math0405014/draft5
1: \magnification=\magstep1 \hoffset .65pt
2: \def\R{I\!\! R}
3: \def\Z{\it Z \!\!\! Z}
4: \def\H {I \!\! H}
5: \def\C{C \!\! \! \!  I\, }
6: \def\I{I\!\!I}
7: \def\J{J\!\!\!J}
8: \def\RP{I\!\! R I\!\! P}
9: \def\Ri{Riemannian }
10: 
11: \input epsf.tex
12: 
13: \font\eightrm=cmr8
14: \parindent=8pt
15: \font\eightrm=cmr8
16: \parindent=8pt
17: \hskip11cm {\bf DRAFT. April 24, 2004}
18: \bigskip
19: \centerline{\bf  Fitting Hyperbolic pants
20: to a three-body problem. }
21: \bigskip
22: \bigskip
23: 
24: {\bf  Abstract.}  Consider the three-body problem
25:  with an  attractive $1/r^2$
26: potential.  Modulo symmetries,
27: the dynamics of the    bounded zero-angular momentum solutions  
28: is equivalent to a  geodesic flow on the  thrice-punctured sphere,
29: or ``pair of pants''.  The sphere is the shape sphere.
30:   The punctures are the binary collisions.  The metric generating
31: the geodesics is the Jacobi-Maupertuis metric.  The metric is 
32: complete, has infinite area, and its ends, the neighborhoods
33: of the punctures,  are asymptotically cylindrical.
34: Our main result is  that when the three masses are  equal then the
35: metric   has negative curvature everywhere except at two
36: points (the Lagrange points).  A corollary
37: of this negativity is the uniqueness of the  $1/r^2$ figure eight,
38:  a complete symbolic dynamics for encoding   
39: the collision-free solutions, and the fact that collision solutions are
40: dense within the bound solutions. 
41:  
42: \vskip .3cm
43: 
44:  
45: 
46:  
47: {\bf 1. Introduction and Results.}
48: 
49: We   study  the  planar three-body problem
50:  with an  attractive $1/r^2$
51: potential.  According to the Lagrange-Jacobi
52: identity (eq. (3.7) below)
53:  every bounded  solution must have  
54: zero energy  and constant moment of interia $I$ , and conversely,
55: if an initial condition has zero energy and $\dot I (0) = 0$
56: then that solution is bounded.  
57: Setting  the moment of inertia $I$ equal to a constant  defines a three-sphere in
58: configuration space.  
59: Rotations act on this sphere according to the
60: Hopf flow so that the quotient of the three-sphere
61: by rotations is the two-sphere or {\it shape sphere}. See figure 1a. Points of this shape sphere 
62: represent oriented similarity classes of triangles.  Newton's equations,
63: for solutions with   $H = 0, \dot I = 0$, 
64: push down to  the shape sphere to yield a   
65: a family of second-order ODEs 
66: parameterized by the angular momentum.  These ODEs
67: have  singularities at the three points  representing
68: the three types of binary collisions.  
69: Upon deleting the collision points
70: we arrive  at dynamics on   the {\it pair of
71: pants}, -- the two-sphere minus three points.  
72: When the angular momentum is zero the resulting  dynamical 
73: system  is,   after a time
74: reparameterization, the geodesic flow for a certain \Ri metric on the
75: pair of pants.  This metric is the (reduced)  Jacobi-Maupertuis metric for energy
76: $0$.    
77: \proclaim Proposition 1. Endow
78: the pair of pants with the Jacobi-Maupertuis
79: metric (equations (3.9a,b) below).   
80:  Modulo rotations, translation, 
81: and scaling, the set of bounded 
82: zero-angular momentum solutions
83: for the $1/r^2$ potential three-body problem are in bijective correspondence 
84: with   geodesics
85: for this metric.   
86: The  metric is complete 
87: and its ends (the deleted neighborhoods of the three
88: binary  collisions) are asymptotic to Euclidean cylinders of positive radii. 
89: 
90: {\bf Proof:}  Section 3.
91: 
92: \vskip .3cm
93: 
94: See figure 1b for a depiction of the pair of pants.  
95: 
96: \eject
97: 
98: \vskip 0.1in
99: 
100: \epsfxsize=2.50in
101: 
102: \epsfbox{sphere4.eps}
103: \vskip 0.2in
104: 
105: {\bf Figure 1a.} The shape sphere 
106: \vskip .1in
107: \epsfxsize=3.00in
108: 
109: \epsfbox{pants4.eps}
110: 
111: \vskip 0.8in
112: 
113: {\bf Figure 1b.} The pair of pants
114: 
115:  
116: 
117:  The Jacobi-Maupertuis  metric depends parametrically on the   masses of the three bodies by way of
118: the potential (eq. 3.1).  Our main result is:
119: \proclaim Theorem 1.  If all three masses are
120: equal then  the
121:  Gaussian curvature for the Jacobi-Maupertuis metric on the pair of pants  is   negative 
122: everywhere  except at the two Lagrange points,  where it is zero.
123: 
124: {\bf Proof.}  Section 4. 
125: 
126: One might hope that  negativity   of 
127: the curvature persists for unequal masses.  It does not.  See section 7.
128: \vskip .44cm
129: 
130: 
131: {\bf 2. Motivation and  Dynamical Consequences.}  
132: 
133: {\bf 2.1. Periodic Orbits and their symbol sequences.}
134: 
135: This work began as  an attempt to give an analytic proof that  the
136: Newtonian ($1/r$ potential) figure eight solution of Moore-Chenciner-Montgomery
137: ([Moore], [ChMont])
138: is unique.   
139: I began with the   easier
140: case of the $1/r^2$ eight.
141:  The figure eight  
142: is a periodic solution   which realizes a certain free homotopy class 
143: on the pair of pants.  Figure eights  exists for all $1/r^a$ potentials, $a > 0$
144: ([CGMS] , [FerrTerr]).  For $a \ge 2$, 
145:  not only is the free homotopy class of the  eight realized,   but 
146: almost   every free homotopy class is realized by a solution. Combining these   facts
147: suggested the approach of this paper, and Theorem 1.
148: 
149: Our pair of pants metric from theorem 1 is neither compact, nor of negative curvature
150: everywhere.  But on a complete, noncompact surface
151: of negative curvature if a free homotopy class  has a geodesic representative,
152: then that representative is unique.  And  uniqueness
153: continues to hold if the curvature vanishes on a discrete set of points.
154: (This theorem is fairly well-known, and proved in  
155: in a more general context in section 6.4 below, and in particular
156: eq (6.4.3.).) We have proved
157:  
158: \proclaim Corollary.  For  the $1/r^2$ 
159: equal-mass zero-angular-momentum three-body problem,  if
160: a solution realizes a given  
161: free homotopy class on the pair of pants,
162: then that solution is unique   modulo rotation and scaling.
163: In particular the eight is unique modulo these
164: symmetries. 
165: 
166: In stating the corollary we begged the question of which classes 
167: are realized.  Every
168: class is realized  with the exception of  those classes which wind around a single end.
169: (See [MontN]).  
170: Gordon [1970] calls these  `bad' or unrealizable classes  `untied' 
171: while the complementary `good',  or realizable classes he  called `tied', being 
172: that they are 
173: `tied' to the collision singularities.  On the pants,
174: a bad class can be represented by  drawing a small circle,
175: or ``anklet''  around one
176: pants leg, and traversing it some number of times.
177: As this ``anklet''  is pushed   down towards the end of the leg
178: its length decreases.  As a result, any   minimizing sequence 
179: of curves realizing such a class ``falls off'' of the leg. (See theorem 3 below.)
180:       
181: 
182: We   follow [MontN] in using syzygies to describe the tied and untied classes.     
183: A {\it syzygy} 
184: is a collinear configuration of  the three bodies.
185: Syzygies come in three flavors,
186: 1,2, and 3, depending on which mass is  between the other two.
187: (We exclude collisions.) 
188: The collinear configurations form the equator of the shape sphere.
189: (Figure 1a.) The three collisions lie on the equator so
190: that deleting them divides the  equator into three arcs,
191: again labelled 1,2,3  according to the   mass in the middle. 
192:    A  curve  on the 
193: shape sphere has  an 
194:  associated syzygy sequence:     list the  syzygies in order. 
195: (Assume that the   syzygy times  are discrete.)
196: The syzygy sequence of a motion of the three bodies
197: is obtained by projecting the motion onto the
198: shape sphere and writing out the syzygy sequence of 
199: the curve resulting  on the shape sphere.   
200:  
201:  Periodic  curves
202: give rise to periodic sequences.    For example, the class
203: in which  1 and 2 circle about each other for ever  while 3 remains far
204: away  has  syzygy sequence $\ldots 121212 \ldots$. 
205: (This is a ``bad'' class as we see later.)   We   subject  syzygy sequences to   the {\it  no stutterring rule}:
206: if  $ij$ are   consecutive letters
207: of the sequence, then   $i \ne j$.  The
208: reason for imposing this rule is that a stutter   can be homotoped
209: away. See figure 2. 
210: 
211:  A letter $j$ with a plus superscript,
212: as in $j^+$, denotes that syzygy $j$  occurs by crossing from the upper
213: to the lower hemisphere of the sphere.  A $j^-$
214: means that  
215: syzygy $j$  occurs by crossing from the lower
216: to the upper hemisphere of the sphere.   Pluses and minuses must alternate, since 
217: the  path will alternate between hemispheres.   In topological
218: terms an arc segment with two consecutive
219: pluses, such as   $k^-i^+ j^+$ lies entirely in
220: the upper hemisphere and   can be homotoped to $k^- j^+$.
221: In dynamical terms such an arc can never occur since at $i^+$
222: it would have to be tangent to the collinear subspace (the equator), but 
223: if a solution is tangent to the collinear  subspace at a point then
224: it lies completely within the collinear subspace.   
225: 
226: It follows from the above considerations, and the topology of the
227: pair of pants, that there is a one-to-one onto correspondence between  
228:  free homotopy classes and  periodic signed non-stuttering syzygy sequences.
229: From now on we will drop the signing indications --the $+, -$ superscripts -- for simplicity. 
230: (Given an unsigned sequence there are only two ways to decorate it with
231: signs.) 
232: The untied (unrealizable)  classes are precisely  those with syzygy sequence
233:   $...1212...$, $2323...$ or $...3131....$.
234:  They correspond to a curve  winding around a single end.  We prove in theorem 3 
235: that there is no bounded zero angular momentum solution which realizes
236: them.    
237:   Excluding these classes is equivalent to  insisting that
238: all three letters occur in the sequence.   Thus  the corollary asserts that  {\bf
239: every periodic non-stuttering syzygy sequence in which all three letters occur
240: is realized by a unique (up to symmetry) relative periodic solution.} 
241: 
242: 
243: \vskip 0.2in
244: 
245: \epsfxsize=2.00in
246: 
247: \epsfbox{stutter.eps}
248: \vskip 0.4in
249: 
250: {\bf Figure 2.} Homotoping away a stutter.
251: 
252: \vskip .4cm
253: 
254: 
255:   
256: {\bf 2.2. Symbolic Dynamics;  aperiodic syzygy  sequences.}
257:  
258: We  move  on to infinite aperiodic syzygy sequences. 
259:  In the rest of this subsection, `the problem'  means 
260:  the 
261: $1/r^2$ 
262: equal-mass zero-angular-momentum three-body problem, and
263: `solution' means a   solution to the problem, i.e. this differential equation.
264: Many of the ideas and results here are adaptations of those
265: pioneered in [Morse] and [Hadamard] XX Ref: use Had? . 
266:  
267: 
268: \proclaim Theorem 2.  Every infinite nonstuttering syzygy sequence with
269: the exception of the untied classes $\ldots ijij \ldots$  is realized by
270: a solution. 
271: 
272: {\bf Proof.} Section 6.2.   The method   is the classical one 
273: [Morse] of approximation
274: by periodic solutions.   
275: 
276: \vskip .1cm
277: 
278: 
279: 
280: \proclaim Theorem 3.  If a  
281: syzygy  sequence ends (begins) with 
282:  $ijij\ldots$  then any bounded solution which realizes
283: this sequence must  end (begin)
284: in  the  $ij$ collision.  
285: The untied sequences $\ldots ijij \ldots$ are not
286: realized by   any solution. 
287: 
288: {\bf Proof.} Section 6.3.
289:  \vskip .1cm
290: 
291: It is perhaps worth remarking that if a solution suffers 
292:   collision 
293: then it does so in   finite Newtonian time, but 
294: infinite `Jacobi time''.  
295: 
296: Inspired by theorem 3, we call the sequences appearing there
297: ``collision sequences''.  In more detail:
298: 
299: \proclaim Definition.  A 
300:   bi-infinite nonstuttering  syzygy sequence 
301: $s = \{s_j\}_{j = -\infty} ^{+\infty}$ 
302: is a forward  {\it collision sequence}
303:  if 
304: one of  its forward tails
305: $\{s_j\}_{j> N}$ contains only two letters.
306: Similarly, we have backward collision sequences.  A
307: collision sequence is one which is either forward
308: or backward collision sequence.  In the contrary case,  all three letters occur
309: in every tail, and the sequence is called {\it collision-free}.
310: 
311:  
312: \vskip .1cm 
313: 
314: Does every solution have a syzygy sequence?
315: If so, is this sequence unique?  In  [MontI]  I showed that
316: every bounded noncollinear zero-angular momentum solution to the Newtonian
317: three-body problem  suffers infinitely many
318: syzygies, provided the solution does not tend to triple collision,
319: {\bf and provided}  
320: binary collisions are counted as syzygies.  (See [Fuji] for another proof.)
321:   That proof  works verbatim for any  $1/r^a$
322: potential, $a > 0$, with the exception that
323: we must exclude   binary collisions. (They cannot
324: be regularized.)   Thus
325: every  bounded  {\bf collision-free} solution  has a syzygy sequence.  
326: That sequence must be nonstuttering in our equal mass case.
327: To prove that there is no stuttering,   use the fact that on a surface of negative
328: curvature any compact geodesic arc is the unique length minimizing curve among all
329: homotopic curves which share its endpoints. (See equation 6.3 and its derivation.)
330: Consequently, an application of  the method of reflection as exposed in
331: [ChM]  rids us of solution arcs representing stutters, i.e. solution arcs which
332:  hit  the same equatorial   arc
333: twice in a row.   These considerations allow us
334: to define a syzygy map from collision-free  sequences to 
335: infinite nonstuttering syzygy sequences. 
336: 
337: \proclaim  Theorem 4.   The
338: syzygy map from bounded solutions to   syzygy sequences
339: is a bijection between the set of   collision-free  solutions,   modulo
340: symmetry and time-translation,  and  the set of  bi-infinite nonstuttering collision-free syzygy sequences,
341: modulo shift.  
342: 
343: {\bf Proof.} Section 6.4.
344:  \vskip .1cm
345: 
346: Finally, we would like to know how much of phase space (the unit tangent
347: bundle of the pair of pants) is taken up
348: by the noncollision solutions.  Not much: 
349: 
350: \proclaim Theorem 5. Solutions tending to   binary collision are dense
351: within the space of all bounded solutions.  Thus the 
352: collision-free solutions have empty interior. 
353: 
354: {\bf Proof.} Section 6.5. 
355:  
356: \vskip .3cm 
357: {\bf Summary.} Putting  the theorems together gives a rather 
358:  complete symbolic dynamical
359: picture of the dyanmics of our problem -- the  zero-angular momentum
360: equal-mass $1/r^2$ three body problem restricted to the
361: bound orbits -- those with $I = const.$.  There are no linearly stable periodic orbits, by theorem 1.  
362: We will use the word ``bounded''
363: in the rest of this  paragraph to mean solutions  which tend to   to collision,
364: as these orbits are precisely the geodesics on the pair of pants which tend to infinity.
365:   
366:  Theorem 4 provides a  complete symbolic dynamics  
367: picture for the    bound  orbits : they are precisely the orbits none
368: of whose tails agree with the tied sequences $ \ldots ijij \ldots$.
369: The  closure of this set of orbits is the recurrent set.  The recurrence
370: set also   coincides  with  the closure of the
371: set of   periodic orbits.  There are unbounded orbits
372: on the frontier of this closure.   The situation is  similar to that of  
373: the recurrent set for the Kepler problem:  
374:   the space of periodic orbits is the recurrent set
375: and  contains
376: the unbounded parabolic orbits.    These unbounded  recurrent orbits   are ``just
377: barely unbound''  in that the collision condition $J_1 ^2 - (m_1 + m_2)  \le 0$  occuring
378: in  the appendix A, inequality (12A) is an equality on these orbits.    The complement of the
379: recurrent set   consists 
380: of orbits tending ``strongly'' to a binary collision. These
381: strongly colliding  orbits  form an open set. (Appendix
382: A.)  Finally,  the density result, theorem 6,  is an analogue of 
383:  what one would like to prove for the honest $1/r$ three-body problem:
384: that the set of solutions tending to infinity (via tight binary pairs)
385: is dense, for fixed energy and angular momentum.  M. Hermann
386: calls this density question  ``the oldest problem in dynamical systems
387: [Hermann] XX.  
388: 
389: 
390: {\bf Loose ends.} 
391: There are some collision orbits
392: which we have  left out of symbol sequence considerations.
393: The collinear solutions are not accounted for.  (Collinear
394: solutions should either  have no syzygy sequence or a continuum
395: of `$i$'s as their sequence, depending on one's taste.)  There are exactly six collinear solutions, two for each of the
396: three collision arcs, the two being related by reversing orientation.    
397: There are also collision orbits which end in collision but without the bodies winding around  infinitely often.
398: They `head straight in' to infinity  down one of the pants legs and their 
399: corresponding syzygy sequences will  truncate in the forward direction, for
400: forward time collision.
401: The simplest of these truncated solutions are the isosceles solutions.  
402: Again,  there are six of these by the same
403:   counting as for colliner solutions.  The
404: isosceles solution  $r_{ij} = r_{ik}$   begins and ends  at the $jk$
405: collision, and has exactly one syzygy in between,  the Euler
406: point in which $i$ is at the midpoint of $j$ and $k$. 
407: Its syzygy sequence is the single letter 
408:  `$i$'. Interpolating between collinear and isosceles is a  
409: one-parameter family of  solutions whose  syzygy sequences truncate.    
410: These interpolating solutions are the $\lambda$-curves
411: of  eq. (3.13), the   curves of constant $\chi$, in
412: the $\lambda, \chi$ coordinate system  there.  
413: I do not know if the syzygy sequences of these  
414: solutions are finite, or one-sided infinite.  
415: 
416: 
417:  
418: \vskip .3cm 
419: 
420: 
421: 
422: {\bf Open questions.} {\bf 1.} Can two distinct collision orbits
423: share the same  syzygy sequence?  
424: 
425: {\bf 2.} Are there any solutions  besides isosceles which have a finite syzygy sequence?
426: If so, can any finite syzygy  sequence occur as the syzygy sequence of
427: some collision orbit?  
428: 
429: \vskip 1cm
430: 
431: \vskip 1cm
432: 
433:  
434: 
435: 
436: \vskip 1cm
437:  
438:  {\bf 3. Set-Up and Proof of Prop. 1} 
439: 
440: Write $x = (x_1, x_2, x_3) \in \R^6$ with $x_i \in \R^2$
441: for the positions of the three bodies,
442: and $r_{ij} = \|x_i - x_j\|$
443: for the distances between them.  The
444: potential is $-U$ where  
445: $$U = \Sigma m_i m_j/ r_{ij} ^2 \hskip 1cm (3.1).$$
446: The $m_i$ are the masses. 
447: Set
448: $$
449: \eqalign{ K  &= \Sigma m_i \| \dot x_i\|^2  \cr
450: &  = \langle \dot x , \dot x \rangle, \hskip 1cm (3.2) 
451: }$$
452: for twice the kinetic energy. The last
453: equality of (3.2) defines  the ``mass inner product'' on the
454: three-body configuration space $\R^6$.  The total energy 
455: $$H = K/2 - U \hskip 1cm (3.3)$$
456:  is constant along solutions.   The 
457: equations of motion, 
458:  $  \ddot x_i = -2 \Sigma_{i \ne j} m_j (x_i -x_j)/
459: r_{ij} ^4$, $i = 1,2,3$,  can be written
460: as the single vector equation
461: $$\ddot x = \nabla U \hskip 1cm (3.4)$$
462: where $\nabla U$ is defined using the
463: mass inner-product: 
464: $dU(x)(v) = \langle \nabla U(x), v \rangle$.
465: 
466: By the standard method of freshman physics, we can,
467: without loss of generality restrict our
468: considerations to motions for which
469:  $$\Sigma m_i x_i = 0 \hskip 1cm (3.5)$$
470: throughout.
471: This constraint defines a four-dimensional
472: real vector space which can be identified
473: with the two-dimensional complex space
474: $\C^2$ in such a way that counterclockwise rotation
475: of a triangle $(x_1, x_2, x_3)$
476: by $\theta$ radians  turns into scalar
477: multiplication of the corresponding complex vector by
478: $exp(i \theta)$. 
479:  Set  
480: $$ 
481: \eqalign{I & = \Sigma m_i m_j r_{ij}^2/ \Sigma m_i \cr
482: & = \langle x , x \rangle 
483: } \hskip 1cm (3.6)
484: $$
485: where the last equality is only true when 
486: the center of mass constraint (3.5) is in place.  
487: Using $\dot I = 2 \langle x ,\dot x \rangle$,
488: $\ddot I = 2 \langle \dot x, \dot x \rangle + 
489: 2 \langle x , \ddot  x \rangle$ and
490: $\langle x, \nabla U (x) \rangle = - 2 U(x)$
491: (by $U$'s homogeneity) 
492: we obtain  the Lagrange-Jacobi identity: 
493:  $$\ddot I = 4 H \hskip 1cm (3.7), $$
494: valid along any solution. 
495: Thus 
496: $I(t) = const.$  along the solution
497: if and only if   $H = 0$  and $\dot I (0) = 0$
498: for that solution.    
499: 
500: 
501: We will call a solution ``bounded'' if
502: the $r_{ij}$ are bounded as functions of time,
503: and  
504: do not simultaneously tend to zero, i.e. to triple collison.    
505: Now $I \to \infty$ if and only if one
506:  of   the $r_{ij}$ tend to infinity,
507: and   $I \to 0$ if all $r_{ij} \to 0$.
508: It follows from (3.7)  that
509: every bounded solution must satisfy $H = 0$,
510: $\dot I (0) = 0$, and $I(t) = const.$.
511:   
512:  The scaling symmetry
513: $x(t) \mapsto \lambda ^{-1/2}x (\lambda t)$
514: takes solutions to solutions, preserves zero energy,
515: and takes $I$ to $I/\lambda$.  Using this scaling, 
516: we may, without loss of generality,
517: assume that  $I = 1$ in studying   bounded solutions.
518: The set $I = 1$, $\Sigma m_i x_i = 0$
519: forms a three-sphere in the $\C^2$.  
520: {\bf We have reduced 
521: the study of the bounded solutions to the $1/r^2$ problem
522: to 
523: a second order dynamics on this three-sphere.}
524: It is well-known 
525: ([Arn] or [AbMar]) that a 
526: dynamics on a constant energy surface 
527: $H  = E$
528: level set is equivalent to   
529: geodesic flow for the Jacobi-Maupertuis metric 
530: $( E + U) ds^2$
531: where $ds^2$ is the kinetic energy metric.
532: In our case, $E = 0$, and 
533: we restrict the kinetic energy to the sphere $I = 1$.  Consequently, 
534: the study of bounded solutions is
535: equivalent to the study of geodesics on
536: the three-sphere under the metric 
537: $ds^2_J =  U ds^2$
538: conformal to the standard metric $ds^2$ on
539: that three-sphere.   
540:  
541: To obtain a metric
542: on the shape sphere, we quotient
543: by rotations.   We review the discussion
544: in [MontN], [ChMont], or [MontR] on
545: this metric. See especially the appendix
546: of [MontR] for explicit computations 
547: and derivations.  The group of rigid rotations acts  on the three-body
548: configuration space   
549: according to scalar multiplication on $\C^2$ by unit modulus complex scalars.
550: Restricting ourselves to   
551: the three-sphere $S^3:= \{ I =1 \}   \subset \C^2$ and forming 
552: the quotient by this rotational action yields the
553: famous Hopf fibration 
554: $$S^3 \to S^2 = S^3/ S^1 \hskip 1cm (3.8) .$$
555: The quotient two-sphere is
556: the  shape sphere ([ChM], [MontR] esp. the appendix).
557: Its points represent oriented similarity classes
558: of triangles. 
559: Both the  dynamics and  the  Jacobi metric $U ds^2$
560: on the three-sphere   descend under the projection (3.8)
561: to the shape sphere  
562: once we fix the value of the total angular momentum.   
563: The total angular momentum of a solution is zero if and only if
564: that solution is orthogonal to the rotational orbits,
565: i.e. orthogonal to the fibers of (3.8).  The projection of 
566: such a zero-angular momentum  
567:  solution under (3.8)
568: is  a geodesic for the quotient metric.
569: (See [Hermann], lemma 4.1. His situation is  more general
570: than ours. Our circle  bundle (3.8) is replaced
571: by a general Riemannian submersion.) 
572: We can write the Jacobi-Maupertuis metric on the
573: shape sphere as  
574: $$ds^2_J = U ds_{shape}^2 \hskip 1cm (3.9a)$$
575:  where $ds_{shape}^2$ is
576: the kinetic-energy induced metric on shape sphere,
577: and $U$ is the (negative) potential (3.1) restricted
578: to $I =1$ and then viewed as a fucntion on the shape-sphere
579: (possible because of its rotation invariance.
580: The shape sphere metric is the   round metric on a sphere of radius $1/2$:
581: $$ds^2 _{shape} = ({1 \over 2})^2 [d \phi ^2 + \cos^2 (\phi) d
582: \theta ^2) ] \hskip 1cm (3.9b)$$
583: where
584:  $\phi$ is the colatitude --
585: the angle from the equator, and $\theta$
586: is labels longitudinal circles on the sphere.
587: \vskip .4cm
588: \vskip .4cm
589: {\bf Proof of Proposition 1.} The 
590: discussion of the last two paragraphs shows
591: that,  modulo rotations, translation, 
592: and scaling, the set of bounded 
593: zero-angular momentum solutions
594: for the negative potential (3.1) are in bijective correspondence 
595: with   geodesics
596: for the Jacobi-Maupertuis metric on the shape sphere
597: minus the three binary collision.   
598:  Under this correspondence the
599: geodesic flow for the metric corresponds,
600: after a time reparameterization, to the
601: the flow defined by Newton's equations. 
602: It remains to verify the claims about the completeness
603: and  that the ends asymptote to cylinders.
604:   
605: \vskip .4cm
606: 
607: {\bf Completeness.}
608: 
609: Let $\rho = \rho_{ij}$ be the spherical
610: distance from  the $ij$ collision point $C_{ij}$,
611: as measured in the spherical metric  $ds^2_{shape}$.
612: Then as $\rho \to 0$ we will show that 
613: $$U = {C^2 \over \rho^2} + O (1)  \hskip 1cm
614: (3.10a) $$
615: for some positive constant $C^2$, while 
616: $$ds^2 _{shape} = d \rho^2 + (\rho^2 + O(\rho^4)) d \chi^2
617:  \hskip 1cm (3.10b)$$
618: where  $\chi$ is  the  angular coordinate based at $\Sigma_{ij}$ 
619: so that $(\rho, \chi)$ are geometric polar
620: coordinates.   It follows that the Jacobi metric has
621: the expansion:
622: $$ ds^2_J =   {C^2 \over \rho^2} (d \rho^2 + \rho^2 d \chi^2)) + O(1) \hskip 1cm
623: (3.10c).$$ It follows
624: that if we  approach the collision $\rho = 0$ along any curve then the
625: length of that curve diverges at least as fast as the integral
626: of $C\sqrt{d \rho^2/ \rho^2} = Cd\rho/\rho$, that is,  it  
627: diverges logarithmically as $C|\log(\rho)|$ as $\rho \to 0$.
628: Consequently any  curve tending towards ``infinity''
629: i.e. to one of the binary collisions,   has infinite length,
630: which proves completeness.  
631:  
632: 
633: To establish (3.10a), it suffices to establish
634: $$r_{ij} = {1 \over {\sqrt{\mu_{ij}}}}  \sin(\rho_{ij})
635: \hskip 1cm (3.11)$$
636: where $\mu_{ij} = m_i m_j/(m_i + m_j )$ is the reduced mass.
637: The other two distances $r_{ik}, r_{jk}$ are bounded
638: away from zero as $r_{ij} \to 0$, due to 
639: the constraint $I =1$
640: (see (3.6)).  Then (3.10a) follows from (3.1)
641:   and
642: the Taylor expansion of $\sin(\rho)$.  The constant
643: $C$ in (3.10a) is $m_i m_j / \sqrt{\mu_{ij}}$.
644: 
645:  To establish (3.11) we work in 
646: the full three dimensional shape space which is the space whose points are
647: oriented congruence classes of planar triangles.  The full shape space is 
648: isometric to the   cone over the shape sphere and consequently
649: distances $d$ in the full shape space can be obtained 
650: from spherical distances together with knowldge of
651: the distance $R = \sqrt{I}$  from the cone point. 
652: Write $d_{ij}$ for the distance in the full shape space between
653: an arbitrary point and   the $ij$ binary collision {\it ray}.
654: Then we have  
655: $$r_{ij}  = {1 \over {\sqrt{\mu_{ij}}}} d_{ij} \hskip 1cm (3.12a) $$
656: and
657: $$d_{ij} =  R \sin(\rho_{ij}) \hskip 1cm (3.12b)  .$$
658: Upon setting $R= 1$, (3.11) follows immediately.
659: Equation (3.12a,b) can be 
660:  found in section 4, equations (4.3.15a,b) of [MontN]. However, note that  there
661: is a typo in eq 4.3.15a. The $\mu_{ij}$ in that equation must be replaced by
662: $\sqrt{\mu_{ij}}$.)   
663:  
664: To get (3.10b)  use the fact that the shape sphere is isometric
665: to the sphere of radius $1/2$ and that the metric on such a sphere is given by 
666: $$ds^2_{shape} = d \rho^2 + [(1/2) \sin(2\rho))]^2 d\chi^2 \hskip 1cm (3.13).$$
667: in spherical-polar coordinates. Then use the Taylor expansion of
668: $(1/2) \sin(2 \rho)$. 
669: 
670: \vskip .4cm 
671: 
672: {\bf Asymptotes to Cylinders.}  We 
673: use a more precise version of the expansion (3.10c).
674: Set 
675: $$d\lambda = - \sqrt{U} d \rho \hskip 1cm (3.14).$$
676: Integrating (3.14) defines   a  function 
677: $\lambda = \lambda(\rho, \chi)$ such thta 
678:  $\lambda \to \infty$
679: as  the collision $\rho = 0$ is approached. 
680: From (3.10a) we have  $d\lambda = - C d \rho / \rho  + O(1)$
681: from which it follows that 
682: $$\rho = e ^{ -C \lambda} + o(\rho).$$
683: The  $(\lambda, \chi)$ are coordinates for the end $\rho = 0$,
684: and in these coordinates
685: $$d\bar s^2 = d \lambda^2 + f(\lambda, \chi)^2 d \chi^2
686: \hskip 1cm (3.15)  $$
687: where, from (3.13) and (3.9a) we have  
688: $$f^2  = ({1 \over 2} \sin (\rho) )^2 U.$$
689: Now $({1 \over 2} \sin  ( 2\rho) )^2 =  \sin^2 \rho \cos^2 \rho$
690: so that from (3.11)  
691: $ ({1 \over 2} \sin (2 \rho) )^2 =  \mu_{ij}
692: r_{ij}^2 \cos^2 (\rho )$
693:  and
694: $$f^2 =  \mu_{ij}  \cos^2 (\rho)\{ m_i m_j  + m_i m_k {{r_{ij}^2} \over
695: {r_{ik}^2}} 
696:  +m_j m_k {{r_{ij}^2} \over {r_{jk}^2}} \} \hskip 1cm (3.16)$$
697: where $ijk$ is a permutation of $123$. 
698: As we approach the collision $\lambda = \infty$
699: we have $r_{ij} \to 0$ while $r_{ik}, r_{jk}$ remain bounded
700: since we are constrained to $I =1$.  Thus
701: $$\lim_{\lambda \to \infty} f = \sqrt{\mu_{ij}  m_i m_j}: = K_{ij} > 0 \hskip 1cm (3.17).$$ 
702: Summarizing:  
703: $$d   s^2 _J  = d \lambda^2 + (K_{ij} ^2 + O( e^{-2 C  \lambda})) d \chi^2 
704: \hskip 1cm (3.18)$$ 
705: which says the metric asymptotes to a Euclidean cylinder of
706: radius
707: $K_{ij}$ as we approach the $ij$ end.  
708: 
709:  QED 
710: \vskip .1cm
711: {\bf Remark.}  It follows from equations (7.12, .13)
712: that   the Gaussian curvature near the end is
713: negative.  But for any metric
714: of the form (3.15) this curvature is equal to $ -{1 \over f}{{\partial ^2
715: f}
716: \over {\partial \lambda ^2}}$.  Thus, for 
717: fixed $\chi$,   the 
718: function $f(\lambda, \chi)$ is a  strictly convex
719: of $\lambda$, for all $\lambda$ from some point on, 
720: and from this point on,   
721: $f(\lambda, \chi)$  monotonically decreases to $K_{ij}$.
722: \vskip .4cm
723:  
724: \vskip 1cm 
725: 
726: {\bf 4.  Curvature. Proof of theorem 1.}
727: 
728: We proceed to the proof of our main result, Theorem 1,
729: the negativity of the Gaussian curvature   when the masses are equal.
730: The computation   proceeds 
731: through a series of lemmas.  The first  is standard and we will
732: not provide the proof. 
733: 
734: \proclaim Lemma 4.1.  Let a  surface 
735: be endowed with   conformally related  metrics
736: $ds^2$ and $d \bar s^2  = U ds^2$.
737: Then their curvatures $K, \bar K$ 
738: are related by
739: $$\bar K = U^{-1} (K - {1 \over 2} \Delta \log(U))$$
740: where the Laplacian $\Delta$ is with respect to the
741: $ds^2$ metric.
742: 
743: 
744: The curvature $K$ of the standard shape metric  $ds^2 = ds^2 _{shape}$  of (3.9b) is 
745: $K = 4$. According to lemma 4.1
746: $$\bar K = 4 - {1 \over 2} \Delta (\log(U) \hskip 1cm (4.1)$$
747: is the desired curvature, the curvature
748:  of  the  Jacobi-Maupertuis metric (3.9a) 
749: of proposition 1 and Theorem 1.
750: A routine computation yields
751: $$\Delta (\log(U)) = {{ U \Delta U - \|\nabla U \|^2} \over
752: {U^2}} \hskip 1cm (4.2) .$$
753: Here, and throughout this section,  $U$ is considered  
754:   as a function on the shape sphere,  
755: $\Delta U$ is its Laplacian with respect to the standard
756: shape space metric metric $ds^2$  
757:    and  
758: $\| \nabla U \|^2$  is the squared length of its gradient with
759: respect to the same metric. 
760: 
761: A key to the subsequent computations
762: is to use the squared length coordinates as in [AlbCh] 
763: $$s_k = r_{ij}^2,  ijk \hbox{ a permutation of } 123,  \hskip 1cm (4.3)$$
764: rather than the lengths $r_{ij}$ themselves. 
765: Write 
766: $$U_{2n} = \Sigma 1/r_{ij}^{2n} = \Sigma 1/s_k^n \hskip 1cm (4.4)$$
767: so that $U = U_2$.
768: \vskip .2cm
769: 
770: \proclaim Lemma 2.
771:   $$\Delta U = 8U_4 \hskip 1cm (4.5) .$$
772: 
773: {\bf Proof. } Section 5.2.
774: 
775: \vskip .2cm
776: 
777: \proclaim Lemma 3.
778: $$\| \nabla U \|^2 = 4 S \hskip 1cm (4.6a)$$
779: where
780: $$S = 2U_6 -  U_4 -{3/2} \Sigma^{\prime} 1/s_i ^2 s_j ^2 +  2 \Sigma^{\prime} 1/s_i  s_j ^2
781: - \Sigma^{\prime} 1/s_i  s_j  
782: \hskip 1cm (4.6a)$$
783: and where ``$ \Sigma^{\prime}$ '' means to sum over all indices 
784: $i, j$ with $i \ne j$. (For example  $ \Sigma^{\prime} s_i s_j = 2s_1 s_2 + 2 s_2 s_3 + 2 s_3 s_1$,
785:  twice the second symmetric polynomial in the $s_i$.)
786: 
787: {\bf Proof. } Section 5.3.
788: 
789: \vskip .2cm
790: 
791: {\bf Proof of the Negativity of the curvature.}
792: Combining the equations (4.1), (4.2), (4.5) and (4.6a) 
793: we find 
794: $$ -  \bar K U^3 =  4 U U_4 -  4 U ^2  -  2 S \hskip 1cm (4.7).$$
795: Expand out the first two terms on the right hand side:
796: $$\eqalign{ U U_4 & = \Sigma 1/s_i \Sigma 1/s_j^2 \cr
797: &= \Sigma 1/s_i ^3 + \Sigma^{\prime} 1/s_i s_j ^2 \cr
798: & = U_6 + \Sigma^{\prime} 1/s_i s_j ^2 .
799: } \hskip 1cm (4.8) 
800: $$
801: while
802: $$\eqalign{ U ^2 & = \Sigma 1/s_i \Sigma 1/s_j \cr
803: &= \Sigma 1/s_i ^2 + \Sigma^{\prime} 1/s_i s_j  \cr
804: & = U_4 + \Sigma^{\prime} 1/s_i s_j  .
805: } \hskip 1cm (4.9) 
806: $$ 
807: Plugging  (4.8), (4.9) and equation (4.6b)  
808:  back into equation  (4.7) yields :  
809: $$ 
810: -  \bar K U^3   = -2 U_4  - 2 \Sigma^{\prime} 1/s_i  s_j  + 3  \Sigma^{\prime} 1/s_i ^2  s_j ^2 
811: \hskip 1cm (4.10a).$$
812: Use the fact that $U_4 + \Sigma^{\prime} 1/s_i  s_j = (\Sigma 1/s_i)^2 = U^2$ 
813: to rewrite the right hand side of (4.10), and divide the resulting equation
814:  by two in order to  obtain
815: $$ 
816: - \bar K U^3   = 3  \Sigma^{\prime} 1/s_i ^2  s_j ^2  - 2 U^2  \hskip 1cm (4.10b)
817: $$
818: Consequently, $\bar K \le  0$ if and only if  
819: $$3(\Sigma^{\prime} 1/s_i ^2 s_j ^2) \ge 2 ( \Sigma 1/s_i)^2 \hskip 1cm (4.11)$$
820: To prove (4.11), multiply both sides of it by $s_1 ^2 s_2 ^3 s_3 ^2$
821: thus arriving at $6 (\Sigma s_i ^2) \ge 2 \sigma_2 ^2$ or
822: $$3 (\Sigma s_i ^2) \ge  \sigma_2 ^2 \hskip 1cm (4.12)$$
823: where $\sigma_2 = s_1 s_2 + s_2 s_3 + s_3 s_1$
824: is the second elementary symmetric polynomial in the $s_i$.
825: (The coefficient $6$ arose because for each pair $ij$ there
826: are two terms in the sum $\Sigma^{\prime}$.  See the parenthetical
827: remark in  lemma 3.)
828: To prove (4.12), remember that we are restricting ourselves
829: to the sphere $I = 1$ and that $I = \Sigma s_i /3$.  Thus we 
830: can homogenize the equation by using that $3 =  (\Sigma s_i )^2 /3$ 
831: on the sphere.  So, the desired inequality now reads:
832: $${{(\Sigma s_i )^2} \over 3}(\Sigma s_i ^2) \ge  \sigma_2 ^2.$$
833: The  two inequalities:
834: $$\Sigma s_i ^2 \ge \sigma_2 \hskip 1cm  \hskip 1cm (4.13A)$$
835: and 
836: $${{(\Sigma s_i )^2} \over 3} \ge \sigma_2 \hskip 1cm (4.13B) $$
837: are classical, with 
838: equality  in {\bf either case} if and only
839: if all the $s_i$ are equal.   Here are the proofs. 
840:  Inequality (4.13A) follows simply upon 
841: rearranging  the inequality 
842: $(s_1 - s_2)^2 + (s_2 - s_3)^2 + (s_3 - s_1)^2 \ge 0$.
843:  Inequality (4.13B) is a special case of a general
844: inequality among the elementary symmetric polynomials evaluated at 
845: positive arguments $s_i$.  See for example the 
846: Encyclopaedia [Math],  
847: App. A, Table 8, inequality (4).  Alternatively, expand
848: out $(\Sigma s_i)^2 = \Sigma s_i^2 + 2 \sigma_2$
849: and use  inequality (4.13A).
850: Multiplying the two inequalities
851: yields the desired inequality (4.12),
852: with equality if and only if we are at the Lagrange points
853: $s_1 =s_2 = s_3$ of the shape sphere.
854: Since (4.12) is equivalent to the curvature inequality, 
855: Theorem 1 is proved, modulo the proofs of lemmas 2 and 3
856: which follow in the next section.   QED
857: 
858: \vskip 1cm
859:   
860: 
861: {\bf 5. Proofs of lemmas 2 and 3.}
862: 
863: {\bf 5.1. Notation.}
864:   
865: 
866: To compute $\Delta U$ and
867: $\|\nabla U \|^2$ we must express
868: the squared distances $s_k = r_{ij}^2$ of (4.3) 
869: in terms of the  spherical coordinates $(\phi, \theta)$ of (3.9b). 
870: In [MontI] I prove that upon restriction to the sphere $I = 1$ 
871: $$s_k = 1 - \cos(\phi) \gamma_k (\theta) \hskip 1cm (5.1.1b)$$
872:  where 
873: $$\gamma_1 (\theta) = \cos(\theta), \gamma_2 (\theta) =
874: \cos(\theta + 2 \pi/3), \gamma_3 (\theta) = \cos(\theta + 4
875: \pi/3) \hskip 1cm (5.1.2).$$ The special angles $\theta = 0, 2 \pi/3, 4 \pi/3$
876: mark the locations of the three binary collision
877: on the equator $\phi = 0$ of collinear triangles. 
878: Later on we will use the 
879: fact that the three planar vectors
880: $(\gamma_k, \gamma_k^{\prime})$, $k =1,2,3$ form the vertices of an equilateral
881: triangle inscribed in the unit circle.  
882: \def\dphi{\partial_{\phi}}
883: \def\dtheta{\partial_{\theta}}
884: {\bf Here and throughout we   write  $\gamma_i ^{\prime}$
885: for the derivative $\dtheta \gamma_i$
886: of $\gamma_i$ with respect to $\theta$.}
887: 
888: \vskip .2cm  
889: 
890: {\bf 5.2. Proof of Lemma 2.} 
891: Write $c = \cos(\phi)$,
892: $\dphi$ for the partial derivative with respect
893: to $\phi$ and $\dtheta$ for the partial derivative
894: with respect to $\theta$.  Then 
895: $$\Delta U = {4 \over c} \dphi (c \dphi U) + {4 \over c^2} (\dtheta
896: ^2 U). \hskip 1cm (5.2.1)$$
897: And
898: $$\dphi (1/s_i) = -s\gamma_i /s_i ^2 \hskip 1cm (5.2.2a)$$
899: $$\dtheta(1/s_i) = c\gamma_i ^{\prime}/s_i ^2 \hskip 1cm (5.2.2b). $$
900: Since  
901: $U = \Sigma 1/s_i$
902: we have
903: $$
904: \eqalign{
905: \dphi(c \dphi U)  & = \dphi c \Sigma (-s\gamma_i)/s_i ^2
906: \cr
907: &= \dphi(-cs \Sigma \gamma_i)/s_i ^2  \cr
908:      &= (-c^2 + s^2)\Sigma \gamma_i /s_i ^2 
909:  + 2c s ^2 \Sigma \gamma_i ^2/ s_i^3}.
910:  $$
911: Thus
912: $$
913: {1 \over c}
914: \dphi(c \dphi U) = ((s^2 - c^2)/c)\Sigma \gamma_i /s_i ^2 
915:  + 2 s^2 \Sigma \gamma_i ^2/ s_i^3 .\hskip 1cm
916: \hskip 1cm (5.2.3)$$
917: And
918: $$\eqalign{{1 \over c^2} \dtheta \dtheta U 
919: & = {1 \over c^2} \dtheta \Sigma c \gamma_i ^{\prime}/s_i ^2
920: \cr
921: & = {1 \over c^2} \Sigma c [\gamma_i ^{\prime \prime}/s_i ^2
922: + 2 (c \gamma_i )^2/ s_i ^3
923: \cr
924: & = {1 \over c} \Sigma \gamma_i ^{\prime \prime}/s_i ^2
925: + 2 \Sigma (\gamma_i ^{\prime})^2/s_i ^3
926: }
927: \hskip 1cm (5.2.4)$$
928: Now $\gamma_i^{\prime \prime} = - \gamma_i$
929: so that
930: $${1 \over c^2} \dtheta \dtheta U =
931: -{1 \over c} \Sigma \gamma_i /s_i ^2 
932:  + 2 \Sigma (\gamma_i ^{\prime})^2/s_i ^3 \hskip 1cm (5.2.5)
933: $$
934: 
935: 
936: Adding 4 times (5.2.4) to 4 times (5.2.5) we get
937: $$\Delta U = 
938: 4 ((s^2 - c^2 - 1)/c ) \Sigma \gamma_i /s_i ^2  
939: +  8 s^2 \Sigma \gamma_i ^2/ s_i^3   + 
940: 8 \Sigma (\gamma_i ^{\prime})^2/s_i ^3 \hskip 1cm (5.2.6) $$
941: Now
942: $s^2 -c^2 - 1 = -2c^2$ so that
943: $$\Delta U = -8 \Sigma c \gamma_i /s_i ^2  
944: + 8 \Sigma [s^2 \gamma_i ^2 + (\gamma_i ^{\prime})^2]/s_i ^3
945: $$
946: Recalling that  $\gamma_i ^2 + (\gamma_i ^{\prime})^2 =1$
947: (the vectors $(\gamma_i,  \gamma_i ^{\prime})$
948: define an equilateral triangle inscribed in the unit circle)
949: we see that we can replace  $(\gamma_i^{\prime})^2$ by $1 - \gamma_i ^2$
950: in order to  obtain
951: $$\eqalign{
952: s^2 \gamma_i ^2 + (\gamma_i ^{\prime})^2
953:  & = (s^2 -1) \gamma_i ^2  +1 \cr
954: & = - c^2 \gamma_i ^2 +1   \cr
955: & = -(1-s_i)^2 +1  \cr
956: & = 2s_i -s_i^2 \cr
957: & = s_i (2 -s_i)
958: }
959: \hskip 1cm (5.2.7) $$
960: Then 
961: $[s^2 \gamma_i ^2 + (\gamma_i ^{\prime})^2]/s_i ^3
962: = (2 -s_i)/s_i^2 = (c \gamma_i +1 )/s_i^2$
963: where I used $1 -s_i = c \gamma_i$
964: It follows that
965: $$
966: \Delta U = -8  \Sigma c \gamma_i /s_i ^2  
967: + 8 \Sigma c \gamma_i /s_i ^2 +  8 \Sigma (1/s_i ^2)
968: = 8U_4
969: $$
970: as claimed.
971:  
972: \vskip .4cm
973: {\bf 5.3. Proof of Lemma 3.}
974: 
975:  
976: We have  
977: $$\eqalign{
978: dU  &= \dphi U d \phi + \dtheta U d \theta \cr
979:    & = ( \Sigma (-s \gamma_i /s_i ^2  )d\phi + ( \Sigma (c \gamma_i ^{\prime})/s_i^2) ) d \theta
980: } \hskip 1cm (5.3.1) 
981: $$
982: Now $\| \nabla U \|^2 = \|d U\|^2$.  The length squared of
983: the covector $dU$ is computed relative to the metric `$g^{ij}$ '
984: induced on covectors, which, from (3.9) is 
985: given at the  point $(\phi, \theta) $
986: by  $\|a d\phi + b d \theta \|^2 = 4( a ^2 + {1 \over c^2} b^2)$,
987: with $c = \cos(\phi)$.
988: It follows that
989: $$
990: \eqalign{
991: \|\nabla U \|^2  & =  4(\Sigma s \gamma_i/s_i^2)^2 +  4 (\Sigma \gamma_i ^{\prime}/s_i ^2 )^2 
992: \cr
993: & =  4 \big( \Sigma s^2 \gamma_i ^2 /s_i^4 + \Sigma^{\prime} s \gamma_i s \gamma_j / s_i^2 s_j^2 
994: + \Sigma \gamma_i ^{\prime ^2} /s_i^4 + \Sigma^{\prime} \gamma_i ^{\prime} \gamma_j ^{\prime} / s_i ^2 s_j ^2 
995: \big)
996: \cr
997: & = 4 ( \Sigma (s^2 \gamma_i ^2 + (\gamma_i ^{\prime 2} ) /s_i ^4 +
998: \Sigma^{\prime} (s^2 \gamma_i  \gamma_j + \gamma_i ^{\prime} \gamma_j ^{\prime})/ s_i^2 s_j^2 ).  
999: } \hskip 1cm (5.3.2) 
1000: $$
1001: Simplify the numerator
1002: in the first summand of the last equation
1003: by  using (5.2.7). 
1004:  We will simplify the numerator
1005: of the second summand  by using  an analogous identity for
1006: $s^2 \gamma_i  \gamma_j + \gamma_i ^{\prime} \gamma_j ^{\prime}$, $i \ne j$.
1007: Indeed, since the vectors $(\gamma_i, \gamma_i ^{\prime})$
1008: form the vertices of an equilateral triangle inscribed within
1009: the unit circle,we have that 
1010: $\gamma_i \gamma_j + \gamma_i ^{\prime} \gamma_j ^{\prime} = -1/2$
1011: for $i \ne j$, since  $-1/2 = \cos(2\pi/3)$ is the cosine
1012: of the central angle defined by any two vertices of an equilateral triangle. Thus 
1013: $$\eqalign{
1014: s^2 \gamma_i  \gamma_j + \gamma_i ^{\prime} \gamma_j ^{\prime}
1015: &= \gamma_i  \gamma_j + \gamma_i ^{\prime} \gamma_j ^{\prime} - c^2 \gamma_i \gamma_j
1016: \cr
1017: &=  -1/2 - (1- s_i)(1-s_j) \cr
1018: &= -3/2 + s_i + s_j - s_i s_j \hskip 1cm  
1019: }. \hskip 1cm (5.3.3)
1020: $$
1021: Plugging (5.3.2) and (5.3.3) into (5.3.1) we get
1022: $$
1023: \eqalign{
1024: \|\nabla U \|^2  & =  4 ( 2 \Sigma 1 /s_i ^3 - \Sigma 1/s_i ^2 
1025: -{3 \over 2} \Sigma^{\prime} 1/ s_i ^2 s_j^2 + 2 \Sigma^{\prime} 1/s_i s_j ^2  - \Sigma^{\prime} 1/s_i s_j )  
1026: \cr
1027: &= 4 S}
1028: $$
1029: as claimed. 
1030: QED
1031: 
1032: 
1033: 
1034: \vskip .4cm
1035: 
1036: {\bf 6.  Dynamical Consequences.} 
1037: 
1038: {\bf 6.1}  
1039: 
1040: 
1041: 
1042: Let $P$ denote the pair of pants.  
1043: Using the spherical shape metric (3.9a),
1044: construct three disjoint circles with centers at
1045: the three binary collision points.  Delete the open discs
1046: bounded by these circles to obtain a   compact region $R \subset P$ 
1047: having for its   boundary  the three disjoint circles.   Refer back to  figure 1.
1048: 
1049:  
1050:  \proclaim Lemma 6.1. Any  arc in $P$ whose (finite) syzygy sequence 
1051: contains all  three letters   must cut through  $R$.
1052: 
1053: {\bf Proof.} Let $c$ be such a `123'  arc.
1054: If one of $c$'s syzygies is within $R$, we are done.
1055: Otherwise, all three syzygies lie within the three excised discs.
1056:  But  all three syzygies cannot be   in the same disc,
1057: since each disc contains exactly  two syzygy types. Thus
1058: $c$ must   must travel  from one disc to the other, and in 
1059:  so doing it cross into $R$.  
1060:  
1061: QED.
1062: 
1063: \vskip .3cm
1064: 
1065: 
1066: {\bf 6.2. Proof of Theorem 2.} Let $s$ be 
1067: syzygy sequence containing all three letters.
1068: Approximate $s$ by a sequence $w^N$, $N = 1,2, 3$ 
1069: of periodic sequences as follows.   
1070: Truncate   $s$ to form the finite even length
1071: subword 
1072: $w_N = s_{-N +1} s_{-N +1} \ldots s_N$.   
1073: Turn this subword into a periodic sequence
1074: $ \ldots w_N w_N w_N \ldots$ by repeating it
1075: in blocks.  If the resulting word
1076: has stutters at the join,  shift the ``window''
1077: we used to form $w^N$ so as to form the word 
1078: $w_{N, j} = s_{-N + j +1} s_{-N + j} \ldots s_{N + j}$
1079: along with its  corresponding  periodic word.  
1080: We can always find a $j$ so that the resulting periodic word,
1081: call it $w^N$,  is non-stuttering.
1082: The sequence 
1083: $w_N$ 
1084: contain all three letters $123$ for all $N$
1085: sufficiently large, since $s$ itself contains all three letters.  
1086: Theorem 2 implies that  the $w_N$, for $N$ large,  
1087: are  represented by a unique geodesic
1088: $\gamma_N$. By lemma 6.1 the $\gamma_N$  must cut through $R$.  
1089: Shifting time (and thus shifting the sequence) if necessary, we may assume that
1090: $\gamma_N (0) \in R$.  Since $R$ is compact, 
1091: so is  the unit tangent bundle to $P$   
1092: (relative to the Jacobi-Maupertuis metric) over $R$.   
1093: The pairs
1094: $(\gamma_N (0), \dot \gamma_N (0)$ lie
1095: in this compact space, so we  
1096: we can find a subsequence of them   which 
1097:   converge to some initial condition $(q,v)$.
1098: The geodesic with initial condition $(q,v)$
1099: realizes the   infinite sequence
1100: $s$. This establishes the existence of a  solution 
1101: realizing the   syzygy sequence $s$.
1102: 
1103: QED
1104: 
1105: \vskip .2 cm
1106: 
1107: FIGURE, SECTION AND EQ RELABELLING NEEDED  BELOW   XX
1108: 
1109: 
1110: {\bf 6.3. Collision sequences. Proof of theorem 3.} 
1111: 
1112: Let $s$ be a collision sequence.  
1113: We may  assume, 
1114: without loss of generality,  that it is a forward
1115: collision sequence, and that the two letters in its  
1116: forward tail are  $1$ and $2$.  The first part of the theorem
1117: asserts that  any solution  which realizes $s$
1118: must satisfy $r_{12} \to 0$ as
1119: the Jacobi time  $t \to \infty$.
1120: (In Newtonian time  the  collision  occurs in finite time.
1121: As the     two bodies get closer,
1122: the third body  effects them less and less.
1123: A straightforward  analysis of the  two-body $1/r^2$
1124: problem shows that 
1125: $\lim \inf r_{12} = 0$ if and only if
1126: $\lim r_{12} =0$.  The perturbation
1127: of the third body does not affect this assertion.
1128: Thus in order to prove the solution realizing $s$ suffers
1129: collision it suffices to show that  
1130: $$\lim \inf r_{12} = 0 \hskip 1cm (6.3.1) .$$
1131: along  the solution. 
1132: 
1133: 
1134: Our proof of (6.3.1)  relies on the fact that on a simply connected complete surface
1135: of  non-negative curvature any compact geodesic arc is the
1136:  unique minimizer between
1137: its endpoints.   On  a complete non-simply connected surface such
1138: as the pair of pants   this implies
1139: that if we have a geodesic arc, then there is no shorter curve
1140: which share  endpoints with that arc and which is homotopic to it
1141: through endpoint-fixing homotopies.    
1142: 
1143: We argue by contradiction.  Suppose that some
1144: solution $\gamma \subset P$ realizes the collision sequence
1145: $s$ but satisfies 
1146: $\lim \inf r_{12} = \delta > 0$.
1147:   We will construct a comparison curve  $c$ which
1148: has the same endpoints as a (long) arc of $\gamma$,  is homotopic
1149: to this arc 
1150: through end-point fixing homotopies, 
1151: but which is shorter than $\gamma$.  The  arc will
1152: be one whose syzygy sequence is $1212 \ldots 12$ with
1153: $N$ repeats of $12$, and $N$ large.  
1154:  See figure 6 for the picture of this arc
1155: and the  shorter comparison curve $c$.  The existence
1156: of $c$ contradicts the minimality of $\gamma$ described in  the previous 
1157: paragraph.
1158:  
1159: 
1160: To  construct $c$ we will use 
1161: the cylindrical  coordinates
1162: $(\lambda, \chi)$ 
1163: of  (3.14a,b), (3.15) associated
1164: to the  $12$-collision end.  We have    
1165: $$\lambda (\rho, \chi) = \int_{\rho_0} ^{\rho} \sqrt{U(s,
1166: \chi)} ds$$
1167: where $(\rho, \chi)$ are spherical-polar coordinates
1168: centered at the collision. 
1169:  The $(\lambda, \chi)$ coordinates are valid 
1170: on the entire sphere minus the `$3$' equatorial arc
1171: and the collision points.
1172:  The coordinate $\lambda$ satifies 
1173:   $\lambda  \to \infty$  as collision
1174: is approached.  The Jacobi metric in these coordinates is 
1175: $$ds^2 = d \lambda^2 + f(\lambda, \chi)^2 d \chi^2
1176: \hskip 1cm (6.3.2)  $$
1177: where  
1178: $$f^2 = {1 \over 2}   \cos^2 (\rho) \{ 1   +
1179:  {{s_{12}} \over
1180: {s_{13}}} + {{s_{12}} \over
1181: {s_{23}}} \} \hskip 1cm  $$
1182: The curvature of any metric of the form (6.3.2) 
1183: is  
1184: $ -{1 \over f}{{\partial ^2 f} \over
1185: {\partial \ \lambda ^2}}$.  Since this
1186: curvature is negative (theorem 2) and since 
1187: $\lim_{\lambda \to \infty} f(\lambda, \chi) = 1/\sqrt{2}$
1188: (eq. 3.18) 
1189: we have that   for each fixed
1190: $\chi$ the function    
1191: $f(\chi, \lambda)$ decreases monotonically to its infimum $1 / \sqrt{2}$
1192: as $\lambda \to \infty$.  
1193: It follows that 
1194: $F(\lambda) = \min_{\chi} f(\lambda, \chi)$
1195: also   decreases monotonically to $1/\sqrt{2}$.
1196: 
1197: Any geodesic arc $\gamma$ on the pair of pants which
1198: realizes the syzygy sequence $12$  
1199: cannot cross either isosceles circle
1200: $r_{12} = r_{23}$ or
1201: $r_{12} = r_{13}$.  This follows from the minimality 
1202: property of the arc, discussed above, and   the reflection principle 
1203: (as in [ChM] and the proof of no stuttering in section 2, between
1204: theorems 3 and 4).
1205: Reflections about the isosceles circles are isometries of the Jacobi metric,
1206: so that any segment of $\gamma$  which crosses, then crosses back, can be reflected,
1207: so as to form a new arc with the same endpoints as $\gamma$, and the same homotopy
1208: type, contradicting uniqueness.   Thus, without loss of generality,
1209: we may assume that our long subarc of $\gamma$ lies entirely in
1210: the union of the regions $r_{12} < r_{13}$ and
1211: $r_{12} < r_{23}$.   In particular,
1212: the $(\lambda, \chi)$ coordinates are valid all along our
1213: arc.  
1214: 
1215: To construct the desired comparison curve $c$ we will  
1216: use the fact that $F$ is monotone
1217: decreasing in $\lambda$  and that the number $N$ of $12$ crossings of  a subarc of $\gamma$
1218: can be taken arbitrarily
1219: large.      Since $\inf r_{12} > 0$, by assumption, we have
1220: that $\Lambda = \sup \lambda < \infty$ along our curve.
1221: Let $\epsilon > 0$ small be given. Choose  points $\gamma(s_N), \gamma (t_N)$
1222: $s_N < t_N$, along the arc 
1223: for which $\lambda (t_n), \lambda (s_n)  >  \Lambda - \epsilon$ 
1224: and such that  the arc $\gamma[s_n, t_n]$ in between realizes
1225: the syzygy sequence $12 \ldots 12$ with $N$ copies of $12$.
1226: Our comparison curve $c$ will go ``straight in'' to collision
1227: until some point with $\lambda_* > \Lambda$, to be determined momentarily,
1228: winds around the $12$ collision point in the same sense as $\gamma$ for the same number of
1229: syzygies at this fixed value $\lambda = \lambda_*$, and then return headed ``straight out''
1230: from collision to the point $\gamma(t_n)$.  See figure 6.  ``Straight in'' and ``straight out''
1231: means that $\chi$ is fixed, and only $\lambda$ varies. During
1232: its `winding around'  journey, $\lambda = \lambda_*$ is
1233: fixed and $\chi$ varies, starting at  $\chi =
1234: \chi(\gamma(s_n))$, increasing so that $c$ 
1235: suffers the   syzygy sequence $1212 \ldots 12$ (N
1236: times), and then stopping   at $\chi = \chi(\gamma(t_n)$ in time for the ``straight out''
1237: return segment.  
1238: 
1239: 
1240: \vskip 0.1in
1241: 
1242: \epsfxsize=3in
1243: 
1244: \epsfbox{collisionfig.eps}
1245: \vskip 0.2in
1246: 
1247: {\bf Figure 6.} Orbit surgery to shorten length of
1248: a collision sequence path
1249: 
1250: FFXX LABEL FIG as per next to last par of this sec.
1251: 
1252: \vskip .1in
1253:  
1254: \vskip 0.8in 
1255: 
1256:  
1257: 
1258: By construction, the curve  $c$ shares endpoints with our arc of $\gamma$,
1259: and is homotopic to it.
1260: It remains to  show that $c$ is shorter  than our arc of $\gamma$.
1261: By the monotonicity and the limiting properties of $f$ we can 
1262: choose $\lambda_* > \Lambda$ so that 
1263: $$max_{\chi}f(\lambda_*, \chi) < F(\Lambda):  = min_{\chi} f(\Lambda, \chi).$$
1264: Choose $N$ so large that
1265: $${{2(\lambda_* - \Lambda) + \epsilon + \pi F(\Lambda) } \over N} < \pi (F(\Lambda) -
1266: max_{\chi}f(\lambda_*,
1267: \chi) ).$$  It follows that 
1268: $$ \pi N  max_{\chi} f(\lambda_*, \chi) + \pi max_{\chi} f(\lambda_*, \chi)   + 2(\lambda_*
1269: - \Lambda) + 2 \epsilon <  N\pi (F(\Lambda)$$
1270: Now consider the cartoon in figure XX or the
1271: construction of $c$.   The `in and out' arcs of
1272: $c$ have length less than $2(\lambda_* - \Lambda) + 2 \epsilon$.
1273: The `around arc' between the $N$ syzyygies has length  less than $\pi N  max_{\chi}
1274: f(\lambda_*,
1275: \chi)$.  (Refer again to figure 6.) The additional $\pi max_{\chi} f(\lambda_*, \chi)$
1276: accounts for the fact that $\chi(s_n)$ and $\chi(t_n)$ need
1277: not be equal.      Thus the  left hand
1278: side of the inequality is greater than the length of $c$.
1279: A similar but simpler analysis shows that the right hand side is smaller than
1280: the length of our arc of $\gamma$.   Thus $c$ is shorter than the arc of $\gamma$,
1281: completing  the proof of first assertion of  theorem 3. 
1282: 
1283: The same analysis shows that the infinite sequence $\ldots 1212 \ldots$
1284: of all $12$'s is never realized.  For such a realization must
1285: be a local minimizer, and the above orbit surgery shows we can always
1286: decrease the length of a path by making it closer to collision.
1287: 
1288: 
1289: 
1290: QED
1291: 
1292: \vskip .3cm 
1293: 
1294: {\bf 6.4. Proof of theorem 4.} 
1295: 
1296: To establish the   uniqueness of the realizing solutions, we 
1297: work on the
1298: universal cover $D$ of the pair of pants $P$. Topologically, $D$ is  
1299:  the Poincare disc and the  
1300: fundamental group  $\Gamma = \pi_1 (P)$ (the free group on two letters) 
1301: acts on $D$ as a Fuchsian group.    See figure 3,
1302:   and also the
1303: book Indra's Pearls [Mumf].  
1304: In  figure 3b   we have drawn in a fundamental domain $P_0$
1305: for $P$ and some of its images   under $\Gamma$. The
1306: boundary of $P_0$ consists of four circular arcs which are lines 
1307: lines relative to the Poincare metric on $D$.  These bounding 
1308: arcs are labelled $1_+, 1_-$ and $2_+, 2_-$. To form $P$
1309: out of $P_0$ glue 
1310: arcs $1_+$ and $1_-$ to form syzygy arc $1$, and glue
1311: arcs $2_+$ and $2_-$ to form syzygy arc $2$.
1312: Syzygy arc $3$ is internal
1313: to the fundamental domain.  In the figure we have dropped the
1314: $+, -$ subscripts on the arcs. 
1315: 
1316: \eject
1317: 
1318: \vskip 0.1in
1319: 
1320: \epsfxsize=3.00in
1321: 
1322: \epsfbox{funddomain.eps}
1323: \vskip 0.1in
1324: 
1325: {\bf Figure 3a.} The fundamental domain
1326: \vskip .2in
1327: \epsfxsize=3.00in
1328: 
1329: \epsfbox{tiling.eps}
1330: 
1331: {\bf Figure 3b.} Some of the tiles.
1332: 
1333: \vskip 0.4in
1334: 
1335: Write $$\pi: D \to P$$
1336: for the covering map, so that the fibers of $\pi$
1337: are copies of $\Gamma$.   We will use the hyperbolic, or   constant 
1338: negative curvature metric on $D$ 
1339: in order to understand the fundamental group $\Gamma$ and its action on $D$.
1340: (With   respect to this   metric  $\pi$ is not a local isometry.)  
1341: $\Gamma$ acts on $D$ with respect to hyperbolic
1342: isometries, i.e. M\"obius transformations. 
1343: $\Gamma$ is freely generated by two elements $a$ and $b$,
1344: one of which, say $a$,  interchanges $1_+$ and $1_-$, and 
1345: the other of which, $b$, 
1346: interchanges $2_+$ and $2_-$. These elements act on $D$
1347: as Mobius transformations.  (Viewed
1348: as acting on the Riemann sphere, they  interchange exteriors with interiors of 
1349: their respective circles. )
1350:  In order to form $P$ out of $P_0$ glue arc $1^+$ to $1_-$ 
1351: by $a$  
1352: and glue  $2_+$ to $2_-$ by $b$.  
1353: The projection under $\pi$ of these boundary arcs form the arcs $1,2$ of the equator
1354: of $P$.
1355: The third arc $3$ is internal to $P_0$, and separates it into
1356: two halves, the northern (+) hemisphere, and southern (-) hemisphere.
1357:       
1358:  The images of $\gamma P_0$ of the fundamental domain 
1359: $P_0$ under elements of $\gamma \in \Gamma$ tile all of $D$. 
1360: Each of the four  boundary arcs of such a tile $\gamma P$  
1361:  is the image under $\gamma$ of a unique
1362: boundary arc of $P_0$ and we continue to  label the tile's boundary arcs  
1363: by the corresponding $P_0$ lables  $1_+, 1_-, 2_+$, or  $2_-$. 
1364: Two  tiles intersect, if at all,   along 
1365: a common boundary arc.  This common
1366: arc must be   a   $j_-$ arc of one tile
1367: and a $j_+$ of the other,   $j =1, 2$. 
1368:    
1369: Suppose  now  that 
1370:  two geodesics   realize the same symbol sequence $s$.
1371: Denote by $\gamma,   c$   the lifts of these geodesics to 
1372:   the universal cover $D$.  After  translating these  curves  by   elements of $\Gamma$ 
1373: we may suppose that  both  begin in the  reference fundamental domain
1374: $P_0$.   
1375: I claim that the syzygy sequence $s$ uniquely specificies
1376: a sequence of contiguous tiles  $\ldots P_{-2}P_{-1} P_0 P_1 P_2 \ldots $
1377: through which   $c$ and $\gamma$ must pass.  
1378: To see this fact, we first note that  each sequence of
1379: three contiguous tiles $P_{i-1} P_i P_{i+1}$
1380: represents either two or three letters of a signed syzygy sequence. See figure 4. 
1381: The sequence is obtained by drawing a curve which crosses from $P_{i-1}$
1382: through $P_i$ and into $P_{i+1}$ in the `most direct'' way.
1383: The curve must enter into $P_i$ across one
1384: of its bounding arcs $1_+, 1_-, 2_+,
1385: 2_-$. The choice of $P_{i-1}$ uniquely specifies
1386: which arc.  It must leave   
1387: across another such arc and the choice of  $P_{i+1}$
1388: uniquely specifies this exit arc.  Along the way it must either
1389: cross $3$ or not.   If no   internal syzygy with arc 3 occurs then the  sequence has two letters
1390: and no `$3$', and otherwise the sequence
1391: does contain the  letter $3$ as the middle letter.    
1392:     Consequently,
1393: both $\gamma, c$ pass through an identical
1394: list of tiling domains, as claimed.  
1395: 
1396: Each tiling domain 
1397: $P_j$ has within
1398: it an inverse image $R_j = \pi^{-1} (R) \cap P_j$  of our  compact domain $R$.  
1399: By lemma 6.1 both $c$ and $\gamma$
1400: must have the property that for infinitely many $j$ we
1401: have that both  $\gamma$ and $c$ 
1402: lie in $R_j$. 
1403: 
1404: \vskip 0.2in
1405: 
1406: \epsfxsize=3.00in
1407: 
1408: \epsfbox{crossings.eps}
1409: \vskip 0.4in
1410: 
1411: {\bf Figure 4.} Syzygies and Tile Crossings.
1412: \vskip .2in
1413:  
1414: 
1415:  Now  lift the Jacobi-Maupertuis metric
1416: from $P$ up to $D$, using the
1417: projection $\pi: D \to P$, thus arriving at a complete
1418: non-negatively curved $\Gamma$-invariant metric
1419: on $D$ for which our two curves
1420: $\gamma$ and $c$ are geodesics.
1421: The curvature of this metric is   zero only
1422: at the discrete set of points $\pi^{-1}(L_{\pm})$,
1423: where $L_{\pm}$ are the Lagrange points of $P$.
1424: And the map $\pi$ is  a local isometry for this metric.   
1425: Write 
1426: $$h(t) = dist(\gamma, c(t)) \hskip 1cm (6.4.1)$$
1427: for the distance between the variable point $c(t)$ on
1428: the curve $c$ 
1429: and the entire geodesic $\gamma$.  
1430: See figure 8.  Now
1431: each $R_j$ has finite diameter $\delta$ because
1432: $R$ is compact, and the   two curves pass
1433: through the $R_j$'s infinitely often, indeed every time 
1434: the letters  $123$ occurs contiguously in $s$. It follows that 
1435:   
1436: $$\lim \inf_{t \to +\infty} h(t) \le \delta 
1437: \hbox{ and } 
1438: \lim \inf_{t \to -\infty} h(t) \le \delta. \hskip 1cm (6.4.2)$$ 
1439: 
1440: We will now show that inequality (6.4.2) is impossible unless  
1441:  the two geodesics are in fact the same, in which case $h(t) =0$ everywhere.
1442: We use the formula
1443: $$d^2 h /dt^2 = -sin(A(t)) \int_{d_t} K ds \hskip 1cm (6.4.3)$$
1444: proved in the following paragraph.
1445: In this formula, 
1446: $d_t $ is  the geodesic realizing the distance
1447: $h(t)$.  It  has one endpoint at the point $c(t)$ on $c$
1448: and the other endpoint on $\gamma$ which it intersects
1449: perpindicularly.  See figure 5. The angle $A(t)$ is the
1450: angle of intersection
1451: between the geodesics $c$ and
1452: $d_t$ at  $c(t)$.  See figure 5. 
1453: This angle satisfies   $0 < A(t) < \pi$, so that $\sin(A(t) > 0$  
1454: The negativity of   $K$ 
1455: (except at a discrete  point set) implies that 
1456: $h(t)$ is strictly convex: $d^2 h /dt^2 > 0$.  
1457:    But any strictly convex function defined
1458: on the real line  tends to infinity  in
1459: one direction or the other. This  contradicts   (6.4.2).
1460: Our two geodesics must be the same.  
1461: 
1462: 
1463: \vskip 0.2in
1464: 
1465: \epsfxsize=3.00in
1466: 
1467: \epsfbox{variationarc.eps}
1468: \vskip 0.4in
1469: 
1470: {\bf Figure 5.} Variation of Distance
1471: \vskip .2in
1472:  
1473: {\bf Derivation of (6.4.3). }
1474: 
1475: The first variation of arclength implies
1476:  $$dh/dt = \cos(A(t)) \hskip 1cm (6.4.4) .$$
1477:  Let
1478: $M(t) \subset D$ denote the quadrilateral whose  edges 
1479: consist of the geodesic arcs $d_0, d_t$
1480: together with  the arcs of $\gamma, c$ which connect
1481: $d_0$ to  $d_t$.  According to  the Gauss-Bonnet theorem,
1482: for any such geodesic quadrilateral $Q$ we have
1483: $$ 2 \pi - (\Sigma \hbox{ interior angles }) = - \int \int_Q  K dA$$
1484: In  the case of $M(t)$ the interior angles are $\pi/2, \pi/2, \pi - A(0)$ and $A(t)$.  
1485: See figure 5 again.  
1486: Thus  
1487: $$A(0)  - A(t)= - \int \int_{M(t)} K(t) dA \hskip 1cm (6.4.5) $$
1488: Differentiating (6.4.4) with respect to $t$ yields  
1489: $$dA/dt = \int_{d_t} K ds. \hskip 1cm (6.4.6) $$
1490: Now differentiate (6.4.4) with respect to $t$, using (6.4.6)  to obtain (6.4.3).
1491:  
1492: \vskip 1cm
1493: 
1494: \magnification=\magstep1 \hoffset .65pt
1495: \def\R{I\!\! R}
1496: \def\Z{\it Z \!\!\! Z}
1497: \def\H {I \!\! H}
1498: \def\C{C \!\! \! \!  I\, }
1499: \def\I{I\!\!I}
1500: \def\J{J\!\!\!J}
1501: \def\RP{I\!\! R I\!\! P}
1502: \def\Ri{Riemannian }
1503: 
1504: 
1505: 
1506: {\bf Proof of theorem 5.}  Let $c$ be a bound orbit and
1507: $s = \{ s_j \}_{j = - \infty} ^{\infty}$ its syzygy sequence.
1508: Approximate $s$ by a family $w^n$, $n = 1,2, 3, \ldots$  
1509: of forward collision sequences by replacing the tails $s_j, j > n$ of $s$
1510: by that of the $12$ collision sequence. Collapse two letters if neccessary when
1511: stutters appear at the `join' $j = n$ of   the replacement. 
1512: The backward tails of
1513: the $w^n$ contain
1514: all three letters because $s$ is collision-free, so  we can apply theorem 2 to realize the $w^n$ by
1515: solutions $\gamma_n$ (not necessarily unique).  By lemma 6.1
1516: the $\gamma_n$ all pass through $R$, so, shifting time if necessary,
1517: we have that the tangent vectors   $v_n = (\gamma_n (0), \dot \gamma_n (0))$ 
1518: are unit vectors with position $\gamma_n (0)$ in   $R$.  
1519: We now argue as in the proof of theorem 3.  By compactness
1520: of the  set of unit
1521: tangent vectors over $R$, we can form a convergent
1522: subsequence of the $v_n$, which we  relabel as  $v_n$, so that $v_n \to v$.
1523: Let $\gamma$ be the curve with initial condition $v$.  
1524: The curves $\gamma_n$  converge, over compact
1525: sets,  to  $\gamma$, so that the syzygy sequences $w^n$ must converge to
1526: the syzygy sequence of $\gamma$.  Thus $\gamma$ and $c$ share the same
1527: syzygy sequence.  By   theorem 4,  $\gamma = c$.  Consequently the
1528: unboounded curves $\gamma_n$ converge to our initial bound curve $c$.
1529: 
1530: QED
1531:  
1532: 
1533: \vskip .3cm 
1534: 
1535: 
1536: {\bf 8. Curvature for  unequal masses.} 
1537: 
1538: In this section is to prove that Theorem 1 is
1539: a special case:  for
1540: most mass distributions the curvature takes on both signs.  
1541: Take the masses to be general positive numbers
1542: $m_i$.  Writing  $$p_i = m_j m_k, s_i = r_{jk}^2$$  for
1543: $ijk$ a   permutation of $123$ we have
1544: $$U = \Sigma p_i /s_i, \hskip .3cm  I = {1 \over M} \Sigma p_i s_i,
1545: \hbox{  
1546: where } M = \Sigma m_i.$$
1547: The Jacobi metric is obtained by multiplying the
1548: shape metric $ds^2_m$ on the shape sphere $I =1$  
1549: by $U$ and restricting it   to the sphere $I =1$.
1550: (The subscript `m' indicates dependence on the masses.)
1551: It is conceptually and computationally more straightforward to identify
1552: the shape sphere  as the space of rays in shape space
1553: and to make $U$ a function on the space of rays by
1554: making it  homogeneous of degree  0  by multiplying it by $I$. Thus, setting
1555: $$\tilde U = I U , $$  our Jacobi metric is
1556: $$ds^2_J =  \tilde U ds^2_m.$$
1557: 
1558: 
1559: In order to compute, we will use the  coordinates  $\phi, \theta, R$  of   [Mont2] 
1560: for  shape space. 
1561: (See, in particular, the notation and computations of 
1562: section 7 there.)
1563: Write $I_1$ for the moment of inertial when
1564: all masses are equal to one:
1565: $$I_1 = \Sigma r_{ij}^2/ 3. \hskip 1cm (7.1) $$
1566: The   $\phi, \theta$ are spherical coordinates
1567: for the   $I_1$ shape sphere, while the radial coordinate  
1568: $R = \sqrt{I}$ for the mass-dependent   $I$ of (3.6). 
1569:  With respect to these coordinates we have that  
1570: $$s_k  =   I_1 (1 - \gamma_k (\theta) \cos(\phi)  \hskip 1cm (7.2).$$
1571: as before
1572: Note that  the coordinates $\theta, \phi$ and the functions $s_k$
1573: do not vary as the masses are changes.  
1574:  The metric $ds^2_m$ when expressed in our coordinates is mass-dependent and
1575: is given by 
1576: $$ds^2 _m = \lambda^2 ds^2_1 \hskip 1cm (7.3a) $$
1577: where
1578: $$ds^2_1 ={1 \over 4}(d \phi ^2 + \cos(\phi)^2 d \theta^2) \hskip 1cm (7.3b)$$
1579: (see [MontI], eq (5.6)  and Prop. 2) is the shape sphere metric when all the
1580: masses are $1$,  and where the conformal factor
1581: $\lambda$ is given  by 
1582: $$\lambda = d(m)I_1/I ; \hskip .2cm d(m) = \sqrt{ 3 m_1 m_2 m_3/ M}  \hskip 1cm (7.3c)$$
1583: (See eq (5.7) of [MontI]).
1584: The metric we are to work with is thus 
1585: $$ds^2 _J  = \tilde U\lambda^2 ds^2_1 \hskip 1cm (7.4).$$
1586: Its curvature is given by lemma 4.1 :
1587: $$\bar K = {{ 1} \over { \tilde U  \lambda^2}} \{ 4 - {1 \over 2} \Delta \log (\tilde U \lambda^2) \} \hskip 1cm (7.5).$$
1588: where the
1589: Laplacian $\Delta$ is with respect to the metric
1590: $ds^2_1$.   The conformal factor is 
1591: $$\eqalign{
1592: \tilde U \lambda^2 & = d(m)^2 I U I_1^2/I^2 \cr
1593: & = d(m) I_1 U (d(m) I_1/I) \cr
1594: & = d(m) \hat U \lambda
1595: }$$
1596: where 
1597: $$\hat U = I_1 U.$$
1598: Then $$\Delta \log \tilde U \lambda^2 = \Delta \log \hat U + \Delta \log \lambda.$$
1599: Set 
1600: $$\hat s_i = s_i/I_1 = (1 - \gamma_k (\theta) \cos(\phi)  $$
1601: so that
1602: $$\hat U = \Sigma p_i/ \hat s_i $$
1603: Since $\hat s_i$ is equal to the variable
1604: $s_i$ as given by eq. (5.1.1b), and since the Laplacian
1605: is the Laplacian of that section, we can continue
1606: to   compute  as per section 5.
1607: We have
1608: $$ \hat U \lambda \bar K =  4 - {1 \over 2} \Delta \log (\hat U) - {1 \over 2} \Delta \log (\lambda) .$$
1609: 
1610: 
1611: To compute the last term $- {1 \over 2} \Delta \log (\lambda)$ of (7.5)
1612: we use the fact that  the metrics   $ds^2_m$ are 
1613: $ds^2_1$  through $\lambda^2$, and that 
1614: both have  curvature $4$. From lemma 4.1 it follows that
1615: $$4 = {1 \over \lambda^2} (4 -  {1 \over 2} \Delta \log \lambda^2 \}$$
1616: or
1617: $$4\lambda^2 - 4 = - \Delta \log \lambda.$$
1618: Equation (7.5) can then be rewritten 
1619: $$ -\hat U^3 \lambda \bar K =  -\hat U^2(4 - {1 \over 2} \Delta \log (\hat U))
1620:  + \hat U^2 (2 - 2 \lambda^2)  \hskip 1cm
1621: (7.6).$$
1622: 
1623: {\bf Computation of $\Delta \hat U$.}
1624:  To ease  notation, we drop the hats for this subsection,
1625: so that   $U$ means the function $\hat U$ and
1626: the   $s_k$ means the function 
1627: $\hat s_k$. 
1628: The   structural form of all the formulae and calculations of section 5
1629: remains intact provided we  insert   the weightings
1630: $p_i$ in the correct places.  
1631: 
1632: We proceed to the computation of $\Delta \log (\hat U))$.  Equations (4.2)
1633: and (5.2.1) continue to hold with $\hat U$ in place of $U$, and equations
1634: (5.2.2a,b) hold with $\hat s_i$ in place of $s_i$.  The analogues of
1635: (5.2.3), (5.2.4) are
1636: $$
1637: {1 \over c}
1638: \dphi(c \dphi U)   = ((s^2 - c^2)/c)\Sigma p_i \gamma_i /s_i ^2 
1639:  + 2 s^2 \Sigma p_i\gamma_i ^2/ s_i^3 \hskip 1cm
1640: \hskip 1cm (7.7a)$$
1641: And
1642: $$ {1 \over c^2} \dtheta \dtheta U 
1643:   = {1 \over c} \Sigma p_i \gamma_i {\prime \prime}/s_i ^2
1644: + 2 \Sigma p_i (\gamma_i ^{\prime})^2/s_i ^3
1645: \hskip 1cm (7.7b)$$
1646: Now note that  algebraic steps going from (5.2.5) to (5.2.7) 
1647: apply term by term, so can be carried over verbatim except that
1648: the $i$th term must be multiplied by $p_i$. Thus (reverting to hats)
1649: $$\Delta \hat U = 8 \hat U_4 \hskip 1cm (7.8a)$$
1650: where $$\hat U_4 = \Sigma p_i/\hat s_i ^2 \hskip 1cm (7.8b).$$
1651: 
1652: The first line of  (5.3.2) becomes 
1653: $$\|\nabla U \|^2  =  4(\Sigma s p_i\gamma_i/s_i^2)^2 +  4 (\Sigma p_i\gamma_i ^{\prime}/s_i ^2 )^2$$
1654: The algebra which follows is essentially the same, leading to
1655: $$\| \nabla U \|^2 = 4 S \hskip 1cm (7.9a)$$
1656: where
1657: $$S = 2 \Sigma p_i ^2/ s_i ^3 - \Sigma p_i ^2/s_i ^3 
1658: -{3/2} \Sigma^{\prime} p_i p_j /s_i ^2 s_j ^2 +  2 \Sigma^{\prime} p_i p_j/s_i  s_j ^2
1659: - \Sigma^{\prime} p_i p_j /s_i  s_j \hskip 1cm (7.9b).$$
1660: Combining (7.8) and (7.9)  according to (4.2) (see also the steps (4.7)-(4.9)) yields the formula: 
1661: $$- U^2 (4 - {1 \over 2} \Delta \log U)  = 3 \Sigma^{\prime} p_i p_j /s_i ^2 s_j ^2 - 2(\Sigma p_i /s_i)^2 
1662: \hskip 1cm (7.10).$$
1663: 
1664: \vskip 1cm 
1665: From earlier, we have   
1666: $$ U^2 (2 - 2 \lambda^2) = 
1667:  2 (\Sigma p_i /s_i)^2 - 2(\Sigma p_i /s_i)^2 d(m)^2 ({1 \over 3} \Sigma
1668: s_i)^2/(\Sigma p_i s_i /M)^2, \hskip 1cm (7.11).$$
1669: (We continue to use $s_i$ in place of $\hat s_i$.) Upon adding (7.10) and (7.11) there
1670: is a   cancellation yielding:  
1671: $$ - U^3 \lambda^2 \bar K =  3 \Sigma^{\prime} p_i p_j /s_i ^2 s_j ^2   - 2(\Sigma p_i /s_i)^2 d(m)^2 ({1 \over 3} \Sigma
1672: s_i)^2/(\Sigma p_i s_i /M)^2.$$
1673: A computation shows that
1674: $$d(m)^2 M^2 = {3 \over 2}\Sigma^{\prime} p_i p_j$$
1675: Recall that the $s_i$ (the previous $\hat s_i$'s) satisfy
1676: ${1 \over 3} \Sigma s_i =1$.  We finally obtain:
1677: $$ - U^3 \lambda^2 \bar K =  3 \{ \Sigma^{\prime} p_i p_j /s_i ^2 s_j ^2  
1678:  - \Sigma^{\prime} p_i p_j {{ (\Sigma p_i /s_i)^2 } \over {(\Sigma p_i s_i )^2}} \} 
1679: \hskip 1cm (7.12).$$
1680: Consequently 
1681: $$\kappa = \sqrt{\Sigma^{\prime} {{p_i p_j} \over{ s_i ^2 s_j ^2}}}  
1682: {{ (\Sigma p_i s_i)} \over {(\Sigma p_i /s_i )}} - \sqrt{  \Sigma^{\prime} p_i p_j }
1683: \hskip 1cm (7.13)$$
1684: governs the sign of the curvature,
1685: with the curvature   $\bar K$   negative if $\kappa$
1686: is positive, positive if $\kappa$ is negative, and zero
1687: if $\kappa$ is zero.  We note that 
1688: $\kappa$, and hence the curvature is zero 
1689: at the Lagrange point $s_1 = s_2 =s_3 =1$,
1690: and that this is true for all choices of the masses $p_i$.
1691: 
1692: \proclaim Theorem 6.  
1693: For a Zariski-dense set of  mass distributions,
1694: the sign of the curvature changes in a neighborhood of the
1695: Lagrange point.
1696: 
1697: Proof. It suffices
1698: to show that for a Zariski-dense set of  mass distributions
1699: the differential  $d \kappa  \ne 0$ at the Lagrange point.
1700: A differential form $\Sigma a_i ds_i$
1701: represents zero on the shape sphere if and only if it is proportional to
1702: $\Sigma ds_i$, the latter being the differential of the constraint
1703: $\Sigma s_i = 3$ satisfied by the $s_i$ (which are the old $\hat s_i$'s). 
1704: A computation shows
1705: that  at  the Lagrange point $s_i =1$ we have  
1706: $$\beta d \kappa = p_1 (p_2 ^2 + p_3 ^2) ds_1 + p_2 (p_1 ^2 + p_3 ^2) ds_2 + p_3 (p_1 ^2 + p_2 ^2) ds_3
1707: \hbox{ mod }  \Sigma ds_i$$
1708: where $\beta$ is a nonzero constant.
1709: We thus want to know whether or not
1710: the equality
1711: $$(p_1 (p_2 ^2 + p_3 ^2), p_2 (p_1 ^2 + p_3 ^2) , p_3 (p_1 ^2 + p_2 ^2)) 
1712: = (\lambda, \lambda, \lambda) \hskip 1cm (**) $$
1713: can be satisfied for some $\lambda$.   
1714: The right hand side of equation (**),
1715: being homogeneous of degree $3$, 
1716:   defines a polynomial map $\RP^2 \to \RP^2$
1717: and we want to know if it is equal to the
1718: constant map   $[1,1,1]$.  Because the map is  polynomial,
1719: if we can exhibit a single point where the  inequality fails
1720: then it must fail on a Zariski-dense set.    Plugging in
1721: $p_1 = p_2 = 1, p_3 = a$ yields
1722: $(p_1 (p_2 ^2 + p_3 ^2), p_2 (p_1 ^2 + p_3 ^2) , p_3 (p_1 ^2 + p_2 ^2))
1723: = (1 + a^2, 1 + a^2, 2a)$
1724: which is not proportional to $(1,1,1)$ unless $a = 1$.
1725: 
1726: QED 
1727: 
1728: \vskip .3cm 
1729: {\bf 
1730: Appendix A}
1731: 
1732: We prove
1733: \proclaim Theorem A.  The set of initial conditions
1734: within $H = 0,  I = 1 , c = 0$
1735: whose solutions tend to a binary collision
1736: of type $ij$ has nonempty interior. This fact
1737: holds for all positive mass distributions. 
1738: 
1739: {\bf Proof of theorem A.}We use Newtonian time and Jacobi
1740: coordinates For notational simplicity,  take $ij = 12$.
1741: The   Jacobi coordinates are
1742: $\zeta_1 = x_1 - x_2$, $\zeta_2 = x_3 - (m_1 x_1 +m_2 x_2)/(m_1 + m_2)$.
1743: The distance to binary collision is
1744: $$r = | \zeta_1| \hskip 1cm (1A).$$
1745: We will exhibit a nonempty open set of
1746: initial conditions at time $t =0$
1747:  for which $r(t) = 0$
1748: for some time $t < O(r(0))$. 
1749: 
1750:  
1751: The Hamiltonian is 
1752: $$H = {1 \over 2} ( \mu_1 | \dot \zeta_1|^2 + 
1753: \mu_2 | \dot \zeta_2 |^2)  -{{m_1 m_2} \over {r^2}} 
1754: - W(\zeta_1, \zeta_2) \hskip 1cm (2A)$$
1755: where $\mu_1 = m_1 m_2 /(m_1 + m_2)$,
1756: $\mu_2 = m_3(m_1 + m_2)/M$, 
1757: $W = m_1 m_3 /s_2 + m_2 m_3/s_1 $
1758: and the squared distances $s_2$, $s_3$  can be expressed 
1759: $|\zeta_2 + a_i
1760: \zeta_1 |^2$ in Jacobi coordinates,  with
1761:  mass-dependent
1762: nonzero constants
1763: $a_i$.   
1764: The interaction term  $W$  satisfies the estimates  
1765: $$|W| \le C_1 + C_2 \epsilon ,\hskip .3cm \hbox{ for } r < \epsilon \hskip 1cm 
1766: (3A), $$ 
1767: $$|{{\partial W} \over {\partial
1768: \zeta_1}}|
1769: \le C_1 + C_2 \epsilon 
1770: \hskip .1cm ,  \hskip .1cm  |{{\partial W} \over {\partial
1771: \zeta_2}}| \le C_1 + C_2 \epsilon , \hskip .3cm 
1772: \hbox{ for  } r  \le \epsilon \hskip 1cm (4A)$$
1773: ($\epsilon$ sufficiently small), where $C_1, C_2$
1774: are constants depending only on the masses.   
1775: 
1776: The equations of motion are 
1777: $$\ddot \zeta_1 = - (m_1 + m_2){{\zeta_1} \over {r^4}}  + {1
1778: \over
1779: \mu_1}{{\partial W} \over {\partial \zeta_1}} \hskip 1cm (5A)$$
1780: and
1781: $$\ddot \zeta_2 =   {1 \over
1782: \mu_2}{{\partial W} \over {\partial \zeta_2}} \hskip 1cm (6A)$$
1783: Write 
1784: $$J_1 = \zeta_1 \wedge \dot \zeta_1 \hskip 1cm (7A)$$
1785: for the angular momentum (up to a factor of $\mu_1$)
1786: of the 12 system.  We compute 
1787: $$\eqalign{\dot J_1 & = \zeta_1 \wedge \ddot \zeta_1 \cr
1788: &= \zeta_1 \wedge{1 \over
1789: \mu_1}{{\partial W} \over {\partial \zeta_1}}}
1790: $$
1791: so that
1792: $$|\dot J_1| \le Cr \hskip 1cm (8A).$$
1793: 
1794: Because $| \dot \zeta_1|^2 = \dot r^2 + J_1^2/ r^2$ we
1795: have that 
1796: $$r^2 H = {1 \over 2} (\mu_1 r^2 \dot r^2 + \mu_1 J_1^2 )
1797: - m_1 m_2  + r^2 ( {1 \over 2} \mu_2 | \dot \zeta_2 |^2
1798: - W ) \hskip 1cm (9A) $$
1799: 
1800: 
1801: Now let $\zeta(t) = (\zeta_1 (t), \zeta_2 (t))$
1802: be a solution satisfying the initial conditions
1803: $r(0) < \epsilon,  H = 0, I =1, J = 0$.
1804: From (6A) and  (4A)
1805: we have that 
1806: $$| \dot \zeta_2 (t)|^2 \le | \dot \zeta_2 (0)|^2 + Ct \hskip 1cm (10A),$$
1807: for $t = O(1)$, provided $r(0) < \epsilon$.
1808: Here $C$ depends only on the masses and $\epsilon$.  
1809:  Letting $r \to 0$, we see from (9A) that if our solution
1810: is to have a collision then
1811: we must have
1812: $$\lim r^2 \dot r^2 + \lim J_1 ^2 - 2 (m_1 + m_2) = 0,  \hskip 1cm (11A)$$
1813:  where we have used $m_1 m_2 /\mu_1 = m_1 + m_2$.
1814: But $r^2 \dot r^2 \ge 0$, so we must have
1815: $$ 2 (m_1 + m_2) -\lim J_1 ^2  \ge 0 \hskip 1cm (12A)$$
1816:  
1817: We  argue   in the reverse. Suppose that   $ 2 (m_1 + m_2) - J_1 (0) ^2 $ is
1818: sufficiently positive at the initial time $t=0$, and that $\dot r (0) < 0$ 
1819: then (11A) forces $r^2 \dot r^2$ to be positive over
1820: a finite  time interval. We will show that, upon  integration, 
1821: this  will  force
1822: $r(t) = 0$ in some finite time  $t = O(\sqrt{r(0)})$.
1823: Note from the bounds (10A)   and the fact that $H= 0$
1824: we have
1825: $$|\mu_1  r^2 \dot r^2 + \mu_1 J_1 (0)^2 )
1826: - 2 m_1 m_2  |  \le  K r(0) \hskip 1cm (13A) $$
1827: for $0 \le t \le 1$ and for as long as $\dot r(t) < 0$. 
1828: Here the constant $K$ depends only on the masses
1829: and $\dot \zeta_2 (0)$.   Dividing by $\mu_1$ and using
1830: $m_1 m_2 /\mu_1 = m_1 + m_2$ we arrive at 
1831: $$|  r^2 \dot r^2 + J_1 (0)^2 )
1832: - 2(m_1  + m_1 ) |  \le  K^* r(0)  \hskip 1cm (14A) $$ 
1833: where $K^* = K/\mu_1$. 
1834: 
1835: We now impose the 
1836: open condition
1837: $$2(m_1 +  m_2)  - J_1 (0)^2 - K^* r(0) > \delta^2 \hskip 1cm 
1838: (15A)
1839: $$
1840: on our initial conditions.  This will be the open condition
1841: of theorem A.  
1842: The positive constant    $\delta$ will
1843: be constrained further
1844: below.     
1845: It follows from (15A) and (14A) that 
1846: $$\delta^2  <  2(m_1 +  m_2) - J_1 (0)^2) - K^* r(0)  \le  r^2 \dot
1847: r^2  \hskip 1cm 
1848: (16A)$$
1849: (16A) together with $\dot r(0) < 0$  forces   $\dot r < 0$ throughout the time interval in
1850: question.   Thus $ - r \dot r > 0 $, and so we can take
1851: square roots of  inequality  (16A) to obtain
1852: $$\delta \le  - r \dot r  \hskip 1cm 
1853: (17A) .$$  Taking negatives and integrating we find
1854: that
1855: $ -  \delta t \ge {1 \over 2} r(t)^2 - {1 \over 2}
1856: r(0)^2$ or 
1857: $$r(0)^2 - 2  \delta t  \ge r(t)^2 \hskip 1cm 
1858: (18A).$$
1859: This forces $r(t) = 0$ for some time
1860: $t$ with $t \le r(0)^2 / 2 \delta $. 
1861: In order that the collision time $t$ is $o(1)$ 
1862: it is sufficient to take $\delta = O(r(0))$.   
1863: 
1864: We have proved that a 12 collision occurs
1865: within a time $t = r(0)^2/ 2 \delta $ 
1866: for all initial conditions satisfying (15A), 
1867: $\dot r(0) < 0$,  and $r(0) < \epsilon$, 
1868: where $\epsilon$ is small enough so that the inequalities  (3A, 4A) are in force.
1869: This set of initial conditions is clearly open.
1870: It remains to show that this set is nonempty.
1871: Consider  the {\it collinear}  solution
1872: having  $H =
1873: 0 = J$ and $I =1$.  (There are precisely two
1874: such solutions, up to   time translation
1875: and rotation, one for each arc of the equator which
1876: ends in the 12 collision.)   These solutions  satisfy $J_1 = 0 $.
1877:    In this case
1878: (15A) reads
1879: $m_1 + m_2  > K r(0) + \delta$.
1880: and so will hold for $r (0)$ small provided
1881: only that $\delta < m_1 + m_2$.
1882: Since the solution tend to collision
1883: it follows that (15A) is eventually in force
1884: along the collinear solution, and hence that
1885: our set of of initial conditions is nonempty.  
1886: 
1887: QED
1888: 
1889: 
1890: \vskip 1cm
1891: 
1892: 
1893: {\bf Acknowledgements.}  
1894: I   dedicate this paper to the memory of my father.
1895: I    acknowledge useful correspondences
1896: with  Toshiaki Fujiwara, Alain Chenciner, Alain Albouy,
1897: and   conversations with  Anatole Katok, Rafe Mazzeo, and with 
1898:   Jeff Xia for pointing out that theorem 1
1899: combined with an earlier
1900: version of theorems 2, 3 and 4 ought to imply theorem 5.  
1901: 
1902: \medskip
1903: 
1904: {\bf References}
1905: 
1906: [AbMar] R. Abraham and J. Marsden, {\bf Foundations of Mechanics}, 
1907: Benjamin-Cummings, [1978]. 
1908: 
1909: [AlbCh]  A. Albouy and A. Chenciner, {\it Le probl\'eme des $n$ corps et les
1910: distances mutuelles}, Invent Math. 131 (1998), no. 1, 151--184.
1911:  
1912: [Arn]  V.I. Arnol'd, {\bf Mathematical Methods of Classical Mechanics},
1913: Springer-Verlag, [1989]. 
1914: 
1915: [Ban] T. Banachiewitz,
1916: {\it Sur un cas particulier du probleme
1917: des trois corps}, CRAS, Paris, 142, (1906), pp 510-512.
1918: 
1919: [CGMS] A. Chenciner, J. Gerver, R. Montgomery R. and C. Sim\'o{\it Simple
1920: choreographies of $N$ bodies: a preliminary study}
1921:  in {\bf Geometry, Mechanics
1922: and Dynamics}, 287--308, Springer,New York, 2002.
1923: 
1924: [ChMont] Chenciner A. and Montgomery R. {\it A remarkable periodic solution of the three-body
1925: problem in the case of equal masses, Annals of Math., 152, pp. 881-901 (2000)}
1926:  
1927: 
1928: [FerrTerr] Davide Ferrario and Susanna Terracini  
1929: {\it  On the Existence of Collisionless Equivariant Minimizers for the Classical
1930: n-body Problem.} Math ArXivs,         math-ph 0302022 [2003]. 
1931: 
1932: 
1933: [Gordon] W. B. Gordon, {\it A minimizing property of keplerian orbits}, American
1934: Journal of Mathematics, vol. 99,  $n^05$,   961-971, (1970).
1935: 
1936: [Fuji]  T. Fujiwara, H. Fukuda,  A. Kameyama,  H. Ozaki, M. Yamada,
1937: {\it  Synchronised Similar Triangles for Three-Body Orbit
1938: with Zero Angular Momentum}, arxiv.org/abs/math-ph/0404056.
1939: 
1940: [Hermann] R. Hermann, {\it On the differential geometry of foliations}, Ann. of
1941: Math. (2), (1959), 445-457 
1942: 
1943: [Math] Mathematical Society of Japan, {\bf Encyclopedic Dictionary of
1944: Mathematics},  by the Mathematical Society of Japan, ed. by S Iyanga and Y Kawada,
1945: translated by K. O. May, The MIT Press, Cambridge, Massachussets, and
1946: London, England, [1977] 
1947: 
1948:  
1949: [MontI]Richard Montgomery, {\it Infinitely Many Syzygies}, 
1950: Archives for  Rational Mechanics and Analysis, 
1951: v. 164 (2002), no. 4, 311--340, 2002. 
1952: 
1953: 
1954: [MontN] Richard Montgomery,  {\it The N-body problem, the braid group, and
1955: action-minimizing periodic orbits}, Nonlinearity, vol. 11, no. 2, 363-376, 1998.
1956: 
1957: [MontR] Richard Montgomery, 
1958: {\it  Geometric Phase of the Three-Body Problem}, Nonlinearity,
1959: vol. 9, no. 5, 1341-1360, 1996.
1960: 
1961:   
1962: [Moore] Cris Moore,  {\it Braids in Classical Gravity},  Physical Review
1963: Letters 70, pp. 3675--3679, (1993).
1964: 
1965: [Morse] H. M. Morse, {\it A one-to-one representation of geodesics on a surface of
1966: negative curvature}, Am. J. Math, {\bf 43}, no. 1, 33-51, 1921.
1967: 
1968: [Mumf] David Mumford, Caroline Series,David Wright, David  
1969: {\bf Indra's pearls.} 
1970: The vision of Felix Klein. 
1971: Cambridge University Press, New York, 2002. 
1972: 
1973: 
1974: [Poin] Poincar\'e, ~H. [1896], {\it Sur les solutions p\'eriodiques et le
1975: principe de moindre action}, .
1976:  {\it C.R.A.S. Paris} {\bf 123}, 915--918.
1977: 
1978: 
1979: 
1980: 
1981: 
1982: \medskip
1983:  
1984: \smallskip
1985: 
1986: %\insert bib
1987: 
1988: 
1989: \bigskip
1990: 
1991: \vfill\eject
1992: \end
1993: 
1994: 
1995: \end
1996: 
1997: 
1998: 
1999: