1:
2: \documentclass[12pt,a4paper,notitlepage]{article}
3: \voffset -2cm
4: \hoffset -1cm
5: \textwidth 16cm
6: \textheight 24cm
7: \usepackage{amsmath,amsfonts,amsthm,amssymb}
8: \usepackage{mathpple}
9: \input mathdef.tex
10: \usepackage[square]{natbib}
11: %\usepackage[notcite]{showkeys}
12: \usepackage{graphicx}
13: \theoremstyle{plain}
14: \newtheorem{theorem}{Theorem}
15: \newtheorem{lemma}{Lemma}
16: \newtheorem{proposition}{Proposition}
17: %\theoremstyle{remark}
18: %\newtheorem{remark}{Remark}
19: \theoremstyle{definition}
20: \newtheorem{example}{Example}
21: \newtheorem{remark}{Remark}
22: %
23: %
24: \begin{document}
25: %
26: \thispagestyle{plain}
27: %
28: \vspace{2em}
29: \hfill\textsf{accepted to: Annales de la Faculte des Sciences de Toulouse}
30: \vspace{2em}
31: \begin{center}
32: {\Large
33: \textbf{Differential Galois Approach to the Non-integrability of the
34: Heavy Top Problem}}\\[3em]
35: %
36: {\large\textbf{Andrzej J.~Maciejewski}}\\[1em]
37: Institute of Astronomy,
38: University of Zielona G\'ora, \\
39: Podg\'orna 50, PL-65--246 Zielona G\'ora, Poland,
40: (e-mail: maciejka@astro.ia.uz.zgora.pl)\\[1em]
41: {\large\textbf{ Maria Przybylska} }\\[1em]
42: INRIA Projet \textsc{Caf\'e},\\
43: 2004, Route des Lucioles, B.~P. 93,
44: 06902 Sophia Antipolis Cedex, France, \\
45: and \\
46: Toru\'n Centre for Astronomy,
47: Nicholaus Copernicus University, \\
48: Gagarina 11, PL-87--100 Toru\'n, Poland,
49: (e-mail: Maria.Przybylska@sophia.inria.fr)
50: \end{center}
51:
52: \vspace{2em}
53:
54: \begin{flushright}
55: \textsf{---To Jean-Pierre Ramis}
56: \end{flushright}
57: \vspace{2em}
58: {\small \textbf{Abstract}\\
59: We study integrability of the Euler-Poisson equations describing
60: the motion of a rigid body with one fixed point in a constant
61: gravity field. Using the Morales-Ramis theory and tools of
62: differential algebra we prove that a symmetric heavy top is
63: integrable only in the classical cases of Euler, Lagrange, and
64: Kovalevskaya and is partially integrable only in the
65: Goryachev-Chaplygin case. Our proof is alternative to that given
66: by Ziglin ({\em Funktsional. Anal. i Prilozhen.}, 17(1):8--23,
67: 1983; {\em Funktsional. Anal. i Prilozhen.}, 31(1):3--11, 95, 1997).}
68: \\[1em]
69: {\small \textbf{R\'esum\'e}\\
70: Nous \'etudions l'int\'egrabilit\'e des \'equations de
71: Euler-Poisson qui d\'ecrivent le mouvement d'un solide rigide avec
72: un point fixe dans un champ gravitationnel constant. En utilisant
73: la th\'eorie de Morales-Ramis et des outils d'alg\`ebre
74: diff\'erentielle, nous prouvons qu'un solide sym\'etrique est
75: int\'egrable seulement dans les cas classiques d'Euler, Lagrange
76: et Kowalevski, et est partiellement int\'egrable seulement dans le
77: cas Goryatchev-Tchaplygin. Notre preuve est une alternative \`a
78: celle donn\'ee par Ziglin ({\em Funktsional. Anal. i Prilozhen.},
79: 17(1):8--23, 1983; {\em Funktsional. Anal. i Prilozhen.},
80: 31(1):3--11, 95, 1997).} \newpage
81:
82: %
83: %----------->
84: %
85:
86: \section{Equations of motion and motivation}
87: %
88: \label{sec:eqandm}
89: %
90: Equations of motion of a rigid body in external fields are usually
91: written in a body fixed frame. Here, we use the following convention.
92: For a vector $\vv$ we denote by $\bV=[V_1,V_2,V_3]^T$ its coordinates
93: in a body fixed frame, and we consider it as a one column matrix. The
94: vector and scalar products of two vectors $\vv$ and $\vw$ expressed in
95: terms of the body fixed coordinates are denoted by $[\bV,\bW]$ and
96: $\rscalar{\bV}{\bW}$, respectively.
97:
98: %
99: We consider a rigid body of mass $m$ located in a constant gravity
100: field of intensity $g$. One point of the body is fixed. The distance
101: between the fixed point and the mass centre of the body is $D$. Assuming
102: that $gD\neq 0$, we choose units in such a way that $\mu:=mgD=1$. The
103: Euler-Poisson equations
104: %
105: \begin{equation}
106: \label{eq:ep}
107: \Dt \bM =[ \bM,\bJ\bM]+ [\bN, \bL], \qquad
108: \Dt \bN = [\bN , \bJ\bM],
109: \end{equation}
110: %
111: describe the rotational motion of the body. In the above equations
112: $\bM$ denotes the angular momentum of the body, $\bN$ is the unit
113: in the direction of the gravity field, and $\bL$ is the
114: unit vector from the centre of mass of the body to the fixed point;
115: $\bJ$ is the inverse of the matrix of inertia, so $\bOmega:=\bJ\bM$ is
116: the angular velocity of the body. The principal moments of inertia
117: $A$, $B$, and $C$ are eigenvalues of $\bJ^{-1}$. For our further
118: consideration it is important to notice that in~\eqref{eq:ep} the body
119: fixed frame is unspecified, so we can choose it according to our
120: needs. A body fixed frame in which $\bJ$ is diagonal is called the
121: principal axes frame. This frame is uniquely defined (up to the
122: numbering of the axes) when $\bJ$ has no multiple eigenvalues. If
123: $\bJ$ has a multiple eigenvalue, e.g. when $A=B$, we say that the body
124: is symmetric. Then the principal axes frame is defined up to a
125: rotation around the symmetry axis.
126:
127: System~\eqref{eq:ep} depends on parameters $A$, $B$, $C$, and $\bL$
128: but physical constraints restrict the allowable values of parameters to
129: a set $\cP\subset\R^6$ defined by the following conditions
130: %
131: \begin{gather*}
132: \label{eq:par}
133: A>0,\quad B>0, \quad C>0, \qquad \rscalar{\bL}{\bL}=1, \\
134: A+B\geq C, \quad B+C\geq A, \quad C + A\geq B.
135: \end{gather*}
136:
137: Euler-Poisson equations possess three first integrals
138: %
139: \begin{gather}
140: \label{eq:H}
141: H = \frac{1}{2}\rscalar{\bM}{\bJ\bM} + \rscalar{\bN}{\bL},\\
142: \label{eq:H1H2}
143: H_1 = \rscalar{\bM}{\bN}, \qquad H_2 = \rscalar{\bN}{\bN}.
144: \end{gather}
145: %
146: It is known that on the level
147: %
148: \begin{equation*}
149: \label{eq:m4}
150: M_\chi = \{ (\bM,\bN)\in \R^6\ | \ H_1 = \chi, H_2 = 1 \}
151: \end{equation*}
152: %
153: the Euler-Poisson equations are the Hamiltonian ones, see e.g.
154: \cite{Arnold:78::,Marsden:94::,Audin:96::}.
155: %
156: \begin{remark}
157: \label{rem:r1}
158: \textsl{ Configuration space of a rigid body with a fixed point is the
159: Lie group $\mathrm{SO}(3,\R)$ (all possible orientations of the body
160: with respect to an inertial frame). Thus, classically, the phase
161: space for the problem is $T^*\mathrm{SO}(3,\R)$. Hence it is a
162: Hamiltonian system with three degrees of freedom and it possesses
163: one additional first integral $H_1$ (first integral $H_2$ is
164: identically equal to one in this formulation). The existence of this
165: first integral is related to the symmetry of the problem (rotations
166: around the direction of the gravity field) and this allows to reduce
167: the system by one degree of freedom. The Euler-Poisson equations
168: can be viewed as an effect of reduction of the system on
169: $T^*\mathrm{SO}(3,\R)$ with respect to this symmetry. The phase
170: space of the reduced problem can be considered as the dual $\efg^*$
171: to Lie algebra $\efg$ of group of rigid motions $G=\R^3
172: \rtimes\mathrm{SO}(3,\R)$. We can identify $(\bN,\bM)$ as element of
173: $\efg^*={\R^{3*}}\rtimes \mathrm{so}(3,\R)^*$ using standard
174: isomorphisms between $ \mathrm{so}(3,\R)$ and $\R^{3}$, and between
175: ${\R^{3*}}$ and $\R^3$. Let $(\bX,\bA)\in G$, with $\bX\in\R^3$ and
176: $\bA \in \mathrm{SO}(3,\R)$. Then the coadjoint action
177: $\mathrm{Ad}^*_{(\bX,\bA)}:\efg^*\rightarrow\efg^* $ is defined by
178: %
179: \[
180: \mathrm{Ad}^*_{(\bX,\bA)}(\bN,\bM)=( \bA\bN,[\bX,\bA \bN] + \bA\bM).
181: \]
182: %
183: Functions $H_1$ and $H_2$ given by~\eqref{eq:H1H2} are orbits
184: invariant, i.e., on each orbit of coadjoint action
185: \[
186: \cO_{(\bN,\bM)}:= \bigcup_{(\bX,\bA)\in G}\{( \bA\bN,[\bX,\bA \bN] + \bA\bM)\}
187: \]
188: they have constant
189: values. They are called Casimirs. Orbits of coadjoint action defined
190: above coincide with $M_\chi$ which is diffeomorphic to $T\bbS^2$. As
191: it is well known, orbits of a coadjoint action are symplectic
192: manifolds equipped with the standard Kostant-Berezin-Souriau-Kirillov
193: symplectic structure \cite{Kirillov:76::,Souriau:70::}. On four
194: dimensional orbits $M_\chi$ the Euler-Poisson equations are
195: Hamiltonian with $H$ given by~\eqref{eq:H} as the Hamilton
196: function. }
197: \end{remark}
198: %
199: Thus, as a Hamiltonian system on $M_\chi$ the Euler-Poisson equations
200: have two degrees of freedom and are integrable on $M_\chi$ if there
201: exists an additional first integral $H_3$ which is functionally
202: independent with $H$ on $M_\chi$. Equivalently, we say that the
203: Euler-Poisson equations are integrable if there exists a first
204: integral $H_3$ which is functionally independent together with $H$,
205: $H_1$ and $H_2$.
206:
207: We say that the Euler-Poisson equations are partially integrable if
208: they are integrable on $M_0$.
209:
210: The known integrable cases are the following:
211: %
212: \begin{enumerate}
213: \item The Euler case (1758) corresponds to the situation when there is
214: no gravity (i.e. when $\mu=0$) or $\bL=\bzero$ (the fixed point of
215: the body is the centre of mass). The additional first integral in
216: this case is the total angular momentum $H_3 = \rscalar{\bM}{\bM}$.
217: \item In the Lagrange case~\cite{Lagrange:1889::} the body is
218: symmetric (i.e. two of its principal moments of inertial are equal)
219: and the fixed point lies on the symmetry axis. The additional first
220: integral in this case is the projection of the angular momentum onto
221: the symmetry axis. If we assume that $A=B$, then in the Lagrange
222: case $L_1=L_2=0$, and $H_3=M_3$.
223: \item In the Kovalevskaya
224: case~\cite{Kowalevski:1888::,Kowalevski:1890::} the body is
225: symmetric and the principal moment of inertia along the symmetry
226: axis is half of the principal moment of inertia with respect to an
227: axis perpendicular to the symmetry axis. Moreover, the fixed point
228: lies in the principal plane perpendicular to the symmetry axis. If
229: $A=B=2C$, then (after an appropriate rotation around the symmetry
230: axis) we have in the Kovalevskaya case $L_2=L_3=0$. The additional
231: first integral has the form
232: %
233: \[
234: H_3 = \left( \frac{1}{2}(M_1^2-M_2^2)+N_1\right)^2 +(M_1M_2+N_2)^2.
235: \]
236: %
237: \item In the Goryachev-Chaplygin case~\cite{ Goryachev:1910::} the body
238: is symmetric and, as in the Kovalevskaya case, the fixed point lies
239: in the principal plane perpendicular to the symmetry axis. If we
240: assume that the third principal axis is the symmetry axis, then in
241: the Goryachev-Chaplygin case we have $A=B=4C$ and $L_2=L_3=0$. In
242: the Goryachev-Chaplygin case equations~\eqref{eq:ep} are integrable
243: only on the level $H_1=0$ and the additional first integral has the
244: following form:
245: \[
246: H_3 =
247: M_3(M_1^2+M_2^2) + M_1N_3.
248: \]
249: \end{enumerate}
250: %
251: For a long time the question if there are other integrable cases of
252: the Euler-Poisson equations except those enumerated above was open,
253: although many leading mathematicians tried to give a positive answer
254: to it.
255:
256: The problem was completely solved by S.L.~Ziglin in a brilliant way.
257: First, in \cite{Ziglin:80::d} he proved the following.
258: %
259: \begin{theorem}[Ziglin, 1980]
260: If $(A-B)(B- C)(C-A)\neq 0$, then the Euler-Poisson system
261: does not admit a real meromorphic first integral which is
262: functionally independent together with $H$, $H_1$ and $H_2$.
263: \end{theorem}
264: %
265: Later, he developed in \cite{Ziglin:82::b} an elegant method, which
266: now is called the Ziglin theory, and using it he proved
267: in~\cite{Ziglin:83::b} the following theorem.
268: %
269: \begin{theorem}[Ziglin, 1983]
270: \label{thm:z2}
271: The complexified Euler-Poisson system for a symmetric body is
272: integrable on $M_0$ with complex meromorphic first integrals only in
273: the four classical cases.
274: \end{theorem}
275: %
276: This result was improved in \cite{Ziglin:97::}.
277: %
278: \begin{theorem}[Ziglin, 1997]
279: \label{thm:z3}
280: The Euler-Poisson system for a symmetric body is integrable on $M_0$
281: with real meromorphic first integrals only in the four classical
282: cases.
283: \end{theorem}
284: %
285: We know that in the Euler, Lagrange and Kovalevskaya cases the
286: Euler-Poisson system is integrable globally, i.e. the additional first
287: integral exists on an arbitrary symplectic manifold $M_\chi$. In the
288: Goryachev-Chaplygin case the Euler-Poisson equations are integrable
289: only on $M_0$, and this fact was also proved by Ziglin.
290:
291: \begin{remark}
292: \textsl{ We do not even try to sketch the very rich history of
293: investigations of the problem of the heavy top. We refer here to
294: books \cite{Golubev:53::,Leimanis:65::,Kozlov:80::,Borisov:01::}
295: and references therein. However, several works where the question
296: of integrability of the problem was investigated are worth
297: mentioning. In \cite{Kozlov:75::a} it was shown that for a
298: non-symmetric body the Euler-Poisson equations do not admit an
299: additional real analytic first integral which depends analytically
300: on a small parameter $\mu$, see also \cite[Ch. III]{Kozlov:80::}.
301: For a symmetric body when the ratio of the principal moments is
302: small enough (i.e. for the case of the perturbed spherical pendulum)
303: the non-existence of an additional real analytic first
304: integral was proved \cite{Kozlov:85::c,Kozlov:86::}. A similar
305: result for a perturbed Lagrange case was shown
306: in~\cite{Dovbysh:90::,Bolotin:90::}. A novel, variational approach
307: to proving the non-integrability was elaborated by S.~V.~Bolotin
308: in~\cite{Bolotin:92::a} where he showed the non-existence of an
309: additional real analytic first integral for a symmetric heavy top
310: for the case when the fixed point lies in the equatorial and the
311: ratio of the principal moments of inertia is greater than 4.}
312: \end{remark}
313:
314: The Ziglin theory is a continuation of the idea of S.~N.~Kovalevskaya
315: who related the (non)integrability with the behaviour of solutions of the
316: investigated system as functions of the complex time. The main object in
317: the Ziglin theory is the monodromy group of variational equations
318: around a particular non-equilibrium solution. As it was shown by
319: Ziglin, if the investigated system possesses a meromorphic first
320: integral, the monodromy group of variational equations possesses a
321: rational invariant. Thus, for an integrable system, the monodromy group
322: cannot be too `rich'. The main difficulty in application of the Ziglin
323: theory is connected with the fact that \emph{`except for a few
324: differential equations, e.g., the Riemann equations,
325: Jordan-Pochhammer equations and generalised hypergeometric
326: equations, the monodromy group has not been determined'},
327: \cite[][p.~85]{Takano:76::}. For other second order
328: differential equations only partial results are known, see e.g.
329: \cite{Churchill:90::b,Churchill:91::a}. Having this in mind, one can
330: notice that the analysis of the variational equations and
331: determination of properties of their monodromy group given by Ziglin
332: in his proof of Theorem~\ref{thm:z2} is a masterpiece. Later, the
333: Ziglin theory was developed and applied for a study of non-integrability of
334: various systems but, as far as we know, nobody used Ziglin's brilliant
335: technique developed in his proof of~Theorem~\ref{thm:z2}.
336:
337:
338: In the nineties the theory of Ziglin was extended by a differential
339: Galois approach. It was done independently by C.~Sim\'o,
340: J.~J.~Morales-Ruiz,
341: J.-P.~Ramis~\cite{Morales:94::b,Morales:99::c,Morales:01::b1,Morales:01::b2}
342: and A.~Braider, R.~C.~Churchill, D.~L.~Rod and M.~F.~Singer
343: \cite{Churchill:95::a,Churchill:96::b}. Nowadays, this approach is
344: called the Morales-Ramis theory. The key point in this theory is to
345: replace an investigation of the monodromy group of variational equations
346: by a study of their differential Galois group. The main fact from this
347: theory is similar to that of Ziglin: the existence of a meromorphic
348: first integral implies the existence of a rational invariant of the
349: differential Galois group of variational equations. Forgetting
350: about differences in hypotheses in main theorems of both theories, the
351: biggest advantage of the Morales-Ramis theory is connected with the
352: fact that applying it, we have at our disposal developed tools and
353: algorithms of differential algebra. Thanks to this fact, it can be
354: applied more easily.
355:
356: We applied the Morales-Ramis theory to study integrability of several
357: systems, see e.g.
358: \cite{Maciejewski:01::i,Maciejewski:01::j,Maciejewski:02::a,Maciejewski:02::b,Maciejewski:02::f,Maciejewski:03::a,Maciejewski:03::e},
359: and we notice that obtaining similar results when working only with
360: the monodromy group is questionable, or, at least difficult. This
361: gives us an idea to reanalyse the Ziglin proof of Theorem~\ref{thm:z2} which
362: is rather long (about 10 pages in~\cite{Ziglin:83::b}). We wanted to
363: present a new, much shorter and simpler proof which is based on the
364: Morales-Ramis theory and tools from differential algebra. In fact, at
365: the beginning, we believed that giving such proof would be a
366: nice and simple exercise but quickly it appeared that we were wrong. A
367: `naive' application of the Morales-Ramis theory leads quickly to
368: very tedious calculations or unsolvable complications. As our aim was
369: to give an `elementary' proof, we put a constrain on the arguments
370: which are allowable in it: no computer algebra. Thus we spent a lot of
371: time analysing sources of difficulties and complications, and the aim
372: of this paper is to present our own version of proofs of
373: Theorem~\ref{thm:z2} and Theorem~\ref{thm:z3}. As we believe, these
374: proofs present the whole power and beauty of the Morales-Ramis
375: theory.
376:
377: The plan of this paper is following. To make it self-contained in the
378: next section we shortly describe basic facts from the Morales-Ramis
379: and Ziglin theory. We collected more specific results about special
380: linear differential equations in Appendix. In Section~\ref{sec:pandve}
381: we derive the normal variational equation. Sections~\ref{sec:Proof2}
382: and~\ref{sec:Proof3} contain our proofs of Theorem~\ref{thm:z2}
383: and~\ref{thm:z3}. In the last section we give several remarks and
384: comments.
385: %
386: %
387: %
388: \section{Theory}
389: %
390: %
391: \label{sec:theory}
392: %
393: %
394: %
395: Below we only mention basic notions and facts concerning the Ziglin
396: and Morales-Ramis theory following
397: \cite{Ziglin:82::b,Churchill:96::b,Morales:99::c}.
398:
399: Let us consider a system of differential equations
400: %
401: \begin{equation}
402: \label{eq:ds}
403: \Dt x = v(x), \qquad t\in\C, \quad x\in M,
404: \end{equation}
405: %
406: defined on a complex $n$-dimensional manifold $M$. If $\varphi(t)$ is
407: a non-equilibrium solution of \eqref{eq:ds}, then the maximal analytic
408: continuation of $\varphi(t)$ defines a Riemann surface $\Gamma$ with
409: $t$ as a local coordinate. Together with system \eqref{eq:ds} we can
410: also consider variational equations (VEs) restricted to $T_{\Gamma}M$,
411: i.e.
412: %
413: \begin{equation*}
414: \label{eq:vds}
415: \dot \xi = T(v)\xi, \qquad \xi \in T_\Gamma M.
416: \end{equation*}
417: %
418: We can always reduce the order of this system by one considering the
419: induced system on the normal bundle $N:=T_\Gamma M/T\Gamma$ of
420: $\Gamma$ \cite{Ziglin:82::b}
421: %
422: \begin{equation}
423: \label{eq:gnve}
424: \dot \eta = \pi_\star(T(v)\pi^{-1}\eta), \qquad \eta\in N.
425: \end{equation}
426: %
427: Here $\pi: T_\Gamma M\rightarrow N$ is the projection. The system of
428: $s=n-1$ equations obtained in this way yields the so-called normal
429: variational equations (NVEs). The monodromy group $\cM$ of system
430: \eqref{eq:gnve} is the image of the fundamental group
431: $\pi_1(\Gamma,t_0)$ of $\Gamma$ obtained in the process of
432: continuation of local solutions of \eqref{eq:gnve} defined in a
433: neighbourhood of $t_0$ along closed paths with the base point $t_0$.
434: By definition, it is obvious that $\cM\subset\mathrm{GL}(s,\C)$. A
435: non-constant rational function $f(z)$ of $s$ variables $z=(z_1,\ldots,
436: z_s)$ is called an integral (or invariant) of the monodromy group if $
437: f(g\cdot z)=f(z) $ for all $g\in\cM$.
438:
439: From the Ziglin theory we need the basic lemma formulated
440: in~\cite{Ziglin:82::b} and then given in an improved form
441: in~\cite{Ziglin:97::}.
442:
443: %
444: \begin{lemma}
445: \label{lem:zig}
446: If system \eqref{eq:ds} possesses a meromorphic first integral defined
447: in a neighbourhood $U\subset M$, such that the fundamental group of
448: $\Gamma$ is generated by loops lying in $U$, then the monodromy group
449: $\cM$ of the normal variational equations has a
450: rational first integral.
451: \end{lemma}
452:
453:
454: If system \eqref{eq:ds} is Hamiltonian, then necessarily $n=2m$ and
455: $M$ is a symplectic manifold equipped with a symplectic form $\omega$.
456: The right hand sides $v=v_H$ of \eqref{eq:ds} are generated by a
457: single function $H$ called the Hamiltonian of the system. For given
458: $H$ vector field $v_H$ is defined by $\omega(v_H,u)=\rmd H\cdot u$,
459: where $u$ is an arbitrary vector field on $M$. Then, of course, $H$ is
460: a first integral of the system. For a given particular solution
461: $\varphi(t)$ we fix the energy level $E=H(\varphi(t))$. Restricting
462: \eqref{eq:ds} to this level, we obtain a well defined system on an
463: $(n-1)$ dimensional manifold with a known particular solution
464: $\varphi(t)$. For this restricted system we perform the reduction of
465: order of variational equations. Thus, the normal variational
466: equations for a Hamiltonian system with $m$ degrees of freedom have
467: dimension $s=2(m-1)$ and their monodromy group is a subgroup of
468: $\mathrm{Sp}(s,\C)$.
469:
470:
471: In the Morales-Ramis theory the differential Galois group $\cG$ of
472: normal variational equations plays the fundamental role, see
473: \cite{Morales:99::c,Morales:01::b1}. For a precise definition of the
474: differential Galois group and general facts from differential algebra
475: see \cite{Kaplansky:76::,Ramis:90::,Beukers:92::,Magid:94::,Put:02::}.
476: We can consider $\cG$ as a subgroup of $\mathrm{GL}(s,\C)$ which acts
477: on fundamental solutions of \eqref{eq:gnve} and does not change
478: polynomial relations among them. In particular, this group maps one
479: fundamental solution to other fundamental solutions. Moreover, it can
480: be shown that $\cM\subset \cG$ and $\cG$ is an algebraic subgroup of
481: $\mathrm{GL}(s,\C)$. Thus, it is a union of disjoint connected
482: components. One of them containing the identity is called the identity
483: component of $\cG$ and is denoted by $\cG^0$.
484:
485: Morales-Ruiz and Ramis formulated a new criterion of
486: non-in\-teg\-ra\-bi\-li\-ty for Hamiltonian systems in terms of the
487: properties of $\cG^0$ \cite{Morales:99::c,Morales:01::b1}.
488: %
489: \begin{theorem}
490: \label{thm:MR}
491: Assume that a Hamiltonian system is meromorphically integrable in
492: the Liouville sense in a neigbourhood of the analytic curve
493: $\Gamma$. Then the identity component of the differential Galois
494: group of NVEs associated with $\Gamma$ is Abelian.
495: \end{theorem}
496: %
497: In most applications the Riemann surface $\Gamma$ associated with the
498: particular solution is open. There are many reasons why it is better to
499: work with compact Riemann surfaces. Because of this, it is customary
500: to compactify $\Gamma$ adding to it a finite number of points at
501: infinity. Doing this we need a refined version of
502: Theorem~\ref{thm:MR}, for details see~
503: \cite{Morales:99::c,Morales:01::b1}. However, in the context of this
504: paper, the thesis of the above theorem remains unchanged if instead
505: of $\Gamma$ and the variational equations over $\Gamma$ we work with
506: its compactification.
507: %
508: %
509: \section{Particular solutions and variational equations}
510: %
511: \label{sec:pandve}
512: %
513: %
514: To apply the Ziglin or the Morales-Ramis theory we have to know a
515: non-equilibrium solution. Let us assume that the fixed point is
516: located in a principal plane. Then, in fact, we can find a one
517: parameter family of particular solutions which describe a pendulum
518: like motion of the body. For a symmetric body this assumption
519: is not restrictive (if necessary we can rotate the principal axes
520: around the symmetry axis).
521:
522: We choose the body fixed frame in the following way. Its first two
523: axes lie in the principal plane where the fixed point is located and
524: the first axis has direction from the fixed point to the centre of
525: mass of the body. We call this frame special. A map given by
526: %
527: \begin{equation*}
528: (\bM,\bN) = (\bR \widetilde \bM, \bR \widetilde \bN),\qquad
529: \bR\in\mathrm{SO}(3,\R),
530: \end{equation*}
531: %
532: transforms equations~\eqref{eq:ep} to the form
533: %
534: \begin{equation*}
535: \label{eq:ept}
536: \Dt \widetilde \bM =[ \widetilde \bM, {\widetilde \bJ}\, {\widetilde \bM}]+
537: [ \widetilde \bN, \widetilde \bL], \qquad
538: \Dt \widetilde \bN = [ {\widetilde \bN} , {\widetilde \bJ}\, {\widetilde \bM}],
539: \end{equation*}
540: %
541: where
542: %
543: \[
544: \widetilde \bJ =\bR^T \bJ \bR,\qquad \widetilde \bL=\bR^T \bL.
545: \]
546: %
547: Now, if symbols with tilde correspond to the principal axes frame,
548: then, taking into account our assumption about the location of the
549: fixed point, we have
550: %
551: \[
552: \widetilde \bJ = \begin{bmatrix} {A}^{-1} &0 &0\\
553: 0 & {B}^{-1} &0\\
554: 0 & 0& C^{-1} \end{bmatrix}, \qquad
555: \widetilde \bL = [ {\widetilde L}_1, 0,{\widetilde L}_3]^T.
556: \]
557: %
558: Taking
559: %
560: \[
561: \bR = \begin{bmatrix}
562: \phantom{-}{\widetilde L}_1 & 0 & {\widetilde L}_3\\
563: 0&1 &0 \\
564: -{\widetilde L}_3 & 0 & {\widetilde L}_1
565: \end{bmatrix},
566: \]
567: %
568: we obtain
569: %
570: \begin{equation*}
571: \bL=[1,0,0]^T, \qquad
572: \bJ = \begin{bmatrix}
573: a & 0& 2d\\
574: 0 & b & 0\\
575: 2d & 0 & c
576: \end{bmatrix},
577: \end{equation*}
578: %
579: where
580: %
581: \begin{gather}
582: \label{eq:ac}
583: a = \frac{{\widetilde L}_1^2}{A} + \frac{{\widetilde L}_3^2}{C}, \qquad
584: c = \frac{{\widetilde L}_1^2}{C} + \frac{{\widetilde L}_3^2}{A}, \notag\\
585: \label{eq:d}
586: 2 d = \left( \frac{1}{C}-\frac{1}{A} \right){\widetilde L}_1 {\widetilde L}_3,
587: \qquad
588: b = \frac{1}{B}.
589: \end{gather}
590: %
591: Thus, the prescribed choice of $\bR$ corresponds to the transformation
592: from the special to the principal frame. Without loss of generality
593: we can put $b=1$. For a symmetric body we assume that $A=B\neq C$.
594: Under this assumption, if $d=0$, then $ {\widetilde L}_1 {\widetilde
595: L}_3=0$, and, in this case, the special frame coincides with the
596: principal frame.
597: %
598:
599: From now on we consider the complexified Euler-Poisson system, i.e. we
600: assume that $(\bM,\bN)\in\C^6$.
601:
602: \subsection{Case $d\neq0$}
603:
604: It is easy to check that manifold
605: %
606: \begin{equation*}
607: \label{eq:inv}
608: \cN:= \{ ( \bM,\bN)\in \C^6\ | \ M_1=M_3=N_2 = 0, N_1^2+N_3^2=1\}\subset M_0,
609: \end{equation*}
610: %
611: is symplectic sub-manifold of $M_0$ diffeomorphic to
612: $T\,\bbS^1_{\C}\subset T\,\bbS^2_{\C}$ (by $\bbS^m_{\C}$ we denote
613: $m$-dimensional complex sphere). Moreover, $\cN$ is invariant with
614: respect to the flow generated by \eqref{eq:ep}. The Euler-Poisson
615: equations restricted to $\cN$ have the following form
616: %
617: \begin{equation}
618: \label{eq:rep}
619: \Dt M_2 = N_3, \qquad
620: \Dt N_1=-M_2N_3, \qquad
621: \Dt N_3= M_2N_1,
622: \end{equation}
623: %
624: and are Hamiltonian with $H_{|\cN}$ as the Hamiltonian function.
625: For each level of Hamiltonian $H_{|\cN}=e:= 2k^2 - 1$, we obtain a
626: phase curve $\Gamma_k$. We restrict our
627: attention to curves corresponding to $e\in(-1,1]$ so that $k\in(0,1]$.
628: A solution of system~\eqref{eq:rep} lying on the level $H_{|\cN}= 2k^2
629: - 1$ we denote by $(M_2(t,k),N_1(t,k),N_3(t,k))$.
630:
631: For a generic value of $k$ phase curve $\Gamma_k$ is an algebraic
632: curve in $\C^3\{M_2,N_1,N_3\}$, and, as intersection of two quadrics
633: %
634: \begin{equation*}
635: 2k^2 - 1 = \frac{1}{2}M_2^2 + N_1, \qquad N_1^2 + N_3^2=1,
636: \end{equation*}
637: %
638: is an elliptic curve (for $k=1$ it is a rational curve). We can
639: compactify it adding two points at infinity which lie in directions
640: $(0,\pm\rmi,1)$. Thus, a generic $\Gamma_k$ can be considered as a
641: torus with two points removed. In our further consideration we work
642: with subfamily $\Gamma_k$ with $k\in(0,1)$. Only in
643: Section~\ref{sec:Proof3} we refer to the phase curve $\Gamma_1$
644: corresponding to $k=1$.
645:
646: Equations~\eqref{eq:rep} describe the pendulum-like motions of the
647: body: the symmetry axis of the body remains permanently in one plane
648: and oscillates or rotates in it around the fixed point.
649:
650: For a point $p=(\bM,\bN)\in M_0$ by $\bv=(\bm,\bn)$ we denote a vector
651: in $T_p M_0$. Variational equations along phase curve $\Gamma_k$ have
652: the following form
653: %
654: %
655: \begin{equation*}
656: \label{eq:varall}
657: \Dt
658: \begin{bmatrix}
659: m_1\\ m_2 \\m_3 \\ n_1 \\ n_2 \\ n_3
660: \end{bmatrix}
661: =
662: \begin{bmatrix}
663: 2 d M_2 & 0 & (c-1)M_2 & 0 & 0 & 0\\
664: 0 & 0 & 0 & 0 & 0 & 1 \\
665: (1-a)M_2 & 0 & -2 d M_2 & 0 & -1 & 0 \\
666: 0 & -N_3 & 0 & 0 & 0&-M_2 \\
667: a N_3 -2 d N_1& 0 & 2 d N_3-c N_1& 0 & 0 & 0 \\
668: 0 & N_1 & 0 & M_2 & 0 & 0
669: \end{bmatrix}
670: \begin{bmatrix}
671: m_1\\ m_2 \\m_3 \\ n_1 \\ n_2 \\ n_3
672: \end{bmatrix},
673: \end{equation*}
674: %
675: where $(M_2,N_1,N_3)=(M_2(t,k),N_1(t,k),N_3(t,k)) \in\Gamma_k$. They have
676: the following first integrals
677: %
678: \begin{equation*}
679: \label{eq:ivar}
680: h = M_2 m_2 + n_1 , \qquad h_1 = N_1m_1 + N_3m_3 + M_2n_2, \qquad
681: h_2 = N_1n_1 + N_3 n_3.
682: \end{equation*}
683: %
684:
685: As it was shown by Ziglin, the normal variational equations are given by
686: %
687: \begin{equation*}
688: \label{eq:nve }
689: \begin{split}
690: \Dt m_1 &= 2 dM_2 m_1 + (c-1)M_2 m_3, \\
691: \Dt m_3 & =(1-a)M_2 m_1 -2 dM_2 m_3 - n_2, \\
692: \Dt n_2 &= ( a N_3 -2 d N_1) m_1 + ( 2 d N_3-c N_1 ) m_3, \\
693: 0 & = N_1m_1 + N_3m_3 + M_2n_2.
694: \end{split}
695: \end{equation*}
696: %
697: We assume that the particular solution is not a stationary point
698: ($M_2(t,k)\equiv0$, $N_1(t,k)\equiv\pm1$, $N_3(t,k)\equiv0$). Under this
699: assumption we reduce the above system to the form
700: %
701: %
702: \begin{equation}
703: \label{eq:nve2}
704: \begin{split}
705: \Dt m_1 &= 2 dM_2 m_1 + (c-1)M_2 m_3, \\
706: \Dt m_3 & =\left[\frac{N_1}{M_2}+(1-a)M_2\right] m_1+
707: \left[\frac{N_3}{M_2} -2 dM_2\right] m_3.
708: \end{split}
709: \end{equation}
710: %
711: %
712: We can write the above system as one second order equation
713: %
714: \begin{equation}
715: \label{eq:rnve}
716: \Dtt m + a_1(t) \Dt m + a_0(t) m = 0, \qquad m \equiv m_1,
717: \end{equation}
718: %
719: with coefficients
720: %
721: \[
722: a_1(t) =- 2\frac{N_3(t,k)}{M_2(t,k)}, \qquad
723: a_0(t) = (1-c)N_1(t,k) + 2 d N_3(t,k) + f M_2(t,k)^2,
724: \]
725: and
726: \[
727: f = (a-1)(c-1)-4 d^2.
728: \]
729: %
730: Now, we make the following transformation of independent variable
731: which is rational parametrisation of the complex circle $\bbS^1_{\C}$
732: %
733: \begin{equation}
734: \label{eq:tran}
735: t \longrightarrow z := \frac{N_3(t,k)}{1+N_1(t,k)}.
736: \end{equation}
737: %
738: Then we obtain
739: %
740: \begin{gather*}
741: N_1 = \frac{1-z^2}{1+z^2}, \qquad N_3 = \frac{2 z}{1+z^2}, \qquad
742: M_2 = \frac{2\dot z}{1+z^2}, \\
743: {\dot z}^2 = \frac{1}{s^2+1}(z^2+1)(z^2-s^2).
744: \end{gather*}
745: %
746: where
747: %
748: \begin{equation*}
749: s = \sqrt{\frac{1-e}{1+e}}=\frac{k'}{k}.
750: \end{equation*}
751: After transformation~\eqref{eq:tran} equation \eqref{eq:rnve} reads
752: %
753: \begin{equation}
754: \label{eq:var}
755: m''+p(z)m'+q(z)m=0,\qquad '=\dfrac{\mathrm{d}}{\mathrm{d}z},
756: \end{equation}
757: with coefficients
758: %
759: \begin{equation*}
760: p(z)= \frac{1}{2}\left[ \frac{3}{z-\rmi} + \frac{3}{z+\rmi} -
761: \frac{1}{z-s} - \frac{1}{z+s}\right],
762: \end{equation*}
763: %
764: and
765: %
766: \begin{equation*}
767: q(z) = \sum_{i=1}^4 \frac{\alpha_i}{(z-z_i)^2} + \frac{\beta_i}{z-z_i},
768: \end{equation*}
769: %
770: where we denote $(z_1, z_2,z_3,z_4)=(\rmi,-\rmi, s, -s)$, and
771: %
772: \begin{gather*}
773: \alpha_1 = \frac{1}{2}(1-c) - f + \rmi d, \qquad \alpha_2 =\alpha_1^*,
774: \qquad \alpha_3 =\alpha_4 = 0, \\
775: \beta_1 = -\frac{2d}{1+s^2} - \rmi \left(f+\frac{c-1}{1+s^2}\right),\qquad
776: \beta_2 = \beta_1^*,\\
777: \beta_3 = \frac{(1-c)(1-s^2) + 4d s}{2s(1+s^2)}, \qquad
778: \beta_4 = \frac{(c-1)(1-s^2) + 4d s}{2s(1+s^2)}.
779: \end{gather*}
780: %
781: We can see that equation \eqref{eq:var} is Fuchsian and it has four
782: regular singular points $z_i$ over the Riemann sphere \CPOne. The
783: infinity is an ordinary point for this equation. We assumed that
784: $k \in(0,1)$, so $s\in (0,\infty)$. Here it is important to notice
785: that for real values of $a$, $c$, and $d$, and for all $s\in
786: (0,\infty)$ equation~\eqref{eq:var} has four distinct regular singular
787: points, i.e, the number of singular points does not depend on $s$. For
788: further calculations we fix $s=1$.
789:
790: Let us note here that transformation~\eqref{eq:tran} is a branched double
791: covering of Riemann sphere $\C\bbP^1\rightarrow \Gamma_k$. Moreover, the
792: branching points of this covering are precisely the four points where
793: equation~\eqref{eq:var} has singularities.
794:
795: Making the following change of the dependent variable
796: %
797: \begin{equation*}
798: m=w\exp\left[-\dfrac{1}{2}\int_{z_0}^z p(\zeta)\mathrm{d}\zeta\right],
799: \end{equation*}
800: %
801: we can simplify \eqref{eq:var} to the standard reduced form
802: \begin{equation}
803: \label{eq:zred}
804: w''=r(z)w,\qquad r(z)=\dfrac{1}{2}p'(z)+\dfrac{1}{4}p(z)^2-q(z),
805: \end{equation}
806: %
807: where $r(z)$ can be
808: written as
809: %
810: \begin{equation}
811: \label{eq:roz}
812: r(z)=\sum_{i=1}^4\left[\dfrac{a_i}{(z-z_i)^2}+\dfrac{b_i}{z-z_i}\right],
813: \end{equation}
814: %
815: with coefficients
816: %
817: \begin{gather*}
818: a_2 = a_1^* = F -\frac{1}{4}+ \rmi d , \qquad a_3=a_4= \frac{5}{16}, \\
819: b_1 = b_2^* = d +\rmi \left( F -\frac{1}{4}\right), \\
820: b_{3,4} = \mp \frac{ 5}{16} -d,\qquad F = f + \frac{c}{2} - \frac{7}{16}.
821: \end{gather*}
822: %
823:
824: %
825: The differences of exponents $\Delta_i=\sqrt{1+4a_i}$ at
826: singular point $z_i$ are the following
827: %
828: \begin{equation}
829: \label{eq:Deltai}
830: \Delta_1 = \Delta_2^* = \sqrt{F -\rmi d}, \qquad
831: \Delta_3=\Delta_4 = \frac{3}{2}.
832: \end{equation}
833: %
834:
835: \subsection{Case $L_3=0$}
836: %
837: Let us assume that $L_3=0$. Then, obviously $d=0$, and, as we already
838: mentioned, the special frame coincides with the principal axes frame.
839: As we consider a symmetric body with $A=B=1$, then additionally we
840: have $f=0$, and
841: %
842: \[
843: F = \frac{c}{2} -\frac{7}{16}, \qquad c =\frac{1}{C}.
844: \]
845: %
846: Thanks to that equation~\eqref{eq:zred} has a simpler form and
847: it can be transformed to a Riemann $P$ equation. Instead of
848: making a direct transformation in~\eqref{eq:zred} it is more convenient
849: to start from equation~\eqref{eq:rnve}. Then, instead of
850: transformation~\eqref{eq:tran} we make the following one
851: %
852: \begin{equation}
853: \label{eq:tran1}
854: t \mapsto z:=N_1(t,k)^2,
855: \end{equation}
856: %
857: and we obtain
858: %
859: \begin{equation}
860: \label{eq:rnveP}
861: \Dtt m + \frac{1}{2}\left( \frac{1}{z-1} + \frac{1}{2z} \right) \Dt m +
862: \frac{1-c}{8} \left(\frac{1}{z-1} -\frac{1}{z} \right) m =0.
863: \end{equation}
864: %
865: For this Riemann $P$ equation the difference of exponents at $z=0$ is
866: $3/4$, at $z=1$ is $1/2$, and at $z=\infty$ is
867: %
868: \begin{equation}
869: \label{eq:delinf}
870: \Delta_\infty = \frac{1}{4}\sqrt{8c-7}.
871: \end{equation}
872: %
873: Let us notice that as in the case $d\neq 0$
874: transformation~\eqref{eq:tran1} is a branched covering of Riemann
875: sphere $\C\bbP^1\rightarrow \Gamma_k$, and the branching points of
876: this covering are precisely the three points where
877: equation~\eqref{eq:rnveP} has singularities.
878: %
879: \subsection{Case $L_3=0$. Second particular solution.}
880: %
881: When $L_3=0$ we have at our disposal another family of particular solutions.
882: Under our assumption $A=B=1$ and $\bL=[1,0,0]^T$, the following manifold
883: %
884: \begin{equation*}
885: \cN_1:= \{ ( \bM,\bN)\in \C^6\ | \ M_1=M_2=N_3 = 0, N_1^2+N_2^2=1\}\subset M_0,
886: \end{equation*}
887: %
888: is invariant with respect to the flow of system~\eqref{eq:ep}.
889: Similarly as $\cN$, manifold $\cN_1$ is diffeomorphic to
890: $T\,\bbS^1_{\C}\subset T\,\bbS^2_{\C}$ and it is a symplectic
891: sub-manifold of $M_0$. The Euler-Poisson equations restricted to
892: $\cN_1$, read
893: %
894: \begin{equation}
895: \label{eq:rep1}
896: \Dt M_3 =- N_2, \qquad
897: \Dt N_1= c M_3N_2, \qquad
898: \Dt N_2=-c M_3N_1.
899: \end{equation}
900: %
901: We consider a family of phase curves $k\mapsto \Gamma^1_k$ of the above
902: equations given by
903: %
904: \begin{equation}
905: \label{eq:lev1}
906: \frac{1}{2}cM_3^2 + N_1 = e, \qquad N_1^2 + N_2^2 = 1,
907: \end{equation}
908: %
909: where $e=2k^2-1$. For $k\in (0,1)$ curves $\Gamma_k^1$ are
910: non-degenerate elliptic curves. Variational equations along phase
911: curve $\Gamma_k^1$ have the following form
912: %
913: %
914: \begin{equation*}
915: \label{eq:var1}
916: \Dt
917: \begin{bmatrix}
918: m_1\\ m_2 \\m_3 \\ n_1 \\ n_2 \\ n_3
919: \end{bmatrix}
920: =
921: \begin{bmatrix}
922: 0 & (c-1)M_3 &0 & 0 & 0 & 0\\
923: (1-c)M_3 & 0 & 0 & 0 & 0 & 1 \\
924: 0 & 0 & 0 & 0 & -1 & 0 \\
925: 0 & 0 & cN_2 & 0 & cM_3 & 0 \\
926: 0 & 0 & -cN_1& -cM_3& 0 & 0 \\
927: -N_2 & N_1 & 0 & 0 & 0 & 0
928: \end{bmatrix}
929: \begin{bmatrix}
930: m_1\\ m_2 \\m_3 \\ n_1 \\ n_2 \\ n_3
931: \end{bmatrix},
932: \end{equation*}
933: %
934: and they have the following first integrals
935: %
936: \begin{equation*}
937: \label{eq:ivar1}
938: \begin{split}
939: h &= cM_3 m_3 + n_1 , \\
940: h_1& = N_1m_1 + N_2m_2 + M_3n_3, \\
941: h_2 &= N_1n_1 + N_2 n_2.
942: \end{split}
943: \end{equation*}
944: %
945: The normal variational equations are given by
946: %
947: \begin{equation*}
948: \label{eq:nve1}
949: \begin{split}
950: \Dt m_1 &= (c-1)M_3 m_2, \\
951: \Dt m_2 &= (1-c)M_3 m_1 + n_3, \\
952: \Dt n_3 &= -N_2 m_1 +N_1 m_2 , \\
953: 0&= N_1m_1 + N_2m_2 + M_3n_3.
954: \end{split}
955: \end{equation*}
956: %
957: Now, the reduction of the above system to the second order equation gives
958: %
959: \begin{equation}
960: \label{eq:nves1}
961: \ddot n + (M_3^2 - N_1) n =0, \qquad n\equiv n_3.
962: \end{equation}
963: %
964: Let us notice that from equations~\eqref{eq:rep1} and \eqref{eq:lev1}
965: it follows that
966: %
967: \begin{equation*}
968: \label{eq:dN1}
969: {\dot N_1}^2 = 2c(e-N_1)(1-N_1^2).
970: \end{equation*}
971: %
972: Thus, putting
973: %
974: \[
975: N_1=\frac{2}{c} v + \frac{e}{3},
976: \]
977: %
978: we obtain the following equation
979: %
980: \begin{equation*}
981: \label{eq:vwp}
982: {\dot v}^2 = 4v^3 -g_2 v - g_3,
983: \end{equation*}
984: %
985: determining the Weierstrass function $v(t)=\wp(t;g_2,g_3)$ with invariants
986: %
987: \begin{equation*}
988: \label{eq:g2g3}
989: g_2 = \frac{1}{3}c^2(e^2+3), \qquad g_3 = \frac{1}{27}c^3 e (e^2 -9) .
990: \end{equation*}
991: %
992: Hence, we can express $N_1$, and $M_3^2$ (using \eqref{eq:lev1}), in
993: terms of the Weierstrass function $\wp(t;g_2,g_3)$. The discriminant and
994: the modular function of $\wp(t;g_2,g_3)$ are following
995: %
996: \begin{equation*}
997: \label{eq:dedi}
998: \begin{split}
999: &\Delta = g_2^3 -27g_3^2 = c^6(e^2-1)^2, \\
1000: &j(g_2,g_3) = j(e)=\frac{g_2^3}{g_2^3 - 27g_3^2}=\frac{(e^2+3)^3}{27(e^2-1)^2}.
1001: \end{split}
1002: \end{equation*}
1003: %
1004: Hence, we can rewrite equation~\eqref{eq:nves1} in the form
1005: of the Lam\'e equation
1006: %
1007: \begin{equation}
1008: \label{eq:nve1lam}
1009: \Dtt n = (\alpha \wp(t;g_2,g_3) + \beta) n ,
1010: \end{equation}
1011: %
1012: where
1013: %
1014: \[
1015: \alpha = 2C(2C+1), \qquad \beta = \frac{e}{3}C(1-4C).
1016: \]
1017: %
1018: It is important to notice here the physical restriction on parameter
1019: $C$, namely, we have $C\in(0,2)$.
1020:
1021: %
1022: %
1023: %
1024: \section{Proof of Theorem~\ref{thm:z2}}
1025: %
1026: \label{sec:Proof2}
1027: %
1028: %
1029: In our proof of Theorem~\ref{thm:z2} we try to be as close as possible
1030: to the proof of Ziglin. Namely, first we show that a necessary condition
1031: for integrability is $\widetilde L_3=0$ (or $\widetilde L_1=0$, but
1032: this gives the already known integrable case of Lagrange). In fact this is
1033: the most difficult part of the proof. Then, we use the second family
1034: of particular solutions and we restrict the possible values of the
1035: principal moment of inertia. Finally, using the first solution, we
1036: limit all allowable values of $C$ to those corresponding to the known
1037: integrable cases.
1038:
1039: We organise the three steps of the proof in the form of three lemmas.
1040: Only the first one is somewhat involved, the remaining two are very
1041: simple.
1042:
1043: The first step is to show that a necessary condition for
1044: integrability is $d=0$, see formulae~\eqref{eq:d}.
1045:
1046:
1047: \begin{lemma}
1048: \label{lem:1}
1049: Let us assume that $d\neq 0$. Then the identity component of the
1050: differential Galois group of equation~\eqref{eq:zred} is not Abelian.
1051: \end{lemma}
1052: %
1053: \begin{proof}
1054: In our proof we use of Lemma~\ref{lem:alg} and~\ref{lem:algc1},
1055: see Appendix. If~\eqref{eq:zred} is reducible then the identity
1056: component $\cG^0$ of its differential Galois group is Abelian in two
1057: cases: when $\cG$ is a subgroup of diagonal group $\cD$ or when it
1058: is a proper subgroup of triangular group $\cT$.
1059:
1060: First we show that $\cG\not\subset\cD$. Let us assume the opposite. Then
1061: there exist two exponential solutions of~\eqref{eq:zred} which have
1062: the following form
1063: %
1064: \begin{equation*}
1065: \label{eq:expsol}
1066: w_l = P_l \prod_{i=1}^4(z-z_i)^{e_{i,l}}, \qquad P_l\in\C[z], \quad l=1,2,
1067: \end{equation*}
1068: %
1069: where $e_{i,l}$ for $l=1,2$ are exponents at singular point $z_i$, i.e.,
1070: %
1071: \[
1072: e_{i,l}\in\left\{ \frac{1}{2}(1 + \Delta_i), \frac{1}{2}(1 -\Delta_i)\right\}.
1073: \]
1074: %
1075: Here $\Delta_i$ for $i=1,\ldots, 4$ are given by \eqref{eq:Deltai}.
1076: The product of these solutions $v=w_1w_2$ belongs to $\C(z)$ and it is
1077: a solution of the second symmetric power of~\eqref{eq:zred}, i.e.
1078: equation~\eqref{eq:ssp} with $r$ given by~\eqref{eq:roz}. This
1079: equation has the same singular points as equation~\eqref{eq:zred}.
1080: Exponents $\rho_{i,l}$ at singular points $z_i$, and at infinity
1081: $\rho_{\infty,l}$ for the second symmetric power of~\eqref{eq:zred}
1082: are given by
1083: %
1084: \[
1085: \rho_{i,l}\in \{1, 1\pm\Delta_i\}, \qquad
1086: \rho_{\infty,l}\in\{-2,-1,0\}, \qquad l=1,2,3,
1087: \]
1088: %
1089: where $\Delta_i$ for $i=1,2,3,4$ are given by~\eqref{eq:Deltai}.
1090: If we write $v = P/Q$ with $P,Q \in \C[z]$ then
1091: \[
1092: Q = \prod_{i=1}^K(z-r_i)^{n_i}, \qquad n_i\in\N, \qquad
1093: r_i\in\{z_1,z_2,z_3,z_4\},
1094: \]
1095: and $ n_i= -\rho_{i,l}\in\N$ for certain $l$. However, if $d\neq0$,
1096: then $\rho_{i,l}$ is not a negative integer for $i=1,2,3,4$ and
1097: $l=1,2,3$. This implies that $Q=1$. Hence, equation~\eqref{eq:ssp} has
1098: a polynomial solution $v=P$, and $\deg P = -\rho_{\infty,l}\leq 2$.
1099: But $v$ is a product of two exponential solutions of the form
1100: ~\eqref{eq:expsol}, so we also have
1101: %
1102: \begin{equation}
1103: \label{eq:exprod}
1104: v = P_1P_2 \prod_{i=1}^4(z-z_i)^{e_{i,m}+e_{i,l}}\in\C[z],
1105: \qquad m,l\in\{1,2\}.
1106: \end{equation}
1107: %
1108: Consequently, ${e_{i,m}+e_{i,l}}$ is a non-negative integer for
1109: $i=1,2,3,4$. As for $d\neq0$, we have $2e_{i,l}\not\in\Z$ for
1110: $i=1,2,3,4$ and $l=1,2$, we deduce that in~\eqref{eq:exprod} $m\neq
1111: l$. But $e_{i,1}+e_{i,2}=1$, for $i=1,2,3,4$.
1112: Thus, we have
1113: \[
1114: \deg v = \deg P =
1115: \deg(P_1P_2)+4\geq 4.
1116: \]
1117: We have a contradiction because we already showed that $\deg P \leq 2$.
1118:
1119: It is also impossible that $\cG$ conjugates to $\cT_m$ for a certain
1120: $m\in\N$ because when $d\neq 0$ exponents for $z_1$ and $z_2$ are not
1121: rational. This implies also that $\cG$ is not finite.
1122:
1123: The last possibility that $\cG^0$ is Abelian occurs when $\cG$ is
1124: conjugated with a subgroup of $D^\dag$. We show that it is impossible.
1125: To this end, we apply the second case of the Kovacic algorithm, see
1126: Appendix. The auxiliary sets for singular points are following
1127: %
1128: \[
1129: E_1=E_2=\{2\}, \qquad E_3=E_4 = \{-1,2,5\}, \qquad E_\infty=\{0,2,4\}.
1130: \]
1131: %
1132: In the Cartesian product $E=E_\infty\times E_1 \times \cdots \times E_4$
1133: we look for such elements $e$ for which
1134: %
1135: \[
1136: d(e):=\frac{1}{2}\left( e_\infty - \sum_{i=1}^4 e_i\right),
1137: \]
1138: %
1139: is a non-negative integer. There are two such elements, namely
1140: \[
1141: e^{(1)}=(4,2,2,-1,-1),\qquad e^{(2)}=(2,2,2,-1,-1).
1142: \]
1143: %
1144: We have $d(e^{(1)})=1$ and $d(e^{(2)})=0$. We have to check if there
1145: exists polynomial $P = p_1 z + p_0$ which satisfies the following
1146: equation
1147: %
1148: \[
1149: P''' + 3\theta P'' +(3 \theta^2 + 3\theta' -4r)P' +
1150: (\theta'' + 3 \theta\theta' + \theta^3 -4 r\theta - 2r')P =0,
1151: \]
1152: %
1153: where
1154: %
1155: \[
1156: \theta = \frac{1}{z-\rmi} + \frac{1}{z+\rmi} -\frac{1}{2}\frac{1}{z-1}-
1157: \frac{1}{2}\frac{1}{z+1}.
1158: \]
1159: %
1160: Inserting $P$ into the above equation we obtain the following system of
1161: linear equations for its coefficients
1162: %
1163: \begin{gather*}
1164: - d p_0 + F p_1 = 0 , \qquad (1-2F)p_0 + 6 d p_1 =0 \\
1165: F p_0 + d p_1 = 0 , \qquad -6 d p_0 + (1-2F) p_1 =0.
1166: \end{gather*}
1167: %
1168: The above system for $p_0$ and $p_1$ has a non-zero solution if $d^2 +
1169: F^2=0$, but for a real $d$ and $F$ it is possible only when $d=0$ and
1170: $F=0$.
1171: \end{proof}
1172: %
1173: As the covering $t\mapsto z$ given by~\eqref{eq:tran} does not change
1174: the identity component of the differential Galois group of the normal
1175: variational equations~\eqref{eq:nve2}, from the above lemma it follows
1176: that if the Euler-Poisson equations are integrable, then
1177: %
1178: \[
1179: 2 d = \left( \frac{1}{C}-\frac{1}{A} \right){\widetilde L}_1 {\widetilde L}_3 =0.
1180: \]
1181: %
1182: For a symmetric body when $A=B\neq C$, the above condition implies
1183: that either ${\widetilde L}_1 = 0$, and this corresponds to the
1184: integrable case of Lagrange, or ${\widetilde L}_3=0$. Hence, we have
1185: to investigate the last case. Notice that in this case the special
1186: frame is the principal axes frame so we have ${\widetilde
1187: \bL}=\bL=[1,0,0]^T$, $a=1/A=1/B=b=1$ and $c=1/C\neq 1$. At this
1188: point it is worth to observe that now the identity component of the
1189: differential Galois group of equation~\eqref{eq:zred} is Abelian for
1190: infinitely many values of $c$. Let us remind here that from the
1191: physical restriction it follows only that $c\in (1/2,\infty )$.
1192: %
1193: \begin{proposition}
1194: \label{prop:1}
1195: If $L_3=0$ then the identity component of the
1196: differential Galois group of equation~\eqref{eq:zred} is Abelian in the
1197: following cases:
1198: %
1199: \begin{gather}
1200: c = 1 + 2l(4l-1), \quad c= 4 + 2l(4l+5), \quad
1201: c = 2 +2l(4l+3,)\\
1202: c= \frac{11}{8} +2l(l+1), \qquad c = \frac{79}{72} + \frac{4}{3}l + 2l^2,
1203: \end{gather}
1204: %
1205: where $l$ is an integer.
1206: \end{proposition}
1207: \begin{proof}
1208: When $L_3=0$ then the identity component of the differential Galois
1209: group of equation~\eqref{eq:zred} is Abelian if and only if the
1210: differential Galois group of equation~\eqref{eq:rnveP} is Abelian
1211: (transformation from \eqref{eq:zred} to \eqref{eq:rnveP} is
1212: algebraic). Then applying Kimura Theorem~\ref{thm:kimura} to
1213: equation~\eqref{eq:rnveP} we easily derive the above values of $c$ for
1214: which the identity component of the differential Galois
1215: group of this equation is Abelian.
1216: \end{proof}
1217: %
1218: Thus, applying the Morales-Ramis or Ziglin theory and using the first
1219: particular solution we cannot prove non-integrability of the
1220: Euler-Poison equations for all values of $c$ listed in the above
1221: proposition. This is why, in the lemma below, we consider normal
1222: variational equations corresponding to the second particular solution.
1223:
1224:
1225:
1226: %
1227: \begin{lemma}
1228: \label{lem:2}
1229: Assume that $C\in(0,2)$ and $C\neq m/4$, for $m=1,\ldots, 7$, then
1230: for almost all $e\in\R$, the differential Galois group of
1231: equation~\eqref{eq:nve1lam} is \SLtwoC.
1232: \end{lemma}
1233: \begin{proof}
1234: We assume first that $e\neq\pm 1$. Then the discriminant of the elliptic
1235: curve associated with $\wp(t;g_2,g_3)$ does not vanish, and we can
1236: apply Lemma~\ref{lem:lame}, see Appendix. We consider successively
1237: three cases from this lemma.
1238:
1239: For the Lam\'e-Hermite case, we have $\alpha = n(n+1)$ for $n\in\Z$.
1240: This implies that $C=n/2$, and hence, as $C\in(0,2)$ we have $C\in\{1/2,1,3/2\}$.
1241:
1242: For the Brioschi-Halphen-Crawford case, we have $\alpha = n(n+1)$ and
1243: $m=n+1/2 \in \N$. Thus, we have
1244: %
1245: \[
1246: C = -\frac{1}{4} +\frac{1}{2}m, \qquad m\in\N.
1247: \]
1248: %
1249: So, this case can occur only when $C\in\{1/4,3/4,5/4,7/4\}$.
1250:
1251: In the Baldassarri case we notice that the mapping
1252: %
1253: \[
1254: \R\backslash\{-1,1\} \ni e \mapsto j(e),
1255: \]
1256: %
1257: is non-constant and continuous. Hence, by Dwork
1258: Proposition~\ref{prop:dwork}, for a fixed $C$ this case can occur only
1259: for a finite number of values of $e$.
1260: \end{proof}
1261:
1262: As $C=1/2$, $C=1$ and $C=1/4$ correspond to the Kovalevskaya, Euler
1263: and Goryachev-Chaplygin cases, respectively, we have to investigate
1264: cases
1265: %
1266: \[
1267: C\in\{3/4,5/4,3/2,7/4\}.
1268: \]
1269: %
1270: To this end we return to equation~\eqref{eq:zred}. As we show, for
1271: $d=0$ it can be transformed to the form~\eqref{eq:rnveP} and, moreover,
1272: this transformation does not change the identity component of its
1273: differential Galois group. We can prove the following.
1274: %
1275: \begin{lemma}
1276: \label{lem:3}
1277: For $C\in\{3/4,5/4,3/2,7/4\}$ the differential Galois group
1278: of~\eqref{eq:rnveP} is \SLtwoC.
1279: \end{lemma}
1280: \begin{proof}
1281: For $C\in\{3/4,5/4,3/2,7/4\}$ the respective values of the difference of
1282: exponents at infinity $\Delta_\infty$ (see
1283: formula~\eqref{eq:delinf}) for equation~\eqref{eq:rnveP} are
1284: following
1285: \[
1286: \frac{1}{4}\left\{ \sqrt{\frac{11}{3}}, \rmi \sqrt{\frac{3}{5}},
1287: \rmi \sqrt{\frac{5}{3}}, \rmi \sqrt{\frac{17}{7}}\right\}.
1288: \]
1289: Now, a direct inspection of possibilities in the Kimura
1290: Theorem~\ref{thm:kimura} shows that Riemann $P$
1291: equation~\eqref{eq:rnveP} with prescribed differences of exponents does
1292: not possess a Liouvillian solution, so its differential Galois group
1293: is \SLtwoC.
1294: \end{proof}
1295: %
1296: Now the proof of Theorem~\eqref{thm:z2} is a simple consequence of the
1297: above three lemmas.
1298: %
1299: %
1300: %
1301: \section{Proof of Theorem~\ref{thm:z3}}
1302: %
1303: \label{sec:Proof3}
1304: %
1305: %
1306: In the proof of Theorem~\ref{thm:z3} we apply the Ziglin
1307: Lemma~\ref{lem:zig} and his idea of its application given in
1308: \cite{Ziglin:97::}.
1309:
1310: The Euler-Poisson equations restricted to $\cN$ possess a
1311: hyperbolic equilibrium at $u=(0,0,0,1,0,0)$. The phase curve
1312: $\Gamma_1$ corresponds to the solution of equations~\eqref{eq:rep} with
1313: $k=1$. It contains two real components which are real phase curves
1314: corresponding to real solutions homoclinic to $u$. Their union is $\Re
1315: \Gamma_1$ and we denote its closure by $\Omega$.
1316: %
1317: \begin{lemma}
1318: \label{lem:eps}
1319: For an arbitrary complex neighbourhood $U\subset\cN$ of $\Omega$ there
1320: exists $\epsilon>0$, such that for $0<1-k<\epsilon$ the fundamental
1321: group $\pi_1(\Gamma_k)$ of phase curve $\Gamma_k$ is generated by
1322: loops lying in $U$.
1323: \end{lemma}
1324: \begin{proof}
1325: The time parametrisation
1326: of $\Gamma_k$ is given by
1327: %
1328: \begin{equation}
1329: \label{eq:exfik}
1330: \begin{split}
1331: M_2(t,k)&:=-2k\cn( t,k), \\
1332: N_1(t,k)&:= 2k^2\sn^2( t,k)-1,\\
1333: N_3(t,k)&:= 2k\sn( t,k)\dn( t,k),
1334: \end{split}
1335: \end{equation}
1336: %
1337: where $\sn(t,k)$, $\cn(t,k)$ and $\dn(t,k)$ denote the
1338: Jacobi elliptic functions of argument $t$ and modulus $k$.
1339: Thus, particular solutions of~\eqref{eq:ep}
1340: %
1341: \begin{equation}
1342: \label{eq:phitk}
1343: \varphi(t,k):= (0,M_2(t,k),0,N_1(t,k),0,N_3(t,k)),
1344: \end{equation}
1345: %
1346: defined by~\eqref{eq:exfik} are single-valued, meromorphic, and double
1347: periodic with periods
1348: %
1349: \begin{equation}
1350: \label{eq:Tik}
1351: T_1(k)= 2K(k)+2\rmi K'(k),\qquad T_2(k)= 2K(k)-2\rmi K'(k),
1352: \end{equation}
1353: %
1354: where $K(k)$ is the complete elliptic integral of the first kind with
1355: modulus $k$, $K'(k):=K(k')$, and $k':=\sqrt{1-k^2}$. In each period
1356: cell they have two simple poles at:
1357: %
1358: \begin{equation*}
1359: t_1(k)= \rmi K'(k), \quad t_2(k)= -\rmi K'(k) \mod (T_1(k),T_2(k)).
1360: \end{equation*}
1361: %
1362: Periods $T_1(k)$ and $T_2(k)$ given by~\eqref{eq:Tik} of
1363: solution~\eqref{eq:phitk} are primitive. Minimal real and imaginary
1364: periods are $T(k)=4K(k)$ and $T'(k)=4\rmi K'(k)$. As a base point
1365: $x(k)\in\Gamma_k$ we choose $x(k)=\varphi(t_0(k),k)$ where
1366: $t_0(k)=K(k)$. Let us notice that from \eqref{eq:exfik} it follows
1367: that
1368: %
1369: \begin{equation}
1370: \label{eq:xk}
1371: \begin{split}
1372: M_2(t_0(k),k) &=0, \\
1373: N_1(t_0(k),k)&=2k^2-1, \\
1374: N_3(t_0(k),k)&=2kk'.
1375: \end{split}
1376: \end{equation}
1377: Now, let
1378: %
1379: \[
1380: \lambda_k, \lambda_k':[0,1]\rightarrow\Gamma_k,
1381: \]
1382: %
1383: be the loops with base point $x(k)$ corresponding to periods $T(k)$
1384: and $T'(k)$, respectively. These loops cross at point
1385: %
1386: \begin{equation*}
1387: \label{eq:xkp}
1388: x'(k)=\varphi(t_0(k)+T(k)/2,k)= \varphi(t_0(k)+T'(k)/2,k ).
1389: \end{equation*}
1390: %
1391: As a results, we obtain four semi-loops with end points $x(k)$ and
1392: $x'(k)$. The fundamental group $\pi_1(\Gamma_k,x(k))$ of $\Gamma_k$ is
1393: generated by these semi-loops, see Figure~\ref{fig:loops}.
1394: %
1395: \begin{figure}[h]
1396: \centering \includegraphics[scale=0.8]{sfig.ps}
1397: \caption{\label{fig:loops}Parallelogram of period with marked loops}
1398: \end{figure}
1399: %
1400: Let us analyse what happens when $k$ tends to 1. From \eqref{eq:xk}
1401: it follows that $x(k)$ tends to $u$ and from \eqref{eq:exfik} we
1402: deduce that loop $\lambda_k$ tends to $\Omega$. To see what happens
1403: with loop $\lambda'_k$ when $k$ tends to 1, let us put $t=t_0(k)+\rmi
1404: \tau$ in formulae \eqref{eq:exfik}. We obtain
1405: %
1406: \begin{equation*}
1407: \begin{split}
1408: M_2(t,k)&= 2\rmi kk'\frac{\sn(\tau,k')}{\dn(\tau,k')},\\
1409: N_1(t,k)&= -1+\frac{2k^2}{\dn^{2}(\tau,k')} \\
1410: N_3(t,k)&= 2kk'\frac{\cn(\tau,k')}{\dn^{2}(\tau,k')}.
1411: \end{split}
1412: \end{equation*}
1413: %
1414: Thus, loop $\lambda'_k$ tends to point $u$ as $k$ tends to 1.
1415: \end{proof}
1416:
1417: Time parametrisation of these
1418: phase curves $\Gamma_k^1$ is given by
1419: %
1420: \begin{equation}
1421: \label{eq:sol2t}
1422: \begin{split}
1423: M_3(t,k)&=\frac{2k}{\omega}\cn(\omega t,k),\\
1424: N_1(t,k)&=2k^2\sn^2(\omega t,k)-1, \\
1425: N_2(t,k)&=2k\sn(\omega t,k)\dn(\omega t,k),
1426: \end{split}
1427: \end{equation}
1428: %
1429: where $\omega^2 = c$. Thus, the second family of particular solutions
1430: of~\eqref{eq:ep}
1431: %
1432: \begin{equation*}
1433: \label{eq:phi1tk}
1434: \varphi_1(t,k):= (0,0,M_3(t,k),N_1(t,k),N_2(t,k), 0),
1435: \end{equation*}
1436: %
1437: defined by~\eqref{eq:sol2t} contains solutions which are
1438: single-valued, meromorphic, and double periodic with periods
1439: %
1440: \begin{equation*}
1441: \label{eq:Tik1}
1442: T_1(k)= 2\frac{K(k)}{\omega}+2\rmi \frac{K'(k)}{\omega},\qquad
1443: T_2(k)= 2\frac{K(k)}{\omega}-2\rmi \frac{K'(k)}{\omega}.
1444: \end{equation*}
1445: %
1446: In each period cell they have two simple poles at:
1447: %
1448: \begin{equation*}
1449: t_1(k)= \rmi \frac{K'(k)}{\omega}, \quad
1450: t_2(k)= -\rmi \frac{K'(k)}{\omega} \mod (T_1(k),T_2(k)).
1451: \end{equation*}
1452: %
1453: Let us notice that for the Euler-Poisson equations restricted to
1454: $\cN_1$ point $u=(0,0,0,1,0,0)$ is also a hyperbolic
1455: equilibrium. The phase curve $\Gamma_1^1$ corresponds to the solution of
1456: equations~\eqref{eq:rep1} with $k=1$. As in the previous case, it
1457: contains two real components which are real phase curves corresponding
1458: to real solutions homoclinic to $u$. Their union is $\Re \Gamma_1^1$
1459: and we denote its closure by $\Omega_1$. Using the same arguments as
1460: in the proof of Lemma~\ref{lem:eps} we can show the following.
1461: %
1462: \begin{lemma}
1463: \label{lem:eps1}
1464: For an arbitrary complex neighbourhood $U\subset\cN_1$ of $\Omega_1$ there
1465: exists $\epsilon>0$, such that for $0<1-k<\epsilon$ the fundamental
1466: group $\pi_1(\Gamma_k^1)$ of phase curve $\Gamma_k$ is generated by
1467: loops lying in $U$.
1468: \end{lemma}
1469: %
1470: Now, to prove Theorem~\ref{thm:z3} let us notice that we showed that,
1471: except for the known integrable cases, the identity component of the
1472: differential Galois group of the normal variational equations
1473: corresponding to $\Gamma_k$ or $\Gamma^1_k$ is not Abelian for almost
1474: all values of $k\in(0,1)$. In fact, in Lemma~\ref{lem:1} we proved
1475: that for $d\neq 0$ the identity component of the differential Galois
1476: group of the normal variation equations corresponding to $\Gamma_k$
1477: with $k=1/\sqrt{2}$ is not Abelian. By Lemma~\ref{lem:par} it is not
1478: Abelian for almost all values of $k\in (0,1)$. Then,
1479: in~Lemma~\ref{lem:3} we proved that the identity component of the
1480: differential Galois group of the normal variation equations
1481: corresponding to $\Gamma_k^1$ is not Abelian for almost all values of
1482: $k\in (0,1)$, except for $C=m/4$, $m=1,\ldots, 7$. Finally, for
1483: $C=m/4$ such that $C\not\in\{1/4, 1/2,1\}$, we showed that the identity
1484: component of the differential Galois group of the normal variation
1485: equations corresponding to $\Gamma_k$ with $k=1/\sqrt{2}$ is not
1486: Abelian. Again, by Lemma~\ref{lem:par} it is not Abelian for almost
1487: all values of $k\in (0,1)$.
1488:
1489: Both normal variational equations corresponding to $\Gamma_k$ and
1490: $\Gamma_k^1$ are Fuchsian. For a Fuchsian equation we know that if
1491: the identity component of its differential Galois group is not
1492: Abelian then its monodromy group does not possess a rational
1493: invariant, see Theorem~3.17 in \cite{Churchill:96::b}.
1494:
1495: Assume now that for $C\not\in\{1/4,1/2,1\}$ the Euler-Poisson
1496: equations possess an additional real meromorphic first integral
1497: defined in a real neighbourhood of $ \Omega\cup\Omega_1$. Then we
1498: can extend this integral to a complex meromorphic one, defined in a
1499: certain complex neighbourhood $U$ of $ \Omega\cup\Omega_1$. Then, by
1500: Lemma~\ref{lem:eps} and \ref{lem:eps1}, we find such $\epsilon>0$
1501: that the fundamental groups of $\Gamma_k$ and $\Gamma^1_k$ with
1502: $0<1-k<\epsilon$, are generated by loops lying entirely in $U$. Then,
1503: from the Ziglin Lemma~\ref{lem:zig}, it follows that both monodromy groups
1504: of normal variational equations corresponding to $\Gamma_k$ and
1505: $\Gamma^1_k$ possess a rational invariant. However, above we showed that
1506: at least for one of them it is not true. A contradiction proves
1507: Theorem~\ref{thm:z3}.
1508:
1509: \section{Remarks and Comments}
1510:
1511: One important difference between the Ziglin and Morales-Ramis theory
1512: is related with the procedure of obtaining the normal variational
1513: equations. Assume that system~\eqref{eq:ds} possesses certain number
1514: of known first integrals $H_i$, such that their differentials $\rmd
1515: H_i$, are linearly independent on $\Gamma$. Then $\rmd H_i\circ
1516: \pi^{-1}$ for $i=1,\ldots,k$ are independent first integrals of
1517: \eqref{eq:gnve}.Their common level
1518: %
1519: \begin{equation*}
1520: N_p:= \{ \eta\in F\,|\, \rmd H_i\circ \pi^{-1}\eta = p_i,\quad p_i\in \C,
1521: \quad i=1,\ldots,k\},
1522: \end{equation*}
1523: %
1524: defines a $m$-dimensional linear bundle over $\Gamma$, where
1525: $m=n-k-1$. Using these integrals we can reduce the order of system
1526: \eqref{eq:gnve}. Namely, we consider the reduced normal variational
1527: equations
1528: %
1529: \begin{equation}
1530: \label{eq:rgnve}
1531: \dot \eta = \pi_\star(T(v)\pi^{-1}\eta), \qquad \eta\in N_p.
1532: \end{equation}
1533: %
1534: %
1535: However after this reduction defined by Ziglin, instead of a linear,
1536: we have an affine bundle over $\Gamma$, equations \eqref{eq:rgnve} are
1537: generally not homogeneous ones, and the monodromy group is a subgroup
1538: of affine transformations of $\C^m$. Till now this construction has
1539: not been translated to the Morales-Ramis theory where we work with a
1540: system of homogeneous equations defined on $N_0$. To realise the
1541: importance of Ziglin reduction let us notice that using only the
1542: reduced equation on $N_0$ it is impossible to prove global
1543: non-integrability of the Goryachev-Chaplygin case. In
1544: \cite{Ziglin:97::} he gave such a non-integrability proof
1545: investigating the reduced normal variational equations on $N_p$ with
1546: $p\neq 0$.
1547:
1548:
1549: The Morales-Ramis theory is coordinate independent, however,
1550: investigating a specific problem, we always have to choose appropriate
1551: coordinates. The form of normal variational equations depends on local
1552: coordinates and this is why their choice is important. It is
1553: especially evident when we investigate a problem connected with a
1554: rigid body. Equations of motion of the heavy top can be written in
1555: many different forms. As we mentioned in Remark~\ref{rem:r1}, the
1556: natural phase space for a rigid body with a fixed point is
1557: $T^*\mathrm{SO}(3,\R)$. There are no `natural' coordinates on
1558: $\mathrm{SO}(3,\R)$, and thus there are no `natural' canonical
1559: coordinates. The most widely used are Androyer-Deprit canonical
1560: coordinates~\cite{Borisov:01::} or the Euler angles and conjugated
1561: momenta. In fact we checked which, from almost all known coordinates
1562: on $T^*\mathrm{SO}(3,\R)$, are most feasible for application of the
1563: Morales-Ramis theory. In our exposition we
1564: work with the Euler-Poisson equations. However, our choice of the
1565: body fixed frame is not conventional. Usually the principal axes
1566: frame is used. To see what is an advantage of our choice, let us
1567: notice that using the principal axes frame we can derive the normal
1568: variational equation in the form similar to~\eqref{eq:rnve}, however,
1569: to put it in the form of an equation with rational coefficient we have
1570: to choose a transformation different than~\eqref{eq:tran}, and as a
1571: result, we obtain, instead of equation~\eqref{eq:zred} possessing four
1572: regular singularities, a much more complicated Fuchsian equation with
1573: seven singular points. Our choice of the body fixed frame appears e.g.
1574: in \cite{Dokshevich:92::}.
1575:
1576: Simplifications of the normal variational equations which occur when
1577: $L_3=0$ need an explanation. In fact, one can observe that although
1578: the Riemann surface $\Gamma_k^1$ for the second particular solution is
1579: a torus with two points removed, see formulae~\eqref{eq:sol2t}, the
1580: normal variational equation~\eqref{eq:nve1lam} corresponding to it
1581: has the form of a Lam\'e equation, so it is defined over a torus with
1582: one point removed. The reason of what happened is symmetry. When
1583: $L_3=0$ the Euler-Poisson equations restricted to $M_0$ are invariant
1584: with respect to an involutive symplectic diffeomorphism $\cJ_1:
1585: M_0\mapsto M_0$ defined by
1586: %
1587: \[
1588: \cJ_1(M_1,M_2,M_3, N_1,N_2,N_3) = (-M_1,M_2,-M_3, N_1,-N_2,N_3).
1589: \]
1590: %
1591: Let us denote
1592: %
1593: \[
1594: M=\{ p\in M_0\,|\, \cJ_1(p)\neq p\}, \qquad \widehat M=
1595: M/\cJ_1,
1596: \]
1597: %
1598: and let $\pi:M\mapsto \widehat M$ be the projection. In the natural way
1599: equations~\eqref{eq:ep} induce Hamiltonian equations on $\widehat M$
1600: with Hamiltonian function $\widehat H= H\circ\pi^{-1}$. Then,
1601: according to Ziglin, see Lemma on page 36 in \cite{Ziglin:82::b}, if
1602: system~\eqref{eq:ep} is integrable, then the induced Hamiltonian system
1603: on $\widehat M$ is also integrable. For the induced system we have
1604: a family of particular solutions $\widehat\varphi_1(t,k) =
1605: \pi\circ\varphi_1(t,k)$. The corresponding Riemann surfaces
1606: $\widehat\Gamma^1_k$ are tori with one point removed.
1607:
1608: A simplification of the normal variational equation for solution
1609: $\varphi(t,k)$ when $L_3=0$ and the fact that we can transform them to a
1610: Riemann $P$ equation (for an appropriate choice of energy) is also
1611: related with symmetry. Namely, when $L_3=0$, system \eqref{eq:ep}
1612: restricted to $M_0$ is also invariant with respect to an involutive
1613: symplectic diffeomorphism $\cJ: M_0\mapsto M_0$ defined by
1614: %
1615: \[
1616: \cJ(M_1,M_2,M_3, N_1,N_2,N_3) = (-M_1,-M_2,M_3, N_1,N_2,-N_3).
1617: \]
1618: %
1619: For symmetry reduction of variational equations see Section~4.2 in
1620: \cite{Churchill:96::b}.
1621:
1622: Let us note that in Lemma~\ref{lem:1} we claim that if $d\neq 0$ then
1623: the identity component of the differential Galois group of
1624: equation~\eqref{eq:zred} is not Abelian. Thus, it can be the whole
1625: group \SLtwoC\ or whole triangular subgroup $\cT$ of \SLtwoC. We do not
1626: know if the second case can occur.
1627:
1628: In the case of first particular solution we do not work with the
1629: elliptic curve $\Gamma_k$ but with the Riemann sphere (minus singular
1630: point) for which $\Gamma_k$ is a covering. The reason of this is that
1631: we have no tool similar to the Kovacic algorithm for a second order
1632: linear differential equation defined on an elliptic curve. However, in
1633: the case of the second particular solution we can work directly on
1634: elliptic curve $\Gamma^1_k$ because, in this case, the normal
1635: variational equation is the Lam\'e equation for which the monodromy
1636: group is know. Of course, in this case we can also work on the Riemann
1637: sphere making well know transformation of the Lam\'e equation to its
1638: algebraic form.
1639:
1640: In his proof of Theorem~\ref{thm:z2} and \ref{thm:z3} Ziglin used the
1641: explicit time parametrisation of particular solutions. First he showed
1642: that if the system is integrable then the monodromy of the normal
1643: variational equations along real periods of a particular solution must
1644: be equal to the identity. Then, using analytical tools he derived the
1645: necessary conditions for the integrability.
1646:
1647: In our exposition we use the explicit time parametrisation of
1648: particular solutions in the proof of Theorem~\ref{thm:z3}. In fact, we
1649: use it only to show explicitly what happens with loops along real and
1650: imaginary periods when $k$ tends to 1. However, one can deduce this
1651: information from the equations defining the elliptic curve. Thus, we
1652: can avoid using explicit time parametrisation at all. We keep it in
1653: the proof of Theorem~\ref{thm:z3} because, as we hope, it makes the
1654: exposition more transparent.
1655:
1656: The physical restriction $C\in(0,2)$ plays crucial role in our, as
1657: well as, in the Ziglin proof. Integrable systems are really rare, hence it
1658: is an interesting question if the Euler-Poisson equations are
1659: integrable for values of parameters which do not satisfy this restriction.
1660:
1661:
1662:
1663: Considering the case $L_3=0$ and the first particular solution Ziglin
1664: showed that the necessary condition for integrability is $c\in\N$. In
1665: our Proposition~\ref{prop:1} there are two families of $c$ such that
1666: $c\in\Q$. The reason why they appears is that we fixed the energy for
1667: the first solution.
1668:
1669:
1670:
1671: In the proof of Lemma~\ref{lem:2} we show
1672: that the Brioschi-Halphen-Crawford case is possible only when
1673: $C\in\{1/4,3/4,5/4,7/4\}$. For this values of $C$ we can calculate the
1674: Brioschi determinant $Q_m(g_2,g_3,\beta)$ defined by~\eqref{eq:bdet}.
1675: Calculations show that it vanishes identically only when $C=1/4$,
1676: i.e., for the Goryachev-Chaplygin case. Thus, in fact, to prove that
1677: for $C\in\{3/4,5/4,7/4\}$ the Euler-Poisson equations are
1678: non-integrable, we can use the second solution. We use the first one
1679: because calculations are simpler.
1680:
1681: \section{Acknowledgements}
1682: %
1683: We are very thankful to Mich\`ele Audin for her remarks, comments,
1684: suggestions and corrections. They allowed us to improve considerably
1685: not only the contents of the paper, but also gave a more clear proof
1686: of our main result.
1687:
1688: We would like to thank Delphine Boucher, Juan J.~Morales-Ruiz,
1689: Jacques-Arthur Weil, Carles Sim\'o, Michael F.~Singer and Felix Ulmer
1690: for discussions and help which allowed us to understand many topics
1691: related to this work. We are very grateful to Mark van Hoeij with whom
1692: we started to discuss some problems concerning this paper in the end
1693: of the previous century. Many important comments by Robert~S.~Maier
1694: concerning the Lam\'e equation are gratefully acknowledged.
1695:
1696: As usual, we thank Zbroja (Urszula Maciejewska) not only for her
1697: linguistic help.
1698:
1699: For the second author this research has been supported by a Marie
1700: Curie Fellowship of the European Community programme Human Potential
1701: under contract number HPMF-CT-2002-02031.
1702: %
1703:
1704:
1705: \section{Appendix}
1706: %
1707: \subsection{Dependence on a parameter}
1708: %
1709: %
1710: Let us consider a second order
1711: differential equation of the following form
1712: %
1713: \begin{equation}
1714: \label{eq:gsoe}
1715: y''= r(z,\varepsilon)y, \qquad '\equiv \frac{\rmd\phantom{z}}{\rmd z} ,
1716: \end{equation}
1717: %
1718: where $r(z,\varepsilon)$ is a rational function with respect to $z$ and
1719: $\varepsilon$, i.e., $r\in\C(\varepsilon)(z)=\C(z,\varepsilon)$. Here
1720: $\varepsilon$ plays the role of a parameter. For a fixed value of
1721: $\varepsilon$ we denote by $\cG^0(\varepsilon)$ the identity component of the
1722: differential Galois group of equation~\eqref{eq:gsoe}. Let $U\subset \C$ denote an open not empty connected set with compact closure. We show the
1723: following.
1724: %
1725: \begin{lemma}
1726: \label{lem:par}
1727: Assume that:
1728: \begin{enumerate}
1729: \item Equation~\eqref{eq:gsoe} is Fuchsian.
1730: \item For $\varepsilon\in U$,
1731: equation~\eqref{eq:gsoe} possesses
1732: $N$ singular points ($N$ does not depend on $\varepsilon$) for
1733: which exponents do not depend on $\varepsilon$.
1734: \item For $\varepsilon_0\in U $,
1735: $\cG^0(\varepsilon_0)$ is not solvable (is not Abelian).
1736: \end{enumerate}
1737: %
1738: Then, except finitely many values of $\varepsilon\in U$,
1739: $\cG^0(\varepsilon)$ is not solvable (is not Abelian).
1740: \end{lemma}
1741: \begin{proof}
1742: From the Kovacic algorithm in the form given
1743: in~\cite{Duval:92::,Morales:99::c} we know that, under our
1744: assumption, if $\cG^0(\varepsilon)$ is solvable, then there exists a
1745: polynomial $P$ (whose degree does not depend on $\varepsilon$) which
1746: is a solution of a linear differential equation $L(y)=0$ with
1747: coefficients in $\C(z,\varepsilon)$. The order of $L(y)=0$ does not
1748: depend on $\varepsilon$. We have a finite number of choices for the
1749: degree of $P$ and a finite number of choices of $L(y)=0$. Finding a
1750: polynomial solution of linear equation $L(y)=0$ reduces to finding a
1751: non-trivial solution of a homogeneous linear system with
1752: coefficients in $\C(\varepsilon)$. But the last problem reduces to
1753: finding common zeros of a finite number of polynomials. We know that
1754: not all of these polynomials vanish identically (otherwise
1755: $\cG^0(\varepsilon_0)$ is solvable). Thus, there is at most a finite
1756: number of values of $ \varepsilon$ for which they vanish
1757: simultaneously. Finally, let us notice that set
1758: \[
1759: \{\varepsilon\in U\,|\, \, \,
1760: \cG^0(\varepsilon)\quad\text{is Abelian} \},
1761: \]
1762: is a subset of
1763: \[
1764: \{\varepsilon\in U \,|\, \, \,
1765: \cG^0(\varepsilon)\quad\text{is solvable} \}.
1766: \]
1767: \end{proof}
1768: %
1769: %
1770: \subsection{Second order differential equations with rational coefficients}
1771: %
1772: Let us consider a second order
1773: differential equation of the following form
1774: %
1775: \begin{equation}
1776: \label{eq:gso}
1777: y''=r y, \qquad r\in\C(z), \qquad '\equiv \frac{\rmd\phantom{z}}{\rmd z} .
1778: \end{equation}
1779: %
1780: For this equation its differential Galois group $\cG$ is an algebraic
1781: subgroup of $\mathrm{SL}(2,\C)$. The following lemma describes all
1782: possible types of $\cG$ and relates these types to forms of solution
1783: of \eqref{eq:gso}, see \cite{Kovacic:86::,Morales:99::c}.
1784: %
1785: \begin{lemma}
1786: \label{lem:alg}
1787: Let $\cG$ be the differential Galois group of equation~\eqref{eq:gso}.
1788: Then one of four cases can occur.
1789: \begin{enumerate}
1790: %
1791: \item $\cG$ is reducible (it is conjugated to a subgroup of triangular
1792: group) ; in this case equation \eqref{eq:gso} has an exponential
1793: solution of the form $y=\exp\int \omega$, where $\omega\in\C(z)$,
1794: %
1795: \item $\cG$ is conjugated with a subgroup of
1796: %
1797: \[
1798: D^\dag = \left\{ \begin{bmatrix} c & 0\\
1799: 0 & c^{-1}
1800: \end{bmatrix} \; \biggl| \; c\in\C^*\right\} \cup
1801: \left\{ \begin{bmatrix} 0 & c\\
1802: c^{-1} & 0
1803: \end{bmatrix} \; \biggl| \; c\in\C^*\right\},
1804: \]
1805: %
1806: in this case equation
1807: \eqref{eq:gso} has a solution of the form $y=\exp\int \omega$, where
1808: $\omega$ is algebraic over $\C(z)$ of degree 2,
1809: %
1810: \item $\cG$ is primitive and finite; in this case all
1811: solutions of equation \eqref{eq:gso} are algebraic,
1812:
1813: \item $\cG= \mathrm{SL}(2,\C)$ and equation \eqref{eq:gso}
1814: has no Liouvillian solution.
1815: %
1816: \end{enumerate}
1817: \end{lemma}
1818: %
1819:
1820: We need a more precise characterisation of case 1 in the above
1821: lemma. It is given by the following lemma, see Lemma~4.2 in
1822: \cite{Singer:93::a}.
1823: %
1824: \begin{lemma}
1825: \label{lem:algc1}
1826: Let $\cG$ be the differential Galois group of
1827: equation~\eqref{eq:gso} and assume that $\cG$ is reducible.
1828: Then either
1829: \begin{enumerate}
1830: \item equation~\eqref{eq:gso} has a unique solution $y$ such that
1831: $y'/y\in\C(z)$, and $\cG$ is conjugate to a subgroup of the
1832: triangular group
1833: \[
1834: \cT = \left\{ \begin{bmatrix} a & b\\
1835: 0& a^{-1}
1836: \end{bmatrix} \, |\, a,b\in\C, a\neq 0\right\}.
1837: \]
1838: Moreover, $\cG$ is a proper subgroup of $\cT$ if and only if there
1839: exists $m\in\N$ such that $y^m\in\C(z)$. In this case $\cG$ is
1840: conjugate to
1841: \[
1842: \cT_m = \left\{ \begin{bmatrix} a & b\\
1843: 0& a^{-1}
1844: \end{bmatrix} \, |\, a,b\in\C, a^m=1\right\},
1845: \]
1846: where $m$ is the smallest positive integer such that $y^m\in\C(z)$, or
1847: \item equation~\eqref{eq:gso} has two linearly independent solutions
1848: $y_1$ and $y_2$ such that $y'_i/y_i\in\C(z)$, then $\cG$ is
1849: conjugate to a subgroup of
1850: \[
1851: \cD = \left\{ \begin{bmatrix} a & 0\\
1852: 0& a^{-1}
1853: \end{bmatrix} \, |\, a\in\C, a\neq 0\right\}.
1854: \]
1855: In this case, $y_1y_2\in\C(z)$. Furthermore, $\cG$ is conjugate to a
1856: proper subgroup of $\cD$ if and only if $y_1^m\in\C(z)$ for some
1857: $m\in\N$. In this case $\cG$ is a cyclic group of order $m$ where $m$
1858: is the smallest positive integer such that $y_1^m\in\C(z)$.
1859: \end{enumerate}
1860: \end{lemma}
1861: %
1862:
1863: In case 2 of the above lemma we know that $v=y_1y_2\in\C(z)$.
1864: Differentiating $v$ three times, and using the fact that $y_i$ satisfies
1865: equation \eqref{eq:gso}, we obtain
1866: %
1867: \begin{equation}
1868: \label{eq:ssp}
1869: v'''= 2r' v + 4r v'.
1870: \end{equation}
1871: %
1872: The above equation is called the second symmetric power of
1873: equation~\eqref{eq:gso}. For applications of symmetric powers of
1874: differential operators to study the existence of Liouvillian
1875: solutions and differential Galois group see e.g.
1876: \cite{Singer:93::a,Singer:95::,Ulmer:96::}.
1877:
1878: %
1879:
1880: To decide if case 2 from Lemma~\ref{lem:alg} occurs we can apply the
1881: Kovacic algorithm. Here we present its part devoted to this case and
1882: adopted to a Fuchsian equation. At first we introduce notation. We
1883: write $r(z)\in\mathbb{C}(z)$ in the form
1884: %
1885: \begin{equation*}
1886: \label{eq:rst}
1887: r(z) = \frac{s(z)}{t(z)}, \qquad s(z),\, t(z) \in \mathbb{C}[z],
1888: \end{equation*}
1889: %
1890: where $s(z)$ and $t(z)$ are relatively prime polynomials and $t(z)$ is
1891: monic. The roots of $t(z)$ are poles of $r(z)$. We denote $\Sigma':=
1892: \{ c\in\mathbb{C}\,\vert\, t(c) =0 \}$ and
1893: $\Sigma:=\Sigma'\cup\{\infty\}$. The order $\ord(c)$ of $c\in\Sigma'$
1894: is equal to the multiplicity of $c$ as a root of $t(z)$, the order of
1895: infinity is defined by
1896: %
1897: \[
1898: \mathrm{ord}(\infty):= \max(0, 4+\deg s - \deg t).
1899: \]
1900:
1901:
1902:
1903: Because we assume that equation~\eqref{eq:gso} is Fuchsian, we have
1904: $\ord(c)\leq 2$ for $c\in\Sigma$. For each $c\in\Sigma'$ we have the
1905: following expansion
1906: \[
1907: r(z) = \frac{a_c}{(z-c)^2} + O\left( \frac{1}{z-c}\right),
1908: \]
1909: and we define $\Delta_c = \sqrt{1+4a_c}$. For infinity we have
1910:
1911: %
1912: \begin{equation*}
1913: r(z)=\dfrac{a_\infty}{z^2}+O\left(\dfrac{1}{z^3}\right),
1914: \end{equation*}
1915: %
1916: and we define $\Delta_\infty = \sqrt{1+4a_\infty}$.
1917:
1918: The algorithm consists of three steps.
1919:
1920: \noindent
1921: \textbf{Step I.}
1922: For $c\in\Sigma'$ such that $\ord(c)=1$ we define
1923: $E_c=\{4\}$;
1924: if $\ord(c)=2 $
1925: \[
1926: E_c := \left\{ 2 , 2(1+\Delta_c),
1927: 2(1-\Delta_c)\right\}\cap\mathbb{Z}.
1928: \]
1929: If $\ord(\infty)<2$ we put $E_\infty=\{ 0,2,4\}$; if
1930: $\ord(\infty)=2$ we define
1931: \[
1932: E_\infty := \left\{2, 2(1+\Delta_\infty),
1933: 2(1-\Delta_\infty)\right\}\cap\mathbb{Z}.
1934: \]
1935: \textbf{Step II.} For each $e$ in the Cartesian product
1936: \[
1937: E:= {E}_{\infty}\times\prod_{c\in\Sigma'}{E}_c,
1938: \]
1939: we compute
1940: \[
1941: d(e) := \frac{1}{2}\left(e_\infty- \sum_{c\in\Sigma'}e_c\right).
1942: \]
1943: We select those elements $e\in E$ for which $d(e)$ is a
1944: non-negative integer. If there are no such elements Case 2 from
1945: Lemma~\ref{lem:alg} cannot occur and the algorithm stops here.
1946:
1947: \noindent
1948: \textbf{Step III.} For each element $e\in{E}$ such that
1949: $d(e)=n\in\mathbb{N}_0$ we define
1950: \[
1951: \theta=\theta(z) = \frac{1}{2}
1952: \sum_{c\in\Sigma'}\frac{e_c}{z-c},
1953: \]
1954: and we search for a monic polynomial $P=P(z)$ of degree $n$ satisfying the
1955: following equation
1956: \[
1957: P''' + 3\theta P'' +(3 \theta^2 + 3\theta' -4r)P' +
1958: (\theta'' + 3 \theta\theta' + \theta^3 -4 r\theta - 2r')P =0.
1959: \]
1960: If such polynomial exists, then equation~\eqref{eq:gso} possesses a
1961: solution of the form $w=\exp\int\omega$, where
1962: \[
1963: \omega^2 + \psi\omega +\frac{1}{2}\psi' + \frac{1}{2}\psi^2 - r =0, \qquad
1964: \psi = \theta + \frac{P'}{P}.
1965: \]
1966: If we do not find such polynomial, then case 2 in Lemma~\ref{lem:alg}
1967: cannot occur.
1968:
1969:
1970: \subsection{Riemann $P$ equation}
1971:
1972: The Riemann $P$ equation \cite{Whittaker:35::} is the most general
1973: second order differential equation with three regular singularities.
1974: If we place, using homography, these singularities at $z=0,1,\infty$,
1975: then it has the form
1976: %
1977: \begin{equation}
1978: \label{eq:riemann}
1979: \begin{split}
1980: \dfrac{\mathrm{d}^2\xi}{\mathrm{d}z^2}&+\left(\dfrac{1-\alpha-\alpha'}{z}+
1981: \dfrac{1-\gamma-\gamma'}{z-1}\right)\dfrac{\mathrm{d}\xi}{\mathrm{d}z}\\
1982: &+
1983: \left(\dfrac{\alpha\alpha'}{z^2}+\dfrac{\gamma\gamma'}{(z-1)^2}+
1984: \dfrac{\beta\beta'-\alpha\alpha'-\gamma\gamma'}{z(z-1)}\right)\xi=0,
1985: \end{split}
1986: \end{equation}
1987: %
1988: where $(\alpha,\alpha')$, $(\gamma,\gamma')$ and $(\beta,\beta')$ are the
1989: exponents at singular points. Exponents satisfy the Fuchs relation
1990: %
1991: \[
1992: \alpha+\alpha'+\gamma+\gamma'+\beta+\beta'=1.
1993: \]
1994: %
1995: We denote differences of exponents by
1996: %
1997: \[
1998: \lambda=\alpha-\alpha',\qquad\nu=\gamma-\gamma',\qquad\mu=\beta-\beta'.
1999: \]
2000: %
2001: For equation \eqref{eq:riemann} the necessary and sufficient
2002: conditions for solvability of the identity component of its
2003: differential Galois group are given by the following theorem due to
2004: Kimura \cite{Kimura:69::}, see also \cite{Morales:99::c}.
2005: %
2006: \begin{theorem}[Kimura]
2007: \label{thm:kimura}
2008: The identity component of the differential Galois group of
2009: equation~\eqref{eq:riemann} is solvable if and only if
2010: \begin{itemize}
2011: \item[A:] at least one of four numbers $\lambda+\mu+\nu$,
2012: $-\lambda+\mu+\nu$, $\lambda-\mu+\nu$, $\lambda+\mu-\nu$ is an odd
2013: integer, or
2014: \item[B:] the numbers $\lambda$ or $-\lambda$ and $\mu$ or $-\mu$ and
2015: $\nu$ or $-\nu$ belong (in an arbitrary order) to some of the
2016: following fifteen families
2017: \begin{center}
2018: \begin{tabular}{|c|c|c|c|c|}
2019: \hline
2020: 1&$1/2+l$&$1/2+s$&arbitrary complex number&\\\hline
2021: 2&$1/2+l$&$1/3+s$&$1/3+q$&\\\hline
2022: 3&$2/3+l$&$1/3+s$&$1/3+q$&$l+s+q$ even\\\hline
2023: 4&$1/2+l$&$1/3+s$&$1/4+q$&\\\hline
2024: 5&$2/3+l$&$1/4+s$&$1/4+q$&$l+s+q$ even\\\hline
2025: 6&$1/2+l$&$1/3+s$&$1/5+q$&\\\hline
2026: 7&$2/5+l$&$1/3+s$&$1/3+q$&$l+s+q$ even\\\hline
2027: 8&$2/3+l$&$1/5+s$&$1/5+q$&$l+s+q$ even\\\hline
2028: 9&$1/2+l$&$2/5+s$&$1/5+q$&$l+s+q$ even\\\hline
2029: 10&$3/5+l$&$1/3+s$&$1/5+q$&$l+s+q$ even\\\hline
2030: 11&$2/5+l$&$2/5+s$&$2/5+q$&$l+s+q$ even\\\hline
2031: 12&$2/3+l$&$1/3+s$&$1/5+q$&$l+s+q$ even\\\hline
2032: 13&$4/5+l$&$1/5+s$&$1/5+q$&$l+s+q$ even\\\hline
2033: 14&$1/2+l$&$2/5+s$&$1/3+q$&$l+s+q$ even\\\hline
2034: 15&$3/5+l$&$2/5+s$&$1/3+q$&$l+s+q$ even\\\hline
2035: \end{tabular}\\[1.5ex]
2036: \end{center}
2037: Here $l,s$ and $q$ are integers.
2038: \end{itemize}
2039: \label{kimura}
2040: \end{theorem}
2041: The solvability conditions are sufficient for our purposes because if
2042: $\cG^0$ is not solvable, then obviously it is not Abelian.
2043:
2044:
2045:
2046: \subsection{Lam\'e equation}
2047:
2048: The Weierstrass
2049: form of the Lam\'e equation is following
2050: %
2051: \begin{equation}
2052: \label{eq:lame2}
2053: \frac{\mathrm{d}^2y}{\mathrm{d}t^2}=(\alpha\wp(t;g_2,g_3)+\beta)y,
2054: \end{equation}
2055: %
2056: where $\alpha$ and $\beta$ are, in general, complex parameters and
2057: $\wp(t;g_2,g_3)$ is the elliptic Weierstrass function with invariants
2058: $g_2$, $g_3$. In other words, $\wp(t;g_2,g_3)$ is a solution of the
2059: differential equation
2060: %
2061: \begin{equation*}
2062: \dot v^2=f(v),\qquad f(v)=4v^3-g_2v-g_3.
2063: \label{wei}
2064: \end{equation*}
2065: %
2066: It is assumed that equation $f(v)=0$ has three different roots, so
2067: %
2068: \begin{equation*}
2069: \Delta=g_2^3 -27g_3^2 \neq 0.
2070: \end{equation*}
2071: %
2072: We recall that the modular function $j(g_2,g_3)$ associated with the
2073: elliptic curve $u^2=4v^3-g_2v-g_3$ is defined as follows
2074: %
2075: \begin{equation*}
2076: \label{eq:j}
2077: j(g_2,g_3)=\frac{g_2^2}{g_2^3 -27g_3^2}.
2078: \end{equation*}
2079: %
2080: Classically the Lam\'e equation is written with parameter $n$
2081: instead of $\alpha$ related by the formula $\alpha=n(n+1)$. We see that
2082: the Lam\'e equation depends on four parameters $(n,\beta,g_2,g_3)$.
2083: The following lemma lists all the cases in which the identity
2084: component of the differential Galois group of Lam\'e
2085: equation~\eqref{eq:lame2} is Abelian, see
2086: \cite[Sec.~2.8.4]{Morales:99::c}.
2087: %
2088: \begin{lemma}
2089: \label{lem:lame}
2090: The identity component of the differential Galois group of Lam\'e
2091: equation~\eqref{eq:lame2} is Abelian only in the following cases:
2092: %
2093: \begin{enumerate}
2094: \item the Lam\'e-Hermite case when $n\in\mathbb{Z}$ and three other
2095: parameters are arbitrary.
2096: \item the Brioschi-Halphen-Crowford case for which
2097: $m:=n+\frac{1}{2}\in\mathbb{N}$, and remaining parameters
2098: $(g_2,g_3,\beta)$ satisfy an algebraic equation
2099: \begin{equation*}
2100: \label{eq:bdets}
2101: Q_{m}\left(g_2,g_3,\beta\right)=0,
2102: \end{equation*}
2103: see below.
2104: \item the Baldassarri case $2n\not\in\Z$, and
2105: \[
2106: n\pm q\in\Z\qquad\text{for some}\quad
2107: q\in\{1/4,1/6,1/10,3/10\},
2108: \]
2109: with additional algebraic restrictions
2110: on $(g_2,g_3,\beta)$.
2111: \end{enumerate}
2112: \end{lemma}
2113: %
2114:
2115: Polynomial $Q_m(g_2,g_3,\beta)$ which appears in the
2116: Brioschi-Halphen-Crowford case, called the Brioschi determinant, is
2117: defined as follows
2118: %
2119: \begin{equation}
2120: \label{eq:bdet}
2121: \left|\begin{array}{ccccccc}
2122: \beta &m-1&0&0&0&\ldots& 0\\
2123: q_{2,1}&\beta &2(m-2)&0&0&\ldots&0 \\
2124: q_{3,1}&q_{3,2}&\beta&3(m-3)& 0& \dots&0\\
2125: 0 &q_{4,2}&q_{4,3}&\beta&4(m-4)& \dots &0 \\
2126: 0 &0&\ddots&\ddots&\ddots&\ldots& 0\\
2127: 0 & 0& \ldots & q_{m-2,m-3} & q_{m-1,m-2} &\beta & m-1\\
2128: 0 & 0& \ldots & 0 &q_{m,m-2}&q_{m,m-1}&\beta
2129: \end{array}
2130: \right|,
2131: \end{equation}
2132: %
2133: where
2134: %
2135: \[
2136: q_{i+1,i} = \frac{g_2}{4}(2m-i)(m-i), \qquad
2137: q_{i+2,i} = \frac{g_3}{4}(2m-i)(2m-i-1).
2138: \]
2139: %
2140:
2141: Algebraic restrictions on $(g_2,g_3,\beta)$ in the Baldassarri case
2142: are involved. Instead of them we use the following proposition which
2143: follows from one unpublished result of B.~Dwork,
2144: see~\cite{Morales:99::c}.
2145: %
2146: \begin{proposition}
2147: \label{prop:dwork}
2148: The Baldassarri case for equation~\eqref{eq:lame2} occurs only for a finite number of values of $j(g_2,g_3)$.
2149: \end{proposition}
2150:
2151:
2152:
2153:
2154:
2155:
2156:
2157:
2158: %
2159: %\bibliographystyle{ajmplain}
2160: %\bibliography{mathreva,ajm,books,ziglin,dgt,morales,churchill,kozlov,oldies}
2161: %
2162: \newcommand{\noopsort}[1]{}\def\cprime{$'$}
2163: \def\cydot{\leavevmode\raise.4ex\hbox{.}}
2164: \begin{thebibliography}{10}
2165:
2166: \bibitem{Arnold:78::}
2167: Arnold, V.~I., \emph{Mathematical Methods of Classical Mechanics}, Graduate
2168: Texts in Mathematics, Springer-Verlag, New York, 1978.
2169:
2170: \bibitem{Audin:96::}
2171: Audin, M., \emph{Spinning tops}, volume~51 of \emph{Cambridge Studies in
2172: Advanced Mathematics}, Cambridge University Press, Cambridge, 1996.
2173:
2174: \bibitem{Churchill:90::b}
2175: Baider, A. and Churchill, R.~C., On monodromy groups of second-order {F}uchsian
2176: equations, \emph{SIAM J. Math. Anal.}, 21(6):1642--1652, 1990.
2177:
2178: \bibitem{Churchill:96::b}
2179: Baider, A., Churchill, R.~C., Rod, D.~L., and Singer, M.~F., On the
2180: infinitesimal geometry of integrable systems, in \emph{Mechanics Day
2181: (Waterloo, ON, 1992)}, volume~7 of \emph{Fields Inst. Commun.}, pages 5--56,
2182: Amer. Math. Soc., Providence, RI, 1996.
2183:
2184: \bibitem{Beukers:92::}
2185: Beukers, F., Differential {G}alois theory, in \emph{From Number Theory to
2186: Physics (Les Houches, 1989)}, pages 413--439, Springer, Berlin, 1992.
2187:
2188: \bibitem{Bolotin:90::}
2189: Bolotin, S.~V., Double asymptotic trajectories and conditions for integrability
2190: of {H}amiltonian systems, \emph{Vestnik Moskov. Univ. Ser. I Mat. Mekh.},
2191: (1):55--63, 1990.
2192:
2193: \bibitem{Bolotin:92::a}
2194: Bolotin, S.~V., Variational methods for constructing chaotic motions in the
2195: dynamics of a rigid body, \emph{Prikl. Mat. Mekh.}, 56(2):230--239, 1992.
2196:
2197: \bibitem{Borisov:01::}
2198: Borisov, A. and Mamaev, I., \emph{Dynamics of rigid body.}, NITS Regulyarnaya i
2199: Khaoticheskaya Dinamika, Izhevsk, 2001, in Russian.
2200:
2201: \bibitem{Churchill:91::a}
2202: Churchill, R.~C. and Rod, D.~L., On the determination of {Z}iglin monodromy
2203: groups, \emph{SIAM J. Math. Anal.}, 22(6):1790--1802, 1991.
2204:
2205: \bibitem{Churchill:95::a}
2206: Churchill, R.~C., Rod, D.~L., and Singer, M.~F., Group-theoretic obstructions
2207: to integrability, \emph{Ergodic Theory Dynam. Systems}, 15(1):15--48, 1995.
2208:
2209: \bibitem{Dokshevich:92::}
2210: Dokshevich, A.~I., \emph{Resheniya v konechnom vide uravnenii
2211: {E}ilera-{P}uassona}, ``Naukova Dumka'', Kiev, 1992.
2212:
2213: \bibitem{Dovbysh:90::}
2214: Dovbysh, S.~A., Splitting of separatrices of unstable uniform rotations and
2215: nonintegrability of a perturbed {L}agrange problem, \emph{Vestnik Moskov.
2216: Univ. Ser. I Mat. Mekh.}, (3):70--77, 1990.
2217:
2218: \bibitem{Duval:92::}
2219: Duval, A. and Loday-Richaud, M., Kovacic's algorithm and its application to
2220: some families of special functions, \emph{Appl. Algebra Engrg. Comm.
2221: Comput.}, 3(3):211--246, 1992.
2222:
2223: \bibitem{Golubev:53::}
2224: Golubev, V.~V., \emph{Lectures on Integration of Equations of Motion of a Rigid
2225: Body about a Fixed Point}, Gosud. Isdat. Teh. Teor. Lit., Moscow, 1953, in
2226: Russian.
2227:
2228: \bibitem{Goryachev:1910::}
2229: Goryachev, D.~I., New integrable cases of integrability of the {E}uler
2230: dynamical equations, \emph{Warsaw Univ. Izv.}, 1910, in Russian.
2231:
2232: \bibitem{Kaplansky:76::}
2233: Kaplansky, I., \emph{An Introduction to Differential Algebra}, Hermann, Paris,
2234: second edition, 1976.
2235:
2236: \bibitem{Kimura:69::}
2237: Kimura, T., On {R}iemann's equations which are solvable by quadratures,
2238: \emph{Funkcial. Ekvac.}, 12:269--281, 1969/1970.
2239:
2240: \bibitem{Kirillov:76::}
2241: Kirillov, A.~A., \emph{Elements of the theory of representations},
2242: Springer-Verlag, Berlin, 1976, translated from the Russian by Edwin Hewitt,
2243: Grundlehren der Mathematischen Wissenschaften, Band 220.
2244:
2245: \bibitem{Kovacic:86::}
2246: Kovacic, J.~J., An algorithm for solving second order linear homogeneous
2247: differential equations, \emph{J. Symbolic Comput.}, 2(1):3--43, 1986.
2248:
2249: \bibitem{Kowalevski:1888::}
2250: Kowalevski, S., Sur le probl\`eme de la rotation d'un corps solide autour d'un
2251: point fixe, \emph{Acta Math.}, 12:177--232, 1888.
2252:
2253: \bibitem{Kowalevski:1890::}
2254: Kowalevski, S., Sur une propri\'et\'e du syst\'eme d'\'equations
2255: diff\'erentielles qui d\'efinit la rotation d'un corps solide autour d'un
2256: point fixe, \emph{Acta Math.}, 14:81--93, 1890.
2257:
2258: \bibitem{Kozlov:75::a}
2259: Kozlov, V.~V., The nonexistence of an additional analytic integral in the
2260: problem of the motion of a nonsymmetric heavy solid body around a fixed
2261: point, \emph{Vestnik Moskov. Univ. Ser. I Mat. Mekh.}, (1):105--110, 1975.
2262:
2263: \bibitem{Kozlov:80::}
2264: Kozlov, V.~V., \emph{Methods of Qualitative Analysis in the Dynamics of a Rigid
2265: Body}, Moskov. Gos. Univ., Moscow, 1980.
2266:
2267: \bibitem{Kozlov:85::c}
2268: Kozlov, V.~V. and Treshchev, D.~V., Nonintegrability of the general problem of
2269: rotation of a dynamically symmetric heavy rigid body with a fixed point. {I},
2270: \emph{Vestnik Moskov. Univ. Ser. I Mat. Mekh.}, (6):73--81, 1985.
2271:
2272: \bibitem{Kozlov:86::}
2273: Kozlov, V.~V. and Treshchev, D.~V., Nonintegrability of the general problem of
2274: rotation of a dynamically symmetric heavy rigid body with a fixed point.
2275: {II}, \emph{Vestnik Moskov. Univ. Ser. I Mat. Mekh.}, (1):39--44, 1986.
2276:
2277: \bibitem{Lagrange:1889::}
2278: Lagrange, J.~L., \emph{M\'ecanique analytique}, volume~2, Chez {L}a {V}euve
2279: {D}esaint, {L}ibraire, Paris, 1788.
2280:
2281: \bibitem{Leimanis:65::}
2282: Leimanis, E., \emph{The General Problem of the Motion of Coupled Rigid Bodies
2283: about a Fixed Point}, Springer-Verlag, Berlin, 1965.
2284:
2285: \bibitem{Maciejewski:01::i}
2286: Maciejewski, A.~J., Non-integrability in gravitational and cosmological models.
2287: {I}ntroduction to {Z}iglin theory and its differential {G}alois extension, in
2288: A.~J. Maciejewski and B.~Steves, editors, \emph{The Restless Universe.
2289: Applications of Gravitational N-Body Dynamics to Planetary, Stellar and
2290: Galactic Systems}, pages 361--385, 2001.
2291:
2292: \bibitem{Maciejewski:01::j}
2293: Maciejewski, A.~J., Non-integrability of a certain problem of rotational motion
2294: of a rigid satellite, in H.~Pr{\c e}tka-Ziomek, E.~Wnuk, P.~K. Seidelmann,
2295: and D.~Richardson, editors, \emph{Dynamics of Natural and Artificial
2296: Celestial Bodies}, pages 187--192, Kluwer Academic Publisher, 2001.
2297:
2298: \bibitem{Maciejewski:02::b}
2299: {Maciejewski}, A.~J., Non-integrability of certain {H}amiltonian systems.
2300: {A}pplications of {M}orales-{R}amis differential {G}alois extension of
2301: {Z}iglin theory, in T.~Crespo and Z.~Hajto, editors, \emph{Differential
2302: Galois Theory}, volume~58, pages 139--150, Banach Center Publication, Warsaw,
2303: 2002.
2304:
2305: \bibitem{Maciejewski:02::f}
2306: {Maciejewski}, A.~J. and {Przybylska}, M., Non-integrability of {ABC} flow,
2307: \emph{Phys. Lett. A}, 303:265--272, 2002.
2308:
2309: \bibitem{Maciejewski:02::a}
2310: {Maciejewski}, A.~J. and {Przybylska}, M., Non-integrability of the {S}uslov
2311: problem, \emph{Regul. Chaotic Dyn.}, 7(1):73--80, 2002.
2312:
2313: \bibitem{Maciejewski:03::e}
2314: Maciejewski, A.~J. and Przybylska, M., Non-integrability of restricted two-body
2315: problems in constant curvature spaces, \emph{Regul. Chaotic Dyn.},
2316: 8(4):413--430, 2003.
2317:
2318: \bibitem{Maciejewski:03::a}
2319: {Maciejewski}, A.~J. and {Przybylska}, M., Non-integrability of the problem of
2320: a rigid satellite in gravitational and magnetic fields, \emph{Celestial
2321: Mech.}, 87(4):317--351, 2003.
2322:
2323: \bibitem{Magid:94::}
2324: Magid, A.~R., \emph{Lectures on Differential {G}alois Theory}, volume~7 of
2325: \emph{University Lecture Series}, American Mathematical Society, Providence,
2326: RI, 1994.
2327:
2328: \bibitem{Marsden:94::}
2329: Marsden, J.~E. and Ratiu, T.~S., \emph{Introduction to Mechanics and Symmetry},
2330: number~17 in Texts in Applied Mathematics, Springer Verlag, New York, Berlin,
2331: Heidelberg, 1994.
2332:
2333: \bibitem{Morales:94::b}
2334: Morales, J.~J. and Sim{\'o}, C., Picard-{V}essiot theory and {Z}iglin's
2335: theorem, \emph{J. Differential Equations}, 107(1):140--162, 1994.
2336:
2337: \bibitem{Morales:99::c}
2338: Morales~Ruiz, J.~J., \emph{Differential {G}alois Theory and Non-Integrability
2339: of {H}amiltonian Systems}, volume 179 of \emph{Progress in Mathematics},
2340: Birkh\"auser Verlag, Basel, 1999.
2341:
2342: \bibitem{Morales:01::b1}
2343: Morales-Ruiz, J.~J. and Ramis, J.~P., Galoisian obstructions to integrability
2344: of {H}amiltonian systems. {I}, \emph{Methods Appl. Anal.}, 8(1):33--95, 2001.
2345:
2346: \bibitem{Morales:01::b2}
2347: Morales-Ruiz, J.~J. and Ramis, J.~P., Galoisian obstructions to integrability
2348: of {H}amiltonian systems. {II}, \emph{Methods Appl. Anal.}, 8(1):97--111,
2349: 2001.
2350:
2351: \bibitem{Ramis:90::}
2352: Ramis, J.-P. and Martinet, J., Th\'eorie de {G}alois diff\'erentielle et
2353: resommation, in \emph{Computer Algebra and Differential Equations}, pages
2354: 117--214, Academic Press, London, 1990.
2355:
2356: \bibitem{Singer:93::a}
2357: Singer, M.~F. and Ulmer, F., Galois groups of second and third order linear
2358: differential equations, \emph{J. Symbolic Comput.}, 16(1):9--36, 1993.
2359:
2360: \bibitem{Singer:95::}
2361: Singer, M.~F. and Ulmer, F., Necessary conditions for {L}iouvillian solutions
2362: of (third order) linear differential equations, \emph{Appl. Algebra Engrg.
2363: Comm. Comput.}, 6(1):1--22, 1995.
2364:
2365: \bibitem{Souriau:70::}
2366: Souriau, J.-M., \emph{Structure des syst\`emes dynamiques}, Ma\^{\i}trises de
2367: math\'ematiques, Dunod, Paris, 1970.
2368:
2369: \bibitem{Takano:76::}
2370: Takano, K. and Bannai, E., A global study of {J}ordan-{P}ochhammer differential
2371: equations, \emph{Funkcial. Ekvac.}, 19(1):85--99, 1976.
2372:
2373: \bibitem{Ulmer:96::}
2374: Ulmer, F. and Weil, J.-A., Note on {K}ovacic's algorithm, \emph{J. Symbolic
2375: Comput.}, 22(2):179--200, 1996.
2376:
2377: \bibitem{Put:02::}
2378: van~der Put, M. and Singer, M.~F., \emph{Galois Theory of Linear Differential
2379: Equations}, volume 328 of \emph{Grundlehren der Mathematischen Wissenschaften
2380: [Fundamental Principles of Mathematical Sciences]}, Springer-Verlag, Berlin,
2381: 2003.
2382:
2383: \bibitem{Whittaker:35::}
2384: Whittaker, E.~T. and Watson, G.~N., \emph{A Course of Modern Analysis},
2385: Cambridge University Press, London, 1935.
2386:
2387: \bibitem{Ziglin:80::d}
2388: Ziglin, S.~L., Branching of solutions and the nonexistence of an additional
2389: first integral in the problem of an asymmetric heavy rigid body in motion
2390: relative to a fixed point., \emph{Dokl. Akad. Nauk SSSR}, 251(4):786--790,
2391: 1980.
2392:
2393: \bibitem{Ziglin:82::b}
2394: Ziglin, S.~L., Branching of solutions and non-existence of first integrals in
2395: {H}amiltonian mechanics. {I}, \emph{Functional Anal. Appl.}, 16:181--189,
2396: 1982.
2397:
2398: \bibitem{Ziglin:83::b}
2399: Ziglin, S.~L., Branching of solutions and non-existence of first integrals in
2400: {H}amiltonian mechanics. {II}, \emph{Functional Anal. Appl.}, 17:6--17, 1983.
2401:
2402: \bibitem{Ziglin:97::}
2403: Ziglin, S.~L., On the absence of a real-analytic first integral in some
2404: problems of dynamics, \emph{Functional Anal. Appl.}, 31(1):3--9, 1997.
2405:
2406: \end{thebibliography}
2407:
2408: \end{document}
2409:
2410: