisfy any
1718: relation $$\sum_{i=1}^{n} c_i\alpha_i=0,\ \ \ \ \ \ c_i\in k,$$ except
1719: the case when $c_1=c_2= ... =c_n$
1720: (see \cite{gi}, Proposition 4, or, in the context of algebraic functions, \cite{pa2}, lemma 2).
1721: Since the monodromy group of an algebraic function coincides with
1722: the Galois group of the equation over $\C(z)$
1723: which defines this function,
1724: it follows that if all $Q(P^{-1}_i(z)),$ $1\leq i \leq n,$ are
1725: different, then equality \e{x} is possible only when
1726: \be \l{ko}
1727: f_1=f_2= ... =f_n.
1728: \ee
1729: On the other hand, for any non-trivial equation $\phi_s(z)=0$
1730: appeared in theorem \ref{t1}
1731: the equality \e{ko} is impossible by construction.
1732: This contradiction completes the proof.
1733:
1734:
1735: \section{Solution of the polynomial moment problem
1736: for polynomials of degree less than
1737: \protect
1738: $10$}
1739:
1740: In this section
1741: %using obtained results combined with some combinatorial
1742: %reasoning involving cacti
1743: we provide a complete solution of
1744: the polynomial moment problem
1745: for polynomials of degree less than
1746: $10.$
1747:
1748:
1749: For an extended cactus $\tilde \lambda_P$ and a path $\Gamma_{a,b}$ define
1750: {\it the
1751: skeleton}
1752: $\hat\Gamma_{a,b}$ of $\Gamma_{a,b}$
1753: as follows.
1754: Draw the path
1755: $\Gamma_{a,b}$ separately from the graph $\tilde\lambda_P$
1756: and erase all its white vertices.
1757: Number the edges of the obtained graph $\hat\Gamma_{a,b}$ so that
1758: the number of an edge $e_k$ coincides with
1759: the number of the star $S_k$ of $\tilde\lambda_P$
1760: for which $e_k\subset S_k.$
1761: The number of edges of
1762: $\hat\Gamma_{a,b}$ is called the length
1763: $l(\hat\Gamma_{a,b})$ of $\hat\Gamma_{a,b}.$
1764: For example, the skeleton
1765: $\hat\Gamma_{a,b}$ of the path $\Gamma_{a,b}$ from Fig. 1
1766: is shown on Fig. 5; here $l(\hat\Gamma_{a,b})=4.$
1767:
1768: \medskip
1769: \epsfxsize=8truecm
1770: \centerline{\epsffile{2.eps}}
1771: \smallskip
1772: \centerline{Figure 5}
1773: \medskip
1774:
1775:
1776: \bt
1777: Let $P(z),q(z)\in \C[z]$, $q(z)\neq 0,$
1778: $a,b\in \C,$ $a\neq b,$ satisfy \e{1}.
1779: Suppose that $\deg P(z)<10.$ Then
1780: either condition \e{2} holds or
1781: %$\deg P(z)=6$ and
1782: there exist linear
1783: functions $L_1(z),L_2(z)$ such that
1784: $$L_2(P(L_1(z)))=T_6(z), \ \ \ \ L_1^{-1}(a)=-\sqrt{3}/2,
1785: \ \ \ \ L_1^{-1}(b)=\sqrt{3}/2,$$
1786: and $$Q(L_1(z))=A(T_3(z))+B(T_2(z))$$
1787: for some $A(z),B(z)\in \C[z].$
1788: \et
1789:
1790:
1791:
1792: \pr First of all observe that any natural number $n<10$ distinct
1793: from 6
1794: is either a prime number or a degree of a prime number.
1795: Therefore, it follows from theorem \ref{prime}
1796: that
1797: it suffices to consider the case
1798: when $\deg P(z) = 6.$ Furthermore, in view of theorem \ref{alo}
1799: we can suppose that the points $a,b$ are critical points of $P(z).$
1800: Finally notice that in order to prove that condition \e{2}
1801: holds for $P(z),$ $q(z)$ satisfying \e{1} with $\deg P(z)=6$
1802: it is enough to establish equality \e{comp}.
1803: Indeed, if $W(a)\neq W(b)$ in \e{comp} then performing the change of
1804: variable $z\rightarrow W(z)$ we see
1805: that \e{1} holds
1806: for $\tilde P(z), \tilde Q(z), W(a), W(b).$
1807: If $\deg W(z)$ equals 3 or 2,
1808: then it follows from theorem
1809: \ref{nepr} that ${\tilde Q(z)=R(\tilde P(z))}$ for some $R(z)\in \C[z]$ and
1810: $\tilde P(W(a))=\tilde P(W(b)).$ Therefore, \e{2} holds
1811: with $W(z)=P(z), \tilde Q(z)=R(z).$ On the other hand, if
1812: $\deg W(z)=6$ in \e{comp} then necessary
1813: $W(a)= W(b)$ since otherwise $\tilde Q(z)$ would be orthogonal to
1814: all powers of $z$ on the segment $W(a),W(b).$
1815: In particular, in view of lemma \ref{lcomp}, we see that
1816: in order to prove that conditions \e{1} and \e{2} are equivalent
1817: it is enough to establish \e{eb}.
1818:
1819:
1820:
1821: Since $\deg P(z)= 6,$ clearly $l(\hat\Gamma_{a,b})\leq 6.$ Moreover,
1822: since the points $a,b$ are critical points of $P(z),$
1823: the valency of the corresponding vertices of $\tilde \lambda_P$ is
1824: at least 2, and, therefore, actually $l(\hat\Gamma_{a,b})\leq 4.$
1825: Consider all possible cases. First of all observe that
1826: the equality $l(\hat\Gamma_{a,b})=1$ is
1827: impossible. Indeed,
1828: in this case theorem \ref{t1} implies that
1829: $Q(P^{-1}_i(z))=0,$ where $i$ is
1830: the number of the unique edge of $\hat\Gamma_{a,b},$
1831: and therefore $q(z)\equiv 0.$
1832: Furthermore, if $l(\hat\Gamma_{a,b})=2$
1833: then,
1834: since adjacent vertices of $\hat\Gamma_{a,b}$ have different
1835: colors,
1836: $\hat\Gamma_{a,b}$ can be of
1837: one from the following two forms shown on Fig. 6.
1838: \vskip 0.2cm
1839: \medskip \epsfxsize=12truecm \centerline{\epsffile{3.eps}} \smallskip
1840: \centerline{Figure 6} \medskip \noindent
1841: In both cases for the middle
1842: vertex $y$ we have $w(y)=1.$
1843: Therefore by theorem \ref{rt} equality \e{comp} holds and
1844: hence conditions \e{1} and \e{2} are equivalent.
1845: Observe, however, that the first configuration shown on Fig. 6 is actually
1846: not realizable since \e{2} implies that $P(a)=P(b).$
1847:
1848: Consider now the case when $l(\hat\Gamma_{a,b})=3.$ It is not
1849: difficult to see that in this case either again $w(y)=1$ for some color
1850: $y$ or $\hat\Gamma_{a,b}$ has the form shown on Fig. 7.
1851:
1852: \medskip
1853: \epsfxsize=8truecm
1854: \centerline{\epsffile{4.eps}}
1855: \smallskip
1856: \centerline{Figure 7}
1857: \medskip
1858: \noindent
1859: Let us examine the last case.
1860: Since for the skeleton shown on Fig. 7
1861: we have $P(a)\neq P(b),$ it follows
1862: from proposition \ref{brc} that
1863: system \ref{e2} holds.
1864: Furthermore, the equality $\deg P(z)= 6$ implies that for at least one
1865: point $s,$ $s\in \{a,b\},$
1866: the following two conditions are satisfied: the multiplicity of the
1867: vertex $s$ of the graph $\tilde \lambda_P$
1868: equals 2 and the connectivity component of $\tilde \lambda_P\setminus s$
1869: which does not contain $\Gamma_{a,b}$ consists of a unique star.
1870: To be definite suppose that $s=a.$
1871: Then, in notation of \ref{cacti},
1872: the first condition implies that \be \l{gra}
1873: \sum_{s=1}^{d_{a}}
1874: Q(P_{a_s}^{-1}(z))=Q(P^{-1}_{i_1}(z))+Q(P^{-1}_{g_x(i_1)}(z))=0 \ee and
1875: the second one that
1876: $g_y(g_x(i_1))=g_x(i_1).$ Therefore, the analytic continuation of
1877: \eqref{gra} along the loop $l_y$ leads to the equality \be \l{gra2}
1878: Q(P^{-1}_{i_2}(z))+Q(P^{-1}_{g_x(i_1)}(z))=0. \ee Now equalities
1879: \eqref{gra},\eqref{gra2} imply that
1880: $Q(P^{-1}_{i_1}(z))=Q(P^{-1}_{i_2}(z))$
1881: %Since this equality implies \e{zae}
1882: and we conclude as above that the configuration
1883: shown on Fig. 7 is not realizable.
1884:
1885: Consider finally the case when $l(\hat\Gamma_{a,b})=4.$
1886: Since $\hat\Gamma_{a,b}$ has 5 vertices,
1887: either $w(y)=1$ for some color
1888: $y$ or $\hat\Gamma_{a,b}$ is two-colored. In the last case
1889: $\hat\Gamma_{a,b}$ has the form shown on Fig. 8
1890: \vskip 0.2cm
1891: \medskip
1892: \epsfxsize=8truecm
1893: \centerline{\epsffile{z.eps}}
1894: \smallskip
1895: \centerline{Figure 8}
1896: \medskip
1897: \noindent
1898: and the corresponding cactus $\tilde \lambda_P$ is a 6-chain
1899: (the cactus with 6 stars of the
1900: maximal diameter).
1901: Furthermore, since $\deg P(z)=6,$ it follows from
1902: the Riemann-Hurwitz formula that
1903: $$\sum_{z\in \C\P^1}({\rm mult}_zP-1)=10.$$ Since
1904: ${\rm mult}_{\infty}P-1=5$ and
1905: the combinatorics of $\tilde \lambda_P$ imply that
1906: $$
1907: \sum_{P(z)=c_x}({\rm mult}_zP-1)=3, \ \ \ \ \ \ \ \
1908: \sum_{P(z)=c_y}({\rm mult}_zP-1)=2,$$
1909: we conclude that $P(z)$ has only two finite critical values $c_x,c_y.$
1910:
1911: It follows from the Riemann existence theorem (see e.g. \cite{cac2})
1912: that a complex polynomial with given
1913: critical values
1914: %$c_x,c_y$
1915: is defined by its cactus up to a
1916: linear change of variable.
1917: On the other hand, it is easy to see using the
1918: formula $T_n(\cos \phi)=\cos n\phi$ that
1919: $T_n(z)$ has only two critical values $-1,1$ and that
1920: all critical points of $T_n(z)$ are simple,
1921: Therefore, the corresponding cactus is a chain.
1922: In particular, for $P(z)=T_6(z)$ the corresponding cactus realized
1923: as the preimage
1924: of the segment $[-1,1]$ (considered as a star connecting
1925: $0$ with points $1$ and $-1$) has the form shown on Fig. 9
1926: (white vertices are omitted).
1927:
1928: %\begin{figure}
1929: \vskip 0.5cm
1930: \medskip
1931: \epsfxsize=9.5truecm
1932: \centerline{\epsffile{zz.eps}}
1933: %\smallskip
1934: \centerline{Figure 9}
1935: \medskip
1936: \noindent
1937: %\end{figure}
1938: Therefore, if we choose linear functions
1939: $L_1(z), L_2(z)$ such that:
1940: $$L_1^{-1}(a)=-\sqrt{3}/2, \ \ \ L_1^{-1}(b)=\sqrt{3}/2,
1941: \ \ \
1942: %critical values of the polynomial
1943: %$L_2(P(L_1(z)))$ are $-1,1,$ and
1944: %$L_2(P(-\sqrt{3}/2))=L_2(P(\sqrt{3}/2))=-1,$
1945: L_2(c_x)=-1,\ \ \ L_2(c_y)=1,$$
1946: the polynomial $L_2(P(L_1(z)))$ will be equal $T_6(z).$
1947:
1948: Finally, the last assertion of the theorem follows from
1949: the main result of the paper \cite{pa3} where
1950: all solutions to \eqref{1} for $P(z)=T_n(z)$
1951: were described.
1952:
1953:
1954:
1955: \bibliographystyle{amsplain}
1956: \begin{thebibliography}{10}
1957:
1958:
1959: \bibitem {bby} M. Blinov, M. Briskin, Y. Yomdin, {\it Local center conditions
1960: for a polynomial Abel equation and cyclicity
1961: of its zero solution}, Journal d'Anal. Math, to appear.
1962:
1963: \bibitem {bfy1} M. Briskin, J.-P. Francoise, Y. Yomdin,
1964: \textit{Une approche au probleme du centre-foyer de Poincare,}
1965: C. R. Acad. Sci., Paris, Ser. I, Math. 326, No.11, 1295-1298 (1998).
1966:
1967:
1968: \bibitem {bfy2} M. Briskin, J.-P.
1969: Francoise, Y. Yomdin,
1970: \textit{Center conditions, compositions of polynomials and moments on
1971: algebraic curve}, Ergodic Theory Dyn. Syst. \textbf{19}, no 5,
1972: 1201-1220 (1999).
1973:
1974:
1975: \bibitem {bfy3} M. Briskin, J.-P. Francoise, Y. Yomdin,
1976: \textit{Center condition II: Parametric and model center problems}, Isr.
1977: J. Math. \textbf{118}, 61-82 (2000).
1978:
1979:
1980: \bibitem {bfy4} M. Briskin, J.-P. Francoise, Y. Yomdin,
1981: \textit{Center condition III: Parametric and model center problems}, Isr.
1982: J. Math. \textbf{118}, 83-108 (2000).
1983:
1984: \bibitem {bfy5} M. Briskin, J.-P. Francoise, Y. Yomdin,
1985: \textit{Generalized moments, center-focus conditions and compositions of
1986: polynomials,} in ``Operator theory, system theory and related topics",
1987: Oper. Theory Adv. Appl., \textbf{123}, 161--185 (2001).
1988:
1989:
1990: \bibitem {bry} M. Briskin, N. Roytvarf, Y. Yomdin,
1991: {\it Center conditions at infinity for Abel differential equation},
1992: preprint, 2003.
1993:
1994:
1995: \bibitem {by} M. Briskin, Y. Yomdin,
1996: {\it Tangential Hilbert problem for Abel equation},
1997: Mosc. Math. J., to appear.
1998:
1999: \bibitem {c} C. Christopher, \textit{Abel equations: composition
2000: conjectures and
2001: the model problem}, Bull. Lond. Math. Soc. \textbf{32}, No.3,
2002: 332-338 (2000).
2003:
2004:
2005: \bibitem {cac1} M. El Marraki, N. Hanusse, J. Zipperer, A. Zvonkin,
2006: \textit{Cacti, braids and complex polynomials}, Sim. Lothar. Combin.
2007: \textbf{37},
2008: Art. B37b, 36 pp, 1996.
2009:
2010:
2011: \bibitem {gi} K. Girstmair, {\it Linear dependence of zeros
2012: of
2013: polynomials and construction of primitive elements}, Manuscripta
2014: Math. 39 (1982), no. 1, 81--97.
2015:
2016: \bibitem{cac2} S. Lando, A. Zvonkin, \textit{Graphs on surfaces and their
2017: applications}, Encyclopaedia of Mathematical
2018: Sciences, \textbf{141}, Low-Dimensional Topology, II,
2019: Springer-Verlag, Berlin, 2004.
2020:
2021:
2022: \bibitem {mus} N.I. Muskhelishvili, {\it Singular integral equations.
2023: Boundary problems of function theory and their application to mathematical
2024: physics,} Dover Publications, Inc., New York, 1992. 447 pp.
2025:
2026: \bibitem {pa1} F. Pakovich, {\it A counterexample to the ``Composition
2027: Conjecture''},
2028: Proc. AMS, 130, no. 12 (2002), 3747-3749.
2029:
2030: \bibitem {pa2} F. Pakovich, {\it On the polynomial moment problem}, Math.
2031: Research Letters
2032: {\bf 10}, (2003), 401-410.
2033:
2034: \bibitem {pa3} F. Pakovich, {\it On polynomials orthogonal to all powers
2035: of a Chebyshev polynomial on a segment}, Isr. J. Math, to appear.
2036:
2037: \bibitem {pa4} F. Pakovich, {\it Polynomial moment problem},
2038: Addendum
2039: to the paper {\it Center Problem for Abel Equation, Compositions of
2040: Functions,
2041: and Moment Conditions} by Y. Yomdin, Mosc. Math. J.,
2042: \textbf{3} (2003), no. 3, 1167-1195.
2043:
2044: \bibitem {pry} F. Pakovich, N. Roytvarf and Y. Yomdin. {\it Cauchy type
2045: integrals
2046: of Algebraic functions,} Isr. J. Math, to appear.
2047:
2048:
2049: \bibitem {ri} J. Ritt, \textit{Prime and composite polynomials,} Trans.
2050: Amer. Math. Soc. \textbf{23}, no. 1, 51--66 (1922).
2051:
2052: \bibitem {ro} N. Roytvarf, \textit{Generalized moments, composition of
2053: polynomials and Bernstein classes}, in "Entire functions in modern
2054: analysis. B.Ya. Levin memorial volume", Isr. Math. Conf. Proc.
2055: \textbf{15}, 339-355 (2001).
2056:
2057:
2058: \bibitem {sch} A. Schinzel, \textit{Polynomials with special regard to
2059: reducibility}, Encyclopedia of Mathematics and Its Applications
2060: \textbf{77}, Cambridge University Press, 2000.
2061:
2062: \bibitem {wi} H. Wielandt, \textit{Finite permutation groups,}
2063: New York and London: Academic Press, 1964.
2064:
2065: \bibitem {y1} Y. Yomdin, \textit{Center Problem for Abel Equation,
2066: Compositions of Functions, and Moment Conditions},
2067: Mosc. Math. J., \textbf{3} (2003), no. 3, 1167-1195.
2068:
2069: \end{thebibliography}
2070:
2071:
2072:
2073:
2074:
2075:
2076: \end{document}
2077: