0710.4085/s.tex
1: \documentclass{amsart} 
2: \usepackage{amssymb}
3: \usepackage{amsmath}
4: \usepackage{amsfonts}
5: \usepackage{amsthm}
6: \usepackage{color}
7: \begin{document} 
8: \input epsf.sty
9: 
10: 
11: \newtheorem{lem}{Lemma}[section]
12: \newtheorem{prop}[lem]{Proposition}
13: \newtheorem{cor}[lem]{Corollary}
14: \newtheorem{thm}[lem]{Theorem}
15: 
16: 
17: 
18: \date{}
19: \subjclass{Primary 30E99; Secondary 34C99; 20C05; 12F10}
20: 
21: 
22: 
23: \keywords{Center-Focus problem, moments, Cauchy type integrals, S-rings}
24: 
25: \address{F. Pakovich, Department of Mathematics,
26: Ben Gurion University,
27: P.O.B. 653,
28: Beer Sheva 84105, Israel}
29: \email{pakovich@math.bgu.ac.il}
30: \address{M. Muzychuk, Department of Mathematics, Netanya Academic College, Kibbutz Galuyot St. 16, Netanya 42365, Israel}
31: \email{muzy@netanya.ac.il}
32: 
33: 
34: \thanks{The research of the first author was supported by the ISF grant no. 979/07}
35: 
36: 
37: 
38: \begin{abstract} In this paper we give a solution of the following ``polynomial moment problem'' which arose about ten years ago in connection with Poincar\'e's center-focus problem: for a given polynomial $P(z)$ to describe polynomials $q(z)$ orthogonal to all powers of $P(z)$ on a segment $[a,b]$. 
39: \end{abstract}
40: 
41: 
42: 
43:  
44: \title{Solution of the polynomial moment problem
45: }  
46: 
47: \author{F. Pakovich, M. Muzychuk}
48: 
49: \maketitle
50: 
51: \date{}
52: 
53: 
54: 
55: \newcommand{\new}[1]{{\emph {#1}}}
56: \newcommand{\aut}{{\sf Aut}}
57: \newcommand{\sym}{{\sf Sym}}
58: \newcommand{\orb}{{\sf Orb}_2}
59: \newcommand{\A}{\mathbb{A}}
60: \newcommand{\C}{\mathbb{C}}
61: \newcommand{\F}{\mathbb{F}}
62: \newcommand{\N}{\mathbb{N}}
63: \newcommand{\Q}{\mathbb{Q}}
64: \newcommand{\Z}{\mathbb{Z}}
65: 
66: \newcommand{\cA}{\mathcal{A}}
67: \newcommand{\cB}{\mathcal{B}}
68: \newcommand{\cR}{\mathcal{R}}
69: \newcommand{\cV}{\mathcal{V}}
70: \newcommand{\cS}{\mathcal{S}}
71: \newcommand{\cX}{\mathcal{X}}
72: \newcommand{\wh}[1]{\widehat{#1}}
73: \newcommand{\sig}{\sigma}
74: \newcommand{\lm}{\lambda}
75: \newcommand{\im}{{\sf Im}}
76: \newcommand{\Ker}{\mathrm{Ker}}
77: \newcommand{\wt}[1]{\widetilde {#1}}
78: \renewcommand{\a}{\tilde{a}}
79: \newcommand{\tR}{\tilde{R}}
80: \newcommand{\irr}{{\sf Irr}}
81: \newcommand{\tr}{\mathrm{Tr}}
82: \newcommand{\un}{\underline}
83: \newcommand{\fX}{\mathfrak X}
84: \newcommand{\Trace}[1]{\mathaccent23{#1}}
85: \newcommand{\New}[1]{{\sl #1}}
86: \newcommand{\und}[1]{\underline{#1}}
87: \newcommand{\sg}[1]{\langle #1\rangle}
88: \newcommand{\Bsets}[1]{{\sf Basic}(#1)}
89: %W\newcommand{\sig}{\sigma}
90: \newcommand{\eps}{\epsilon}
91: \newcommand{\misha}[1]{\textcolor{red}{#1}}
92: 
93: 
94: 
95: 
96: 
97: 
98: \def\be{\begin{equation}}
99: \def\ee{\end{equation}}
100: \def\bs{$\square$ \vskip 0.2cm}
101: \def\d{{\rm d}} 
102: \def\D{{\rm D}} 
103: \def\I{{\rm I}} 
104: \def\C{{\mathbb C}} 
105: \def\N{{\mathbb N}} 
106: \def\P{{\mathbb P}}
107: \def\Q{{\mathbb Q}}
108: \def\Z{{\mathbb Z}}
109: \def\R{{\mathbb R}} 
110: \def\ord{{\rm ord}}
111: \def\qed{$\ \ \Box$ \vskip 0.2cm}
112: 
113: \def\e{\eqref}
114: \def\phi{{\varphi}}
115: \def\v{{\varepsilon}} 
116: \def\deg{{\rm deg\,}} 
117: \def\Det{{\rm Det}}
118: \def\dim{{\rm dim\,}} 
119: \def\Ker{{\rm Ker\,}} 
120: \def\Gal{{\rm Gal\,}}
121: \def\St{{\rm St\,}} 
122: \def\exp{{\rm exp\,}} 
123: \def\cos{{\rm cos\,}}
124: \def\card{{\rm card\,}} 
125: \def\diag{{\rm diag\,}} 
126: \def\GCD{{\rm GCD }}
127: \def\LCM{{\rm LCM }}
128: \def\mod{{\rm mod\ }}
129: 
130: \def\bp{\begin{proposition}}
131: \def\ep{\end{proposition}}
132: \def\bt{\begin{theorem}}
133: \def\et{\end{theorem}}
134: \def\be{\begin{equation}}
135: \def\l{\label}
136: \def\ee{\end{equation}}
137: \def\bl{\begin{lemma}}
138: \def\el{\end{lemma}}
139: \def\bc{\begin{corollary}}
140: \def\ec{\end{corollary}}
141: \def\pr{\noindent{\it Proof. }}
142: \def\note{\noindent{\bf Note. }}
143: \def\bd{\begin{definition}}
144: \def\ed{\end{definition}}
145: 
146: \newtheorem{theorem}{Theorem}[section]
147: \newtheorem{lemma}[theorem]{Lemma}
148: \newtheorem{definition}[theorem]{Definition}
149: \newtheorem{corollary}[theorem]{Corollary}
150: \newtheorem{proposition}[theorem]{Proposition}
151: 
152: 
153: 
154: 
155: 
156: 
157: 
158: 
159: \section{Introduction}
160: 
161: 
162: 
163: 
164: In this paper we solve the following
165: ``polynomial moment problem'':
166: {\it for given $P(z)\in \C[z]$ and distinct $a,b\in \C$ 
167: to describe $q(z)\in \C[z]$ such that
168: \be \l{1}
169: \int^b_a P^i(z)q(z)\d z=0 
170: \ee
171: for all $i\geq 0$.} 
172: 
173: 
174: The polynomial moment problem was posed in the 
175: series of papers \cite{bfy1}-\cite{bfy4}
176: in connection with the
177: center problem for the Abel differential equation 
178: \be \l{ab}
179: \frac{\d y}{\d z}=p(z)y^2+q(z)y^3.
180: \ee
181: with polynomial
182: coefficients $p(z), q(z)$ in the complex domain.
183: For given $a,b\in \C$ the center problem for the Abel equation
184: is to find necessary and sufficient conditions on $p(z),q(z)$ which imply
185: the equality $y(b)=y(a)$ for any solution $y(z)$ of \eqref{ab} with $y(a)$ small enough.
186: This problem is closely related to the classical 
187: Center-Focus problem of Poincar\'e and
188: has been studied in many recent papers 
189: (see e.g. \cite{bby}-\cite{c}, \cite{y1}).
190: 
191: The center problem for the Abel equation is connected with the 
192: polynomial moment problem in several ways.
193: For example, it was shown in \cite{bfy3}
194: that for the parametric version  
195: $$
196: \frac{\d y}{\d z}=p(z)y^2+\varepsilon q(z)y^3
197: $$
198: of \eqref{ab} the ``infinitesimal'' center conditions with respect to $\varepsilon$ 
199: reduce to moment equations \eqref{1} with $P(z)=\int p(z) \d z.$
200: On the other hand, it was shown in \cite{bry} that ``at infinity'' (under an appropriate projectivization of the 
201: parameter space) the system of equations on the
202: coefficients of $q(z)$, describing the center set of \eqref{ab} for fixed $p(z)$,
203: also reduces to \eqref{1}.
204: Many other results concerning connections between 
205: the center problem and the polynomial moment problem can be found in \cite{bry}.
206: These results convince that a thorough 
207: description of solutions of system \eqref{1} is an important step 
208: in the understanding of the center problem for the Abel equation.
209: 
210: 
211: There exists a natural condition on $P(z)$ and $Q(z)=\int q(z)\d z$
212: which reduces equations \eqref{1}, \eqref{ab} to similar equations
213: with respect to polynomials of smaller degrees. Namely, suppose that
214: there exist polynomials $\tilde P(z),$ $\tilde Q(z),$ $W(z)$ with $\deg W(z)>1$ such that
215: \be \l{2}
216: P(z)=\tilde P(W(z)), \ \ \ \ \ 
217: \ Q(z)=\tilde Q(W(z)).
218: \ee
219: Then after the change of variable $w=W(z)$
220: equations \eqref{1} transform to the equations  
221: \be \l{zam}
222: \int^{W(b)}_{W(a)} \tilde P^i(w)\tilde Q^{\prime}(w)\d w=0, 
223: \ee
224: while equation \eqref{ab} transforms to the equation 
225: \be \l{ab1}
226: \frac{\d\tilde y}{\d w}=\tilde P^{\prime} (w)\tilde y^2+\tilde Q^{\prime}(w)\tilde y^3.
227: \ee 
228: 
229: 
230: 
231: 
232: Furthermore, if the polynomial $W(z)$ in \eqref{2} 
233: satisfies the equality 
234: \be \l{w} W(a)=W(b), 
235: \ee 
236: then the Cauchy theorem implies that the polynomial $\tilde Q^{\prime}(w)$ is a solution of system \eqref{zam} and hence the polynomial $q(z)=Q^{\prime}(z)$ is a solution of system \eqref{1}.
237: Similarly, since any solution $y(z)$ of equation \eqref{ab} is the pull-back 
238: \be \l{pb} y(z)=\tilde y(W(z))
239: \ee
240: of a solution $\tilde y(w)$ 
241: of equation \eqref{ab1}, if $W(z)$ satisfies \eqref{w} then equation
242: \eqref{ab} has a center. 
243: This justifies the following definition:
244: a center for equation
245: \eqref{ab} or a solution of system \eqref{1} is called
246: {\it reducible} if there exist polynomials 
247: $\tilde
248: P(z),$ $\tilde Q(z),$ $W(z)$ such that
249: conditions 
250: \eqref{2}, \eqref{w} hold.
251: The main conjecture concerning the center problem for the Abel equation
252: (``the composition conjecture for the Abel equation"), supported by the results obtained 
253: in the papers cited above, states 
254: that any center for the Abel equation is reducible 
255: (see \cite{bry}
256: and the bibliography there).
257: 
258: By analogy with the composition conjecture it was suggested 
259: (``the composition conjecture for the polynomial moment problem") 
260: that, under the additional assumption $P(a)=P(b),$
261: any solution of (1) is reducible.
262: This conjecture was shown to be 
263: true in many cases. For instance, 
264: if $a,b$ are not critical points of $P(z)$   
265: (\cite{c}), if $P(z)$ is indecomposable (\cite{pa2}), and in some other 
266: special cases (see e. g. \cite{bfy3}, \cite{pp}, \cite{pry}, \cite{ro}).
267: Nevertheless, 
268: in general the composition conjecture for the polynomial moment problem 
269: fails to be true. 
270: Namely, it was shown in \cite{pa1} that if $P(z)$ has several ``compositional right factors'' $W(z)$ such that $W(a)=W(b)$, then
271: it may happen that the sum of reducible solutions corresponding to these factors is a non-reducible solution.
272:  
273: It was conjectured in \cite{pa4} that actually {\it any} non-reducible solution of \eqref{1} is a sum of reducible ones.
274: Since compositional factors $W(z)$ of a polynomial $P(z)$ can be defined explicitly,
275: such a description of non-reducible solutions of \eqref{1}
276: would be very helpful, especially   
277: for 
278: applications to the 
279: Abel equation
280: (cf. \cite{bry}).
281: However, until now
282: this conjecture was verified only in a single special case 
283: (see \cite{pa3}). 
284: 
285: Meanwhile, another necessary and sufficient condition 
286: for a polynomial $q(z)$ to be a solution of \eqref{1} was constructed in \cite{pp}. 
287: Namely, it was shown in \cite{pp} that there exists a finite system of equations 
288: \be \l{su}
289: \sum_{i=1}^nf_{s,i}Q(P^{-1}_{i}(z))=0, \ \ \ \ \ \ f_{s,i} \in \Z,  \ \ \ \ \ \ 1\leq s\leq k, 
290: \ee 
291: where $Q(z)=\int q(z)\d z$ and 
292: $P^{-1}_{i}(z),$ 
293: $1\leq i \leq n,$ are branches of the algebraic function  
294: $P^{-1}(z)$, 
295: such that \eqref{1} holds if and only if \eqref{su} holds.
296: Moreover, this system was constructed 
297: explicitly with the use of a special planar tree $\lambda_P$ 
298: which represents the monodromy group $G_P$ of the algebraic function $P^{-1}(z)$ in a combinatorial way.
299: By construction, points $a,b$ are vertices of $\lambda_P$ and system 
300: \eqref{su} reflects the combinatorics of the path connecting $a$ and $b$ on $\lambda_P$.
301: 
302: 
303: 
304: A finite system of equations 
305: \eqref{su} 
306: is more convenient for a study than initial infinite system of equations \eqref{1}.
307: In particular, in many cases the analysis of \eqref{su} permits to conclude that for given $P(z),a,b$ any solution 
308: of \eqref{1} is reducible (see \cite{pp}).
309: In this paper we develop
310: necessary algebraic and analytic techniques which allow us to describe 
311: solutions of \eqref{su} in the general case
312: and to prove that any solution of \eqref{1} is a sum of reducible ones.
313: So, our main result is the following theorem.
314: 
315: \bt \l{os} A non-zero polynomial $q(z)$ is a solution of system (1) 
316: if and only if $Q(z)=\int q(z)\d z$ can be represented as a sum of
317: polynomials $Q_j(z)$ such that
318: \be \l{cc}
319: P(z)=\tilde P_j(W_j(z)), \ \ \
320: Q_j(z)=\tilde Q_j(W_j(z)), \ \ \ {\it and} \ \ \ W_j(a)=W_j(b)
321: \ee
322: for some polynomials $\tilde P_j(z), \tilde Q_j(z), W_j(z)$.
323: \et
324: 
325: 
326: Note that since conditions of the theorem 
327: impose no restrictions on
328: the values of $P(z)$ at the points $a,b$ the theorem implies in particular 
329: that non-zero solutions of \eqref{1} exist if 
330: and only if the equality
331: $P(a)=P(b)$ holds. Indeed, if $P(a)=P(b)$ then
332: for any $\tilde Q(z)\in \C[z]$ the polynomial $Q(z)=\tilde Q(P(z))$ is a solution of \eqref{1} since we can set
333: $W(z)=P(z)$ in \eqref{2}, \eqref{w}. On the other hand, if $Q(z)$ is a solution of \eqref{1} then equalities \eqref{cc} imply that $P(a)=P(b)$. 
334: 
335: The paper is organized as follows. 
336: In the second section we give a detailed 
337: account of definitions and previous results related to the polynomial moment 
338: problem.   
339: In particular, starting from system \eqref{su}, we introduce a linear subspace $M_{P,a,b}$ of 
340: $\mathbb Q^n$ 
341: generated by the vectors 
342: $$(f_{s,\sigma(1)}, f_{s,\sigma(2)},\, ...\,, f_{s,\sigma(n)}), \ \ \ \sigma\in G_P, \ \ \ 1\leq s \leq k,$$
343: and study its basic properties. 
344: 
345: It follows from the definition that $M_{P,a,b}$ is invariant 
346: with respect to the permutation matrix representation of 
347: the group 
348: $G_P$.
349: In the third section of the paper, written entirely in the framework of the group theory,
350: we describe a general structure of such subspaces.
351: More generally, 
352: we describe subspaces of $\mathbb Q^n$ 
353: invariant with respect to the 
354: permutation matrix representation of a permutation group $G$ 
355: of degree $n$, containing a cycle of length $n$. Roughly speaking, we show that the structure of invariant subspaces of
356: $\mathbb Q^n$ for such $G$ depends on imprimitivity systems of $G$ only. 
357: We believe that this result is new and interesting by itself. 
358: 
359: Finally, in the fourth section, using the description of $G_P$-invariant subspaces of $\mathbb Q^n$  
360: and results and techniques of \cite{pp}, we prove Theorem \ref{os}.
361: 
362: 
363: \vskip 0.2cm
364: \noindent{\bf Acknowledgments}. The authors would
365: like to thank C. Christopher, J. P. Fran\-\c{c}oise, L. Gavrilov, G. Jones, M. Klin, Y. Yomdin, and W. Zhao 
366: for discussions of different questions
367: related to the subject of this paper. The authors also are grateful to the anonymous 
368: referee for valuable comments and suggestions.
369: 
370: 
371: 
372: \section{Preliminaries
373: } In this section we collect basic definitions and results concerning the polynomial moment problem.
374: In order to make the paper self-contained we outline
375: proofs of main statements.
376: 
377: 
378: \subsection{Criterion for $\hat H(t)\equiv 0$}
379: For $P(z),Q(z)\in \C[z]$ and 
380: a path $\Gamma_{a,b}\subset \C$ connecting different points $a,b$ of $\C$  
381: let
382: $H(t)=H(P,Q,\Gamma_{a,b},t)$ be a function defined on $\C\P^1\setminus P(\Gamma_{a,b})$ by the integral
383: \be \l{cau}
384: H(t)= \int_{\Gamma_{a,b}} \frac{Q(z)P^{\prime}(z)dz}{P(z)-t}.
385: \ee
386: Notice that although integral \eqref{cau} depends on $\Gamma_{a,b}$ the Cauchy theorem implies that if $\tilde \Gamma_{a,b}\subset \C$ is another path connecting $a$ and $b$,
387: then for all $t$ close enough to infinity the equality $$H(P,Q,\tilde \Gamma_{a,b},t)=H(P,Q,\Gamma_{a,b},t)$$
388: holds. Therefore, the Taylor expansion of $H(t)$ at infinity and the corresponding germ $\hat H(t)$ do
389: not depend on the choice of $\Gamma_{a,b}.$
390: 
391: After the change of variable
392: $z\rightarrow P(z)$
393: integral \eqref{cau} transforms to the Cauchy type integral
394: \be \l{ci} H(t)=
395: \int_{\gamma}\frac{g(z)dz}
396: {z-t}\,,
397: \ee where $\gamma=P(\Gamma_{a,b})$ and
398: $g(z)$ is obtained by the analytic continuation
399: along $\gamma$ of a germ of the algebraic function $Q(P^{-1}(z)).$ 
400: Clearly, integral representation \eqref{ci}
401: defines an analytic function in each domain of the complement
402: of $\gamma$ in $\C\P^1$. Notice that for any choice of $\Gamma_{a,b}$
403: the function defined in the domain containing infinity is the analytic continuation of the germ $\hat H(t).$
404: 
405: 
406: 
407: \bl [\cite{pp}]\l{lem} Assume that $P(z),$ $q(z)\in \C[z]$ 
408: and $a,b\in \C,$ $a\neq b,$ satisfy 
409: \be \l{ux} \int_{\Gamma_{a,b}} P^i(z)q(z)\d z=0, \ \ \ i\geq 0, \ee
410: and let $Q(z)$ be a polynomial defined by the equalities \be \l{urur} Q(z)=\int q(z) \d z, \ \ \ Q(a)=0.\ee
411: Then for the germ $\hat H(t)$ defined near infinity by integral \eqref{cau}  
412: the equality $\hat H(t)\equiv 0$ holds.
413: \el
414: 
415: \pr Indeed, for all $i\geq 1$ by integration by parts we have:
416: \be \l{ururu} 
417: \int_{\Gamma_{a,b}} P^i(z)q(z)\d z=P^i(b)Q(b)-P^i(a)Q(a)
418: -i\int_{\Gamma_{a,b}} P^{i-1}(z)Q(z)P^{\prime}(z)\d z.\ee
419: Furthermore, $Q(a)=0$ implies $Q(b)=0$ in view of \eqref{ux} taken for $i=0$. Therefore, if \eqref{ux} holds then  
420: all the integrals appearing in the right part of \eqref{ururu} vanish. 
421: On the other hand, these integrals 
422: are coefficients of the Taylor expansion of $-\hat H(t)$ at infinity.
423: \qed
424: 
425: Lemma \ref{lem} shows that the polynomial moment problem reduces to the problem of 
426: finding conditions on $Q(z)$ under which the equality $\hat H(t)\equiv 0$ holds.
427: On the other hand, we will show below (Corollary \ref{zzaa}) that if $\hat H(t)\equiv 0$ holds for some polynomial $Q(z)$ then 
428: \eqref{ux} holds for $q(z)=Q'(z).$ 
429: A condition of a general character for $\hat H(t)$ to vanish was given in the paper \cite{pry}
430: in the context of the theory of Cauchy type integrals of algebraic functions.
431: Subsequently, in the paper \cite{pp} was proposed a construction
432: which permits to obtain conditions for
433: the vanishing of $\hat H(t)$ in a very explicit form.
434: Briefly, the idea of \cite{pp} is to choose the integration path $\Gamma_{a,b}$
435: in such a way that its image under the mapping $P(z)\,:\, \C\P^1\rightarrow \C\P^1$ does not divide the Riemann sphere.
436: 
437: The construction of the paper \cite{pp} uses
438: a special graph $\lambda_P,$
439: embedded into the Riemann sphere, defined as follows
440: (see \cite{pp}).
441: Let $S$ be
442: a ``star'' joining a non-critical value $c$ of a polynomial $P(z)$ of degree $n$ with all its {\it finite} critical values $c_1,c_2, ... ,c_k$ by non intersecting oriented arcs
443: $\gamma_1, \gamma_2, ... ,\gamma_k$. Define $\lambda_P$
444: as a preimage of $S$ under the map $P(z)\,:\, \C\P^1\rightarrow \C\P^1$ (see Fig. 1).
445: \begin{figure}[ht]
446: %\medskip
447: \epsfxsize=12truecm
448: \centerline{\epsffile{1.eps}}
449: \smallskip
450: \centerline{Figure 1}
451: %\medskip
452: \end{figure}
453: More precisely, define
454: vertices of $\lambda_P$ as
455: preimages
456: of the points $c$ and $c_s,$ $1\leq s \leq k,$
457: and edges of $\lambda_P$ as preimages of the arcs
458: $\gamma_s,$ $1 \leq s \leq k.$
459: Furthermore, for each $s,$ $1\leq s \leq k,$ mark vertices of $\lambda_P$ which are preimages of the point $c_s$ by the number $s.$ Finally, define a {\it star} of $\lambda_P$ as a subset of edges of $\lambda_P$ consisting of edges adjacent to some
460: non-marked vertex.
461: 
462: By construction, the restriction of $P(z)$ on $\C\P^1\setminus \lambda_P$ is a covering of the topological punctured disk $\C\P^1\setminus \{S\cup \infty\}$ and therefore $\C\P^1\setminus \lambda_P$ is a disjointed union of punctured disks (see e.g. \cite{fors}). Moreover, since the preimage of infinity under $P(z)$ consists of a unique point, $\C\P^1\setminus \lambda_P$ consists of a unique disk and hence the graph $\lambda_P$ is a tree.
463: 
464: Set $C=\{c_1,c_2,
465: ... ,c_k\}$ and let $U\subset \C$ be a simply connected domain
466: such that $S\setminus C\subset U$ and $U\cap C=\emptyset$. Then in $U$
467: there exist $n$ single-valued analytic 
468: branches of the algebraic function $P^{-1}(z)$ inverse to $P(z).$ We will denote these branches by $P^{-1}_i(z),$ $1\leq i \leq n.$
469: The stars
470: of $\lambda_P$ may be naturally identified with branches of $P^{-1}(z)$ in $U$ as follows:
471: to the branch $P^{-1}_i(z),$ $1\leq i \leq n,$
472: corresponds the star $S_i,$ $1\leq i \leq n,$
473: such that $P^{-1}_i(z)$ maps bijectively the interior of $S$ to the interior of $S_i.$
474: 
475: 
476: Under the analytic continuation along a closed curve  
477: the set $P^{-1}_i(z),$ $1\leq i \leq n,$ transforms 
478: to itself. This induces a homomorphism \be \l{homo} \pi_1(\C\P^1\setminus \{C\cup \infty\},c)\rightarrow S_n\ee whose 
479: image $G_P$ is called the monodromy
480: group of $P(z)$. 
481: Notice that if $\omega_{\infty}$ and $\omega_i$, $1\leq i \leq k,$ are loops around $\infty$ and
482: $c_i$, $1\leq i \leq k,$ respectively, such that $\omega_1\omega_2\dots \omega_k\omega_{\infty}=1$ in 
483: $\pi_1(\C\P^1\setminus  \{C\cup \infty\},c),$ then the elements $g_i$, $1\leq i \leq k,$ of $G_P$,
484: which are the images of $\omega_i$, $1\leq i \leq k,$ under homomorphism \eqref{homo},
485: generate $G_P$ and satisfy the equality $g_1g_2\dots g_kg_{\infty}=1,$ where $g_{\infty}$ is the element of $G_P$ which is the image of $\omega_{\infty}$.
486: 
487: Having in mind the identification of
488: the set of stars of $\lambda_P$ with the set of branches of $P^{-1}(z),$
489: the permutation $g_s,$ $1\leq s \leq k,$ can be
490: identified with a permutation $\hat g_s,$
491: $1\leq s \leq k,$
492: acting
493: on the set of stars of $\lambda_P$ in the following way: $\hat g_s$ sends the
494: star $S_i,$ $1\leq i \leq n,$ to the ``next'' star under a counterclockwise
495: rotation around the vertex of $S_i$ colored by the $s$th color. For example,
496: for the tree shown on Fig. 1 we have: $g_1=(1)(2)(37)(4)(5)(6)(8),$
497: $g_2=(1)(2)(3)(47)(56)(8),$ $g_3=(1238)(4)(57)(6).$ Notice that since $P(z)$ is a polynomial, the permutation
498: $g_{\infty}$ is a cycle of
499: length $n.$ We always will assume that the numeration of branches of $P^{-1}(z)$ in $U$ is chosen in such a way
500: that $g_{\infty}$ coincides with the cycle $(1\,2\,...\,n)$. Clearly, such a numeration is
501: defined uniquely up to a choice of $P_1^{-1}(z)$.
502: 
503: 
504: 
505: The tree constructed above is known under the name of ``constellation'' or ``cactus'' and is closely related to
506: what is called a ``dessin d'enfant" (see \cite{lz}
507: for further details and other versions of this construction).
508: Notice that the Riemann existence theorem implies that
509: a polynomial $P(z)$ is defined by $c_1,c_2, ... ,c_k$
510: and $\lambda_P$ up to a composition $P(z)\rightarrow P(\mu (z)),$ where $\mu(z)$ is a linear function.
511: 
512: It follows from the definition that the points $a$ and $b$ are vertices of $\lambda_P$ if and only if $P(a)$ and $P(b)$
513: are critical values of $P(z)$. For our purposes however it is more convenient
514: to define the tree $\lambda_P$ so that
515: the points $a,b$ always would be its vertices.
516: So, in the case when $P(a)$ or $P(b)$ (or both of them) is not a
517: critical value of $P(z)$ we modify the construction as follows. Define $c_1,c_2, ... ,c_{k}$ as the set of all finite critical values
518: of $P(z)$ {\it supplemented} by $P(a)$ or $P(b)$ (or by both of them), and set as above
519: $\lambda_P=P^{-1}\{S\},$ where $S$ is a star
520: connecting $c$ with $c_1,c_2, ... ,c_{k}$
521: (we suppose that $c$ is chosen
522: distinct from $P(a),P(b)$). Clearly, $\lambda_P$ is still a tree and
523: the points $a,b$ are vertices of $\lambda_P.$
524: 
525: Since
526: $\lambda_P$ is connected and has no cycles there exists a unique oriented path $\mu_{a,b}
527: \subset \lambda_P$
528: joining the point $a$ with the point $b$.
529: Furthermore, it follows from the definition of $\lambda_P$ that if we set
530: $\Gamma_{a,b}=\mu_{a,b}$ then after the change of variable $z\rightarrow P(z)$
531: integral \eqref{cau}
532: reduces to the sum of integrals
533: \be \l{f}
534: H(t)=\sum_{s=1}^{k}\int_{\gamma_s}\frac{\phi_s(z)}{z-t}\, \d z,
535: \ee
536: where each $\phi_s(z),$ $1\leq s \leq k,$ is a linear combination
537: of the functions $Q(P^{-1}_{i}(z)),$ $1\leq i \leq n,$ in $U$.
538: Namely, \be \l{piz}
539: \phi_s(z)=\sum_{i=1}^nf_{s,i}Q(P^{-1}_{i}(z)),
540: \ee
541: where $f_{s,i}\neq 0$ if and only if the path
542: $\mu_{a,b}$ goes through the star
543: $S_i$ across its $s$-vertex.
544: Furthermore, if when going along $\mu_{a,b}$
545: the $s$-vertex of $S_i$
546: is followed by the center of $S_i$ then $f_{s,i}=-1$ otherwise
547: $f_{s,i}=1$.
548: For example, for the graph $\lambda_P$ shown on Fig. 1 and
549: the path $\mu_{a,b}\subset \lambda_P$ pictured by the fat
550: line we have:
551: \begin{gather*}
552: \phi_1(z)=-Q(P^{-1}_{2}(z))+Q(P^{-1}_{3}(z))- Q(P^{-1}_{7}(z)),\\
553: \phi_2(z)=Q(P^{-1}_{7}(z))-Q(P^{-1}_{4}(z)),\\
554: \phi_3(z)=Q(P^{-1}_{2}(z))-Q(P^{-1}_{3}(z))+ Q(P^{-1}_{4}(z)).
555: \end{gather*}
556: 
557: 
558: Notice that the number $k$ in \eqref{f} coincides with the number of critical values $s$ of $P(z)$ such that
559: the path $\Gamma_{a,b}$ passes through at least one vertex
560: colored by the $s$-th color.
561: Note also that
562: equations \e{piz}
563: are linearly dependent. Indeed,
564: for each $i,$ $1\leq i \leq n,$ such that there exists an index $s,$
565: $1\leq s
566: \leq k,$ with $f_{s,i}\neq 0$ there exist exactly two such
567: indices $s_1,s_2$, and $c_{s_1,i}=-c_{s_2,i}.$ Therefore, the equality
568: $$\sum_{s=1}^{k}\phi_s(t)=0$$
569: holds in $U$.
570: 
571: 
572: 
573: \bt [\cite{pp}]\l{t1}
574: Let $P(z),Q(z)\in \C[z]$ and $a,b\in \C,$ $a\neq b.$
575: Then $\hat H(t)\equiv 0$ if and only if
576: $\phi_s(z)\equiv 0$ for any $s,$ $1\leq s \leq k.$
577: \et
578: 
579: \pr Formula \eqref{f} defines the analytic continuation of
580: $\hat H(t)$ to the domain
581: $\C\P^1\setminus S$. In particular, $\hat H(t)\equiv 0$
582: if and only if $H(t)\equiv 0$ in $\C\P^1\setminus S$.
583: On the other hand, by the well-known boundary property of
584: Cauchy type integrals (see e.g. \cite{mus}), for any $s,$ $1\leq s \leq k,$ and any interior point $z_0$ of $\gamma_s$
585: we have:
586: \be \l{cuu}2\pi \sqrt{-1} \,\phi_s(z_0)=
587: \lim_{t \to z_0}\!\!\!\!\,^+H(t)-\lim_{t \to z_0}
588: \!\!\!\!\,^-H(t),
589: \ee
590: where the limits are taken when $t$ approaches $z_0$ from 
591: the ``right''  (resp. ``left'') side of $\gamma_s$. Therefore, if $H(t)\equiv 0$ in $\C\P^1\setminus S$,
592: then the limits in \eqref{cuu} equal zero and hence
593: $\phi_s(z)\equiv 0$ for any $s,$ $1\leq s \leq k.$
594: 
595: Finally,
596: if \be \l{sss}\phi_s(z)\equiv 0, \ \ \ \ 1\leq s \leq k,\ee then
597: it follows directly from formula \e{f} that
598: $H(t)\equiv 0$. \qed
599: 
600: 
601: 
602: 
603: 
604: \subsection{Subspace $M_{P,a,b}$}
605: For any element $\sigma\in G_P$ the
606: equality $\phi_s(z)=0,$ $1\leq s \leq k,$ implies by the analytic continuation
607: the equality
608: $$
609: \sum_{i=1}^nf_{s,i}Q(P^{-1}_{\sigma(i)}(z))=0.
610: $$
611: Therefore, replacing $\sigma$ by $\sigma^{-1}$ we see that Theorem \ref{t1} implies that
612: $\hat H(t)\equiv 0$ if and only if
613: for any $\sigma\in G_P$ and $s,$ $1\leq s \leq k,$
614: the equality
615: $$
616: \sum_{i=1}^nf_{s,\sigma(i)}Q(P^{-1}_{i}(z))=0
617: $$ holds.
618: 
619: Denote by $M_{P,a,b}$ the subspace of $\mathbb Q^n$ generated by the vectors
620: $$(f_{s,\sigma(1)}, f_{s,\sigma(2)},\, ...\,, f_{s,\sigma(n)}), \ \ \ 1\leq s \leq k, \ \ \ \sigma\in G_P.$$
621: Abusing the notation we usually will not distinguish an element
622: of $M_{P,a,b}$ and the corresponding equation connecting branches
623: of $ Q(P^{-1}(z))$.
624: For example, instead of using
625: the notation
626: \be \l{gi} (0,0,\, ... \,, 1, \, ...\,, 0,0,\, ... \,, -1,\, ... \,,0,0) \ee
627: for an element of $M_{P,a,b}$ we simply will use the equality
628: \be \l{compp} Q(P^{-1}_{i_1}(z))=Q(P^{-1}_{i_2}(z)),\ee
629: for corresponding $i_1\neq i_2,$ $1\leq i_1,i_2\leq n.$
630: 
631: Equality \eqref{compp}
632: is the simplest example of the equality
633: $\phi_s(z)=0,$ $1\leq s \leq k,$ 
634: and is equivalent to the statement that $P(z)$ and $Q(z)$ have a non-trivial ``compositional right factor''
635: (cf. \cite{c}, \cite{ro}, \cite{pa2},
636: \cite{pry}, \cite{pp}).
637: 
638: \bl \l{lll} Let $P(z),Q(z)\in \C[z]$. Then
639: the equalities
640: \be \l{comp} P(z)=\tilde P(W(z)), \ \ \ \ \ \ Q(z)=\tilde Q(W(z))\ee
641: hold for some $\tilde P(z), \tilde Q(z), W(z) \in \C[z]$ with $\deg W(z)>1$ if and only if equality \eqref{compp} holds for some $i_1\neq i_2,$ $1\leq i_1,i_2\leq n.$
642: Furthermore, $Q(z)=\tilde Q(P(z))$ for some $\tilde Q(z)\in \C[z]$ if and only if 
643: all the functions
644: $Q(P^{-1}_{i}(z)),$ $1\leq i \leq n,$ are equal between themselves.
645: 
646: \el
647: \pr Let $d(Q(P^{-1}))$ be a number of {\it different} functions in the collection of functions $Q(P^{-1}_i(z)),$ $1\leq i \leq n.$
648: Since any algebraic relation over $\C$ between $Q(p^{-1}(z))$ and $z,$ where $p^{-1}(z)$ is a branch of the algebraic function
649: $P^{-1}(z)$ in $U$, supplies
650: an algebraic relation between $Q(z)$ and $P(z)$ and vice versa,
651: we have:
652: $$d(Q(P^{-1}))=[\C(Q,P):\C(P)]=[\C(z):\C(P)]/[\C(z):\C(Q,P)]=$$ $$=n/[\C(z):\C(Q,P)].$$
653: Therefore, \be \l{degrr} [\C(z):\C(Q,P)]=n/d(Q(P^{-1})).\ee
654: It follows now from the L\"{u}roth theorem that $d(Q(P^{-1}))<n$ if and only if   
655: \eqref{comp} holds for some {\it rational} functions $\tilde P(z),$ $\tilde Q(z),$ $W(z)$ with
656: $\deg W(z)>1.$ Furthermore, if $d(Q(P^{-1}))=1$ then 
657: \eqref{degrr} implies that $Q(z)=\tilde Q(P(z))$ for some $\tilde Q(z)\in \C(z).$
658: 
659: Observe now that, since $P(z),$ $Q(z)$ are polynomials, without loss of generality
660: we may assume that $\C(Q,P)=\C(W)$ for some {\it polynomial} $W(z).$
661: Indeed, since $P(z)$ is a polynomial the equality $P(z)=U(V(z)),$ where $U(z),$ $V(z)$ are rational functions,
662: implies that $U(z)$ has a unique pole and that the preimage
663: of this pole under $V(z)$ consists of infinity only. This implies that $V(z)=\mu(W(z))$
664: for some polynomial $W(z)$ and M\"obius transformation $\mu(z)$, and it is clear that the fields $\C(V(z))$ and $\C(W(z))$ coincide. Finally, if $W(z)$ is a polynomial then obviously $\tilde P(z),$ $\tilde Q(z)$ also are polynomials. \qed
665: 
666: 
667: 
668: Since \eqref{comp} implies that
669: $$
670: \int^b_a P^i(z)q(z)\d z=
671: \int^{W(b)}_{W(a)} \tilde P^i(W)\tilde Q^{\prime}(W)\d W,
672: $$
673: Lemma \ref{lll} shows that
674: if the subspace $M_{P,a,b}$ contains an element of the form \eqref{compp},
675: then any solution $q(z)$ of the polynomial moment problem for $P(z)$ is either reducible or the
676: ``pull-back'' $q(z)=\tilde q(W(z))W'(z)$ of a solution $\tilde q(z)$ of the polynomial
677: moment problem for a compositional left factor $\tilde P(z)$ of $P(z)$ and the points $\tilde a=W(a)$ and $\tilde b =W(b).$
678: 
679: If $a,b$ are not critical points of $P(z)$ then $M_{P,a,b}$ always contains elements of form \eqref{compp}. In general case however a more delicate conclusion is true.
680: Denote by
681: $P_{a_1}^{-1}(z),$
682: $P_{a_2}^{-1}(z), ... , P_{a_{d_a}}^{-1}(z)$
683: (resp. $P_{b_1}^{-1}(z),$ $P_{b_2}^{-1}(z),$ ... ,
684: $P_{b_{d_b}}^{-1}(z)$)
685: branches
686: of $P^{-1}(z)$ in $U$
687: which map points close
688: to $P(a)$ (resp. to $P(b)$)
689: to points close to $a$ (resp. $b$).
690: In particular, the number $d_a$ (resp. $d_b$)
691: equals the multiplicity of the point $a$ (resp. $b$)
692: with respect to $P(z)$. The proposition below was proved in \cite{pry} and by a different method in \cite{pp}.
693: Below we give a proof following \cite{pp}.
694: 
695: 
696: 
697: \bp [\cite{pry}, \cite{pp}] \l{brc} If $P(a)=P(b)$ then $M_{P,a,b}$
698: contains the element
699: \be \l{e1}
700: \frac{1}{d_{a}}\sum_{s=1}^{d_{a}}
701: Q(P_{a_s}^{-1}(z))=\frac{1}{d_{b}}\sum_{s=1}^{d_{b}}
702: Q(P_{b_s}^{-1}(z)).
703: \ee
704: On the other hand, if $P(a)\neq P(b)$ then $M_{P,a,b}$
705: contains the elements
706: \be \l{e2}
707: \frac{1}{d_{a}}\sum_{s=1}^{d_{a}}
708: Q(P_{a_s}^{-1}(z))=0, \ \ \ \ \ \ \ \ \ \ \frac{1}{d_{b}}\sum_{s=1}^{d_{b}}
709: Q(P_{b_s}^{-1}(z))=0.
710: \ee
711: \ep
712: 
713: \pr Suppose first that $P(a)=P(b).$ Without loss of generality assume that $P(a)= P(b)=c_1$ and consider the relation
714: $$\phi_1(z)=\sum_{i=1}^nf_{1,i}Q(P^{-1}_{i}(z))=0$$ corresponding to $c_1$.
715: Let $i,$ $1\leq i \leq n,$ be an index such that $f_{1,i}\neq 0$
716: and $x$ be a vertex of the star $S_i$ such that $P(x)=c_1.$
717: It follows from the definition of $\phi_i(z),$ $1\leq i \leq k,$ that if $x\neq a,b$ then there exists an index
718: $j$ such
719: that $x$ also is a vertex of the star $S_{j}$ and
720: $f_{1,j}=-f_{1,i}.$ Furthermore, we have $j=g_1^l(i)$
721: for some natural number $l$ (see Fig. 2).
722: \begin{figure}[ht]
723: %\medskip
724: \epsfxsize=5truecm
725: \centerline{\epsffile{2.eps}}
726: \smallskip
727: \centerline{Figure 2}
728: %\medskip
729: \end{figure}
730: Therefore, $\phi_1(z)$ has the form
731: $$\phi_1(z)=-Q(P^{-1}_{i_a}(z))+ $$
732: $$
733: Q(P^{-1}_{i_1}(z))-Q(P^{-1}_{g_1^{l_1}(i_1)}(z))+
734: \, ... \, +
735: Q(P^{-1}_{i_r}(z))-Q(P^{-1}_{g_1^{l_r}(i_r)}(z))$$
736: $$+Q(P^{-1}_{i_b}(z))=0,$$
737: where $i_a$ (resp. $i_b$) is an index such that
738: $a\subset S_{i_a}$
739: (resp. $b\subset S_{i_b}$), $i_1,i_2,\, ... \, i_r$
740: are some other indices, and $l_1,l_2,\, ... \, l_r$
741: are some natural numbers.
742: 
743: 
744: For each $s\geq 0$ the equality
745: $$-Q(P^{-1}_{g^s_1(i_a)}(z))+ $$
746: $$
747: Q(P^{-1}_{g^s_1(i_1)}(z))-Q(P^{-1}_{g_1^{l_1+s}(i_1)}(z))+
748: \, ... \, +
749: Q(P^{-1}_{g^s_1(i_r)}(z))-Q(P^{-1}_{g_1^{l_r+s}(i_r)}(z))$$
750: $$+Q(P^{-1}_{g^s_1(i_b)}(z))=0$$ holds by the analytic continuation
751: of the equality
752: $\phi_1(z)=0.$ Summing now these equalities from $s=0$ to $s=r-1,$ where $r$ is the order of the element $g_1$ in the group $G_P,$ and taking into account that
753: for any $i,$ $1\leq i \leq n,$ and any natural number $l$ we have:
754: $$\sum_{s=0}^{r-1}
755: Q(P^{-1}_{g^s_1(i)}(z))=\sum_{s=0}^{r-1} Q(P^{-1}_{g_1^{l+s}(i)}(z)),$$
756: we obtain equality \e{e1}.
757: 
758: In order to prove the proposition in the case when $P(a)\neq P(b)$ it is enough to
759: examine in a similar way the relations $\phi_1(z)=0$ and $\phi_2(z)=0,$ where $P(a)=c_1,$
760: $P(b)=c_2.$
761: \qed
762: 
763: \bc \l{zzaa} Let $P(z),Q(z)\in \C[z]$ and $a,b\in \C,$ $a\neq b.$
764: Then $\hat H(t)\equiv 0$ implies that \eqref{ux} hold for $q(z)=Q^{\prime}(z).$
765: \ec
766: \pr Indeed, if $P(a)=P(b)$ then 
767: equating the limits of both parts of equality \eqref{e1} as $z$ approaches to $P(a)=P(b)$ 
768: we see that $Q(a)=Q(b)$. On the other hand, if $P(a)\neq P(b)$ then it follows from equalities \eqref{e2} in a similar way that 
769: $Q(a)=Q(b)=0$.
770: In both case it follows from \eqref{ururu} that \eqref{ux} holds.  
771: \qed   
772: 
773: 
774: Recall that we assume that the numeration of branches $P_i^{-1}(z),$
775: $1\leq i \leq n,$ in $U$ is chosen in such a way that the permutation $g_{\infty}\subset G_P$ coincides with the cycle $(1\,2\,...\,n)$.
776: The proposition below describes the position of branches appearing in Proposition \ref{brc}
777: with respect to this numeration. More precisely, we describe the mutual position on the unit circle of the sets
778: $$V(a)=\{ \v_n^{a_1},\v_n^{a_2}, ... , \v^{a_{d_a}}_n \}\ \ \ \ \
779: {\rm and} \ \ \ \ \ V(b)=\{\v_n^{b_1},\v_n^{b_2}, ... ,\v_n^{b_{d_b}}\},$$
780: where $\varepsilon_n=exp(2\pi \sqrt{-1}/n)$.
781: 
782: Let us introduce the following
783: definitions.
784: Say that two sets of points $X,Y$ on the unit circle $S^1$ are
785: {\it disjointed} if there exist $s_1, s_2 \in S^1$
786: such that one of two connected components of $S^1\setminus \{s_1,
787: s_2\}$ contains all points from $X$ while the other connected component of $S^1\setminus \{s_1,
788: s_2\}$ contains all points from $Y.$
789: Say that $X,Y$ are
790: {\it almost disjointed} if $X\cap Y$ consists of a single point $s_1$
791: and there exists a point $s_2\in S^1$ such that
792: one of two connected components of $S^1\setminus \{s_1,
793: s_2\}$ contains all points from $X\setminus s_1$ while the other connected component of $S^1\setminus \{s_1,
794: s_2\}$ contains all points from $Y\setminus s_1.$
795: 
796: 
797: \bp [\cite{pp}] \l{mono}
798: The sets $V(a)$ and $V(b)$ are disjointed or almost disjointed.
799: Furthermore,
800: if $P(a)= P(b)$ then $V(a)$ and $V(b)$ are disjointed.
801: \ep
802: \pr Consider first the case when $P(a)=P(b)=c_1.$
803: Let $\hat U$ be a simply-connected domain, containing no critical values of $P(z)$,
804: such that $U\subset \hat U$ and $\infty\in \partial \hat U$. Any branch of $P^{-1}(z)$ in $U$ can be extended analytically to $\hat U$ and we will assume that the numeration of branches of $P^{-1}(z)$ in $\hat U$ is induced by the numeration of
805: branches of $P^{-1}(z)$ in $U.$
806: Furthermore,
807: let $M\subset \hat U$ be a simple curve connecting points $c_1$
808: and $\infty$ and $\Omega=P^{-1}\{M\}$ be the preimage
809: of $M$ under the map $P(z)\,:\,\C\P^1\rightarrow
810: \C\P^1.$
811: It is convenient to consider $\Omega$ as a bicolored graph
812: embedded into the Riemann sphere. Namely, we define
813: black vertices of $\Omega$ as preimages of
814: $c_1,$ a
815: unique white vertex of
816: $\Omega$ as the preimage of $\infty,$ and
817: edges of $\Omega$ as preimages of $M$ (see Fig. 3).
818: \begin{figure}[ht]
819: \medskip
820: \epsfxsize=10truecm
821: \centerline{\epsffile{G1.eps}}
822: \smallskip
823: \centerline{Figure 3}
824: \medskip
825: \end{figure}
826: The edges of $\Omega$ may be identified with branches of $P^{-1}(z)$ in $\hat U$
827: as follows:
828: to the branch $P^{-1}_i(z),$ $1\leq i \leq n,$
829: corresponds the edge $e_i$
830: such that $P^{-1}_i(z)$ maps bijectively the interior of $M$ to the interior of $e_i.$
831: In particular, the ordering of branches of $P^{-1}(z)$ in $\hat U$ induces the ordering of edges of $\Omega.$
832: Since the multiplicity of the vertex $\infty$ equals $n$ and $\Omega$
833: has $n$ edges, $\Omega$ is connected.
834: 
835: Let $E_a$
836: (resp. $E_b$)
837: be a union of edges of
838: $\Omega$ which are adjacent to the vertex $a$ (resp. $b$).
839: It follows from the bijectivity of branches of $P^{-1}(z)$ on the interior of $M$
840: that if $D$ is a domain from the collection of domains $\C\P^{1}\setminus E_a$ such that $b\in D$, then $D$ contains the
841: whole set $E_b\setminus \infty.$ Now the proposition follows from the observation that
842: the cyclic ordering of edges of $\Omega$, induced by
843: the cyclic ordering of branches of $P^{-1}(z)$ in $\hat U$, coincides
844: with the cyclic ordering of edges of $\Omega$, induced by the orientation of $\C\P^1$
845: in a neighborhood of infinity.
846: 
847: In the case when $P(a)\neq P(b)$ the proof is modified as follows.
848: Take two
849: simple curves $M_1,$ $M_2\subset \hat U$ connecting the point $\infty$ with the points $P(a)$ and $P(b)$ correspondingly
850: and consider the preimage $P^{-1}\{M_1\cup M_2\}$
851: as a graph
852: $\Omega$ embedded into the Riemann sphere. The vertices of $\Omega$
853: fall into three sets: the first one consists of a unique vertex
854: which is the preimage of $\infty,$ the second one consists of vertices
855: which are preimages of $P(a),$ and the third one consists of vertices
856: which are preimages of $P(b).$ Similarly, the edges of $\Omega$
857: fall into two sets: the first one consists of edges
858: which are preimages of $M_1$ and the second one consists of edges
859: which are preimages of $M_2$
860: (see Fig. 4).
861: \begin{figure}[ht]
862: \medskip
863: \epsfxsize=9truecm
864: \centerline{\epsffile{G2.eps}}
865: \smallskip
866: \centerline{Figure 4}
867: \medskip
868: \end{figure}
869: 
870: Each of two sets of edges of $\Omega$ may be identified with branches of $P^{-1}(z)$ in $\hat U$
871: as follows:
872: to the branch $P^{-1}_i(z),$ $1\leq i \leq n,$
873: corresponds the edge $e^1_i$ from the first set (resp. the edge $e^2_i$ from the second set)
874: such that $P^{-1}_i(z)$ maps bijectively the interior of $M_1$ (resp. of $M_2$) to the interior of $e^1_i$
875: (resp. of $e^2_i$).
876: The ordering of branches of $P^{-1}(z)$ in $\hat U$ induces the ordering of edges of $\Omega$
877: in each of two sets. Clearly, this
878: ordering coincides with the natural ordering induced by the
879: orientation of $\C\P^1.$ Furthermore, it is easy to see that
880: when going round infinity in the counterclockwise direction the edge
881: $e^1_i,$ $1\leq i \leq n,$ is
882: followed by the edge $e^2_i.$
883: 
884: Let $E_a^1$
885: (resp. $E_b^2$)
886: be a union of edges from the first (resp.
887: the second) set
888: $\Omega$ which are adjacent to the vertex $a$ (resp. $b$). The bijectivity of branches of $P^{-1}(z)$ on the interior of $M_1$ and $M_2$
889: implies that if $D$ is a domain from the collection of domains $\C\P^{1}\setminus E_a^1$ such that $b\in D$, then $D$ contains the
890: whole set $E_b^2\setminus \infty.$ 
891: Taking into account that
892: for any $k,$ $1\leq i \leq n,$
893: the edge $e^1_i$ is
894: followed by $e^2_i,$ this implies
895: that $V(a)$ and $V(b)$ are disjointed or almost disjointed.
896: \qed
897: 
898: 
899: 
900: \noindent{\bf Remark.} Since $Q(P^{-1}_{i}(z))$, $1\leq i \leq n,$ are branches of an algebraic function,
901: relations \eqref{sss} are examples of linear relations between roots of an algebraic equation over the field $\C(z).$ 
902: A general algebraic approach to such relations, 
903: over an arbitrary field, was developed in the papers \cite{gi1}, \cite{gi2}. In particular,
904: it follows from Theorem 1 of \cite{gi2}
905: that a necessary and sufficient condition for the existence of at least one solution $Q(z)$ of \eqref{sss},
906: such that the functions $Q(P^{-1}_{i}(z))$, $1\leq i \leq n,$ are distinct between themselves, is that the subspace $M_{P,a,b}$ does not contain elements of form \eqref{gi}. An equivalent form of this condition is that the subspace $M_{P,a,b}$ does not contain any of subspaces $V_d^{\perp},$ $d\in D(G_P),$ which are defined below.
907: Notice however that the method of \cite{gi2} does not provide any information about the
908: description or the actual finding of these solutions.
909: 
910: 
911: \section{\l{cycy} Permutation matrix representations of groups containing a full cycle}
912: \subsection{Invariant subspaces and the centralizer ring}
913: 
914: The construction of $M_{P,a,b}$ implies that $M_{P,a,b}$ is an invariant subspace
915: of $\Q^n$ with respect to the so called {\it permutation matrix representation}
916: of the group $G_P$ on $\Q^n.$
917: By definition, the
918: permutation matrix representation of a transitive permutation group $H\subseteq S_n$
919: on $\Q^n$ is a homomorphism $R_H:\, H\rightarrow GL_n(\Q)$
920: which associates to
921: $h\in H$ a permutation matrix $R_H(h)\in GL_n(\Q)$ the elements $r_{i,j},$ $1\leq i,j\leq n,$ of  
922: which satisfy $r_{i,j}=1$ if $i=j^h$ and $r_{i,j}=0$ otherwise. In other words,
923: $$\begin{pmatrix}x_1\\x_2\\ \vdots \\x_n\end{pmatrix}=R_H(h)
924: \begin{pmatrix}x_{1^h}\\ x_{2^h}\\ \vdots \\ x_{n^h}\end{pmatrix}.$$
925: Note that $\Q^n$ admits a $R_H$-invariant scalar product $(x, y):=\sum_{i=1}^n x_iy_i$.
926: 
927: 
928: The goal of this section is to provide a full description of the invariant subspaces of $\Q^{n}$
929: with respect to the permutation matrix representation of $G_P$. More general, we classify all invariant subspaces of $\Q^{n}$
930: with respect to the permutation matrix representation of an arbitrary  group
931: $G\subseteq S_n$ containing the cycle $(1\,2\,...\,n)$. {\it In the following $G$ will always denote such a group.}
932: 
933: 
934: Recall that a subset $B$ of $X=\{1,2,\dots ,n\}$
935: is called a {\it block} of a transitive permutation group $H\subseteq S_n$ if for each $h\in H$ the set
936: $B^h$ is either disjoint or is equal to $B$ (see e.g. \cite{wi}). For a block $B$ the set
937: ${\mathcal B}:=\{B^h\,|\,h\in H\}$ forms a partition of $X$ into a disjoint union
938: of blocks of equal cardinality which is called an {\it imprimitivity system} of $H$.
939: Each permutation group $H\subseteq S_n$ has two {\it trivial} imprimitivity systems:
940: one formed by singletons and another formed by the whole $X$. A permutation group
941: is called {\it primitive} if it has only trivial imprimitivity systems. Otherwise it
942: is called {\it imprimitive}.
943: 
944: 
945: For each $d\,|\,n$ we denote by $V_d$ the subspace of $\Q^{n}$
946: consisting of $d$-periodic vectors. The fact that the group $G$ contains the cycle $(1,...,n)$
947: implies easily the following statement.
948: 
949: \bl\label{p_1} Any imprimitivity system for $G$ coincides with the residue classes modulo $d$
950: for some $d\,|\,n$. Furthermore, for given $d$ such classes
951: form an imprimitivity system for $G$
952: if and only if the subspace $V_d$ is $G$-invariant.\qed
953: \el
954: 
955: Denote by $D(G)$ the set of all divisors of $n$ for which $V_d$ is $G$-invariant. Clearly, $1,n\in D(G)$. Notice that $D(G)$ is a
956: lattice with respect to the ope\-rations $\land, \lor,$ where
957: $d\land f := \gcd(d,f)$ and $d\lor f := {\rm lcm}(d,f)$. Indeed, for an element $x\in X$
958: the intersection of two blocks containing $x$ and corresponding to $d,f\in D(G)$ is a block which
959: corresponds to $d\lor f$. On the other hand,
960: the intersection of two invariant subspaces $V_d,V_f$ is
961: an invariant subspace which is equal to $V_{d\land f}.$
962: 
963: 
964: 
965: 
966: We say that $d\in D(G)$ {\it covers} $f\in D(G)$
967: if $f\,|\,d,$ $f<d,$ and there is no $x\in D(G)$ such that $f<x<d$ and $f\vert x,$ $x\vert d$.
968: Now we are ready to formulate the main result of this section.
969: 
970: \begin{thm}\label{p_main} Each $R_G$-irreducible subspace of $\Q^{n}$ has the form
971: $$U_d:=V_d\cap \left(V_{f_1}^\perp\cap ...\cap V_{f_\ell}^\perp\right),$$ where $d\in D(G)$ and
972: $f_1,...,f_\ell$ is a complete set of elements of $D(G)$ covered by $d$. The subspaces $U_d$ are mutually orthogonal and
973: every $R_G$-invariant subspace of $\Q^{n}$ is a direct sum of some $U_d$ as above.
974: \end{thm}
975: 
976: The proof of this theorem splits into 
977: several steps and is given below.
978: We start from recalling some basic 
979: facts of the representations theory
980: which we will use afterward (see e.g. \cite{kir}).
981: 
982: First, any representation $T_H:\, H\rightarrow GL_n(k)$ of a finite group $H$ over a field $k$ of characteristic not dividing
983: $\vert H\vert$ 
984: is completely reducible,
985: that is $k^n$ is a direct sum of $T_H$-invariant irreducible  subspaces (Maschke's theorem).
986: Furthermore, irreducible subspaces
987: of a completely reducible representation $T_H:\, H\rightarrow GL_n(k)$
988: are in one-to-one correspondence with minimal idempotents of the {\it centralizer ring} $V_{k}(T_H)$.
989: Recall that $V_{k}(T_H)$ consists of all matrices $A\in M_n(k)$ which commute
990: with every $T_H(h), h\in H.$ Furthermore, a matrix $E$ is called
991: an idempotent if $E\neq 0$ and $E^2 = E$.
992: Two idempotents $E,F$ are called {\it orthogonal} if $EF = FE = 0$. Finally, an idempotent
993: $E\in V_{k}(T_H)$ is called {\it minimal} if it can not be presented as
994: a sum of two orthogonal idempotents from $V_{k}(T_H)$. Under this notation the correspondence above is obtained as follows: to a minimal idempotent $E\in V_{k}(T_H)$ corresponds an irreducible subspace $V=\im\{E\}.$
995: 
996: 
997: 
998: In general, the decomposition of $k^n$
999: into a sum of $T_H$-invariant irreducible subspaces is not uniquely defined. Nevertheless,
1000: if \be \l{dec}
1001: k^n=V_1^{\oplus a_1}\oplus\dots\oplus V_r^{\oplus a_r}\ee is a decomposition such that
1002: $V_i,$ $1\leq i \leq r,$ are pairwise non-isomorphic 
1003: $T_H$-invariant irreducible subspaces of $k^n$,
1004: then the
1005: subspaces $V_i^{\oplus a_i},$ $1\leq i \leq r,$ are defined uniquely. They
1006: correspond to the
1007: minimal idempotents of the {\it center} $C(V_{k}(T_H))$ of 
1008: the centralizer ring $V_{k}(T_H).$ Furthermore,
1009: $V_{k}(T_H)$ is commutative if and only if
1010: $a_i = 1$ for all $i,$ $1\leq i\leq r$.
1011: Notice that if $V_{k}(T_H)$ is commutative and the space $k^n$ admits a $T_H$-invariant scalar product
1012: then all $T_H$-invariant irreducible subspaces of $k^n$ are mutually orthogonal. Indeed, for any representation $T_H:\, H\rightarrow GL_n(k)$, which admits an invariant scalar product, $k^n$ can be decomposed into a sum of 
1013: $T_H$-invariant irreducible subspaces
1014: \be \l{dec1}
1015: k^n=V_1\oplus\dots\oplus V_r\ee with mutually orthogonal $V_i$. On the other hand, if $V_{k}(T_H)$ is commutative then a decomposition of $T_H$
1016: into a sum of $T_H$-invariant irreducible subspaces is uniquely defined and therefore coincides with \eqref{dec1}.
1017: 
1018: For the permutation matrix representation $R_H:\, H\rightarrow GL_n(k)$ of a transitive permutation group $H\subseteq S_n$ instead of the notation $V_{k}(R_H)$ we will use simply the symbol $V_{k}(H).$
1019: Below we will show (Proposition \ref{corre}) that for any group $G$ as above the ring
1020: $V_{\Q}(G)$ is isomorphic to a subring of the group algebra of a cyclic group
1021: and hence is
1022: commutative. Therefore, the above remarks imply the following statement.
1023: 
1024: 
1025: 
1026: 
1027: \bp \label{idempotent}
1028: An $R_G$-invariant subspace $W\subset{\Q}^n$ is irreducible
1029: if and only if there exists a minimal idempotent
1030: $E\in V_{\Q}(G)$ such that $\im\{E\} = W$. $R_G$-invariant irreducible subspaces of $\Q^n$ are mutually orthogonal and every $R_G$-invariant subspace is a
1031: direct sum of some $W$ as above. \qed
1032: \ep
1033: 
1034: For each transitive permutation group $H\subseteq S_n$ we can construct
1035: some special basis of $V_\C(H)$ via orbits of the stabilizer
1036: $H_1$ of the point 1 as follows. To each orbit $\Delta$ of $H_1$ associate a matrix $A^{\Delta}$,
1037: where $A_{i,j}^{\Delta}=1$ if there exist $h\in H$, $\delta \in \Delta$ such that $1^h=j,$ $\delta^h=i,$ and
1038: $A_{i,j}^{\Delta}=0$ otherwise. In particular, for the first column
1039: of $A^{\Delta}$ the equality $A_{i,1}^{\Delta}=1$ holds if and only if $i\in \Delta.$
1040: It turns out that the matrices $A^{\Delta}$ form a basis of $V_\C(H)$ (\cite{wi}, Theorem 28.4). Furthermore,
1041: since by construction the matrices $A^{\Delta}$ are contained in
1042: $M_n(\Q)$ they form a basis of $V_\Q(H)$.
1043: We summarize the properties of $A^{\Delta}$ in the proposition below (see \cite{wi}, \S 28).
1044: 
1045: 
1046: 
1047: \bp \l{alg} The matrices $A^{\Delta}$ satisfy the following conditions:
1048: \begin{itemize}
1049: 
1050: \item[(1)] $A^{\Delta}$ form a basis of the algebra $V_\Q(H)$ as of a $\Q$-module;
1051: 
1052: \item[(2)] If $\Delta_1\neq \Delta_2$ then the ones of $A^{\Delta_1}$ and $A^{\Delta_2}$
1053: do not occur in the same place. On the other hand, $\sum_{\Delta}A^{\Delta}$ is a matrix all the entries of
1054: which are ones;
1055: 
1056: \item[(3)] For each orbit $\Delta$ there exists an orbit $\Gamma$ such that
1057: $(A^{\Delta })^T=A^{\Gamma}.$ \qed 
1058: 
1059: \end{itemize}
1060: 
1061: \ep
1062: Notice that the property (3) implies that for the first row of
1063: $A^{\Delta}$ the equality $A_{1,j}^{\Delta}=1$ holds if and only if $j\in \Gamma$.
1064: Furthermore, it is easy to see that the mapping $\Delta\rightarrow \Gamma$ defines an involution
1065: on the set of orbits of $H_1.$
1066: 
1067: \pagebreak
1068: 
1069: \subsection{Schur rings}
1070: 
1071: \subsubsection{Isomorphism between $S_{\Q}(G)$ and $V_\Q(G)$}
1072: In order to construct the minimal idempotents of $V_{\Q}(G)$ we will use
1073: so called {\it Schur rings} introduced by Schur in his classical paper \cite{schu} for the investigation of
1074: permutation groups
1075: containing a regular subgroup $C$.
1076: Since in this paper $C$ always will be a cyclic group, in the following we will restrict our attention to
1077: this case only (see \cite{wi} for the account of the Schur method in the general case).
1078: 
1079: The idea of the Schur approach can be described as follows.
1080: If $G\subseteq S_n$ contains the cycle $c:=(1\,2\, ...\,n)$
1081: then elements of the set $\left\{1,2,\dots,n\right\}$ can be identified with elements of the cyclic group $C$ generated
1082: by $c$ as follows: to the element $i$ corresponds the element of $C$ which transforms $1$ to $i$.
1083: Therefore, we can consider $G$ as a permutation group acting on its subgroup $C.$
1084: After such an identification we can ``multiply'' elements of the set $\left\{1,2,\dots,n\right\}$ and
1085: this multiplication agrees with the action of $G$ in the following sense: if $h, g \in C$ then $h^g=hg$.
1086: Furthermore, identifying any two subsets of $\left\{1,2,\dots,n\right\}$
1087: with the corresponding elements of the group algebra $\Q[C]$ we can define their ``pro\-duct''
1088: as the product of these elements in $\Q[C]$. The remarkable result of Schur is that under
1089: such a multiplication the orbits of the stabilizer $G_1$
1090: form a basis of some subalgebra
1091: of $\Q[C]$.
1092: To make this statement precise let us introduce the following definition.
1093: 
1094: For $T\subseteq C$
1095: denote by $T^{\left(-1\right)}$ the set of elements of $C$ inverse to the elements of $T$
1096: and by $\und{T}$ the formal sum $\sum_{h\in T} h$. The elements of $\Q[C]$ of the form $\und{T}$ for
1097: some $T\subseteq C$
1098: are called {\it simple quantities} (\cite{wi}).
1099: \begin{definition}\label{s-ring}
1100: A subalgebra $\cA$ of the group algebra $\Q [C]$ is called a
1101: {\it Schur ring} or an {\it S-ring} over $C$
1102: if it satisfies the following axioms:
1103: 
1104: \begin{itemize}
1105: 
1106: \item[(S1)] $\cA$ as a $\Q$-module has a basis consisting
1107:          of simple quantities
1108:         $\und{T_0},\dots,\und{T_d}$, where $T_0=\{e\}$,
1109: 
1110: \item[(S2)] $T_i\cap T_j =\emptyset$ for $i\neq j$ and
1111:  $\bigcup_{j=0}^d T_j=C$,
1112: 
1113: \item[(S3)] For each $i\in\{0,1,\dots,d\}$ there exists
1114: $i'\in\{0,1,\dots,d\}$ such that \linebreak
1115: $T_{i'} =T_i{}^{(-1)}$.
1116: 
1117: \end{itemize}
1118: \end{definition}
1119: 
1120: 
1121: It is easy to see that the basis $\und{T_0},\dots,\und{T_d}$ satisfying (S1) and (S2) is 
1122: unique. Such a basis is called the {\it standard basis} of $\cA$.
1123: The number
1124: $d+1$ is called the {\it rank} of $\cA$. The sets $T_i$, $0\leq
1125: i\leq d$, are called the {\it basic sets} of $\cA$. Finally, the
1126: notation $\cA=\langle  \und{T_0},\dots,\und{T_d} \rangle$ is
1127: used if $\cA$ is an S-ring over $C$ whose basic sets are
1128: $T_0,\dots,T_d$.
1129: We also write $\Bsets{\cA}$ for the set
1130: $\{T_0,\dots,T_d\}$.
1131: Notice that if $\tilde \cA$ is an $S$-ring which is a subring of
1132: $\cA$ then its basic sets are some unions of basic sets of $\cA$. There
1133: are two {\it trivial} S-rings, namely
1134: $\langle \und{e},\und{C\setminus\{e\}}\rangle$ and $\Q[C]$.
1135: 
1136: 
1137: 
1138: \bp \l{corre} To any group $G$ corresponds a Schur ring $S_{\Q}(G)$ the basic sets of which
1139: are the orbits of the stabilizer $G_1$. Moreover,
1140: $S_{\Q}(G)$ and $V_\Q(G)$ are isomorphic as $\Q$-algebras.
1141: 
1142: \ep
1143: 
1144: The Proposition \ref{corre} is a particular case of Theorem 28.8 in \cite{wi}. It
1145: implies in particular that in order to describe
1146: the minimal idempotents of $V_{\Q}(G)$ it is enough to describe the ones of $S_{\Q}(G)$. Since however
1147: for this purpose
1148: an explicit construction of the isomorphism between $S_{\Q}(G)$ and $V_\Q(G)$
1149: is needed, we give below a short proof of Proposition \ref{corre} which
1150: is based on Proposition \ref{alg}
1151: 
1152: \vskip 0.2cm
1153: \noindent{\it Proof of Proposition \ref{corre}.} First of all observe that
1154: since $G$ contains $c$
1155: each matrix $M\in V_\Q(G)$ is necessarily a {\it circulant} that is
1156: each row vector of $M$
1157: is cyclically shifted for one element to the right relative to the preceding row vector, in other words
1158: \be \l{cirrr} M_{i,j} = M_{1,j-i+1 \, \mod n}.\ee
1159: 
1160: Define now a mapping $\psi: V_\Q(G)\rightarrow \Q[C]$
1161: by the formula $$\psi(M):=\sum_{j=1}^{n} M_{1,j}c^{j-1}$$ and show
1162: that $\psi$ is an algebra monomorphism.
1163: Indeed, for any $M,N\in V_\Q(G)$ we have:
1164: $$
1165: \psi(MN) = \sum_{\ell=1}^{n} (MN)_{1,\ell}c^{\ell-1} = \sum_{\ell=1}^{n} \sum_{i=1}^{n}
1166: M_{1,i}N_{i,\ell} c^{\ell-1}=$$
1167: $$=
1168: \sum_{\ell=1}^{n} \sum_{i=1}^{n} M_{1,i}N_{1,\ell-i+1} c^{\ell-1}=
1169: \sum_{i=1}^{n} \sum_{j=1}^{n}M_{1,i}N_{1,j}c^{i+j-2} =$$ $$ =\left(\sum_{i=1}^{n} M_{1,i}c^{i-1}\right)
1170: \left(\sum_{j=1}^{n} N_{1,j}c^{j-1}\right) =
1171: \psi(M)\psi(N).
1172: $$
1173: Thus $\psi$ is an algebra homomorphism. Furthermore, $\psi$ is injective since
1174: any matrix $M\in V_\Q(G)$ is defined by its first row in view of \eqref{cirrr}.
1175: 
1176: Clearly, the image of $V_\Q(G)$ is a subalgebra $S_{\Q}(G)$ of $\Q[C]$. Furthermore, by construction, the basis of this subalgebra consists of the orbits of the stabilizer
1177: $G_1.$ The properties S1, S2 of $S_{\Q}(G)$ are obvious. Finally, since
1178: any matrix from $V_\Q(G)$ is a circulant, it follows from the third part of Proposition \ref{alg} that $\Delta^{(-1)}=\Gamma$. \qed
1179: 
1180: 
1181: For $d$ dividing
1182: $n$ denote by $C_d$ a unique subgroup of $C$ of order $d$. For a Schur ring
1183: $\cA$ denote by $D(\cA)$ a set consisting of all divisors of $n$ for which $\und{C_d}\in\cA $.
1184: 
1185: \bl\label{p_D}
1186: $d\in D(G)\iff n/d\in D(S_{\Q}(G))$.
1187: \el
1188: \pr Let $d\in D(G)$. Then $C_{n/d}$ under the identification of the set $\{1,2, \dots, n\}$ with $C$
1189: corresponds to
1190: the set $X=\{1,d+1,2d+1, \dots, n-d+1\}$ and therefore
1191: is a block of $G$ containing 1. This implies that $C_{n/d}$ is a union of some $G_1$-orbits,
1192: say $T_0,...,T_\ell$. Hence
1193: $\und{C_{n/d}}= \und{T_0}+\und{T_1}+\dots+\und{T_\ell}$
1194: and therefore $\und{C_{n/d}}\in S_{\Q}(G)$.
1195: 
1196: 
1197: 
1198: Let now $n/d\in D(S_{\Q}(G))$. Then $\psi^{-1}(\und{C_{n/d}})\in V_\Q(G)$.
1199: It follows from the definition of $\psi$ that $\psi^{-1}(\und{C_{n/d}})$
1200: is a circulant matrix $M$ such that $M_{1,i}=1$ if $i\in X$ and
1201: $0$ otherwise. Since
1202: $M \in V_\Q(G)$ the subspace ${\rm Im}(M)$
1203: is $G$-invariant. On the other hand, it is easy to see that
1204: $ {\rm Im}(M)= V_d$. Therefore, $d\in D(G)$ by Lemma \ref{p_1}. \qed
1205: 
1206: \subsubsection{Rational $S$-rings}
1207: The automorphism group
1208: of $C$ is isomorphic to the multiplicative group $\Z_n^*$.
1209: Namely, to the element $m\in\Z_n^*$ corresponds the automorphism $g\mapsto g^m,g\in C.$
1210: Extending this action onto $\Q[C]$
1211: by linearity we obtain an action of $\Z_n^*$ on the group algebra $\Q[C]$:
1212: $$\alpha=\sum_{g\in C} \alpha_g g\ \ \longrightarrow \ \alpha^{(m)}:=\sum_{g\in C} \alpha_g g^m.$$ An element $\alpha\in\Q[C]$ is called {\it rational} if $\alpha=\alpha^{(m)}$ for any $m \in \Z_n^*$.
1213: Note that the mappings $\alpha\mapsto\alpha^{(m)},$ $m \in \Z_n^*$,
1214: are automorphisms of $\Q[C]$. Moreover, these mappings
1215: are automorphisms of any S-ring $\cA$ over $C$
1216: (see \cite{wi}, Theorem 23.9). In particular,
1217: for each $m\in\Z_n^*$ and $T\subseteq C$ we have
1218: $$
1219: T\in\Bsets{\cA}\iff T^{(m)}\in\Bsets{\cA},
1220: $$
1221: where for a subset $T\subset C$ by $T^{(m)}$ is denoted the set of $m$-th powers of
1222: elements of $T$.
1223: 
1224: 
1225: Recall that the set of all irreducible complex representations of $C$ consists of $n$ one-dimensional representations
1226: (characters)
1227: $\chi_0,...,\chi_{n-1}$ where
1228: $$\chi_\ell(c^j):=e^{2\pi \sqrt{-1} \ell j/n}, \ \ 0\leq j,\ell\leq n-1.$$
1229: We will keep the same notation for the extensions of $\chi_0,...,\chi_{n-1}$ by linearity
1230: on $\Q[C]$.
1231: The rational elements of an $S$-ring $\cA$ admit the following characterization.
1232: 
1233: \bl\l{re}
1234: An element $\alpha\in\Q[C]$ is rational if and only if
1235: $\chi_l(\alpha)\in\Q$ for all $l,$ $0 \leq l \leq n-1.$
1236: \el
1237: 
1238: \pr For an element $\alpha=\sum_{j=1}^n h_j c^j$ of $\Q[C]$ the condition
1239: that $\chi_l(\alpha)\in\Q$ for all $l,$ $0 \leq l \leq n-1,$ is equivalent to the condition that
1240: $\chi_l(\alpha),$ $0\leq l \leq n-1,$
1241: is invariant with respect to the action of the Galois group $\Gamma$ of the extension
1242: $(\Q(e^{2\pi \sqrt{-1} /n}):\Q)$. The group $\Gamma$ is isomorphic to $\Z_n^*$.
1243: Namely, to the element $m\in \Z_n^*$
1244: corresponds the element $\sigma_m\in \Gamma$ which transforms $e^{2\pi \sqrt{-1} /n}$ to $e^{2\pi \sqrt{-1} m /n}$.
1245: We have:
1246: $$ \sigma_m(\chi_\ell(\alpha)) = \sigma_m(\chi_\ell(\sum_{j=1}^n h_j c^j))
1247: =\sigma_m(\sum_{j=1}^n h_j e^{2\pi \sqrt{-1} \ell j/n})
1248: =
1249: $$$$=\sum_{j=1}^n h_j e^{2\pi \sqrt{-1} m\ell j/n}=
1250: \chi_{\ell}(\sum_{j=1}^n h_j c^{mj}) = \chi_\ell(\alpha^{(m)}).$$
1251: Therefore, for $\ell,$ $0\leq \ell \leq n-1,$ and
1252: $m\in\Z_n^*$ the equality
1253: $\sigma_m(\chi_\ell(\alpha))=\chi_\ell(\alpha)$ is equivalent to the equality
1254: $\chi_\ell(\alpha^{(m)})=\chi_\ell(\alpha)$. Since
1255: for $\alpha, \beta \in \Q[C]$ the equality  $\chi_{\ell}(\alpha)=\chi_{\ell}(\beta)$ holds
1256: for all $\ell,$ $0\leq \ell \leq n-1,$
1257: if and only if $\alpha=\beta,$ we conclude that $\chi_\ell(\alpha)\in\Q$
1258: for all $\ell,$ $0 \leq \ell \leq n-1,$ if and only if $\alpha$ is rational.\qed
1259: 
1260: An $S$-ring $\cA$ is called {\it rational} if all its elements are rational.
1261: Clearly, $\cA$ is rational if and only if $T^{(m)}=T$ for all $T\in\Bsets{\cA}$ and $m\in \Z_n^*$.
1262: Any rational $S$-ring is
1263: a subring of some universal rational $S$-ring $W.$
1264: To construct $W$ observe that the orbits of the action of $\Z_n^*$ on
1265: $C$ are parametrized by the divisors of $n$ as follows: an orbit $O_m,$ $m\vert n,$ consists of all generators of
1266: the group $C_m$. It turns out that the vector space spanned by $\und{O_m},$ $m\vert n$,
1267: is a rational $S$-ring $W$ (\cite{schu}). Furthermore, any rational $S$-ring $\cA$ is
1268: a subring of $W.$ Indeed, since any element of the standard basis of a rational $S$-ring $\cA$ is
1269: invariant with respect to the action of $\Z_n^*,$ such an element is
1270: a union of some $O_m,$ $m\vert n.$ Therefore,
1271: $\cA$ is
1272: a subring of $W.$
1273: 
1274: Denote by $D_n$
1275: the lattice of all divisors of $n$ with respect to the
1276: operations $\land, \lor.$ The statement below describes the rational S-rings.
1277: 
1278: \bp \l{mu} (\cite{mu})\label{p_rational}
1279: An S-ring $\cA$ over $C$ is rational if and only if there exists a sublattice $D$ of $D_n$
1280: with $1,n\in D$ such that $\und{C}_d,\, d\in D$, is a basis
1281: of $\cA$. \qed
1282: \ep
1283: 
1284: Notice that the basis $\und{C_d},$ $d\in D$, is not a standard
1285: basis of $\cA$ in the sense of definition \ref{s-ring}.
1286: 
1287: To any $S$-ring $\cA$ one can associate a rational $S$-ring
1288: $\Trace{\cA}$, called the {\it rational closure} of $\cA$,
1289: which is constructed as follows.
1290: Introduce an equivalence relation on $\Bsets{\cA}$ setting
1291: $S\sim T$ if there exists
1292: $m\in\Z_n^*$ such that $S=T^{(m)}$. For $T\in\Bsets{\cA}$ set $$\Trace{T}:=\bigcup\{T^{(m)}\,|\,m\in\Z_n^*\}$$ and denote by $\Trace{\cA}$ the $\Q$-module spanned by $\und{\Trace{T}},$ $T\in\Bsets{\cA}$.
1293: 
1294: 
1295: \bp \l{mmp} (\cite{schu}) $\Trace{\cA}$ is an S-ring consisting of all rational elements of $\cA.$
1296: \ep
1297: 
1298: The Proposition \ref{mu} allows us to describe a rational closure of an arbitrary S-ring.
1299: 
1300: \bp \label{p_r_closure2} Let $\cA$ be an S-ring over $C$. Then $\und{C}_d\,,d\in D(\cA)$, is a basis
1301: of $\Trace{\cA}$.
1302: \ep
1303: \begin{proof} By Proposition ~\ref{p_rational} $\Trace{\cA}$
1304: is spanned by vectors $\und{C}_d,\, d\in D$,
1305: for a certain sublattice $D$ of $D_n$. It remains to prove that $D=D(\cA)$. The inclusion
1306: $D\subseteq D(\cA)$ follows from the following line
1307: $$
1308: d\in D \implies \und{C_d}\in\Trace{\cA}\subseteq \cA\implies \und{C_d}\in \cA\implies d\in D(\cA).
1309: $$
1310: 
1311: Conversely, pick an arbitrary $f\in D(\cA)$. Then $\und{C_f}\in\cA$. Furthermore, since $$\und{C_f}=\sum_{t\in D_f} \und{O_t}\ ,$$ the element
1312: $\und{C_f}$ is rational and therefore
1313: $\und{C_f}\in\Trace{\cA}$. This means that $\und{C_f}$ is
1314: a linear combination of $\und{C_d},\, d\in D$.
1315: Therefore, in order to prove that $\und{C_f} = \und{C_d}$ for suitable $d\in D$ it is enough to show that the simple quantities $\und{C_d},\, d\in D_n,$ are linearly independent.
1316: 
1317: In order to prove the last statement assume that \be \l{asd} \sum_d l_d\,\und{C_d}=0\ee and let $M$ be a maximal number $d$
1318: for which $l_d\neq 0$. Clearly, any element $u$ of $C$ which generates $C_M$ can not be an element of $C_d$ for $d<M.$ But then $u$ appears in the left part of equality \eqref{asd} only once with the coefficient $l_d\neq 0$. This is a contradiction and therefore
1319: $\und{C_d},\, d\in D_n,$ are linearly independent.
1320: \end{proof}
1321: 
1322: 
1323: \subsection{Proof of Theorem \ref{p_main}}
1324: Similarly to the definition given above for the elements of $D(G)$
1325: say that for an $S$-ring $\cA$ an element $d\in D(\cA)$ covers an element $f\in D(\cA)$
1326: if $f\,|\,d,$ $f<d,$ and there is no $x\in D(\cA)$ such that $f<x<d$ and $f\vert x,$ $x\vert d$. 
1327: 
1328: Set $$\sig_d:=\frac{1}{d}\und{C_d}, \ \ \ d\in D(\cA).$$ It follows from  
1329: \begin{equation}\label{eq_idem}
1330: \sig_f\sig_d = \sig_d\sig_f=\sig_{f\lor d}
1331: \end{equation} that $\sig_d, d\in D(\cA),$ are idempotents of the
1332: algebra $\cA$. Nevertheless, they are not pairwise orthogonal.
1333: 
1334: 
1335: \bp \label{p_idempotents} An element of an S-ring $\cA$ over $C$ is a
1336: minimal idempotent of $\cA$ if and only if it has the form
1337: \be \label{minimal_idem}
1338: \eps_d = \sig_d\prod_{i=1}^\ell(1-\sig_{f_i}),
1339: \ee
1340: where $d\in D(\cA)$ and $f_1,...,f_\ell$ is a complete set of elements of $D(\cA)$
1341: covering $d$.
1342: \ep
1343: \begin{proof}
1344: Let us show first that $\eps_d,d\in D(\cA),$ are pairwise orthogonal idempotents.
1345: Since each $\sig_d,$ $d\in D_n$, is an idempotent, we have:
1346: $$
1347: \eps_d^2 = \sig_d^2\prod_{i=1}^\ell(1-\sig_{f_i})^2 =\sig_d\prod_{i=1}^\ell(1-2\sig_{f_i}+\sig_{f_i}^2)=
1348: \sig_d\prod_{i=1}^\ell(1-\sig_{f_i}) = \eps_d.
1349: $$
1350: Therefore, in order to show that $\eps_d$ is an idempotent we only must check that $\eps_d\neq 0.$
1351: In view of \eqref{eq_idem},
1352: after opening the brackets in \eqref{minimal_idem} we obtain
1353: a linear combination of $\sigma_f$ in which $\sig_d$ appears
1354: with the coefficient one. Since $\sigma_d,$ $d\in D_n$, are linearly independent this implies that
1355: $\eps_d\neq 0.$
1356: 
1357: 
1358: Let us check now the orthogonality. Take two distinct $m,d\in D(\cA),$
1359: where it is assumed that $d<m,$ and
1360: consider the product $\eps_d\eps_m$. Let $f_1,...,f_\ell$ and $n_1,...,n_k$
1361: be complete sets of elements of $D(\cA)$ which cover $d$ and $m$ respectively.
1362: By \eqref{eq_idem} we have:
1363: $$
1364: \eps_d\eps_m = \sig_d\prod_{i=1}^\ell(1-\sig_{f_i})\cdot\sig_m \prod_{j=1}^k(1-\sig_{n_j})
1365: =\sig_d \sig_m\prod_{i=1,j=1}^{i=\ell,j=k}(1-\sig_{f_i})(1-\sig_{n_j}) =$$
1366: \be \l{lop} =\sig_{d\lor m}\prod_{i=1,j=1}^{i=\ell,j=k}(1-\sig_{f_i})(1-\sig_{n_j}) \ee
1367: Since $d\,|\,d\lor m$ and $d < d\lor m$, there exists an element $f_i\in D(\cA)$ which
1368: covers $d$ and divides $d\lor m$. For such an element
1369: $(1 -\sig_{f_i})\sig_{d\lor m} = 0$ and
1370: this implies the vanishing of the right-hand side of \eqref{lop}.
1371: 
1372: Since the idempotents $\eps_d,d\in D(\cA),$ are pairwise orthogonal
1373: they are linearly independent elements of $\cA$. Furthermore, since Proposition \ref{p_r_closure2} implies that 
1374: $\eps_d \in \Trace{\cA}$ for any $d\in D(\cA)$ and
1375: \be \l{dim} \dim(\Trace{\cA}) = |D(\cA)|,\ee the idempotents $\eps_d,d\in D(\cA),$ form a basis of $\Trace{\cA}$ which consists of
1376: pairwise orthogonal idempotents.
1377: This implies that any minimal idempotent $\eps$ of $\Trace{\cA}$ coincides with some $\eps_d,d\in D(\cA).$ Indeed, since $\eps_d,d\in D(\cA),$ form a basis of $\Trace{\cA}$ there exist numbers $a_d, d\in D(\cA)$, such that $\eps=\sum_{d\in D(\cA)} a_d\eps_d.$ 
1378: Furthermore, since $\eps$ is an idempotent, for any $d\in D(\cA)$ the coefficient
1379: $a_d$ equals either $1$ or $0.$ Therefore, if $\eps$ is minimal then $\eps=\eps_d$ for some $d\in D(\cA).$
1380: 
1381: 
1382: 
1383: Finally, observe that the sets of minimal idempotents of $\Trace{\cA}$ and $\cA$ coincide. Indeed,
1384: if $\eps$ is any idempotent of $\cA$ then $\eps^2 = \eps$ implies that $\chi_i(\eps)\in\{0,1\}$ for all $i,$ $0\leq i \leq n-1.$ Therefore, by Proposition \ref{mmp},
1385: $\eps\in \Trace{\cA}.$ Furthermore, if $\eps$ is minimal in $\cA$ then obviously it is also minimal
1386: in $\Trace{\cA}$. On the other hand, any minimal idempotent of  $\Trace{\cA}$ remains a minimal idempotent in $\cA$ since all idempotents of $\cA$ are contained in $\Trace{\cA}.$
1387: \end{proof}
1388: 
1389: \noindent{\it Proof of Theorem \ref{p_main}.}
1390: By Proposition \ref{idempotent} any $R_G$-irreducible invariant subspace $W$ of $\Q^n$ corresponds to
1391: a minimal idempotent $E\in V_\Q(G)$ such that ${\sf Im}\{E\}=W$.
1392: Furthermore,
1393: since $\psi$ is an isomorphism between $V_\Q(G)$ and $S_\Q(G)$, the element
1394: $\psi(E)$ is a minimal idempotent of $S_\Q(G)$ and therefore,
1395: by Proposition \ref{p_idempotents}, $\psi(E) = \eps_d$ for some $d\in D(S_\Q(G))$.
1396: Thus $W$ is $R_G$-irreducible invariant subspace of $\Q^n$ if and only if there exist
1397: $d\in D(S_\Q(G))$
1398: such that
1399: \be \l{plm}
1400: W = {\sf Im} \{\psi^{-1}(\eps_d)\}=
1401: {\sf Im}\left\{\psi^{-1}(\sig_d)\Pi_{i=1}^\ell(I-\psi^{-1}(\sig_{f_i}))\right\}.
1402: \ee
1403: 
1404: Observe now that if two idempotent matrices $A,$ $B$ commute
1405: then for the matrix $C=AB=BA$
1406: the equality
1407: $${\sf Im}\{C\}= {\sf Im}\{A\}\cap{\sf Im}\{B\}$$ holds. Indeed, it is clear
1408: that $${\sf Im}\{C\}\subseteq {\sf Im}\{A\}\cap{\sf Im}\{B\}.$$ On the other hand, if $z\in{\sf Im}\{A\}\cap{\sf Im}\{B\}$ then
1409: $z=Ax=By$ for some vectors $x,y$ and \be \l{idem}Az=A(Ax)=Ax=z, \ \ \ Bz=B(By)=By=z.\ee It follows that $Cz=A(Bz)=Az=z$ and hence $z\in {\sf Im}\{C\}$.
1410: Since Lemma \ref{corre} implies that $V_\Q(G)$ is commutative it follows now from \eqref{plm} that
1411: $$
1412: W={\sf Im}\left\{\psi^{-1}(\sig_d)\right\}\cap \left(\bigcap_{i=1}^\ell{\sf Im}\left\{(I-\psi^{-1}(\sig_{f_i}))\right\}\right).
1413: $$
1414: 
1415: It was observed in the proof of Lemma \ref{p_D} that
1416: ${\sf Im}(\psi^{-1}(\sig_d))=V_{n/d}$.
1417: Furthermore, since the image of any idempotent matrix consists of
1418: its invariant vectors we have
1419: ${\sf Im}\{I -\psi^{-1}(\sig_d)\}={\sf Ker}\{\psi^{-1}(\sig_d)\}$. On the other hand,
1420: since the matrix $\psi^{-1}(\sig_d)$ is symmetric,
1421: ${\sf Ker}\{\psi^{-1}(\sig_d)\}={\sf Im}\{\psi^{-1}(\sig_d)\}^\perp.$
1422: Therefore,
1423: $$
1424: W = V_{n/d}\cap V_{n/f_1}^\perp\cap ... \cap V_{n/f_\ell}^\perp.
1425: $$
1426: Finally, Lemma \ref{p_D} implies that $n/d\in D(G)$ and
1427: $n/f_1,...,n/f_\ell$ is a complete set of elements of $D(G)$ covered by $n/d$.
1428: Hence, $W= U_{n/d}.$
1429: 
1430: 
1431: \vskip 0.2cm
1432: \noindent{\bf Remark.} If $G$ does not contain a full cycle, then Theorem \ref{p_main} fails to be true.
1433: A simple example is provided by the group $S_5$ acting on two element subsets of
1434: $\{1,2,3,4,5\}$. One can verify that in this way we obtain a primitive permutation group $G$ on $10$ points which
1435: yields a permutation matrix representation $\rho_G$ of dimension $10$. However, the collection of $\rho_G$-invariant irreducible subspaces of $\Q^{10}$ is distinct from the collection  $U_1,U_{10}$ since  
1436: $U_{10}$ is a direct sum
1437: of two irreducible $\rho_G$-invariant subspaces of dimensions $4$ and $5$.
1438: 
1439: 
1440: Notice also that Theorem \ref{p_main}
1441: is not true for representations over $\C$.
1442: In order to see this it is enough to take as $G$ any cyclic group.
1443: 
1444: 
1445: \section{Description of $Q(z)$ satisfying $\phi_s(t)=0$}
1446: 
1447: 
1448: \subsection{Geometry of $M_{P,a,b}$}
1449: In notation of Section \ref{cycy} set $$W =V_{f_1}^\perp\cap ...\cap V_{f_\ell}^\perp,$$ where
1450: $f_1,...,f_\ell$ is the set of all elements of $D(G_P)$
1451: distinct from $n$.
1452: Notice that since $n\in D(G_P)$ covers any other element of $D(G_P)$,
1453: the subspace $W$ coincides with the subspace $U_n$ from
1454: Theorem \ref{p_main} and therefore is $G_P$-invariant irreducible subspace of $\mathbb Q^n$.
1455: 
1456: Theorem \ref{p_main} together with Proposition \ref{mono} imply the following
1457: important geometric property of $M_{P,a,b}.$
1458: 
1459: \bp \l{pp} The subspace $M_{P,a,b}$ contains the subspace $W.$
1460: \ep
1461: 
1462: \pr Indeed, since by construction $M_{P,a,b}$ is a $G_P$-invariant subspace of $\Q^n$, it follows from
1463: Theorem \ref{p_main} that either $M_{P,a,b}$ contains
1464: $W$ or is orthogonal to $W.$ In the last case $M_{P,a,b}$ also would be orthogonal
1465: to the complexification $W^{\C}$ of $W.$
1466: Therefore, in order to prove the proposition
1467: it is enough to find vectors
1468: $\vec w\in W^{\C}$ and $\vec v\in M_{P,a,b}$
1469: such that $(\vec v,\vec w)\neq 0.$
1470: 
1471: In order to find such $\vec w$ observe that the vectors
1472: $$\vec w_i=(1,\varepsilon_n^j, \varepsilon_n^ {2j}, \,...\,,
1473: \varepsilon_n^{(n-1)j}),$$ $1\leq  j \leq n,$
1474: where $\varepsilon_n=exp(2\pi \sqrt{-1}/n),$
1475: form an orthogonal basis of $\C^n.$ Furthermore, for $d\vert n$
1476: vectors $\vec w_j$ for which $(n/d)\,\vert \,j$
1477: form a basis of $V_d^{\C}.$ Therefore, the vector
1478: $\vec w_1$ is orthogonal to $V_{f}^{\C}$
1479: for any $f\in D(G_P),$ $f\neq n,$ and hence
1480: $\vec w_1\in W^{\C}.$ Set $\vec w={\vec w}_1.$
1481: 
1482: Consider now two cases. Suppose first that $P(a)= P(b)$ and show that in this case
1483: for the vector $\vec v\in M_{P,a,b}$ corresponding to equation \eqref{e1}
1484: the inequality $(\vec v,\vec w)\neq 0$ holds.
1485: Indeed, the equality $(\vec v,\vec w)= 0$ is equivalent to the equality
1486: $$\sum_{s=1}^{d_a}\varepsilon_n^{a_s}/d_a =
1487: \sum_{s=1}^{d_b}\varepsilon_n^{b_s}/d_b$$
1488: which in its turn is equivalent to the statement
1489: that the ``centers of mass'' of the sets $V(a)$ and $V(b)$
1490: coincide. But this contradicts to Proposition \ref{mono} since
1491: the center of mass of a system of points in $\C$
1492: is inside of the convex envelope of this system and therefore
1493: the centers of mass of disjointed sets must be distinct.
1494: 
1495: Similarly, if $P(a)\neq P(b)$ then $ (\vec v,\vec w)\neq 0$ for at least
1496: one of two vectors corresponding to equations \eqref{e2}.
1497: Indeed, otherwise
1498: $$\sum_{s=1}^{d_a}\varepsilon_n^{a_s}/d_a =0, \ \ \ \ \
1499: \sum_{s=1}^{d_b}\varepsilon_n^{b_s}/d_b=0$$ that
1500: contradicts again to Proposition \ref{mono} since the fact that
1501: the sets $V(a)$ and $V(b)$ are almost disjointed implies that at least
1502: one of these sets is contained in an open half plane bounded by a line
1503: passing through the origin and therefore has the center of mass
1504: distinct from zero.
1505: $\ \ \Box$
1506: 
1507: 
1508: \subsection{Puiseux expansions of $Q(P^{-1}(z))$}
1509: Let $\hat U\subset \C$ be a domain as in the proof of Proposition
1510: \ref{mono}. Then, taking into account
1511: our convention about the numeration of branches of
1512: $P^{-1}(z)$, at points of $\hat U$ close enough to infinity
1513: the function $Q(P^{-1}_i(z))$, $1\leq i \leq n,$
1514: is represented by the converging series
1515: \be \l{ps2}
1516: Q(P^{-1}_i(z))=\sum_{k=-m}^{\infty}
1517: s_k\varepsilon_n^{(i-1)k}z^{-\frac{k}{n}},
1518: \ee
1519: where $z^{\frac{1}{n}}$ denotes some fixed branch of the algebraic function
1520: inverse to $z^n$ in $\hat U.$
1521: Therefore, any relation of the form
1522: \be \l{x}
1523: \sum_{i=1}^{n}f_iQ(P^{-1}_i(z))=0, \ \ \ \ \ \ f_i\in \C,
1524: \ee
1525: is equivalent to the system
1526: \be \l{vot}
1527: \sum_{i=1}^{n}f_is_k\varepsilon^{k(i-1)}_n=0, \ \ \ k\geq -m.
1528: \ee In particular, in view of Theorem \ref{t1}, the equality $\hat H(t)\equiv 0$ implies that 
1529: for any $k\geq -m$ such that the coefficient $s_k$ of series \eqref{ps2} distinct from zero
1530: the vector $\vec w_k$ is orthogonal to $M_{P,a,b}$.
1531: This fact together with Proposition \ref{pp} imply the following
1532: statement (cf. \cite{pp}, Theorem 4.1).
1533: 
1534: 
1535: \bp \l{la} Let $Q(z)$ be a polynomial such that
1536: $\hat H(t)\equiv 0.$ Then for any $k\geq -m$ such that the coefficient $s_k$ of series \eqref{ps2} is distinct from zero
1537: there exists $f\in D(G_P),$ $f\neq n,$ such that
1538: $(n/f)\,\vert\, k.$
1539: \ep
1540: 
1541: \pr Indeed, if $s_k\neq 0$ then it follows from
1542: \eqref{vot} that the vector $\vec w_k$ is orthogonal to $M_{P,a,b}^{\C}$ and
1543: therefore by Proposition \ref{pp} is orthogonal to $W^{\C}$.
1544: Since the subspace $(W^{\C})^{\perp}$ is generated by the vectors $\vec w_j,$ $(n/f)\,\vert \,j,$ $f\in D(G_P),$ $f\neq n,$ this implies that $\vec w_k$ is a linear combination of these vectors
1545: and hence coincides with one of them since
1546: the vectors $\vec w_i,$ $1\leq i \leq n,$ are linearly independent. Therefore, 
1547: $(n/f)\,\vert \,k$ for some $f\in D(G_P),$ $f\neq n.$ \qed
1548: 
1549: 
1550: 
1551: For $f\in D(G_P)$, $f\neq n,$
1552: set
1553: $$\psi_{f}(z)=\sum_{\substack {k\geq -m \\ k\equiv 0\, \mod n/f}}s_{k} z^{-\frac{k}{n}},$$
1554: where $s_k,$ $k\geq -m,$ are coefficients of series \eqref{ps2}. Clearly, $\psi_{f}(z)$ is
1555: an analytic function in $\hat U.$
1556: 
1557: 
1558: \bl \l{lape} For any $f\in D(G_P),$ $f\neq n,$ there exists $S_f(z)\in\C[z]$ such that \be \l{polk} \psi_{f}(z)=S_f(P_1^{-1}(z)).\ee
1559: Furthermore, we have:
1560: \be \l{xru} P(z)=A_1(B_1(z)), \ \ \ S_f(z)=R_1(B_1(z))\ee for some $A_1(z),B_1(z),R_1(z)\in \C[z]$ with $\deg B_1(z)>1.$
1561: \el
1562: 
1563: \pr First, observe that since
1564: $$1+(\varepsilon_n^{k})^f+(\varepsilon_n^{k})^{2f}+\dots +(\varepsilon_n^k)^{n-f}$$
1565: equals $n/f$ if $n\vert(fk)$ and zero otherwise, it follows from \eqref{ps2} that
1566: the equality
1567: \be \l{ra}\left(\frac{n}{f}\right)\psi_{f}(z)=
1568: Q(P^{-1}_{1}(z))+
1569: Q(P^{-1}_{f+1}(z))+Q(P^{-1}_{2f+1}(z))
1570: + ... + Q(P^{-1}_{n-f+1}(z))
1571: \ee holds.
1572: 
1573: Let now $\Omega_P$ be a field generated by all branches of $P^{-1}(z)$ 
1574: considered as elements of some fixed algebraic closure of $\C(z)$. Recall that
1575: the Galois group of the extension $[\Omega_P:\C(z)]$
1576: is permutation equivalent to the group $G_P$ and  
1577: under the Galois correspondence
1578: to the stabilizer of $P^{-1}_1(z)$ in $G_P$ corresponds the invariant subfield $\C(P^{-1}_1(z))$ of $\Omega_P$.
1579: Since $f\in D(G_P),$
1580: the collection of branches appearing in the right part of equality \eqref{ra} is a block of an imprimitivity system of
1581: $G_P$ containing $P^{-1}_1(z)$. Therefore, equality \eqref{ra} implies that
1582: the function $\psi_{f}(z)\in \Omega_P$ is invariant with respect
1583: to the action of the stabilizer of $P^{-1}_1(z)$ in $G_P$ and hence
1584: is contained in the field $\C(P^{-1}_1(z))$.
1585: So, there exists a rational function $S_f(z)$ such that equality \eqref{polk} holds.
1586: Furthermore, since the analytic continuation of the right side of \eqref{ra}
1587: has no poles in $\C$ the function $S_f(z)$ is a polynomial.
1588: Finally, since branches appearing in the right part of equality \eqref{ra} form a block,
1589: it is easy to see that 
1590: $$S_f(P_1^{-1}(z))=S_f(P_{lf+1}^{-1}(z)), \ \ \ 1\leq l \leq n/f-1,$$ 
1591: and hence the last part of the lemma follows from Lemma \ref{lll}.
1592: \qed
1593: 
1594: 
1595: \subsection{Proof of Theorem 1.1} In view of Theorem \ref{t1} we essentially must 
1596: show that the conclusion of the theorem holds for any non zero polynomial $Q(z)$ such that 
1597: $\hat H(t)\equiv 0.$ So, abusing the notation, below we will mean by a solution of the polynomial moment problem such a polynomial $Q(z)$.
1598: The proof is by induction on the number $i(P)$ of
1599: imprimitivity systems of the group $G_P$.
1600: If $i(P)=2$, that is if $G_P$ has only trivial imprimitivity systems, then Proposition \ref{la} implies that for any non-zero coefficient $s_j,$ $j\geq m,$ of  \eqref{ps2} the number $k$ is a multiple of $n$.
1601: Therefore, all the functions $Q(P^{-1}_i(z)),$ $1\leq i \leq n,$ are equal between themselves and hence $Q(z)=R(P(z))$ for some polynomial $R(z)$ by 
1602: Lemma \ref{lll}. Furthermore, necessarily
1603: $P(a)=P(b)$. Indeed,
1604: otherwise after the change of variable $z=P(z)$ we would
1605: obtain that the polynomial $R(z)$ is orthogonal to all powers
1606: of $z$ on the segment  $[P(a),P(b)]$. However, for
1607: $$P(z)=z, \ \ Q(z)=R(z), \ \ a=P(a), \ \ b=P(b)$$
1608: any of relations \eqref{e2} reduces to the
1609: equality $R(z)\equiv 0$ in contradiction with the condition $Q(z)\not\equiv 0$
1610: (of coarse instead of Proposition 2.1 we also
1611: could use the Weierstrass theorem).
1612: Therefore,
1613: if $i(P)=2$ then all solutions of the polynomial moment problem for $P(z)$ are reducible
1614: (cf. \cite{pa2}, Theorem 1 and \cite{pp}, Theorem 5.3).
1615: 
1616: Suppose now that the theorem is proved for all $P(z)$ with
1617: $i(P)<l$ and let $Q(z)$ be a non-zero solution of the polynomial moment problem for a polynomial $P(z)$
1618: of degree $n$ with $i(P)=l.$
1619: If $Q(z)=R(P(z))$ for some polynomial $R(z)$ then one can show as above that
1620: $P(a)=P(b)$ and 
1621: $Q(z)$ is reducible.
1622: Otherwise there exists a non-zero coefficient $s_{j_1},$ $j_1\geq m,$ of expansion
1623: \eqref{ps2} such that $j_1$ is not a multiple of $n$. By Proposition \ref{la} this implies that
1624: there exists $f_1\in D(G_P),$ $f_1\neq n,$ such that
1625: $(n/f_1) \,\vert\, j_1$.
1626: Furthermore, by Lemma \ref{lape} there exists a polynomial
1627: $S_1(z)$ such that
1628: $\psi_{f_1}(z)=S_1(P^{-1}_1(z))$ and equalities
1629: $$ P(z)=A_1(B_1(z)), \ \ \ S_1(z)=R_1(B_1(z))$$ hold for some $A_1(z),B_1(z),R_1(z)\in \C[z]$ with $\deg B_1(z)>1.$
1630: 
1631: Define a polynomial $T_1(z)$ by the equality $T_1(z)=Q(z)-S_1(z)$. Then for any $i,$ $1\leq i \leq n,$ we have:
1632: $$Q(P^{-1}_i(z))=S_1(P^{-1}_i(z))+T_1(P^{-1}_i(z)).$$
1633: Since by
1634: construction the intersection of the supports of the series
1635: $S_1(P^{-1}(z))$ and $T_1(P^{-1}(z))$ is empty, if the series $Q(P^{-1}_i(z)),$ $1\leq i \leq n,$ satisfy some linear relation over $\C$ then the series $S_1(P^{-1}_i(z)),$ $1\leq i \leq n,$ and $T_1(P^{-1}_i(z)),$
1636: $1\leq i \leq n,$ also satisfy this relation. It follows now from Theorem
1637: \ref{t1} that
1638: each of germs defined in a neighborhood of infinity by the integrals
1639: $$
1640: \hat H_1(t)= \int_{\Gamma_{a,b}} \frac{S_1(z)P^{\prime}(z)dz}{P(z)-t}\,, \ \ \ \
1641: \hat F_1(t)= \int_{\Gamma_{a,b}} \frac{ T_1(z)P^{\prime}(z)dz}{P(z)-t}\,,
1642: $$
1643: vanishes or in other words the polynomials $S_1(z)$ and $R_1(z)$ are solutions of the polynomial moment problem for $P(z).$
1644: Moreover, by
1645: construction the Puiseux series of $T_1(P^{-1}(z))$
1646: contains no non-zero coefficients with indices which are multiple of
1647: $n/ f_1$. In particular, this implies that all coefficients of $T_1(P^{-1}(z))$ whose indices are multiple of $n$
1648: vanish and hence
1649: $T_1(z)$ may not have the form $T_1(z)=R(P(z))$ for some $R(z)\in \C[z]$ unless $T_1(z)\equiv 0.$
1650: 
1651: If $T_1(t)\neq 0$ then arguing as above we conclude that
1652: there exist $f_2\in D(G_P),$ $f_2\neq f_1,$ $f_2\neq n$, and
1653: polynomials $S_2(z), T_2(z), R_2(z), A_2(z), B_2(z) \in \C[z]$ with
1654: $\deg B_2(z)>1$ such that the following conditions hold:
1655: $$T_1(P^{-1}(z))=S_2(P^{-1}(z))+T_2(P^{-1}(z)),$$
1656: $$ P(z)=A_2(B_2(z)), \ \ \ S_2(z)=R_2(B_2(z)),$$
1657: the germs
1658: $$
1659: \hat H_2(t)= \int_{\Gamma_{a,b}} \frac{S_2(z)P^{\prime}(z)dz}{P(z)-t}\,, \ \ \ \
1660: \hat F_2(t)= \int_{\Gamma_{a,b}} \frac{T_2(z)P^{\prime}(z)dz}{P(z)-t}\,
1661: $$ vanish,
1662: and the Puiseux expansion of $T_2(P^{-1}(z))$
1663: contains no non-zero coefficients whose indices are multiple of
1664: $n/ f_1$ or $n/ f_2$.
1665: 
1666: Since the number of divisors of $n$ is finite, continuing in this way, after a finite number of steps we will arrive to
1667: a decomposition of the function $Q(z)$ into a sum of polynomials $S_s(z),$ $1\leq s \leq r,$
1668: $$Q(z)=S_1(z)+S_2(z)+\dots +S_r(z)$$ such that
1669: the germs
1670: $$
1671: \hat H_s(t)= \int_{\Gamma_{a,b}} \frac{S_s(z)P^{\prime}(z)dz}{P(z)-t}\,, \ \ \ \
1672: 1 \leq s \leq r,
1673: $$
1674: vanish and $$P(z)=A_s(B_s(z)), \ \ \ \
1675: S_s(z)=R_s(B_s(z)), \ \ \ \ 1 \leq s \leq r, $$
1676: for some $R_s(z), A_s(z), B_s(z) \in \C[z]$ with $\deg B_s(z)>1.$
1677: 
1678: 
1679: In order to finish the proof it is enough to show any polynomial $S(z)$ from the collection $S_s(z),$ $1\leq s \leq r,$ is a sum of reducible solutions
1680: of the polynomial moment problem for $P(z)$. So, take some $S(z)$ and let
1681: $R(z), A(z), B(z),$ $\deg B(z)> 1,$ be polynomials
1682: such that
1683: $$P(z)=A(B(z)), \ \ \ \
1684: S(z)=R(B(z)). $$
1685: If $B(a)=B(b)$ then $S(z)$ itself is a reducible
1686: solution. Otherwise, since
1687: $$\int_{\Gamma_{a,b}} \frac{S(z)P^{\prime}(z)dz}{P(z)-t}
1688: = \int_{B(\Gamma_{a,b})}\frac{R(z)A^{\prime}(z)dz}{A(z)-t},$$ we conclude that
1689: the polynomial
1690: $R(z)$ is a solution
1691: of the polynomial moment problem for the polynomial $A(z)$ (and the points $B(a),$ $B(b)$). Since the condition
1692: $\deg B(z)>1$ implies that $i(A)<i(P)$
1693: it follows from the induction assumption that
1694: there exist polynomials $V_{1}(z),$ $V_{2}(z), \dots , V_{j}(z)$ such that
1695: $$R(z)= V_{1}(z)+V_{2}(z)+ \dots + V_{j}(z)$$
1696: and $$V_{e}(z)=\tilde V_{e}(U_{e}(z)),\ \ \
1697: A(z)=\tilde A_{e}(U_{e}(z)), \ \ \  U_{e}(B(a))=U_{e}(B(b)),$$
1698: for some  $\tilde V_{e}(z),\tilde A_{e}(z), U_{e}(z) \in \C[z],$ $1 \leq e\leq j.$
1699: 
1700: Set now $$E_e(x)=V_e(B(x)), \ \ \ W_e(z)=U_{e}(B(z)),\ \ \ 1 \leq e\leq j.$$ Then 
1701: $$S(z)=E_1(z)+E_2(z)+\dots +E_j(z),$$ where 
1702: for each
1703: $e,$ $1\leq e \leq j,$ we have:
1704: $$E_e(z)= \tilde V_{e}(W_e(z)),\ \ \ P(z)=\tilde A_e(W_e(z)), \ \ \ W_e(a)=W_e(b).$$
1705: Therefore,  
1706: $S(z)$ is a sum of reducible solutions. \qed
1707: 
1708: \vskip 0.2cm
1709: \noindent{\bf Remark.} Theorem 1.1 implies that if for a given polynomial $P(z)$ the corresponding polynomial moment problem has non-reducible solutions, then $P(z)$ has at least one ``double decomposition''
1710: $$P=A\circ B=C\circ D$$
1711: such that $$B(z)\notin \C(D(z)), \ \ \ D(z)\notin \C(B(z)).$$ 
1712: Notice that this condition
1713: is quite restrictive. Namely, the results of Engstrom \cite{en} and Ritt \cite{ri}
1714: imply that if polynomials $A,B,$ $C,D$ satisfy the equation
1715: $$ A\circ B= C\circ D $$
1716: then there exist polynomials
1717: $\hat A, \hat B, \hat C, \hat D, U, V$ such that
1718: $$A=U\circ \hat A, \ \  C=U\circ \hat C,\ \ B=\hat B\circ V, \ \  D=\hat D \circ V, \ \ \hat A\circ \hat B=\hat C\circ \hat D,$$ and
1719: up to a possible replacement of
1720: $\hat A$ by $\hat C$ and $\hat B$ by $\hat D$ either
1721: $$\hat A\circ \hat B\sim z^n \circ z^rR(z^n),  \ \ \ \ \ \ \hat C\circ \hat D
1722: \sim  z^rR^n(z) \circ z^n,$$
1723: where $R(z)$ is a polynomial, $r\geq 0,$ $n\geq 1,$ and
1724: $\GCD(n,r)=1,$ or $$\hat A\circ \hat B \sim T_n \circ T_m, \ \ \ \ \ \ \hat C\circ \hat D\sim T_m \circ T_n,$$
1725: where $T_n(z),T_m(z)$ are the corresponding Chebyshev polynomials, $n,m\geq 1,$ and $\GCD(n,m)=1.$
1726: 
1727: Notice however that a polynomial $P(z)$ may have more than {\it one} 
1728: double decomposition satisfying the condition above.
1729: Indeed, for example for any distinct prime divisors $p_1,p_2$ of a number $n$
1730: we have
1731: $$T_n(z)=T_{n/p_1}(T_{p_1}(z))=T_{n/p_2}(T_{p_2}(z))$$ and 
1732: $$T_{p_1}(z)\notin \C(T_{p_2}(z)), \ \ \ T_{p_2}(z)\notin \C(T_{p_1}(z).$$
1733: %into a composition of two polynomials equals the number
1734: %of divisors of $n$ since for any $d_1,d_2\geq 1$ the equality $T_{d_1d_2}=T_{d_1}\circ T_{d_2}$ holds.
1735: It would be interesting to investigate what
1736: conditions should be imposed on the collection $P(z),a,b$
1737: in order to conclude that any solution of the polynomial moment problem for $P(z)$ can be represented as a sum
1738: {\it at most} $r$ reducible solutions, where $r\geq 1$ is a fixed number.
1739: 
1740: 
1741: 
1742: 
1743: 
1744: 
1745: 
1746: \bibliographystyle{amsplain}
1747: \begin{thebibliography}{10}
1748: 
1749: 
1750: \bibitem {bby}  M. Blinov, M. Briskin, Y. Yomdin, {\it Local center conditions
1751: for a polynomial Abel equation and cyclicity
1752: of its zero solution}, in ``Complex analysis and dynamical systems II'', Contemp. Math., AMS, Providence, RI, 2005, 65-82.
1753: 
1754: \bibitem {bfy1} M. Briskin, J.-P. Fran\c{c}oise, Y. Yomdin,
1755: \textit{Une approche au probleme du centre-foyer de Poincar\'e,}
1756: C. R. Acad. Sci., Paris, Ser. I, Math. 326, No.11, 1295-1298 (1998).
1757: 
1758: 
1759: \bibitem {bfy2} M. Briskin, J.-P.
1760: Fran\c{c}oise, Y. Yomdin,
1761: \textit{Center conditions, compositions of polynomials and moments on
1762: algebraic curve}, Ergodic Theory Dyn. Syst. \textbf{19}, no 5,
1763: 1201-1220 (1999).
1764: 
1765: 
1766: \bibitem {bfy3} M. Briskin, J.-P. Fran\c{c}oise, Y. Yomdin,
1767: \textit{Center condition II: Parametric and model center problems}, Isr.
1768: J. Math. \textbf{118}, 61-82 (2000).
1769: 
1770: 
1771: \bibitem {bfy4} M. Briskin, J.-P. Fran\c{c}oise, Y. Yomdin,
1772: \textit{Center condition III: Parametric and model center problems}, Isr.
1773: J. Math. \textbf{118}, 83-108 (2000).
1774: 
1775: \bibitem {bfy5} M. Briskin, J.-P. Fran\c{c}oise, Y. Yomdin,
1776: \textit{Generalized moments, center-focus conditions and compositions of
1777: polynomials,} in ``Operator theory, system theory and related topics",
1778: Oper. Theory Adv. Appl.,  \textbf{123}, 161--185 (2001).
1779: 
1780: 
1781: \bibitem {bry} M. Briskin, N. Roytvarf, Y. Yomdin,
1782: {\it Center conditions at infinity for Abel differential equation}, Ann. Math., to appear.
1783: 
1784: \bibitem {by} M. Briskin, Y. Yomdin,
1785: {\it Tangential version of Hilbert 16th problem for the Abel equation},
1786: Mosc. Math. J., \textbf{5}, (2005), no. 1, 23-53.
1787: 
1788: \bibitem {c} C. Christopher, \textit{Abel equations: composition
1789: conjectures and
1790: the model problem}, Bull. Lond. Math. Soc. \textbf{32}, No.3,
1791: 332-338 (2000).
1792: 
1793: \bibitem {en} H. Engstrom, \textit{Polynomial substitutions,} Amer. J. Math. 63, 249-255 (1941).
1794: 
1795: \bibitem {fors} O. Forster, {\it
1796: Lectures on Riemann surfaces}, Graduate Texts in Mathematics, Vol. 81. New York - Heidelberg -Berlin: Springer-Verlag.
1797: 
1798: 
1799: \bibitem {gi1} K. Girstmair, {\it Linear dependence of zeros
1800: of
1801: polynomials and construction of primitive elements}, Manuscripta
1802: Math. 39 (1982), no. 1, 81--97.
1803: 
1804: 
1805: \bibitem {gi2} K. Girstmair, {\it Linear relations between roots of polynomials},
1806: Acta Arith. 89, No.1, 53-96 (1999)
1807: 
1808: 
1809: \bibitem{kir} A. Kirillov, \textit{
1810: Elements of the theory of representations,}
1811: Grundlehren der mathematischen Wissenschaften. 220, Springer-Verlag. (1976).
1812: 
1813: \bibitem{lz} S. Lando, A. Zvonkin, \textit{Graphs on surfaces and their applications},
1814: Encyclopaedia of Mathematical Sciences, 141. Low-Dimensional Topology,
1815: II. Springer-Verlag, Berlin, 2004.
1816: 
1817: \bibitem{mus} N. Muskhelishvili, \textit{Singular Integral Equations},
1818: P. Noordhoff N.V., Groningen, 1953.
1819: 
1820: \bibitem{mu} M.E. Muzychuk,
1821: \textit{The structure of rational Schur rings over cyclic groups,}
1822: Europ. J. of Combin., 14 (1993), 479-490.
1823: 
1824: \bibitem {pa1} F. Pakovich, {\it A counterexample to the ``Composition
1825: Conjecture''},
1826: Proc. AMS, 130, no. 12 (2002), 3747-3749.
1827: 
1828: \bibitem {pa2} F. Pakovich, {\it On the polynomial moment problem}, Math.
1829: Research Letters
1830: {\bf 10}, (2003), 401-410.
1831: 
1832: \bibitem {pa4} F. Pakovich, {\it Polynomial moment problem},
1833: Addendum
1834: to the paper {\it Center Problem for Abel Equation, Compositions of
1835: Functions,
1836: and Moment Conditions} by Y. Yomdin, Mosc. Math. J.,
1837: \textbf{3} (2003), no. 3, 1167-1195.
1838: 
1839: \bibitem {pa3} F. Pakovich, {\it On polynomials orthogonal to all powers
1840: of a Chebyshev polynomial on a segment}, Isr. J. Math,
1841: Vol. 142 (2004), pp. 273--283.
1842: 
1843: \bibitem {pry} F. Pakovich, N. Roytvarf and Y. Yomdin. {\it Cauchy type
1844: integrals
1845: of Algebraic functions,} Isr. J. Math., Vol. 144 (2004), pp. 221-291.
1846: 
1847: \bibitem {pp} F. Pakovich, {\it On polynomials orthogonal to all powers of a
1848: given polynomial on a segment}, Bull. Sci. Math.  129  (2005),  no. 9,
1849: 749--774.
1850: 
1851: 
1852: \bibitem {ri} J. Ritt, \textit{Prime and composite polynomials,} Trans.
1853: Amer. Math. Soc.  \textbf{23}, no. 1, 51--66 (1922).
1854: 
1855: \bibitem {ro} N. Roytvarf, \textit{Generalized moments, composition of
1856: polynomials and Bernstein classes}, in "Entire functions in modern
1857: analysis. B.Ya. Levin memorial volume", Isr. Math. Conf. Proc.
1858: \textbf{15}, 339-355 (2001).
1859: 
1860: 
1861: \bibitem {sch} A. Schinzel, \textit{Polynomials with special regard to
1862: reducibility}, Encyclopedia of Mathematics and Its Applications
1863: \textbf{77}, Cambridge University Press, 2000.
1864: 
1865: 
1866: \bibitem {schu} I. Schur, \textit{Zur Theorie der einfach transitiven Permutationsgruppen},
1867: Sitzun\-gsber. Preu\ss. Akad. Wiss., Phys.-Math. Kl. 1933, No.18/20, 598-623 (1933).
1868: 
1869: \bibitem {wi} H. Wielandt, \textit{Finite permutation groups,}
1870: New York and London: Academic Press, 1964.
1871: 
1872: \bibitem {y1}  Y. Yomdin, \textit{Center Problem for Abel Equation,
1873: Compositions of Functions, and Moment Conditions},
1874: Mosc. Math. J., \textbf{3} (2003), no. 3, 1167-1195.
1875: 
1876: \end{thebibliography}
1877: 
1878: 
1879: 
1880: 
1881: \end{document}