math0701174/bft.tex
1: %
2: %%
3: \documentclass[a4paper]{article}
4: \usepackage{geometry}
5: \input{assiomi.tex}
6: 
7: \geometry{a4paper,scale={0.65,0.80}}
8: 
9: % \RequirePackage{showkeys} %%%DRAFT%%%
10: \RequirePackage{amsmath}
11: \RequirePackage{amssymb}
12: \RequirePackage{amsthm}
13: \usepackage{epsfig}
14: \usepackage{graphicx}
15: \usepackage[dvips]{psfrag}
16: \usepackage[latin1]{inputenc}
17: \newcommand{\SSS}{\mathbb{S}}
18: \newcommand{\RR}{\mathbb{R}}
19: \newcommand{\NN}{\mathbb{N}}
20: \newcommand{\QQ}{\mathbb{Q}}
21: \newcommand{\ZZ}{\mathbb{Z}}
22: \newcommand{\CC}{\mathbb{C}}
23: \newcommand{\eps}{\varepsilon}
24: \newcommand{\todo}{....\marginpar{\fbox{\large{....}}}}
25: \newcommand{\action}{{\mathcal A}}
26: \newcommand{\cont}{{\mathcal C}}
27: % \newcommand{\elli}{{\mathcal E}}
28: \newcommand{\dint}{\displaystyle \int}
29: \newcommand{\dlim}{\displaystyle \lim}
30: \newcommand{\ds}{\displaystyle}
31: \newcommand{\dt}{\, dt}
32: 
33: % new set of axioms...
34: \newcommand{\uztag}{(U0)}
35: \newcommand{\uz}{\ref{a:U0}}
36: 
37: \newcommand{\uutag}{(U1)}
38: \newcommand{\uu}{\ref{a:U1}}
39: 
40: \newcommand{\udtag}{(U2)}
41: \newcommand{\ud}{\ref{a:U2}}
42: 
43: \newcommand{\udhtag}{(U2)$_{\mbox{h}}$}
44: \newcommand{\udh}{\ref{a:U2h}}
45: 
46: \newcommand{\udltag}{(U2)$_{\mbox{l}}$}
47: \newcommand{\udl}{\ref{a:U2l}{}}
48: 
49: \newcommand{\uthtag}{(U3)$_{\mbox{h}}$}
50: \newcommand{\uth}{\ref{a:U3h}}
51: 
52: \newcommand{\utltag}{(U3)$_{\mbox{l}}$}
53: \newcommand{\utl}{\ref{a:U3l}}
54: 
55: \newcommand{\uqhtag}{(U4)$_{\mbox{h}}$}
56: \newcommand{\uqh}{\ref{a:U4h}}
57: 
58: \newcommand{\uqltag}{(U4)$_{\mbox{l}}$}
59: \newcommand{\uql}{\ref{a:U4l}}
60: 
61: \newcommand{\uctag}{(U5)}
62: \newcommand{\uc}{\ref{a:U5}}
63: 
64: \newcommand{\ustag}{(U6)}
65: \newcommand{\us}{\ref{a:U6}}
66: 
67: \newcommand{\useitag}{(U6)}
68: \newcommand{\usei}{\ref{a:U6}}
69: 
70: \newcommand{\usettetag}{(U7)}
71: \newcommand{\usette}{\ref{a:U7}}
72: 
73: \newcommand{\usetteltag}{(U7)$_{\mbox{l}}$}
74: \newcommand{\usettel}{\ref{a:U7l}}
75: 
76: \newcommand{\usettehtag}{(U7)$_{\mbox{h}}$}
77: \newcommand{\usetteh}{\ref{a:U7h}}
78: 
79: 
80: \newcommand{\minus}{{\smallsetminus}}
81: \newcommand{\from}{\colon\thinspace}
82: 
83: 
84: % \swapnumbers
85: 
86: 
87: \theoremstyle{plain}
88: \newtheorem{theorem}{Theorem}
89: \newtheorem{lemma}{Lemma}
90: \newtheorem{corollary}[lemma]{Corollary}
91: \newtheorem{proposition}[lemma]{Proposition}
92: \newtheorem*{painleve}{Painlev\'e's Theorem}
93: \newtheorem*{vonzeipel}{Von Zeipel's Theorem}
94: 
95: \theoremstyle{definition}
96: \newtheorem{remark}[lemma]{Remark}
97: \newtheorem{definition}[lemma]{Definition}
98: \newtheorem{example}{Example}
99: 
100: \numberwithin{equation}{section}
101: \numberwithin{lemma}{section}
102: 
103: \renewcommand{\labelenumi}{\theenumi}
104: \renewcommand{\theenumi}{{\it(\roman{enumi})}}
105: 
106: %===========================
107: \title{On the singularities of generalized solutions \\to $n$--body type problems%
108: \footnote{This work is partially supported by Italy MIUR, national project
109: ``Variational Methods and Nonlinear Differential Equations''.}}
110: %===========================
111: \author{Vivina Barutello\footnote{Supported by Istituto Nazionale di Alta
112: Matematica.}\and Davide L.~Ferrario \and Susanna Terracini%
113: \footnote{\protect\address}
114: }
115: 
116: \newcommand{\address}{{Universit\`a di Milano Bicocca,}\\
117: {Dipartimento di Matematica e Applicazioni,}\\
118: {via  Cozzi 53, 20125 Milano.}\\
119: \small{e-mail {\ttfamily vivina.barutello@unimib.it,davide.ferrario@unimib.it,susanna.terracini@unimib.it}.}
120: }
121: 
122: 
123: %===========================
124: \begin{document}
125: %===========================
126: \maketitle
127: %===========================
128: \begin{abstract}
129: \noindent The validity of Sundman-type asymptotic estimates for collision solutions
130:  is established for a wide class of dynamical systems with singular forces, including the classical $N$--body problems with Newtonian, quasi--homogeneous and logarithmic potentials. The solutions are meant in the generalized sense of Morse (locally --in space and time-- minimal  trajectories with respect to compactly supported variations) and their uniform limits. The analysis includes the extension of the Von Zeipel's Theorem and the proof of isolatedness of collisions. Furthermore, such asymptotic analysis is applied to prove the absence of collisions for  locally minimal trajectories.
131: \end{abstract}
132: %===========================
133: \noindent 2000 {\it Mathematics Subject Classification.} \\
134: \noindent {\it Keywords.} Singularities in the $N$--body problem, locally
135: minimizing trajectories, collisionless solutions, logarithmic potentials.
136: %===========================
137: \section{Introduction}\label{sec:intro}
138: %===========================
139: %\cite{DPS1,DPS2,DS,diacu_book}
140: Many systems of interacting bodies of interest in Celestial and other areas of classical 
141: Mechanics have the form
142: \begin{equation} \label{DS_intro}
143: m_i\ddot x_i = \frac{\partial U}{\partial x_i}(t,x), \qquad i=1,\ldots,n
144: \end{equation}
145: where the  forces $\frac{\partial U}{\partial x_i}$ are undefined 
146: on a singular set $\Delta$. This is for example the set of collisions
147: between two or more particles in the $n$--body problem.  Such singularities play
148: a fundamental role in the phase portrait (see, e.g. \cite{diacu_book}) and strongly influence the global orbit
149: structure, as they can be held responsible, among others,
150: of the  presence of chaotic motions (see, e.g. \cite{Dev_survey}) and of motions
151: becoming unbounded in a finite time \cite{MM,xia}.
152: 
153: Two are the major steps in the analysis of the impact of the singularities 
154: in the $n$--body problem:  the first consists in performing the asymptotic
155: analysis along a single collision  (total or partial)  trajectory  and
156: goes back, in the classical case, to the works by  Sundman (\cite{sundman}), 
157: Wintner (\cite{wintner}) and, in
158: more recent years by Sperling, Pollard, Saari, Diacu and other authors
159: (see for instance \cite{polsaa1,polsaa2,saa72,sperling,elbialy,diacureg}).
160: The second step consists in blowing--up the singularity by a
161: suitable change of coordinates introduced by McGehee in \cite{mcgehee1} 
162: and replacing it by an invariant
163: boundary --the collision manifold-- where the flow can be extended in a
164: smooth manner. It turns out that, in many 
165: interesting applications, the flow on the collision manifold has a simple
166: structure: it is a gradient--like, Morse--Smale flow featuring a few stationary
167: points and heteroclinic connections (see, for instance, the surveys \cite{Dev_survey, moeckel88}).
168: The analysis of the extended flow allows us to obtain a full picture of the
169: behavior of solutions 
170: near the singularity, despite the flow fails to be fully regularizable (except
171: in a few cases).
172: 
173: The geometric approach, via the McGehee coordinates and the
174: collision manifold, 
175: can be successfully applied also to obtain asymptotic estimates
176: in some cases, such as  the 
177: collinear three--body problem (\cite{mcgehee1}), the anisotropic Kepler problem 
178: (\cite{Dev1,Dev2,DPS2,DS}),
179: the three--body problem both in the planar isosceles case (\cite{Dev3}) and  the
180: full perturbed three-body, as described in \cite{Dev_survey,diacuab}.
181: Besides the quoted cases, however, one needs
182: to establish the asymptotic estimates before
183: blowing--up the singularity, in order to prove convergence of the blow--up
184: family. The reason is quite 
185: technical and mainly rests in the fact the a singularity of the $n$--body
186: problems needs not  be
187: isolated, for the possible occurrence of  partial collisions in a 
188: neighborhood of the total collision. 
189: In the literature, this problem has been usually overcame by extending the flow
190: on partial collisions via some regularization technique (such as Sundman's, in \cite{Dev3}, 
191: or Levi--Civita's in \cite{LC}). Such a
192: device works well only when partial collisions are binary, which are the only
193: singularities to be globally removable.
194: Thus, the extension of the geometrical analysis to the full $n$--body problems
195: finds a strong theoretical obstruction:
196: partial collisions must be regularizable,  what is known to hold true only in few cases.
197: Other interesting cases in which the geometric method is not effective are that  of
198: quasi--homogeneous potentials
199: (where there is a lack of regularity for the extended flow)
200: and that of logarithmic potentials (for the failure of the blow--up technique).
201: 
202: In this paper we extend the classical asymptotic estimates near collisions in three
203: main directions.
204: \begin{enumerate}
205: \item 
206: We take into account of a very general notion of solution for the dynamical
207: system~\eqref{DS_intro},
208: which fits particularly well to solutions found by variational techniques. Our notion of
209: solution includes, besides all classical noncollision trajectories, all the  
210: \emph{locally minimal solutions} (with respect to compactly supported variations) 
211: which are often termed  minimal the sense of Morse. Furthermore, we include in the set of 
212: \emph{generalized solutions} all the limits of classical and locally minimal solutions.
213: \item We extend our analysis to a wide class of potentials including not only
214: homogeneous and quasi-homogeneous potentials, but also those with weaker
215: singularities of logarithmic type.
216: \item We allow potentials to strongly depend  on time
217: (we only require its time derivative 
218: to be controlled by the potential itself -- see assumption \ref{a:U1}). 
219: In this way, for instance, we 
220: can take into account models where masses vary in time.
221: \end{enumerate}
222: 
223: Our main results on the asymptotics near total collisions (at the origin) are Theorems 
224: \ref{central_conf} and~\ref{central_conf2} (for quasi-homogeneous potentials) and 
225: Theorems~\ref{central_conf_log} and~\ref{central_conf2_log} 
226: (when the potential is of logarithmic type) which extend the classical 
227: Sundman--Sperling asymptotic estimates (\cite{sundman,sperling}) in the directions above
228: (see also \cite{diacuab,DPS1}).
229: 
230: As a consequence of the asymptotic estimates, the
231: presence of a total collision prevents the occurrence of partial ones for
232: neighboring times.
233: 
234: This observation plays a central role when extending the
235: asymptotic estimates to the full $n$--body problem  since it allows us to reduce
236: from partial (even simultaneous) collisions to total ones by decomposing the
237: system in colliding clusters. Our results also lead to the extension of 
238: the concept of singularity for the dynamical system~\eqref{DS_intro} to the 
239: class of generalized solutions.
240: We shall prove an extension of the Von Zeipel's Theorem:
241: when the moment of inertia is bounded then every singularity of a generalized solution
242: admits a limiting configuration, hence all singularities are collisions.
243: The results on total collisions are then fully extended 
244: to partial ones in Theorem~\ref{thm:partial}.
245: 
246: A further motivation for the study of generalized solutions 
247: comes from the variational approach to the study of 
248: selected  trajectories to the $n$-body problem.
249: Indeed the exclusion of collisions is a major problem in the 
250: application of variational techniques as it results in the recent
251: literature, where many different arguments have been introduced to prove that the trajectories found in such a way
252: are collisionless (see \cite{amco93,ABT,BFT,BT2004,BC-Z,MR2032484,chenkyoto,CM,F3,FT,montgomery96,montgomery_contmath,mont_prepr,serter1,terven,andrea_thesis}).
253: As a  first application we shall be able to extend some of these techniques in order to 
254: prove that action minimizing trajectories are free of collisions for a wider class of 
255: interaction potentials. For example in the case of quasi--homoge\-neous potentials,
256: once collisions are isolated, the blow-up technique can be successfully applied 
257: to prove that locally minimal solutions are, in many circumstances, {\em free of collisions}.
258: In order to do that we can use the method 
259: of {\em averaged variations} introduced by Marchal and developed in \cite{marchal,Ch2,FT}. 
260: It has to be noticed that, when dealing with logarithmic--type potentials, the blow-up technique
261: is not available since  converging
262: blow-up sequences do not exists; we can anyway prove that the average 
263: over all possible variations is negative by taking advantage of the
264: harmonicity of the function $\log|x|$ in $\RR^2$. With this result we can then 
265: extend to quasi--homogeneous and logarithmic potentials all the analysis 
266: of the (equivariant) minimal trajectories carried in \cite{FT}.
267: 
268: Besides the direct method, other variational techniques --Morse and minimax theory--
269: have been applied to the search of periodic solutions in singular problems
270: (\cite{amco93,bahrab,mate95,riahi}). In the quoted papers, however 
271: only the case of \emph{strong force} interaction (see \cite{G}) has been
272: treated. Le us consider a  sequence of solutions to penalized problems
273: where an infinitesimal sequence of strong force terms is added to the potential:
274: then its limit enjoys the same conservation laws as the generalized solutions. Hence our
275: main results apply also to this class of trajectories.
276: We believe that our study can be usefully applied to
277: develop a Morse Theory that takes into account 
278: the topological contribution of collisions.
279: Partial results in this direction are given in \cite{BS,riahi2}, where the contribution
280:  of collisions to the Morse Index is computed.
281: 
282: The paper is organized as follows:
283: 
284: \tableofcontents
285: 
286: %===========================
287: \section{Singularities of locally minimal solutions}\label{sec:sing}
288: %===========================
289: %% __HERE__ 
290: \subsection{Locally minimal solutions}\label{sec:locminsol}
291: %===========================
292: We fix a metric on the \emph{configuration space} $\RR^k$
293: and we denote by $I(x)=|x|^2$ the \emph{moment of inertia} associated 
294: to the configuration $x \in
295: \RR^{k}$ and
296: \[
297: \mathcal{E}:=\{ x \in \RR^{k} : |x|^2 = 1 \}
298: \]
299: the inertia ellipsoid.
300: We define the radial and ``angular'' variables associated to $x \in \RR^{k}$ as
301: \begin{equation}
302: \label{rs}
303: r := |x| = I^{\frac12}(x) \in [0,+\infty), \quad
304: s := \frac{x}{|x|} \in {\mathcal E}.
305: \end{equation}
306: 
307: 
308: We consider the dynamical system 
309: \begin{equation} \label{DS}
310: \ddot x = \nabla U(t,x),
311: \end{equation}
312: on the time interval $(a,b)\subset \RR$, $-\infty \leq a < b \leq +\infty$.
313: Here $U$ is a positive time-dependent potential function 
314: $U \from   (a,b) \times \left(\RR^{k}\minus \Delta\right)  \rightarrow \RR^+$,
315: and it is supposed to be of class $\mathcal C^1$ on  its domain; 
316: by $\nabla U$ we denote  its gradient
317: with respect to the given metric.
318: \begin{remark}
319: In the case of $n$-body type systems as described in~\eqref{DS_intro}, given $m_1,\ldots,m_n$, $n \geq 2$ positive real numbers, 
320: we define the scalar product induced by the {\it mass metric} 
321: on the \emph{configuration space} $\RR^{nd}$ between $x=(x_1,\ldots,x_n)$ 
322: and $y=(y_1,\ldots,y_n)$, as 
323: \begin{equation} \label{mass_prod}
324: x \cdot y = \sum_{i=1}^{n}m_i \langle x_i,y_i \rangle,
325: \end{equation}
326: where $\langle\cdot,\cdot\rangle$ is the scalar product in $\RR^{d}$.
327: We denote by $|\cdot|$ the norm induced by the mass scalar
328: product~\eqref{mass_prod}. Then 
329: $\nabla U(t,x)$ denotes the gradient of the potential, in the mass metric, with respect to the spatial variable $x$,
330: that is:
331: \[
332: \nabla U(t,x)=M^{-1} \frac{\partial U}{\partial x}(t,x),
333: \]
334: where $\left(\frac{\partial U}{\partial x}\right)_i=\frac{\partial U}{\partial x_i}$,
335: $i=1,\ldots,n$, and $M=[M_{ij}]$, 
336: $M_{ij}=m_i\delta_{ij}{\mathbf 1}_d$ (${\mathbf 1}_d$ is the $d$-dimensional
337: identity matrix) for every $i,j=1,\ldots,n$.
338: \end{remark}
339: 
340: Furthermore we suppose that $\Delta$ is a singular set for $U$ of an attractive
341: type, in the sense that
342: \begin{assiomi}
343: \item[\uztag]\label{a:U0}
344: $\displaystyle\lim_{x \to \Delta} U(t,x) =
345: +\infty$, uniformly in $t$.
346: \end{assiomi}
347: Borrowing the terminology from the study of the singularities of the $n$--body problem, 
348: the set $\Delta$ will be often referred as \emph{collision set} and it is
349: required to be a \emph{cone}, that is\label{cone}
350: \[
351: x \in \Delta \quad \implies \quad \lambda x\in\Delta, \quad \forall \lambda\in
352: \RR.
353: \]
354: We observe that being a cone implies that  $0 \in \Delta$. 
355: When $x(t^*) \in \Delta$ for some $t^*\in(a,b)$ we will say that $x$ has
356: an \emph{interior collision} at $t=t^*$ and that $t^*$ is a
357: \emph{collision instant} for $x$. When $t^* = a$ or $t^* = b$ (when
358: finite) we will talk about a \emph{boundary collision}. In particular, if
359: $x(t^*)=0\in \Delta$, we will say that $x$ has a \emph{total collision at the origin} at
360: $t=t^*$.
361: A collision instant $t^*$ is termed
362: \emph{isolated} if there exists $\delta>0$ such that, for every
363: $t \in (t^*-\delta,t^*+\delta)\cap (a,b)$, $x(t)\notin \Delta$.
364: 
365: We consider the following assumptions on the potential $U$:
366: 
367: \begin{assiomi}
368: \item[\uutag] 
369: \label{a:U1}
370: \text{There exists a constant $C_1 \geq 0$ such that, for every $(t,x) \in (a,b)
371: \times\left( \RR^{k}\minus \Delta\right)$},
372: \[
373: \left|\frac{\partial U}{\partial t}(t,x)\right| \leq C_1 \left(U(t,x)+1\right).
374: \]
375: \item[\udtag] 
376: \label{a:U2}
377: \text{There exist constants $\tilde \alpha \in (0,2)$ and $C_2 \geq 0$ such
378: that}
379: \[
380: \nabla U(t,x)\cdot x + \tilde \alpha U(t,x) \geq - C_2.
381: \]
382: \end{assiomi}
383: 
384: % \begin{remark}
385: % When $U$ is homogeneous of degree $-\alpha$ in its second
386: % variable (for instance when $U(t,x)=1/|x|^\alpha$), condition \ud\,
387: % is satisfied and the equality is achieved with $\tilde \alpha=\alpha$ and $C_2=0$. 
388: % Furthermore, \ud\, is satisfied when $U(t,x)=\log(1/|x|)$ for every value of\enspace
389: % $\tilde \alpha$.
390: % \end{remark}
391: 
392: We then define the \textit{lagrangian action functional} on the interval $(a,b)$
393: as
394: \begin{equation}
395: \label{action}
396: \action(x,[a,b]) := \int_a^b K(\dot{x}) + U(t,x) \dt,
397: \end{equation}
398: where 
399: \begin{equation}
400: \label{kin}
401: K(\dot{x}) :=  \frac{1}{2} |\dot{x}|^2,
402: \end{equation}
403: is the \textit{kinetic energy}. We observe that $\action(\cdot,[a,b])$ is
404: bounded and ${\cal C}^2$ on the Hilbert space $H^1\left( (a,b),\RR^{k}\minus
405: \Delta \right)$. In terms of the variables $r$ and $s$ introduced
406: in~\eqref{rs}, 
407: the action functional reads as
408: \[
409: \action(rs,[a,b]):=\int_a^b\frac12\left(\dot r^2 + r^2|\dot
410: s|^2\right)+U(t,rs)\dt
411: \]
412: and the corresponding Euler--Lagrange equations, whenever $x \in H^1\left(
413: (a,b),\RR^{k}\minus \Delta \right)$, are
414: \begin{equation}
415: \label{eq:EL}
416: \begin{split}
417: -\ddot r + r|\dot s|^2 + \nabla U(t,r s) \cdot s & = 0 \\
418: -2r\dot r \dot s -r^2 \ddot s + r \nabla_{T} U(t,r s) & = \mu s
419: \end{split}
420: \end{equation}
421: where $\mu= r^2|\dot s|^2$ is the Lagrange multiplier due to the presence of the
422: constraint $|s|^2 = 1$ and the vector $\nabla_{T} U(t,r s)$ is the tangent
423: components to the ellipsoid ${\mathcal E}$ of the gradient $\nabla U(t,r s)$, that is
424: $\nabla_{T} U(t,rs)=\nabla U(t,r s)-\nabla U(t,r s)\cdot s$.
425: 
426: \begin{definition}\label{def:locally_minimal_sol}
427: A path $x \in H^1_{loc}\left((a,b),\RR^{k}\right)$ is a \emph{locally minimal
428: solution} for the dynamical system~\eqref{DS} if,
429:  for every $t_0 \in (a,b)$, there exists $\delta_0 >0$ such that the restriction
430: of $x$ to the interval $I_0=[t_0-\delta_0,t_0+\delta_0]$, is a local minimizer
431: for
432: $\action(\cdot,I_0)$ with respect to compactly supported variations (fixed-ends).
433: \end{definition}
434: 
435: \begin{remark}\label{rem:3}
436: We observe that a priori a locally minimal solution $x$ can have a large collision 
437: set, $x^{-1}(\Delta)$; this set, though of Lebesgue measure zero,  can very well 
438: admit many accumulation points. For this
439: reason the Euler--Lagrange equations~\eqref{eq:EL} and the dynamical 
440: system~\eqref{DS} do not hold for a locally minimal trajectory.
441: \end{remark}
442: \begin{remark}
443: When the potential is of class ${\cal C}^2$ outside $\Delta$ then every classical
444: noncollision solution in the interval $(a,b)$ is a a locally minimal solution.
445: \end{remark}
446: \begin{definition}\label{def:generalized_sol}
447: A path $x$ is a \emph{generalized solution} for the dynamical system~\eqref{DS} 
448: if there exists a sequence $x_n$ of locally minimal solutions such that
449: \begin{enumerate}
450: \item $x_n\to x$ uniformly on compact subsets of $(a,b)$;
451: \item for almost all  $t \in (a,b)$ the associated total energy
452: $h_n(t) := K(\dot x_n(t))-U(t,x_n(t))$ converges.
453: \end{enumerate}
454: \end{definition}
455: 
456: We say that a (classical, locally minimal, generalized) solution $x$ on the interval $(t_1,t_2)$,
457: has a {\em singularity} at  $t_2$ (finite) if it is not possible to extend $x$ as a (classical,
458: locally minimal, generalized) solution to a larger interval $(t_1,t_3)$ with $t_3>t_2$.
459: 
460: In the framework of classical solutions to $n$-body systems, 
461: the classical Painlev\'e's Theorem  (\cite{painleve,Dpainleve}) asserts
462: that the existence of a singularity
463: at a finite time $t^*$ is equivalent to the fact that the minimal 
464: of the mutual distances becomes infinitesimal as $t\to t^*$.
465: This fact reads as:
466: 
467: \begin{painleve}\label{teo:Painleve}
468: Let $\bar x$ be a classical solution for the $n$-body dynamical system
469: on the interval $[0,t^*)$. If $\bar x$ has a singularity at $t^*<+\infty$, then the
470: potential associated to the problem diverges to $+\infty$ as $t$ approaches $t^*$.
471: \end{painleve}
472: 
473: Painlev\'e's Theorem does not necessarily imply that a collision (i.e. that is a singularity
474: such that all mutual distances have a definite limit) occurs when
475: there is a singularity at a finite time; indeed this two facts are equivalent 
476: only if each particle approaches a definite configuration
477: (on this subject we refer to \cite{polsaa1,polsaa2,saa72}).
478: This result has been stated by Von Zeipel in 1908 (see \cite{vonzeipel} and also \cite{mcgehee2})
479: and definitely proved by Sperling in 1970 (see \cite{sperling}): 
480: in the $n$-body problem the occurrence of singularities (in finite time) 
481: which are not collisions is then equivalent to the existence of an unbounded motion.
482: 
483: \begin{vonzeipel}\label{teo:Zeipel}
484: If $\bar x$ is a classical solution for the $n$-body dynamical system
485: on the interval $[0,t^*)$ with a singularity at $t^*<+\infty$ and
486: $\lim_{t \to t^*}I(\bar x(t))<+\infty$, then $\bar x(t)$ has a definite limit
487: configuration $x^*$ as $t$ tends to $t^*$.
488: \end{vonzeipel}
489: We will come back later on the proof of this result (in Corollary~\ref{cor:2}
490: and in Section~\ref{sec:partial}). To our purposes, we give the following definition.
491: 
492: \begin{definition}\label{def:singularity}
493: We say that the (generalized) solution $\bar x$ for 
494: the dynamical system~\eqref{DS} has a {\em singularity} at $t=t^*$ if
495: \[
496: \lim_{t \to t^*}U(t,\bar x(t))=+\infty.
497: \]
498: \end{definition}
499: 
500: \begin{definition}\label{def:collision}
501: The singularity $t^*$ is a said to be a \emph{collision} 
502: for the locally minimal 
503: solution $\bar x$ if it admits a limit configuration as $t$ tends to $t^*$.
504: \end{definition}
505: 
506: %=======================================================
507: \subsection{Approximation of locally minimal solutions}
508: %=======================================================
509: Let $\bar x$ be a locally minimal solution on the interval $(a,b)$ and let
510: $I_0\subset(a,b)$ be an interval 
511: such that $\bar x$ is a (local) minimizer for $\action(\cdot,I_0)$ with respect
512: to compactly supported variations. Generally local minimizers need not to be
513: isolated; we illustrate below a penalization argument to select a particular
514: solution from the possibly large set of local minimizers.
515: To begin with, we define the auxiliary functional on the space $H^1\left( I_0,
516: \RR^{k}\right)$
517: \begin{equation}
518: \label{baraction}
519: \bar \action(x,I_0) := 
520: \int_{I_0} K(\dot{x}) + U(t,x) +\dfrac{|x-\bar x|^2}{2}\dt.
521: \end{equation}
522: When the interval $I_0$ is sufficiently
523: small, $\bar x$ is actually the global minimizer for the penalized functional
524: $\bar \action(\cdot,I_0)$
525: defined in~\eqref{baraction}. Of course we may assume that
526: \begin{equation}\label{bddA}
527: \bar \action(\bar x,I_0) = \action(\bar x,I_0) < + \infty,
528: \end{equation}
529: which is equivalent to require that $\bar \action(\cdot,I_0)$ takes a finite
530: value at least at one point.
531: 
532: \begin{proposition} \label{propo:globalmin}
533: Let $\bar x$ be a locally minimal solution on the interval $(a,b)$, let
534: $\delta_0>0$ and $t_0 \in (a,b)$ be such that $\bar x$ is a local minimizer for
535: $\action(\cdot,I_0)$ where  $I_0=[t_0-\delta_0,t_0+\delta_0]\subset(a,b)$.
536: Then there exists $\bar \delta = \bar \delta(\bar x)>0$ such that whenever
537: $\delta_0 \leq \bar \delta$, $\bar x$ is the unique global minimizer for $\bar
538: \action(\cdot,I_0)$.
539: \end{proposition}
540: 
541: \begin{proof}
542: For every $x \in H^1_{loc}\left( I_0, \RR^{k}\right)$ the 
543: inequality $\action(x,I_0)
544: \leq \bar \action(x,I_0)$ holds true,
545: and it is an equality only if $x=\bar x$. Since
546: $\bar x$ is a local minimizer for $\action(x,I_0)$, one easily infers, by a
547: simple convexity argument, the existence of $\eps > 0$ such that
548: \[\|x-\bar x\|_\infty < \eps\quad\implies\quad\action(\bar x,I_0) \leq
549: \action(x,I_0).\]
550: We conclude that, for every $x \in H^1_{loc}\left( I_0, \RR^{k}\right)$,
551: such that $0<\|x-\bar x\|_\infty < \eps$,  the following chain 
552: of inequalities holds
553: \[
554: \bar \action(\bar x,I_0) = \action(\bar x,I_0) \leq \action(x,I_0) < \bar
555: \action(x,I_0);
556: \]
557: hence $\bar x$ is a strict local minimizer for $\bar\action(\cdot,I_0)$,
558: independently on 
559: $\delta_0$.
560: 
561: In order to complete the proof we show that $\bar \action(\bar x,I_0) < \bar
562: \action(x,I_0)$ also for those functions 
563: $x \in H^1_{loc}\left( I_0, \RR^{k}\right)$ such that $\|x-\bar x\|_\infty \geq
564: \eps$, provided $\delta_0$ is sufficiently small. Indeed, since  the Sobolev
565: space $H^1_{loc}\left( I_0, \RR^{k}\right)$ is embedded in the space of
566: absolutely continuous functions, we can compute, by H\"older inequality,
567: \begin{equation} \label{eq:dis}
568: \begin{split}
569: |(x-\bar x)(t)|&\leq \int_{I_0} |\dot x(s)|ds + \int_{I_0}|\dot{\bar x}(s)|ds \\
570:                &\leq \sqrt{2\delta_0} \left( \sqrt{\int_{I_0} |\dot x(s)|^2ds} 
571:                      + \sqrt{\int_{I_0}|\dot{\bar x}(s)|^2ds}\right).
572: \end{split}
573: \end{equation}
574: By taking the supremum at both sides of~\eqref{eq:dis} it follows that 
575: \[
576: \dfrac{\|x-\bar x\|_\infty}{\sqrt{2\delta_0}} - \sqrt{\int_{I_0}|\dot{\bar
577: x}(s)|^2\, ds} \leq
578: \sqrt{\int_{I_0} |\dot x(s)|^2\, ds},
579: \]
580: and therefore, for every $x \in H^1\left( I_0, \RR^{k}\right)$,
581: \begin{equation} \label{eq:dis2}
582: \begin{split}
583: \bar \action(x,I_0) \geq \int_{I_0} |\dot x(s)|^2\, ds 
584: &\geq \left( \dfrac{\|x-\bar x\|_\infty}{\sqrt{2\delta_0}}
585:              -\sqrt{\int_{I_0}|\dot{\bar x}(s)|^2\, ds}\right)^2\\
586: &\geq \left( \dfrac{\eps}{\sqrt{2\delta_0}} -\sqrt{\int_{I_0}|\dot{\bar
587: x}(s)|^2\, ds}\right)^2
588: \end{split}
589: \end{equation}
590: Hence, by choosing $\delta_0$ such that 
591: $2\delta_0 < \eps\left(\sqrt{\int_{I_0}|\dot{\bar x}(s)|^2ds}+\sqrt{\bar
592: \action(\bar x,I_0)}\right)^{-2}$, it follows that 
593: \[
594: \bar \action(x,I_0) \geq \left( \dfrac{\eps}{\sqrt{2\delta_0}} - 
595: \sqrt{\int_{I_0}|\dot{\bar x}(s)|^2ds}\right)^2 > \bar \action(\bar x,I_0) 
596: \]
597: also for those paths $x \in H^1_{loc}\left( I_0, \RR^{k}\right)$ such that
598: $\|x-\bar x\|_\infty \geq \eps$.
599: This concludes the proof.
600: \end{proof}
601: 
602: We now wish to approximate the singular potential $U$ with a family of smooth
603: potentials $U_{\eps}\from (a,b)\times \RR^{k} \to \RR^+$, 
604: depending on a parameter
605: $\eps > 0$.
606: To this aim consider  the function
607: \[
608: \eta(s) = 
609: \begin{cases}
610: s & \text{if $s\in [0,1]$} \\
611: \dfrac{-s^2 +6s - 1}{4} & \text{if $s\in [1,3]$} \\
612: 2 & \text{if $s \geq 3$;}
613: \end{cases}
614: \]
615: notice that
616: $\eta\in\cont^{1}\left(\RR^+,\RR^+\right) $ and,  
617: % \begin{equation*}
618: % \label{eq:eta}
619: % \eta(s)=s \, \text{ when } \, s \in \left[0,{1}\right], \qquad \eta(s)=2 \,
620: % \mbox{ when } s \geq 3
621: % \end{equation*}
622: for every $s \in [0,+\infty)$,
623: \begin{equation*}
624: % \label{eta_ineq}
625: \dot\eta(s) \, s \leq \eta(s) \qquad \text{and} \qquad \dot\eta(s)\leq 1.
626: \end{equation*}
627: Now let us define, for $\eps >0$,
628: \[
629: \eta_\eps(s) := \frac{1}{\eps}\eta(\eps s);
630: \]
631: then the following inequalities hold for every $s \in [0,+\infty)$
632: \begin{equation}\label{etaeps_ineq}
633: \dot\eta_\eps(s) \, s \leq \eta_\eps(s), \qquad \text{and} \qquad
634: \dot\eta_\eps(s)\leq 1.
635: \end{equation}
636: By means of the family $\eta_\eps$ we can regularize the potential $U$ in the
637: following way: 
638: \begin{equation}
639: \label{def:Ueps}
640: U_{\eps}(t,x) = \left\{
641: \begin{array}{lll}
642: \displaystyle
643: \eta_\eps\left(U(t,x)\right), &   & \mbox{if } x \in \RR^{k}\minus
644: \Delta,\\ 
645: {2}/{\eps},                   &   & \mbox{if } x \in \Delta.
646: \end{array} \right.
647: \end{equation}
648: It is worthwhile to understand that each $U_\eps(t,x)$ coincides with $U(t,x)$
649: whenever $U(t,x) \leq {1}/{\eps}$; in fact
650: \[
651: \eta_\eps(s)=\frac{1}{\eps}\eta(\eps s)=s
652: \]
653: whenever $\eps s \in [0,1]$, that is $s \in [0,1/\eps]$.
654: Next, we  consider the associated family of boundary value problems on the
655: interval $I_0 \subset (a,b)$
656: \begin{equation}
657: \label{DSbar}
658: \left\{
659: \begin{array}{l}
660: \ddot x = \nabla U_\eps(t,x) + (x-\bar x), \\
661: x\vert_{ \partial I_0 } = \bar x \vert_{\partial I_0} 
662: \end{array}\right.
663: \end{equation}
664: where, as usual, $\nabla U_\eps(t,x)$ is the gradient, in the mass metric, with respect
665: to the spatial variable $x$. Solutions of~\eqref{DSbar} are critical points of the
666: action functional
667: \begin{equation}
668: \label{baractioneps}
669: \bar \action_\eps(x,I_0) := \int_{I_0} K(\dot{x}) + U_\eps(t,x) +\dfrac{|x-\bar
670: x|^2}{2}\dt.
671: \end{equation}
672: We observe that $\bar \action_\eps(\cdot,I_0)$ is bounded and $\cont^2$ on
673: $H^1_{loc}\left( I_0, \RR^{k}\right)$, since $U_\eps$ is smooth on the whole
674: $\RR^{k}$.
675: We also remark that the infimum of $\bar \action_\eps(\cdot,I_0)$ is achieved,
676: for $\bar \action_\eps(\cdot,I_0)$ is a positive and coercive functional on 
677: $H^1_{loc}\left( I_0, \RR^{k}\right)$.
678: 
679: In the next proposition we prove that a locally minimal solution 
680: has the fundamental property to be the limit of a
681: sequence of global minimizers for the approximating functionals $\bar
682: \action_\eps(\cdot,I_0)$, provided the interval $I_0 \subset (a,b)$ is chosen so
683: small  that the restriction of the minimal solution to $I_0$ is the unique
684: global minimizer for $\bar\action(\cdot,I_0)$.  This result is crucial, indeed, as observed in 
685: Remark~\ref{rem:3}, the Euler--Lagrange equations and the dynamical system 
686: hold for a locally minimal solution; we will anyway be able to use the ones corresponding
687: to the approximating global minimizers (for the regularized problems) to prove the fundamental
688: properties of locally minimal (and generalized) solutions in the rest of the paper.
689: 
690: \begin{proposition}\label{propo:conv}
691: Let $\bar x$ and $I_0$ be given by Proposition~\ref{propo:globalmin}.
692: Let $\epsilon>0$ and $x_\eps$ be a global minimizer for $\bar
693: \action_{\eps}(\cdot,I_0)$. Then, up to subsequences, as $\eps \to 0$,
694: \begin{enumerate}
695: \item[(i)] $U_{\eps}(t,x_{\eps}) \to U(t,\bar x)$ almost everywhere and in $L^1$;
696: \item[(ii)] $x_{\eps}\to \bar x$ uniformly; 
697: \item[(iii)] $\dot x_{\eps}\to \dot{\bar x}$ in $L^2$;
698: \item[(iv)] $\dot x_{\eps}\to \dot{\bar x}$ almost everywhere;
699: \item[(v)] $\dfrac{\partial U_{\eps}}{\partial t}(t,x_{\eps}) \to \dfrac{\partial
700: U}{\partial t}(t,\bar x)$ almost everywhere and in $L^1$.
701: \end{enumerate}
702: \end{proposition}
703: \begin{proof}
704: As we have already observed, for every $\eps>0$, the potential $U_\eps$
705: coincides with $U$ on the sublevel $\{(t,x):U(t,x)\leq 1/\eps\}$ and, by its
706: definition, for every $(t,x) \in I_0\times \RR^{k}\minus\Delta$
707: \[
708: U_\eps(t,x) \leq U(t,x).
709: \]
710: Therefore 
711: \[
712: \bar \action_\eps(x,I_0) \leq \bar \action(x,I_0)
713: \]
714: for every $x \in H^1_{loc}\left(I_0,\RR^{k}\right)$.
715: It follows from~\eqref{bddA} that
716: \begin{equation}\label{bddineq}
717: \bar \action_\eps(x_\eps,I_0) = \inf_{x \in H^1_{loc}} \bar \action_\eps(x,I_0)
718: \leq \bar \action(\bar x,I_0) < +\infty,
719: \end{equation}
720: which implies the boundedness of the family $\left\{\int_{I_0}|\dot{x}_\eps|^2
721: +|x_\eps-\bar x|^2\right\}_\eps$.
722: Hence we deduce the existence of a sequence $\left(x_{\eps_n} \right)_{\eps_n}
723: \subset (x_\eps)_\eps$ such that $\left(\dot x_{\eps_n} \right)_{\eps_n}$
724: converges weakly in $L^2$ and uniformly to some limit $\tilde x$. In addition we
725: observe that 
726: \[
727: \lim_{\eps_n \to 0}U_{\eps_n}(t,x_{\eps_n}(t)) = U(t,\tilde x(t))
728: \]
729: for every $t \in I_0$, regardless the finiteness of $U(t,\tilde x(t))$.\\
730: From~\eqref{bddineq} we also deduce the boundedness of the following integrals 
731: \[
732: \int_{I_0}U_{\eps_n}(t,x_{\eps_n}) \,dt \leq \bar \action_{\eps_n}(x_{\eps_n},I_0)
733: < +\infty.
734: \]
735: and therefore, since the sequence $\left( U_{\eps_n}(t,x_{\eps_n})
736: \right)_{\eps_n}$
737: is positive, by applying Fatou's Lemma  one deduces that 
738: \[
739: \int_{I_0} U (t,\tilde x) \leq \liminf \int_{I_0}
740: U_{\eps_n}(t,x_{\eps_n})<+\infty.
741: \]
742: Hence from the weak semicontinuity of the norm in $L^2$ (the sequence
743: $(\dot x_{\eps_n})_{\eps_n}$ converges weakly in $L^2$ to $\tilde x$) we obtain the
744: inequalities
745: \begin{equation*}
746: \bar \action (\tilde x,I_0) \leq \liminf \bar \action_{\eps_n}(x_{\eps_n},I_0)
747: \leq \bar \action(\bar x,I_0)
748: \end{equation*}
749: which contradict Proposition~\ref{propo:globalmin}, unless $\tilde x = \bar x$
750: and
751: \begin{equation}\label{disA}
752: \liminf \bar \action_{\eps_n}(x_{\eps_n},I_0) = \bar \action(\bar x,I_0).
753: \end{equation}
754: Therefore we deduce the $L^2$-convergence of the sequence $\left(\dot x_{\eps_n}
755: \right)_{\eps_n}$
756: and its convergence almost everywhere to $\dot{\bar x}$, up to subsequences.
757: From~\eqref{disA} it follows  also that
758: \begin{equation}\label{conv_int}
759: \lim_{\eps_n \to 0}\dint_{I_0} U_{\eps_n}(t,x_{\eps_n}) =\dint_{I_0}U(t,\bar x).
760: \end{equation}
761: From the convergence almost everywhere of $(U_{\eps_n}(t,x_{\eps_n}))_{\eps_n}$
762: together with~\eqref{conv_int} we conclude its convergence in $L^1$ to $U(t,\bar
763: x)$.
764: 
765: We now turn to the convergence of the sequence
766: $(\varphi_n(t))_{\eps_n}=\left(\dfrac{\partial U_{\eps_n}}{\partial
767: t}(t,x_{\eps_n})\right)_{\eps_n}$. To this aim, we observe that condition \ref{a:U1}
768: together with~\eqref{etaeps_ineq} imply the following chain of inequalities
769: \[
770: \begin{split}
771: \left|\frac{\partial U_{\eps_n}}{\partial t}(t,x_{\eps_n}(t))\right| 
772: & = \dot\eta_{\eps_n}\left(U(t,x_{\eps_n}(t))\right)\left|\frac{\partial
773: U}{\partial t}(t,x_{\eps_n}(t))\right| \\
774: & \leq C_1 \dot\eta_{\eps_n}\left( U(t,x_{\eps_n}(t))\right) 
775: \left(U(t,x_{\eps_n}(t))+1\right)\\
776: & \leq C_1 \left(\eta_{\eps_n}\left( U(t,x_{\eps_n}(t))\right)+1\right) \\
777: & = C_1 \left(U_{\eps_n}(t,x_{\eps_n}(t))+1\right).
778: \end{split}
779: \]
780: We already know that $U_{\eps_n}(t,x_{\eps_n}(t))$ converges in $L^1$. 
781: This implies the finiteness almost everywhere of $(\varphi_n(t))_{\eps_n}$ and
782: hence its almost everywhere 
783: convergence is due to the uniform convergence of
784: $(x_{\eps_n})_{\eps_n}$.
785: We obtain the $L^1$ convergence of $(\varphi_n(t))_{\eps_n}$ to $\dfrac{\partial
786: U}{\partial t}(t,\bar x)$ from the Dominated Convergence Theorem.
787: \end{proof}
788: 
789: %==============================================
790: \subsection{Conservation laws}
791: \label{subsec:conlaws}
792: %==============================================
793: Now the sequence of solutions to the regularized problems are used to prove
794: the conservation of the energy for locally minimal solutions.
795: 
796: \begin{proposition}\label{propo:convenergy}
797: Let $\bar x$ and $I_0$ be given by Proposition~\ref{propo:globalmin}. 
798: Then the energy associated to\enspace$\bar x$
799: \begin{equation}
800: \label{eq:energy}
801: h\from I_0 \to \RR, \qquad h(t) := K(\dot{\bar x}(t)) - U(t,\bar x(t)) 
802: \end{equation}
803: is of class $W^{1,1}$ on $I_0$ and its weak derivative is
804: \[
805: \dot h(t)= \frac{\partial U}{\partial t}(t,\bar x).
806: \]
807: \end{proposition}
808: %
809: \begin{proof}
810: Let $(x_{\eps})_{\eps}$ be the sequence of global minimizers for the
811: corresponding functionals $\bar \action_{\eps}(\cdot,I_0)$ convergent to $\bar
812: x$ whose existence is proved in Proposition~\ref{propo:conv}.
813: Let $h_{\eps}$ be the energy associated to $x_{\eps}$, that is
814: \begin{equation}
815: \label{eq:energyeps}
816: h_{\eps} \from  I_0 \to \RR,  \qquad 
817: h_{\eps}(t) := K({\dot x}_{\eps}(t))-U_{\eps}(t,x_{\eps}(t))+\frac{1}{2}|\bar
818: x(t)- x_{\eps}(t)|^2
819: \end{equation}
820: From Proposition~\ref{propo:conv} we immediately deduce that the sequence
821: $(h_{\eps})_{\eps}$ converges pointwise to $h(\bar x)$; moreover 
822: from~\eqref{bddineq} and~\eqref{eq:energyeps} 
823: we have
824: \[
825: \int_{I_0} |h_{\eps}(t)|\dt \leq \bar\action_{\eps}\left(x_{\eps},I_0\right) <
826: \bar\action\left(\bar x,I_0\right).
827: \]
828: From the Dominated Convergence Theorem we obtain that the sequence
829: $(h_{\eps})_{\eps}$ converges in $L^1$ to the integrable function $h$.
830: 
831: We still have to prove that $h$ admits weak derivative. To this end, let us
832: consider a test function $\varphi \in \cont_0^\infty(I_0)$; we can  write
833: \[
834: \begin{split}
835: \int_{I_0}h(t)\dot \varphi(t)\dt 
836: & = \lim_{\eps \to 0} \int_{I_0}h_{\eps}(t)\dot \varphi(t)\dt \\
837: & = \lim_{\eps \to 0} -\int_{I_0}\frac{\partial U_{\eps}}{\partial
838: t}(t,x_{\eps}(t))\varphi(t)\dt.
839: \end{split}
840: \]
841: In consequence of Proposition~\ref{propo:conv}, the sequence $\left(\frac{\partial
842: U_{\eps}}{\partial t}(t,x_{\eps}(t))\right)_\eps$ converges to
843: $\frac{\partial U}{\partial t}(t,\bar x(t))$ in $L^1$; then
844: \begin{equation}\label{limdU}
845: \lim_{\eps \to 0} \int_{I_0}\frac{\partial U_{\eps}}{\partial t}(t,x_{\eps}(t))
846: \varphi(t)\dt
847: =  \int_{I_0}\frac{\partial U}{\partial t}(t,\bar x(t)) \varphi(t)\dt, \qquad 
848: \forall \varphi \in \cont_0^\infty(I_0)
849: \end{equation}
850: and hence
851: \[
852: \int_{I_0}h(t)\dot \varphi(t) \dt 
853: = - \int_{I_0}\dfrac{\partial U}{\partial t}(t,\bar x)\varphi(t) \dt, \qquad 
854: \forall \varphi \in \cont_0^\infty(I_0)
855: \]
856: which means that 
857: $\frac{\partial U}{\partial t}(t,\bar x)$ is the weak derivative of
858: $h(\bar x)$.
859: \end{proof}
860: The next corollary follows straightforwardly.
861: \begin{corollary}\label{cor:en_abs_cont}
862: The energy associated to a locally minimal solution on the interval $(a,b)$ is
863: in the Sobolev space $W^{1,1}_{loc}\left((a,b),\RR\right)$.
864: \end{corollary}
865: 
866: We now investigate the behavior of the moment of inertia of a locally minimal solution
867: when a singularity occurs
868: (see Definition~\ref{def:singularity}). The results contained 
869: in Proposition~\ref{propo:Iconvex} and Corollary~\ref{cor:I''>0} are the natural extension
870: of the classical Lagrange--Jacobi inequality to locally minimal solutions (see \cite{wintner}).
871: 
872: \begin{proposition}\label{propo:Iconvex}
873: Let $\bar x$ be a locally minimal solution and $I_0$ be given by Proposition
874: \ref{propo:globalmin}.
875: Then 
876: \begin{equation} \label{disconv}
877: \frac{1}{2}\int_{I_0}I(\bar x(t))\ddot \varphi (t)\dt \geq 
878: \int_{I_0} \left[2h(\bar x(t))+(2-\tilde\alpha)U(t,\bar x(t))-C_2\right]\varphi (t)\dt
879: \end{equation}
880: for every $\varphi \in \cont_0^\infty\left(I_0,\RR\right)$, $\varphi(t)\geq 0$.
881: \end{proposition}
882: \begin{proof}
883: Let $(x_{\eps})_{\eps}$ be the sequence of global minimizers for the
884: corresponding functionals $\bar \action_{\eps}(\cdot,I_0)$ convergent to $\bar
885: x$ whose existence is proved in Proposition~\ref{propo:conv}. 
886: When we compute the second derivative of the moment of inertia of $x_{\eps}$ we
887: obtain
888: \[
889: \begin{split}
890: \frac{1}{2}\ddot I(x_{\eps}(t)) & = |\dot x_{\eps}(t)|^2 + \ddot x_{\eps}(t)
891: \cdot x_{\eps}(t)\\
892: &= 2h_{\eps}(t)+2U_{\eps}(t,x_{\eps}(t))-|\bar x(t)-x_{\eps}(t)|^2 \\
893: &\hspace*{3.5cm} + \left[\nabla U_{\eps}(t,x_{\eps}(t))+(x_{\eps}(t)-\bar
894: x(t))\right]\cdot x_{\eps}(t)\\
895: &= 2h_{\eps}(t)+2U_{\eps}(t,x_{\eps}(t))+\bar x(t)\cdot(x_{\eps}(t)-\bar x(t)) \\ 
896: &\hspace*{3.5cm} + \dot\eta_{\eps}\left(U(t,x_{\eps})\right) \nabla
897: U(t,x_{\eps}(t)) \cdot x_{\eps}(t)
898: \end{split}
899: \]
900: hence, by assumption \ref{a:U2} on the potential $U$ and
901: inequality~\eqref{etaeps_ineq}, it follows that 
902: \begin{equation}\label{dis:I''eps}
903: \begin{split}
904: \frac{1}{2}\ddot I(x_{\eps}(t)) &\geq 2h_{\eps}(t)+2U_{\eps}(t,x_{\eps}(t)) +
905: \bar x(t)\cdot(x_{\eps}(t)-\bar x(t))\\
906: &\hspace*{4cm}-\dot\eta_{\eps}\left(U(t,x_{\eps})\right)\left[\alpha
907: U(t,x_{\eps})+C_2\right]\\
908: &\geq 2h_{\eps}(t) + (2-\tilde\alpha)U_{\eps}(t,x_{\eps}(t))
909: +\bar x(t)\cdot(x_{\eps}(t)-\bar x(t)) -C_2
910: \end{split}
911: \end{equation}
912: for some $\tilde \alpha\in (0,2)$ and $C_2>0$.
913: Therefore, since $x_{\eps} \in \cont^2(I_0)$, for every $\varphi \in
914: \cont_0^\infty\left(I_0,\RR\right)$, $\varphi(t)\geq 0$
915: \[
916: \frac{1}{2}\int_{I_0}I(x_{\eps}(t))\ddot \varphi (t)\dt =
917: \frac{1}{2}\int_{I_0}\ddot I(x_{\eps}(t))\varphi (t)\dt
918: \]
919: and, from~\eqref{dis:I''eps},
920: \[
921: \begin{split}
922: &\frac{1}{2}\int_{I_0}I(x_{\eps}(t))\ddot \varphi (t)\dt \\ 
923: & \hspace{1cm} \geq 
924: \int_{I_0} \left[ 2h_{\eps}(t)+(2-\tilde\alpha)U_{\eps}(t,x_{\eps}(t))+\bar
925: x(t)\cdot(x_{\eps}(t)-\bar x(t))-C_2 \right]\varphi (t)\dt.
926: \end{split}
927: \]
928: We conclude by passing to the limit as $\eps \to 0$ in~\eqref{dis:I''eps} and
929: using the $L^1$-convergences proved in Propositions~\ref{propo:conv} and
930: \ref{propo:convenergy}.
931: \end{proof}
932: 
933: The next corollaries follow directly.
934: \begin{corollary}[\textbf{Lagrange--Jacobi inequality}]\label{cor:I''>0}
935: Let $\bar x$ be given by Proposition~\ref{propo:globalmin}.
936: Then the following inequality holds in the distributional sense
937: \[
938: \frac{1}{2}\ddot I(\bar x(t)) \geq 
939: 2h(t)+(2-\tilde \alpha)U(t,\bar x(t))-C_2, \qquad \forall t \in (a,b).
940: \]
941: \end{corollary}
942: 
943: \begin{corollary}\label{cor:14}
944: Let $\bar x$ be given by Proposition~\ref{propo:globalmin}.
945: Then its moment of inertia is convex on $I_0$ whenever $\bar x$ has a singularity
946: in $t_0$ and $\delta_0$ is small enough.
947: \end{corollary}
948: \begin{proof}
949: Whenever $\eps$ and $\delta_0$ are sufficiently small, the right hand side 
950: of inequality~\eqref{dis:I''eps} is
951: strictly positive, indeed $h_{\eps}(t)$ is bounded, $x_\eps$ converges to $\bar
952: x$ uniformly and $U_{\eps}(t,x_{\eps}(t))$ diverges to $+\infty$. Whenever
953: $\eps$ is small enough we conclude that
954: \[
955: \ddot I(x_{\eps}(t)) > 0
956: \]
957: and hence $I(x_{\eps})$ are strictly convex functions in a neighborhood of
958: $t_0$. Since the sequence $I(x_{\eps})$ uniformly converges to $I(\bar x)$ we
959: conclude that  also $I(\bar x)$ is convex on the interval\enspace$I_0$.
960: \end{proof}
961: 
962: We now investigate 
963: the possibility that a sequence of singularities accumulates
964: at the right bound of the interval $(a,b)$; in this section we will
965: suppose that $b<+\infty$.
966: 
967: \begin{lemma}
968: \label{le:1}
969: Let $\bar x$ be given in Proposition~\ref{propo:globalmin}, $h$ be its energy
970: defined in~\eqref{eq:energy} and fix $\tau \in (a,b)$ be such that 
971: \begin{equation}\label{eq:lambda}
972: \lambda := \frac{2-\tilde\alpha}{2}-C_1(b-\tau)
973: \end{equation}
974: is a strictly positive constant. Then there exists a constant $K>0$ such that 
975: \begin{equation}
976: \label{eq:lemma1}
977: \left|\int_{\tau}^{t} h(s) ds\right| \leq \left(
978: \frac{2-\tilde\alpha}{2}-\lambda\right) \int_{\tau} ^{t} U(s,\bar x(s))ds + K, \qquad
979: \forall t \in (\tau,b).
980: \end{equation}
981: \end{lemma}
982: %
983: \begin{proof}
984: Since $h$ is absolutely continuous on every interval $[\tau,t] \subset (a,b)$
985: (Corollary~\ref{cor:en_abs_cont}) we have
986: \[
987: |h(t)| \leq |h(\tau)| + \int_{\tau}^t |\dot{h}(\xi)| \,d\xi, \qquad \forall t \in
988: (\tau,b).
989: \]
990: From Proposition~\ref{propo:convenergy} and assumption \ref{a:U1} we obtain
991: \begin{equation}
992: \label{bddh}
993: \begin{split}
994: |h(t)|
995: & \leq |h(\tau)| + \int_{\tau}^t \left|\frac{\partial U}{\partial
996: \xi}(\xi,\bar x(\xi))\right| \,d\xi \\
997: & \leq |h(\tau)| + C_1\int_{\tau}^t \left(U(\xi,\bar x(\xi))+1\right) \,d\xi
998: \end{split}
999: \end{equation}
1000: and integrating both sides of the inequality on the interval $[\tau,t]$
1001: \[
1002: \int_{\tau}^{t}|h(s)|ds \leq |h(\tau)|(t-\tau)+C_1\frac{(t-\tau)^2}{2}
1003: + C_1\int_{\tau}^t ds \int_{\tau}^s U(\xi,\bar x(\xi)) \,d\xi.
1004: \]
1005: Since $U$ is positive, the integral $\int_{\tau}^s U(\xi,\bar x(\xi)) \,d\xi$
1006: increases in the variable $s$, hence we conclude
1007: \[
1008: \left|\int_{\tau}^{t} h(s) ds\right| \leq \int_{\tau}^{t}|h(s)|ds \leq K +
1009: C_1(b-\tau)\int_{\tau}^t U(\xi,\bar x(\xi)) \,d\xi.
1010: \]
1011: where $K:=|h(\tau)|(t-\tau)+C_1 (t-\tau)^2/2$.
1012: \end{proof}
1013: %
1014: \begin{lemma}
1015: \label{le:2}
1016: Let $\bar x$ be given in Proposition~\ref{propo:globalmin} and $\tau$ be chosen
1017: as in Lemma~\ref{le:1}. Suppose that there exist $\delta,C>0$ such that
1018: \[
1019: I(\bar x(t)) \leq C, \mbox{ for every } t \in (b-\delta,b) 
1020: \]
1021: and 
1022: \begin{equation}
1023: \label{eq:liminf}
1024: \liminf_{t\to b^-} \dot I(\bar x(t)) \leq C.
1025: \end{equation}
1026: Then there exists $\tau \in (a,b)$ such that
1027: \[
1028: \int_{\tau} ^{b} U(t,\bar x(t)) \dt < +\infty.
1029: \]
1030: \end{lemma}
1031: %
1032: \begin{proof}
1033: If $b$ is not a singularity for $\bar x$,
1034: the assertion follows from assumption~\eqref{bddA}.
1035: Otherwise, 
1036: it follows from \eqref{eq:liminf} that 
1037: % since $\liminf_{t\to b^-} \dot I(\bar x(t)) \leq C$ 
1038: there exists
1039: an increasing sequence $(t_n)_n$ such that
1040: \[
1041: t_n \to b \mbox{ as } n\to+\infty\quad \mbox{and} \quad  \dot I(\bar x(t_n)) \leq C, \forall n.
1042: \]
1043: Now let $N$ be an integer
1044: such that $t_N \in (b-\delta,b)$ and the constant $\lambda$
1045: defined in~\eqref{eq:lambda}, with $\tau = t_N$, is strictly positive.
1046: Hence, for every index $n>N$,
1047: \[
1048: 2C \geq \dot I(\bar x(t_n)) - \dot I(\bar x(t_N))
1049: = \int_{t_N}^{t_n} \ddot I(\bar x(t)) \dt.
1050: \]
1051: Corollary~\ref{cor:I''>0} implies that
1052: \[
1053: C \geq 2 \int_{t_N}^{t_n} h(t) \dt + (2-\tilde\alpha)\int_{t_N}^{t_n} U(t,\bar x(t))
1054: \dt - C_2(t_n-t_N).
1055: \]
1056: We now apply Lemma~\ref{le:1} to deduce that 
1057: \begin{equation}
1058: \label{eq:end_le2}
1059: 2\lambda \int_{t_N}^{t_n} U(t,\bar x(t)) \dt \leq C_2(t_n-t_N) + C + 2K.
1060: \end{equation}
1061: Since $\lambda>0$ is fixed, as $n \to +\infty$ the proof is completed.
1062: \end{proof}
1063: %
1064: \begin{corollary} \label{cor:2} 
1065: Let $\bar x$ be a generalized solution on $(a,b)$. 
1066: Suppose that 
1067: \begin{equation}\label{eq:limitazioni}
1068: \limsup_{t\to b^-}I(\bar x(t))<+\infty,\qquad\mbox{and}\qquad
1069: \liminf_{t\to b^-} \dot I(\bar x(t)) <+\infty.
1070: \end{equation} 
1071: Then, if $-\infty<a<\tau<b<+\infty$ there hold
1072: \begin{enumerate}
1073: \item[(i)] $\displaystyle \int_{\tau}^{b} U(t,\bar x(t)) \dt < +\infty$;
1074: \item[(ii)] $\displaystyle \left|\int_{\tau}^{b} h(t) \dt \right| < +\infty$;
1075: \item[(iii)] $\displaystyle \int_{\tau}^{b} K(\dot{\bar x}(t)) \dt < +\infty$;
1076: \item[(iv)] $\displaystyle \|h\|_\infty < +\infty$ on $[\tau,b)$.
1077: \item[(v)] $\displaystyle \lim_{t\to b^-} \bar x(t)$ exists.
1078: \end{enumerate}
1079: \end{corollary}
1080: \begin{proof}
1081: We first prove the assertions in the case of locally minimal solutions.
1082: The boundedness of the first integral follows from the assumption of local
1083: boundedness of the action functional on a locally minimal trajectory,
1084: assumption~\eqref{bddA}, and from Lemma~\ref{le:2}. Concerning the second one, we use
1085: Corollary~\ref{cor:en_abs_cont} and inequality~\eqref{eq:lemma1}; {\it (iii)}
1086: follows straightforwardly from {\it (i)}, {\it (ii)} and the definition of the energy 
1087: $h$.
1088: The boundedness of $\| h\|_\infty$ on $(a,b)$ follows from Corollary
1089: \ref{cor:en_abs_cont} and inequality~\eqref{bddh}.
1090: To deduce {\it (v)} it is sufficient to remark that, from {\it (iii)},
1091: $\bar x$ is  H\"{o}lder-continuous on $(a,b)$.
1092: 
1093: In order to extend the proof to
1094: generalized solutions, we first remark that all the constants and bounds
1095: appearing in the 
1096: proof 
1097: above do not depend on the specific solution $\bar x$, but only on the potential
1098: and the limits in equations~\ref{eq:limitazioni}, 
1099: and the total energy
1100: valued at single instant $h(\tau)$ of the interval. Hence the assertions {\it (i)}, {\it (ii)}
1101: still hold true when passing to pointwise limit such that the energy $h(\tau)$ is bounded. 
1102: The other assertions then follow from the fist two.
1103: \end{proof}
1104: %
1105: \begin{remark}
1106: In Corollary~\ref{cor:2} {\it (v)}, 
1107: Von Zeipel's Theorem is proved, for generalized solutions, under the additional 
1108: assumption~\eqref{eq:liminf}. The proof will be completed in Section~\ref{sec:partial}.
1109: \end{remark}
1110: \begin{remark}
1111: From inequality~\eqref{bddh} we can easily understand that in Definition
1112: \ref{def:generalized_sol} the convergence of the energy of the approximating sequence of
1113: locally minimal solutions can be assumed only at one point.
1114: \end{remark}
1115: 
1116: %=======================================================
1117: \section{Asymptotic estimates at total collisions}\label{sec:asymp}
1118: %=======================================================
1119: 
1120: The purpose of this section is to deepen the analysis of the asymptotics of
1121: generalized solutions as they approach a total collision at the origin; to this aim, 
1122: we introduce some further hypothesis on the potential $U$.
1123: Though we will perform all the analysis in a left neighborhood of the collision
1124: instant, the analysis concerning right neighborhoods is the exact analogue.
1125: 
1126: We recall that $\bar x$ has a {\emph total collision at the origin} at $t=t^*$
1127: if $\lim_{t\to t^*}\bar x(t)=x(t^*)=0$. Since by our assumptions $0$ belongs to the singular set 
1128: $\Delta$ of the potential, assumption~\ref{a:U0} reads that a total collision instant 
1129: is a singularity for $\bar x$.
1130: 
1131: The results proved in \S~\ref{subsec:conlaws} have some relevant consequences
1132: in the case of total collisions at the origin; in particular Corollary~\ref{cor:14} now reads
1133: \begin{corollary}\label{cor:Icont}
1134: Let $\bar x$ be given by Proposition~\ref{propo:globalmin}.
1135: If $|\bar x(t_0)|=0$, then there exists $\delta_0>0$ such that $I(\bar x)$ is
1136: continuous on $I_0=[t_0-\delta_0,t_0+\delta_0]$, it admits weak derivative
1137: almost everywhere, the function $\dot I(\bar x)$ is monotone increasing and
1138: $\dot I(\bar x) \in BV(I_0)$.
1139: Furthermore the following inequalities hold in the distributional sense
1140: \[
1141: \begin{aligned}
1142: \ddot I(\bar x(t)) > 0 & \qquad    \forall t \in \bar I_0 \\
1143: \dot I(\bar x(t)) < 0 &  \qquad  \forall t \in (t_0-\delta_0,t_0)\\
1144: \dot I(\bar x(t)) > 0 &  \qquad  \forall t \in (t_0,t_0+\delta_0).
1145: \end{aligned}
1146: \]
1147: \end{corollary}
1148: 
1149: Furthermore, since $I(\bar x(t))\geq 0$, and 
1150: $I(\bar x(t^*)) = 0$  if and only if $\bar x$ has a total collision at the origin at $t=t^*$, from
1151: Corollary~\ref{cor:14} one can 
1152: deduce that, whenever a total collision occurs at $t=t_0$, 
1153: no other total collisions take place in the interval $I_0$.
1154: Concerning the occurrence of total collision at the boundary of the interval 
1155: $(a,b)$, we argue as in Lemmata~\ref{le:1} and~\ref{le:2} and we use
1156: the convexity of the function $I$ to deduce that also boundary total collisions are isolated.
1157: It is worthwhile noticing that this fact does not prevent, at this stage, the occurrence of 
1158: infinitely many other singularities in a neighborhood of a total collision at the origin.
1159: We summarize these remarks in the next theorem.
1160: \begin{theorem}\label{thm:totcolliso}
1161: Let $\bar x$ be a generalized solution for the dynamical system~\eqref{DS}.
1162: Suppose that $-\infty<a<b<+\infty$ and that there exists $t_0 \in [a,b]$ such
1163: that $|\bar x(t_0)|=0$. Then there exists $\delta>0$ such that, for every
1164: $t\in(t_0-\delta,t_0+\delta)\cap[a,b]$, $t\neq t_0$,
1165: we have $|\bar x(t)|\neq 0$.
1166: \end{theorem}
1167: 
1168: In terms of the radial variable $r$ Corollary~\ref{cor:Icont} and Theorem
1169: \ref{thm:totcolliso} state that whenever $r(t_0)=0$ for some $t_0\in (a,b)$ ($t_0$
1170: can coincide with $a$ or $b$ when finite) then there exists $\delta >0$ such
1171: that
1172: \begin{equation}\label{eq:isol_coll}
1173: \begin{split}
1174: & r(t) \neq 0, \qquad \dot r(t) < 0, \qquad \forall t \in (t_0-\delta,t_0)\\
1175: & r(t) \neq 0, \qquad \dot r(t) > 0, \qquad \forall t \in (t_0,t_0+\delta).
1176: \end{split}
1177: \end{equation}
1178: We moreover rewrite the bounded energy function as
1179: \begin{equation}
1180: \label{eq:energy_rs}
1181: h(t)=\frac{1}{2}\left( \dot r^2 + r^2|\dot s|^2 \right) - U(t,rs).
1182: \end{equation}
1183: Similarly, denoting by 
1184: $(x_\eps)_\eps$ the sequence of global minimizers for $\bar
1185: \action_\eps(\cdot,[t_0-\delta,t_0+\delta])$ converging to the locally minimal
1186: collision solution $\bar x$ whose existence is proved in Proposition
1187: \ref{propo:conv},  we define, for every $\eps$, 
1188: \[
1189: r_\eps := |x_\eps| \in \RR \qquad \text{and} \qquad s := \frac{x_\eps}{|x_\eps|}
1190: \in {\mathcal E}
1191: \]
1192: and we write the energy in~\eqref{eq:energyeps} as 
1193: \[
1194: h_{\eps}(t)=\frac{1}{2}\left( \dot r_\eps^2 + r_\eps^2|\dot s_\eps|^2 \right) -
1195: U_\eps(t,r_\eps s_\eps)+\frac{1}{2}|rs - r_\eps s_\eps|^2.
1196: \]
1197: Furthermore the approximating action functional and the corresponding
1198: Euler--Lagrange equations in the new variables are respectively 
1199: \[
1200: \bar \action_\eps(r_\eps s_\eps,[t_0-\delta,t_0+\delta]) :=
1201: \int_{t_0-\delta}^{t_0+\delta} \frac{1}{2}\left( \dot r_\eps^2 + r_\eps^2|\dot
1202: s_\eps|^2 \right) + U_\eps(t,r_\eps s_\eps)+\frac{1}{2}|rs - r_\eps s_\eps|^2
1203: \dt
1204: \]
1205: and 
1206: \begin{equation}
1207: \label{eq:ELeps}
1208: \begin{split}
1209: -\ddot r_\eps + r_\eps|\dot s_\eps|^2 + \nabla U_\eps(t,r_\eps s_\eps) \cdot
1210: s_\eps -(rs - r_\eps s_\eps)\cdot s_\eps & = 0 \\
1211: -2r_\eps\dot r_\eps \dot s_\eps -r_\eps^2 \ddot s_\eps + r_\eps \nabla_{T}
1212: U_\eps(t,r_\eps s_\eps) -r_\eps(rs - r_\eps s_\eps)& = \mu_\eps s_\eps,
1213: \end{split}
1214: \end{equation}
1215: where $\mu_\eps = r_\eps^2|\dot s_\eps|^2-r_\eps(rs - r_\eps s_\eps)\cdot
1216: s_\eps$ is the Lagrange multiplier due to the presence of the constraint
1217: $|s_\eps|^2 = 1$ and the vector $\nabla_{T} U_\eps(t,r_\eps s_\eps)$ is the
1218: tangent components to the ellipsoid ${\mathcal E}$ of the gradient $\nabla
1219: U_\eps(t,r_\eps s_\eps)$. A similar approximation procedure will be implicitly
1220: done for generalized solutions.
1221: 
1222: \bigskip
1223: 
1224: To proceed with the analysis of the asymptotic behavior near total collisions at the origin we need 
1225: some stronger conditions on the potential $U$ when the radial variable $r$ tends to 0.
1226: These additional conditions includes quasi--homogeneous potential and logarithmic ones, 
1227: in the following analysis, however we will treat separately the two different cases.
1228: 
1229: %----------------------------------------
1230: \subsection{Quasi-homogeneous potentials}
1231: %----------------------------------------
1232: In this section we impose some stronger assumptions on the behavior of the potential
1233: when $|x|$ is small. The following conditions are trivially satisfied by $\alpha$-homogeneous 
1234: potentials and mimic the behavior of combination of such homogeneous potentials:
1235: \begin{assiomi}
1236: \item[\udhtag]
1237: \label{a:U2h}
1238: There exist $\alpha \in (0,2)$, $\gamma>l$ and $C_2 \geq 0$ such that
1239: \[
1240: \nabla U(t,x)\cdot x + \alpha U(t,x) \geq - C_2 |x|^\gamma U(t,x),
1241: \]
1242: whenever $|x|$ is small.
1243: \end{assiomi}
1244: 
1245: % \begin{remark}
1246: % We observe that when $U$ is homogeneous of degree $-\alpha$ in its second
1247: % variable (for instance when $U(t,x)=1/|x|^\alpha$), the equality in condition
1248: % \udh\, is achieved with $C_2 =0$; this assumption is satisfied also when we take
1249: % into account potentials of the form $U(t,x)=U_\alpha(x)+U_\beta(x)$ where
1250: % $U_\alpha$ is homogeneous of degree $-\alpha$, $U_\beta$ is homogeneous of
1251: % degree $-\beta$, $0<\beta < \alpha$, and both  $U_\alpha$ and $U_\beta$ are
1252: % positive. Indeed in this case  condition \udh\, is verified (with the strict
1253: % inequality) with $\gamma = C_2 =0$.
1254: % On the other hand when we consider a negative homogeneous singular perturbation
1255: % $U_\beta$ (we think for instance to the potential $U(t,x)=1/|x|^{\alpha} -
1256: % 1/|x|^{\beta}$, with $0<\beta < \alpha$) we have that \udh\, holds, when $|x|$
1257: % is sufficiently small, with $C_2=\alpha-\beta$ and $0<\gamma < \alpha-\beta$.
1258: % \end{remark}
1259: \begin{remark}
1260: \ref{a:U2h} implies \ref{a:U2} (for small values of $|x|$); in fact, 
1261: by choosing $2>\tilde\alpha>\alpha>0$, one obtains
1262: \[
1263: \begin{split}
1264: \nabla U(t,x)\cdot x + \tilde\alpha U(t,x) & =
1265: \nabla U(t,x)\cdot x + \alpha U(t,x) + (\tilde\alpha-\alpha) U(t,x) \\
1266: & \geq - C_2 |x|^\gamma U(t,x)+(\tilde\alpha-\alpha) U(t,x),
1267: \end{split}
1268: \]
1269: and the last term remains bounded below as $|x|\to0$ since
1270: $\tilde\alpha-\alpha>0$.
1271: \end{remark}
1272: 
1273: Furthermore we suppose the existence of a function $\tilde U$ defined and of
1274: class $\cont^1$ on
1275: $(a,b)\times \left({\mathcal E}\minus \Delta\right)$ such that
1276: \begin{equation}
1277: \label{tildeU}
1278: \inf_{(a,b)\times \left({\mathcal E}\minus \Delta\right)}\tilde U(t,s)  >0
1279: \quad \text{and} \quad \lim_{s\to {\mathcal E}\cap\Delta}\tilde U(t,s)=+\infty \qquad
1280: \mbox{uniformly in } t.
1281: \end{equation}
1282: The potential $U$ is then supposed to verify the following condition uniformly
1283: in the variables $t$ and $s$ (on the compact subsets of $(a,b)\times
1284: \left({\mathcal E}\minus \Delta\right)$):
1285: 
1286: \begin{assiomi}
1287: \item[\uthtag]
1288: \label{a:U3h}
1289: $\dlim_{r \to 0} r^\alpha U(t,x) = \tilde U(t,s)$.
1290: \end{assiomi}
1291: 
1292: \begin{remark}
1293: In \ref{a:U2h} and \ref{a:U3h} the value of $\alpha$ must be the same. We shall refer to potentials satisfying such assumptions as quasi--homogeneous (cf.~\cite{diacuab}).
1294: \end{remark}
1295: 
1296: 
1297: \begin{lemma}
1298: \label{le:integ}
1299: Let $\bar x$ be a generalized solution, let $t_0 \in (a,b]$ be a total
1300: collision instant and let $\delta$ be given in Theorem~\ref{thm:totcolliso}. Then, 
1301: for every $\alpha'\in(\alpha,2)$, we have 
1302: \[
1303: \int_{t_0-\delta}^{t_0}-r^{\alpha'}\frac{\dot r}{r}U(t,rs)\dt < +\infty,
1304: \]
1305: where $\alpha \in (0,2)$ is the constant fixed in assumption \ref{a:U2h}.
1306: \end{lemma}
1307: %
1308: \begin{proof}
1309: We consider the function
1310: \[
1311: \Gamma_{\alpha'} (t) := r^{\alpha'} \left( \dfrac{1}{2}r^2|\dot s|^2 - U(t,rs)
1312: \right), 
1313: \quad \alpha' \in (\alpha,2);
1314: \]
1315: Replacing in~\eqref{eq:energy_rs} we have 
1316: \[
1317: \Gamma_{\alpha'} (t) = h(t)r^{\alpha'} -\dfrac{1}{2}\dot r^2 r^{\alpha'} \leq
1318: h(t)r^{\alpha'};
1319: \]
1320: since $h$ is bounded (see Corollary~\ref{cor:2}, {\em(iv)}) and $r$ tends to $0$, we
1321: conclude that the function $\Gamma_{\alpha'}$ is bounded above on the interval
1322: $[t_0-\delta,t_0]$. 
1323: We consider the corresponding functions (still bounded above) for the
1324: approximating problems:
1325: \[
1326: \begin{split}
1327: \Gamma_{\alpha',\eps} (t) &= r_\eps^{\alpha'} \left( \dfrac{1}{2}r_\eps^2|\dot
1328: s_\eps|^2 - U_\eps(t,r_\eps s_\eps) +\frac{1}{2}|rs-r_\eps s_\eps|^2\right)\\
1329: &= h_\eps(t)r_\eps^{\alpha'} -\dfrac{1}{2}\dot r_\eps^2 r_\eps^{\alpha'} \leq
1330: h_\eps(t)r_\eps^{\alpha'},
1331: \end{split}
1332: \]
1333: and we observe that the sequence $\left(\Gamma_{\alpha',\eps}\right)_\eps$
1334: converges almost everywhere and $L^1$ to $\Gamma_{\alpha'}$, as $\eps \to 0$.
1335: We compute the derivative of $\Gamma_{\alpha',\eps} (t)$ with respect to  time as
1336: \begin{equation}
1337: \begin{split}
1338: \label{eq:dGamma1eps}
1339: &\frac{d}{dt}\Gamma_{\alpha',\eps} (t)= 
1340:  \frac{2+\alpha'}{2}r_\eps^{1+\alpha'}\dot r_\eps |\dot s_\eps|^2\\
1341: &\qquad\qquad\qquad + r_\eps^{2+\alpha'} \dot s_\eps \cdot \ddot s_\eps 
1342: + \alpha'r_\eps^{\alpha'-1}\dot r_\eps\left[\frac12|rs-r_\eps s_\eps|^2 -
1343: U_\eps(t,r_\eps 
1344:   s_\eps)\right]\\
1345: & \;\; - r_\eps^{\alpha'}\left[ \frac{\partial U_\eps}{\partial t}(t,r_\eps
1346: s_\eps) 
1347: + \nabla U_\eps(t,r_\eps s_\eps)(\dot r_\eps s_\eps+r_\eps\dot s_\eps) \right] 
1348: + r_\eps^{\alpha'}(rs-r_\eps s_\eps)\frac{d}{dt}(rs-r_\eps s_\eps).
1349: \end{split}
1350: \end{equation}
1351: Now 
1352: we multiply the Euler--Lagrange equation~\eqref{eq:ELeps}$_2$ by $\dot
1353: s_\eps$ to obtain 
1354: (we recall that $\nabla_T U_\eps(t,r_\eps s_\eps)\cdot \dot s_\eps = \nabla
1355: U_\eps(t,r_\eps s_\eps)\cdot \dot s_\eps$ since $s_\eps$ and $\dot s_\eps$ are orthogonal)
1356: \begin{equation}
1357: \label{eq:EL2s'}
1358: r_\eps^2 \ddot s_\eps\cdot \dot s_\eps = - 2r_\eps\dot r_\eps |\dot s_\eps|^2 +
1359: r_\eps \nabla U_\eps(t,r_\eps s_\eps)\cdot \dot s_\eps - r_\eps r s\cdot s_\eps.
1360: \end{equation}
1361: Replacing~\eqref{eq:EL2s'} in~\eqref{eq:dGamma1eps} we have
1362: \begin{multline}
1363: \label{eq:gamma'3eps}
1364: \frac{d}{dt}\Gamma_{\alpha',\eps}(t) = -\frac{2-\alpha'}{2} r^{1+\alpha'} \dot
1365: r_\eps |\dot s_\eps|^2 - \alpha' r_\eps^{\alpha'-1} \dot r_\eps U_\eps(t,r_\eps
1366: s_\eps) 
1367: - r_\eps^{\alpha'} \frac{\partial U_\eps}{\partial t}(t,r_\eps s_\eps)\\
1368: - r_\eps^{\alpha'-1} \dot r_\eps \nabla U_\eps(t,r_\eps s_\eps)\cdot (r_\eps
1369: s_\eps)
1370: - r^{\alpha'+1}_\eps r s \cdot s_\eps  \\
1371: + \frac{\alpha'}{2}r^{\alpha'-1}_\eps\dot r_\eps|rs-r_\eps s_\eps|^2
1372: + r_\eps^{\alpha'}(rs-r_\eps s_\eps)\frac{d}{dt}(rs-r_\eps s_\eps).
1373: \end{multline}
1374: We now combine assumptions \ref{a:U1}, \ref{a:U2h} and~\eqref{etaeps_ineq} to obtain the
1375: following inequalities
1376: \[
1377: \begin{split}
1378: - r_\eps^{\alpha'} \frac{\partial U_\eps}{\partial t}(t,r_\eps s_\eps) 
1379: & = - r_\eps^{\alpha'}\dot\eta_\eps(U(t,r_\eps s_\eps))\frac{\partial
1380: U}{\partial t}(t,r_\eps s_\eps)\\
1381: & \geq - C_1 
1382: r_\eps^{\alpha'}\dot\eta_\eps(U(t,r_\eps s_\eps))\left(U_\eps(t,r_\eps s_\eps)+1\right)\\
1383: & \geq - C_1r_\eps^{\alpha'}\left(\eta_\eps(U(t,r_\eps s_\eps))+1\right)\\
1384: & \geq -C_1 r_\eps^{\alpha'} \left(U_\eps(t,r_\eps s_\eps)+1\right),
1385: \end{split}
1386: \]
1387: \[
1388: \begin{split}
1389: - r_\eps^{\alpha'} \frac{\dot r_\eps}{r_\eps}\nabla U_\eps(t,r_\eps s_\eps)\cdot
1390: (r_\eps s_\eps)& =
1391: - r_\eps^{\alpha'} \frac{\dot r_\eps}{r_\eps} \dot\eta_\eps(U(t,r_\eps
1392: s_\eps))\nabla U(t,r_\eps s_\eps)\cdot (r_\eps s_\eps) \\
1393: & \geq - r_\eps^{\alpha'} \frac{\dot r_\eps}{r_\eps} \dot\eta_\eps(U(t,r_\eps
1394: s_\eps))U(t,r_\eps s_\eps)\left[-\alpha-C_2r_\eps^\gamma\right] \\
1395: & \geq r_\eps^{\alpha'} \frac{\dot r_\eps}{r_\eps} U_\eps(t,r_\eps
1396: s_\eps)\left[\alpha+C_2r_\eps^\gamma \right].
1397: \end{split}
1398: \]
1399: Finally, by replacing in~\eqref{eq:gamma'3eps}, we obtain 
1400: \[
1401: \frac{d}{dt}\Gamma_{\alpha',\eps}(t) \geq \Psi_{\alpha',\eps}(t)
1402: \]
1403: where
1404: \[
1405: \begin{split}
1406: \Psi_{\alpha',\eps}(t) &= 
1407: -\frac{2-\alpha'}{2} r_\eps^{1+\alpha'} \dot r_\eps |\dot s_\eps|^2
1408: -(\alpha'-\alpha) r_\eps^{\alpha'} \frac{\dot r_\eps}{r_\eps} U_\eps(t,r_\eps
1409: s_\eps)
1410: - C_1 r_\eps^{\alpha'} \left(U_\eps(t,r_\eps s_\eps)+1\right) \\
1411: & + C_2 r_\eps^{\alpha'+\gamma} \frac{\dot r_\eps}{r_\eps} U_\eps(t,r_\eps
1412: s_\eps)
1413: - r^{\alpha'+1}_\eps r s \cdot s_\eps\\
1414: & + \frac{\alpha'}{2}r^{\alpha'-1}_\eps\dot r_\eps|rs-r_\eps s_\eps|^2
1415: + r_\eps^{\alpha'}(rs-r_\eps s_\eps)\frac{d}{dt}(rs-r_\eps s_\eps).
1416: \end{split}
1417: \]
1418: Since $\gamma>0$ and $r_\eps \to 0$ as $t \to t_0$, for every $\eps >0$, there
1419: exists a positive $\lambda_\eps\leq (\alpha'-\alpha)/2$ such that
1420: \[
1421: C_2r_\eps^{\alpha'+\gamma} U_\eps(t,r_\eps s_\eps) \leq 
1422: \lambda_\eps r_\eps^{\alpha'} U_\eps(t,r_\eps s_\eps)
1423: \]
1424: whenever $\delta$ is small enough; furthermore, since $- \frac{2-\alpha'}{2}
1425: r_\eps^{1+\alpha'} \dot r_\eps |\dot s_\eps|^2$ is positive, we have
1426: \[
1427: \begin{split}
1428: \Psi_{\alpha',\eps}(t) & \geq
1429: - (\alpha'-\alpha-\lambda_\eps) r_\eps^{\alpha'} \frac{\dot r_\eps}{r_\eps}
1430: U_\eps(t,r_\eps s_\eps) 
1431: - C_1 r_\eps^{\alpha'} \left(U_\eps(t,r_\eps s_\eps)+1\right) \\
1432: &- r^{\alpha'+1}_\eps r s \cdot s_\eps
1433: + \frac{\alpha'}{2}r^{\alpha'-1}_\eps\dot r_\eps|rs-r_\eps s_\eps|^2
1434: + r_\eps^{\alpha'}(rs-r_\eps s_\eps)\frac{d}{dt}(rs-r_\eps s_\eps).
1435: \end{split}
1436: \]
1437: Therefore, for every $\eps$, the function $\Psi_{\alpha',\eps}$ is larger than 
1438: the sum of a positive term 
1439: \[
1440: - (\alpha'-\alpha-\lambda_\eps) r_\eps^{\alpha'}
1441: \frac{\dot r_\eps}{r_\eps} U_\eps(t,r_\eps s_\eps),
1442: \]
1443: an integrable term 
1444: \[
1445: - C_1 r_\eps^{\alpha'} \left(U_\eps(t,r_\eps s_\eps)+1\right)
1446: \]
1447: and
1448: a remainder
1449: \[
1450: - r^{\alpha'+1}_\eps r s \cdot s_\eps
1451: + \frac{\alpha'}{2}r^{\alpha'-1}_\eps\dot r_\eps|rs-r_\eps s_\eps|^2
1452: + r_\eps^{\alpha'}(rs-r_\eps s_\eps)\frac{d}{dt}(rs-r_\eps s_\eps)
1453: \]
1454: converging uniformly to $- r^{\alpha'+2}$ as $\eps$ tends to 0.
1455: 
1456: We can then conclude that, for every $\eps$
1457: \begin{multline*}
1458: \Gamma_{\alpha',\eps}(t_0) - \Gamma_{\alpha',\eps}(t_0-\delta) \geq
1459: -\int_{t_0-\delta}^{t_0}(\alpha'-\alpha-\lambda_\eps) r_\eps^{\alpha'}
1460: \frac{\dot r_\eps}{r_\eps} U_\eps(t,r_\eps s_\eps) \dt\\
1461: - C_1 \int_{t_0-\delta}^{t_0}r_\eps^{\alpha'} \left(U_\eps(t,r_\eps s_\eps)+1\right)dt\\
1462: + \int_{t_0-\delta}^{t_0}\left[-r^{\alpha'+1}_\eps r s \cdot s_\eps  
1463: + \frac{\alpha'}{2}r^{\alpha'-1}_\eps\dot r_\eps|rs-r_\eps s_\eps| 
1464: + r_\eps^{\alpha'}(rs-r_\eps s_\eps)\frac{d}{dt}(rs-r_\eps s_\eps)\right]\dt.
1465: \end{multline*}
1466: The right hand side of the last inequality is bounded above 
1467: because of the boundedness of 
1468: $\Gamma_{\alpha',\eps}$.
1469: Passing to the limit as $\eps \to 0$, from Proposition~\ref{propo:conv} and the
1470: boundedness above of the function $\Gamma_{\alpha'}$ it follows that
1471: \begin{multline*}
1472: \Gamma_{\alpha'}(t_0) - \Gamma_{\alpha'}(t_0-\delta) \geq
1473: -\int_{t_0-\delta}^{t_0}(\alpha'-\alpha-\lambda)r^{\alpha'}\frac{\dot r}{r}
1474: U(t,rs)\dt\\
1475: - C_1 \int_{t_0-\delta}^{t_0} r^{\alpha'} \left(U(t,rs)+1\right)\dt
1476: -\int_{t_0-\delta}^{t_0}r^{\alpha'+2}\dt,
1477: \end{multline*}
1478: where $\lambda\leq (\alpha'-\alpha)/2$ is the limit, up to subsequences, of the
1479: bounded sequence $(\lambda_\eps)_\eps$.
1480: Now we recall that, by Lemma~\ref{le:2}, $U(t,rs)$ is integrable; this fact implies the
1481:  integrability of the function
1482: $r^{\alpha'} U(t,rs)-r^{\alpha'+2}$ and hence the existence of a constant $K$ such that
1483: \[
1484: \int_{t_0-\delta}^{t_0} - r^{\alpha'} \frac{\dot r}{r} U(t,rs) dt \leq K <
1485: +\infty.
1486: \]
1487: \end{proof}
1488: %
1489: \begin{lemma}[{\bf Monotonicity Formula}]
1490: \label{le:Gamma}
1491: Let $\bar x$ be a generalized solution, let $t_0 \in (a,b]$ be a total
1492: collision instant and let $\delta>0$ be the constant 
1493: obtained in Theorem~\ref{thm:totcolliso}. Then
1494: the function 
1495: \begin{equation}
1496: \label{eq:function_Gamma}
1497: \Gamma_\alpha (t) := r^\alpha \left[ \frac12 r^2|\dot s|^2 - U(t,rs) \right]
1498: \end{equation}
1499: is bounded on $[t_0-\delta,t_0)$ and
1500: \begin{equation}\label{eq:ineq:dGamma}
1501: \begin{split}
1502: \Gamma_\alpha (t)  \geq \Gamma_\alpha(t_0-\delta) 
1503: &- \int_{t_0-\delta}^{t}\frac{2-\alpha}{2} r^{1+\alpha} \dot r |\dot s|^2 \,d\xi\\
1504: &- C_1 \int_{t_0-\delta}^{t} r^\alpha \left(U(\xi,rs)+1\right)\,d\xi
1505:  + C_2 \int_{t_0-\delta}^{t} r^{\alpha+\gamma} \frac{\dot r}{r} U(\xi,rs) \,d\xi
1506: \end{split}
1507: \end{equation}
1508: where $t\in[t_0-\delta,t_0]$.
1509: \end{lemma}
1510: \begin{proof}
1511: Replacing in~\eqref{eq:energy_rs} the expression of the function $\Gamma_\alpha$
1512: we have 
1513: \[
1514: \Gamma_\alpha (t) = h(t)r^\alpha - \dot r^2r^\alpha \leq h(t)r^\alpha;
1515: \]
1516: since $h$ is bounded (see Corollary~\ref{cor:2}) and $r$ tends to $0$, we
1517: conclude that the function $\Gamma_\alpha$ is bounded above.
1518: Using the same approximation arguments described in Lemma~\ref{le:integ}, we
1519: obtain~\eqref{eq:ineq:dGamma}.
1520: From Lemma~\ref{le:integ} we deduce the integrability of the negative function 
1521: $r^{\alpha+\gamma} \frac{\dot r}{r} U(t,rs)$.
1522: Hence, since $-\frac{2-\alpha}{2} r^{1+\alpha} \dot r |\dot s|^2$
1523: is positive and both $r^\alpha U(t,rs)$ and $r^{\alpha+\gamma} \frac{\dot r}{r}
1524: U(t,rs)$ are integrable (Lemma~\ref{le:integ}), the boundedness below of
1525: the function $\Gamma_\alpha$ follows from~\eqref{eq:ineq:dGamma}.
1526: \end{proof}
1527: 
1528: \begin{corollary}
1529: \label{cor:lim2}
1530: In the same setting of Lemma~\ref{le:Gamma} we have 
1531: $\displaystyle \int_{t_0-\delta}^{t_0}-r^{1+\alpha}\dot r |\dot s|^2 < +\infty$.
1532: \end{corollary}
1533: \begin{proof}
1534: It follows from the boundedness above of the function
1535: $\Gamma_\alpha$ 
1536: and inequality~\eqref{eq:ineq:dGamma} since the terms 
1537: $- C_1 \int_{t_0-\delta}^{t} r^\alpha \left(U(t,rs)+1\right) dt$ and
1538: $ C_2 \int_{t_0-\delta}^{t} r^{\alpha+\gamma} \frac{\dot r}{r} U(t,rs)\,dt$ 
1539: are
1540: negative.
1541: \end{proof}
1542: 
1543: \begin{lemma} \label{le:lim1}
1544: Let $\varphi(t):= -\dot r(t) r^{\alpha/2}(t)$, $t \in [t_0-\delta,t_0]$. 
1545: Then there exist two constants depending on $\alpha$, $c_{1,\alpha} \leq
1546: c_{2,\alpha}$, such that
1547: for all  $t \in [t_0-\delta,t_0]$
1548: \[
1549: c_{1,\alpha} \leq \varphi(t) \leq c_{2,\alpha}. 
1550: \]
1551: \end{lemma}
1552: \begin{proof}
1553: Since the energy function $h$ is bounded (see Corollary~\ref{cor:2}, {\em(iv)}) 
1554: and we assume that $r$ tends to $0$ as $t$ tends to $t_0$,
1555: the function $r^\alpha h(t)=\dfrac{1}{2}\varphi^2(t) + \Gamma_\alpha(t)$ 
1556: is also bounded and by Lemma~\ref{le:Gamma} we can deduce that 
1557: $\varphi(t)=\sqrt{2\left[r^\alpha h(t) - \Gamma_\alpha(t) \right]}$ is bounded
1558: below and above by a pair of constants $0\leq c_{1,\alpha} \leq c_{2,\alpha}$ on
1559: the interval $[t_0-\delta,t_0]$.
1560: \end{proof}
1561: 
1562: \begin{corollary}
1563: \label{cor:int2}
1564: In the same setting of Lemma~\ref{le:Gamma} we have 
1565: $\displaystyle \lim_{t \to t_0}\int_{t_0-\delta}^{t} \frac{1}{r^{\alpha/2
1566: +1}} = +\infty$.
1567: \end{corollary}
1568: \begin{proof}
1569: We can write the boundedness above of the function $\varphi$ (proved in
1570: Lemma~\ref{le:lim1}) as 
1571: \begin{equation}
1572: \label{eq:in}
1573: -\frac{\dot r}{r} \leq \frac{c_{2,\alpha}}{r^{\alpha/2 +1}}, \quad t \in
1574: [t_0-\delta,t_0].
1575: \end{equation}
1576: Integrating inequality~\eqref{eq:in} on the interval $[t_0-\delta, t]$, when $t
1577: \to t_0$, we obtain 
1578: \[
1579: \lim_{t \to t_0} c_{2,\alpha}\int_{t_0-\delta}^{t}\frac{d\xi}{r^{\alpha/2 +1}} 
1580: \geq \lim_{t \to t_0} \int_{t_0-\delta}^{t} -\frac{\dot r}{r}d\xi
1581: = \log r(t_0-\delta)-\lim_{t \to t_0} \log r(t) = +\infty
1582: \]
1583: since $r$ tends to 0 as $t \to t_0$.
1584: \end{proof}
1585: 
1586: \begin{lemma}
1587: \label{le:lim2}
1588: The lower bound $c_{1,\alpha}$ 
1589: of the function $\varphi$ defined in Lemma~\ref{le:lim1} can be chosen
1590: strictly positive, that is $c_{1,\alpha} > 0$.
1591: \end{lemma}
1592: \begin{proof}
1593: We start proving an estimate above of the derivative of the function
1594: $\varphi$. With this purpose we consider the approximating sequence
1595: $(\varphi_\eps)_\eps$ where 
1596: \[
1597: \varphi_\eps (t)=-\dot r_\eps(t) r_\eps^{\alpha/2}(t)
1598: \]
1599: and, for every $\eps>0$, we compute the first derivative of the smooth function
1600: $\varphi_\eps$ and we use the Euler--Lagrange equation~\eqref{eq:ELeps}$_1$ 
1601: for the approximating problem to obtain
1602: \begin{equation*}
1603: \begin{split}
1604: \dot\varphi_\eps(t)
1605: & =-\frac{\alpha}{2}r_\eps^{\alpha/2-1}\dot r_\eps^2-r_\eps^{\alpha/2}\ddot
1606: r_\eps \\
1607: & = -\frac{\alpha}{2} r_\eps^{\alpha/2 -1}\dot r_\eps^2 -
1608: r_\eps^{\alpha/2+1}|\dot s_\eps|^2   -r_\eps^{\alpha/2-1} \nabla U_\eps(t,r_\eps
1609: s_\eps)\cdot (r_\eps s_\eps)+r_\eps^{\alpha/2}(rs-r_\eps s_\eps)\cdot s_\eps.
1610: \end{split}
1611: \end{equation*}
1612: Arguing as in the proof of Lemma~\ref{le:integ} we use assumptions \ref{a:U2h}
1613: and~\eqref{etaeps_ineq} to deduce
1614: \begin{equation*}
1615: \begin{split}
1616: \dot\varphi_\eps(t)
1617: & \leq r_\eps^{\alpha/2-1}\left[-\frac{\alpha}{2}\dot r_\eps^2-r_\eps^2|\dot
1618: s_\eps|^2  +(\alpha + C_2r_\eps^\gamma) U_\eps(t,r_\eps s_\eps)+(rs-r_\eps
1619: s_\eps)\cdot (r_\eps s_\eps) \right] \\
1620: & = \frac{1}{r_\eps^{\alpha/2 +1}}\Big[\frac{2-\alpha}{2}\varphi_\eps^2(t) -
1621: 2r_\eps^\alpha h_\eps(t)-(2-\alpha)r_\eps^{\alpha}U_\eps(t,r_\eps s_\eps) +
1622: C_2r_\eps^{\alpha+\gamma} U_\eps(t,r_\eps s_\eps) \\
1623: & \hspace{3cm} + r_\eps^\alpha(rs-r_\eps s_\eps)\cdot (rs) \Big]
1624: \end{split}
1625: \end{equation*}
1626: and then, for every $ \in (t_0-\delta,t_0)$,
1627: \begin{equation*}
1628: \begin{split}
1629: \varphi_\eps(t)
1630: & \leq \varphi_\eps^0 +\int_{t_0-\delta}^{t}
1631: \frac{1}{r_\eps^{\alpha/2 +1}}\Big[\frac{2-\alpha}{2}\varphi_\eps^2(\xi) -
1632: 2r_\eps^\alpha h_\eps(\xi)-(2-\alpha)r_\eps^{\alpha}U_\eps(\xi,r_\eps s_\eps)\\
1633: & \hspace{3cm} + C_2r_\eps^{\alpha+\gamma} U_\eps(\xi,r_\eps s_\eps) 
1634:                + r_\eps^\alpha(rs-r_\eps s_\eps)\cdot (rs) \Big]d\xi
1635: \end{split}
1636: \end{equation*}
1637: where $\varphi_\eps^0=\varphi_\eps(t_0-\delta)$.
1638: As $\eps\to 0$, from Proposition~\ref{propo:conv} we have 
1639: \begin{multline*}
1640: \varphi(t) \leq \varphi^0 \\+\int_{t_0-\delta}^{t}
1641: \frac{1}{r^{\alpha/2 +1}}\left[\frac{2-\alpha}{2}\varphi^2(\xi) - 2r^\alpha
1642: h_\eps(\xi)-(2-\alpha)r^{\alpha}U(\xi,rs) + C_2 r^{\alpha+\gamma} U(\xi,rs)
1643: \right]d\xi,
1644: \end{multline*}
1645: where $\varphi^0=\varphi(t_0-\delta)$.
1646: Since $\gamma>0$ there exists $\lambda \in (0,2-\alpha)$, such that
1647: \begin{equation*}
1648: \varphi(t) \leq \varphi^0 +\int_{t_0-\delta}^{t}
1649: \frac{1}{r^{\alpha/2 +1}}\left[\frac{2-\alpha}{2}\varphi^2(\xi) + C(\xi)
1650: -(2-\alpha-\lambda)r^{\alpha}U(\xi,rs) \right]d\xi,
1651: \end{equation*}
1652: where $C(t)$ is such that $|C(t)| \to 0$ as $t \to t_0$ and $2r^\alpha h(t) \leq
1653: C(t)$ on $[t_0-\delta,t_0]$. 
1654: Furthermore, the uniform convergence assumed in condition \ref{a:U3h} implies that,
1655: denoting by
1656: $\tilde U_0$ the minimal value assumed by $\tilde U$ on the ellipsoid
1657: ${\mathcal E}$, there exist two positive constants $k_1,k_2 >0$ such that
1658: \begin{equation*}
1659: \varphi(t) \leq \varphi^0 +\int_{t_0-\delta}^{t}\frac{k_1}{r^{\alpha/2
1660: +1}}\left( \varphi^2(\xi) - k_2 \tilde U_0 \right)d\xi
1661: \end{equation*}
1662: whenever $\delta$ is sufficiently small.
1663: We will conclude showing that necessarily $\varphi^2(t) \geq k_2 \tilde U_0$ and
1664: then choosing $c_{1,\alpha}:= \sqrt{k_2 \tilde U_0} > 0$.
1665: 
1666: By the sake of contradiction we suppose the existence of $\hat t$ such that 
1667: $\varphi^2(\hat t) < k_2 \tilde U_0$; 
1668: then $\varphi^2 - k_2 \tilde U_0 < 0$ in a neighborhood of $\hat t$ and 
1669: \begin{equation*}
1670: \varphi(t) \leq \varphi(\hat t) +\int_{\hat t}^{t}\frac{k_1}{r^{\alpha/2
1671: +1}}\left( \varphi^2(\xi) - k_2 \tilde U_0 \right)d\xi < \varphi(\hat t)
1672: \end{equation*}
1673: for every $t \in (\hat t,t_0)$.
1674: We deduce the existence of a strictly positive constant $\hat k$ such that, 
1675: for every $t \in (\hat t,t_0)$,
1676: \begin{equation*}
1677: \varphi(t)-\varphi(\hat t) \leq - \hat k \int_{\hat
1678: t}^{t}\frac{d\xi}{r^{\alpha/2 +1}}
1679: \end{equation*}
1680: Since the right hand side tends to $-\infty$ as $t$ approaches $t_0$ (see
1681: Corollary~\ref{cor:int2}), the last inequality contradicts the boundedness of
1682: the function $\varphi$.
1683: \end{proof}
1684: 
1685: \begin{corollary}
1686: \label{cor:stime_r}
1687: There exist two strictly positive constants $0<k_{1,\alpha} \leq k_{2,\alpha}$
1688: such that 
1689: \[
1690: k_{1,\alpha}(t_0-t)^{\frac{2}{\alpha+2}} \leq r(t) \leq
1691: k_{2,\alpha}(t_0-t)^{\frac{2}{\alpha+2}},
1692: \]
1693: whenever $t \in [t_0-\delta_0,t_0]$.
1694: \end{corollary}
1695: \begin{proof}
1696: The statement follows from Lemmata~\ref{le:lim1} and~\ref{le:lim2} with 
1697: $k_{i,\alpha} := \left(\frac{\alpha+2}{2}c_{i,\alpha}
1698: \right)^{\frac{2}{\alpha+2}}$, $i=1,2.$
1699: \end{proof}
1700: 
1701: \begin{corollary}
1702: \label{cor:limGamma}
1703: There exists $b>0$ such that 
1704: \[
1705: \lim_{t \to t_0^-} \Gamma_\alpha(t) = -b \qquad \mbox{and}\qquad
1706: \lim_{t \to t_0^-} \dot r^2 r^\alpha = 2b.
1707: \]
1708: \end{corollary}
1709: \begin{proof}
1710: Since $\Gamma_\alpha$ is bounded and inequality~\eqref{eq:ineq:dGamma} holds,
1711: $\Gamma_\alpha$ admits a limit when $t$ tends to $t_0$ from the right.
1712: We call this limit $-b \in \RR$.
1713: Since $\Gamma_\alpha (r,s) = h(t)r^\alpha - \dfrac{1}{2}\dot r^2 r^\alpha$, the
1714: energy $h$ is bounded  and $r$ tends to $0$ as $t \to t_0$ we conclude that $\dot r^2 r^\alpha$ 
1715: converges to $2b$ and, using Lemma~\ref{le:lim2} we
1716: deduce that $b>0$.
1717: \end{proof}
1718: 
1719: \begin{theorem}
1720: \label{central_conf}
1721: Let $\bar x$ be a generalized solution for the dynamical system~\eqref{DS},
1722: let $t_0 \in (a,b)$ (if $b<+\infty$ $t_0$ can coincide with $b$) be a total
1723: collision instant and let $\delta>0$ be the constant
1724: obtained in Theorem~\ref{thm:totcolliso}. 
1725: Let $r,s$ be the new variables defined in~(\ref{rs}); if the potential $U$
1726: satisfies assumptions~\ref{a:U0}, \ref{a:U1}, \ref{a:U2h}, \ref{a:U3h} then the following
1727: assertions hold
1728: \begin{itemize}
1729: \item[(a)] $\dlim_{t \rightarrow {t_0^-}} r^\alpha U(t,rs)=b$, where $b$ is
1730: the strictly positive constant introduced in Corollary~\ref{cor:limGamma};
1731: \item[(b)] there is a positive constant $K$ 
1732: such that, as $t$ tends to $t_0$, 
1733: \[
1734: \begin{aligned}
1735: r(t) & \sim [K
1736: ({t_0}-t)]^{\frac{2}{2+\alpha}}  \\ 
1737: \dot r(t) & \sim -\frac{2K}{2+\alpha}
1738: [K({t_0}-t)] ^{\frac{-\alpha}{2+\alpha}};
1739: \end{aligned}
1740: \]
1741: \item[(c)] $\ds \lim_{t \rightarrow t_0^-} |\dot s(t)|({t_0}-t)=0$;
1742: \item[(d)] for every real positive sequence $(\lambda_n)_n$, such that
1743:            $\lambda_n \rightarrow 0$ as $n \rightarrow +\infty$ we have   
1744:            \[
1745:            \lim_{n \rightarrow +\infty} |s({t_0}-\lambda_n) - s({t_0}-\lambda_n
1746: t)|=0, \quad \forall t > 0.
1747:            \]
1748: \end{itemize}
1749: \end{theorem}
1750: 
1751: \begin{remark}
1752: Condition 
1753: {\it (a)} of Theorem~\ref{central_conf} together with assumptions \ref{a:U3h} on $U$ 
1754: and~\eqref{tildeU} on $\tilde U$ imply that, if $\bar x$ is generalized solution
1755: and $|\bar x(t_0)|=0$, then there exist $\delta>0$ such that, for every $t\in(t_0-\delta,t_0)$,
1756: $\bar x(t)\notin\Delta$, i.e., in a (left) neighborhood 
1757: of the total collision instant no other collision is allowed: neither total nor partial. 
1758: As a consequence, in such a neighborhood, the generalized solution $\bar x$ satisfies the
1759: dynamical system~\eqref{DS} and the corresponding variables $(r,s)$ verify the Euler--Lagrange
1760: equations~\eqref{eq:EL}.
1761: \end{remark}
1762: 
1763: \begin{proof}[Proof of Theorem \ref{central_conf}]
1764: We begin by 
1765: proving statement \emph{(a)}.
1766: The boundedness of the function $\Gamma_\alpha$ together with
1767: inequality~\eqref{eq:ineq:dGamma} imply the integrability of the function
1768: $r^{\alpha+1}\dot r |\dot s|^2$ on the interval $[t_0-\delta,t_0]$ (see
1769: Corollary~\ref{cor:lim2}). Furthermore, since the integral of $\dot r/r$ on the
1770: same interval diverges to $-\infty$ we conclude that
1771: \[
1772: \liminf_{t \to t_0^-} r^{\alpha+2}|\dot s|^2 = 0
1773: \]
1774: and from~\eqref{eq:function_Gamma} together with Corollary~\ref{cor:limGamma}
1775: \begin{equation}
1776: \label{eq:liminfU}
1777: \liminf_{t \to t_0^-} r^{\alpha} U(t,rs) = b.
1778: \end{equation}
1779: It remains to prove that also $\ds \limsup_{t \to t_0^-} r^{\alpha} U(t,rs) =
1780: b$.
1781: Suppose, 
1782: for the sake of contradiction, the existence of a strictly positive
1783: $\eps$
1784: such that 
1785: \begin{equation}
1786: \label{eq:limsupU}
1787: \limsup_{t \to t_0^-} r^{\alpha} U(t,rs) = b + 3\eps.
1788: \end{equation}
1789: Using assumption \ref{a:U3h} we have that~\eqref{eq:liminfU} and~\eqref{eq:limsupU}
1790: are respectively equivalent to 
1791: \[
1792: \liminf_{t \to t_0^-} \tilde U(t,s) = b \quad \mbox{ and } \quad 
1793: \limsup_{t \to t_0^-} \tilde U(t,s) = b + 3\eps
1794: \]
1795: and Corollary~\ref{cor:limGamma} implies the existence of $t_{\eps}$ such that
1796: $\Gamma_\alpha(t) \geq -b-\eps/2$ whenever $t \in (t_{\eps} , t_0]$. We can then
1797: define the set 
1798: \[
1799: {\cal U} := \left \{ t \in (t_{\eps} , t_0) : \tilde U(t,s(t))\geq b+\eps
1800: \right\}.
1801: \]
1802: We define two non-empty subsets of the ellipsoid ${\mathcal E}$ as 
1803: \[
1804: A:= \left\{ s(t) : \tilde U(t,s(t)) \leq b+\eps \right\} \quad \mbox{ and }
1805: \quad
1806: B:= \left\{ s(t) : \tilde U(t,s(t)) \geq b+2\eps \right\};
1807: \]
1808: since $\eps > 0$ the quantity 
1809: \[
1810: d:=\mathrm{dist}(A,B)=\inf_{s_1 \in A, s_2 \in B} |s_1-s_2|
1811: \]
1812: is strictly positive and there exists a sequence $(t_n)_{n\geq 0} \subset
1813: [t_0-\delta,t_0]$, such that
1814: \[
1815: \begin{split}
1816: & t_n \to t_0 \mbox{ as } n \to +\infty  \\
1817: & s(t_{2k}) \in \partial A \quad \mbox{and} \quad s(t_{2k+1}) \in \partial B
1818: \quad \mbox{ for every } k \in \NN \\
1819: & b+\eps \leq \tilde U(t,s(t)) \leq b+2\eps, \quad \mbox{for every } t \in
1820: (t_{2k},t_{2k+1}) \mbox{ and } k \in \NN.
1821: \end{split}
1822: \]
1823: Hence $(t_{2k},t_{2k+1}) \subset {\cal U}$, for every $k$, and
1824: from the definition of the function $\Gamma_\alpha$ in~\eqref{eq:function_Gamma}
1825: we have that
1826: \begin{equation}
1827: \label{eq:ass}
1828: r^{\alpha+2}|\dot s|^2 \geq \eps \mbox{ in the intervals } (t_{2k},t_{2k+1}). 
1829: \end{equation}
1830: We now estimate the integral on $(t_{2k},t_{2k+1})$ of the integrable (on
1831: $[t_0-\delta,t_0]$) function $r^{\alpha+1}\dot r|\dot s|^2$ using~\eqref{eq:ass}
1832: and Corollary~\ref{cor:stime_r}
1833: \begin{equation}
1834: \label{eq:est1}
1835: \begin{split}
1836: \int_{t_{2k}}^{t_{2k+1}} -\frac{\dot r}{r}r^{\alpha+2}|\dot s|^2 dt 
1837: & \geq \eps \int_{t_{2k}}^{t_{2k+1}} - \frac{\dot r}{r} dt 
1838: = \eps\log \frac{r(t_{2k})}{r(t_{2k+1})} \\
1839: &\geq \frac{2\eps}{2+\alpha}\log
1840: \frac{c_{1,\alpha}(t_0-t_{2k})}{c_{2,\alpha}(t_0-t_{2k+1})}.
1841: \end{split}
1842: \end{equation}
1843: On the other hand, using H\"older inequality, we have 
1844: \begin{equation}
1845: \label{dis:1}
1846: d^2 \leq |s(t_{2k+1})-s(t_{2k})| \leq \left( \int_{t_{2k}}^{t_{2k+1}} |\dot s| dt \right)^2 \leq 
1847:          \int_{t_{2k}}^{t_{2k+1}} -r^{\alpha+2}\frac{\dot r}{r}|\dot s|^2 dt 
1848:          \int_{t_{2k}}^{t_{2k+1}} \frac{dt}{-r^{\alpha+1}\dot r} 
1849: \end{equation}
1850: and from Lemma~\ref{le:lim1} and Corollary~\ref{cor:stime_r}, we obtain
1851: \begin{equation}
1852: \label{dis:2}
1853: \begin{split}
1854: \int_{t_{2k}}^{t_{2k+1}} \frac{dt}{-r^{\alpha+1}\dot r} & =
1855: \int_{t_{2k}}^{t_{2k+1}} \frac{1}{-r^{\alpha/2}\dot r}\frac{1}{r^{\alpha/2+1}}dt
1856: \\ 
1857: & \leq \frac{2}{2+\alpha}\frac{1}{c^2_{1,\alpha}} \int_{t_{2k}}^{t_{2k+1}}
1858: \frac{dt}{t_0-t} =
1859: \frac{2}{2+\alpha}\frac{1}{c^2_{1,\alpha}} \log \frac{t_0-t_{2k}}{t_0-t_{2k+1}}.
1860: \end{split}
1861: \end{equation}
1862: Combining~\eqref{dis:1} and~\eqref{dis:2} we obtain
1863: \begin{equation}
1864: \label{eq:est2}
1865: \int_{t_{2k}}^{t_{2k+1}} -r^{\alpha+2}\frac{\dot r}{r}|\dot s|^2 dt \geq 
1866: \frac{2+\alpha}{2}d^2 c_{1,\alpha}^2 \left[ \log
1867: \frac{t_0-t_{2k}}{t_0-t_{2k+1}}\right]^{-1}.
1868: \end{equation}
1869: From the estimates~\eqref{eq:est1} and~\eqref{eq:est2} we deduce
1870: \[
1871: \int_{t_{2k}}^{t_{2k+1}} - r^{\alpha+2}\frac{\dot r}{r}|\dot s|^2 dt \geq 
1872: \frac{\eps}{2+\alpha}\log
1873: \frac{c_{1,\alpha}(t_0-t_{2k})}{c_{2,\alpha}(t_0-t_{2k+1})} +
1874: \frac{2+\alpha}{4}d^2 c_{1,\alpha}^2 \left[ \log
1875: \frac{t_0-t_{2k}}{t_0-t_{2k+1}}\right]^{-1}.
1876: \]
1877: Summing on the index $k$ and recalling that the positive function $-\dot r
1878: r^{\alpha+1}|\dot s|^2$ has a finite integral on $[t_0-\delta,t_0]$ (Corollary
1879: \ref{cor:lim2}) we have
1880: \begin{equation}
1881: \label{mult:rhs}
1882: \begin{split}
1883: +\infty & > \int_{t_0-\delta}^{t_0} -\dot r r^{\alpha+1}|\dot s|^2 dt 
1884: > \sum_{k \geq 0} \int_{t_{2k}}^{t_{2k+1}} -\dot r r^{\alpha+1}|\dot s|^2 dt \\
1885: & \geq \frac{\eps}{2+\alpha}\sum_{k \geq 0}\log
1886: \frac{c_{1,\alpha}(t_0-t_{2k})}{c_{2,\alpha}(t_0-t_{2k+1})} +
1887: \frac{2+\alpha}{4}d^2 c_{1,\alpha}^2 \sum_{k \geq 0}\left[ \log
1888: \frac{t_0-t_{2k}}{t_0-t_{2k+1}}\right]^{-1}.
1889: \end{split}
1890: \end{equation}
1891: Since $c_{2,\alpha}/c_{1,\alpha}$ is 
1892: bounded  (see Lemma~\ref{le:lim2}),
1893: for the last term in~\eqref{mult:rhs} to be  finite it 
1894: is necessary that 
1895: \begin{equation}
1896: \label{eq:limits}
1897: \lim_{k \to+\infty}\frac{t_0-t_{2k}}{t_0-t_{2k+1}}=\frac{c_{2,\alpha}}{c_{1,\alpha}}
1898: \quad \mbox{and} \quad
1899: \lim_{k \to +\infty}\frac{t_0-t_{2k}}{t_0-t_{2k+1}}=+\infty.
1900: \end{equation}
1901: This is a contradiction, hence we conclude that 
1902: \[
1903: \limsup_{t \to t_0} r^\alpha U(t,rs)=b
1904: \]
1905: and, after replacing the value in~\eqref{eq:function_Gamma},
1906: \begin{equation}
1907: \label{eq:limspunto}
1908: \lim_{t \to t_0} r^{\alpha+2}|\dot s|^2 =0.
1909: \end{equation}
1910: 
1911: \noindent To prove \emph{(b)},
1912: from Corollary~\ref{cor:limGamma} we obtain
1913: \[
1914: \lim_{t \to t_0^-}\frac{r(t)^{\alpha/2 + 1}}{(\alpha/2 + 1)(t_0-t)} =
1915: \lim_{t \to t_0^-}-r(t)^{\alpha/2}\dot r(t) = \sqrt{2b};
1916: \]
1917: we then conclude by 
1918: defining $\displaystyle K:=\frac{2+\alpha}{2}\sqrt{2b}$.
1919: The second estimate follows directly.
1920: % of Theorem \ref{central_conf}
1921: % directly follows from \emph{(b)}.
1922: 
1923: \noindent Part \emph{(c)}
1924: directly follows from~\eqref{eq:limspunto} and \emph{(b)}.
1925: 
1926: \noindent We conclude by
1927: proving  statement \emph{(d)}.
1928: If $t=1$ there is nothing to prove. Suppose $t>0$, $t\neq 1$, and
1929: consider a sequence $(\lambda_n)_n$, $\lambda_n \rightarrow 0$;
1930: let $N$ be such that $\lambda_n < \delta/\max(1,t)$, $\forall n \geq N.$
1931: Whenever $t>1$, for every $n\geq N$, we have
1932: \[
1933: t_0-\delta < t_0 -\lambda_nt < t_0 -\lambda_n < t_0
1934: \]
1935: and
1936: \[
1937: \begin{split}
1938: |s({t_0}-\lambda_n)-s({t_0}-\lambda_n t)| 
1939: &\leq \int_{{t_0}-\lambda_n t}^{{t_0}-\lambda_n}|\dot s|du \\
1940: &\leq \left( \int_{{t_0}-\lambda_n t}^{{t_0}-\lambda_n}r^{1+\alpha/2}| 
1941: \dot s|^2du\right)^{1/2} 
1942: \left(\int_{{t_0}-\lambda_n
1943: t}^{{t_0}-\lambda_n}\frac{du}{r^{1+\alpha/2}}\right)^{1/2}
1944: \end{split}
1945: \]
1946: It is not restrictive to suppose $t>1$: indeed, when $t \in(0,1)$, we obtain an equivalent 
1947: estimate by permuting the integration bounds.
1948: From Corollary~\ref{cor:lim2} and Lemmata~\ref{le:lim1} and~\ref{le:lim2} we obtain
1949: \[
1950: + \infty > \int_{t_0-\delta}^{t_0} r^{1+\alpha} \dot r |\dot s|^2 du 
1951:       \geq \int_{t_0-\delta}^{t_0} c_{1,\alpha}r^{1+\alpha/2}|\dot s|^2 du.
1952: \]
1953: Then, since the constant $c_{1,\alpha}$ is strictly positive, we have
1954: \[
1955: \lim_{n \rightarrow +\infty} \int_{{t_0}-\lambda_n
1956: t}^{{t_0}-\lambda_n}r^{1+\alpha/2}|\dot s|^2 du = 0.
1957: \]
1958: Moreover, as $n$ tends to $+\infty$, the second integral
1959: $\int_{{t_0}-\lambda_n t}^{{t_0}-\lambda_n}r^{-(1+\alpha/2)} < +\infty$; 
1960: indeed both integration bounds tend to $t_0$ and the asymptotic estimate proved
1961: in 
1962: \emph{(b)} holds. 
1963: Hence, as $\lambda_n \rightarrow 0$
1964: \[
1965: \begin{split}
1966: \lim_{n\to+\infty}\int_{{t_0}-\lambda_n
1967: t}^{{t_0}-\lambda_n}\frac{du}{r^{1+\alpha/2}}  & =
1968: \lim_{n\to+\infty}\left[\int_{{t_0}-\lambda_n
1969: t}^{{t_0}-\lambda_n}C\frac{du}{(t_0-u)}  +o(1)\right]\\
1970: & = C \lim_{n\to+\infty} [\log(\lambda_n)-\log(\lambda_n t)+o(1)] = -C\log t
1971: \end{split}
1972: \]
1973: that is bounded since $t$ is fixed and 
1974: $C=\left[\dfrac{\sqrt{2b}(\alpha+2)}{2}\right]^{-(\alpha+2)/2}$.
1975: \end{proof}
1976: 
1977: \begin{theorem}\label{central_conf2}
1978: In the same setting of Theorem~\ref{central_conf},
1979: assume that the potential 
1980: $U$ verifies the further assumption
1981: \begin{assiomi}
1982: \item[\uqhtag]
1983: \label{a:U4h}
1984: $\ds \lim_{r \to 0} r^{\alpha+1}\nabla_{T} U(t,x)=\nabla_{T}\tilde U(t,s)$.
1985: \end{assiomi}
1986: Then 
1987: \[
1988: \dlim_{t \rightarrow {t_0}} \mathrm{dist} \left({\cal C}^b,s(t)\right)=
1989: \dlim_{t \rightarrow {t_0}} \inf_{\bar s \in {\cal C}^b}|s(t) - \bar s | = 0,
1990: \]
1991: where ${\cal C}^b$ is the set of central configurations for $\tilde U$ at level $b$, namely
1992: the subset of critical points of the restriction of 
1993: $\tilde U$  to the ellipsoid ${\mathcal E}$:
1994: \begin{equation}
1995: \label{eq:cal_C}
1996: {\cal C}^b := \left\{ s : \tilde U(t_0,s) = b, \nabla_{T} \tilde U(t_0,s) = 0 \right\}.
1997: \end{equation}
1998: \end{theorem}
1999: 
2000: \begin{remark}
2001: When $U$ is homogeneous, as in the classical keplerian potential, then 
2002: $\tilde U$ is simply the restriction of $U$ 
2003:  on ${\mathcal E}$ and Theorem~\ref{central_conf2} asserts
2004: that the angular component $s$ of the motion tends to a set of central configurations.
2005: \end{remark}
2006: 
2007: \begin{proof}
2008: Since in {\it (a)} of Theorem~\ref{central_conf} we have already proved that 
2009: $\lim_{t \to t_0} \tilde U(t,s(t))=b$, it remains to show that 
2010: \[
2011: \lim_{t \to t_0^-} |\nabla_{T} \tilde U(t,s(t))| = 0
2012: \]
2013: that, using condition \ref{a:U4h}, is equivalent to 
2014: \[
2015: \lim_{t \to t_0^-} r^{\alpha + 1}|\nabla_{T} U(t,rs)| = 0.
2016: \]
2017: We now consider the Euler--Lagrange equation~\eqref{eq:EL}$_2$
2018: multiplied by $r^{\alpha}$
2019: \[
2020: -2 r^{\alpha + 1}\dot r\dot s - r^{\alpha + 2}\ddot s + r^{\alpha + 1}\nabla_{T}
2021: U(t,rs) = r^{\alpha + 2}|\dot s|^2 s;
2022: \]
2023: since $r^{\alpha + 1}\dot r\dot s=r^{\alpha/2 + 1}\dot r r^{\alpha/2}\dot s$ is the product
2024: of a bounded term with an infinitesimal one (see equation~\eqref{eq:limspunto} and
2025: Lemma~\ref{le:lim1}), while $|r^{\alpha + 2}|\dot s|^2 s| = r^{\alpha + 2}|\dot s|^2$
2026: tends to 0 for~\eqref{eq:limspunto}, we claim that
2027: \begin{equation}\label{eq:ddot_s}
2028: \lim_{t \to t_0^-} r^{\alpha + 2}\ddot s = 0.
2029: \end{equation}
2030: We perform the time rescaling (cf.~McGehee's change of coordinates in~\ref{eq:mcgehee2}) 
2031: \begin{equation}\label{eq:mcgehee1}
2032: \tau = \int_{t_0-\delta}^{t} \frac{d\xi}{r^{\alpha/2+1}}
2033: \end{equation}
2034: which maps the interval $[t_0-\delta,t_0)$ into $[0,+\infty)$ (see Corollary~\ref{cor:int2}).
2035: If 
2036: the prime~$\ {}'\ $ denotes
2037: the derivative with respect to the new variable $\tau$, 
2038: then~\eqref{eq:ddot_s} is equivalent to
2039: \begin{equation}\label{eq:ddot_s2}
2040: \lim_{\tau \to +\infty} s''(\tau) = 0
2041: \end{equation}
2042: and the limit~\eqref{eq:limspunto} reads simply 
2043: \begin{equation}\label{eq:ddot_s3}
2044: \lim_{\tau \to +\infty} |s'(\tau)|^2 = 0.
2045: \end{equation}
2046: Suppose now,
2047: for the sake of contradiction,
2048: that there exists a sequence $(\tau_n)_n$ 
2049: such that $\tau_n \to +\infty$ as $n \to +\infty$ and 
2050: \[
2051: \lim_{n\to+\infty} \nabla_{T} \tilde U(\tau_n,s(\tau_n)) = \lim_{n\to+\infty}  s''(\tau_n)= 
2052: \sigma
2053: \]
2054: for some $\sigma \neq 0$.
2055: Since the ellipsoid ${\mathcal E}$ is compact, up to subsequences,
2056: $\left(s(\tau_n)\right)_n$ converges to some $\bar s$. 
2057: Furthermore, from Theorem~\ref{central_conf} we know that
2058: $\tilde U\left(\tau_n,s(\tau_n)\right)$ tends to the finite limit $b$ as 
2059: $n \to +\infty$, hence
2060: $(t_0,\bar s)$ is a regular point both for $\tilde U$ and for $\nabla_T\tilde U$.
2061: We moreover remark that, since the limit~\eqref{eq:ddot_s3} holds, for every fixed positive constant
2062: $h>0,$ there holds
2063: \[
2064: s(\tau) \to \bar s, \qquad \mbox{uniformly on } [\tau_n,\tau_n+h], \; \mbox{for every }n
2065: \]
2066: and also
2067: \[
2068: \sup_{\tau \in [\tau_n,\tau_n+h]}|\nabla_T \tilde U(\tau,s(\tau))-\sigma| \to  0, 
2069: \qquad \mbox{as }n \to +\infty.
2070: \]
2071: We can then compute
2072: \[
2073: \begin{split}
2074: s'(\tau_n +h)-s'(\tau_n) &= \int_{\tau_n}^{\tau_n+h} s''(\tau) d\tau \\
2075: & = \int_{\tau_n}^{\tau_n+h} \nabla_T \tilde U(\tau,s(\tau)) d\tau + o(1) \\
2076: & = h \sigma + o(1) \qquad \mbox{as } n\to +\infty.
2077: \end{split}
2078: \]
2079: We obtain the contradiction
2080: \[
2081: 0 = \lim_{n \to +\infty}|s'(\tau_n +h)-s'(\tau_n)| = h |\sigma| \neq 0.
2082: \]
2083: \end{proof}
2084: 
2085: %--------------------------------
2086: \subsection{Logarithmic potentials}
2087: %--------------------------------
2088: Aim of this section is  to extend  the asymptotic estimates of 
2089: Theorem~\ref{central_conf}
2090: to potentials having
2091: logarithmic singularities.
2092: We follow the same scheme and we still work in a left neighborhood of a total 
2093: collision instant $t_0$, $(t_0-\delta,t_0)$. The main differences concern the
2094:  monotonicity formul\ae\/ (Lemmata~\ref{le:integ_log} and~\ref{le:Gamma_log}).
2095: 
2096: In this setting, we suppose the existence of a continuous function 
2097: \begin{equation} \label{eq:M}
2098: M\from 
2099: (a,b)\to\RR \quad \mbox{ such that $\dot M(t)$ is bounded on $(t_0-\delta,t_0)$}
2100: \end{equation}
2101: and we replace conditions \ref{a:U2h} and \ref{a:U3h} with
2102: \begin{assiomi}
2103: \item[\udltag]
2104: \label{a:U2l}
2105: There exist $\gamma>0$ and $C_2 \geq 0$ such that
2106: \[
2107: \nabla U(t,x)\cdot x + M(t) \geq - C_2 |x|^\gamma U(t,x),
2108: \]
2109: whenever $|x|$ is small.
2110: \item[\utltag]
2111: \label{a:U3l}
2112: $\ds \lim_{|x| \to 0} \left[U(t,x) + M(t)\log |x| \right] = \tilde U(t,s)$, 
2113: uniformly in $t$,
2114: \end{assiomi}
2115: where $\tilde U$, as in the quasi--homogeneous case, is of class $\cont^1$
2116: on $(a,b)\times ({\mathcal E} \minus \Delta)$ and verifies~\eqref{tildeU}.
2117: 
2118: \begin{remark}
2119: \ref{a:U2l} implies \ref{a:U2} (for small value of $|x|$) for every $\tilde \alpha \in (0,2)$.
2120: \end{remark}
2121: 
2122: \begin{remark}\label{rem:int_log}
2123: From Corollary~\ref{cor:2} and assumption~\ref{a:U3l} it follows
2124: that
2125: the positive function $-M(t)\log |x|+\tilde U(t,s)$ is integrable
2126: in a neighborhood of a total collision at the origin.
2127: \end{remark}
2128: 
2129: We now prove the analogue of Lemmata~\ref{le:integ} and~\ref{le:Gamma} in the setting of
2130: logarithmic--type potentials.
2131: \begin{lemma}
2132: \label{le:integ_log}
2133: Let $\bar x$ be a generalized solution, let $t_0 \in (a,b]$ be a total
2134: collision instant and let $\delta$ be given in Theorem~\ref{thm:totcolliso}.
2135: Let $\gamma$ be the positive exponent appearing in \ref{a:U2h}, then 
2136: \begin{equation}
2137: \label{eq:integ_log}
2138: \int_{t_0-\delta}^{t_0}-r^{\gamma}\frac{\dot r}{r}U(t,rs)dt < +\infty.
2139: \end{equation}
2140: \end{lemma}
2141: \begin{proof}
2142: We define the functions
2143: \begin{equation}
2144: \label{eq:function_Gamma_log}
2145: \Gamma_{log} (r,s) := \dfrac{1}{2}r^2|\dot s|^2 - \left[ U(t,rs) + M(t)\log r
2146: \right]
2147: \end{equation}
2148: and
2149: \begin{equation}
2150: \label{eq:function_TGamma}
2151: \tilde \Gamma_{log} (r,s) := r^{\gamma}\Gamma_{log};
2152: \end{equation}
2153: since 
2154: \[
2155: \tilde \Gamma_{log} (r,s) = r^{\gamma}\left[ h(t)-\frac12 \dot r^2 - M(t)\log r
2156: \right]
2157: \leq r^{\gamma}h(t) - r^{\gamma}M(t)\log r,
2158: \]
2159: then $\tilde \Gamma_{log}$ is bounded above, indeed $h$ is bounded, $M$
2160: continuous and, since $\gamma>0$, $\lim_{r \to 0} r^{\gamma}\log r=0$. 
2161: We now proceed exactly as in the proof of Lemma~\ref{le:integ}: we omit here
2162: the approximation argument and
2163: we formally compute the time derivative of $\tilde \Gamma_{log}$
2164: \[
2165: \frac{d}{dt}\tilde \Gamma_{log} (r,s) = \gamma r^{\gamma-1}\dot r \Gamma_{log}
2166: (r,s) + r^\gamma \frac{d}{dt} \Gamma_{log} (r,s).
2167: \]
2168: Using the Euler--Lagrange equation~\eqref{eq:EL}$_2$, we obtain
2169: \[
2170: \frac{d}{dt}\Gamma_{log} (r,s) = 
2171: -r \dot r |\dot s|^2 -\frac{\partial U}{\partial t}(t,rs)
2172: - \frac{\dot r}{r}\nabla U(t,rs)\cdot(rs)
2173: - \dot M(t)\log r - M(t)\frac{\dot r}{r}.
2174: \]
2175: From assumptions \ref{a:U1} and \ref{a:U2l} we deduce that 
2176: \begin{equation}
2177: \label{eq:stima_dGammalog}
2178: \frac{d}{dt}\Gamma_{log} (r,s) 
2179: \geq - r \dot r |\dot s|^2 - C_1 \left(U(t,rs)+1\right)
2180: + C_2 r^{\gamma}\frac{\dot r}{r}U(t,rs) - \dot M(t)\log r
2181: \end{equation}
2182: and then 
2183: \begin{multline}
2184: \label{eq:stima_dTGamma}
2185: \frac{d}{dt}\tilde \Gamma_{log} (r,s) \geq 
2186: - \frac{2-\gamma}{2}r^{\gamma+1} \dot r |\dot s|^2 
2187: - \gamma r^\gamma \frac{\dot r}{r}U(t,rs) 
2188: - \gamma r^\gamma \frac{\dot r}{r}M(t)\log r\\
2189: - C_1 r^\gamma U(t,rs) - C_1 r^\gamma + C_2 r^{2\gamma}\frac{\dot r}{r} U(t,rs)
2190: - \dot M(t)r^\gamma \log r.
2191: \end{multline}
2192: The first term in~\eqref{eq:stima_dTGamma} is positive, since~\eqref{eq:isol_coll}
2193: holds; moreover, since $r$ tends to $0$ as
2194: $t$ approaches $t_0$, there exist $\eps \in (0,\gamma)$ and 
2195: $\delta_0\in (0,\delta]$ such that
2196: \begin{equation}
2197: \label{eq:ultimaGT}
2198: -\gamma r^\gamma \frac{\dot r}{r}U(t,rs)+ C_2 r^{2\gamma}\frac{\dot r}{r}
2199: U(t,rs) \geq -(\gamma-\eps) r^\gamma \frac{\dot r}{r}U(t,rs) \geq 0
2200: \end{equation}
2201: on $(t_0-\delta_0,t_0)$.
2202: The remaining terms in~\eqref{eq:stima_dTGamma} are integrable 
2203: functions, indeed the last term $\dot M(t)r^\gamma \log r$ is bounded as $r$ tends to 0
2204: (see~\eqref{eq:M}),
2205: $r^{\gamma}U \leq U$ and $U$ is integrable and we have the following estimate
2206: \[
2207: -\gamma r^\gamma \frac{\dot r}{r}M(t)\log r \geq 
2208: -\gamma r^{\gamma-1} \dot r\log r \max_{t \in [t_0-\delta,t_0]} M(t)
2209: \]
2210: and
2211: \[
2212: \int_{t_0-\delta}^{t_0}\gamma r^{\gamma-1}\dot r\log r dt =  -r_0^\gamma\log r_0 +
2213: \int_{t_0-\delta}^{t_0} r^{\gamma-1}\dot r dt
2214: = r_0^\gamma \left(-\log r_0+\frac{1}{\gamma}\right) < +\infty
2215: \]
2216: where $r_0=r(t_0-\delta)$.
2217: Hence the right hand side of~\eqref{eq:stima_dTGamma} is the sum of an
2218: integrable function with a
2219: positive one; since the $\tilde \Gamma_{log} (r,s)$ is bounded above
2220: from~\eqref{eq:ultimaGT}
2221: we have the estimate in~\eqref{eq:integ_log}.
2222: \end{proof}
2223: 
2224: \begin{lemma}[{\bf Monotonicity Formula}]
2225: \label{le:Gamma_log}
2226: The function $\Gamma_{log}$ defined in~\eqref{eq:function_Gamma_log} is bounded
2227: on $[t_0-\delta,t_0]$.
2228: \end{lemma}
2229: \begin{proof}
2230: We consider the expression of the derivative of $\Gamma_{log}$ with respect to
2231: the time variable computed in~\eqref{eq:stima_dGammalog}.
2232: Using Lemma~\ref{le:integ_log}, the integrability of the function $U$ and
2233: Remark~\ref{rem:int_log} we deduce the boundedness below 
2234: (in a left neighborhood of $t_0$) 
2235: of the function $\Gamma_{log}$ being the right hand side 
2236: of~\eqref{eq:stima_dGammalog} the sum of a positive function with an integrable one.
2237: 
2238: To prove the boundedness above of $\Gamma_{log}$ we cannot use the
2239: boundedness of the energy function, indeed in this case we can just estimate 
2240: $\Gamma_{log}(r,s) + M(t)\log r = h(t) - \frac12 \dot r ^2 $.
2241: By the sake of contradiction suppose that $\Gamma_{log}$ diverges to $+\infty$
2242: as 
2243: $t$ tends to $t_0$; since  $U(t,rs) + M(t)\log r$ converges uniformly to
2244: $\tilde U(t,s)$ as
2245: $t$ tends to $t_0$ and $\tilde U(t,s)$ is a positive function, if $\Gamma_{log}$
2246: diverges to $+\infty$ 
2247: \begin{equation}
2248: \label{assurdo}
2249: \exists t_1 \in (t_0-\delta,t_0) \mbox{ such that }  \forall \, t \in (t_1,t_0), \quad
2250: r^2 |\dot s|^2 > \max_{t \in [t_0-\delta,t_0]} M(t).
2251: \end{equation}
2252: From assumption~\eqref{assurdo} we have
2253: \begin{equation}
2254: \label{assurdo2}
2255: \begin{split}
2256: \int_{t_0-\delta}^{t_0} -\frac{\dot r}{r}\left( r^2 |\dot s|^2 - M(t) \right)dt & =
2257: \int_{t_0-\delta}^{t_1} -\frac{\dot r}{r}\left( r^2 |\dot s|^2 - M(t) \right)dt +
2258: \int_{t_1}^{t_0} -\frac{\dot r}{r}\left( r^2 |\dot s|^2 - M(t) \right)dt \\
2259: & \geq \text{constant} - \lim_{t \to t_0} \log r(t) = +\infty.
2260: \end{split}
2261: \end{equation}
2262: We now define the function 
2263: \[
2264: \Omega_{log}(r,s) := \Gamma_{log}(r,s) + M(t)\log r = h(t) -\frac12 \dot r^2
2265: \]
2266: that is bounded above.
2267: When we compute its derivative with respect to the time variable we obtain
2268: the sum of a positive function with an integrable one (we use
2269: assumption~\eqref{assurdo}
2270: and Lemma~\ref{le:integ_log}), indeed
2271: \[
2272: \begin{split}
2273: \frac{d}{dt}\Omega_{log}(r,s) 
2274: &= \frac{d}{dt}\Gamma_{log}(r,s) + \dot M(t)\log r + M(t)\frac{\dot r}{r}\\
2275: &\geq -\frac{\dot r}{r} \left[ r^2 |\dot s|^2 -M(t) \right] 
2276: - C_1 \left(U(t,rs)+1\right) + C_2 r^{\gamma}\frac{\dot r}{r}U(t,rs).
2277: \end{split}
2278: \]
2279: We can then conclude the boundedness of $\Omega_{log}$ on the interval $[t_0-\delta,t_0]$
2280: and from the estimate on its derivative we have
2281: \[
2282: \int_{t_0-\delta}^{t_0} -\frac{\dot r}{r}\left( r^2 |\dot s|^2 - M(t) \right)dt < +\infty
2283: \]
2284: that contradicts~\eqref{assurdo2}.
2285: We conclude that the function $\Gamma_{log}$ is also bounded above.
2286: \end{proof}
2287: 
2288: \begin{corollary}
2289: \label{cor:lim_gamma_log}
2290: As $t$ tends to $t_0$ the limit of the function $\Gamma_{log}$ exists finite
2291: and 
2292: \[
2293: \lim_{t \to t_0^+} -\frac{\dot r^2}{2\log r}=M_0
2294: \]
2295: where $M_0:=M(t_0)$.
2296: \end{corollary}
2297: \begin{proof}
2298: We argue as in the proof of Corollary~\ref{cor:limGamma}
2299: to show that the function $\Gamma_{log}$ has a finite limit as 
2300: $t$ tends to $t_0$. Since $\Gamma_{log} = h(t)- \frac12 \dot r^2 - M(t)\log r$,
2301: we conclude dividing by $\log r$ using the boundedness of the function $h$.
2302: \end{proof}
2303: 
2304: \begin{theorem}
2305: \label{central_conf_log}
2306: Let ${\bar x}$ be a generalized solution for the dynamical system~\eqref{DS}
2307: and let $t_0\in(a,b)$ (in the case $b<+\infty$, $t_0$ can coincide with $b$) be a total
2308: collision instant.
2309: Let $r,s$ be the new variables defined in~(\ref{rs}); if the potential $U$ satisfies
2310: assumptions \ref{a:U0}, \ref{a:U1}, \ref{a:U2l}, \ref{a:U3l} then the following assertions hold
2311: \begin{itemize}
2312: \item[(a)] $\displaystyle \lim_{t \rightarrow {t_0}^-} \left[ U(t,rs) + M(t)\log r
2313: \right] = - \lim_{t \rightarrow {t_0}^-} \Gamma_{log}(r,s) = b$;
2314: \item[(b)]
2315: as $t$ tends to ${t_0}$, 
2316: \[\begin{aligned}
2317: r(t) & \sim
2318: ({t_0}-t)\sqrt{-2M_0\log({t_0}-t)} \\
2319: \dot r(t) & \sim
2320: -\sqrt{-2M_0\log({t_0}-t)}; 
2321: \end{aligned}\]
2322: % \item[(b)] as $t$ tends to ${t_0}$, $\displaystyle r(t) \sim
2323: % ({t_0}-t)\sqrt{-2M_0\log({t_0}-t)}$;
2324: % \item[(c)] as $t$ tends to ${t_0}$, $\displaystyle \dot r(t) \sim
2325: % -\sqrt{-2M_0\log({t_0}-t)}$;
2326: \item[(c)] $\displaystyle \lim_{t \rightarrow {t_0}^-} |\dot
2327: s(t)|({t_0}-t)\sqrt{-2M_0\log({t_0}-t)} = 0$;
2328: \item[(d)] for every real positive sequence $(\lambda)_n$, such that
2329: $\lambda_n \rightarrow 0$ as $n \rightarrow +\infty$ we have
2330: \[
2331: \lim_{n \rightarrow +\infty} |s({t_0}-\lambda_n) - s({t_0}-\lambda_n t)|=0, 
2332: \quad \forall t > 0.
2333: \]
2334: \end{itemize}
2335: \end{theorem}
2336: \begin{proof}
2337: \emph{(a)} The proof is essentially the same of for 
2338: Theorem~\ref{central_conf}.
2339: 
2340: \noindent \emph{(b)} From Corollary~\ref{cor:lim_gamma_log} we deduce that 
2341: \[
2342: \dot r(t) \sim -\sqrt{-2M_0\log r(t)}\quad \mbox{ as } t \mbox{ tends to }
2343: t_0.
2344: \]
2345: We define $R(t) := ({t_0}-t)\sqrt{-2M_0\log({t_0}-t)}$ and we remark that, as
2346: $t$ tends to $t_0$
2347: \[
2348: -\log R(t) = -\log({t_0}-t)-\log\left(\sqrt{-2M_0\log({t_0}-t)}\right) \sim
2349: -\log({t_0}-t)
2350: \]
2351: and
2352: \[
2353: \dot R(t) = -\sqrt{-2M_0\log({t_0}-t)}
2354:             +\frac{M_0}{\sqrt{-2M_0\log({t_0}-t)}}\sim -\sqrt{-2M_0\log
2355: R(t)}.
2356: \]
2357: Our aim is then to prove that the function $r(t)$ is asymptotic to $R(t)$ 
2358: as $t$ tends to $t_0$. We define the following functions
2359: \[
2360: f(\xi) := -\sqrt{-2M_0\log \xi} \quad \mbox{ and } \quad
2361: \Phi(\xi):=\int_0^\xi \frac{d\eta}{f(\eta)}, \quad \xi \in(0,1] 
2362: \]
2363: and we remark that $\Phi(0)=0$ and $\Phi$ is a strictly decreasing function on
2364: $[0,1]$. Moreover
2365: \[
2366: \dot r(t) \sim f(r(t)), \quad \dot R(t) \sim f(R(t)) \mbox{ as } t \mbox{ tends
2367: to } t_0
2368: \]
2369: or equivalently
2370: \[
2371: \lim_{t \rightarrow t_0}\frac{d}{dt}\Phi(r(t)) = \lim_{t \rightarrow t_0}
2372: \frac{d}{dt}\Phi(R(t)) = 1.
2373: \]
2374: Since the function $\Phi(\xi)$ decreases in $\xi$ and $r(t)$, 
2375: $R(t)>0$ decreases in
2376: $t$ (when we stay close to
2377: the collision instant) we have that the functions $\Phi(r(t))$ and $\Phi(R(t))$
2378: are negative 
2379: on $(t_0-\delta_0,t_0)$, vanishes at $t_0$ (since $r(t_0)=R(t_0)=0$) and increase in the
2380: variable $t$.
2381: Furthermore fixed $\bar t < t_0$, the following property holds 
2382: \begin{equation}
2383: \label{eq:prop_Phi}
2384: \begin{split}
2385: \frac{d}{dt}\Phi(r(t)) \leq 1 \leq \frac{d}{dt}\Phi(R(t)), \, \forall t \in
2386: (\bar t,t_0)
2387: &\quad \Rightarrow \quad \Phi(r(t)) \geq \Phi(R(t)), \, \forall t \in (\bar
2388: t,t_0) \\
2389: &\quad \Rightarrow \quad r(t) \leq R(t),\, \forall t \in (\bar t,t_0).
2390: \end{split}
2391: \end{equation}
2392: For every $\epsilon > 0$, we consider the functions 
2393: \[
2394: \begin{split}
2395: & R^+_{\eps} (t):= (1+\eps)R(t),\\
2396: & R^-_{\eps} (t):= (1-\eps)R(t).
2397: \end{split}
2398: \]
2399: Since $\dot R(t) \sim f(R(t))$, we deduce that in a left neighborhood of $t_0$
2400: \begin{equation}
2401: \label{eq:log_pepenult}
2402: \begin{split}
2403: &\dot R^+_{\eps} (t) = (1+\eps)\dot R(t) \leq \left( 1 +
2404: \frac{\eps}{2}\right)f(R(t))
2405:                        \leq \left( 1 + \frac{\eps}{2}\right)f(R^+_{\eps}(t)),\\
2406: &\dot R^-_{\eps} (t) = (1-\eps)\dot R(t) \geq \left( 1 -
2407: \frac{\eps}{2}\right)f(R(t))
2408:                                               \geq \left( 1 -
2409: \frac{\eps}{2}\right)f(R^-_{\eps}(t)),
2410: \end{split}
2411: \end{equation}
2412: indeed 
2413: \[
2414: f(R(t)) = -\sqrt{-2M_0\log(R(t))} \leq
2415: -\sqrt{-2M_0\log(1+\eps)-2M_0\log(R(t))}=f(R^+_{\eps}(t))
2416: \]
2417: and similarly 
2418: \[
2419: f(R(t)) \geq -\sqrt{-2M_0\log(1-\eps)-2M_0\log(R(t))}=f(R^-_{\eps}(t)). 
2420: \]
2421: 
2422: From~\eqref{eq:log_pepenult} we then obtain
2423: \begin{equation}
2424: \label{eq:log_penult}
2425: \frac{d}{dt} \Phi(R^+_{\eps}(t)) \geq 1 + \frac{\eps}{2} \quad \mbox{and} \quad
2426: \frac{d}{dt} \Phi(R^-_{\eps}(t)) \leq 1 - \frac{\eps}{2}.
2427: \end{equation}
2428: Moreover, since $\dot r(t) \sim f(r(t))$, again in a left neighborhood of $t_0$
2429: we have that
2430: \begin{equation}
2431: \label{eq:log_ult}
2432: \left( 1 + \frac{\eps}{2}\right)f(r(t)) \leq \dot r(t) \leq \left( 1 -
2433: \frac{\eps}{2}\right)f(r(t))
2434: \end{equation}
2435: and dividing~\eqref{eq:log_ult} for the negative function $f(r(t))$ and
2436: comparing the resulting inequalities with~\eqref{eq:log_penult} we have
2437: \[
2438: \frac{d}{dt} \Phi(R^-_{\eps}(t)) \leq \frac{d}{dt} \Phi(r(t)) \leq \frac{d}{dt}
2439: \Phi(R^+_{\eps}(t)).
2440: \]
2441: From~(\ref{eq:prop_Phi}) we deduce that, in a neighborhood of the collision
2442: instant $t_0$,
2443: the following chain of inequalities holds
2444: \[
2445: (1-\eps) \leq \frac{r(t)}{R(t)} \leq (1+\eps).
2446: \]
2447: The second estimate follows directly.
2448: 
2449: \noindent \emph{(c)} From the result proved in \emph{(a)} we have that $\lim_{t
2450: \to t_0}r|\dot s|=0$;
2451: we conclude using \emph{(b)}.
2452: 
2453: \noindent \emph{(d)} As in the proof of Theorem~\ref{central_conf}, 
2454: if $t=1$ there is nothing to prove. We then chose $t>0$, $t\neq 1$, 
2455: a sequence $(\lambda_n)_n$, $\lambda_n \rightarrow 0$ and
2456: $N$, sufficiently large, such that $\lambda_n < \delta/\max(1,t)$, $\forall n \geq N$.
2457: We then obtain
2458: \begin{eqnarray*}
2459: |s({t_0}-\lambda_n)-s({t_0}-\lambda_n t)| 
2460: & \leq & \int_{{t_0}-\lambda_n t}^{{t_0}-\lambda_n}|\dot s(u)|du \\
2461: & \leq & \left( \int_{{t_0}-\lambda_n t}^{{t_0}-\lambda_n}-r(u)\dot r(u)|\dot
2462: s(u)|^2du \right)^{\frac{1}{2}}
2463: \left( \int_{{t_0}-\lambda_n t}^{{t_0}-\lambda_n}-\frac{du}{r(u)\dot r(u)}\right)^{\frac{1}{2}}.
2464: \end{eqnarray*}
2465: The boundedness of the $\Gamma_{log}$ and the estimate on its derivative
2466: in~\eqref{eq:stima_dGammalog} 
2467: imply the boundedness of the integral $\int_0^{t_0}r \dot r|\dot s|^2$ and then
2468: \begin{equation*}
2469: \lim_{n \rightarrow +\infty} \int_{{t_0}-\lambda_n t}^{{t_0}-\lambda_n}-r(u)\dot
2470: r(u)|\dot s(u)|^2du =0.
2471: \end{equation*}
2472: Moreover, as $n$ tends to $+\infty$, from \emph{(b)} and \emph{(c)} we have 
2473: $r(u)\dot r(u) \sim -2M_0(t_0-u)\log (t_0-u)$, hence 
2474: \begin{equation*}
2475: \lim_{n \rightarrow +\infty}
2476: \int_{{t_0}-\lambda_n t}^{{t_0}-\lambda_n}\frac{du}{r(u)\dot r(u)} 
2477: = \frac{1}{M_0}\lim_{n \rightarrow +\infty}\log \frac{\log \lambda_n t}{\log \lambda_n} = 0.
2478: \end{equation*}
2479: The proof is now complete.
2480: \end{proof}
2481: 
2482: The behavior of the angular part is conserved also for logarithmic potential 
2483: and the following result can be proved following the proof of 
2484: Theorem~\ref{central_conf2}.
2485: \begin{theorem}\label{central_conf2_log}
2486: In the same setting of Theorem~\ref{central_conf_log},
2487: assuming furthermore that the potential $U$ verifies
2488: \begin{assiomi}
2489: \item[\uqltag]
2490: \label{a:U4l}
2491: $\ds \lim_{r \to 0} r\nabla_{T} U(t,x)=\nabla_{T}\tilde U(t,s)$,
2492: \end{assiomi}
2493: then there holds
2494: \[
2495: \dlim_{t \rightarrow {t_0}} \mathrm{dist} \left({\cal C}^b,s(t)\right)=
2496: \dlim_{t \rightarrow {t_0}} \inf_{\bar s \in {\cal C}^b}|s(t) - \bar s | = 0
2497: \]
2498: where ${\cal C}^b$ is the central configuration subset defined in~\eqref{eq:cal_C}.
2499: \end{theorem}
2500: 
2501: %===========================
2502: \section{Partial collisions}
2503: \label{sec:partial}
2504: %===========================
2505: This section is devoted to the study of the singularities which are not total collision at the origin.
2506: At first we shall prove the existence of a limiting configuration for bounded trajectories, that is
2507: the Von Zeipel's Theorem  (stated on page \pageref{teo:Zeipel}). This fact allows the reduction from partial to
2508: total collisions through a change of coordinates. To carry on the analysis we shall extend the clustering argument proposed by McGehee in \cite{mcgehee2} 
2509: to prove the Von Zeipel's Theorem. To this aim we need to introduce some further assumptions on the potential $U$ and its singular 
2510: set $\Delta$.
2511: More precisely we suppose that
2512: \begin{equation} \label{eq:strucDelta}
2513: \Delta = \bigcup_{\mu\in\mathcal M} V_\mu,
2514: \end{equation}
2515: where the $V_\mu$'s are distinct linear subspaces of $\RR^{k}$ and $\mathcal M$ 
2516: is a finite set;
2517: observe that the set $\Delta$ is a cone, as required on page \pageref{cone}. We endow the 
2518: family of the $V_\mu$'s with the inclusion partial ordering and we assume the family to be 
2519: closed with respect to  intersection
2520: (thus we are assuming that $\mathcal{M}$ is a semilattice
2521: of linear subspaces of $\RR^k$: it is the intersection semilattice generated 
2522: by the arrangement of maximal subspaces $V_\mu$'s).
2523: With each $\xi\in\Delta$ we associate 
2524: \[
2525: \mu(\xi)=\min\{\mu\;:\; \xi\in V_\mu\}\qquad \text{i.e.,}
2526: \qquad V_{\mu(\xi)}=\bigcap_{\xi\in V_\mu} V_\mu.
2527: \]
2528: Fixed $\mu\in {\mathcal M}$ we define the  set of collision configurations satisfying
2529: \[
2530: \Delta_\mu = \left\{ \xi \in \Delta : \mu(\xi)=\mu \right\}
2531: \]
2532: and we observe that this is an open subset of $V_\mu$ and its closure 
2533: $\overline{\Delta_\mu}$ is 
2534: $ V_\mu$. We also notice that the map $\xi\to \dim(V_{\mu(\xi)})$ is lower semicontinuous.
2535: 
2536: We denote by $p_\mu$ the orthogonal projection onto $V_\mu$ and we write
2537: \[
2538: x=p_\mu(x)+w_\mu(x),
2539: \]
2540: where, of course,  $w_\mu=\mathbb I-p_\mu$.
2541: 
2542: We assume that, near the collision set, the potential depends, roughly, only on the 
2543: projection orthogonal to the collision set: more precisely we assume
2544: 
2545: \begin{assiomi} 
2546: \item[\uctag]
2547: \label{a:U5}
2548: For every $\xi\in\Delta$, there is $\varepsilon>0$ such that
2549: \[
2550: \displaystyle U(t,x) -U(t,w_{\mu(\xi)}(x))= W(t,x) 
2551: \in {\cal C}^1\left((a,b)\times B_\eps(\xi)\right),
2552: \]
2553: where $B_\eps(\xi)=\{x\;:\: |x-\xi|<\varepsilon\}$.
2554: \end{assiomi}
2555: 
2556: %
2557: % We moreover assume the boundedness of the potential away from the singularities
2558: % \begin{enumerate}
2559: % \item[\us] $\displaystyle \forall \eps>0, \;\forall C>0, \; \exists M>0 :
2560: % \left(d(x,\Delta)>\eps \mbox{ and } I(x)\leq C 
2561: % \implies |\nabla U(t,x)| \leq M\right)$.
2562: % where $d(x,\Delta)$ denotes the distance of $x$ from $\Delta$.
2563: % \end{enumerate}
2564: 
2565: \begin{theorem}\label{thm:partial}
2566: Let $\bar x$ be a generalized solution for the dynamical system~\eqref{DS} on the 
2567: bounded interval $(a,b)$.
2568: Suppose that the potential $U$ satisfies assumptions \ref{a:U0}, \ref{a:U1}, \ref{a:U5}, and 
2569: \ref{a:U2h}, \ref{a:U3h}, \ref{a:U4h} (or \ref{a:U2l}, \ref{a:U3l}, \ref{a:U4l}).\\
2570: If $\bar x$
2571: is bounded on the whole interval $(a,b)$ then 
2572: \begin{enumerate}
2573: \item[(a)] $\bar x$ has a finite number of  singularities which are collisions 
2574: (the Von Zeipel's Theorem holds).
2575: \item[(b)] Furthermore, if $t^* \in \bar x^{-1}(\Delta)$ is a collision instant, 
2576: $x^*$ the limit configuration  of $\bar x$ as $t$ tends to $t^*$ and
2577: $\mu^*=\mu(x^*) \in {\mathcal M}$, then 
2578:  $r_{\mu^*}=|w_{\mu^*}(\bar x)|$, $s_{\mu^*}=w_{\mu^*}(\bar x)/r_{\mu^*}$ and $U_{\mu^*}=
2579:  U(t,w_{\mu^*}(\bar x))$ satisfy the asymptotic estimates given in
2580: Theorems~\ref{central_conf} and~\ref{central_conf2} 
2581: (or Theorems~\ref{central_conf_log} and~\ref{central_conf2_log} 
2582: when \ref{a:U2l}, \ref{a:U3l} and \ref{a:U4l} hold).
2583: \end{enumerate}
2584: \end{theorem}
2585: 
2586: \begin{proof}
2587: Let $\bar x$ be a generalized solution with a 
2588: singularity at $t=t^*$ (see Definition~\ref{def:singularity}) and $\Delta^*$ its 
2589: $\omega$-limit set, that is
2590: \[
2591: \Delta^* = \left\{ x^* : \exists (t_n)_n \mbox{ such that } t_n \to t^*
2592: \mbox{ and } \bar x(t_n)\to x^* \right\}.
2593: \]
2594: It is well known that the  $\omega$-limit of a bounded trajectory is a compact and connected set. From the Painlev\'e's Theorem (on page \pageref{teo:Painleve}) we have the inclusion
2595: \[
2596: \Delta^*  \subset \Delta.
2597: \]
2598: Von Zeipel's Theorem asserts that whenever  $\bar x$ remains
2599: bounded as $t$ approaches $t^*$, then the $\omega$-limit set of $\bar x$ 
2600: contains just one element, that is $\Delta^* = \{ x^* \}$.
2601: 
2602: In view of Corollary~\ref{cor:2}, where we proved the Theorem in the case 
2603: $\liminf_{t \to t^*} \dot I(\bar x(t))$ $< +\infty$, we are left with the case
2604: when
2605: \begin{equation*}
2606: \lim_{t \to t^*} \dot I(\bar x(t)) = +\infty.
2607: \end{equation*}
2608: From this and our  assumptions it follows that $I(\bar x(t))$ is a 
2609: definitely increasing and bounded 
2610: function. Hence it admits a limit
2611: \begin{equation}
2612: \label{hp:liminf}
2613: \lim_{t \to t^*} I(\bar x(t)) = I^*.
2614: \end{equation}
2615: 
2616: We perform the proof of Von Zeipel's Theorem in two steps.
2617: 
2618: \bigskip
2619: 
2620: \noindent 
2621: {\em Step 1. We suppose that $\mu\left(\Delta^*\right) = \{\mu^*\}$ for some 
2622: $\mu^* \in {\mathcal M}$ and we show that $\Delta^* = \{ x^* \}$}.
2623: 
2624: \medskip
2625: 
2626: \noindent 
2627: 
2628: As  $\Delta^*$ is a compact and 
2629: connected subset of $V_{\mu^*}$, we have the following inclusions 
2630: \[
2631: \Delta^* \subset \Delta_{\mu^*} \subset V_{\mu^*}.
2632: \]
2633: % Since $\Delta^*$ is compact and  $\Delta_{\mu^*}$ is open in $V_{\mu^*}$, 
2634: % there exists $\eps>0$ such that
2635: % the $\eps$-neighborhood of $\left(\Delta^*\right)^\eps$ does not intersect
2636: % $V_{\mu^*}\minus\Delta_{\mu^*}$.
2637: We consider the orthogonal projections
2638: \[
2639: p(t)=p_{\mu^*}(\bar x(t)),\qquad\qquad w(t)=w_{\mu^*}(\bar x(t)).
2640: \]
2641: Since we have assumed that $\mu(\Delta^*)=\{\mu^*\}$, then
2642: \begin{equation}\label{eq:bar_coor}
2643: \lim_{t \to t^*} w(t) = 0,
2644: \end{equation}
2645: our aim is now to prove that
2646: \[
2647: \lim_{t \to t^*} p(t) = x^*.
2648: \]
2649: 
2650: Projecting on $V_{\mu^*}$ the equations of motion, we obtain from \ref{a:U5}
2651: \begin{equation}\label{eq:DScenter}
2652: -\ddot p=p_{\mu^*}\left(\nabla U(t,\bar x(t))\right)=p_{\mu^*}\left(\nabla W(t,\bar x(t))\right)
2653: \end{equation}
2654: where $\nabla W$ is globally bounded as $t\to t^*$. Indeed, fixed $\eps>0$,
2655: %such that $B_\eps\left( \Delta^* \right) \cap \left( V_{\mu^*}\minus 
2656: %\Delta_{\mu^*}\right) = \emptyset$, 
2657: there exists $\delta>0$ such that 
2658: $\bar x(t)\in B_\eps\left( \Delta^* \right)$ whenever $t \in (t^*-\delta,t^*)$, and
2659: from assumption \ref{a:U5} and the compactness of $\Delta^*\subset\Delta_{\mu^*}$ 
2660: we deduce the boundedness of the right hand 
2661: side of~\eqref{eq:DScenter}. From this fact we easily deduce the existence of a limit for $(p(t))$ as 
2662: $t$ tends to $t^*$.
2663: A word of caution must be entered at this point. As $\bar x$ is a generalized 
2664: solution to~\eqref{DS}, the
2665: equation of motions are not available, because of the possible occurence of collisions, and 
2666: therefore they can not be projected on $V_{\mu^*}$. Nevertheless, exploiting the 
2667: regularization method exposed in Section~\ref{sec:sing} and projecting the regularized 
2668: equations, one can easily obtain the validity of~\eqref{eq:DScenter} after passing to the 
2669: limit.
2670: \bigskip
2671: 
2672: \noindent 
2673: {\em Step 2. There always exists $\mu^* \in {\cal M}$ such that 
2674: $\mu\left(\Delta^*\right) = \{\mu^*\}$.}
2675: 
2676: \medskip
2677: 
2678: \noindent Let $\mu^*$ be the element of $\mu\left(\Delta^*\right)$ associated with the
2679: subspace $V_{\mu^*}$ having \emph{minimal dimension}. Since the function 
2680: $\xi\to \dim(V_{\mu(\xi)})$ is lower semicontinuous, 
2681: the minimality of the dimension has as a main implication that 
2682: $\Delta_{\mu^*}\cap\Delta^*$ is compact. Hence the function $\nabla W$ appearing in \ref{a:U5} can be
2683: though to be globally bounded in a neighborhood of $\Delta_{\mu^*}\cap\Delta^*$.
2684: In other words, when considering the orthogonal projections
2685: $p(t)=p_{\mu^*}(\bar x(t))$ and $w(t)=w_{\mu^*}(\bar x(t))$, 
2686: as a major consequence of the minimality of the dimension $\mu^*$ we find the following
2687: implication: 
2688: \begin{equation}\label{eq:min1}
2689: \exists M>0, \; \exists \eps>0 : |w(t)|^2 < \eps
2690: \implies|p_{\mu^*}\left(\nabla W(t,\bar x)\right)| \leq M .
2691: \end{equation}
2692: We now compute the second derivative (with respect to the time $t$) 
2693: of the function $|p(t)|^2$
2694: \[
2695: \dfrac{d^2}{d t^2} |p(t)|^2  = 2\ddot p(t)\cdot p(t)
2696: + 2 \dot p(t)\cdot \dot p(t)  \geq 
2697: -2 p_{\mu^*}\left(\nabla W(t,\bar x(t))\right)\cdot p(t)
2698: \]
2699: Thus, from the projected motion equation ~\eqref{eq:DScenter} and from~\eqref{eq:min1} we infer
2700: \begin{equation}\label{eq:min2}
2701: \exists K>0, \; \exists \eps>0 : |w(t)|^2 < \eps
2702: \implies  \dfrac{d^2}{d t^2}|p(t)|^2 \geq -K .
2703: \end{equation}
2704: We now argue by contradiction, supposing that $\mu\left(\Delta^*\right) \neq \{\mu^*\}$. 
2705: Then
2706: \begin{equation}\label{eq:inertia2}
2707: 0 = \liminf_{t \to t^*}|w(t)|^2 < \limsup_{t \to t^*}|w(t)|^2.
2708: \end{equation}
2709: Since, obviously, the total moment of inertia splits as
2710: \[
2711: I(\bar x(t))=|p(t)|^2+|w(t)|^2,
2712: \]
2713: from~\eqref{eq:inertia2} and~\eqref{hp:liminf} we deduce that
2714: \begin{equation}\label{eq:HPassurdo}
2715: I^* = \limsup_{t \to t^*}|p(t)|^2 > \liminf_{t \to t^*}|p(t)|^2
2716: \end{equation}
2717: and from~\eqref{eq:HPassurdo} together with~\eqref{eq:min2} we have
2718: \[
2719: \exists K>0, \; \exists \eps>0 \;:\;  |p(t)|^2 \geq I^* - \eps
2720: \implies  \dfrac{d^2}{d t^2} |p(t)|^2\geq -K .
2721: \]
2722: Let $(t_n^0)_n$ and $(t_n^*)_n$ be two sequences such that, fixed $\eps>0$
2723: \[
2724: \begin{split}
2725: & t_n^*<t_n^0<t_{n+1}^* \quad \forall n\\
2726: & t_n^0 \to t^* \quad t_n^* \to t^* \mbox{ as } n \to +\infty \\
2727: & f(t_n^*) \to I^* \mbox{ as } n \to +\infty  \mbox{ and } f'(t_n^*)=0, \quad \forall n \\
2728: & t_n^0=\inf\{t>t^*_n : |p(t)|^2 \leq I^*-\eps\}, \quad \forall n. 
2729: \end{split}
2730: \]
2731: Hence $|p(t_n^0)|^2-|p(t_n^*)|^2 = \dfrac{d}{d t^2}|p(\xi)|^2(t_n^0-t_n^*)^2/2 \geq -K(t_n^0-t_n^*)^2/2$ and then 
2732: \[
2733: -\eps \geq \frac{-K}{2}(t_n^0-t_n^*)^2 \quad \mbox{or} \quad
2734: (t_n^0-t_n^*)^2 \geq \frac{2\eps}{K}
2735: \]
2736: in contradiction with the assumptions that both sequences $(t_n^0)_n$ and $(t_n^*)_n$
2737: tend to the finite limit $t^*$.
2738: This concludes the proof of the Von Zeipel's Theorem. Next we prove isolatedness of collision instants.
2739: 
2740: To this aim, let us select  $t^* \in \partial\left(\bar x^{-1}(\Delta)\right)$  a collision instant
2741: such that the dimension of $V_{\mu(\bar x(t^*))}$ is minimal among all dimensions of collision configurations $V_{\mu(\bar x(t))}$
2742: in $(t^*-\delta,t^*+\delta)$ for some $\delta>0$. As before, let us split the components of the trajectory $\bar x(t)= p(t)+w(t)$ on $V_{\mu^*}$ and its
2743: orthogonal complement.
2744: 
2745: Since $\mu^*$ is minimal (see~\eqref{eq:min1}), we already know from the previous discussion that the equations of motion 
2746: projected on the subspace $V_{\mu^*}$ (equation~\eqref{eq:DScenter}) are not singular; 
2747: on the other hand, by \ref{a:U5}, the trajectories in the orthogonal coordinates $w$ are 
2748: generalized solutions to a dynamical system of the form
2749: \begin{equation}\label{eq:DSproj}
2750: -\ddot w = \nabla U(t,w)+\nabla W(t,p(t)+w).
2751: \end{equation}
2752: Now, since $w(t)$ has a total collisions at the origin at $t^*$, we can apply the results of Section 
2753: \ref{sec:asymp}. More precisely, at first we deduce 
2754: from Theorem~\ref{thm:totcolliso} that $t^*$ is isolated in the set of collisions $\Delta_{\mu^*}$; 
2755: furthermore from Corollary~\ref{cor:2} we deduce the boundedness of the action 
2756:  and the energy. Finally we conclude applying 
2757: Theorems~\ref{central_conf},~\ref{central_conf2} (or Theorems~\ref{central_conf_log}, 
2758: \ref{central_conf2_log} when \ref{a:U2l}, \ref{a:U3l} and \ref{a:U4l} hold) to the projection $w$. 
2759: In particular from (a) in Theorem~\ref{central_conf} (or Theorem~\ref{central_conf_log})
2760: we obtain that every collision is isolated and hence, whenever the interval $(a,b)$ is finite,
2761: the existence of a finite number of collisions.
2762: \end{proof}
2763: 
2764: %==============================
2765: \section{Absence of collisions for locally minimal path}
2766: \label{sec:blowup}
2767: %==============================
2768: 
2769: As a matter of fact, solutions to the Newtonian $n$--body problem which are
2770: minimals for the action are, very likely, free of any collision. This  was
2771: discovered  in \cite{serter1} for a class of periodic three--body problems and,
2772: since then, widely exploited in the literature concerning the variational
2773: approach to the periodic $n$--body problem. In general, the proof goes by the
2774: sake of the contradiction and involves the construction of a suitable variation
2775: that lowers the action in presence of a collision.  A recent breakthrough in
2776: this direction is due of the neat idea, due to C. Marchal in 
2777: \cite{marchal}, 
2778: of averaging over a family of variations parameterized on a sphere. The method
2779: of averaged variations for Newtonian potentials has been developed and exposed
2780: in \cite{Ch2}, and then extended to $\alpha$--homogeneous potentials and
2781: various constrained minimization problems in \cite{FT}. This argument can be
2782: used in most of the known cases to prove that minimizing trajectories are
2783: collisionless. In this section we prove the absence of collisions for locally
2784: minimal solutions when the potentials have  quasi--homogeneous or logarithmic
2785: singularities.
2786: 
2787: We consider separately the quasi--homogeneous and the logarithmic cases;
2788: indeed in the first case one can exploit the blow--up technique as
2789: developed in Section 7 of \cite{FT};
2790: in \S~\ref{sec:blowup_hom} we will just recall the main steps of this
2791: arguments.  On the other hand, when dealing with logarithmic potentials, the
2792: blow--up technique is no longer available and we conclude proving directly some
2793: averaging estimates that can be used to show the nonminimality of large classes
2794: of colliding motions.
2795: 
2796: %==============================
2797: \subsection{Quasi--homogeneous potentials}
2798: \label{sec:blowup_hom}
2799: %==============================
2800: Let $\tilde U$ be the $\cont^1$ function defined on $(a,b) \times ({\cal E}\minus\Delta)$
2801: introduced on page \pageref{tildeU}; we extend its definition on the whole 
2802: $(a,b) \times (\RR^{k}\minus\Delta)$ in the following way
2803: \[
2804: \tilde U(t,x)= |x|^{-\alpha}\tilde U(t,x/|x|).
2805: \]
2806: Fixed $t^*$ (in this section we will consider a locally minimal trajectory $\bar x$ 
2807: with a collision at $t^*$) in this section, with an abuse of notation, we denote
2808: \begin{equation}
2809: \label{eq:tildeUristr}
2810: \tilde U(x)=\tilde U(t^*,x).
2811: \end{equation}
2812: Of course, the function $\tilde U$ is homogeneous of degree $-\alpha$
2813: on $\RR^{k}\minus\Delta$.
2814: \begin{theorem}\label{theo:nocoll}
2815: In addition to \ref{a:U0}, \ref{a:U1}, \ref{a:U2h}, \ref{a:U3h}, \ref{a:U4h}, \ref{a:U5}, 
2816: assume that, for a given $\xi\in\Delta$ 
2817: \begin{assiomi}
2818: \item[\useitag]
2819: \label{a:U6}
2820: there is a $2$--dimensional linear subspace of $V_{\mu(\xi)}^{\perp}$, say $W$,
2821: where $\tilde U$ is rotationally invariant;
2822: %\[\tilde U(e^{i\theta}w)=\tilde U(w) , \qquad\forall w\in W, \forall \theta\in[0,2\pi];\]
2823: \item[\usettehtag]
2824: \label{a:U7h}
2825: for every $x\in\RR^k$ and $\delta \in W$ there holds
2826: \[
2827: \tilde U(x+\delta) \leq \tilde U\left(\left(\dfrac{\tilde U(\pi_W(x))}{\tilde U(x)}\right)^{1/\alpha}\pi_W(x)+\left(\dfrac{\tilde U(x)}{\tilde U(\pi_W(x))}\right)^{1/\alpha}\delta\right)
2828: \]
2829: where $\pi_W$ denotes the orthogonal projection onto $W$.
2830: \end{assiomi}
2831: Then generalized solutions do not have collisions at the 
2832: configuration $\xi$ at the time $t^*$.
2833: \end{theorem}
2834: 
2835: \begin{figure}
2836: \caption{Potential levels, with $\lambda=\left(\dfrac{\tilde U(\pi_W(x))}{\tilde U(x)}\right)^{1/\alpha}>1$}
2837: \begin{center}
2838: \psfrag{x}[c]{$x$}
2839: \psfrag{xpd}{$x+\delta$}
2840: \psfrag{x}[c]{$x$}
2841: \psfrag{W}[c]{$W$}
2842: \psfrag{px}[c]{$\pi_W(x)$}
2843: \psfrag{xpx}{$\lambda\pi_W(x)+\lambda^{-1}\delta$}
2844: \psfrag{lampx}[c]{$\lambda\pi_W(x)$}
2845: \psfrag{Ux}[c]{$\tilde U(x)$}
2846: \psfrag{Ulamx}[c]{$\tilde U(\lambda \pi_W(x) + \lambda^{-1}\delta)$}
2847: \psfrag{del}{$\delta$}
2848: \psfrag{lamdel}{$\lambda^{-1}\delta$}
2849: \includegraphics[height=0.3\textheight]{ovale}
2850: \end{center}
2851: \end{figure}
2852: 
2853: 
2854: \begin{remark}
2855: Some comments on assumptions \ref{a:U6} and \ref{a:U7h} are in order. Of course, as our potential 
2856: $\tilde U$ is homogeneous of degree $-\alpha$ the function 
2857: \[
2858: \varphi(x)=\tilde U^{-1/\alpha}(x)
2859: \]
2860: is a non negative, homogeneous of degree one function, having now $\Delta$ as zero set. In most 
2861: of our  applications $\varphi$ will be indeed a quadratic form.  
2862: Assume that $\varphi^2$ splits in the following way:
2863: \[
2864: \varphi^2(x)=K|\pi_W(x)|^2+\varphi^2(\pi_{W^\perp}(x))
2865: \]
2866: for some positive constant $K$.
2867: Then \ref{a:U6} and \ref{a:U7h} are satisfied. Indeed, denoting $w=\pi_W(x)$ and $z=x-w$ we have,
2868: for every $\delta \in W$,
2869: \[
2870: \begin{split}
2871: \varphi^2(x+\delta)&=K\vert w+\delta\vert^2+\varphi^2(z)\\
2872: & = K\left|{\frac{\varphi(x)}{\varphi(w)}}w +
2873: {\frac{\varphi(w)}{\varphi(x)}}\delta\right|^2 + 
2874: K\frac{\varphi^2(z)}{\varphi^2(x)}|\delta|^2\\
2875: &\geq K\left|{\frac{\varphi(x)}{\varphi(w)}}w +
2876: {\frac{\varphi(w)}{\varphi(x)}}\delta\right|^2\\
2877: & = \varphi^2\left({\frac{\varphi(x)}{\varphi(w)}}w +
2878: {\frac{\varphi(w)}{\varphi(x)}}\delta\right),
2879: \end{split}
2880: \]
2881: which is obviously equivalent to \ref{a:U7h}. Therefore, we have the following Proposition.
2882: \end{remark}
2883: \begin{proposition}\label{prop:usette}
2884: Assume $\tilde U(x)=\mathcal Q^{-\alpha/2}(x)$ for some non negative quadratic form 
2885: $\mathcal Q(x) = \langle Ax,x\rangle$. Then assumptions~\ref{a:U6} and~\ref{a:U7h} are satisfied whenever $W$ is 
2886: included in an eigenspace of $A$ associated with a multiple eigenvalue.
2887: \end{proposition}
2888: 
2889: \begin{remark}
2890: Given two potentials satisfying \ref{a:U6} and  \ref{a:U7h}  for a common
2891: subspace $W$, their sum enjoys the same properties. On the other hand, if they
2892: do not admit a common subspace $W$, their sum does not satisfy \ref{a:U6} and
2893: \ref{a:U7h}.
2894: \end{remark}
2895: 
2896: % Let us start the proof or Theorem~\ref{theo:nocoll}.
2897: \begin{proof}[Proof or Theorem~\ref{theo:nocoll}]
2898: % Assume not and 
2899: Let  $\bar x(t)$ be a generalized solution with a collision at the
2900: time $t^*$, i.e.~ $\bar x(t^*)=\xi \in \Delta$; 
2901: up to time-translation we assume that  the collision
2902: instant is $t^*=0$. 
2903: Furthermore, using the same arguments needed in the proof of the Von Zeipel's
2904: Theorem in Section~\ref{sec:partial}, we can suppose that $\xi=0$.  We consider
2905: the case of a boundary collision  (interior collisions can be treated in a
2906: similar way).  Then Theorem~\ref{central_conf} 
2907: ensures the existence of $\delta_0>0$ such that no other collision occurs in some interval
2908: $[0,\delta_0]$.
2909: 
2910: We consider the family of rescaled generalized solutions 
2911: \[
2912: \bar x^{\lambda_n}(t) := \lambda_n^{-\frac{2}{2+\alpha}}\bar x(\lambda_n t),
2913: \quad t \in [0,\delta_0/\lambda_n].
2914: \]
2915: where  $\lambda_n \rightarrow 0$ as $n \rightarrow +\infty$. 
2916: From the asymptotic estimates of Theorem~\ref{central_conf2} we know that 
2917: the angular part $(s(\lambda_n))_n$ converges, up to subsequences, to some central 
2918: configuration $\bar s$, in particular $\bar s$ is in the $\omega$--limit of $s(t)$.
2919: 
2920: For any $\bar s$ in the $\omega$--limit of $s(t)$, 
2921: a (right) blow-up of $\bar x$ in $t=0$ is a path defined, for $t \in [0,+\infty)$, as
2922: \begin{equation}
2923: \label{eq:blow-up}
2924: \bar q(t) := \zeta t^\frac{2}{2+\alpha},\qquad\zeta=K\bar s,
2925: \end{equation}
2926: where the constant $K>0$ is determined by part \emph{(b)} of
2927: Theorem~\ref{central_conf}. We note that the blow--up is a homothetic solution to the 
2928: dynamical system associated with the homogeneous potential $\tilde U$ and that 
2929: it has zero energy (the blow--up is parabolic).
2930: If $s(\lambda_n) \to \bar s$ as $n \to+\infty$,
2931: from Theorem~\ref{central_conf}, we obtain straightforwardly 
2932: the pointwise convergence of $\bar x^{\lambda_n}$ to the blow up $\bar q$ and the
2933: $H^1$-boundedness of $\bar x^{\lambda_n}$ implies
2934: its uniform convergence on compact subsets of $[0,+\infty)$. Furthermore 
2935: the convergence holds locally in the $H^1([0,+\infty))$--topology. Finally
2936: also the 
2937: sequence $\dot {\bar x}^{\lambda_n}$ converges uniformly on every interval $[\eps,T]$,
2938: with arbitrary $0<\eps<T$.
2939: 
2940: The following fact has been proven in \cite{FT}, Proposition 7.9.
2941: 
2942: \begin{lemma}
2943: Let $\bar x$ be a locally minimizing trajectory with a total collision at $t=0$ and let $\bar 
2944: q$ be its blow--up in $t=0$. Then $\bar q$ is a locally minimizing trajectory for the 
2945: dynamical system associated with the homogeneous potential $\tilde U$ introduced in 
2946: \eqref{eq:tildeUristr}.
2947: \end{lemma}
2948: 
2949: We will conclude the proof showing that $\bar q$ cannot be a locally 
2950: minimizing trajectory for the  dynamical system associated with $\tilde U$.
2951: Following \cite{FT}, we now introduce a class of suitable variations as follows:
2952: 
2953: \begin{definition}
2954: \label{def:stand_var}
2955: The standard variation associated with $\delta \in \RR^k\minus\{0\}$ is defined as 
2956: \begin{equation*}
2957: v^\delta(t)= \left\{
2958: \begin{array}{ll}
2959: \delta & \mbox{if} \,\,\, 0 \leq |t| \leq T-|\delta| \\
2960: (T-t)\frac{\delta}{|\delta|} & \mbox{if} \,\,\, T-|\delta| \leq |t| \leq T \\
2961: 0 & \mbox{if} \,\,\, |t| \geq T, \\
2962: \end{array}\right.
2963: \end{equation*}
2964: for some positive $T$.
2965: \end{definition}
2966: 
2967: We wish to estimate the action differential corresponding to a standard variation. 
2968: To this aim we give the next definition.
2969: \begin{definition} The displacement potential differential associated with $\delta \in \RR^k$ 
2970: is defined as:
2971: $$
2972: S(\zeta,\delta)=\int_0^{+\infty} 
2973: \left( \tilde U(\zeta t^{2/(2+\alpha)}+\delta) -
2974: \tilde U( \zeta t^{2/(2+\alpha)})  \right) dt
2975: $$
2976: where $\bar q(t) = \zeta t^{2/(2+\alpha)} $ is a blow-up of $\bar x$ in $t=0$.
2977: \end{definition}
2978: The quantity $S(\zeta,\delta)$ represents the \emph{potential differential} needed for 
2979: displacing the  colliding trajectory originarily traveling along the $\zeta$--direction  
2980: to  the point $\delta$.
2981: It has been proven in \cite{FT} Proposition 9.2, that the function $S$ represents 
2982: the limiting behavior, as $\delta\to0$, of the whole action differential:
2983: \[
2984: \Delta \mathcal{A}^{\delta} := 
2985: \int_{-\infty}^{+\infty} \left[K(\dot{\bar q}+\dot v^\delta)+\tilde U(\bar q+v^\delta) - 
2986: K(\dot{\bar q})-\tilde U(\bar q)\right] dt.
2987: \]
2988: Indeed, the fundamental estimate holds: 
2989: \begin{lemma}\label{lemma:act_decreas}
2990: Let $\bar q=  \zeta t^{2/(2+\alpha)}$ be a blow--up
2991: trajectory and $v^\delta$ any standard variation. 
2992: Then, as $\delta \rightarrow 0$
2993: \begin{equation*}
2994: \Delta {\mathcal A}^{\delta} 
2995: = |\delta|^{1-\alpha/2}  S\left(\zeta, \frac{\delta}{|\delta|}\right) + O(|\delta|).
2996: \end{equation*}
2997: \end{lemma}
2998: 
2999: We observe that, from the homogeneity of $\tilde U$ it follows that
3000: \begin{equation}\label{eq:homog}
3001: S(\lambda\xi,\mu\delta)=  
3002: \left|\lambda\right|^{-1-\alpha/2}|\mu|^{1-\alpha/2}S(\xi,\delta)
3003: \end{equation}
3004: (see \cite[(8.2)]{FT}) and hence, 
3005: if $\tilde U$ is invariant under rotations, the sign of $S$ depends only on the 
3006: angle between $\xi$ and $\delta$. To deal with the isotropic case (which is not the case 
3007: here), the following function was introduced in \cite{FT}:
3008:  $$\Phi_\alpha(\vartheta)=\int_0^{+\infty} 
3009: \frac{1}{\left( t^{\frac{4}{\alpha+2}} 
3010: -2\cos \vartheta t^{\frac{2}{\alpha+2}} 
3011: +1 \right)^{\alpha /2}} - 
3012: \frac{1}{ t^{\frac{2\alpha}{\alpha+2}}} dt.
3013: $$
3014: The value of $\Phi_\alpha(\vartheta)$ ranges from  positive to negative values, 
3015: depending on $\vartheta$ and $\alpha$. Nevertheless,  it is always negative, 
3016: when averaged on a circle. 
3017: Indeed, the following inequality was obtained in \cite[Theorem 8.4]{FT}.
3018: 
3019: \begin{lemma}\label{fund_ineq} For any $\alpha \in(0,2)$ there holds
3020: \[
3021: \dfrac{1}{2\pi}\int_0^{2\pi}\Phi_\alpha(\vartheta)d\vartheta<0.
3022: \]
3023: \end{lemma}
3024: 
3025: This inequality will be a key tool in proving the following averaged estimate:
3026: \begin{lemma}\label{lemma:ineq}
3027: Assume \ref{a:U6} and \ref{a:U7h}, then, if $\SSS$ is the unitary circle of $W$, for any 
3028: $\zeta\in\RR^k\minus\{0\}$ the following inequality holds
3029: \[
3030: % \frac{1}{|\SSS|}
3031: \int_{\SSS} S(\zeta,\delta)d\delta<0\;.
3032: \]
3033: As a consequence,
3034: \[
3035: \forall \zeta\in\RR^k\minus\{0\}\;\exists \delta=\delta(\zeta)\in\SSS\;:\; 
3036: S(\zeta,\delta(\zeta))<0\;.
3037: \]
3038: \end{lemma}
3039: \begin{proof} 
3040: As a first obvious application of Lemma~\ref{fund_ineq} we obtain the assertion \emph{for any 
3041: $\zeta\in W\minus \{0\}$}. Indeed, by~\eqref{eq:homog} and \ref{a:U6} we easily obtain
3042: \[
3043: \zeta\in W \minus \{0\} \quad \Longrightarrow \quad
3044: S(\zeta,\delta) = K\left|\zeta\right|^{-1-\alpha/2}\Phi_\alpha(\vartheta),
3045: \]
3046: where $K$ is a positive constant and $\vartheta$ denotes the angle between $\zeta$ 
3047: and $\delta$.
3048: 
3049: Now we prove the assertion for any $\zeta \neq 0$ in the configuration space.
3050: It follows from  the homogeneity of $\tilde U$ that
3051: \[
3052: \tilde U\left(\left(\dfrac{\tilde U(\pi_W(\zeta))}{\tilde 
3053: U(\zeta)}\right)^{1/\alpha}\pi_W(\zeta)\right)=\tilde U(\zeta).
3054: \]
3055: Hence \ref{a:U7h} implies, for every $\delta\in\SSS$,
3056: \[S(\zeta,\delta)\leq S\left(\left(\dfrac{\tilde U(\pi_W(\zeta))}{\tilde 
3057: U(\zeta)}\right)^{1/\alpha}\pi_W(\zeta),\left(\dfrac{\tilde U(\zeta)}{\tilde 
3058: U(\pi_W(\zeta))}\right)^{1/\alpha}\delta\right).
3059: \]
3060: Hence~\eqref{eq:homog} implies
3061: \[S(\zeta,\delta)\leq \left(\dfrac{\tilde U(\zeta)}{\tilde U(\pi_W(\zeta))}\right)^{2/\alpha} 
3062: S(\pi_W(\zeta),\delta),
3063: \]
3064: and thus 
3065: \[
3066: % \frac{1}{|\SSS|}
3067: \int_{\SSS} S(\zeta,\delta)\,d\delta
3068: \leq 
3069: \left(\dfrac{\tilde U(\zeta)}{\tilde 
3070: U(\pi_W(\zeta))}\right)^{2/\alpha}
3071: % \frac{1}{|\SSS|}
3072: \int_{\SSS} 
3073: S(\pi_W(\zeta,\delta))\,d\delta<0\;.
3074: \]
3075: \end{proof}
3076: \noindent\emph{End of the Proof of Theorem~\ref{theo:nocoll}.}  
3077: To conclude the proof, according with Lemma~\ref{lemma:ineq} we chose 
3078: $\delta=\delta(\zeta)\in  W\minus\{0\}$ with the property that 
3079: $S(\zeta,\delta(\zeta)/|\delta(\zeta)|)<0$. 
3080: As a consequence of Lemma~\ref{lemma:act_decreas}, we can 
3081: lower the value of the action of $\bar q$ by performing the standard variation 
3082: $v^{\delta(\zeta)}$, provided the norm of $|\delta(\zeta)|$ is sufficiently small
3083: (in order to apply Lemma \ref{lemma:act_decreas}). 
3084: Hence $\bar q$ can not be locally minimizing for the action.
3085: \end{proof}
3086: 
3087: As we have already noticed, the class of potentials satisfying \ref{a:U6} and \ref{a:U7h} is not 
3088: stable with respect to the sum of potentials. In order to deal with a class of potentials
3089: which is closed with respect to the sum, we introduce the following variant of Theorem 
3090: \ref{theo:nocoll}.
3091: 
3092: \begin{theorem}\label{theo:nocoll_2}
3093: In addition to \ref{a:U0}, \ref{a:U1}, \ref{a:U2h}, \ref{a:U3h}, \ref{a:U4h}, \ref{a:U5}, 
3094: assume that $\tilde U$ has the form
3095: \[
3096: \tilde U(x)=\sum_{\nu=1}^N\frac{K_\nu}{\left({\rm dist}(x,V_\nu)\right)^\alpha}
3097: \]
3098: where $K_\nu$ are positive constants and $V_\nu$ is a family of linear subspaces, 
3099: with ${\rm codim}(V_\nu)\geq 2$, for every $\nu=1,\dots,N$.
3100: Then locally minimizing trajectories do not have collisions at the time $t^*$.
3101: \end{theorem}
3102: 
3103: \begin{proof}
3104: Following the  arguments of the proof of Theorem~\ref{theo:nocoll}, the
3105: assertion will be proved once we show, as in Lemma~\ref{lemma:ineq}, that, for
3106: every index $\nu$, there holds
3107: \begin{equation*}\label{eq:int_sfera}
3108: % \frac{1}{|\SSS^{k-1}|}
3109: \int_{\SSS^{k-1}} S_\nu(\zeta,\delta) \,d\delta<0,
3110: \end{equation*}
3111: where, of course, we denote 
3112: \[
3113: S_\nu(\zeta,\delta)=\int_0^{+\infty} \left(
3114: {\rm dist}(\zeta t^{2/(2+\alpha)}+\delta, V_\nu)^{-\alpha} -
3115: {\rm dist}(\zeta t^{2/(2+\alpha)}, V_\nu)^{-\alpha} \right) dt
3116: \]
3117: and $\SSS^{k-1}$ is the unit sphere of the configuration space $\RR^k$. 
3118: This is an elementary consequence of Lemma \ref{lemma:ineq} and the  fact 
3119: that the function $S_\nu(\zeta,\delta)$ only depends on the projection of $\zeta$ 
3120: orthogonal to $V_\mu$ and has rotational invariance on $V_\nu^\perp$. 
3121: Thus the integral of $S_\nu$ over the sphere is a positive multiple of its integral 
3122: on any circle $\SSS$ orthogonal to $V_\nu$.
3123: \end{proof}
3124: 
3125: %==============================
3126: \subsection{Logarithmic type potentials}
3127: \label{sec:blowup_log}
3128: %==============================
3129: In this section we prove the equivalent to Theorems \ref{theo:nocoll} and \ref{theo:nocoll_2}
3130: suitable for logarithmic type potentials. Concerning the quasi-homogeneous 
3131: case we have seen that a crucial role is played by the construction of a blow--up 
3132: function which minimizes a limiting problem.
3133: Before starting, let us highlight the reasons why, when dealing
3134: with logarithmic potentials,  a blow-up limit can not exist.
3135: Indeed, the natural scaling
3136: should be $\bar x^{\lambda_n}(t) := \lambda_n^{-1}\bar x(\lambda_n t)$,
3137: which does not converge, since
3138: \[
3139: \lim_{\lambda_n \to 0} \bar x^{\lambda_n}(t) =
3140: \lim_{\lambda_n \to 0} \frac{r(\lambda_n t)s(\lambda_n t)}
3141:                            {\lambda_n t\sqrt{-2M(0)\log(\lambda_n t)}} 
3142: t\sqrt{-2M(0)\log(\lambda_n t)}
3143: =+\infty
3144: \]
3145: for every $t>0$. On the other, hand, looking at~\ref{eq:blow-up},
3146: the (right) blow--up should be, up to a change of time
3147: scale,
3148: \begin{equation}
3149: \label{eq:blow-up_log}
3150: \bar q(t) := t\bar s, \quad i \in {\bf k},
3151: \end{equation}
3152: where $\bar s$ is a central configuration for the system
3153: limit of a sequence $s(\lambda_n)$ where  $(\lambda_n)_n$ is such that
3154: $\lambda_n \rightarrow 0$.
3155: The blow up function defined in~(\ref{eq:blow-up_log})
3156: is the pointwise limit of the normalized sequence 
3157: \[
3158: \bar x^{\lambda_n}(t) := \frac{1}{\lambda_n \sqrt{-2M(0)\log\lambda_n}}\bar
3159: x(\lambda_n t).
3160: \]
3161: Unfortunately the path  in~(\ref{eq:blow-up_log}) is not locally minimal for the
3162: limiting problem, indeed since, the sequence $(\ddot{\bar x}^{\lambda_n})_n$ 
3163: converges to $0$ as $n$ tends to $+\infty$, the blow-up in~(\ref{eq:blow-up_log})
3164: minimizes only the kinetic part of the action functional.
3165: 
3166: We shall overcome this difficulty by proving the averaged estimate 
3167: in a direct way from the asymptotic  estimates of Theorem~\ref{central_conf_log}
3168: and assuming~\eqref{eq:hp_pot_log} on the potential $U$.
3169: As we have done for the quasi--homogeneous case, we extend the function $\tilde U$,
3170: introduced in assumption \ref{a:U3l}, to the whole $(a,b)\times \RR^k\minus \Delta$
3171: in the natural way 
3172: \begin{equation}
3173: \label{eq:tildeUlog}
3174: \tilde U(t,x) = \tilde U(t,s) - M(t)\log|x|,
3175: \end{equation}
3176: where $M$ has been introduced in \eqref{eq:M}.
3177: \begin{theorem}\label{theo:nocoll_log}
3178: In addition to \ref{a:U0}, \ref{a:U1}, \ref{a:U2l}, \ref{a:U3l}, \ref{a:U4l}, \ref{a:U5} 
3179: assume the potential $U$ to be of the form
3180: \begin{equation}
3181: \label{eq:hp_pot_log}
3182: U(t,x) = \tilde U(t,x) + W(t,x)
3183: \end{equation}
3184: where $\tilde U$ satisfies \eqref{eq:tildeUlog}
3185: and $W$ is a bounded $\cont^1$ function on $(a,b) \times \RR^{k}$.
3186: Furthermore assume that, for a given $\xi\in\Delta$ , $\tilde U$ satisfies \ref{a:U6} and
3187: \begin{assiomi}
3188: \item[\usetteltag]
3189: \label{a:U7l}
3190: for every $x\in\RR^k$ and $t \in (a,b)$ there holds
3191: \[
3192: \tilde U(t,x) = -\frac12 M(t)\log\left(|\pi_Wx|^2 + \psi^2(\pi_{W^\perp}x)\right)
3193: \]
3194: where $\pi_W$ and $\pi_{W^\perp}$ denote the orthogonal projections onto $W$ and $W^\perp$,
3195: $\psi$ is $\cont^1$ and 
3196: homogeneous of degree $1$.
3197: % $\psi(0)=0$.
3198: \end{assiomi}
3199: Then locally minimizing trajectories do not have collisions at the configuration $\xi$ at the time $t^*$.
3200: \end{theorem}
3201: 
3202: \begin{proof}
3203: As in the proof of Theorem~\ref{theo:nocoll},
3204: we consider a generalized solution $\bar x$ and we first reduce to 
3205: the case of an isolated total collision at the origin occurring at the time $t=0$.
3206: From Theorem~\ref{central_conf_log} we deduce the existence of $\delta_0>0$
3207: such that no other collision occur in $[-\delta_0,\delta_0]$, hence 
3208: we perform a local variation on the trajectory of $\bar x$
3209: that removes the collision and makes the action decrease.
3210: 
3211: % To this aim, we first observe that,
3212: % from Theorem~\ref{central_conf_log}, it follows that 
3213: % $r(t)=|\bar x(t)|$ is asymptotic to $t \sqrt{-2M_0 \log t}$ as $t$ tends to $0^+$; 
3214: % since $t \sqrt{-2M_0 \log t}>t$ whenever $t<e^{-1/2M_0}$, we can choose $\delta_0$ small enough
3215: % and $\eps <\delta_0/2$, to have
3216: % \begin{equation}\label{eq:>eps}
3217: % r(t) > \eps \qquad \mbox{for every } t \in [\delta_0-\eps,\delta_0].
3218: % \end{equation}
3219: % Fixed $\eps >0$ verifying \eqref{eq:>eps}
3220: % and $\delta \in \RR^d$ such that $|\delta|=\eps$, 
3221: Consider now the 
3222: {standard variation} $v^\delta$, defined at page \pageref{def:stand_var},
3223: on the interval $[0,\delta_0]$ (i.e., in Definition \ref{def:stand_var} 
3224: $T$ is replaced by $\delta_0$).
3225: % \begin{equation}
3226: % \label{eq:variation_mu}
3227: % v^{\delta,\eps}:[0,\delta_0]\to\RR^{k}, \qquad
3228: % v^{\delta,\eps}(t) := 
3229: % \left\{
3230: % \begin{array}{ll}
3231: % \xi,                                        & \mbox{if } t \in [0,\delta_0-\eps]  \\
3232: % \displaystyle(\delta_0-t)\frac{\xi}{\eps},  & \mbox{if } t \in [\delta_0-\eps,\delta_0]
3233: % \end{array}
3234: % \right.
3235: % \end{equation}
3236: % where $\xi = (\eps\delta,0,\ldots,0) \in \RR^{k}$.
3237: %
3238: Let $\Delta^{\delta} {\cal A}$ denote the difference 
3239: $$
3240: \Delta^{\delta} {\cal A} :=
3241: {\cal A}(\bar x + v^{\delta},[0,\delta_0]) - {\cal A}(\bar x,[0,\delta_0]);
3242: $$
3243: generally speaking, this difference can be positive or negative, depending
3244: on the choice of $\delta$.
3245: Our goal is to prove that, when \emph{averaging over a suitable set of standard 
3246: variations}, the action lowers.
3247: Hence  $\Delta^{\delta} {\cal A}$
3248: must be negative for at least one choise of $\delta$ and the path $\bar x$ 
3249: can not be a local minimizer for the action.
3250: 
3251: We can write $\Delta^{\delta} {\cal A}$ as the sum of three terms
3252: \begin{equation}
3253: \label{eq:A_var_mu}
3254: \Delta^{\delta} {\cal A} = \int_{0}^{\delta_0} \Delta^{\delta}{\cal K}(t) \,dt 
3255:                    + \int_0^{\delta_0} \Delta^{\delta}{\cal U}(t) \,dt
3256:                    + \int_{0}^{\delta_0} \Delta^{\delta}{\cal W}(t) \,dt 
3257: \end{equation}
3258: where $\Delta^{\delta}{\cal K}(t)$, $\Delta^{\delta}{\cal U}(t)$ and
3259: $\Delta^{\delta}{\cal W}(t)$ are respectively the variations of the kinetic energy,
3260: of the singular potential $\tilde U$ and of the smooth part of the potential, $W$.
3261: More precisely  since the first derivative of the function $v^{\delta}$
3262: vanishes everywhere on $[0,\delta_0]$, except on $[\delta_0-|\delta|,\delta_0]$, we compute
3263: \begin{equation}
3264: \label{eq:K_var_mu}
3265: \Delta^{\delta}{\cal K}(t) :=
3266: \left\{
3267: \begin{array}{ll}
3268: \displaystyle
3269: 0, & \mbox{if } t \in [0,\delta_0-|\delta|], \\
3270: \displaystyle\frac{1}{2}( \left| \dot{\bar x} - \delta/|\delta|\right|^2 - |\dot{\bar x}|^2 )
3271:    & \mbox{if } t \in [\delta_0-|\delta|,\delta_0].
3272: \end{array}
3273: \right.
3274: \end{equation}
3275: Similarly
3276: \[
3277: \Delta^{\delta}{\cal U}(t) := \tilde U(t,\bar x+v^{\delta}) - \tilde U(t,\bar x)
3278: \qquad \mbox{and}
3279: \qquad \Delta^{\delta}{\cal W}(t) := W(t,\bar x+v^{\delta}) - W(t,\bar x).
3280: \]
3281: We now evaluate separately the mean values 
3282: of the tree terms of $\Delta^{\delta} {\cal A}$ over the circle $S^{|\delta|}$ of radius $|\delta|$ in $W$.
3283: \begin{lemma}
3284: \label{le:media_cin}
3285: There holds
3286: \begin{equation}
3287: \frac{1}{2\pi|\delta|}\int_{S^{|\delta|}} \int_0^{\delta_0}
3288: (\Delta^{\delta}{\cal K} + \Delta^{\delta} {\cal W})\,dt\,d\delta = O(|\delta|).
3289: \end{equation}
3290: \end{lemma}
3291: \begin{proof}
3292: From~\eqref{eq:K_var_mu} we obtain
3293: \begin{equation*}
3294: \int_0^{\delta_0} \Delta^\delta{\cal K}(t) dt =
3295: \int_{\delta_0-|\delta|}^{\delta_0}
3296: \frac{1}{2}(\left| \dot{\bar x} -\delta/|\delta|\right|^2 - |\dot{\bar x}|^2) dt
3297: = \frac{1}{2}\left(|\delta|-2\int_{\delta_0-|\delta|}^{\delta_0}
3298: \dot{\bar x}(t)\cdot\frac{\delta}{|\delta|} dt \right),
3299: \end{equation*}
3300: hence 
3301: \begin{equation*}
3302: \left|\int_0^{\delta_0} \Delta^\delta{\cal K}(t) dt \right| \leq O(|\delta|),
3303: \end{equation*}
3304: which does not depend on the circle $S^{|\delta|}$ where $\delta$ varies. 
3305: Concerning the variation of the $\cont^1$ function $W$ we have
3306: \begin{equation*}
3307: \begin{split}
3308: \left| \int_0^{\delta_0} \Delta^\delta{\cal W}(t) \,dt \right| & =
3309: \left| \int_0^{\delta_0-|\delta|} \Delta^\delta{\cal W}(t) \,dt \right| +
3310: \left| \int_{\delta_0-|\delta|}^{\delta_0} \Delta^\delta{\cal W}(t) \right| \,dt \\
3311: & \leq W_1 |\delta| (\delta_0-|\delta|) + 2W_2 |\delta| = O(|\delta|),
3312: \end{split}
3313: \end{equation*}
3314: where $W_1$ is a bound for 
3315: $\left|\frac{\partial W}{\partial x}(t,\bar x+\lambda v^{\delta})\right|$, 
3316: with $\lambda \in [0,1]$ and $t \in [0,\delta_0-|\delta|]$ while $W_2$ is an upper bound for
3317: $|W(t,x)|$.
3318: \end{proof}
3319: 
3320: In order to estimates the variation of the potential part,
3321: $\Delta^\delta{\cal U}(t)$, we prove the next two technical lemmata.
3322: Let us start with recalling an equivalent version of the mean value property for the
3323: fundamental solution of the planar Laplace equation.
3324: 
3325: \begin{lemma}
3326: \label{le:pot:2}
3327: Fixed $z>0$, for every $y\in\RR$ such that $y\geq 2z$, we have 
3328: \[
3329: \frac{1}{2\pi }\int_{0}^{2\pi} \log(y+2z\cos\vartheta) \,d\vartheta =
3330: \log\frac{y+\sqrt{y^2-4z^2}}{2}.
3331: \]
3332: \end{lemma}
3333: \begin{proof}
3334: Since $y\geq 2z$, then $\frac{y+\sqrt{y^2-4z^2}}{2}\geq z$. Let $x\in\RR^2$ be such
3335: that 
3336: $|x|=\frac{y+\sqrt{y^2-4z^2}}{2}$, then $y=(|x|^2+z^2)/|x|$ and for every $\delta \in S^z$, 
3337: where $S^z$ is the circle of radius $z$, we have
3338: \[
3339: \begin{split}
3340: |x+\delta|^2 &= |x|^2 + z^2 +2z|x|\cos \vartheta \\
3341:              &= |x|\left( \frac{|x|^2 + z^2}{|x|} + 2z\cos \vartheta\right) = 
3342:                 |x|(y + 2z\cos \vartheta).
3343: \end{split}
3344: \]
3345: We have, as the logarithm is the fundamental solution to the Laplace equation on the plane,
3346: \begin{equation}
3347: \label{le:soluz_fond}
3348: \frac{1}{2\pi z}\int_{S^z} \log|x+\delta|^2 \,d\delta = \max\{\log|x|^2, \log z^2\} = 
3349: \left\{
3350: \begin{array}{ll}
3351: \log|x|^2 ,        & \mbox{if }|x| >    z \\
3352: \log z^2 ,  & \mbox{if }|x| \leq z.
3353: \end{array}
3354: \right..
3355: \end{equation}
3356: Consequently, when computing
3357: \[
3358: \begin{split}
3359: \int_{S^z}\log|x+\delta|^2 d\delta &= \int_{S^z}\log|x| d\delta +
3360: z\int_0^{2\pi}\log(y + 2z\cos \vartheta)d\vartheta \\
3361: &= 2\pi z\log|x| + z\int_0^{2\pi}\log(y + 2z\cos \vartheta)d\vartheta,
3362: \end{split}
3363: \]
3364: we find
3365: \[
3366: 2\pi z \log|x|^2 = 2\pi z\log|x|+z\int_0^{2\pi}\log(y + 2z\cos \vartheta)d\vartheta.
3367: \]
3368: We conclude replacing $|x|=\frac{y+\sqrt{y^2-4z^2}}{2}$.
3369: \end{proof}
3370: 
3371: Now we consider the averages of the potential with respect to a circle in $W$ (here we assume implicitly that $d\geq 3$).
3372: 
3373: \begin{lemma}
3374: \label{le:pot:3}
3375: Fixed $|\delta|>0$, for every circle of radius $|\delta|$, $S^{|\delta|} \subset W$, 
3376: for every $x\in\RR^d$ and every $t \in [0,\delta_0]$, there holds
3377: \[
3378: \frac{1}{2\pi|\delta|}\int_{S^{|\delta|}} \left(\tilde U(x+\delta)-\tilde U(x)\right)\,d\delta  \leq 
3379: % \left{
3380: % \begin{array}{l}
3381: \begin{cases}
3382: 0    \text{\ \ ( if $\ds |\pi_W x|^2+\psi^2(\pi_{W^\perp} x) > |\delta|^2$)},\\
3383: \frac{M(t)}{2}\log(|\pi_W x|^2+\psi^2(\pi_{W^\perp} x))-\log(|\delta|^2) \text{\ \ (otherwise).}
3384: \end{cases}
3385: % \end{array}
3386: % \right.
3387: \]
3388: \end{lemma}
3389: \begin{proof}
3390: We consider the orthogonal decomposition of $x$, 
3391: $x=\pi_W x+\pi_{W^\perp} x$, and we term $u:=|\pi_W x|$ and $\eps:=\psi(\pi_{W^\perp} x)$.
3392: Since whenever $\delta \in W$ we have
3393: \[
3394: |\pi_W (x+\delta)|^2+\psi^2(\pi_{W^\perp} x) = u^2  + |\delta|^2 +2u|\delta|\cos \vartheta + \eps^2\geq 0,
3395: \]
3396: when $\cos \vartheta = -1$ we have $\frac{u^2  + |\delta|^2 + \eps^2}{u|\delta|} \geq 2$
3397: and, using Lemma \ref{le:pot:2} and equation \eqref{le:soluz_fond}, we compute
3398: \[
3399: \begin{split}
3400: \frac{1}{2\pi |\delta|}&\int_{S^{|\delta|}} \log(|\pi_W (x+\delta)|^2+\psi^2(\pi_{W^\perp} x)) \,d\delta \\
3401: & =\frac{1}{2\pi}\int_0^{2\pi } \log(u^2+\eps^2+|\delta|^2+2u|\delta|\cos \vartheta)
3402: \,d\vartheta\\
3403: & =\frac{1}{2\pi}\int_0^{2\pi} \log\left(\frac{u^2 + \eps^2 + |\delta|^2}{u|\delta|} + 2\cos
3404: \vartheta\right) \,d\vartheta + \log (u|\delta|)\\
3405: & =\log\left(\frac{u^2 + \eps^2 + |\delta|^2 + \sqrt{(u^2 + \eps^2 +
3406: |\delta|^2)^2-4u^2|\delta|^2}}{2}\right)\\
3407: & \geq\log\left(\frac{u^2 + \eps^2 + |\delta|^2 + \sqrt{(u^2 + \eps^2 + |\delta|^2)^2 - 4u^2|\delta|^2 -
3408: 4\eps^2|\delta|^2}}{2}\right)\\
3409: & = \log \left(\frac{u^2 + \eps^2 + |\delta|^2 + |u^2 + \eps^2 - |\delta|^2|}{2}\right)\\
3410: & = \max\left( \log(|\pi_W x|^2+\psi^2(\pi_{W^\perp} x),\log(|\delta|^2)\right)
3411: \end{split}
3412: \]
3413: and the assertion easily follows.
3414: \end{proof}
3415: 
3416: \begin{lemma}
3417: \label{le:media_pot1}
3418: Let $S$ be the circle of radius $|\delta|$ on $W$;
3419: then, as $|\delta| \rightarrow 0$
3420: \begin{equation}
3421: \frac{1}{2\pi|\delta|}\int_{S^{|\delta|}} \int_0^{\delta_0} \Delta^{\delta} {\cal U} \, dt \, d\delta 
3422: < -K|\delta|\sqrt{-\log|\delta|}, \qquad K>0.
3423: \end{equation}
3424: \end{lemma}
3425: 
3426: \begin{proof}
3427: Let $S^{|\delta|}$ be the circle of radius $|\delta|$ on $W$,
3428: we apply Fubini-Tonelli's Theorem and we argue as in the proof of 
3429: Lemma~\ref{le:pot:3} to have
3430: \[
3431: \begin{split}
3432: & \frac{1}{2\pi|\delta|}\int_{S^{|\delta|}} \int_{0}^{\delta_0} \Delta^{\delta}{\cal U}(t) dt \,  d\delta
3433: = \int_{0}^{\delta_0} \frac{1}{2\pi|\delta|} 
3434: \int_{S^{|\delta|}} \tilde U(\bar x+v^\delta) - \tilde U(\bar x) \, d\delta dt\\
3435: &= \frac{M^*}{2} \int_{0}^{\delta_0} 
3436: \left\{-\max\left[ \log(|\pi_W\bar x|^2 + \psi^2(\pi_{W^\perp}\bar x)), \log |v^\delta|^2\right]
3437: + \log(|\pi_W\bar x|^2 + \psi^2(\pi_{W^\perp}\bar x))\right\}\, dt
3438: \end{split}
3439: \]
3440: where $M^* = \max_t |M(t)|$. We then straightforwardly deduce that, for every $S^{|\delta|}\subset W$
3441: \[
3442: \frac{1}{2\pi|\delta|}\int_{S^{|\delta|}} \int_{0}^{\delta_0} \Delta^{\delta}{\cal U}(t) 
3443: dt \,  d\delta < 0.
3444: \]
3445: In order to estimate more precisely this quantity, we observe that 
3446: \begin{equation}\label{eq:le:media_pot1}
3447: \int_{0}^{\delta_0} \frac{1}{2\pi|v^\delta|} 
3448: \int_{S^{|v^\delta|}} \tilde U(\bar x+v^\delta) - \tilde U(\bar x) \, d\delta dt
3449: \leq
3450: \int_A 
3451: \log\frac{|\pi_W\bar x|^2 + \psi^2(\pi_{W^\perp}\bar x)}{|\delta|^2} \, dt%d\delta dt
3452: \end{equation}
3453: where
3454: \[
3455: A := \left\{ t \in [0,\delta_0-|\delta|] : |\pi_W\bar x|^2 + \psi^2(\pi_{W^\perp}\bar x) < |\delta|^2\right\}.
3456: \]
3457: Furthermore, there exists  a strictly positive constant $C$ such that
3458: \[
3459: C r^2 < |\pi_W x|^2 + \psi^2(\pi_{W^\perp} x) < C^{-1} r^2
3460: \]
3461: where, as usual, we denote $r^2=|\pi_W x|^2 + |\pi_{W^\perp} x|^2$ the radius of $x$. 
3462: The left inequality follows from Theorem~\ref{central_conf_log} indeed the existence of
3463: a finite limit of $\tilde U(t,s(t))$ prevents the projection $|\pi_W x|^2$
3464: and the function $\psi^2(\pi_{W^\perp} x)$ to be both infinitesimal with  $r^2$.
3465: The right inequality follows from the continuity of $\psi$.
3466: From \eqref{eq:le:media_pot1} and the asymptotic estimates of Theorem~\ref{central_conf_log} 
3467: we conclude that, as $|\delta| \rightarrow 0$
3468: \[
3469: \begin{split}
3470: \frac{1}{2\pi|\delta|}\int_{S^{|\delta|}} \int_{0}^{\delta_0} \Delta^{\delta}{\cal U}(t) 
3471: dt \,  d\delta 
3472: & \leq \int_{_{t:r(t)<|\delta|/\sqrt{C}}}
3473: \log\frac{r^2(t)}{C|\delta|^2} \, dt \\
3474: & \sim \int_0^{|\delta|/\sqrt{C}}
3475: 2\frac{\log (r/\sqrt{C}|\delta|)}{-\sqrt{-\log r}}  dr \\
3476: % & < \int_0^{|\delta|/\sqrt{C}}
3477: % 2\frac{\log r}{-\sqrt{-\log r}}  dr \\
3478: & < - 2\int_0^{|\delta|/\sqrt{C}} \sqrt{-\log r} dr
3479: <  - K|\delta| \sqrt{-\log |\delta|}
3480: \end{split}
3481: \]
3482: for some positive $K$, since $-\sqrt{-\log r}$ is an increasing function on the interval
3483: $[0,|\delta|]$.
3484: \end{proof}
3485: 
3486: \noindent\emph{End of the Proof of Theorem \ref{theo:nocoll}.}
3487: Let $S^{|\delta|}$ be a circle in $W$ with radius $|\delta|$ and $\Delta^{\delta} {\cal A}$
3488: the variation of the action functional defined in~(\ref{eq:A_var_mu}), then 
3489: from Lemmata \ref{le:media_cin} and  \ref{le:media_pot1}
3490: we conclude that, as $|\delta|$ tends to $0$
3491: \[
3492: \frac{1}{2\pi|\delta|}\int_{S^{|\delta|}} \Delta^{\delta} {\cal A} d\delta \leq O(|\delta|) 
3493: - K|\delta| \sqrt{-\log |\delta|} < 0.
3494: \]
3495: \end{proof}
3496: 
3497: %===========================%
3498: % DIM. PROPRIETA' LOGARITMO %
3499: %===========================%
3500: % \begin{remark}
3501: % From Theorem \ref{thm:var<0_mu} we easily deduce that, as $\eps$ tends to $0$,
3502: % \begin{equation*}
3503: % \frac{1}{4\pi}\int_{S^2} \Delta^{\delta,\eps} {\cal A} \, d\delta < 0,
3504: % \end{equation*}
3505: % where $S^2 = \{ z  \in \RR^3 : |z|=1 \}$; indeed
3506: % \begin{equation*}
3507: % \frac{1}{4\pi}\int_{S^2} \Delta^{\delta,\eps} {\cal A} \, d\delta 
3508: % = \frac{1}{4\pi} \int_0^\pi \sin \varphi \int_{S^1} \Delta^{\delta,\eps} {\cal A}
3509: % \, d\delta \, d\varphi 
3510: % \leq \frac{1}{4} \int_{S^1} \Delta^{\delta,\eps} {\cal A} \, d\delta < 0.
3511: % \end{equation*}
3512: % \end{remark}
3513: 
3514: Of course, likewise to Theorem~\ref{theo:nocoll_2}, there holds
3515: 
3516: \begin{theorem}\label{theo:nocoll_log_2}
3517: In addition to \ref{a:U0}, \ref{a:U1}, \ref{a:U2l}, \ref{a:U3l}, \ref{a:U4l}, \ref{a:U5}, assume $\tilde U$ be of the form
3518: \[
3519: \tilde U(x)=-\sum_{\nu=1}^N K_\nu\log\left({\rm dist}(x,V_\nu)\right)
3520: \]
3521: where $K_\nu$ are positive constants and $V_\nu$ is a family of linear subspaces, with ${\rm codim}(V_\nu)\geq 2$, 
3522: for every $\nu=1,\dots,N$.
3523: Then locally minimizing trajectories do not have collisions at the time $t^*$.
3524: \end{theorem}
3525: 
3526: %================================================================
3527: \subsection{Neumann boundary conditions and $G$--equivariant minimizers}
3528: %================================================================
3529: 
3530: As a final comment of this Section, we remark that, in our framework, the analysis allows 
3531: to prove that minimizers to the fixed--ends (Bolza) problems are free of collisions: indeed all 
3532: the variations of our class have compact support. However,
3533: other type  of boundary conditions (generalized Neumann) can be treated in the same way. 
3534: Indeed, consider a trajectory which is a (local) minimizer of the action among
3535: all paths satisfying the boundary conditions
3536: \[ x(0)\in X^0\qquad x(T)\in X^1,\]
3537: where $X^0$ and $X^1$ are two given linear subspaces of  the configuration
3538: space. Consider a (locally) minimizing  path $\bar x$: of course it has not
3539: interior collisions.  In order to exclude boundary collisions we have to be
3540: sure that the class of variations preserve the boundary condition; this can be
3541: achieved by restricting to $X^i$ the points $\delta$ appearing in the standard
3542: variations.  Hence, to complete the averaging argument, one needs
3543: assumptions~\ref{a:U6} and \ref{a:U7h} or \ref{a:U7l}  to be fulfilled also by
3544: the restriction of the potential to the boundary subspaces $X^i$.  This point
3545: of view differs from that of \cite{chen}, where the boundary subspaces can not
3546: be chosen arbitrarily in the configuration space. The argument in \cite{chen},
3547: already introduced in \cite{FT}, does not involve any averaging on the boundary
3548: but relies upon a suitable choice of a standard variation whose projection is
3549: extremal. 
3550: 
3551: The  analysis of boundary conditions was a key point in the paper \cite{FT},
3552: where symmetric periodic trajectories where constructed by reflections about
3553: given subspaces. 
3554: By Theorems~\ref{theo:nocoll} and~\ref{theo:nocoll_log}  one can obtain the
3555: absence of collisions also for $G$--equivariant (local) minimizers, provided
3556: the group $G$ satisfies the {\em Rotating Circle Property} introduced in
3557: \cite{FT} (see Example~\ref{exe:gruppo1}). Hence, existence of $G$--equivariant
3558: collisionless periodic solutions can be proved for the wide class of symmetry
3559: groups described in \cite{FT,F3,BFT}, for a much larger class of interacting
3560: potentials, 
3561: including quasi--homogeneous and logarithmic ones. On the other hand,
3562: Theorems~\ref{theo:nocoll_2} and~\ref{theo:nocoll_log_2} can be applied to
3563: prove that $G$--equivariant minimals are collisionless for many relevant
3564: symmetry groups violating the rotating circle property, such as the groups of
3565: rotations in \cite{F2}; indeed, the idea of averaging on spheres having maximal
3566: dimension has been borrowed from that paper (cf.~Example~\ref{exe:gruppo2}).
3567: 
3568: %The crucial assumptions for these extensions are \uc\, and \usei\enspace concerning the interaction 
3569: %between each pair of  particles, and~\eqref{eq:strucDelta} on the structure of the singular set.
3570: %Indeed they allow the reduction from partial collisions to total ones, in order to avoid 
3571: %each collision from the trajectory of an equivariant minimizer.
3572: 
3573: %----------------------------------
3574: \section{Examples and further remarks}\label{sec:comments}
3575: %----------------------------------
3576: 
3577: We now discuss various examples of classes of potentials which fullfill our assumptions.
3578: 
3579: \begin{example}[Homogeneous isotropic potentials]\label{ex:0}
3580: The simplest example of function satisfying  all our assumptions 
3581: \ref{a:U0},
3582: \ref{a:U1},
3583: \ref{a:U2h},
3584: \ref{a:U3h},
3585: \ref{a:U4h},
3586: \ref{a:U5},
3587: \ref{a:U6} and 
3588: \ref{a:U7h}
3589: is the $\alpha$-homogeneous one-center problem:
3590: \[
3591: U_\alpha(x) =  \frac{1}{|x|^\alpha},
3592: \]
3593: and its associated $n$--body problem:
3594: \[
3595: U_\alpha(x) = \sum_{\substack{i<j\\  i,j=1}}^n \frac{m_i m_j}{|x_i-x_j|^\alpha}.
3596: \]
3597: Assumptions \ref{a:U0} and \ref{a:U1} are trivially satisfied since $U$ is positive, diverges to 
3598: $+\infty$ when $x$ approaches $\Delta = \{x \in \RR^{nd}: x_i=x_j \mbox{ for some }i\neq j\}$,
3599: and does not depend on time.
3600: Furthermore in both \ref{a:U2} and \ref{a:U2h} the equality is achieved
3601: with $\tilde \alpha=\alpha$ and $C_2=0$. Since $U$ is homogeneous of degree $-\alpha$,
3602: in \ref{a:U3h} and \ref{a:U4h} the function $\tilde U$ coincides with $U$. \ref{a:U5} and \ref{a:U6} are trivially satisfied, while~\ref{a:U7h} holds by virtue of Proposition~\ref{prop:usette}
3603: \end{example}
3604: 
3605: \begin{example}[Logarithmic potentials]\label{ex:2}
3606: Our results apply also to logarithmic singularities of type
3607: \[
3608: U_{log}(x) = \sum_{\substack{i<j\\ i,j=1}}^{n} m_i m_j\log\frac{1}{|x_i-x_j|};
3609: \]
3610: indeed \ref{a:U2} is in this case satisfied for every value of\enspace $\tilde \alpha$
3611: and \ref{a:U2l}, \ref{a:U3l} and \ref{a:U4l} are verified with $C_2=0$.
3612: 
3613: Dynamical systems of type~\eqref{DS} with logarithmic interactions arise in the study of vortex flows in fluid mechanics, and,
3614: precisely, in the analysis of systems of $n$ 
3615: almost--parallel vortex filaments, under a linearized version of the LIA 
3616: self--interaction assumption (see \cite{KPV, KMD}).
3617: \end{example}
3618: 
3619: \begin{example}[Anisotropic $n$--body potentials] Consider potentials having the form
3620: \[
3621: U(t,x) = \sum_{\substack{i<j\\ i,j=1}}^{n} U_{i,j}(t,x_i-x_j),
3622: \]
3623: where the interaction potentials $U_{i,j}$ have a singularity at zero, of
3624: homogeneous or logarithmic type, but do depend on the angle. Typical examples
3625: are the Gutzwiller potentials \cite{Gu}. Notice that the total potential
3626: satisfies assumptions \ref{a:U0}, \ref{a:U1}, 
3627: and \ref{a:U2h}, \ref{a:U3h}, \ref{a:U4h} (or \ref{a:U2l}, \ref{a:U3l}, \ref{a:U4l}) provided each of the $U_{i,j}$'s
3628: do. It not difficult to see that also~\eqref{eq:strucDelta} and \ref{a:U5}
3629: hold (in the $n$--body case), while \ref{a:U6} and \ref{a:U7h} or \ref{a:U7l}
3630: do not. Hence we can not exclude the presence of collisions for locally
3631: minimizing paths, though the results about isolatedness and the asymptotic
3632: estimates are still available. More generally, we can deal with potentials of
3633: the form
3634: \[
3635: U_\alpha(r s) = r^{-\alpha}\tilde U(s),
3636: \]
3637: where $\tilde U:{\mathcal E}\setminus\Delta \to \RR$ is positive and admits an arbitrary
3638:  singular set on the ellipsoid ${\mathcal E}=\{I=1\}$, provided 
3639:  \[\lim_{s\to\Delta}\tilde U(s)=+\infty\;.\]
3640: It is worthwhile noticing that as a consequence of Theorem~\ref{central_conf}, a total collision
3641: trajectory will not interact, definitively, with the singularities of $\tilde U$.
3642: \end{example}
3643: 
3644: The class of potentials satisfying our assumptions is clearly stable with respect to 
3645: the addition of arbitrary perturbations of class $\mathcal C^1$. Therefore, we are mainly 
3646: interested in the analysis of those perturbations which are singular themselves. 
3647: 
3648: \begin{example}[$N$--body potentials with time--varying masses]\label{ex:0.5}
3649: Although the potentials in the previous examples  do not depend on time, our assumptions
3650: allow an effective time--dependence of the potentials.  For instance, we can choose positive 
3651: and bounded $\cont^1$ functions $m_i(t)$, $i=1,\ldots,n$.
3652: 
3653: Obviously, the simplest example is the class of $\alpha$-homogeneous $n$-body problem
3654: \[
3655: U_\alpha(t,x) = \sum_{\substack{i<j\\ i,j=1}}^{n} \frac{m_i(t) m_j(t)}{|x_i-x_j|^\alpha},\qquad 0<\alpha<2.
3656: \]
3657: Assumptions \ref{a:U0} and \ref{a:U1} are trivially satisfied since $U$ is positive, 
3658: diverges to  $+\infty$ when $x$ approaches 
3659: $\Delta = \{x \in \RR^{nd}: x_i=x_j \mbox{ for some }i\neq j\}$,
3660: and does not depend on time.
3661: Furthermore in both \ref{a:U2} and \ref{a:U2h} the equality is achieved
3662: with $\tilde \alpha=\alpha$ and $C_2=0$. Since $U$ is homogeneous of degree $-\alpha$,
3663: in \ref{a:U3h} and \ref{a:U4h} the function $\tilde U$ coincides with $U$.
3664: \end{example}
3665: 
3666: \begin{example}[Quasi--homogeneous potentials]\label{ex:1}
3667: We can also handle homogeneous perturbations of degree $-\beta$ of the potential $U_\alpha$
3668: \[
3669: U(x) = U_\alpha(x) + \lambda U_\beta(x) \qquad\qquad 0<\beta<\alpha<2.
3670: \]
3671: Indeed, when $\lambda>0$ condition \ref{a:U2h} is verified (with the strict
3672: inequality) with $\gamma = C_2 =0$, while, when $\lambda<0$, then \ref{a:U2} holds, when $|x|$
3673: is sufficiently small, with $C_2=\alpha-\beta$ and $0<\gamma < \alpha-\beta$.
3674: 
3675: As pointed out in \cite{diacuab} (where the case $\beta=1$ and $\alpha>1$ was treated), quasi--homogeneous potentials generalize classical potentials such as Newton, Coulomb, Birkhoff, Manev and many others. Therefore, the range of physical applications
3676: of quasi--homogeneous potentials spans from celestial mechanics and atomic physics to chemistry and crystallography. It is worthwhile noticing that the collision problem for quasi--homogeneous potentials exhibit an interesting and peculiar lack of regularity. Indeed, a classical framework for the study of collisions is given by the McGehee coordinates \cite{mcgehee1}
3677: (here and below we assume, for simplicity of notations, all the masses be equal to one):
3678: \begin{equation*}
3679: \begin{split}
3680: r&=|x|=I^{1/2}\\
3681: s&=\dfrac{x}{r}\\
3682: v&=r^{\alpha/2}(y\cdot s)\\
3683: u&=r^{\alpha/2}(y-y\cdot s s).
3684: \end{split}
3685: \end{equation*}
3686: After a reparametrization of the time--variable (see~\ref{eq:mcgehee1}):
3687: \begin{equation}\label{eq:mcgehee2}
3688: d\tau=r^{-1-\alpha/2}dt,
3689: \end{equation}
3690: the equation of motions become (here $^\prime$ denotes differentiation with respect to the new time variable $\tau$):
3691: \begin{equation*}
3692: \begin{split}
3693: r^\prime&=rv\\
3694: v^\prime&=\dfrac{\alpha}{2}v^2+|u|^2-r^{\alpha-\beta}\lambda U_\beta(s)-\alpha U_\alpha(s)\\
3695: s^\prime&=u\\
3696: u^\prime&=\left(\dfrac{\alpha}{2}-1\right)vu-|u|^2s+r^{\alpha-\beta}\lambda\left(U_\beta(s)s-\nabla U_\beta(s)\right)+\alpha U_\alpha(s)s+\nabla U_\alpha(s)\;.\\
3697: \end{split}
3698: \end{equation*}
3699: The field depends on $r$ in a non smooth manner, unless $\alpha-\beta\geq 1$ (this last condition was indeed assumed in \cite{diacuab}). Hence the flow \emph{can not be continuously extended} to the total collision manifold $C=\{(r,s,v,u)\;:\; r=0,\,\frac{1}{2}(|u|^2+v^2)-2U_\alpha=0\}$. Another peculiar feature of this system is that the monotonicity of the variable $v$ can not be ensured close to the collision manifold. As a consequence, the usual analysis of collision and near collision motions can not be extended to this case.
3700: \end{example}
3701: 
3702: \begin{example}[$N$--body potential reduced by a symmetry group satisfying the rotating circle property] 
3703: The paper \cite{terven} deals with \label{exe:gruppo1}
3704: minimal trajectories to the spatial $2N$--body problem under the \emph{hip--hop symmetry}, 
3705: where the configuration is constrained at all time to form a regular antiprism.   
3706: This problem has three degrees of freedom and the reduced potential of a 
3707: configuration generated by the  point of coordinates $(u,\zeta)\in\CC\times\RR\simeq\RR^3$
3708: decomposes as 
3709: $$
3710: U(u,\zeta)=\frac{K(N)}{|u|^\alpha}+U_0(u,\zeta),
3711: $$
3712: where 
3713: \begin{eqnarray}
3714: K(N)&=&\sum\limits_{k=1}^{N-1}\frac{1}{\sin^\alpha(\frac{k\pi}{N})}, \nonumber
3715: \\
3716: U_0(u,\zeta)&=& 
3717: \sum\limits_{k=1}^{N} 
3718: \frac{1}{\left(\sin^2\left(\frac{(2k-1)\pi}{2N}\right)|u|^2+\zeta^2\right)^{\frac{\alpha}{2}}}, \nonumber
3719: \end{eqnarray}
3720: The first term  comes from the interaction among points of the same $N$--agon and is 
3721: singular at simultaneous partial
3722: collisions on the $\zeta$--axis. The second term, $U_0(u,\zeta)$, comes from the interaction 
3723: between the the upper and lower $N$--agons and is singular only at the origin.
3724: One easily verifies that all the assumptions are satisfied, including, again by
3725: Proposition~\ref{prop:usette}, \ref{a:U6} and \ref{a:U7h}. In general, one
3726: easily verifies that, for a given symmetry group $G$ of the $N$--body problem,
3727: it is equivalent to say that $\ker\tau$ has the rotating circle property and
3728: that the reduced potential verifies \ref{a:U6} and \ref{a:U7h}.
3729: \end{example}
3730: 
3731: \begin{example}[$N$--body potential reduced by a symmetry group not satisfying the rotating 
3732: circle property] Consider the \label{exe:gruppo2} symmetry groups generated by rotations 
3733: introduced in \cite{F2}: the configuration is, at all time,
3734: an orbit of a group $Y$ of rotations about given lines in the $3$--dimensional space. 
3735: When $Y$ is a finite group, the reduced potential takes the form required in 
3736: Theorem~\ref{theo:nocoll_2} and minimizers can be shown to be free of collision. 
3737: \end{example}
3738: 
3739: %  \bibliographystyle{plain}
3740: %  \bibliography{bibliog}
3741: % \end{document}
3742: \begin{thebibliography}{10}
3743: 
3744: \bibitem{amco93}
3745: A.~Ambrosetti and V.~Coti~Zelati.
3746: \newblock {\em Periodic solutions of singular {L}agrangian systems}.
3747: \newblock Progress in Nonlinear Differential Equations and their Applications,
3748:   10. Birkh\"auser Boston Inc., Boston, MA, 1993.
3749: 
3750: \bibitem{ABT}
3751: G.~Arioli, V.~Barutello, and S.~Terracini.
3752: \newblock A new branch of mountain pass solutions for the choreographical
3753:   \emph{3}-body problem.
3754: \newblock {\em Comm. Math. Phys.}, 268(5):439--463, 2006.
3755: 
3756: \bibitem{bahrab}
3757: A.~Bahri and P.H. Rabinowitz.
3758: \newblock Periodic solutions of {H}amiltonian systems of {$3$}-body type.
3759: \newblock {\em Ann. Inst. H. Poincar\'e Anal. Non Lin\'eaire}, 8(6):561--649,
3760:   1991.
3761: 
3762: \bibitem{BFT}
3763: V.~Barutello, D.L. Ferrario, and S.~Terracini.
3764: \newblock Symmetry groups of the planar $3$-body problem and action--minimizing
3765:   trajectories.
3766: \newblock Arxiv:math.DS/0404514, preprint (2004).
3767: 
3768: \bibitem{BS}
3769: V.~Barutello and S.~Secchi.
3770: \newblock Morse index properties of colliding solutions to the $n$-body
3771:   problem.
3772: \newblock Arxiv:math/0609837, preprint (2006).
3773: 
3774: \bibitem{BT2004}
3775: V.~Barutello and S.~Terracini.
3776: \newblock Action minimizing orbits in the {$n$}-body problem with simple
3777:   choreography constraint.
3778: \newblock {\em Nonlinearity}, 17(6):2015--2039, 2004.
3779: 
3780: \bibitem{BC-Z}
3781: U.~Bessi and V.~Coti~Zelati.
3782: \newblock Symmetries and noncollision closed orbits for planar {$N$}-body-type
3783:   problems.
3784: \newblock {\em Nonlinear Anal.}, 16(6):587--598, 1991.
3785: 
3786: \bibitem{chen}
3787: K.-C. Chen.
3788: \newblock Action-minimizing orbits in the parallelogram four-body problem with
3789:   equal masses.
3790: \newblock {\em Arch. Ration. Mech. Anal.}, 158(4):293--318, 2001.
3791: 
3792: \bibitem{MR2032484}
3793: K.-C. Chen.
3794: \newblock Variational methods on periodic and quasi-periodic solutions for the
3795:   {$N$}-body problem.
3796: \newblock {\em Ergodic Theory Dynam. Systems}, 23(6):1691--1715, 2003.
3797: 
3798: \bibitem{Ch2}
3799: A.~Chenciner.
3800: \newblock Action minimizing solutions of the newtonian $n$-body problem: from
3801:   homology to symmetry.
3802: \newblock In {\em Proceedings of the ICM, Peking}, 2002.
3803: 
3804: \bibitem{chenkyoto}
3805: A.~Chenciner.
3806: \newblock Simple non-planar periodic solutions of the $n$-body problem.
3807: \newblock In {\em Proceedings of the NDDS Conference, Kyoto}, 2002.
3808: 
3809: \bibitem{CM}
3810: A.~Chenciner and R.~Montgomery.
3811: \newblock A remarkable periodic solution of the three body problem in the case
3812:   of equal masses.
3813: \newblock {\em Ann. of Math.}, 152(3):881--901, 2000.
3814: 
3815: \bibitem{Dev1}
3816: R.~L. Devaney.
3817: \newblock Collision orbits in the anisotropic {K}epler problem.
3818: \newblock {\em Invent. Math.}, 45(3):221--251, 1978.
3819: 
3820: \bibitem{Dev2}
3821: R.~L. Devaney.
3822: \newblock Nonregularizability of the anisotropic {K}epler problem.
3823: \newblock {\em J. Differential Equations}, 29(2):252--268, 1978.
3824: 
3825: \bibitem{Dev3}
3826: R.~L. Devaney.
3827: \newblock Triple collision in the planar isosceles three-body problem.
3828: \newblock {\em Invent. Math.}, 60(3):249--267, 1980.
3829: 
3830: \bibitem{Dev_survey}
3831: R.~L. Devaney.
3832: \newblock Singularities in classical mechanical systems.
3833: \newblock In {\em Ergodic theory and dynamical systems, I (College Park, Md.,
3834:   1979--80)}, volume~10 of {\em Progr. Math.}, pages 211--333. Birkh\"auser
3835:   Boston, Mass., 1981.
3836: 
3837: \bibitem{diacureg}
3838: F.~Diacu.
3839: \newblock Regularization of partial collisions in the {$N$}-body problem.
3840: \newblock {\em Differential Integral Equations}, 5(1):103--136, 1992.
3841: 
3842: \bibitem{Dpainleve}
3843: F.~Diacu.
3844: \newblock Painlev\'e's conjecture.
3845: \newblock {\em Math. Intelligencer}, 15(2):6--12, 1993.
3846: 
3847: \bibitem{diacuab}
3848: F.~Diacu.
3849: \newblock Near-collision dynamics for particle systems with quasihomogeneous
3850:   potentials.
3851: \newblock {\em J. Differential Equations}, 128(1):58--77, 1996.
3852: 
3853: \bibitem{diacu_book}
3854: F.~Diacu.
3855: \newblock Singularities of the {$N$}-body problem.
3856: \newblock In {\em Classical and celestial mechanics (Recife, 1993/1999)}, pages
3857:   35--62. Princeton Univ. Press, Princeton, NJ, 2002.
3858: 
3859: \bibitem{DPS1}
3860: F.~Diacu, E.~P{\'e}rez-Chavela, and M.~Santoprete.
3861: \newblock Central configurations and total collisions for quasihomogeneous
3862:   {$n$}-body problems.
3863: \newblock {\em Nonlinear Anal.}, 65(7):1425--1439, 2006.
3864: 
3865: \bibitem{DS}
3866: F.~Diacu and M.~Santoprete.
3867: \newblock On the global dynamics of the anisotropic {M}anev problem.
3868: \newblock {\em Phys. D}, 194(1-2):75--94, 2004.
3869: 
3870: \bibitem{DPS2}
3871: F. Diacu, E. P{\'e}rez-Chavela, and M. Santoprete.
3872: \newblock The {K}epler problem with anisotropic perturbations.
3873: \newblock {\em J. Math. Phys.}, 46(7):072701, 21, 2005.
3874: 
3875: \bibitem{elbialy}
3876: M.S. ElBialy.
3877: \newblock Collision singularities in celestial mechanics.
3878: \newblock {\em SIAM J. Math. Anal.}, 21(6):1563--1593, 1990.
3879: 
3880: \bibitem{F2}
3881: D.L. Ferrario.
3882: \newblock Transitive decomposition of symmetry groups for the $n$--body
3883:   problem.
3884: \newblock Arxiv:math.DS/0603684, preprint (2006).
3885: 
3886: \bibitem{F3}
3887: D.L. Ferrario.
3888: \newblock Symmetry groups and non-planar collisionless action-minimizing
3889:   solutions of the three-body problem in three-dimensional space.
3890: \newblock {\em Arch. Ration. Mech. Anal.}, 179(3):389--412, 2006.
3891: 
3892: \bibitem{FT}
3893: D.L. Ferrario and S.~Terracini.
3894: \newblock On the existence of collisionless equivariant minimizers for the
3895:   classical {$n$}-body problem.
3896: \newblock {\em Invent. Math.}, 155(2):305--362, 2004.
3897: 
3898: \bibitem{G}
3899: W.B. Gordon.
3900: \newblock A minimizing property of {K}eplerian orbits.
3901: \newblock {\em Amer. J. Math.}, 99(5):961--971, 1977.
3902: 
3903: \bibitem{Gu}
3904: M.~Gutzwiller.
3905: \newblock The anisotropic kepler problem in two dimensions.
3906: \newblock {\em J. Mathematical Phys.}, 14:139--152, 1973.
3907: 
3908: \bibitem{KPV}
3909: C.~Kenig, G.~Ponce, and L.~Vega.
3910: \newblock On the interaction of nearly parallel vortex filaments.
3911: \newblock {\em Comm. Math. Phys.}, 243:471--483, 2003.
3912: 
3913: \bibitem{KMD}
3914: R.~Klein, A.~Majda, and K.~Damodaran.
3915: \newblock Simplified equations for the interaction of nearly parallel vortex
3916:   filaments.
3917: \newblock {\em J. Fluid Mech.}, 288:201--248, 1995.
3918: 
3919: \bibitem{LC}
3920: T.~Levi~Civita.
3921: \newblock Sur la r\'egularization du probl\`eme des trois corps.
3922: \newblock {\em Acta Math.}, 42:44--, 1920.
3923: 
3924: \bibitem{mate95}
3925: P.~Majer and S.~Terracini.
3926: \newblock On the existence of infinitely many periodic solutions to some
3927:   problems of {$n$}-body type.
3928: \newblock {\em Comm. Pure Appl. Math.}, 48(4):449--470, 1995.
3929: 
3930: \bibitem{marchal}
3931: C.~Marchal.
3932: \newblock How the method of minimization of action avoids singularities.
3933: \newblock {\em Celestial Mech. Dynam. Astronom.}, 83(1-4):325--353, 2002.
3934: \newblock Modern celestial mechanics: from theory to applications (Rome, 2001).
3935: 
3936: \bibitem{MM}
3937: J.N. Mather and R.~McGehee.
3938: \newblock Solutions of the collinear four body problem which become unbounded
3939:   in finite time.
3940: \newblock In {\em Dynamical systems, theory and applications (Rencontres,
3941:   Battelle Res. Inst., Seattle, Wash., 1974)}, pages 573--597. Lecture Notes in
3942:   Phys., Vol. 38. Springer, Berlin, 1975.
3943: 
3944: \bibitem{mcgehee1}
3945: R.~McGehee.
3946: \newblock Triple collision in the collinear three-body problem.
3947: \newblock {\em Invent. Math.}, 27:191--227, 1974.
3948: 
3949: \bibitem{mcgehee2}
3950: R.~McGehee.
3951: \newblock von {Z}eipel's theorem on singularities in celestial mechanics.
3952: \newblock {\em Exposition. Math.}, 4(4):335--345, 1986.
3953: 
3954: \bibitem{moeckel88}
3955: R.~Moeckel.
3956: \newblock Some qualitative features of the three-body problem.
3957: \newblock In {\em Hamiltonian dynamical systems (Boulder, CO, 1987)}, volume~81
3958:   of {\em Contemp. Math.}, pages 1--22. Amer. Math. Soc., Providence, RI, 1988.
3959: 
3960: \bibitem{montgomery96}
3961: R.~Montgomery.
3962: \newblock The geometric phase of the three-body problem.
3963: \newblock {\em Nonlinearity}, 9(5):1341--1360, 1996.
3964: 
3965: \bibitem{montgomery_contmath}
3966: R.~Montgomery.
3967: \newblock Action spectrum and collisions in the planar three-body problem.
3968: \newblock In {\em Celestial mechanics (Evanston, IL, 1999)}, volume 292 of {\em
3969:   Contemp. Math.}, pages 173--184. Amer. Math. Soc., Providence, RI, 2002.
3970: 
3971: \bibitem{mont_prepr}
3972: R.~Montgomery.
3973: \newblock Fitting {H}yperbolic pants to a three-body problem.
3974: \newblock {\em Ergodic Theory Dynam. Systems}, 25(3):921--947, 2005.
3975: 
3976: \bibitem{painleve}
3977: P.~Painlev\'e.
3978: \newblock {\em Le\c{c}ons sur la th\'eorie analytique des \'equations
3979:   diff\'erentielles}.
3980: \newblock Hermann, Paris, 1897.
3981: 
3982: \bibitem{polsaa1}
3983: H.~Pollard and D.G. Saari.
3984: \newblock Singularities of the {$n$}-body problem. {I}.
3985: \newblock {\em Arch. Rational Mech. Anal.}, 30:263--269, 1968.
3986: 
3987: \bibitem{polsaa2}
3988: H.~Pollard and D.G. Saari.
3989: \newblock Singularities of the {$n$}-body problem. {II}.
3990: \newblock In {\em Inequalities, II (Proc. Second Sympos., U.S. Air Force Acad.,
3991:   Colo., 1967)}, pages 255--259. Academic Press, New York, 1970.
3992: 
3993: \bibitem{riahi2}
3994: H.~Riahi.
3995: \newblock Study of the generalized solutions of {$n$}-body type problems with
3996:   weak forces.
3997: \newblock {\em Nonlinear Anal.}, 28(1):49--59, 1997.
3998: 
3999: \bibitem{riahi}
4000: H.~Riahi.
4001: \newblock Study of the critical points at infinity arising from the failure of
4002:   the {P}alais-{S}male condition for {$n$}-body type problems.
4003: \newblock {\em Mem. Amer. Math. Soc.}, 138(658):viii+112, 1999.
4004: 
4005: \bibitem{saa72}
4006: D.~G. Saari.
4007: \newblock Singularities and collisions of {N}ewtonian gravitational systems.
4008: \newblock {\em Arch. Rational Mech. Anal.}, 49:311--320, 1972/73.
4009: 
4010: \bibitem{serter1}
4011: E.~Serra and S.~Terracini.
4012: \newblock Collisionless periodic solutions to some three-body problems.
4013: \newblock {\em Arch. Rational Mech. Anal.}, 120(4):305--325, 1992.
4014: 
4015: \bibitem{sperling}
4016: H.J. Sperling.
4017: \newblock On the real singularities of the {$N$}-body problem.
4018: \newblock {\em J. Reine Angew. Math.}, 245:15--40, 1970.
4019: 
4020: \bibitem{sundman}
4021: K.~F. Sundman.
4022: \newblock M\'emoire sur le probl\`eme des trois corps.
4023: \newblock {\em Acta Math.}, 36:105--179, 1913.
4024: 
4025: \bibitem{terven}
4026: S.~Terracini and A.~Venturelli.
4027: \newblock Symmetric trajectories for the $2n$-body problem with equal masses.
4028: \newblock {\em Arch. Rational Mech. Anal.}, 2006.
4029: \newblock To appear.
4030: 
4031: \bibitem{andrea_thesis}
4032: A.~Venturelli.
4033: \newblock {\em Application de la minimisation de l'action au Probl\`eme des $N$
4034:   corps dans le plan et dans l'espace}.
4035: \newblock PhD thesis, University Paris VII, Paris, 2002.
4036: 
4037: \bibitem{wintner}
4038: A.~Wintner.
4039: \newblock {\em The {A}nalytical {F}oundations of {C}elestial Mechanics}.
4040: \newblock Princeton Mathematical Series, v. 5. Princeton University Press,
4041:   Princeton, N. J., 1941.
4042: 
4043: \bibitem{xia}
4044: Z.~Xia.
4045: \newblock The existence of non collision singularities in newtonian systems.
4046: \newblock {\em Ann. of Math.}, 135:411--468, 1992.
4047: 
4048: \bibitem{vonzeipel}
4049: H.~von Zeipel.
4050: \newblock Sur les singularités du problème des $n$ corps.
4051: \newblock {\em Ark. Math. Astr. Fys.}, 4:1--4, 1908.
4052: 
4053: \end{thebibliography}
4054: 
4055: \end{document}
4056: 
4057: 
4058: