Sei sulla pagina 1di 23

THE 3-BODY PROBLEM

A series of 3 lectures by Prof. R. Montgomery


8-10 may 1997
CIMAT, Guanajuato, Mexico

Outline
1. Background. The 3- body and N-body equations, their symmetries, and
constants of motions. Some of the basic problems. The Kepler case. The central
moment of inertia I. The difference between positive and negative energies.
Reduced space.
2. Reduced dynamics. Details of the reduced space and reduced dynamics.
Variational methods. Results after ’mollifying’ potential.
3. Collisions. Collisions and their regularizations. Prospects.
4. Literature. There are a vast number of references. I list a few I like.

1 Background
Imagine three objects moving in space. The objects could be three stars, or a
star, a planet and a satellite, or the nucleus of a Helium atom, together with its
two electrons. We idealize the objects as points. The spatial positions of these
points are represented by three vectors x1 , x2 , x3 . (We have dispensed with
writing little arrows over the xs and hope this is not confusing.) The masses of
the objects are denoted m1 , m2 , m3 . They are positive. We refer to the objects
as “particles” or “stars” or “masses”.
The objects move so that their positions xi are functions of time t. If the
masses are astronomical objects, then to a good degree of approximation they
move according to Newton’s laws:

d2 x1 x2 − x1 x3 − x1
m1 2
= m1 m2 + m1 m3
dt kx2 − x1 k kx3 − x1 k

d2 x2 x1 − x2 x3 − x2
m2 = m2 m1 + m2 m3
dt2 kx1 − x2 k kx3 − x2 k
d2 x3 x1 − x3 x2 − x3
m3 = m3 m1 + m3 m2
dt2 kx1 − x3 | kx2 − x3 k
The right-hand sides should be multiplied through by a physical constant de-
noted G, Newton’s universal constant of gravitation. But we are really mathe-
maticians, so we are allowed to set G = 1.
What is the three-body problem? Before Poincare, it was to solve these
equations as explicitly as possible, in the same sense that the harmonic oscillator
2
problem ( ddt2x = −x) or the Kepler problem ( N = 2 below) is solved. Poincare
showed this is impossible.

1
The history around his discovery is ironic and complicated. Mittag-Leffler
formulated a series of prize problems for King Oscar of Sweden. One was a
version of the three-body problem: “Given a system of arbitrarily many mass
points that attract each other according to Newton’s laws, try to find, under the
assumption that no two points ever collide, a representation of the coordinates
of each point as a series in variable that is some known function of time and
for all of whose values the series converges uniformly.” Poincare did not solve
the problem, but won the prize for his work on it. Poincare summarized this
work in his book “Les Méthodes Nouvelles de la Mécanique Céleste”. After
Poincare, in the 1920s, Sundman actually solved the King Oscar problem for
three masses! However his power series solution converges so incredibly slowly,
and contains so little extractable information that it seems to be irrelevant to
understanding the qualitative nature of the dynamics! For a discussion of these
matters we refer the reader to the recent Intelligencer article of Diacu, and to
the book ‘Celestial Encounters’ by Diacu and Holmes.
Today, when people refer to the three-body problem, they usually mean
the qualitative study of these equations. This contains within a number of
open sub-problems. The ones we are interested in here refer mostly to periodic
orbits.
Problem 1: [Another Poincare conjecture] The set of periodic orbits are
dense within the space of all bounded motions.
Here is a slight variant of this problem. The energy, defined momentarily,
is constant along any solution. If the solution is bounded then its energy is
necessarily negaitive. Fix a negative energy. Are the periodic orbits dense
within this energy hypersurface?
Problem 2: [A ’homotopy’ density conjecture.] The planar configuration
space, with collisions (xi = xj , i 6= j) is homotopic to the product of a two-
sphere minus three points with a circle. As such it has an enormous fundamental
group. Is there a periodic orbit in each free homotopy class?
I have proved this conjecture is valid, but I had to “cheat”. What I did was
change the forces, that is, the right hand sides of the equations. I only needed to
make the change when two of the particles get very close. Then the added forces
must grow like 1/r3 or stronger, where r is the distance between the particles.
One thing I would like to do here is explain this result.
Geometrically, the three masses form a triangle. So the equations define a
dynamical system on the space of triangles. Instead of asking how the triangle
evolves in space, we will focus instead on how its congruence class evolves.
This change of viewpoint will significantly reduce the dimensionality of the
system of equations and simplify the overall picture of the subject.

1.1 N bodies
We find it helpful to put the three-body problem within the more general context
of the N-body problem. Then there are N point masses m1 , . . . , mN moving in

2
a d-dimensional Euclidean space. We will be most interested in the cases d = 2
and 3. The locations of the masses are written x1 , x2 , . . . , xN . Let
rij = kxi − xj k
denote the distances betweem mass i and j. We suppose given N2 scalar


functions fij : IR+ → IR, called the two-particle potentials. Form the potential
function
V (x1 , . . . , xN ) = Σi6=j fij (rij )
Then the equations are
d2 xi ∂V
mi 2
=− ; i = 1, . . . , N.
dt ∂xi
We will refer to these as either the N-body equations or Newton’s equa-
tions.
m m
To recover the three-body equations take N = 3 and set fij = − riij j .
Warning: Later on , we are going to “cheat” by adding to the standard gravita-
tional two-body potential a term of the form −
r 2 . These so-called “strong-force”
potentials are ubiquituous in the variational approaches to the problem.
There are two standard formalisms within which to investigate the N-body
problem, the Hamiltonian formalism and the Lagrangian formalism. Each has
its own “most important” function. To define these functions, first define the
total kinetic energy
1
T = Σmi kẋi k2
2
where dots over things represent their time derivative. The Hamiltonian is the
function
H = T + V.
It represents the total energy of the N-body system. The Lagrangian is the
function
L = T − V.
Both of these are to be thought of as functions of x1 , . . . , xN , ẋ1 , . . . , ẋN . That
is to say, they are real-valued functions on the 2N -fold product of IRd .
To explain the significance of these functions, suppose that γ(t) = (x1 (t), . . . , xN (t))
is a solution to the N-body equations. Then
d
H(x1 (t), . . . , xN (t), ẋ1 (t), . . . , ẋN (t)) = 0.
dt
A function such as H which is constant along solutions is called a constant
of the motion. To explain the significance of L, we form the action
Z
S(γ) = Ldt.
γ

Then, we have

3
Background Theorem 1 A curve γ in (IRd )N is a solution to Newton’s equa-
tion if and only if dS(γ) = 0, (in the classical notation δS
δγ (γ) = 0).
We are being deliberately sloppy regarding end-point conditions on the varia-
tions of γ in the theorem. These are important, but we will come to them later
if we need them.
Remark. This is often called Hamilton’s principle of least action. The re-
d ∂L ∂L
sulting differential equations dt ∂ ẋi = ∂xi are referred to as the Euler-Lagrange
equations. For some history of the principle see [?].
...
To better understand our methods, and the direction in which we are going,
we will place these two formalisms in their modern differential-geometric con-
text. (See Abraham-Marsden, Arnol’d, Sen,... for sometimes excessive details.)
Set Q = Q(N, d) = (IRd )N . Q is refered to as configuration space. The
curve γ = {x1 , . . . , xN } is a curve here. Differentiating the curve yields the
curve (x1 , . . . , xN , ẋ1 , . . . , ẋN ) in the tangent bundle T Q of Q. L is a function
here, so that the Lagrangian formalism lives on the tangent bundle.
The Hamiltonian formalism lives on the cotangent bundle. Set:
pi = mi ẋi , i = 1, . . . , N.
We view (p1 , . . . , pN ) as an element of the cotangent bundle, Tx∗ Q at x ∈ Q.
Then T can be rewritten T = 12 Σ m1i kpi k2 as a function on T ∗ Q, so we may think
of H as a function here. (Please excuse the two uses of T. They are too steeped
in tradition to change.) The N-body equations can be rewritten as Hamilton’s
equations:
∂H
ẋi =
∂pi
∂H
ṗi = − .
∂xi
The kinetic energy T is one-half of a positive-definite quadratic form of
an inner product on Q. This inner product, like any inner product, induces
an invertible linear map, Q → Q∗ between Q and its dual space. This map is
called the Legendre transformation (associated to L) and is given by the relation
pi = mi ẋi between the momenta pi and the velocities ẋi . Note that T Q = Q×Q
and T ∗ Q = Q × Q∗ .
We may think of this inner-product as a Riemannian metric on Q, namely
ds2 = Σmi kdxi k2 .
Both the Lagrangian and Hamilitonian formalisms survive intact if we replace
our configuration space Q by an arbitrary Riemannian manifold. The Rieman-
nian metric defines the kinetic energy and one picks a function on Q to replace
our potential V . The N-body equations are then replaced by
∇γ̇ γ̇ = −∇V

4
where ∇ is the Levi-Civita connection. These are sometimes called Newton’s
equations on a manifold. The formulas for L and H, as well as Hamilton’s
principle and equations, are the same.
For the last century or so, the Hamiltonian formalism has dominated dis-
cussions of the three and N-body problem. Recently Lagrangian or variational
methods have gained power, and our recent work is based more on these.
There is another variational principle which yields yet a third formulation
of mechanics. It is the oldest of them all, and probably the strangest and most
wonderful. It is attributed to Mauperituis (1746) although Euler has instances of
it earlier (1744). (See [?] for history. ) In it one fixes the energy but not the time.
Let ds2 denote the kinetic energy metric ( d2 sγ (γ̇, γ̇) = 2T (γ̇, γ̇)). Form a new
metric: (E−V )ds2 , called the Mauperituis metric. It is is conformally equivalent
to the original kinetic-energy metric, except that it is only to be defined on the
region E ≥ V . For historical reasons, this subset ΣE := {x : E ≥ V (x)} of
configuration space Q is called the Hill’s region.
The Euler-Maupertuis least-action principle asserts that the solutions to
Newton’s equations with energy E are in one-to-one correspondence with “geodesics”
for this Maupertuis metric. We have to be a bit careful about what “geodesic”
means, since any curve satisfying E = V has zero length. Having E = V means
that the velocities of all particles are zero, so we should not allow a curve to
move around on this surface. Using Newton’s equations one finds (by a Taylor
expansion) that any solution of energy E which satisfies E = V at some instant
moves orthogonally to the surface E = V in such a way that V is decreasing.
So we insist that our Mauperitis geodesics have this property. The meaning of
‘geodesic’ for segments of curves x(t) interior to E = V ( satisfying E > V (x(t)))
is as usual: they extremize arclength.
Now the Maupertuis metric is often called the Jacobi metric. I don’t know
why. I’m sure the history is interesting. In any case I will probably follow the
current trend and refer to the Maupertuis metric as the Jacobi metric, and the
corresponding lengths the “Jacobi actions”.
It is quite interesting to note that in some vague sense the Jacobi action
principle and the Lagrange action principle are Legendre transforms of each
other, with energy E and time t being R the Legendre- dual variables. More
specifically, set SL (x0 , x1 ; t) = inf Ldt where the infimum is taken over all
paths connecting the R√ configurations x0 , x1 to each other in time t. And let
SJ (x0 , x1 ; E) = inf E − V ds where the infimum is taken over those paths
with energy E joining x0 to x1 . Here we must assume the two points lie in the
Hill’s region. Or, I suppose, we could set SJ = ∞ if they do not, I’m not sure. In
any case, the Legendre transform of a function f (t) is defined to be the function
Leg(f )(E) = inft (f (t) + Et). In many circumstances, SL and SJ are Legendre
transforms. (The positions are treated as fixed parameters in performing the
Legendre transform.) I recently learned this fact from the book by Gutzwiller
entitled Chaos in Classical and Quantum Mechanics.

5
Exercise. For a “free” particle, i.e. the case V = 0, compute both SL and
SJ in terms of the Riemannian distance d(x0 , x1 ) between the points, and show
that they are indeed Legendre transforms of each other.

1.2 Symmetries and constants of the motion.


Let E(d) denote the group of isometries of d-dimensional Euclidean space. An
element g of E(d) can be decomposed into a linear, or rotational, part R, and
translation part d ( for displacement). Then g(xi ) = Rxi + d. Here R is a d × d
orthogonal matrix and d is a vector in IRd .
Lemma 1 Let g ∈ E(d) be any fixed isometry of d-space, and t0 any real
number. If (x1 (t), . . . , xN (t)) is a solution to Newton’s equations, then so is
(g(x1 (t − t0 )), . . . , g(xN (t − t0 )).
This lemma can be proved by direct computation. The crucial fact is that
both the kinetic energy T , and potential energy V are invariant under the
action of E(d), and do not explicitly depend on the time t. This lemma can be
formulated, and even more easily proved, in the general context of Riemannian
geometry.
The lemma says that the Lie group  E(d) × IR acts on solutions. The di-
mension of this Lie group is N = d+1 2 + 1. For each one of these dimensions
there is an associated constant fi = fi (x, ẋ), i = 1, . . . , N of the motion. The
constant for time translations has already been written: it is the total energy
H. The constants associated to translation xi 7→ xi + d are the components of
the total linear momentum
P = Σmi ẋi = Σpi .
d

The constants associated to rotations xi 7→ Rxi are the 2 components of the
total angular momentum:
J = Σmi xi ∧ ẋi = Σxi ∧ pi .
We are to think of J as taking values in the Lie algebra o(d) of the orthogonal
group. To do this, we use the fact that for an inner product space V there
is a canonical identification between o(V ), the lie algebra of skew-symmetric
matrices, and Λ2 V , the space of bi-vectors. This identification is obtained by
letting x ∧ p correspond to the v 7→ 21 xhp, vi − phx, vi. The cases of real interest
to us are dimensions d = 2 and 3. In these cases we have further identifications
o(2) = IR and o(3) = IR3 . Using these identifications we get, for d = 2,
x ∧ p = x1 p2 − x2 p1 (d = 2)
where x = (x1 , x2 ), p = (p1 , p2 ) are the components of the two-vectors. For
d = 3 we get
x ∧ y = x × y (d = 3),

6
the usual vector cross-product of three-vectors.
We recall what we mean when we say these are constants of the motion:

Ḣ = 0, Ṗ = 0, J˙ = 0

along any solution curve to Newton’s equations.


Remark The relation between symmetries and constants of the motion is a
general phenomenon in mechanics. In the Lagrangian formalism it is called
Noether’s theorem. In the Hamiltonian formalism it has been formalized by
the concept of momentum map. We refer the interested reader to Abraham-
Marsden, Guillemin-Sternberg, Marsden-Ratiu, or Arnol’d.

1.3 N = 2; The Kepler problem.


There are only two masses m1 and m2 , located at x1 and x2 . Set x = x1 − x2 ,
r = kxk, M = m1 + m2 and µ = mm11+m m2
2
. µ is called the reduced mass. We
assume that the center of mass is the origin: m1 x1 + m2 x2 = 0. Then we may
solve for the positions xi in terms of the relative position x to get: x1 = m
M x
2

−m1
and x2 = M x. We find that the angular momentum is J = µx ∧ ẋ. This is
a constant bivector, and so defines a two-dimensional plane P with area form.
The time-dependent vectors x(t), ẋ(t) associated to any solution with angular
momentum J must lie on this plane. So the problem is effectively planar.
Let r, θ denote polar coordinates on this plane relative to an orthonormal
basis e1 , e2 . So write x = r(cosθe1 + sin(θ)e2 ). Then we find that:

J = µr2 θ̇e1 ∧ e2 .

Thus
pθ = µr2 θ̇
is constant. As a result of this expression, the angular momentum is sometimes
referred to as the area integral. This is because if we view the formula for pθ
as a one-form then it corresponds to µr2 dθ and d(r2 dθ) = 2dx ∧ dy = twice the
area form on P . Here x, y are Cartesian coordinates on P . It follows that the
integral of pθ over any path is proportional to the area swept out by the path.
Here “swept out” means that we close the path up by joining its beginning and
end to the origin by rays. Write A(t1 , t2 ) for the area swept out by traversing
the path for t1 ≤ t ≤ t2 . Since pθ is constant, we have

2µA(t1 , t2 ) = pθ (t2 − t1 ).

This is Kepler’s second law: equal areas are swept out by the motion in equal
times.
The kinetic energy of a trajectory is
µ 2
T = (ṙ + r2 θ̇2 ),
2

7
so its total energy is
H = T + f (r),
where f = f12 is the two-body potential. Observe that p2θ = µ2 r4 θ̇2 . It follows
that if we set
1
Vef f (r) = f (r) + 2 p2θ
µr
then the radial motion r(t) satisfies the differential equation

d2 r dVef f
µ 2
=− .
dt dr
µ 2
Alternatively, we have H = 2 ṙ + Vef f (r), so that
2
ṙ2 = (H − Vef f (r)).
µ
Either way, the radial motion of r is determined by Vef f (r) [Insert figure??].
The angular motion θ(t) is obtained by plugging this radial solution into the
expression for the areal integral pθ .
As is well known, if we take f = γ/r to be the Newtonian gravitational
potential, then the relative position vector, and hence each individual position
vector xi , describes a a conic section on the plane P with focus at the origin. A
traditional way to obtain these conics is to let u = 1r denote the inverse distance
and then to think of it as a function of the polar angle θ. We can convert the
d d dt dθ pθ
differential equation for r into one for u(θ) by using dθ = dt dθ and dt = µr 2 .
2 dV
One readily computes that ddθu2 = − c12 du ef f
where c = pµθ . Arnol’d attributes
this fact to Clairut (sp??). In the Newtonian potential case the function Vef f
is quadratic in u so that this is a linear inhomogeneous function. It is solvable
in terms of trig functions (H < 0) or exponential functions (H > 0).
The conic is degenerate if pθ = 0. The two masses move along a straight
line and collide at the origin.
For pθ 6= 0 the conic is an ellipse or circle if H < 0, a hyperbola if H > 0,
and a parabola if H = 0. When H is negative the eccentricity of the ellipse is
given by
2c2 H
e2 = 1 + .
γ2
It is given in “inverse” polar coordinates by
γ
u= (1 + e cos(θ − θ0 )),
c2
where θ0 is a constant angle describing the orientation of the ellipse relative to
our coordinate system on P . The semi-major axis a of the ellipse satisfies
γ
a= ,
2|H|

8
and the period of the ellipse is computed to satisfy

a3 γ
= ,
T2 4π 2
which is known as Kepler’s third law.
It is impossible to find a closed form formula for the Keplerian trajectory in
terms of elementary functions of time. The standard procedure is to introduce
an auxilary angular variable called the eccentric anomaly. It is a transcendental
function of time, and an elementary function of the usual polar angle θ. (The
usual polar angle is also called the true anomaly.) Power series and analytic
function theory can be brought to bear to explore these relationships.
There is an enormous literature on the Kepler problem! We recommend the
treatment in Arnol’d, (Encyclopaedia of Dynamical Systems III), Landau and
Lifshitz, and Gutzwiller. One of the topics we have not discussed is the hidden
o(4) symmetry of the planar Kepler problem. We refer the reader to the book
of Guillemin-Sternberg (Variations on a Theme of Kepler) for this. There are
also beautiful short papers by Moser and by Milnor on the subject.

1.4 Moment of Inertia


Recall that our configuration space Q is an inner-product space with norm-
square equal to:
kxk2 = Σmi kxi k2 .
This function is also called the central moment of inertia, denoted I, so that

I = kxk2 .

The reason behind this terminology is that if a planar system of N-bodies rotates
about its center of mass with a constant angular velocity ω, this being a vector
perpindicular to the plane of motion, then we have ẋi = ω × xi and the kinetic
energy of this motion is T = 21 Ikωk2 .
This moment of inertia satisfies three basic identities. Let U = −V (a
positive function) and h the constant value of the energy H = T − U for a given
motion. Then the three identities are
d2 I
= 4h + 2U,
dt2
1 2
I= Σa<b ma mb rab ,
M
2IT ≥ kJk2 .
In the second identity M = Σma is the total mass and rab = kxa − xb k is the
distance between the ath and bth particle. The first identity is referred to as
Sundman’s identity, or as the Lagrange-Jacobi identity. The second identity

9
is sometimes called Lagrange’s identity. As with most name-brand formulas
in celestial mechanics, these identities are probably due to none of these three
mathematicians. Only dedicated historians will be able to cipher out who it is
due to.
Proof of the first identity. Note that
dI
= 2Σmi hxi , ẋi i,
dt
so that
d2 I d2 xi
2
= 2Σmi kẋi k2 + 2Σmi hxi , 2 i.
dt dt
Now use
d2 xi 1 ∂U
=
dt2 mi ∂xi
to get
d2 I ∂U
2
= 4T + +2Σhxi , i,
dt ∂xi
where U is the negative of the potential as above. It follows from Euler’s identity
that if U is homogeneous of degree α then

d2 I
= 4T + 2αU,
dt2
or
d2 I
= 4h + (4 + 2α)U.
dt2
We are interested in the Newtonian potential case α = −1, in which case we
obtain the identity:
d2 I
= 4h + 2U.
dt2
Proof of the second identity.
We have raa = 0, so that if we sum over all a, b we get
2 2
Σma mb rab = 2Σa<b ma mb rab .

On the other hand


2
rab = kxa k2 − 2xa · xb + kxb k2 ,
so that
2
Σma mb rab = Σma mb kxa k2 −2Σma mb xa ·xb +Σma mb kxb k2 = 2M I−2(Σma xa )·(Σmb xb ).

But the center of mass is zero so that the last term is zero and we finally get
2
2Σa<b ma mb rab = 2M I,

10
as desired.
Proof of the third identity.
Recall that
kxa ∧ ẋa k2 = kxa k2 kẋa k2 − (xa · ẋa )2 ,
so we have
kxa ∧ ẋa k ≤ kxa kkẋa k.
It follows that
√ √
kJk ≤ Σma kxa ∧ ẋa k ≤ Σ ma kxa k ma kẋa k.

Applying Cauchy-Schwartz to this last sum yields


√ √
kJk ≤ I 2T ,

as desired.
Application of the first identity.
If the energy h is non-negative then I is convex function of time. If h > 0,
then it is strictly convex. It follows that if a positive-energy solution x(t) exists
for all time, then the function I(x(t)) must tend to infinity in both positive
and negative times. In other words,such a solution consists of the N particles
moving in from infinity, interacting for some bounded period of time, and then
receding again to infinity. (This does not exclude the possibility of such things
as tight binary stars, in which two of the masses orbit around each other in a
tight near-Keplerian orbit while the rest of the masses recede away.)

Definition 1 Let us call a solution x(t) bounded if it exists for all time t, and
if there exists a positive constant I0 such that kx(t)k2 < I0 for all time.

We can then restate the result of the above paragraph as


Theorem 1 (Jacobi) Any bounded solution to Newton’s N-body equations has
negative energy.
These are the solutions which we are most interested in.
Remark This contrast between positive and negative energy persists into quan-
tum mechanics. The classical N-body problem is replaced by the problem of
understanding the spectrum and spectral decomposition of a certain differential
operator, the Schrodinger operator with potential −U , on Q. Every discrete
eigenvalue of this operator is negative. The positive spectrum is all continuous,
corresponding to ‘waves’ coming in from infinity and receding back again.
Combining the three identities we can prove a fundamental result attributed
to Sundman, or to Weierstrass.

11
Theorem 2 If a motion tends to n-tuple collision in finite time then its angular
momentum must be zero.

Proof Suppose that limt→t0 x(t) = 0 where 0 denotes the N-tuple collision.
Observe that x = 0 if and only if I = 0. So it will suffice to show that if
limt→t0 I(t) = 0 then J = 0.
Now h = T − U so that
kJk2
4h + 2U = 2h + 2T ≥ 2h +
2I
by the third identity. Thus
d2 I kJk2
2
≥ + 2h.
dt 2I
We will show momentarily that I˙ < 0 in some neighborhood of t0 . Assume this
fact momentarily and multiply the previous inequality by −2I˙ to get
d ˙2 d ˙
− (I ) ≥ −2kJk2 log(I) − 4hI.
dt dt
d ˙2 2 ˙
(We have used dt (I ) = 2I˙ ddt2I and d
dt log(I) = II .) Integrating then

˙ 2 + I(a)
˙ 2 ≥ 2kJk2 log I(a)
−I(t) + 4h(I(a) − I(t)),
I(t)
where a < t < t0 are any two times. Re-arranging and dropping some positive
terms
˙ 2 − 4h((I(a) − I(t)) ≥ 2kJk2 log I(a) .
I(a)
I(t)
˙ 2 −4hI(a),
Fix a and let t → t0 . Then the left hand side tends to the constant I(a)
but the right hand side tends to ∞ if kJk 6= 0. This contradicts the inequality,
and proves that we must have J = 0 if I(t) → 0.
Finally, to see that I˙ < 0 near t = t0 , use the first identity again and
U (0) = +∞ to get that the 2nd time derivative of I is uniformly positive near
t = t0 . This combined with the fact that I is tending to its minimum tells us
that it is decreasing monotonically with I˙ < 0.

1.5 Eventual fate


Motions can be classified according to their eventual fate or their ancient history.
If there is no triple collision and if the energy is negative then there are only three
possibilities for the eventual fates. Either the the motion remains bounded ((B),
for bounded), or two of the masses form a tight binary while the third moves
away from them ((T) for tight), or the motion oscillates with masses alternately
coming close together, then moving very far away (O for oscillatory). Let us be
more precise:

12
Background Theorem 2 Every three-body motion without triple collision ex-
hibits one of the three following behaviours as t → ∞.
• Bound (B+ ): There is a positive constant C such that for all mass pairs
ij and positive times t > 0 we have rij (t) < C.
• Tight binary (Tij+ ): For one pair ij there is a positive constant C such
that for for all positive times t > 0 we have rij (t) < C, while for the
other two pairs, ik and jk, we have limt→+∞ rik = limt→+∞ rjk = +∞.
We think of this as if the i and j masses form a a ‘binary star’ which is
travelling away from mass k.
• Oscillatory (Os+ ): There is a pair ij such that lim supt→+∞ rij (t) = ∞
while lim inf t→+∞ rij < ∞
We have an analogous theorem for the distant past, t → −∞, in which case
we use subscripts “−”, as in B− , T− , Os− , to denote the distant past history
of an orbit. This theorem is attributed to Chazy, 1922. He actually found
seven classes of eventual fates. Only three of his classes can occur for negative
energies. For example, if the energy is positive it can happen that all interpar-
ticle distances rij tend to infinity, whereas this can never happen for negative
energies.
Following Arnol’d, we construct a 3 by 3 matrix, with entries indexed by
B, T and Os, according to what is known about the negative-energy motions
with these attributes.

B T+ Os+
 + 
B− + 0 0
T−  0 + 0 
Os− 0 0 ?
All entries occur for some trajectory. An entry with a ‘+’ means the set of
orbits of this type has positive measure. An entry with a zero means an entry of
this type has zero measure. The question mark for the ‘Os-Os’ entry means that
the measure of these orbits is unknown. See Arnol’d for references concerning
the entries.
Our interest is mainly in the B− − B+ entry, for any periodic orbit is of this
type. The ‘+’ in this entry was one of the first major applications of the KAM
theorem (Arnol’d, 1963).

2 The reduced dynamics


People use the term “reduction” to mean any procedure to simplify a system
by getting rid of the easily understood variables, thereby distilling a problem
to its essentials. Currently, reduction has a precise meaning within symplectic

13
geometry and Hamiltonian dynamics. (This is the symplectic or Marsden-
Weinstein reduction.) We will instead pursue a Riemannian version of reduc-
tion.
We begin by abstracting the N-body problem. We are given a Riemannian
manifold Q together with a Lie group G of isometries and a (negative of a) po-
tential function U , which is invariant under G. Then we can formulate Newton’s
equations on Q, and if γ is any solution, so is gγ for g ∈ G. Our immediate
goal is to reformulate Newton’s equations on Q as equations on the quotient
space Q/G. If G is Abelian we can realize this goal. The result is a family of
Newton-like equations, the family being parameterized by the possible values of
the constants of the motion – the angular momentum.
We recall that a point in the quotient space Q/G corresponds to a G- orbit
in Q. The quotient space is a metric space with the distance between two points
being the distance between their corresponding orbits in Q. This metric will
‘almost’ be that of a Riemannian metric as we will see. The points where it fails
to be Riemannian correspond to group orbits which are smaller than normal –
that is whose points have extra symmetry.
We will now follow this reduction procedure for the case of planar three-
body problem, ie Q = R2 × R2 × R2 and G = E(2). In this case the quotient
space Q/G is the space of congruence classes of triangles. So we write
C for the quotient space Q/G. We know from high-school geometry that C is
a three-dimensional set, being parameterized by the three side-lengths of the
triangle (SSS) or by an angle, side and angle (ASA), etcetera. But we were
not taught the metric on C in high school. It is a quite remarkable metric
and we will need to understand it well, with explicit coordinates and analytic
expressions.
We proceed analytically. We have already gotten rid of translations by going
to center of mass coordinates. The resulting configuration space is a real four-
dimensional inner product space with squared norm kxk2 = Σma kxa k2 = I(x).
If we think of it as a Riemannian manifold, its metric is then ds2 = Σma kdxa k2 ,
i.e. twice the kinetic energy. We may diagonalize the quadratic form I
I = |z1 |2 + |z2 |2 .
by a linear change of coordinates (x1 , x2 ) 7→ (z1 , z2 ). We think of the za as
complex variables whose modulus is denoted |z|. The standard diagonalization
is centuries old and is due to Jacobi. Choose two of the masses, say 1 and
2. Construct their relative position vector ζ1 = x2 − x1 and center of mass
vector, X12 = (m1 x1 + m2 x2 )/M12 with M12 = m1 + m2 . Now construct the
vector x3 − X12 joining this center of mass to the 3rd mass. Jacobi computed
I = µ1 |ζ1 |2 + µ2 |ζ2 |2 , with µ1 = mM1 m 2
and µ1 = M12 m3
M √, M = m1 + m2 + m3
12 √
being the total mass. If follows that z1 = µ1 ζ1 , z2 = µ2 ζ2 are diagonalizing
variables.
We are now identifying the inner product space Q with C2 , two copies of the
complex plane, together with its standard Hermitian structure. Rigid rotations

14
of the triangle x by an angle θ correspond to (z1 , z2 ) 7→ (eiθ z1 , eiθ z2 ). So, our
quotient space C = C2 /S1 . Its metric structure is well-known, being intimately
related to the Hopf fibration S 3 → S 2 . In order to describe it we will need the
following definition:

Definition 2 Let (X, d2 sX ) be a Riemannian manifold. Then the cone over


X, denoted C(X) is the usual topological cone: C(X) = ([0, ∞) × X)/(0 × X)
together with the metric dr2 + r2 d2 sX . Here r ∈ [O, ∞), and the quotient by
0 × X means that all points in [0, ∞) × X of the form (0, x) are to be identified
with each other. The resulting single point with r = 0 in C(X) is called the cone
point.

Example The cone over a circle of radius r0 ≤ 1 is isometric to the standard


cone made out of a piece of paper. The circumference of the circle is 2πr0 so
the cone is constructed from the wedge 0 ≤ θ ≤ 2πr0 (in polar coordinates) of
the plane and gluing the edges together. This cone isometric to the standard
Euclidean plane if r0 = 1. If r0 = 21 then it is the cone obtained by taking a
single sheet of paper and rolling it into a cone by gluing along an edge. The later
space is also isometric to the quotient of the usual plane by the group x → −x
of half-twists. In general the cone can be expressed in Euclidean three-space as
1
the surface x2 + y 2 = a2 z 2 , z ≥ 0, with r02 = 1+a 2 . If r0 > 1 then it cannot be

isometrically embedded in three-space.

Background Theorem 3 The quotient configuration space C = Q/G = C2 /S1


for the planar three-body problem is the space of congruence classes of (weighted)
planar triangles. It is isometric to the cone C(S 2 ( 21 )) over the sphere S 2 ( 12 ) of
radius 12 . The cone point corresponds to triple collision. The distance R be-
p
tween any given point w ∈ C and the cone point is I(x) where x ∈ Q is any
representative of the congruennce class w. C can in turn be identified with IR3
with either of the following metrics
1
ds2 = |dw|2
4|w|
or
1
ds2 = dR2 + R2 [dφ2 + sin2 φ dθ2 ].
4
In this first expression w denotes a vector in IR3 , and | · | the standard Euclidean
norm. The variables w ∈ C = IR3 and z ∈ C2 can be related by writing: w =
(w1 , w2 , w3 ) = (w1 , w2 + iw3 ) and setting w1 = 12 (|z1 |2 − |z2 |2 ), w2 + iw3 = z1 z̄2 .
The variables (R, φ, θ) are standard spherical coordinates in IR3 .
In particular, the orthogonal group O(3) of linear distance preserving maps
of IR3 is also an isometry group of C.

15
Sketch of proof. C2 = IR4 is isometric to the cone C(S 3 ) over the stan-
dard 3-sphere S 3 of unit radius. (This is a re-interpretation of standard spher-
ical coordinates.) In general, if G acts isometrically on Riemannian Y then
C(Y )/G = C(Y /G) as metric spaces. Thus C2 /S1 = C(S3 /S1 ). The fact that
S 3 /S 1 = S 2 is well-known. It is less well-known that the radius of this S 2 is
1 3 3 1 2
2 . The quotient map S → S /S = S is the Hopf fibration. It is given by
our coordinate expression for the map z 7→ w,uupon restring to the 3-sphere
I(z) = 1. The fact that the S 2 has radius 12 corresponds to the fact that by
travelling half-way around the S 3 along any great circle, we connect a point z
to its antipodal point −z. Now −z represents the triangle z after rotation by
180 degrees, since −1 = eiπ . So these two points lie on the same G-orbit, and
the resulting path on the S 2 has travelled full-circle. If the original half-great
circle is chosen orthogonal to the S 1 orbits on S 3 , then the projected curve w(t)
traverses a full great circle. We have travelled all the way around the ‘earth’
on a great circle in our reduced two-sphere: spherical distances are halved upon
going to the quotient. QED
Remark This theorem can be found in the paper of Iwai. It was also known
to Lemaitre, and Deprit. Probably it was known in some guise to Euler and
Lagrange.
Random facts regarding this metric, and coordinates on C.
The height coordinate w3 is essentially the area A of the triangle. Precisely:
r
m1 m2 m3 2A
w3 = √ .
M I
It is important that this is a signed area. The area of triangle and its reflection
are negatives of each other. The operation of reflection of triangles in the plane
corresponds to the map (w1 , w2 , w3 ) 7→ (w1 , w2 , −w3 ) of C which is an isometry.
The equatorial plane w3 = 0 is the fixed point set of this reflectional isometry,
and it corresponds to triangles of zero area, which is to say, configurations where
all three masses lie on a single line. This equatorial plane is separated into three
sectors by the binary collisions, the triangles two of whose vertices are equals.
These form three dilational rays in the cone. The angles between these rays
are computable function of the masses. We refer to Hsiang or Lemaitre for the
explicit functions. In spherical coordinates the equatorial plane is defined by
φ = 0 and the three binary collision rays are defined by the further specification
θ = θij . We are interested in solutions without collisions, so that we
want to avoid these rays as we move about in C. (If we use the Jacobi
z’s, obtained by focussing on masses 1 and 2, then θ12 = π, corresponding to
the 12 collision. The precise choice of xy axis however will not be important as
the rotation group acts isometrically on C.)
The interparticle distances rij are well-defined functions on C. In fact, away
from the equatorial plane they serve as alternative coordinates. They are related

16
to our spherical coordinates by:
s
mi + mj
q
rij = R 1 − sin φ cos(θ − θij ).
2mi mj

Lemaitre has expressions for the metric in these rij coordinates. Hsiang has
expressions for the metric in terms of the invariants kx1 k2 , kx2 k2 , kx3 k3 .
We now return to our original “immediate” goal: rewriting Newton’s equa-
tions on C. It turns out that this is very simple for angular momentum zero,
but more complicated for nonzero angular momentum. We pause momentarily
to describe what we mean by “zero angular momentum” in the general context.
A differentiable curve q(t) in Q will be said to have zero angular momentum if
it is everywhere orthogonal to the G-orbits.
Exercise Show that for the N-body problem this notion of ‘zero angular mo-
mentum’ is equivalent to the earlier notion, J(x, ẋ) = 0 where J = Σma xa ∧ ẋa
is as before

Background Theorem 4 The projection of any zero angular momentum so-


lution q(t) on Q to a curve on C = Q/G satisfies Newton’s equations ∇ċ ċ = ∇U
where ∇ is the Levi-Civita connections for the metric on C and U is the neg-
ative of the original G-invariant potential, but now viewed as a function on C.
Conversely, every solution to these Newton’s equations is the projection of such
a curve on Q.

We may return to the reasoning underlying this background theorem later.


But right now we have a sufficiently detailed description of the reduced geometry
and its dynamics to describe our main goal more precisely. Observe that the
homotopy type of the reduced configuration space C for the planar three-body
problem is that of two-sphere minus three points. So its fundamental group is
enormous: it is the free group on two letters.
Goal : Let C ∗ = C \ collisions. Given a free homotopy class in C ∗ , find a
reduced periodic zero-angular momentum collision-free solution in this homotopy
class which minimizes the reduced action among all loops in C ∗ in this class.
We recall that the set of free homotopy classes of a space X is in one-to-
one correspondence with the set of conjugacy classes of its fundamental group,
π1 (X). These conjugacy classes for our X = C ∗ can be nicely represented in
terms of eclipses or syzygies. We say that an “eclipse” has occured at a
certain instant t0 of time if at that instant the three masses are collinear. We
say that the eclipse is of type j , j = 1, 2, 3, if at the eclipse time mass j lies
between the other two masses. Use the letters A, B,C to stand for eclipses
of type 1, 2 or 3. Imagine a 3 body motion x(t) in which the masses never
collide, and imagine its projected curve w(t) moving in the three-dimensional
space C. As we will see, if x(t) is a zero-angular momentum solution to the

17
three-body equations then it is either everywhere transverse to the collinear
submanifold (the equatorial plane), or lies completely within it. Curves w lying
within the equatorial plane cannot represent nontrivial homotopy classes., So
we may suppose that w(t) is everywhere transverse to the equatorial plane. The
three sectors of this plane are to be labelled A, B, C corresponding to the three
types of eclipses. Every time the curve crosses sector α, write down the letter
α. If it crosses from the lower hemisphere into the upper hemisphere put a
superscript + over the letter. If it crosses from the upper hemisphere into the
lower hemisphere put a superscript − over it. In this way we have associated to
any solution curve a word of the form A+ B − C + A− . . .. in the letters A, B, C.
The subscripts must be arranged so that a + is always followed by a - and a
- by a +. Now we will subject the words to the “spelling” rule that we may
never have the same letter occurring two times in a row. That is to say, if two
A’s appear in a row we erase them. For example, the word C + A− A+ B− is to
be replaced by C + B − . This rule encodes the fact that any loop which dips
across the equator without ‘looping’ around a collision is homotopic to one in
which this extraneous loop is deleted. As we hope to show later, this rule has a
variational counterpart in terms of the principle of least action, or Mauperitus’
principle. If the curve w(t) is periodic then the associated word will repeat
after some period. The corresonding finite word will also have an even number
of letters, since it must cross the equator an even number of times to close up.
This representation of free homotopy classes in terms of words is unique up to
cyclic order of the letters: ABCB = BABC = CBAB = BABC. (Changing
the order corresponds to changing the starting point on the closed curve w(t).
Reversing the sense of time, i.e., inverting the homotopy class, corresponds to
reversing the parity of all the superscripts. But since the dynamics (with J = 0)
is time-reversible, if we have a periodic solution for one homotopy class we have
one for the inverse class. So we can forget about the signs. Our goal then, is to
solve:

The homotopy word problem Given any finite word β in three letters of
even length subject to our spelling rule, find a solution to Newton’s equations
with zero angular momentum whose sequences of eclipses realizes this word.

An even more interesting thing to do would be to solve the same problem


for bi-infinite words!
In any case, we are able to solve the problem, provided we cheat.
Definition 3 We say that the potential V = −Σfij (rij ) is a strong force
potential provided the fij are positive for r > 0 and that there are positive
(perhaps very small) numbers δ , c such that r < δ implies that fij (r) > c/r2 .

18
Thus the Manev potential f = 1/r + /r2 is a strong force potential. More
generally, we could replace  by a bump function (r) with arbitrarily small
support r < δ near r = 0.

Theorem 3 If the potential is a strong force potential then our homotopy word
problem is solvable for all words β except those words which only contain two of
the three letters.

Remark The Newtonian gravitational potential is not a strong force potential.


Almost every paper in which variational methods are applied to the N-body
problem makes the assumption that the force is strong in our sense, or something
morally equivalent to it. The main remaining problem for us is to relax the
strong force assumption so as to allow for the Newtonian potential.
Discussion re:
tight binaries ...
kepler orbits at infinity...
braids...
...

2.1 Geometric miscellany



Returning momentarily to the general conical metric, we refer to the ∂r direc-
tion as the dilational direction, or the radial direction, or the homothety
direction. It corresponds to dilating the triangle z. A moving frame computa-
tion allows us to compute the sectional curvatures of the conical metric relative
to that of the original metric, for any Riemannian manifold X. In the conical

metric, the sectional curvature of any plane containing the radial direction ∂r
is zero. A two-plane orthogonal to this direction can be viewed as a two-plane
in Tx X. Its sectional curvature as a two-plane in T C(X) equals r12 (KX − 1)
where KX denotes its sectional curvature in X. Thus, in our case the sectional
∂ ∂ ∂ ∂
curvatures of planes of the form ∂r ∧ ∂φ and ∂r ∧ ∂θ is zero, whereas the sec-
tional curvature of the plane tangent to the sphere R = R0 is R32 . In contrast,
0
if we first restrict the conical metric to the sphere R = R0 and then compute
the curvature we get R42 .
0
Given any curve in X we have an associated submanifold C(γ) in C(X).
Its induced Riemannian metric is that of two-dimensional cone, and so is a flat
metric.
One of the overall goals in this development is to reduce all the
questions regarding the three-body problem to questions regarding
curves on the two-sphere.
Remarks In the three-body case, the sphere S, over which C is a cone, is
described by Moeckel [?] and Saari [?]. This sphere S is essentially the triple

19
collision manifold appearing in the McGehee transformation. In McGehee trans-
formation based literature the metric structure of the sphere or of C are not
used. The metric structure of C was uncovered by Iwai [?] for general N, al-
though it was very likely known to earlier authors. I used it in a paper on
the falling cat problem [?]. It is discussed in the survey article by Reinsch and
Littlejohn [?]. Hsiang is, to our knowledge, the first author to propose using the
metric structure of C to say something serious about the three-body problem.

3 Collisions
The flow on Q × Q = IR2d(N −1) defined by Newton’s N-body equations is not
complete. There are two sources for infinities in the solutions. One is spatial
infinity, corresponding to one or more of the rij → ∞. The other is the infinite
forces occuring in the right-hand side of the equations when one of the rij → 0.
We say a binary collision occurs along a solution if exactly two of the masses
collide while the rest remain bounded away from each other. These collisions are
relatively innocuous. Any solution with a single binary collision can be analyt-
ically extended through the collision by rescaling time and defining a branched
cover of configuration space, branched along the binary collision manifold. This
procedure is uniform with respect to nearby trajectories, resulting in a well-
defined flow on the branched configuration space. We will discuss the procedure
later.
Triple collisions cannot be regularized. This is closely related to chaos in the
three-body problem.
For the N-body problem, N ≥ 5 there are solution singularities which are
more complicated than simple collisions. Recently Jeff Xia proved that when
N = 5 there are solutions x(t), a ≤ t < t0 , t0 a finite time, whose blow up is of
the following type: lim supt→t0 rij (t) = +∞, while lim inf t→t0 rij (t) = 0.
For the three-body problem such blow-ups cannot occur. The only solutions
which cannot be analytically continued past some time t0 are those which tend
to triple collision at t0 . This is the fundamental result of Sundman. Diacu
[?] argues that for this reason Sundman has “solved” the three-body problem,
at least in the sense of the original problem posed by Mittag-Leffler. It follows
from this and the Sundman-Weierstrass theorem, that if a solution has non-zero
angular momentum then its domain of definition can be taken to be all time.

4 Literature
Straight 3 body stuff:
V. Arnold, Kozlov and Neishtadt, mathematical Aspects of Classical
and Celestial Mechanics, 2nd ed., ch. 2, in the series Encyclopaedia of
Mathematical Sciences, Springer, , 1997.

20
G. Birkhoff, Dynamical Systems ch. 8, §11, AMS, (1927).
F. Diacu, The planar isosceles problem for Maneff ’s gravitational law J.
Math. Phys., v. 34, no. 12, 5671-5690, (1993).
T. Levi-Civita, Sur La Réqularisation du Problème des Trois Corps, Acta
Math., v. 42, 99-144; esp. p. 105, eq.(12); (1921).
C. Marchall,The Three-body problem,Elsevier pub., 1990.
too much stuff, and some of it wrong, but still useful.
G. Maneff, La gravitation et l’énergie au zéro, Comptes Rendus, v. 190,
1374-1377, (1930).
R. Moeckel, Some qualitative features of the three-body problem in Hamil-
tonian Dynamical Systems, proceedings, AMS Contemporary Math,
Vol. 81, ed. K. Meyer and C. Saari, 1-22, American Math. Society, (1988).
H. Poincaré, Les Methodes Nouvelles de la Mécanique Céleste v. 1,
ch. 3, Paris, Gauthier-Villars et fils, (1892)
see also
introduction to English translation by D. Goroff, p. I 42. AIP, History of
Modern Physics and Astronomy v. 13, (1993).
This is the source!
C. Siegel, Lectures on the Singularities of the Three body problem
Tata Institute, vol. 42., 1967.
K. Sundman, Mémoire sur Le Problème des Trois Corps, Acta Math., v. 36,
105-179, (1912).
A. Wintner, The analytical foundations of celestial mechanics, Prince-
ton Univ. Press, (1941).
Z. Xia, Arnold Diffusion and Oscillatory Solutions in the Planar Three-Body
Problem, J. Diff. Eq. , v. 110, 289-321, (1994).
Variational approach to periodic orbits:
In almost all of these the ‘strong force’ assumption, which excludes the
original Newtonian gravitational potential is made. Often it is a bit hidden,
eg. as ‘assumption V6’.
A. Ambrosetti, V. Zelati, Periodic Solutions of Singular Lagrangian
Systems, Birkhäuser, Boston, 1993.
P. Majer and S. Terracini, Periodic solutions to some problems of n-body
type, Arch. Rat. Mech., and references therein, 124, no. 4, 381-404, (1993).
A. Bahri and P. Rabinowitz, Periodic solutions of Hamiltonian systems of
3-body type , Annales Inst. Henri Poincaré, Analyse non linéaire, v. 8, no.
6,561-649, (1991).
W. Gordon, A minimizing property of Keplerian orbits, Am J. Math., v. 99,
no. 5, pp961-971, 1977.
This contains many of the main ideas of my periodic N body paper, and the
results in a primitive form. The new ingredients I add are a detailed under-
standing of the reduced configuration space, and the relation to braids.

21
R. Mañe, Global Variational Methods in Conservative Dynamics,
18th Colóquio Brasileiro de Matemática, available from IMPA,
L. Sbano, On the Lack of Coercivity of the reduced Action-functional for zero
total angular momentum in the planar three body problem, SISSA/ISAS preprint
185/94/FM, (1994).
Sbano proves that if we slightly perturb a path going through collision,
keeping the endpoints fixed, then the action can be made to decrease.
Topology and Symbolic dynamics:
V.I. Arnol’d, The Cohomology of the Colored Braid Group, Math Zametki
5, no. 2, 227-231; English translation: Math. Notes. 5, 138-140 (1969).
Apparently first computation of this cohomology.
P. Boyland and C. Golé, Lagrangian Systems on Hyperbolic Manifolds; Dy-
namical Stability in Lagrangian Systems SUNY StonyBrook Inst for Math. Sci.
preprint 1996/1, (1996).
They show that if a compact manifold admits a hyperbolic metric then
all natural mechanical systems on it have embedded within them subsystems
conjugate to the hyperbolic geodesic flow.
J. Hadamard, Les surfaces à courbures opposeées et leur linges géodesiques,
J. de Math., ser. 5, v. 4, (1898).
Perhaps the first paper using symbolic dynamics.
C. Series, ch. 5 of: Ergodic Theory, Symbolic Dynamics, and Hy-
perbolic Spaces, T. Bedford, M. Keane, and C. Series ed., Oxford U. Press,
(1991).
Historical:
F. Diacu and P. Holmes, Celestial Encounters, Princeton Univ. Press,
1997.
fun reading
F. Diacu The Solution of the N-body problem, Intelligencer, 1996 or 1997,
Geometric, or ‘Riemannian Submersion’ approaches:
A. Guichardet, On rotation and vibration motions of molecules, Ann. Inst.
H. Poincare, Phys. Theor., 40, no.3, 329-342, (1984).
W-Y Hsiang, Geometric Study of the Three-Body Problem I, report PAM-
620, Center for Pure and Applied Math, University California, Berkeley, 1994.
W-Y Hsiang and E. Straume, , Kinematic Geometric of triangles with given
mass distribution,report PAM-636, Center for Pure and Applied Math, Univer-
sity California, Berkeley, 1995.
W-Y Hsiang, Kinematic Geometric of mass triangles and the three-body prob-
lem in quantum mechanics, report PAM-682, Center for Pure and Applied Math,
University California preprint, Berkeley, 1996.
T. Iwai, A geometric setting for classical molecular dynamics, Ann. Inst.
Henri Poincare, vol. 47, no. 2, pp 199-219. (1987).

22
T. Iwai, A geometric setting for internal motions of the quantum three-body
problem J. Math. Phys., v. 28, 1315-1226, (1987).
R. Montgomery, The geometric phase of the three-body problem, Nonlinearity,
(1996).
R. Montgomery, The N-body problem, the braid group, and action-minimizing
periodic solutions., preprint (submitted to Nonlinearity.
M. Reinsch and R. Littlejohn, Gauge fields in the separation of rotations and
internal motions in the n-body problem, Rev. Mod. Phys. , maybe 1996?
D. Saari, From rotation and inclination to zero configurational velocity sur-
faces, I, a natural rotating coordinate system, Celestial Mechanics, 33, 299-318,
(1984).
D. Saari, Symmetry in n-particle systems, in Hamiltonian Dynamical
Systems, proceedings, AMS Contemporary Math, Vol. 81, ed. K.
Meyer and C. Saari, 23-42, esp. 26-27, American Math. Society, (1988).
In
more geometry in the three-body
Sydlowsid, Heller, Sasin, Geometry of spaces with the Jacobi metric, J. Math.
Phys., 37, 1, 1996
Sydlowsid, Curvature of gravitationally bound mechanics systems, J. Math.
Phys., 35, 4, 1994
The author gives expressions for the sectional and Ricci curvatures for the
Jacobi metric (unreduced) and shows that as N goes to infinity it is dominated
by negative curvature.

23

Potrebbero piacerti anche