1: \documentclass[12pt]{article}
2: \usepackage{amsmath}
3: \usepackage{amssymb}
4: \usepackage{mathtext}
5: \usepackage{graphicx}
6:
7: \textwidth=15cm \textheight=21cm
8:
9: \begin{document}
10:
11: \section*{LIE ALGEBRAS IN VORTEX DYNAMICS AND \\
12: CELESTIAL MECHANICS --- IV}
13: {\bf 1) Classificaton of the algebra of $n$ vortices on a plane\\
14: 2) Solvable problems of vortex dynamics\\
15: 3) Algebraization and reduction in a three-body problem}\footnote{REGULAR AND
16: CHAOTIC DYNAMICS V.4, No. 1, 1999\\
17: {\it Received March 22, 1999\\
18: AMS MSC 76C05}}\bigskip
19:
20: \begin{centering}
21: A.\,V.\,BOLSINOV\\
22: Faculty of Mechanics and Mathematics,\\
23: Department of Topology and Aplications\\
24: M.\,V.\,Lomonosov Moscow State University\\
25: Vorob'ievy Gory, Moscow, Russia, 119899 \\
26: E-mail: bols@difgeo.math.msu.su\medskip\\
27: A.\,V.\,BORISOV\\
28: Faculty of Mechanics and Mathematics,\\
29: Department of Theoretical Mechanics
30: Moscow State University\\
31: Vorob'ievy gory, 119899 Moscow, Russia\\
32: E-mail: borisov@uni.udm.ru\medskip\\
33: I.\,S.\,MAMAEV\\
34: Laboratory of Dynamical Chaos and Non Linearity,\\
35: Udmurt State University\\
36: Universitetskaya, 1, Izhevsk, Russia, 426034\\
37: E-mail: rcd@uni.udm.ru\\
38: \end{centering}
39:
40: \begin{abstract}
41: The work~\cite{BPavlov} introduces a naive description of dynamics of
42: point vortices on a plane in terms of variables of distances and areas
43: which generate Lie--Poisson structure. Using this approach a
44: qualitative description of dynamics of point vortices on a plane and
45: a sphere is obtained in the works~\cite{BorisovLebedev,BorisovLeb}.
46: In this paper we consider more formal constructions of the general problem of~$n$
47: vortices on a plane and a sphere. The developed methods of algebraization
48: are also applied to the classical
49: problem of the reduction in the three-body problem.
50: \end{abstract}
51:
52: \section{Classification of the algebra of $n$ vortices on a plane}
53:
54: \subsection{Vortex algebra and Lie pencils}
55:
56: Let us introduce coordinates of vortices ${\bf r}_k(x_k,y_k),\;k=1,\dots,n$
57: by complex variables $z_k= x_k+iy_k,\;k=1,\ldots,n.$
58: The set ${\bf z}=(z_1,\dots,z_n)$ defines a vector in
59: a complex linear space ${\mathbb C}^n$ with an Hermitian form
60: \begin{equation}
61: \label{Bol1}
62: ({\bf z,\bf w})_\Gamma=\sum_{i=1}^n \Gamma_i z_i \overline{w}_i,\quad
63: \Gamma_i\in{\mathbb R}\,,
64: \end{equation}
65: where $\Gamma_i$ are vorticities.
66:
67: %\wfig*[7]{brq}
68:
69: The imaginary part
70: of (\ref{Bol1}) gives the
71: symplectic form corresponding to the Poisson
72: bracket $\{x_i,y_j\}=\frac{\delta_{ij}}{\Gamma_i}$.
73: The Hamiltonian~$\mathcal H$ and integrals of motion $P,\,Q,\,I$ of
74: vortex system can be written as
75: \begin{gather} \mathcal
76: H=\frac{1}{16\pi}\sum\limits_{i,j=1}^n \Gamma_i \Gamma_j \ln |z_i-z_j|^2\,,
77: \label{new1}\\ P+iQ=\sum\limits_{i=1}^n \Gamma_i
78: z_i=({\bf z},{\bf z}_0)_{\Gamma}\,, \qquad I=\sum\limits_{i=1}^n \Gamma_i
79: |z_i|^2=({\bf z,z})_{\Gamma}\,, \label{new2} \end{gather}
80: here
81: ${\bf z}_0=(1,1,\ldots,1)$ is a constant vector.
82:
83: As relative variables we introduce the squares
84: of mutual
85: distances and doubled areas of
86: triangles spanned on three
87: vortices (see Fig.~1)
88: as in~\cite{BPavlov}
89: \begin{equation}
90: \label{Bol3/2}
91: \begin{array}{c}
92: M_{lm} =({\bf r}_i-{\bf r}_j)^2=(x_{i}-x_{j})^2+(y_{i}-y_{j})^2,\\[8pt]
93: \Delta_{ijk}=({\bf r}_j-{\bf r}_i)\wedge ({\bf r}_k-{\bf r}_i)=
94: (x_{j}-x_{i})(y_{k}-y_{i})-(x_{k}-x_{i})(y_{j}-y_{i}).
95: \end{array}
96: \end{equation}
97: New variables, as it was shown~\cite{BPavlov},
98: generate a linear bracket, which also
99: linearly depends on inverse in\-ten\-sities~$1/\Gamma_i$
100: \begin{align}
101: \{M_{ij}, M_{kl}\}={}&4\Bigl(\frac{1}{\Gamma_i}\delta_{ik}
102: -\frac{1}{\Gamma_j}
103: \delta_{jk}\Bigr)\Delta_{ijl}+4\Bigl(\frac{1}{\Gamma_i}\delta_{il}
104: -\frac{1}{\Gamma_j}\delta_{jl}\Bigr)\Delta_{ijk},\notag\\
105: \{M_{ij}, \Delta_{klm}\}={}&
106: \Bigl(\frac{1}{\Gamma_i}\delta_{ik}-\frac{1}{\Gamma_j}
107: \delta_{jk}\Bigr)(M_{li}-M_{im}+M_{mj}-M_{jl})+\notag\\
108: {}\span+\Bigl(\frac{1}{\Gamma_i}\delta_{il}
109: -\frac{1}{\Gamma_j}\delta_{jl}\Bigr)(M_{mi}-M_{ik}+M_{kj}-M_{jm})+\notag\\
110: {}\span+\Bigl(\frac{1}{\Gamma_i}\delta_{im}-
111: \frac{1}{\Gamma_j}\delta_{jm}\Bigr)(M_{ki}-M_{il}+M_{lj}-M_{jk}),\notag\\
112: \label{Plane7}
113: \{\Delta_{ijk},\Delta_{lmn}\}={}&
114: \frac{\delta_{il}}{\Gamma_i}(\Delta_{jkn}-\Delta_{jkm})+
115: \frac{\delta_{im}}{\Gamma_i}(\Delta_{jkl}-\Delta_{jkn})+{}\\
116: {}\span+\frac{\delta_{in}}{\Gamma_i}(\Delta_{jkm}-\Delta_{jkl})+
117: \frac{\delta_{jl}}{\Gamma_j}(\Delta_{ikm}-\Delta_{ikn})+{}\notag\\
118: {}\span+\frac{\delta_{jm}}{\Gamma_j}(\Delta_{ikn}-\Delta_{ikl})+
119: \frac{\delta_{jn}}{\Gamma_j}(\Delta_{ikl}-\Delta_{ikm})+{}\notag\\
120: {}\span+\frac{\delta_{kl}}{\Gamma_k}(\Delta_{ijn}-\Delta_{ijm})+
121: \frac{\delta_{km}}{\Gamma_k}(\Delta_{ijl}-\Delta_{ijn})+
122: \frac{\delta_{kn}}{\Gamma_k}(\Delta_{ijm}-\Delta_{ijl}).\notag
123: \end{align}
124:
125: Let us set invariant relations corresponding to real motions~\cite{BPavlov}
126: \begin{equation}
127: F_{ijkl}=\Delta_{ijk}+\Delta_{ikl}-\Delta_{lij}-\Delta_{ljk}=0\,,
128: \label{Plane11}
129: \end{equation}
130: \vspace{-7mm}
131: \begin{equation}
132: F_{ijk}={(2\Delta_{ijk})}^2+M_{ij}^2+M_{jk}^2+
133: M_{ik}^2-2(M_{ij}M_{jk}+M_{ij}M_{ik}+M_{jk}M_{ik})=0\,,
134: \label{plane}
135: \end{equation}
136: Relations (\ref{Plane11}) mean
137: that the quadrangle spanned on the vortices $ijkl$ can be constructed
138: of triangles by two
139: ways (Fig.~1). Equations~(\ref{plane}) are Heron formulas
140: expressing the area of
141: a triangle via its sides.
142:
143: Let us note that for the structure (\ref{Plane7})
144: the Jacobi
145: identity is fulfilled only on the submanifold defined by the first set of
146: geometrical relations~(\ref{Plane11}).
147: Below, we consider the
148: relations~(\ref{Plane11}) to be fulfilled on default. Thus the
149: brackets~(\ref{Plane7}) define Lie--Poisson structure, which wee shall call
150: {\it vortex bracket}.
151:
152:
153: The bracket (\ref{Plane7}) admits the linear Casimir function
154: \cite{BPavlov}
155: \begin{equation}
156: \label{Plane10}
157: D=\sum_{i,j=1}^N \Gamma_i \Gamma_j M_{ij}=2\left(
158: I\left(\sum\limits_{i=1}^n \Gamma_i\right)-Q^2-P^2\right)\,.
159: \end{equation}
160:
161: %%%%%%%%%%%%%%%
162:
163: To define the real type of the corresponding Lie
164: algebra we indicate an explicit isomorphism with some Lie
165: pencil~\cite{Bolsinov}.
166:
167: {\bf Remark 1.} {\it The compactness conditions of the
168: real form~(\ref{Plane7}), depending on intensities,
169: imply that all mutual distances are bounded from above
170: at any moment of time.}
171:
172: We consider a space of skew Hermitian $n\times n$ matrices and
173: subspace $L$ in it generated by matrices of the form:
174: \begin{equation}
175: \label{Bol2} \begin{array}{c} \Delta_{lmk} = \begin{matrix} \mbox{}\\ l \\
176: \mbox{} \\ m \\ \mbox{} \\ k \\ \mbox{}\end{matrix} \left( \arraycolsep=3pt
177: \begin{array}{ccccccc} \ddots & & & & & & \\[-3pt] & 0
178: & & 1 & & -1 & \\[-3pt] & &\ddots& & & &
179: \\[-3pt] &-1 & & 0 & & 1 & \\[-3pt] & & &
180: &\ddots& & \\[-3pt] & 1 & & -1 & & 0 &
181: \\[-3pt] & & & & & &\ddots \end{array} \right),
182: \\ M_{lm} = \begin{matrix} \mbox{}\\ l \\ \mbox{} \\ m \\ \mbox{}
183: \end{matrix} \left( \arraycolsep=8pt
184: \begin{array}{ccccc} \ddots & &
185: & & \\[-3pt] & i & & -i & \\[-3pt] & &\ddots&
186: & \\[-3pt] &-i & & i & \\[-3pt] & & &
187: &\ddots \end{array} \right). \end{array} \end{equation}
188:
189: Let us describe the main properties of this subspace which can be verified by a
190: straightforward calculation.
191:
192: {\bf Proposition 1.}
193: \begin{itemize}
194: \item[1)]
195: {\it The $L$ is a Lie subalgebra,}
196:
197: \item[2)] {\it This subalgebra is isomorphic to $u(n-1)$,}
198:
199: \item[3)]
200: {\it The center of this subalgebra is generated by the element} %%%%
201: $\sum\limits_{m,l} M_{ml}$,
202:
203: \item[4)] {\it The elements $\Delta_{lms}$ generate a
204: subalgebra in $L$ isomorphic to $so(n-1)$,}
205:
206: \item[5)] {\it The
207: %%% The %%%
208: following relation holds
209: %%%% holds вместо exists %%%%
210: for any} $k,l,m,s$:
211: $$ \Delta_{klm} + \Delta_{kms} +
212: \Delta_{lks} + \Delta_{mls} = 0, $$
213:
214: \item[6)] {\it Commutation relations in
215: %%% убрать algebra %%%%
216: $L$ coincide with the vortex bracket
217: (\ref{Plane7}) for the case when all intensities are the same (and equal
218: to unity). }\end{itemize}
219:
220: Therefore, (\ref{Bol2}) determines
221: an $n$-dimensional (unitary) linear representation of Lie algebra
222: ${u(n-1)}$. There are no such irreducible representations,
223: that is why it splits into a sum of the standard
224: ${n-1}$-dimentional representation and one-dimensional
225: %%%one-dimensional %%%%
226: trivial representation. This splitting is organized
227: as follows.
228:
229: {\bf Proposition 2.}
230: {\it The representation space $V = \Bbb C^n$ splits
231: into a direct sum of invariant subspaces
232: $V = V_1 \oplus V_2$, where $V_1 ={\mathbb C}^{n-1}$ is given by
233: $z_1 + z_2 + \dots + z_n = 0$ and a one-dimensional subspace $V_2 =
234: {\mathbb C}$
235: is spanned on the vector ${\bf z}_0=(1,1,\dots, 1)$.}
236:
237: Let us turn to the case of arbitrary intensities $\Gamma_1,\dots,\Gamma_n.$
238: Consider the Lie pencil on the algebra of
239: skew Hermitian $n\times n$ matrices generated by commutators
240:
241: \begin{equation}
242: \label{eqqq*} [X, Y]_{\Gamma^{-1}} = X\Gamma^{-1} Y - Y\Gamma^{-1} X,
243: \end{equation} where $\Gamma^{-1}$ is a real diagonal matrix of the form
244: \begin{equation} \label{eq46_3} \Gamma^{-1} = \left( \begin{matrix} \frac
245: {1}{\Gamma_1} & & & \\ & \frac {1}{\Gamma_2} & & \\ & & \ddots
246: & \\ & & & \frac {1}{\Gamma_n} \end{matrix} \right). \end{equation}
247:
248: The remarkable fact is that the subalgebra $L$ is closed
249: with respect to the commutator $[\cdot,\cdot]_{\Gamma^{-1}}$.
250: Thus the family of commutators%%%%%
251: ~\eqref{eqqq*} generates some Lie pencil on $L$. Moreover,
252: restricting the commutator $[\cdot,\cdot]_{\Gamma^{-1}}$ on $L$ we
253: obtain the Lie algebra isomorphic to the vortex
254: algebra~$L_\Gamma$ corresponding to intensities $\Gamma_1, \dots,
255: \Gamma_n$. Thus one can deduce a symmetric isomorphism between the family
256: of vortex algebras and a rather simple Lie pencil.
257:
258: Using this construction we shall describe
259: properties of vortex algebras.
260:
261: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
262: {\bf Proposition 3.}
263: {\it For positive $\Gamma_i$ the vortex algebra is isomorphic
264: to~${u(n-1)}$.}
265:
266: {\it Proof.}
267:
268: The corresponding isomorphism is constructed as follows. From the beginning
269: let all intensities be equal to unity. All matrices~\eqref{Bol2}
270: are skew Hermitian and satisfy the following property~$A{\bf z}_0=0$,
271: where~${\bf z}_0$ is the vector with coordinates
272: ~$(1,\,1,\,\ldots,\,1)$. In other words, this vector is invariant
273: under the vortex algebra.
274: Therefore its orthogonal completment, i.~e., the hyperplane~$V_1=
275: \{{\bf z} \mid\sum z_i = 0\}$ is also invariant.
276:
277:
278: We consider all skew Hermitian matrices possessing this
279: property (that is, mapping ${\bf z}_0$ into zero) %%%%.
280: They form the subalgebra~$u(n-1)$. Thus the
281: vortex algebra in
282: %%% the %%%
283: case of equal intensities is embedded into~$u(n-1)$.
284: But their dimensions coincide, therefore the
285: algebras also coincide.
286:
287: This can be shown more explicitly by choosing %%%%
288: another basis in space $\Bbb C^n$. Consider a basis $\bf e_1, \bf e_2, \dots
289: , \bf e_n$ where $\bf e_1, \dots , \bf e_{n-1}$ is an %%%
290: orthonormal
291: basis in the hyperplane $V_1$ (see Proposition~2) and
292: ${\bf e}_n = (1/\sqrt {n},\dots, 1/\sqrt {n})$. Rewriting
293: matrices~(\ref{Bol2}) from the vortex algebra in the new basis one can see
294: that they take the form
295: \begin{equation}
296: \label{new3_j}
297: \left( \begin{matrix} A_{n-1} & \begin{matrix} 0 \\
298: \vdots \\ 0 \end{matrix} \\ \begin{matrix} 0 & \hdots & 0 \end{matrix} & 0
299: \end{matrix} \right),
300: \end{equation}
301: where $A_{n-1}$ is a skew Hermitian
302: $(n-1)\times (n-1)$ matrix. In this basis the isomorphism of the vortex algebra with $u(n-1)$ is evident.
303: A defect of this basis is that this basis can't be made symmetric with
304: respect to all vortices.
305:
306: Let us generalize these arguments to the case of arbitrary
307: intensities.
308: Transition to the new basis in~${\mathbb C}^n$ defines the conjugation of
309: matrices
310: in algebra~(\ref{Bol2}) of the form $A = C A' C^{-1}$, where $C$ is
311: the transformation matrix.
312: (In our case we can assume it to be real and
313: orthogonal ~$C^{-1}=C^T$). With this substitution the commutator of the
314: vortex algebra transforms to the following form:
315: $$ \begin{aligned}
316: A\Gamma^{-1} B - B \Gamma^{-1} A& = CA' C^{-1} \Gamma^{-1}
317: C B' C^{-1} - CB' C^{-1}
318: \Gamma^{-1} C A' C^{-1} =\\
319: \mbox{}&= C [A', B']_{\Gamma '} C^{-1},
320: \end{aligned}
321: $$
322: where
323: $$
324: \Gamma'= C^{-1} \Gamma^{-1} C =
325: \left( \begin{matrix}
326: \Gamma'_{n-1} & \begin{matrix} \gamma_{1n} \\ \vdots \\ \gamma_{n-1, n}
327: \end{matrix} \\
328: \begin{matrix} \gamma_{1n} & \hdots & \gamma_{n-1, n} \end{matrix} & \gamma_{n,n}
329: \end{matrix} \right).
330: $$
331:
332: Here $\Gamma'_{n-1}$ is the symmetric $(n-1)\times (n-1)$ matrix
333: corresponding to the restriction of the form~$\Gamma^{-1}$ onto %%%
334: the hyperplane $V_1$. This follows from the relation $C^{-1} = C^T$.
335:
336: Taking into account that the last row
337: and column of matrices $A$ and $B$ vanish one finds that the vortex
338: commutator transforms to the form
339: $$ [A'_{n-1} ,
340: B'_{n-1}]_{\Gamma'_{n-1}}. $$
341:
342: Let us note that the matrix~$\Gamma'_{n-1}$ is positively defined and real,
343: therefore one can take the root~$(\Gamma'_{n-1})^{1/2}$.
344: Then the substitution
345: $$
346: A'_{n-1} =
347: (\Gamma'_{n-1})^{-1/2}A''_{n-1}(\Gamma'_{n-1})^{-1/2}
348: $$
349: reduces the commutator
350: $[A'_{n-1} , B'_{n-1}]_{\Gamma'_{n-1}}$ to the standard one. The matrices
351: remain skew Hermitian. This argument shows also that the
352: vortex pencil under consideration is a subpencil of the standard pencil on
353: the space of skew Hermitian matrices.$\blacksquare$
354:
355: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
356:
357: The main result following from this construction is that the properties of
358: the vortex Lie algebra corresponding to parameters
359: $\Gamma_1, \dots, \Gamma_n$ are completely determined by properties of
360: the bilinear form~$\Gamma_{n-1}'$. This form is the
361: restriction of the
362: form~$\Gamma^{-1}$ onto the subspace
363: $V_1 = \{ z_1 + \dots + z_N =
364: 0\}$ (simply by the signature of this restriction).
365:
366: {\bf Remark 2.}
367: {\it Using the method of proof of Proposition 3 it is not
368: difficult to show that in general case the vortex algebra is isomorphic
369: to the algebra~$u(p,q)$ with~$p+q=n-1$.}
370:
371: Let us find conditions for which the %%%
372: algebra $L_\Gamma$ is compact, i.~e.,
373: isomorphic to the Lie algebra $u(n-1)$.
374: The necessary and sufficient condition is that $\Gamma_{n-1}'$
375: is a form of fixed sign. In this case there are the
376: following possibilities (we suppose that intensities are finite
377: and differ from zero).
378:
379: \begin{itemize}
380: \item[1)] The $\Gamma^{-1}$ is a form of fixed
381: sign, i.~e., all $\Gamma_i$
382: are simultaneously positive or negative.
383: \item[2)] The form $\Gamma^{-1}$ has the signature $(n-1, 1)$, all
384: $\Gamma_i$ except for one are positive.
385: Indeed the condition of positive definiteness %%%
386: of the restriction on~$V_1$
387: requires that the form $\Gamma^{-1}$ is negative defined onto %%%
388: a one-dimensional orthogonal complement to the $V_1$ with respect to~$\Gamma$ %%%.
389: This condition is easy to find if one notices that the orthogonal
390: complement to $V_1$ in the sense of~(\ref{eq46_3}) is a one
391: dimensional subspace spanned on the vector $v_\Gamma =
392: (\Gamma_1, \dots , \Gamma_n)$. It has the form
393: $$
394: (v_\Gamma , v_\Gamma )_{\Gamma^{-1}} = \sum \Gamma_i < 0.
395: $$
396: \item[3)] Similarly, in the case of the signature $(1, n-1)$.
397: \end{itemize}
398:
399: Finally we have the following
400:
401: {\bf Proposition 4.}
402: {\it The vortex Lie algebra $L_\Gamma$ is compact only in the following
403: cases:}
404: \begin{itemize}
405: \item[1)] {\it All intensities have the same sign.
406: \item[2)] All intensities except for one are positive and
407: $\sum\limits_{i=1}^n \Gamma_i < 0$.
408: \item[3)] All intensities except for one are negative and
409: $\sum\limits_{i=1}^n \Gamma_i > 0$.}
410: \end{itemize}
411:
412: {\bf Remark 3.} {\it
413: This proposition coincides with the one which was proved in
414: the case of three vortices \cite{BorisovLebedev}.}
415:
416: If the vortex algebra is not
417: ``semi-simple", the conditions are defined similarly.
418: %%%% я не понял смысла перевода %%%.
419: It happens iff the form
420: $\Gamma'_{n-1}$ is degenerate on $V_1$. From linear algebra it follows that
421: this condition equals to the requirement that the
422: orthogonal complement to $V_1$ lies in
423: $V_{1}$. It means that
424: $v_\Gamma \in V_1$, i.~e.
425: $$
426: \Gamma_1 + \ldots + \Gamma_n = 0.
427: $$
428:
429: Let us note one more point. The subspace $\Delta \subset L$ spanned on vectors
430: $\Delta_{lmn}$ is a subalgebra for any algebra of the pencil.
431: Its compactness conditions are the same as those
432: for the whole $L_\Gamma$, i.~e., the form $\Gamma_{n-1}'$ should be
433: positive definite. Thus the vortex algebra $L_\Gamma$ is
434: compact iff the subal\-gebra~$\Delta$ is compact.
435:
436: \subsection{Reduction with symmetries and singular orbits}
437:
438: Now we explain the origin
439: the nature of Lie pencils connected with the vortex algebra and describe
440: the (singular)
441: symplectic orbits of the algebra~(\ref{Plane7}) corresponding to real motions.
442: Consider a transition to relative variables~(\ref{Bol3/2}) from the point
443: of view of reduction with symmetries \cite{MarsdenW}.
444:
445: Hamiltonian function~\eqref{new1} and equations of motion of vortices are
446: invariant under the action of the group~$E(2)$. This action can be represented in
447: the form
448: \begin{equation}
449: \label{eqq_dop1}
450: g\colon{\bf z}\to e^{i\varphi}{\bf z}+a{\bf z}_0\,,\quad g\in E(2)\,,
451: \end{equation}
452: where~${\bf z}_0=(1,\,\ldots,\,1),\;a=a_x+ia_y$.
453: Parameters~$\varphi,\,a_x,\,a_y$ define
454: the translation and rotation corresponding to the element~$g\in E(2)$.
455:
456: The action~\eqref{eqq_dop1} is the {\em non-Hamiltonian} group action~\cite{Arnold}.
457: Indeed integrals of motion~\eqref{new2} cor\-res\-pon\-ding to translations
458: $P,\,Q$ and
459: to rotations $I$ form a Poisson structure that differs from Lie--Poisson structure
460: of the algebra~$e(2)$ by a constant, cocycle.
461: Obviously this cocycle is unremovable and standard reduction
462: with the momentum is impossible~\cite{Arnold, ArnoldGivental, MarsdenW}.
463: To execute the reduction in algebraic form we use the momentum map in some other
464: way.
465:
466: Consider the action of linear transformation
467: preserving the
468: form~$(\cdot,\cdot)_{\Gamma}$~\eqref{Bol1}.
469: The corresponding linear operators
470: (matrices) form the group which is isomorphic to~$U(p,q)$, $p+q=n$ and satisfies
471: relations
472: $$
473: X\Gamma X^+=\Gamma\,,\qquad \Gamma=\mbox{diag}(\Gamma_1,\,\ldots,\,\Gamma_n)\,.
474: $$
475: Operators of the corresponding Lie algebra~$A\in u(p,q)$ are defined by
476: \begin{equation}
477: \label{eqq_dop2}
478: A\Gamma+\Gamma A^+=0\,.
479: \end{equation}
480:
481: Since $\Gamma^+=\Gamma$ from~\eqref{eqq_dop2},
482: %%%убрать "it follows that" и поставить запятую%%%
483: the matrix~$A\Gamma$ is skew Hermitian. After the
484: substitution $\phi(A)=A\Gamma$ the standard commutator transforms into the
485: commutator
486: $[\cdot,\cdot]_{\Gamma^{-1}}$~\eqref{eqqq*}:
487: $$
488: \phi([A,B])=[\phi(A),\phi(B)]_{\Gamma^{-1}}\,.
489: $$
490: The advantage of this approach is that for any~$\Gamma_i$ the
491: algebra~\eqref{eqq_dop2} is represented on the same space of skew Hermitian
492: matrices. However, instead of the standard commutator it is necessary to
493: use~$[\cdot,\cdot]_{\Gamma^{-1}}$. In this case a family %%%
494: of Poisson
495: brackets~$\{\cdot,\cdot\}_{\Gamma^{-1}}$ appears on
496: the corresponding coalgebra~$u^*(p,q)$.
497:
498: Linear vector fields corresponding to operators~\eqref{eqq_dop2} in complex form
499: can be written as
500: $$
501: V_A=A{\bf z}=\phi(A)\Gamma^{-1}{\bf z}\,.
502: $$
503: These fields are Hamiltonian~\cite{ArnoldGivental} with
504: \begin{equation}
505: \label{eqq_dop3}
506: H_A=\frac{i}{2}(A{\bf z},{\bf z})_\Gamma=\frac i2\overline{{\bf z}}^TA
507: \Gamma{\bf z}=\frac
508: i2(\phi(A){\bf z},{\bf z})=\frac{i}{2}\overline{{\bf z}}^T\phi(A){\bf z}\,.
509: \end{equation}
510: The bracket~$\{\cdot,\cdot\}_{\Gamma^{-1}}$ of quadratic
511: Hamiltonian functions is
512: \begin{equation}
513: \label{eqq_dop4}
514: \{H_{\phi(A)}({\bf z}),H_{\phi(B)}({\bf z})\}_{\Gamma^{-1}}=\frac {i}{2}\bigl(
515: [\phi(A),\phi(B)]_{\Gamma^{-1}}{\bf z},{\bf z}\bigr),
516: \end{equation}
517: and it corresponds to the commutator~$[\cdot,\cdot]_{\Gamma^{-1}}$.
518:
519: We can describe the momentum map~$\mu({\bf z})$
520: explicitly using
521: the standard identification~$u(n)$ and~$u^*(n)$ with the help of the form
522: $\langle A,B\rangle=\mbox{Tr}AB$
523: $$
524: \mbox{Tr}\mu({\bf z})\phi(A)=\mbox{Tr}\frac {i}{2}\overline{{\bf z}}^T\phi(A){\bf z}
525: =\mbox{Tr}\left(\frac {i}{2}{\bf z}\overline{{\bf z}}^T\right)\phi(A).
526: $$
527: Thus
528: \begin{equation}
529: \label{eqq_dop5}
530: \mu({\bf z})=\frac {i}{2}{\bf z}\overline{{\bf
531: z}}^T=\frac{i}{2}\begin{pmatrix}
532: z_1\overline {z}_1&\cdots&z_1\overline {z}_n\\
533: \vdots&\ddots&\vdots\\
534: z_n\overline {z}_1&\cdots&z_n\overline {z}_n
535: \end{pmatrix}.
536: \end{equation}
537: The formula~\eqref{eqq_dop5} defines a map onto a singular symplectic orbit
538: of the algebra~$u(p,q)$ with the
539: com\-mu\-ta\-tor~$[\cdot,\cdot]_{\Gamma^{-1}}$. Now the
540: integral~(\ref{new2})~$I=\sum\Gamma_i z_i\overline{z}_i$ is a Casimir
541: function, therefore the reduction with it is carried out. In case of all
542: positive (negative) intensities the orbit is topologically homeomorphic
543: to~${\mathbb C}P^{n-1}$, because under the map~(\ref{eqq_dop5}) all points
544: of the form ~$e^{i\varphi}{\bf z}$ (orbits of the rotation group
545: action~\eqref{eq46_3}) stick together.
546:
547: Let us carry out the reduction with the remained integrals~$P,Q$~\eqref{new2}
548: on the reduced space of matrices~(\ref{eqq_dop5}).
549: Due to non-commutativity of~$P,Q$ we can reduce only one degree of
550: freedom~\cite{ArnoldKozlovNei}.
551: The constant vector field corresponding to the translation in the
552: direction~$a=a_x+ia_y$
553: is generated by the
554: linear Hamiltonian
555: %%%убрать "of the form"%%%
556: $$
557: H_a=({\bf z},\Gamma a{\bf z}_0)=({\bf z},a{\bf z}_0)_\Gamma\,.
558: $$
559: It's easy to show that Hamiltonians $H_A$~\eqref{eqq_dop3} commuting with~$H_a$
560: are generated by matrices~$A$ for which
561: $$
562: A\Gamma {\bf z}_0 =\phi(A){\bf z}_0=0\,.
563: $$
564: So~$\phi(A)$ belong to the subspace~$L$ defined in the previous section
565: (matrices~(\ref{eqq_dop5}) don't belong to this space).
566:
567: Now it is
568: %%%Now it is%%%
569: easy to see that squares of mutual distances and areas~(\ref{Bol3/2})
570: also admit
571: a natural representation of the form
572: \begin{align*}
573: M_{ij}(x,y)&=|z_i-z_j|^2=i(M_{ij}{\bf z},{\bf z})\,,\\
574: \Delta_{ijk}(x,y)&=i(\Delta_{ijk}{\bf z},{\bf z})\,,
575: \end{align*}
576: where~$M_{ij},\,\Delta_{ijk}$ are matrices~\eqref{Bol2}. The
577: momentum map~$\mu({\bf z})$ into the corresponding algebra~$u(p',q')$, $p'+q'=n-1$
578: has the form
579: \begin{equation}
580: \label{eq_dop_*}
581: \mu({\bf z})=\frac{i}{2}\left({\bf z}-\frac{({\bf z},{\bf z}_0)\Gamma{\bf
582: z}_0}{\sum\Gamma_i}\right) \left(\overline{\bf
583: z}-\frac{({\bf z},{\bf z}_0)\Gamma\overline{\bf
584: z}_0}{\sum\Gamma_i}\right)^T\,.
585: \end{equation}
586: Matrices~(\ref{eq_dop_*}) satisfy the relation~$\mu({\bf z}){\bf
587: z}_0=0$, i.~e.,
588: %%%после "i.\,e." обычно ставится запятая%%%
589: they belong to the subspace~$L$ defined in the previous
590: section. In compact case the appropriate
591: orbit is homeomorphic
592: to~${\mathbb C}P^{n-2}$
593: and corresponds to the reduced phase space (after reducing two degrees of
594: freedom).
595: The rank of matrices~(\ref{eq_dop_*}) is equal to the unit. This is their
596: characteristic property.
597:
598: {\bf Remark 4.}
599: {\it
600: The structure and compactness conditions of the reduced phase space can
601: be found by investigation (by methods of analytical geometry) of a common
602: level surface of the first integrals~(\ref{new2}). It is
603: %%%It is%%%%
604: interesting that
605: the algebraic approach also allows to solve this pure geometrical
606: problem.}
607:
608: \subsection{Symplectic coordinates}
609: Let us describe an algorithm of construction of symplectic coordinates
610: for the reduced system of~$n$ vortices in the
611: case of compact
612: algebra~$(u(n-1))$. The method of reduction of the vortex algebra (for
613: this case) to the standard representation has been described above. In
614: this case elements of the algebra are represented by skew Hermitian
615: $(n-1)\times(n-1)$ matrices with an ordinary matrix commutator:
616: \begin{equation}
617: \label{eqq_dp1} A=\begin{pmatrix}
618: ih_1&x_{12}+iy_{12}&\cdots&x_{1,n-1}+iy_{1,n-1}\\
619: -x_{12}+iy_{12}&ih_2&\cdots&x_{2,n-1}+iy_{1,n-1}\\
620: \hdotsfor{4}\\
621: -x_{1,n-1}+iy_{1,n-1}&-x_{2,n-1}+iy_{2,n-1}&\cdots&ih_{n-1}.
622: \end{pmatrix}
623: \end{equation}
624: Let us define a Poisson map~${\mathbb R}^{2(n-1)}\to u^*(n-1)$ onto
625: a singular orbit~$u^*(n-1)$ by formulas
626: \begin{equation}
627: \label{eqq_dp2}
628: \begin{aligned}
629: h_i&=p_i\,,\\
630: x_{ij}&=\sqrt{h_ih_j}\sin(\varphi_i-\varphi_j)\,,\\
631: y_{ij}&=\sqrt{h_ih_j}\cos(\varphi_i-\varphi_j)\,,
632: \end{aligned}
633: \qquad i,\,j=1,\,\ldots,\,n-1\,.
634: \end{equation}
635: The canonical bracket~$\{\varphi_i,p_j\}=\delta_{ij}$ turns in the standard
636: Lie--Poisson bracket on~$u^*(n-1)$. The orbit via variables~$p_i,\,\varphi_i$
637: is written as
638: $$
639: p_1+p_2+\cdots+p_{n-1}=C=const\,.
640: $$
641: This orbit is formed by matrices of rank~1
642: ($2\times 2$~minors
643: equal zero).
644: To introduce symplectic coordinates we carry out a symplectic transformation
645: \begin{equation}
646: \label{new**}
647: \begin{aligned}
648: r_0&=p_1+p_2+\cdots+p_{n-1}\,,\quad&s_0&=\varphi_1\,,\\
649: r_1&=p_2+\cdots+p_{n-1},\quad&s_1&=-\varphi_1+\varphi_2\,,\\
650: &\vdots&{}&\vdots\\
651: r_{n-2}&=p_{n-2}\,,\quad&s_{n-2}&=-\varphi_{n-1}+\varphi_{n-2}\,.
652: \end{aligned}
653: \end{equation}
654: It is easy to see that the map~\eqref{eqq_dp2} doesn't depend on~$s_0$,
655: therefore $s_1,\, \ldots,\,s_{n-2}$, $r_1,\,\ldots,\,r_{n-2}$ define
656: symplectic coordinates of~$2(n-2)$-dimensional orbit $\mathbb CP^{n-2}$.
657:
658: {\bf Remark 5.}
659: {\it (Co)algebra~$(u^*)u(n-1)$
660: admits a linear
661: change of variables~$h_i\mapsto h_i+ \lambda D,\;
662: D=\sum\limits_{i=1}^{n-1}h_i,\;\lambda=const$ preserving commutator %%%%
663: relations. In this connection,
664: the orbits given by matrices~$F\in u(n)$ of rank~1 map into %%%
665: orbits for which the
666: matrix~$(A-\lambda D)$ has rank 1.}
667:
668: {\it The orbit given by Heron relations~(\ref{plane}) (in compact case) after
669: the %%%
670: transition from variables~$M,\triangle$ to standard variables of
671: algebra~$u(n-1)$ turns out among the above indicated orbits for
672: some~$\lambda$.}
673:
674: \subsection{Canonical coordinates of a reduced four-vortex system.
675: Poincar\'e section}
676: Using the above described algorithm we shall point out
677: canonical coordinates explicitly for a four-vortex problem of equal intensities.
678:
679: Consider the
680: coordinates on the algebra~$u(3)$ corresponding to the matrix representation
681: of the form~(\ref{eqq_dp1})
682: \begin{equation}
683: \label{vort1}
684: A_3=\begin{pmatrix}
685: ih_1&x_3+iy_3&-x_2+iy_2\\
686: -x_3+iy_3&ih_2&x_1+iy_1\\
687: x_2+iy_2&-x_1+iy_1&ih_3,\\
688: \end{pmatrix}
689: \end{equation}
690: In accordance with the above arguments these coordinates
691: are expressed linearly via areas and
692: squares of distances
693: \begin{equation}
694: \label{vort2}
695: x_i=\sum_r\delta^r_{(x_i)}\Delta_r{(z)}\mbox{, }y_i=\sum_{k<l}m_{(y_i)}^{kl}
696: M_{kl}(z)\mbox{, } h_i=\sum_{k<l}m_{(h_i)}^{kl}M_{kl}(z),\\
697: i=1,\dots ,3\,.
698: \end{equation}
699:
700: To define coefficients~$\delta_{(x_i)}^k,m_{(y_i)}^{kl}
701: m_{(h_i)}^{kl}$ let us use the Proposition~3 and a matrix realization
702: (\ref{Bol2}) of elements~$\Delta_r,M_{kl}$,
703: where~$\Delta_1=\Delta_{234}$, $\Delta_2=-\Delta_{134}$,
704: $\Delta_3=\Delta_{124}$, $\Delta_4=-\Delta_{123}$. We choose the
705: orthogonal matrix~$C$, which reduces matrices of the vortex algebra~\eqref{vort2}
706: by the transformation~$C^TAC$ to the form~(\ref{new3_j}) as:
707: \begin{equation}
708: \label{vort3}
709: C=\begin{pmatrix}
710: -\frac{1}{2}&-\frac{1}{2}&-\frac{1}{2}&\frac{1}{2}\\
711: \frac{1}{2}&\frac{1}{2}&-\frac{1}{2}&\frac{1}{2}\\
712: \frac{1}{2}&-\frac{1}{2}&\frac{1}{2}&\frac{1}{2}\\
713: -\frac{1}{2}&\frac{1}{2}&\frac{1}{2}&\frac{1}{2}
714: \end{pmatrix}.
715: \end{equation}
716: Let us set one of coordinates~$h_i,x_i,y_i$ in the obtained matrix~$A_3$ is equal
717: to~1
718: and the others are equal to~0 and solve the system of linear equations.
719: Thus we find corresponding
720: coefficients~$\delta_{(x_i)}^k,m_{(y_i)}^{kl},m_{(h_i)}^{kl}$. In this case
721: \begin{equation}
722: \label{vort4}
723: \begin{aligned}
724: x_1&=\frac{1}{4}(-\Delta_1+\Delta_2+\Delta_3-\Delta_4)\mbox{, }&
725: x_2&=\frac{1}{4}(-\Delta_1+\Delta_2-\Delta_3+\Delta_4)\,,\\
726: x_3&=\frac{1}{4}(-\Delta_1-\Delta_2+\Delta_3+\Delta_4)\mbox{, }&
727: y_1&=\frac{1}{4}(M_{14}-M_{23})\,,\\
728: y_2&=\frac{1}{4}(M_{13}-M_{24})\mbox{, }&
729: y_3&=\frac{1}{4}(M_{12}-M_{34})\,,\\
730: h_1&=\frac{1}{4}(M_{12}+M_{34}+M_{13}+M_{24}-M_{14}-M_{23})\,,\span\span\\
731: h_2&=\frac{1}{4}(M_{12}+M_{34}-M_{13}-M_{24}+M_{14}+M_{23})\,,\span\span\\
732: h_3&=\frac{1}{4}(-M_{12}-M_{34}+M_{13}+M_{24}+M_{14}+M_{23})\,.\span\span
733: \end{aligned}
734: \end{equation}
735: Denoting canonical coordinates~(\ref{new**}) as~$(g,G,h,H)$, we have
736: \begin{equation}
737: \label{vort5}
738: \begin{array}{c}
739: \begin{aligned}
740: x_1&=-\sqrt{(H-G)G}\sin g\mbox{, }&
741: x_2&=\sqrt{(D-H)G}\sin (h+g)\,,\\[3pt]
742: x_3&=-\sqrt{(D-H)(H-G)}\sin h\mbox{, }&
743: y_1&=\sqrt{(H-G)G}\cos g\,,\\[3pt]
744: y_2&=\sqrt{(D-H)G}\cos (h+g)\mbox{, }&
745: y_3&=\sqrt{(D-H)(H-G)}\cos h\,,
746: \end{aligned}
747: \medskip\\
748: h_1=D-H,\qquad h_2=H-G,\qquad h_3=G,
749: \end{array}
750: \end{equation}
751: where~$D=\frac{1}{4}\sum_{kl}M_{kl}$ is a constant of the Casimir function~$u(3)$,
752: in this connection $0<G<H<D$.
753:
754: The Hamiltonian in new variables can be represented in the form
755: \begin{equation}
756: \label{vort6}
757: \begin{aligned}
758: {\cal H}={}&\frac{1}{4\pi}\Bigl\{\ln\bigl((D-G)^2-4(D-H)(H-G)\cos^2h\bigr)+{}\\
759: {}&+\ln\bigl((D-H+G)^2-4(D-H)G\cos^2(h+g)\bigr)+{}\\
760: {}&+\ln\bigl(H^2-4(H-G)\cos^2g\bigr)\Bigr\}\,.
761: \end{aligned}
762: \end{equation}
763:
764: Let us construct the Poincar\'e map on a plane~$(g,G)$ for different values
765: of the energy~$E ={\cal H}(h,H,g,G)$. Surfaces of energy function in
766: figures~2--4
767: for fixed~${g=g_0}$, ${G=G_0}$, $f(h,H)={\cal
768: H}(h,H,g,G)$ show the complex structure of the isoenergetic surface.
769:
770: We define an intersecting plane by the equation~$H=H_*$. The value~$H_*$
771: should be chosen so that~$\forall g_0,G_0<H_*$ equation~$f(h,H_*)=E$
772: has the unique solution with positive (negative)
773: deri\-va\-tion~$\frac{dH}{dt}$. As is obvious from the presented
774: Fig.~2--4, that it is fulfilled not for all
775: $H_*$ (analogous conditions for the intersecting plane
776: $h=h_*$ are practically never fulfilled).
777:
778: The maximal possible energy~$E_T=-4\ln2\simeq -2.77\ldots$ corresponds
779: to the Tompson configuration for~$D=1$, in this connection~$H=G=\frac{1}{2}$.
780: In this case
781: the phase portrait on a plane~$(g,G)$ consists of the only
782: line~$G=\frac{1}{2}$. Phase portraits for the energies~$ E<E_T$ are
783: represented in Fig.~5--10.
784: Regions where the motion is impossible~($f(h,H_*)=E$ has no solutions)
785: corresponds to painted regions in Fig.~5--10.
786: Figures show that with decreasing of the energy the stochastic layer increases
787: first
788: occupying all plane (Fig.~7, 8). Then it decreases
789: and remains only near unstable solutions and
790: sepa\-ra\-tri\-ces~(Fig.~9, 10).
791: In the limit~$E\to -\infty$ one of pairs
792: of vortices merges and we obtain an integrable problem that is
793: the system of three vortices.
794:
795: \subsection{Lax--Heisenberg representation}
796: As a result of the reduction we can write equations of motion
797: on the orbit of the %%%
798: coadjoint representation of the Lie
799: algebra $u(p,q)$. This orbit is singular and consists of matrices of the
800: form $$ L = \frac{i}{2} {\bf z \bar \bf z}^{\text {T}}\,, $$ where $\sum z_i =
801: 0$ (in a system connected with the center of vorticity).
802:
803: According to a general principle (because of the simplicity
804: of the %%%
805: corresponding Lie algebra~\cite{BolsBor})
806: we can rewrite equations in Lax--Heisenberg form:
807: $$
808: \frac {dL}{dt} = [L, A]\,,
809: $$
810: where $A=dH(L)$ is the %%%
811: differential of the %%%
812: Hamiltonian of the form
813: $$
814: H(L) = \sum \Gamma_i \Gamma_j \log M_{ij}(L)\,.
815: $$
816:
817: Here $M_{ij}$ is interpreted as an element of
818: the %%%
819: Lie algebra
820: $u(n-1)$ and $M_{ij}(L)$~ is a standard pairing between the algebra and
821: coalgebra. In our case the identification $M_{ij}(L)=\pm \mbox{Tr} M_{ij} L =
822: |z_i - z_j|^2$ is fulfilled.
823:
824: The explicit form for the differential of
825: the %%%
826: Hamiltonian is given by the formula:
827: $$
828: dH(L) = \sum \Gamma_i \Gamma_j \frac {1}{ M_{ij}(L)} M_{ij}\,,
829: $$
830: i.~e., $dH(L)$ is a linear combination of matrices $M_{ij}$.
831: Thus the expression for the matrix ~$A$ has the form
832: $$
833: A = dH(L) = i \left( \begin{matrix}
834: \sum_j \frac {\Gamma_1 \Gamma_j}{|z_1 - z_j|^2} &
835: -\frac {\Gamma_1 \Gamma_2}{|z_1 - z_2|^2} & \ldots &
836: -\frac {\Gamma_1 \Gamma_n}{|z_1 - z_n|^2} \\[5pt]
837: -\frac {\Gamma_2 \Gamma_1}{|z_2 - z_1|^2} &
838: \sum_j \frac {\Gamma_2 \Gamma_j}{|z_2 - z_j|^2} & \ldots &
839: -\frac {\Gamma_2 \Gamma_n}{|z_2 - z_n|^2} \\[5pt]
840: \vdots & \vdots & \ddots & \vdots \\
841: -\frac {\Gamma_n \Gamma_1}{|z_n - z_1|^2} &
842: \ldots &
843: -\frac {\Gamma_n \Gamma_{n-1}}{|z_n - z_{n-1}|^2} &
844: \sum_j \frac {\Gamma_n \Gamma_j}{|z_2 - z_j|^2}
845: \end{matrix} \right).
846: $$
847:
848: If all intensities coincide with each other and
849: equal to 1, the matrix
850: $A$ is simplified:
851: $$
852: A = i \left( \begin{matrix}
853: \sum_j \frac {1}{|z_1 - z_j|^2} &
854: -\frac {1}{|z_1 - z_2|^2} & \ldots &
855: -\frac {1}{|z_1 - z_n|^2} \\[5pt]
856: -\frac {1}{|z_2 - z_1|^2} &
857: \sum_j \frac {1}{|z_2 - z_j|^2} & \ldots &
858: -\frac {1}{|z_2 - z_n|^2} \\[5pt]
859: \vdots & \vdots & \ddots & \vdots \\
860: -\frac {1}{|z_n - z_1|^2} &
861: \ldots &
862: -\frac {1}{|z_n - z_{n-1}|^2} &
863: \sum_j \frac {1}{|z_2 - z_j|^2}
864: \end{matrix} \right).
865: $$
866:
867: In case of different intensities the additional procedure is required
868: to reduce the commutator to the standard form.
869:
870: \subsection{Stationary configurations}
871: In terms of $L-A$ pair the stationary configuration is naturally described.
872: The stationary condition is equivalent to
873: the fact that the matrices~$L$ and~$A$ commute:
874: $[{L},{A}]=0$. In our case it means that
875: $$
876: \frac {i}{2} A{\bf z}\bar {\bf z}^{\text{T}} = \frac {i}{2} {\bf z} \bar {\bf z}^{\text{T}}
877: A =
878: -\frac {i}{2} {\bf z} (\bar {A{\bf z}})^{\text{T}} .
879: $$
880:
881: Denoting ${\boldsymbol b}=A{\bf z}$, one has
882: ${\boldsymbol b}\bar {\bf z}^{\text{T}} = - {\bf z} \bar {\boldsymbol b}^{\text{T}}$.
883: It can be shown that it is possible only if $\boldsymbol b =
884: i\lambda {\bf z}$, $\lambda \in \Bbb R$.
885: It means that ${\bf z}$ is an eigenvector of a matrix
886: $A$ (with an imaginary eigenvalue).
887:
888: As a result we obtain a rather natural set of relations
889: $$
890: \left( \begin{matrix}
891: \sum_j \frac {1}{|z_1 - z_j|^2} &
892: -\frac {1}{|z_1 - z_2|^2} & \ldots &
893: -\frac {1}{|z_1 - z_n|^2} \\
894: -\frac {1}{|z_2 - z_1|^2} &
895: \sum_j \frac {1}{|z_2 - z_j|^2} & \ldots &
896: -\frac {1}{|z_2 - z_n|^2} \\
897: \vdots & \vdots & \ddots & \vdots \\
898: -\frac {1}{|z_n - z_1|^2} &
899: \ldots &
900: -\frac {1}{|z_n - z_{n-1}|^2} &
901: \sum_j \frac {1}{|z_2 - z_j|^2}
902: \end{matrix} \right)
903: \left( \begin{matrix} z_1 \\ z_2 \\ \vdots \\ z_n \end{matrix} \right) = \lambda
904: \left( \begin{matrix} z_1 \\ z_2 \\ \vdots \\ z_n \end{matrix} \right).
905: $$
906: This can be rewritten in a simpler form
907: \begin{equation}
908: \label{bolsin*}
909: \sum_j \frac {1}{z_k - z_j} = \lambda \bar z_k\mbox{, where } k=1, \dots, n.
910: \end{equation}
911:
912: Equations~(\ref{bolsin*}) can be obtained directly from
913: the %%%
914: equations of motion under the condition that every point rotates around
915: the origin with the same velocity~$\lambda$.
916:
917: Stationary configurations have been studied in several works
918: (see~\cite{ArefH,Campbell,Melesh}). Apparently new results can be obtained
919: with the help of the following arguments.
920:
921: \begin{itemize}
922: \item[1.] Stationary configurations can be interpreted as eigenvectors of
923: a matrix $A$.
924: \item[2.] Stationary configurations can be interpreted as singular points
925: of Hamiltonian on an orbit (which is on $\mathbb{C}P^{n-1}$). As a Hamiltonian
926: one can take a function $\tilde H =
927: \prod_{k\ne j} |z_k - z_j|^2$. It is a positive function which vanishes
928: on a submanifold of the ``collapse". \end{itemize}
929:
930: %\fig{vihri}
931: Using these arguments and by the investigation of a commutativity condition
932: $[{L,A}]=0$, it would be interesting to give a theoretical explication of
933: configurations from the ``Los-Alamos" catalogue.
934: For these configurations stable states of rotation
935: realize on concentric circles
936: (``atomic covers" by Kelvin) (see Fig.~11), where for the system of
937: 11 vortices two
938: possibilities~$11=2+9$ or $11=3+8$)~\cite{Aref2} are pointed out.
939: As a rule these configurations possess the same type of symmetry (a rotational one
940: or have a plane of symmetry). Recently in a short note in~``Nature"~\cite{ArefV}
941: non-symmetric stationary configurations have been indicated for the vortex
942: system of equal intensities.
943:
944: Theoretically the simplest problem is to determine a number of collinear
945: configurations in depen\-dence on the ratio of intensities. In celestial mechanics
946: (for positive masses~$m_i$) the answer is given by the Multon theorem.
947: In accordance with this theorem there is the only (rotating) collinear
948: configuration
949: for any permutation of masses~$m_1, \dots, m_n$ $(m_i>0)$
950: (the proof of this theorem is contained in~\cite{Smeil}). For the three-vortex
951: system
952: the number of collinear configurations depends on the type of the algebra
953: defined by Poisson brackets \cite{BorisovLebedev,BorisovLeb}.
954: Apparently there is such a connection in general case
955: (for the system of~$n$ vortices).
956:
957: \section{Solvable problems of vortex dynamics}
958: In this section we consider in
959: parallel the problems of dynamics of point vortices on a plane and sphere
960: \cite{Bog2, BPavlov}. The position of vortices on a sphere in the "absolute" space
961: is defined in cartesian~${\bf r}_i (x_i, y_i, z_i) $ or spatial polar
962: coordinates~$ (\theta_i, \varphi_i, \, i=1.. N) $
963: \[
964: x_i=R\sin \theta_i \cos \varphi_i \,,\qquad
965: y_i=R\sin \theta_i \sin \varphi_i \,,\qquad
966: z_i=R\cos \theta_i \,,
967: \]
968: where $R $ is the radius of a sphere. Let us remind (see \cite{BPavlov}), that
969: the Hamiltonian function and Poisson bracket are defined by relations
970: \[
971: \begin{aligned}
972: {}\span
973: H = -\frac{1}{8 \pi} \sum\limits _{i < j} \Gamma_i \Gamma_j \ln 4R^2 \sin^2
974: \frac{\theta _{ij}}{2} \,, \\
975: {}\span
976: \{\varphi_i, \cos \theta_j \} =\frac{1}{R^2
977: \Gamma_i} \delta _{ij} \,, \qquad
978: i, j=1, \ldots, N \,,
979: \end{aligned}
980: \]
981: where $ \theta _{ij} $ is the angle between $i $-th and $j $-th vortices, and $
982: \Gamma_i $ is a vortex intensity. For relative variables on a sphere the same
983: labels --- $M _{ij} $, $ \Delta _{ijk} $ are used, however, their meaning is
984: different, than for a plane~\cite{BPavlov}. So $M _{ij} $ is a length of a
985: chord connecting the vortices, and $ \Delta _{ijk} $ is equal to the ratio of
986: volume of a tetrahedron spanned on three vortices and the center of a sphere
987: to~$\frac{1}{4} R $.
988: \begin{equation} \label{c1}
989: \begin{aligned}
990: M_{ij}=&{}(x_i-x_j)^2+(y_i-y_j)^2+(z_i-z_j)^2\,,\\
991: \Delta_{ijk}=&{}\frac{1}{R}({\bf r} _i{, \bf r} _j\times{\bf r} _k) \,.
992: \end{aligned}
993: \end{equation}
994:
995: \subsection{The particular case of the problem of $n$ vortices, the reduction
996: to the problem of ${(n-1)}$ vortices}
997:
998: There is a special case of the problem of $n$ vortices on a plane and a sphere,
999: for that case the problem can be reduced to the system of $(n-1)$ vortices with
1000: the same algebra of Poisson brackets, but reduced Hamiltonian function.
1001: The procedure of a reduction in Hamiltonian exposition corresponds to
1002: restriction of the system on a Poisson submanifold \cite{Fomenko2}, which is
1003: defined in this case by involution conditions of integrals of motion.
1004: For vortices on a plane the integrals are given by relations~(\ref{new2}),
1005: necessary conditions of reduction accept the form
1006: \begin{equation}
1007: \label{Plane25.1}
1008: \sum _{i=1} ^N\Gamma_i=0 \,,\quad Q=P=0 \,.
1009: %, \quad\Rightarrow\quad D=0\,.
1010: \end{equation}
1011: For a sphere integrals of motion are obtained similarly:
1012: \begin{equation}
1013: \label{Sp5}
1014: \begin{gathered}
1015: F_1 = R\sum_{i=1}^N\Gamma_{i}\sin\theta_{i}\cos\varphi_{i}\,,\quad
1016: F_2=R\sum_{i=1}^N\Gamma_{i}\sin\theta_{i}\sin\varphi_{i}\,,\quad
1017: F_3 = R\sum _{i=1} ^N\Gamma _{i} \cos\theta _{i}\,,
1018: \end{gathered}
1019: \end{equation}
1020: and conditions of reduction
1021: \begin{equation}
1022: \label{Plane25.2}
1023: F _{1} =F _{2} =F _{3} =0\,.
1024: % \quad\Leftrightarrow\quad D=D _{\rm max} =2\left (R\sum\Gamma_i\right) ^2.
1025: \end{equation}
1026:
1027: {\bf Remark 6.}{\it
1028: The equations (\ref{Plane25.1}) have the following geometrical meaning: each
1029: vortex is situated at the center of vorticity of all remaining vortices.
1030: Really, expressing for example $\displaystyle\Gamma_N=-\sum_{i=1}^{N-
1031: 1}\Gamma_i$ from (\ref{Plane25.1}), we have found that the absolute coordinates
1032: of the $N-$th vortex are defined by expressions}
1033: $$
1034: {\bf r}_N=\frac{\sum\limits_{i=1}^{N-1}\Gamma_i{\bf r} _i}{\sum\limits _{i=1}
1035: ^{N-1} \Gamma_i}\,.
1036: $$
1037: {\it The geometrical meaning of relations (\ref{Plane25.2}) on a sphere is a little
1038: different, it means, that the center of vorticity of a system of $n$ vortices
1039: coincides with the geometrical center of a sphere.}
1040:
1041: {\it The mentioned below reduction happens because there is a possibility of
1042: calculation of positions of~$n$ vortices basing on positions of~$(n-1)$
1043: vortices.}
1044:
1045: For a determination of additional invariant relations for relative variables we
1046: shall use identities~\cite{Melesh}
1047: \begin{equation}
1048: \label{Plane25.3}
1049: \frac {1}{2}\sum _{k=1} ^N\Gamma_k (M _{jk} -M _{ik}) =
1050: P (x_i-x_j) +Q (y_i-y_j)
1051: \end{equation}
1052: for plane which are fulfilled only under condition of $\displaystyle
1053: \smash{\sum_{i=1}^N}\Gamma_i=0$, that is also valid for the case of sphere~\cite{BPavlov}
1054: \begin{equation}
1055: \label{Plane25.4}
1056: \frac {1}{2}\sum _{k=1} ^N\Gamma_k (M _{jk} -M _{ik}) =
1057: R\bigl(F_{1}(x_i-x_j)+F_{2}(y_i-y_j)+F_{3}(z_i-z_j)\bigr)\,.
1058: \end{equation}
1059:
1060: Using conditions (\ref{Plane25.1}), (\ref{Plane25.2}), we have obtained
1061: $C_N^2$ invariant relations for $M_{ij}$
1062: \begin{equation}
1063: \label{Plane25.5}
1064: \sum _{k=1} ^N\Gamma_k (M _{jk} -M _{ik}) =0\,.
1065: \end{equation}
1066: These relations we shall supplement by relations for $\Delta_{ijk}$, which also
1067: follow from (\ref{Plane25.1}), (\ref{Plane25.2}). It is possible to show by a
1068: straightforward calculation that the complete set of these relations determines a
1069: Poisson submanifold with the structure~(\ref{Plane7}) and its generalization for a
1070: sphere \cite{BPavlov}.
1071:
1072: Using representation for $M_{ij}$ in absolute coordinates (\ref{Bol3/2})
1073: and~(\ref{c1}), it is simple to verify that among the equations~(\ref
1074: {Plane25.5}) only $n-1$ are linearly independent. Using these equations, we can
1075: express squares of distances from one of the vortices ($n$th vortex) to all
1076: the others~$M_{kn},\; k=1,\ldots,n-1$ via mutual distances between
1077: $n-1$ vortices~$M_{ij}, \; i,j=1,\ldots,n-1$.
1078: Then substituting these expressions into initial Hamiltonian, we shall obtain a
1079: system with a bracket of the problem of~$n-1$ vortices and reduced
1080: Hamiltonian function.
1081:
1082: For the case of four vortices the explicit formulas of squares of distances
1083: from the first three vortices up to the fourth vortex $M_{14}, M_{24}, M_{34}$
1084: have
1085: a form
1086: \begin{equation}
1087: \label{Plane25.6}
1088: \hookrightarrow
1089: M_{14}=\frac{\Gamma_3^2M_{13}+\Gamma_2^2M_{12}+ \Gamma_2\Gamma_3 (M _{12} -M _
1090: {23} +M _{13})}{
1091: ( \Gamma_1 +\Gamma_2 +\Gamma_3) ^2}\,,
1092: \end{equation}
1093: the reduced Hamiltonian is obtained by substitution (\ref{Plane25.6}) into
1094: initial Hamiltonian (\ref{new1}).
1095:
1096: The system of three vortices is integrable (independently on a Hamiltonian
1097: given on al\-geb\-ra~(\ref{Plane7})), the described case corresponds to a special
1098: case of integrability of problem of four vortices (origina\-ting in work by
1099: Kirchhoff~\cite{Kirh}).
1100: The phase portraits of this system at various intensities are presented in the
1101: previous work \cite{BorisovLeb}.
1102:
1103: {\bf Remark 7.} {\it
1104: The indicated special cases of integrability correspond to a situation at
1105: which one of integrals reaches the extreme value. It is obvious, that in this
1106: case the system necessarily has additional invariant relations. For integrable
1107: systems that situation produces an additional degeneration. Delone case for
1108: Kovalevskaya top is one of examples. In this case the Kovalevskaya integral, which is
1109: the sum of two full squares will vanish, and the two-dimensional tori degenerate
1110: into one-dimensional torus(periodic and asymptotic solutions).}
1111:
1112: \subsection{Particular solutions in the problem of four vortices}
1113: General equations of motion of vortices~\cite{BPavlov} with some restrictions on
1114: intensities~$\Gamma_i$ allow a finite group of symmetries, the elements of the
1115: group are the permutations and reflections in some planes. Such discrete
1116: symmetries do not produce the general integrals of motion and do not allow
1117: reducing the order of a system. However, presence of these symmetries is
1118: connected to invariant submanifolds. The solution on that submanifolds can, as a
1119: rule, be obtained in quadratures~\cite{Melesh}.
1120:
1121: Let us consider two problems of dynamics of four vortices on a plane and a sphere
1122: possessing two types of symmetry --- central symmetry and reflection symmetry
1123: (for a plane, central symmetry and axial symmetry).
1124:
1125: \noindent{\bf a. A centrally symmetric solution at $\bf {D=0}$.} Equations
1126: of motion
1127: of four vortices on sphere (flat case is obtained with the passage to the
1128: limit~$R\to\infty$) with condition~$\Gamma_1 =\Gamma_3,$ $ \Gamma_2 =\Gamma_4$
1129: allow invariant relations
1130: \begin{equation}
1131: \label{Plane34}
1132: \begin{array}{cc}
1133: M _{14} =M _{23} =M _{14} =M_1 \,,\quad &M_{12}=M_{34}=M_3\,,\\
1134: \Delta_{234}=\Delta_{124}=\Delta_{1}\,,\quad &\Delta_{213}= \Delta _{314} =
1135: \Delta _{2} \,.
1136: \end{array}
1137: \end{equation}
1138: (The relations~(\ref{Plane34}) don't define the Poisson submanifold).
1139:
1140: %\wfig*{ivanw}
1141:
1142: The equations~(\ref{Plane34}) have the following geometrical meaning: a
1143: centrally symmetric configuration of vortices (the parallelogram), keeps this
1144: symmetry in all moments (see Fig.~12).
1145:
1146: The analysis of a centrally symmetric solution in absolute variables through
1147: explicit quadratures is carried out by D.\,N.\,Gorya\-chev~\cite{Goryachev1}
1148: (see also~\cite{ArefPoint}). However, he has not made clear the qualitative
1149: properties of motion and has introduced the very confusing classification. Let us
1150: introduce the qualitative analysis of relative motion.
1151:
1152: The equations describing evolution of sides $M_1, M_3$ and diagonals
1153: $M_{13}=M_2 $, $M_{24}=M_4$ of the parallelogram, have a form
1154: \begin{equation}
1155: \label{Plane35}
1156: \begin{array}{l}
1157: \dot M_{1}=4\Gamma_1\Delta_2\left(\frac{1}{M_2}-\frac{1}{M_3}\right)
1158: + 4\Gamma_2\Delta_1\left(\frac{1}{M_4}-\frac{1}{M_3}\right)\,,\\[10pt] \dot
1159: M_{3}=4\Gamma_1\Delta_2\left(\frac{1}{M_1}-\frac{1}{M_2}\right)
1160: + 4\Gamma_2\Delta_1\left(\frac{1}{M_1}-\frac{1}{M_4}\right)\,,\\[10pt] \dot
1161: M_{2}=8\Gamma_2\Delta_2\left(\frac{1}{M_3}-\frac{1}{M_1}\right)\,,
1162: \\ [10pt] \dot
1163: M_{4}=8\Gamma_1\Delta_1\left(\frac{1}{M_3}-\frac{1}{M_1}\right)\,.
1164: \end{array}
1165: \end{equation}
1166:
1167:
1168: For a sphere (for a plane $R\to \infty $) the geometrical relations between
1169: $M, \Delta $ can be written as
1170: \begin{equation}
1171: \label{cen3}
1172: \begin{array}{l}
1173: \Delta_1\left(4R^2-M_2\right)+\Delta_2\left(4R^2-M_1-M_3\right)=0\,,\\[6pt]
1174: \Delta_1\left(4R^2-M_1-M_3\right)+\Delta_2\left(4R^2-M_4\right)=0\,.\\[6pt]
1175: \end{array}
1176: \end{equation}
1177: System~(\ref{cen3}) is solvable with respect to $ \Delta_1, \Delta_2$ under
1178: condition of
1179: \begin{equation}
1180: \label{cen4}
1181: 2(M_1+M_3)-(M_2+M_4)+\frac{1}{4R^2}\left(M_2M_4-(M_1+M_3)^2\right)=0\,. \end
1182: {equation}
1183: The linear integral of the equations~(\ref{Plane35}) corresponding to the Casimir
1184: function (\ref{Plane10}) has a form
1185: \begin{equation}
1186: \label{cen5}
1187: D=2\Gamma_1\Gamma_2(M_1+M_3)+\Gamma_1^2M_2+\Gamma_2^2M_4\,.
1188: \end{equation}
1189:
1190: With the help of relations~(\ref{cen3}) and regularization of time, the
1191: system~(\ref{Plane35}) can be reduced to two inhomogeneous equations describing
1192: evolution of sides of the parallelogram~$M_1,\,M_3$.
1193:
1194: For simplicity we consider the limited case~$D=0$, that is also a necessary
1195: condition of a collapse~\cite{NovikovSed}. Geometrical interpretations on a
1196: plane and a sphere are a little different, therefore we shall consider these cases
1197: separately.\medskip
1198:
1199: 1) P\,l\,a\,n\,e.
1200:
1201: Relations (\ref{cen3}),(\ref{cen4}) for a plane ($R\to\infty $) have a form
1202: \begin{equation}
1203: \label{cen6}
1204: \begin{array}{c}
1205: \Delta_1 =-\Delta_2 =\Delta \,, \\ [6pt]
1206: 2 (M_1+M_3) - (M_2+M_4) =0 \,.
1207: \end{array}
1208: \end{equation}
1209: Taking into account, that in (\ref{cen5}) $D=0$, we found the equation,
1210: describing a trajectory of system on plane $M_1, M_3 $ (sides of a
1211: parallelogram)
1212: \begin{equation}
1213: \label{cen8}
1214: \frac{dM_1}{dM_3} = -\frac
1215: {M_1\left(2\Gamma_1\Gamma_2(M_1+M_3)+(\Gamma_1^2+\Gamma_2^2)M_3\right)}
1216: {M_3\left(2\Gamma_1\Gamma_2(M_1+M_3)+(\Gamma_1^2+\Gamma_2^2)M_1\right)}\,.
1217: \end{equation}
1218: The solution of this equation has a form
1219: \begin{equation}
1220: \label{cen9}
1221: \begin{array}{c}
1222: \left(M_1-M_3\right)^2=\left(M_1+M_3\right)^2 -
1223: C\left(M_1+M_3\right)^{-2\alpha}\,,\\[6pt]
1224: \alpha=\frac{\Gamma_1^2+\Gamma_2^2}{4\Gamma_1\Gamma_2}\,,\qquad C =const \,.
1225: \end{array}
1226: \end{equation}
1227: (Index $\alpha<0$, because with $D=0$ intensities $\Gamma_1, \Gamma_2$ have
1228: different signs).
1229:
1230:
1231: In a quadrant $M_1 > 0, \, M_2 > 0 $ (which corresponds to physical area)
1232: three types of trajectories~(\ref{cen9}) are possible depending on $\alpha$:
1233: \begin{itemize}
1234: \item [$ 1^\circ.$] $-1<\alpha<0$~--- all trajectories are closed, started from
1235: the origin of coordinates, tangenting axes $OM_1 $ $OM_2 $ (see. Fig.~13, a),
1236: a) (inhomogeneous collapse).
1237: \item [$ 2^\circ.$] $\alpha <-1$~--- all trajectory are curves asymptotically
1238: approaching to coordinate axes (see Fig.~13,b).
1239: \item [$ 3^\circ.$] $\alpha =-1$~--- all trajectories are straight lines moving
1240: through the origin of coordinates (see Fig.~13,c) (homogeneous
1241: collapse)~\cite{NovikovSed}.
1242: \end{itemize}
1243:
1244: The physical area on a plane~$ M_1,M_3$ is defined by the part of a positive
1245: quadrant~$(M_1>0$, ${M_3>0)}$, for which the inequality~$\Delta^2>0$ is fulfilled.
1246: When the trajectory reached a boundary~${\Delta^2=0}$, the sign in the
1247: equation~(\ref{cen8}) should be changed (reflection) that corresponds to the
1248: same trajectory passing in the opposite direction. It is easy to show, that the
1249: equation~$\Delta^2=0$ is defined by two straight lines on a plane~$M_1,M_3$,
1250: which are situated inside a quadrant ~ $ M_1 > 0, \, M_3 > 0 $ for any
1251: intensities~$\Gamma_1, \Gamma_2$. Hence, except for a case~$3^\circ$ vortices
1252: are moving in the bounded area without collisions and scattering.
1253: In the system in case~$3^\circ$ the homogeneous collapse of all vortices or
1254: homogeneous scattering happen.
1255:
1256: For the problem on a plane each point of the trajectory on the plane~$M_1,M_3$
1257: correspond to two configurations of vortices, which differ only by
1258: permutation:~$1\leftrightarrow 3$, or~$2\leftrightarrow4$ (see
1259: Fig.~12).
1260:
1261: {\bf Remark 8.} {\it
1262: The condition of a homogeneous collapse ($\alpha=-1$) from the analysis of
1263: motion in absolute variables is obtained in~\cite{NovikovSed}. As
1264: result of
1265: quasi-homogeneity of the equations the condition of a homogeneous collapse can
1266: be obtained by investigation of existence conditions of solutions of the
1267: form $M_i=C_i t^{\xi},\, C_i=const$ of the system.}
1268:
1269: {\bf Remark 9.} {\it
1270: For a plane the system~(\ref{Plane35}) with a regularization}
1271: $$
1272: d\tau =\frac{4\Delta}{M_1M_2M_3M_4} \, dt
1273: $$
1274: {\it can be reduced to a homogeneous Hamiltonian system}
1275: \begin{align}
1276: \frac{dM_1}{d\tau}&=\Gamma_1M_1M_4(M_3-M_2)+\Gamma_2M_1M_2(M_4-M_3)\,,\\[-3pt]
1277: \frac{dM_3}{d\tau}&=\Gamma_1M_3M_4(M_2-M_1)+\Gamma_2M_2M_3(M_1-M_3)\,,\\[-3pt]
1278: \frac{dM_2}{d\tau}&=2\Gamma_2M_2M_4(M_1-M_3)\,,\\[-3pt]
1279: \frac{dM_4}{d\tau}&=2\Gamma_1M_2M_4(M_3-M_1)
1280: \end{align}
1281: {\it with a Poisson bracket of the form}
1282: \begin{align}
1283: \{M_1,M_2\}&=\frac{1}{\Gamma_1}M_1M_2M_3M_4\,,\quad&\{M_1,M_3\}&=
1284: \frac12\left(\frac{1}{\Gamma_1}-\frac{1}{\Gamma_2}\right)M_1M_2M_3M_4\,,\notag\\[-3pt
1285: ] \label{cencen}
1286: \{M_2,M_3\}&=\frac{1}{\Gamma_1}M_1M_2M_3M_4\,,\quad&\{M_1,M_4\}&= -\frac1
1287: {\Gamma_2} M_1M_2M_3M_4\,, \\[-3pt]
1288: \{M_2,M_4\}&=0,\quad&\{M_3,M_4\}&=
1289: \frac1{\Gamma_2} M_1M_2M_3M_4\notag
1290: \end{align}
1291: {\it and Hamiltonian}
1292: $$
1293: H=2\Gamma_1\Gamma_2\ln M_1 +\Gamma_1^2\ln M_2+2\Gamma_1\Gamma_2\ln M_3 +
1294: \Gamma_2^2\ln M_4\,.
1295: $$
1296:
1297: {\it The rank of a bracket~(\ref{cencen}) is equal to two, hence dividing the
1298: bracket on~$M_1M_2M_3M_4$, it can be reduced to a constant without violation of
1299: the Jacobi identity.}
1300: \medskip
1301:
1302: 2) S\,p\,h\,e\,r\,e.
1303:
1304: For a centrally symmetric configuration on a sphere we also consider a
1305: projection of trajectories to a plane $M_1,M_3$ (see Fig.~14).
1306: The form of physical area defined by inequalities $\Delta_i^2>0,\,i=1,2$
1307: varies. Moreover, there are two differences from a case of plane connected with the
1308: nonlinearity of the equation (\ref{cen4}). At first, each point $M_1,M_3$
1309: correspond to two solutions of the equation (\ref{cen4}), and therefore two
1310: various (spatial) parallelograms with the given sides on a sphere, these
1311: parallelograms are not reduced to each other by permutation~$1\leftrightarrow
1312: 3\;(2\leftrightarrow 4)$. Secondly, the system~(\ref{Plane35}) is not reduced
1313: to two quasi-homogeneous equations.
1314:
1315: Let us express~$\Delta_1$ and $\Delta_2$ from~(\ref{cen3}) with preservation of
1316: homogeneity
1317: \begin{equation}
1318: \label{regul1}
1319: \Delta_1 = (M_4-M_1-M_3) \Delta \,,\qquad
1320: \Delta_2 = (M_2-M_1-M_3) \Delta \,.
1321: \end{equation}
1322: Also we shall make change of the time
1323: $$
1324: dt=\smash{\frac{4\Delta}{M_1M_3}}\,d\tau\,.
1325: $$
1326: So we obtained the system describing evolution of
1327: variables~$M_1$,~$M_2$,~$M_3$,~$M_4$ \vspace{-2mm}
1328: \begin{equation}
1329: \label{evolut_*}
1330: \begin{aligned}
1331: M'_1&=M_1\left(\Gamma_1\frac{(M_2-M_1-M_3)(M_3-M_2)}{M_2}
1332: +\Gamma_2
1333: \frac{(M_4-M_1-M_3)(M_3-M_4)}{M_4}\right)\,,\\
1334: M'_3&=M_3\left(\Gamma_1\frac{(M_2-M_1-M_3)(M_2-M_1)}{M_2}
1335: +\Gamma_2
1336: \frac{(M_4-M_1-M_3)(M_4-M_1)}{M_4}\right)\,,\\
1337: M'_2&=2\Gamma_2(M_2-M_1-M_3)(M_1-M_3)\,,\\
1338: M'_4&=2\Gamma_1(M_4-M_1-M_3)(M_1-M_3)\,,
1339: \end{aligned}
1340: \end{equation}
1341: here primes denote derivation on new time $\tau$.
1342:
1343: Let us introduce a linear change of variables reducing system (\ref{evolut_*})
1344: to two differential equations, which has the most simple form.
1345:
1346: Let us choose variables $x, y, M, N $:
1347: \begin{equation}
1348: \label{perem_*}
1349: \begin{aligned}
1350: x&=\Gamma_1M_2+\Gamma_2M_4\,,\qquad & y&=\Gamma_1^2M_2-\Gamma_2^2M_4\,,\\
1351: M&=M_1+M_3\,,\qquad & N&=M_1-M_3\,.
1352: \end{aligned}
1353: \end{equation}
1354: With the help of a relation (\ref{cen5}) where $D=0$ and (\ref{cen4}) which have
1355: a form
1356: \begin{equation}
1357: \label{1a}
1358: x=-{\frac{y^2}{16R^2\Gamma_1\Gamma_2(\Gamma_1+\Gamma_2)}}\,,
1359: \end{equation}
1360: we exclude $x, M $ from~(\ref{evolut_*}). In outcome we obtain the equations
1361: of evolution $y, N $\vspace{-1mm}
1362: \begin{align}
1363: \label{eqq}
1364: \frac{dy}{d\tau}=&\,4(\Gamma_1+\Gamma_2)yN\,,\notag\\
1365: \frac{dN}{d\tau}=&-\frac{1}{4(\Gamma_1+\Gamma_2)^3\Gamma_1^2
1366: \Gamma_2^2}\biggl[\frac{2ky-\Gamma_1(\Gamma_1+\Gamma_2)}{\Gamma_2
1367: (ky-\Gamma_1(\Gamma_1+\Gamma_2))}\Bigl(4\Gamma_1^3\Gamma_2^2(\Gamma_1 +
1368: \Gamma_2) ^4N^2-\notag \\
1369: &-y^2(2ky-\Gamma_1^2+\Gamma_2^2)(2ky(\Gamma_1+2\Gamma_2) -
1370: \Gamma_1(\Gamma_1+3\Gamma_2)(\Gamma_1+\Gamma_2))\Bigr)+\notag\\
1371: &+\frac{2ky+\Gamma_2(\Gamma_1+\Gamma_2)}{\Gamma_1
1372: (ky+\Gamma_2(\Gamma_1+\Gamma_2))}\Bigl(4\Gamma_1^2\Gamma_2^3(\Gamma_1 +
1373: \Gamma_2) ^4N^2- \\[-3pt]
1374: &-y^2(2ky-\Gamma_1^2+\Gamma_2^2)(2ky(\Gamma_2+2\Gamma_1)
1375: +\Gamma_2(\Gamma_2+3\Gamma_1)(\Gamma_1+\Gamma_2))\Bigr)\biggr]\,.\notag \end
1376: {align}
1377: hereinafter $k=1/R^2 $.
1378:
1379:
1380: The projection of trajectory~$y(\tau),\, N(\tau)$ on the plane $M_1,M_3$ is
1381: defined by the formulas $M_1=\frac 12 (M+N),M_3=\frac 12 (M-N)$, where
1382: \begin{equation}
1383: \label{2a}
1384: M=\frac{\Gamma_2-\Gamma_1}{2\Gamma_1\Gamma_2(\Gamma_1+\Gamma_2)}y+
1385: \frac{y^2}{16R^2\Gamma_1\Gamma_2(\Gamma_1+\Gamma_2)^2}\,,\qquad
1386: (\Gamma_1\Gamma_2 < 0) \,.
1387: \end{equation}
1388: The equation (\ref{2a}) allows real solutions only for
1389: $M\le M_{max}=-\frac{(\Gamma_1-\Gamma_2)^2}{\Gamma_1\Gamma_2}R^2.$ Therefore, in
1390: case of a sphere the physically accessible area on a plane is defined by
1391: inequalities
1392: \begin{gather}
1393: M_1+M_3\le M _{max} \,,
1394: \label{4a}
1395: \\[4pt]
1396: \begin{aligned}
1397: 4\Delta_1^2=&\,2(M_1M_4+M_3M_4+M_1M_3)-M_1^2-M_3^2-M_4^2-\frac{1}{R^2}M_1M_3M_
1398: 4 \ge 0 \,, \\
1399: 4\Delta_2^2=&\,2(M_1M_2+M_3M_2+M_1M_3)-M_1^2-M_3^2-M_2^2-\frac{1}{R^2}M_1M_3M_
1400: 2\ge 0 \,.
1401: \end{aligned}
1402: \label{3a}
1403: \end{gather}
1404:
1405: In the area (\ref{3a}) the equation (\ref{2a}) has two roots, therefore we have
1406: on a plane $M_1,M_3$ a projection of two different areas of possible positions of
1407: vortices defined by inequalities (\ref{4a}) and equations~(\ref{2a}).
1408: If the areas (\ref{4a}) do not reach a straight line $M_1+M_3=M_{max}$
1409: (Fig.~14 b, c), the point, beginning to move inside one area,
1410: remains there in all moments. When the point reached boundaries ($ \Delta_i=0$),
1411: it is necessary to change a sign of time, and the trajectory is passed in the
1412: opposite direction.
1413: In a case when the straight line $M_1+M_3=M_{max} $ passes through the interior
1414: of areas~(\ref{4a}) (Fig.~14 a) the trajectory, that had reached
1415: that line, passes from one area in another and is described by other solution of
1416: the equation (\ref{2a}). Characteristics of phase portraits in case of a sphere
1417: is the occurrence new motionless points, absent in a flat case.
1418: If the second necessary condition of a collapse is realized
1419: ($\Gamma_1^2+\Gamma_2^2=-4\Gamma_1\Gamma_1$), we can see on a Fig.~14
1420: c), that some of trajectories starting at the origin of coordinates goes
1421: to origin again and do not reach the boundary of area (collinear positions) and
1422: some trajectories goes back to origin only after the reaching of boundary. It
1423: means, that the collapse remains homogeneous only near the origin of
1424: coordinates.
1425:
1426:
1427:
1428: \noindent{\bf b. A reflective-symmetric solution}.
1429: For a system of two cooperating vortex pairs~$\Gamma_1=-\Gamma_4$, ${\Gamma_2=-
1430: \Gamma_3}$ the general system of four vortices on a sphere and on a plane also
1431: allows invariant relations
1432: %\wfig{ivane}
1433: \begin{equation}
1434: \label{zerc1}
1435: \begin{array}{cc}
1436: M _{12} =M _{34} =M_1\,, \quad &M_{13}=M_{24}=M_{3}\,,\\[5pt] \Delta _{123} =
1437: \Delta _{234} = \Delta_1\,, \quad &\Delta_{124}=\Delta_{134}= \Delta _{2}\,.
1438: \end{array}
1439: \end{equation}
1440: These equations have the following geometrical meaning. The vortices located at
1441: the initial momentum in tops of a trapezoid will form a trapezoid during all time
1442: of motion (see Fig.~15) \cite{Melesh}. The given configuration has a reflection
1443: (axial) symmetry. The equations describing the evolution of sides and diagonals
1444: of a trapezoid have a form
1445: \begin{equation}
1446: \label{zerc2}
1447: \begin{aligned}
1448: \dot M_1&=4\Gamma_2\Delta_1\left(\frac{1}{M_4}-\frac{1}{M_3}\right)
1449: +4\Gamma_1\Delta_2\left(\frac{1}{M_2}-\frac{1}{M_3}\right)\,,\\
1450: \dot M_3&=4\Gamma_2\Delta_1\left(\frac{1}{M_4}-\frac{1}{M_1}\right)
1451: +4\Gamma_2\Delta_2\left(\frac{1}{M_2}-\frac{1}{M_1}\right)\,,\\
1452: \dot M_2&=8\Gamma_2\Delta_2\left(\frac{1}{M_1}-\frac{1}{M_3}\right)\,,\\ \dot
1453: M_4&=8\Gamma_1\Delta_1\left(\frac{1}{M_1}-\frac{1}{M_3}\right)\,. \end
1454: {aligned}
1455: \end{equation}
1456: Here~$ M_2=M_{14}, \; M_{4}=M_{23}$ --- of the bases of a trapezoid.
1457:
1458: The geometrical relations between~$M,\,\Delta$ in this case have the identical
1459: form on a plane and sphere:
1460: \begin{equation}
1461: \label{zerc3}
1462: \begin{aligned}
1463: \Delta_1M_4+\Delta_2(M_3-M_1)&=0\,,\\
1464: \Delta_1(M_3-M_1)+\Delta_2M_2&=0\,,\\
1465: \end{aligned}
1466: \end{equation}
1467: the condition of their resolvability with respect to $ \Delta_1, \, \Delta_2 $ has a
1468: form
1469: \begin{equation}
1470: \label{zerc4}
1471: M_2M_4- (M_1-M_3) ^2=0 \,.
1472: \end{equation}
1473: The integral of the momentum has a form
1474: \begin{equation}
1475: \label{zerc5}
1476: D=2\Gamma_1\Gamma_2(M_1-M_3)-\Gamma_1^2M_4-\Gamma_2^2M_2\,.
1477: \end{equation}
1478: From the equations~(\ref{zerc2}--\ref{zerc4}) it follows that
1479: trajectories of a system in space~$ M_1,M_2,M_3,M_4$ coincide for cases
1480: of a plane and sphere. The difference between these problems consists
1481: in form of physical areas defined by inequalities~(\ref{4a}).
1482:
1483: With the help of the equations~(\ref{zerc3}) we shall
1484: express~$\Delta_1,\Delta_2 $ by
1485: $$
1486: \Delta_1=(M_2-M_1+M_3)\Delta,\quad\Delta_2=-(M_4-M_1+M_3)\Delta\,,
1487: $$
1488: and also we shall exclude~$\Delta$ with the help of regularizing substitution of
1489: time
1490: $$
1491: dt =\frac{4\Delta}{M_1M_3} d\tau \,.
1492: $$
1493: To reduce the regularized equation to the system of two equations, we make the
1494: substitution~(\ref{perem_*}). As a result of the existence of an integral of
1495: the moment between the variables~$ x, y $ the following relation is fulfilled
1496: \begin{equation}
1497: \label{zerc6}
1498: \frac{4\Gamma_1\Gamma_2}{\Gamma_1+\Gamma_2}Dx+y^2+\frac{2(\Gamma_1\Gamma_2)}
1499: {\Gamma_1+\Gamma_2}Dy+D^2=0\,.
1500: \end{equation}
1501: Using~(\ref{zerc6}) and equation~(\ref{zerc4}), we find
1502: \begin{equation}
1503: \label{zerc7}
1504: N=-\frac{y^2}{4\Gamma_1\Gamma_2D}+\frac{D}{4\Gamma_1\Gamma_2}\,.
1505: \end{equation}
1506: With~$ D\ne0 $ from relations~(\ref{zerc5}),~(\ref{zerc6}) and~(\ref{zerc7})
1507: all~$ M_1, M_2, M_3, M_4 $ can be expressed through~$ M, y $. For the
1508: variables~$M, y $ we shall get the equations of the form
1509: \begin{equation}
1510: \label{zerc8}
1511: \begin{array}{c}
1512: \frac{dy}{d\tau}=\frac{(y^2-D^2)(y(\Gamma_1+\Gamma_2)+D(\Gamma_1-\Gamma_ 2))}
1513: {4\Gamma_1\Gamma_2D}\,, \\[12pt]
1514: \begin{aligned}
1515: \frac{dM}{d\tau}=&\,\frac{1}{32D^2\Gamma_1\Gamma_2^3(D-y)}\left[D(\Gamma_2 -
1516: \Gamma_1)-y(\Gamma_2+\Gamma_1)\bigr)\bigl(16D^2\Gamma_1^2\Gamma_2^2M^2 +
1517: \right. \\[5pt]
1518: &\left.+8(D-y)^2D\Gamma_2^2M-(D^2-y^2)^2\right]-\\[10pt]
1519: &-\frac{1}{32D^2\Gamma_1^3\Gamma_2(D+y)}
1520: \left[D(\Gamma_2-\Gamma_1)-y(\Gamma_2+\Gamma_1))
1521: (16D^2\Gamma_1^2\Gamma_2^2M_2+\right.\\[5pt]
1522: &\left.+8(D+y)^2D\Gamma_1^2M-(D^2-y^2)^2\right]\,.
1523: \end{aligned}
1524: \end{array}
1525: \end{equation}
1526: By analogy with a centrally symmetric solution, we shall consider a projection of
1527: a trajectory to a plane~$ M_1, M_3 $. Besides inequalities~(\ref{4a}) the
1528: physical area is defined by an additional condition (corollary of the
1529: relation~(\ref{zerc7}))
1530: \begin{equation}
1531: \label{zerc9}
1532: D (D-4\Gamma_1\Gamma_2 N) > 0 \,.
1533: \end{equation}
1534: Two various points of a phase space (two solutions of the
1535: equation~(\ref{zerc7})) are projected in each point on a plane~$ M_1, M_3 $,
1536: satisfying to inequalities~(\ref{4a}),~(\ref{zerc9}). It is more evident if we
1537: imagine that two different areas defined by an inequality~(\ref{4a}) are glued
1538: together along a line~$M_1-M_3=N_{0}$. When a point reached a
1539: boundary~$\Delta_i=0$, it is reflected back and moves on the same trajectory,
1540: and at reaching boundary~(\ref{zerc9}) the point passes from one area to
1541: another (see. Fig.~16 a, 16 b).
1542: In figures for convenience we have unrolled the areas, which have been glued
1543: together along a line~$ M_1-M_3=N_0 $, where~$ N_0 $ is a solution of the
1544: equation $D-4\Gamma_1\Gamma_2 N_0=0$.
1545: The difference of a plane from a sphere is exhibited in form of physical areas.
1546: For a plane (see Fig.~16 a) two types of trajectories exist:
1547: \begin{enumerate}
1548: \item Trajectories are touched once boundaries~$\Delta_i=0 $ and then leaving
1549: to infinity;
1550: \item Trajectories stayed between boundaries~$\Delta_i=0 $.
1551: \end{enumerate}
1552: Different motions of vortices correspond to these cases: in the first case the
1553: vortices run up, passing only once through a collinear configuration, in second
1554: case vortex pairs alternately pass one through another ({\it leapfrog of
1555: Helmholtz}), remaining on bounded distance. Only trajectories of the second type
1556: exists for a sphere. The analysis of leapfrog on a plane was carried out by
1557: other method in~\cite{Melesh}.
1558:
1559: Let us introduce also, for completeness, the equation of motion and geometrical
1560: interpretation at the zero moment~$ D=0 $ (necessary condition of a
1561: collapse~\cite{NovikovSed}).
1562:
1563: According to~(\ref{zerc6}) in this case~$ y=0 $, and all mutual
1564: distances~$M_1, \dots, M_4 $ can be expressed via variables~$ x, M $ by
1565: formulas
1566: $$
1567: \begin{array}{c}
1568: M_2=\frac{\Gamma_2x}{\Gamma_1(\Gamma_1+\Gamma_2)}\,,\quad
1569: M_4=\frac{\Gamma_1x}{\Gamma_2(\Gamma_1+\Gamma_2)},\quad N =\frac{x}{\Gamma_1
1570: +\Gamma_2}\,, \\ [10pt]
1571: M_1 =\frac12 (M+N), \quad M_3 =\frac12 (M-N)\,.
1572: \end{array}
1573: $$
1574: The equations of motion for~$ m = (\Gamma_1 +\Gamma_2) M$ and $x$ have a form
1575: \begin{equation}
1576: \label{zerc10}
1577: \begin{aligned}
1578: \frac{dx}{d\tau}&=-2x^2\,,\\
1579: \frac{dm}{d\tau}&=\frac12\left(\frac{\Gamma_1}{\Gamma_2}+
1580: \frac{\Gamma_2}{\Gamma_1}\right)m^2-2xm-\frac12\left(\frac{\Gamma_1}
1581: {\Gamma_2}+\frac{\Gamma_2}{\Gamma_1}\right)x^2\,.
1582: \end{aligned}
1583: \end{equation}
1584: The trajectory, defined by the system~(\ref{zerc10}), is obtained from equation
1585: \begin{equation}
1586: \label{zerc11}
1587: m=x\frac{Cx ^{4/a} -1}{Cx ^{4/a} -1}\,, \quad a =\frac12\left (\frac{\Gamma_1}
1588: {\Gamma_2}+\frac{\Gamma_2}{\Gamma_1}\right)\,,
1589: \end{equation}
1590: where~$ C $ is a constant of an integration. The equation, defining the form of
1591: area~$ (\Delta_i > 0)$ on a plane~$ m, x $ has a form
1592: \begin{equation}
1593: \label{zerc12}
1594: \Gamma_1\Gamma_2\left(x(m-ax)-\frac{(m^2-x^2)x}{8R^2(\Gamma_1+\Gamma_2)^2}
1595: \right) > 0\,.
1596: \end{equation}
1597: Analyzing~(\ref{zerc11}), (\ref{zerc12}) near to the origin of
1598: coordinates~$m=x=0 $ it is possible to conclude that for reflectional
1599: symmetrical solution the simultaneous collapse of four vortices is impossible.
1600:
1601: The discussed above integrable systems of vortex dynamics allow the complete
1602: enough analysis with the help of qualitative research of dynamical systems on a
1603: plane. At the same time application of classical methods of an explicit solution
1604: with the help of the theory of special (Abelian) functions produces very bulky
1605: expressions, and that expressions are not permitting to make any of
1606: representation about the real motion~\cite{Goryachev1}.
1607:
1608: \section { Algebraization and reduction of the three-body problem }
1609:
1610: The reduction in the three-body problem of celestial mechanics
1611: is for the first time systematically considered in the lectures of Jacobi
1612: ~\cite {Jacobi}. He dwelt upon barycentric coordinates, which permit
1613: to ignore uniform motion of the centre of masses, and also to the
1614: procedure of elimination of momentum (elimination of a knot).
1615: Constructively this process reducing in a plane (spatial) case in a system with
1616: three (four) degrees of freedom was executed by Radout, Brunce, Charlier,
1617: Lie, Levi-Chevita and Whittaker, the review of their researches contains in the~\cite
1618: {Whitt, Sharle}. The most natural and complete reduction of order
1619: in the three-body problem belongs to van Kampen (E.\,P.\,van Kampen) and
1620: Wintner (A.\,Wintner)~\cite {Camp}.
1621: This problem is also considered in the work ~\cite {Wint}. All these classical
1622: results are based on the theory of canonical transformations and connected with bulky calculations.
1623: Here we give more geometrical method of the reduction of order in a flat three-body
1624: problem, which could be easily generalized to a spatial~$n$-body problem. It is connected with a preliminary
1625: algebraization of the reduced system and subsequent introduction of canonical
1626: coordinates on a symplectic orbit. In contrast with the classical approach this
1627: procedure uses (algebraic) Poisson structure of the reduced system only and has no connection with Hamiltonian
1628: (and therefore it can be derived for other
1629: potentials, which depend on mutual distance between bodies). The algorithm
1630: of reduction under our consideration is universal enough and uses only algebraic methods.
1631: Undoubdedly, the study of the algebraic form of~$n$-body problem
1632: (not mentioning more natural symmetrization) has large prospects for celestial mechanics.
1633:
1634: \subsection {An algebraization of a system}
1635: The Hamiltonian of a plane three-body problem can be written in the form
1636: \begin {equation}
1637: \label {New1/2}
1638: H = \frac {1} {2} \sum _ {i=1} ^k \frac {{\bf p} _i^2} {m_i} + \frac {1} {2} \sum
1639: _ {i\ne j} ^3 U (| {\bf r} _i- {\bf r} _j |) \,
1640: \end {equation}
1641: where $ {\bf r} _i (x_i, y_i), \, {\bf p} _i (p _ {ix}, p _ {iy}) $ are
1642: two-dimensional vectors of position and momentums of particles, for which
1643: components ($p _ {ix}, \, x_i, \, p _ {iy}, \, y_i $) of Poisson brackets are
1644: canonical.
1645:
1646: %\fig {bormam}
1647: Let us choose new variables, which characerize relative dynamics of particles
1648: as the quadratic forms by canonical variables:
1649: \begin {enumerate}
1650: \item Quadrates of mutual distances
1651: \begin {equation}
1652: \label {New1}
1653: \begin {aligned}
1654: M_1&=|{\bf r} _3- {\bf r} _2 | ^ 2, \\
1655: M_2&=|{\bf r} _1- {\bf r} _3 | ^ 2, \\
1656: M_3&=|{\bf r} _2- {\bf r} _1 | ^ 2.
1657: \end {aligned}
1658: \end {equation}
1659: \item Scalar products of momentums and neighbouring sides of a triangle
1660: formed by points (see Fig.~17)
1661: \begin {equation}
1662: \label {New2}
1663: \begin {aligned}
1664: X_2&=({\bf r} _1- {\bf r} _3, {\bf p}_1)\mbox{, }&
1665: X_3&=({\bf r} _2- {\bf r} _1, {\bf p} _1)\mbox{, }\\
1666: Y_1&=({\bf r} _3- {\bf r} _2, {\bf p}_2)\mbox{, }&
1667: Y_3&=({\bf r} _2- {\bf r} _1, {\bf p} _2)\mbox{, }\\
1668: Z_1&=({\bf r} _3- {\bf r} _2, {\bf p}_3)\mbox{, }&
1669: Z_2&=({\bf r} _1- {\bf r} _3, {\bf p} _3)\,. \\
1670: \end {aligned}
1671: \end {equation}
1672:
1673: \end {enumerate}
1674:
1675:
1676: Since the Hamiltonian (\ref {New1/2}) is dependent only on absolute
1677: quantity of momentums and mutual distances, it can be expressed through
1678: relative variables (\ref {New1}--\ref {New2}). Moreover these variables
1679: form a Lie-Poisson bracket
1680: \begin {equation}
1681: \label {New3}
1682: \begin {aligned}
1683: \{X_2,X_3\}&=X_2+X_3\mbox{, }& \{X_2,Y_1\}&=0\mbox{, }& \{X_2,Y_3\}&=-Y_1 - Y_3\,, \\
1684: \{X_2,Z_1\}&=X_2+X_3\mbox{, }& \{X_2,Z_2\}&=-X_2-Z_2\mbox{, }& \{X_3,Y_1\}&=-X_2-X_3\,,\\
1685: \{X_3,Y_3\}&=X_3+Y_3\mbox{, }& \{X_3,Z_1\}& = 0\mbox{, }& \{X_3,Z_2\}&=Z_1+Z_2\,, \\
1686: \{Y_1,Y_3\}&=-Y_1-Y_3\mbox{, }& \{Y_1,Z_1\}&=Y_1+Z_1\mbox{, }& \{Y_1,Z_2\}&=-Y_1-Y_3\,,\\
1687: \{Y_3,Z_1\}&=-Z_1-Z_2\mbox{, }& \{Y_3,Z_2\}&= 0\mbox{, }& \{Z_1,Z_2\}&=Z_1+Z_2\,,\\
1688: \{X_2,M_1\}&= 0\mbox{, }& \{X_2,M_2\}&=-2M_2\mbox{, }& \{X_2,M_3\}&=M_1-M_2-M_3\,, \\
1689: \{X_3,M_1\}&=0\mbox{, }& \{X_3,M_2\}&=-M_1+M_2+M_3\mbox{, }& \{X_3,M_3\}&=2M_3\,,\\
1690: \{Y_1,M_1\}&=2M_1\mbox{, }& \{Y_1,M_2\}&=0\mbox{, } &\{Y_1,M_3\}&=-M_2+M_1+M_3\,,\\
1691: \{Y_3,M_1\}&=M_2-M_1-M_3\mbox{, } &\{Y_3,M_2\}&=0\mbox{, }& \{Y_3,M_3\}&=-2M_3\,,\\
1692: \{Z_1,M_1\}&=-2M_1\mbox{, }& \{Z_1,M_2\}&=M_3-M_1-M_2\mbox{, }& \{Z_1,M_3\}&=0\,, \\
1693: \{Z_2,M_1\}&=-M_3+M_1+M_2\mbox{, }& \{Z_2,M_2\}&=2M_2\mbox{, }& \{Z_2,M_3\}& = 0\,, \\
1694: \{M_1,M_2\}&=0\mbox{, }& \{M_3,M_2\}&= 0\mbox{, }&\{M_1,M_3\}& = 0\,.
1695: \end {aligned}
1696: \end {equation}
1697:
1698: It can be shown, that variables $M, X, Y, Z $ commute with an integral of the total angular momentum of a system with respect to an arbitrary point and quadrate of total momentum. Hence, map\-ping~{(\ref {New1}--\ref {New2})} corresponds to the
1699: reduction with these integrals, while the rank of an initial canonical bracket
1700: (equal 12) reduces four units down.
1701: Thus rank of a bracket (\ref {New3}) is equal to eight, and the Casimir function
1702: is the quadrate of total momentum (which is expressed through relative variables (\ref {New1}--\ref {New2}) in contrast to the angular momentum).
1703:
1704: {\bf Remark 10.} {\it The method of the algebraization of a dynamical
1705: system which was used here and in the vortex dynamics is a special case of
1706: a general method. It is based on the fact that the quadratic functions of
1707: variables $p, x $ form the algebra~ $ sp (n) $ with respect to canonical
1708: symplectic structure~ $\omega =\sum _ {i} dp_i\wedge dx_i $ ~ \cite
1709: {ArnoldGivental}}. {\it There exists a certain subalgebra in ~ $ sp (n) $,
1710: and its generators commute with integrals of motion of a system to each
1711: particular problem (Hamiltonian).}
1712:
1713: \subsection {Barycentric coordinates and Poisson
1714: submanifolds}
1715: The above mentioned reduction concerns an arbitrary inertial system,
1716: for which two projections of total momentum and total angular
1717: momentum form in general case a noncommutative set of integrals.
1718: In the system of the centre of inertia (barycentric frame) the total momentum is equal to zero, hence set of integrals is commutative and the reduction for one more degree of freedom is possible. To reduce the order in the algebraic form it is necessary to restrict a system on a submanifold of zero total
1719: momentum in the algebra (\ref{New3}).
1720:
1721: We choose new generators of the algebra~(\ref {New3}).
1722: They correspond to eigenvectors of Killing form~\cite {Barut}
1723: \begin {equation}
1724: \label {New4}
1725: \begin {gathered}
1726: \begin {aligned}
1727: S_1 &= \frac{2}{\sqrt{3}}(X_2+X_3+Y_1+Y_3+Z_1+Z_2)\,,\qquad &
1728: S_2 &= \frac {1} {4} (Y_3+Z_2-X_2-Y_1) \,, \\
1729: S_3 &= \frac {1} {4\sqrt {3}} (2X_3-Y_3+2Z_1-Z_2+X_2-Y_1) \,, \qquad &
1730: S_4 &= \frac {1} {6} (X_2-X_3-Y_1+Y_3+Z_1-Z_2) \,, \\
1731: S_5 &= X_2 - Y_1 - Y_3 + Z_2 \,,\qquad &
1732: S_6 &= - X_2 - X_3 + Y_1 + Z_1 \,, \\
1733: \end {aligned} \\
1734: N_1 = -\frac {1} {\sqrt {6}} (M_1+M_2+M_3) \,, \qquad
1735: N_2 = \frac {1} {\sqrt {2}} (M_1 - M_3) \,, \qquad
1736: N_3 = -\frac {1} {\sqrt {6}} (M_1 -2M_2 + M_3) \,.
1737: \end {gathered}
1738: \end {equation}
1739:
1740: The variables $S_5, S_6 $ are proportional to projections of total momentum
1741: on two sides of a triangle (see Fig.~17). A linear span~$ S_5, S_6 $
1742: forms an ideal of algebra (\ref {New3}), hence submanifold of zero
1743: momentum
1744: \begin {equation}
1745: \label {New4.5}
1746: S_5=0 \,,\qquad S_6=0
1747: \end {equation}
1748: is Poisson submanifold. Restricting a system on~(\ref {New4.5}) and ordering
1749: remaining variables as follows~$ {\bf x} = (S_1, \, S_2, \, S_3, \, S_4,\,N_1, \, N_2, \, N_3) $, we
1750: obtain the table of commutators
1751: \begin {equation}
1752: \label {New5}
1753: \| \{x_i, x_j \} \| =
1754: \left (\begin {array} {ccccccc}
1755: 0 & S_3 & -S2 & 0 & 0 & N_3 & -N_2 \\
1756: -S_3 & 0 & -S_1 & 0 & -N_3 & 0 & - N_1 \\
1757: S2 & S_1 & 0 & 0 & N_2 & N_1 & 0 \\
1758: 0 & 0 & 0 & 0 & -N_1 & -N_2 & -N_3 \\
1759: 0 & N_3 & -N_2 & N_1 & 0 & 0 & 0 \\
1760: -N_3 & 0 & -N_1 & N_2 & 0 & 0 & 0 \\
1761: N_2 & N_1 & 0 & N_3 & 0 & 0 & 0
1762: \end {array} \right).
1763: \end {equation}
1764: The Casimir function of structure (\ref {New5}) coincides with quadrate of
1765: the total angular momentum with respect to the centre of masses
1766: \begin {equation}
1767: \label {New6}
1768: M_z^2=4\frac {\left < {\bf S, N} \right > ^2} {\left < {\bf N, N} \right >} \,
1769: \end {equation}
1770: where~$ \left < {\bf a}, {\bf b} \right > =a_1b_1-a_2b_2-a_3b_3 $ is a scalar
1771: product in the Minkowski space.
1772: Numerator and denominator of this fraction are Casimir functions of a
1773: subalgebra~$ so (1,2) \otimes_s\mathbb{R}^3 $ with generators
1774: $ (S_1, S_2, S_3, N_1, N_2, N_3) $.
1775: Moreover~$ \left < {\bf N}, {\bf N} \right > =2\Delta^2 $, where $ \Delta $ is
1776: the square of a triangle formed by particles.
1777:
1778: The algebra ~ $ l_7 $, corresponding to a bracket (\ref {New5}) is the
1779: semidirect sum of subalgebra formed by elements $ {S_i, \, i=1,
1780: \ldots, 4} $ and three-dimensional commutative ideal: $l_7 = (so (1,2) \oplus
1781: R) \oplus_s R^3 $.
1782:
1783: The quadrates of momentum and mutual distances can be written as
1784: \begin{equation}
1785: \label {New7}
1786: \begin {gathered}
1787: p_k^2 = \frac {1} {3} \frac {\left < {\bf e} _k, {\bf N} \right > S_4^2
1788: - 2\left < {\bf e} _k, {\bf S}, {\bf N} \right > S_4 +
1789: 2\left < {\bf e} _k, {\bf S} \right > \left < {\bf S}, {\bf N} \right > \left <
1790: {\bf e} _k, {\bf N} \right > \left < {\bf
1791: S}, {\bf S} \right >} {\left < {\bf N}, {\bf N} \right >} \, \\
1792: M_k =\langle {\bf e} _k, {\bf N} \rangle \mbox{, } k=1, \, 2, \, 3,
1793: \end {gathered}
1794: \end {equation}
1795: where
1796: $ {\bf S} = (S_1, S_2, S_3), \, {\bf N} = (N_1, N_2, N_3) $ and vectors $ {\bf e}
1797: _k $ have the form
1798: $$
1799: {\bf e}_1=\frac{1}{\sqrt{6}}(-2,-\sqrt{3},1)\mbox{, } {\bf e} _2 =\frac
1800: {1} {\sqrt {6}} (-2,0, -2)\mbox{, } {\bf
1801: e} _3 =\frac {1} {\sqrt {6}} (-2, -\sqrt {3}, 1)
1802: $$
1803: and satisfy the relation
1804: $ \langle {\bf e} _i, {\bf e} _j\rangle=1-\delta _ {ij} $. Here~$ \langle {\bf a},
1805: {\bf b}, {\bf c} \rangle $ is a determinant matrix which is formed by components of
1806: vectors
1807: $ \bf a, b, c $.
1808:
1809: The Hamiltonian has the form
1810: \begin {equation}
1811: \label {New9}
1812: H = \frac {p_1^2} {m_1} + \frac {p_2^2} {m_2} + \frac {p_3^2} {m_3} + \frac
1813: {m_1m_2} {\sqrt {M_3}} + \frac {m_2m_3} {\sqrt {M_1}} + \frac {m_1m_3} {\sqrt{M_2}}\,.
1814: \end {equation}
1815:
1816: \subsection {Orbits and symplectic coordinates} In algebra (\ref {New5}) the
1817: orbits appropriate to physical motions of a system are divided in
1818: two types.
1819: \begin {enumerate}
1820:
1821: \item Regular six-dimentional orbits are surfaces of a level of Casimir
1822: function~(\ref {New6}). The quadrate of area of a triangle is nonnegative,
1823: therefore for physical symplectic sheets~$ \left < {\bf N, N} \right > \ge 0 $.
1824: Topologi\-cal\-ly such orbit is diffeomorphic~$TL^2\times \mathbb{R} ^
1825: {+}\times \mathbb{R} $. Here $TL^2 $ ~ is a tangent bundle of a
1826: pseudosphere $ \left < {\bf N, N} \right > =c_1 $, its ``radius" has
1827: non-negative values $c_1\in \mathbb{R} ^ {+} = [0, \infty) $, and the last factor
1828: corresponds to a linear span $S_4 $.
1829:
1830: \item It is possible to show (analyzing a rank of a matrix (\ref {New5})),
1831: that four-dimentional orbits diffeomor\-phic to tangent bundle to
1832: a cone $ \left < {\bf N}, {\bf N} \right > =0 $ pass through points satisfying
1833: the equations ${ \left < {\bf N}, {\bf N}
1834: \right > =0}, \left < {\bf S}, {\bf N} \right > =0 $. These orbits correspond to
1835: the motion of particles along a straight line (the area~$\Delta$ is equal to zero).
1836: \end {enumerate}
1837:
1838: {\bf Remark 11.} {\it
1839: The orbits passing through points with negative value of
1840: quadrate of area~$ \left < {\bf N}, {\bf N} \right > < 0 $ or with zero
1841: quadrates of distances have no physical sense.}
1842:
1843: We construct symplectic coordinates on a regular orbit similar to
1844: Anduier-Depri coordinates in the rigid body dynamics~\cite {BE}.
1845: For this purpose we will apply the following algorithm.
1846:
1847: For algebra (\ref {New5}) we shall consider the following consequence of the
1848: inserted subalgebras $so (1,2) \subset so (1,2) \oplus_s R^3\subset l_7 $.
1849:
1850: On a subalgebra $so (1,2) $ we shall choose $L=S_1 $ as an action.
1851: For a Hamiltonian field of vectors with the Hamilton function~$ \mathcal H=L $
1852: $$
1853: \frac {dS_1} {dl} =0 \,,\qquad \frac {dS_2} {dl} =S_3 \,,\qquad \frac {dS_3} {dl}
1854: = -S_2\,
1855: $$
1856: time parameter along an integral curve $l $ is a desired angular variable
1857: canonically conjugate $L $: ${\{l, L \}=1}$.
1858: Choosing constants of an integration so that the relation
1859: $ \{S_2, S_3 \} =-S_1 $ can be fulfilled, we obtain
1860: \begin {equation}
1861: \label {New10}
1862: S_1=L \,,\qquad S_2 =\sqrt {L^2-G^2} \cos l \,, \qquad S_3 =\sqrt {L^2-G^2}
1863: \sin l \,.
1864: \end {equation}
1865:
1866: We shall choose the Casimir function $G =\sqrt {S_1^2-S_2^2-S_3^2} $ of subalgebra
1867: $so (1,2) $ as a Hamiltonian $ \mathcal H=G $ on a subalgebra
1868: $so (1,2) \oplus_s R^3 $. Then we integrate the equation that is linear
1869: with respect to~$N_i $
1870: $$
1871: \frac {dN_1} {dg} = \frac {S_3 N_2 - S_2 N_3} {G}\,, \quad
1872: \frac {dN_2} {dg} = \frac {S_3 N_1 - S_1 N_3} {G}\,, \quad
1873: \frac {dN_3} {dg} = \frac {-S_2 N_1 + S_1 N_2} {G}\,,
1874: $$
1875: {\sloppy We find dependence of constants of integration on $L, l, G $ from
1876: commutators of the subalgebra ${so (1,2) \oplus_s R^3}$.
1877: \begin{equation}
1878: \label {New11}
1879: \begin {array} {l}
1880: N_1 =\frac {HL} {G^2}
1881: +\frac{\sqrt{L^2-G^2}}{G}\sqrt{\frac{H^2}{G^2}- \Delta^2} \cos g\,, \\
1882: N_2 =\frac {H} {G^2} \sqrt {L^2 - G^2} \cos l
1883: +\frac{L}{G}\sqrt{\frac{H^2}{G^2}-\Delta^2}\cos g\cos l - \sqrt {\frac {H^2}
1884: {G^2} - \Delta^2} \sin g\sin l\,, \\
1885: N_3 =\frac {H} {G^2} \sqrt {L^2 - G^2}
1886: \sin l +\frac{L}{G}\sqrt{\frac{H^2}{G^2}-\Delta^2}\cos g\sin l + \sqrt
1887: {\frac {H^2} {G^2} - \Delta^2} \sin g\cos l\,,
1888: \end {array}
1889: \end {equation}
1890: where $H=\left<{\bf S},{\bf N}\right>,\;\Delta=\left<{\bf N},{\bf N}\right>$.
1891:
1892: }
1893:
1894: We shall choose as a last action variable~$S=S_4 $.
1895: As a result we obtain the following expressions
1896: \begin {equation}
1897: \label {New12}
1898: \begin {aligned}
1899: N_1&=e^s\Bigl(\frac{M_z L} {G^2}
1900: +\frac{\sqrt{L^2\,{-}\,G^2}}{G}\sqrt{\frac{M_z^2}{G^2}-1}
1901: \cos g\Bigr), \\
1902: N_2&=e^s\Bigl(\frac{M_z}{G^2}\sqrt{L^2{-}G^2}\cos l
1903: +\frac{L}{G}\sqrt{\frac{M_z^2}{G^2}-1}\cos g\cos l - \sqrt {\frac {M_z^2}
1904: {G^2} - 1} \sin g\sin l \Bigr), \\
1905: N_3&=e^s\Bigl(\frac{M_z}{G^2}\sqrt{L^2-G^2}\sin l
1906: +\frac{L}{G}\sqrt{\frac{M_z^2}{G^2}-1}\cos g\sin l + \sqrt {\frac {M_z^2}
1907: {G^2} - 1} \sin g\cos l \Bigr), \\
1908: S_4&=S\,,
1909: \end {aligned}
1910: \end {equation}
1911: here $M_z $ is a constant of total momentum with respect to a system of the
1912: centre of masses. The variables $S_1, S_2, S_3 $ express by the formulas
1913: (\ref {New10}).
1914:
1915: It is possible to introduce the other pair of canonical variables~$x,\,y $,
1916: instead of~$s,\,S $, by the formulas~$ e^s=x, \, S=xy $.
1917:
1918: {\bf Remark 12.} {\it Using the above mentioned algebraic structure of
1919: three-body problem, it is also easy to reconstruct the results of classics
1920: in a problem of reduction of order. However, obtained expressions will
1921: remain bulky enough. Remind that in researches of Brunce, Whittaker and
1922: van Kampen as positional variables the mutual distances are chosen.
1923: Reduction of the order which has been carried out by Charlier, uses
1924: distances of three bodies from the general centre of inertia.}
1925:
1926: Using the algebraic form of the equations of the three-body problem
1927: it is easy to investigate its particular solutions. One of such
1928: solutions was founded by Lagrange. In this case all of three bodies are
1929: situated in tops of an equilateral triangle (which in an even more special
1930: case does not change sizes). The second solution belongs to Euler and
1931: defines collinear stationary configurations. All these solutions can be
1932: defined by a method, explained in \cite {Wint, Smeil}, and also one can
1933: obtain a system of appropriate invariant relations which are not determined,
1934: Poisson manifolds.
1935:
1936: It is easy to show (see ~\cite {Wint}), that there are only two ``rigid-body" stationary configurations in the three-body problem --- Euler and
1937: Lagrangian. Collinear configurations are investigated in detail in the~ $ n $-body problem. By the Multon theorem their number is equal
1938: to~$\frac {n!} {2} $~\cite {Smeil}.
1939:
1940: The topological proof of the Multon theorem is given with the use of the Morse theory.
1941: In the paper~\cite {Smeil} series of hypotheses about the amount of noncollinear configurations in a~$ n $-body
1942: problem are also expressed. As far as we know the majority of them are not proved until now. We hope that the algebraic form of the reduced system offered in this application, will allow to achieve further progress in this problem.
1943:
1944: The autors are grateful to N.\,N.\,Simakov for the help in numerical computations. In Russian outcomes
1945: of this paper are introduced in the book A.\,V.\,Borisov,
1946: I.\,S.\,Mamaev ``Poisson structures and Lie algebras in Hamiltonian
1947: mechanics".
1948:
1949:
1950:
1951:
1952:
1953:
1954: \begin{thebibliography}{99}
1955:
1956: \bibitem{ArefH}
1957: H.\,Aref.
1958: {\em On the equilibrium and stability of a row of point vortices}. J. Fluid
1959: Mech., v.~290, 1995, p.~167--191.
1960:
1961: \bibitem{Aref2}
1962: H.\,Aref.
1963: {\em Integrable, chaotic and turbulent vortex motion in
1964: two-dimensional flows}.
1965: Ann. Rev. Fluid Mech., v.~15, 1983, p.~345--389.
1966:
1967: \bibitem{ArefPoint}
1968: H.\,Aref.
1969: {\em Point vortex motions with a center of symmetry}. Phys. Fluids, 1982,
1970: v.~25 (12), p.~2183--2187.
1971:
1972: \bibitem{ArefV}
1973: H.\,Aref, D.\,L.\,Vainstein.
1974: {\em Point vertices exhibit asymmetric equilibria}. Nature, v.~392, 1998,
1975: 23~April, p.~769--770.
1976:
1977: \bibitem{ArnoldKozlovNei}
1978: V.\,I.\,Arnold, V.\,V.\,Kozlov, A.\,I.\,Neishtadt.
1979: {\em Mathematical aspects of classical and celestial mechanics}.
1980: 2nd ed. Encyclopedia of Mathematical Sciences,
1981: v.~3 (Dynamical systems III), New York, Springer, 1996.
1982:
1983: \bibitem{Arnold}
1984: V.\,I.\,Arnold.
1985: {\em Mathematical methods of classical mechanics.}
1986: New York, Springer, v.~60., 1984.
1987:
1988: \bibitem{ArnoldGivental}
1989: V.\,I.\,Arnold, A.\,B.\,Givental.
1990: {\em Symplectic geometry}.
1991: M., VINITI, v.~4, 1985, p.~5--140.
1992:
1993: \bibitem{Barut}
1994: A.\,Barut, R.\,Raczka. {\em Theory of {G}roup {R}epresentations
1995: and {A}pplications}. {P}{W}{N}. {P}olish {S}cientific {P}ublishers, 1977.
1996:
1997: \bibitem{Bog2}
1998: V.\,A.\,Bogomolov.
1999: {\em Dynamics of the vortisity on a sphere}.
2000: Izv. AN USSR, Mekh. zhid. gaza, No. 6, 1977, p.~57--65.
2001:
2002: \bibitem{Bolsinov}
2003: A.\,V.\,Bolsinov.
2004: {\em Compatible Poisson brackets on Lie algebras and the completeness
2005: of families of involutive functions}
2006: Izv. AN USSR, ser. math., v.~55, 1991, No. 1, p.~68--92.
2007:
2008: \bibitem{BolsBor}
2009: A.\,V.\,Bolsinov, A.\,V.\,Borisov.
2010: {\em Lax-representation and compatible Poisson brackets on Lie algebras}.
2011: (to be published, 1999).
2012:
2013: \bibitem{BE}
2014: A.\,V.\,Borisov, K.\,V.\,Emelyanov.
2015: {\em Non-integrability and stochastisity in rigid-body dynamics}.
2016: Izhevsk, Izd. Udm. Univ., 1995.
2017:
2018: \bibitem{BPavlov}
2019: A.\,V.\,Borisov, A.\,E.\,Pavlov.
2020: {\em Dynamics and {S}tatics of {V}ortices on a {P}lane and a {S}phere
2021: --- I}. Reg. \& Ch. Dynamics, v.~3, 1998, No. 1, p.~28--39.
2022:
2023: \bibitem{BorisovLebedev}
2024: A.\,V.\,Borisov, V.\,G.\,Lebedev.
2025: {\em Dynamics of {T}hree {V}ortices on a {P}lane and a {S}phere ---
2026: II. General compact case}.
2027: Reg. \& Ch. Dynamics, v.~3, 1998, No. 2, p.~99--114.
2028:
2029: \bibitem{BorisovLeb}
2030: A.\,V.\,Borisov, V.\,G.\,Lebedev.
2031: {\em Dynamics of three vortices on a plane and a sphere ---
2032: III. Noncompact case. Problem of collapse and scattering}.
2033: Reg. \& Ch. Dynamics, v.~3, 1998, No. 4, p.~76--90.
2034:
2035: \bibitem{Campbell}
2036: L.\,J.\,Campbell, R.\,M.\,Ziff. {\em Vortex patterns and energies in a rotation
2037: superfluid}. Phys. Rev. B., v.~20, 1979, No. 5, p.~1886--1901.
2038:
2039: \bibitem{Goryachev1}
2040: D.\,N.\,Goryachev.
2041: {\em On some cases of the motion of straightline vortices}.
2042: Moskwa, 1898.
2043:
2044:
2045: \bibitem{Fomenko2}
2046: A.\,T.\,Fomenko.
2047: {\em Symplectic geometry.}
2048: М.: МGU, 1988.
2049:
2050: \bibitem{Jacobi}
2051: C.\,G.\,J.\,Jacobi.
2052: {\em Vorlesungen \"uber Dynamik}.
2053: Aufl. 2. Berlin: G. Reimer, 1884, 300~S.
2054:
2055: \bibitem{Camp}
2056: E.\,R.\,van\,Kampen, A.\,Wintner. {\em On a symmetrical canonical reduction of
2057: the problem of three bodies}. Amer. J. Math., v.~59, 1937, No. 1, p.~153--166.
2058:
2059:
2060: \bibitem{Kirh}
2061: G.\,Kirchhoff.
2062: {\em Vorlesungen \"uber mathematische Physik}.
2063: Leipzig, 1891, 272~S.
2064:
2065:
2066: \bibitem{MarsdenW}
2067: J.\,Marsden, A.\,Weinstein. {\em Reduction of Symplectic manifolds with
2068: symmetry}. Rep. on Math. Phys., v.~5, 1974, No. 5, p.~121--130.
2069:
2070:
2071: \bibitem{Melesh}
2072: V.\,V.\,Meleshko, M.\,Yu.\,Konstantinov.
2073: {\em Dynamics of vortex structures}.
2074: Kiev, Naukova Dumka, 1993.
2075:
2076: \bibitem{NovikovSed}
2077: E.\,A.\,Novikov, Yu.\,B.\,Sedov. {\em Collapse of vortices}.
2078: ZETF, v.~77, No. 2(8), 1979, p.~588--597.
2079:
2080: \bibitem{Sharle}
2081: C.\,L.\,Charlier. {\em Die Mechanik des Himmels}.
2082: Walter de Gruyter \& Co. 1927.
2083:
2084: \bibitem{Smeil}
2085: S.\,Smeil.
2086: {\em Topology and mechanics}.
2087: Uspekhi mat. nauk, v.~27, No. 2, 1972,
2088: p.~77--133.
2089:
2090: \bibitem{Whitt}
2091: E.\,T.\,Whittaker. {\em A treatise on the analytical dynamics}.
2092: Ed. 3-d. Cambridge Univ. Press., 1927.
2093:
2094: \bibitem{Wint}
2095: A.\,Wintner. {\em The analytical foundation of celestial mechanics}.
2096: Princeton Univ. Press, 1941.
2097: \end{thebibliography}
2098:
2099: \begin{figure}[ht!]
2100: $$
2101: \includegraphics{brq.eps}
2102: $$
2103: \caption{}
2104: \end{figure}
2105:
2106: \begin{figure}[ht!]
2107: $$
2108: \includegraphics{energyp.eps}
2109: $$
2110: \caption{$G=0.3,\;g=\pi/2$}
2111: \end{figure}
2112:
2113: \begin{figure}[ht!]
2114: $$
2115: \includegraphics{energyt.eps}
2116: $$
2117: \caption{$G=0.3,\;g=\pi/4$}
2118: \end{figure}
2119:
2120: \begin{figure}[ht!]
2121: $$
2122: \includegraphics{energyq.eps}
2123: $$
2124: \caption{$G=0.3,\;g=0$}
2125: \end{figure}
2126:
2127: \begin{figure}[ht!]
2128: $$
2129: \includegraphics{simpleq.eps}
2130: $$
2131: \caption{$H=0.5,\;E=-3.2$}
2132: \end{figure}
2133:
2134: \begin{figure}[ht!]
2135: $$
2136: \includegraphics{simplee.eps}
2137: $$
2138: \caption{$H=0.5,\;E=-3.3$}
2139: \end{figure}
2140:
2141: \begin{figure}[ht!]
2142: $$
2143: \includegraphics{simpler.eps}
2144: $$
2145: \caption{$H=0.5,\;E=-3.4$}
2146: \end{figure}
2147:
2148: \begin{figure}[ht!]
2149: $$
2150: \includegraphics{simplew.eps}
2151: $$
2152: \caption{$H=0.5,\;E=-3.6$}
2153: \end{figure}
2154:
2155: \begin{figure}[ht!]
2156: $$
2157: \includegraphics{simplet.eps}
2158: $$
2159: \caption{$H=0.5,\;E=-6.0$}
2160: \end{figure}
2161:
2162: \begin{figure}[ht!]
2163: $$
2164: \includegraphics{simpley.eps}
2165: $$
2166: \caption{$H=0.5,\;E=-7.5$}
2167: \end{figure}
2168:
2169: \begin{figure}[ht!]
2170: $$
2171: \includegraphics{vihri.eps}
2172: $$
2173: \caption{}
2174: \end{figure}
2175:
2176: \begin{figure}[ht!]
2177: $$
2178: \includegraphics{ivanw.eps}
2179: $$
2180: \caption{}
2181: \end{figure}
2182:
2183: \begin{figure}[ht!]
2184: \begin{center}
2185: { \small
2186: \begin{tabular}{ccc}
2187: \includegraphics{cent.eps} & \includegraphics{ceny.eps} &
2188: \includegraphics{cenu.eps}\\
2189: a) $\begin{aligned}\Gamma_1&=-1.5,\\
2190: \Gamma_2&=2.0.
2191: \end{aligned}$&
2192: b) $\begin{aligned}\Gamma_1&=-0.183,\\
2193: \Gamma_2&=2.0. \end{aligned}$&
2194: c) $\begin{aligned}\Gamma_1&=-0.183,\\
2195: \Gamma_2&=0.683.\end{aligned}$
2196: \end{tabular}}
2197: \end{center}
2198: \caption{}
2199: \end{figure}
2200:
2201: \begin{figure}[ht!]
2202: \begin{center}
2203: { \small
2204: \begin{tabular}{ccc}
2205: \includegraphics{cenq.eps} & \includegraphics{cenw.eps} &
2206: \includegraphics{cene.eps}\\
2207: a) $\begin{aligned}\Gamma_1&=-0.183,\\
2208: \Gamma_2&=2.0,\\
2209: k&=0.6.\end{aligned}$&
2210: b) $\begin{aligned}\Gamma_1&=-1.5,\\
2211: \Gamma_2&=2.0,\\ k&=0.6.\end{aligned}$&
2212: c) $\begin{aligned}\Gamma_1&=-0.183,\\
2213: \Gamma_2&=0.683,\\
2214: k&=0.6.\end{aligned}$
2215: \end{tabular}} \vspace{-5mm}
2216: \end{center}
2217: \caption{}
2218: \end{figure}
2219:
2220: \begin{figure}[ht!]
2221: $$
2222: \includegraphics{ivane.eps}
2223: $$
2224: \caption{}
2225: \end{figure}
2226:
2227: \begin{figure}[ht!]
2228: \begin{center}
2229: { \small
2230: \begin{tabular}{cc}
2231: \includegraphics {zerce.eps} & \includegraphics{zercy.eps}\\
2232: $\begin{aligned}\Gamma_1&=2.0,\;&\Gamma_2&=2.0\\
2233: k&=0.0,\;&D&=-2.0.\end{aligned}$&
2234: $\begin{aligned}\Gamma_1&=2.0,\;&\Gamma_2&=2.0\\
2235: k&=0.9,\;&D&=-2.0.\end{aligned}$\\[15pt]
2236: \includegraphics {zercq.eps} & \includegraphics {zercr.eps}\\
2237: $\begin{aligned}\Gamma_1&=-1.0,\;&\Gamma_2&=2.0\\
2238: k&=0.0,\;&D&=-2.0.\end{aligned}$&
2239: $\begin{aligned}\Gamma_1&=-1.0,\;&\Gamma_2&=2.0\\
2240: k&=0.9,\;&D&=-2.0.\end{aligned}$\\[15pt]
2241: \includegraphics{zercw.eps} & \includegraphics {zerct.eps}\\
2242: $\begin{aligned}\Gamma_1&=-1.8,\;&\Gamma_2&=2.0\\
2243: k&=0.0,\;&D&=-2.0.\end{aligned}$&
2244: $\begin{aligned}\Gamma_1&=-1.8,\;&\Gamma_2&=2.0\\
2245: k&=0.9,\;&D&=-2.0.\end{aligned}$\\[15pt]
2246: a)&b)
2247: \end{tabular}} \vspace{-5mm}
2248: \end{center}
2249: \caption{}
2250: \end{figure}
2251:
2252: \begin{figure}[ht!]
2253: $$
2254: \includegraphics{bormam.eps}
2255: $$
2256: \caption{}
2257: \end{figure}
2258:
2259: \end{document}
2260: