Sie sind auf Seite 1von 15

Journal of Dynamics and Differential Equations, Vol. 6, No. 1.

1994

Linear Stability of Relative Equilibria


with a Dominant Mass
Richard M o e c k e l t

Received November 23, 1992

A criterion for the linear stability of relative equilibria of the Newtonian n-body
problem is found in the case when n - 1 of the masses are small. Several stable
periodic orbits of the problem are presented as examples.
KEY WORDS: Celestial mechanics; relative equilibria; stability.

1. I N T R O D U C T I O N
In 1772, Lagrange discovered his remarkable equilateral periodic solutions
of the planar three-body problem (4). For any choice of the three masses,
there is a periodic solution for which the configuration of the bodies is
always an equilateral triangle which rotates rigidly about its center of mass.
Later, Routh showed that these periodic orbits are linearly stable if one
mass is much larger than the other two (9).
In 1859, James Clerk Maxwell published his study of the rings of
Saturn (5, 6). As a first approximation, he treats the ring as a rigidly rotating regular polygon of n equal masses. He carries out a complete analysis
of the linear stability of this periodic orbit and finds that such a ring is
linearly stable provided the central mass is sufficiently large compared
to the masses in the ring. This is indeed the case, but only under the
additional assumption that n I> 7. A minor error in computing one of the
characteristic exponents caused Maxwell to miss this necessary condition.
Both the equilaterial triangle and the regular n-gon with a central
mass are examples of relative equilibria, that is, configurations which

School of Mathematics, University of Minnesota, Minneapolis, Minnesota 55455. E-mail:


rick@math.umn.edu
37
1040-7294/94/~100-0037S07.00/09 1994 Plenum Publishing Corporation

38

Moeckel

become equilibria of Newton's differential equations in uniformly rotating


coordinates. The concept of relative equilibrium occurs in Lagrange's work.
However, for n i> 4 it is very difficult to find them, much less to analyze
their linear stability. The exceptions are the highly symmetrical relative
equilibria, like the regular polygon with a central mass, which occur when
the masses are equal. Some progress has been made in finding and analyzing the stability of relative equilibria of the four-body problem (1, 2, 8, 10).
All of the linearly stable relative equilibria of which the author is
aware occur when one mass is much larger that the others. Hall studied the
limiting situation when one mass interacts with n infinitisimal ones, the
so-called (I +n)-body problem (3). He derives equations which a configuration of n + 1 bodies must satisfy if it is the limit of a sequence of
relative equilibria with n masses tending to zero. These equations are
shown to be the equations for critical points of a certain potential function.
When the n small masses are equal and when n is sufficiently large, Hall
shows that the only possible limiting configuration for which the bodies do
not coalesce is Maxwell's ring. However, for small n, other configurations
are possible. An example is Lagrange's equilateral triangle.
The goal of this paper is to give a criterion for the linear stability of
relative equilibria of the (n + 1)-body problem with n small but not
necessarily equal masses. Suppose the small masses tend to zero in such a
way that their ratios converge to nonzero real numbers and that the configurations converge to a relative equilibrium of the (1 + n)-body problem
without coalescing. Then the rehtive equilibria with sufficiently small
masses are linearly stable provided that the limiting configuration is a
nondegenerate minimum of Hall's potential function.
It turns out that Maxwell's ring is a minimum if and only if n I> 7.
Since there is always at least one minimum, this suggests the existence of
other linearly stable relative equilibria for n ~<6. By means of a numerical
search for minima, Hall found configurations which are generalizations of
Lagrange's equilateral triangle in that the n-small masses are gathered
together on one side of the central body. Some of these are illustrated in
Section 4. Numerical computation of their Hessian's shows that these
configurations are nondegenerate minima. The results below show that the
relative equilibria which converge to these limiting configurations are
linearly stable when the masses are small enough.

2. RELATIVE EQUILIBRIA
Consider the Newtonian (n + 1)-body problem in the plane. Let the
mass and position of the ith body be denoted m~ and qiER 2, i = O . . . n .

Linear Stability of Relative Equilibda

39

Introducing the momenta, P~=mi(li, the equations of motion can be


written in Hamiltonian form,

= n,(q, p) = M - ~
p = --Hq(q, p)=VU(q)
where q = ( q o ..... qn) e R 2"+~, p = ( p o ..... Pn)~R2~+2,
Hamiltonian function,

H(q, p)= 8 9

H(q,p)

is the

U(q)

and U(q) is minus the Newtonian potential function,

U(q) = E mirn/
i < j rr
Here rr = l q ~ - qj[ is the Euclidean distance between two of the bodies.
A relative equilibrium is a configuration x which becomes an equilibrium of Newton's equations in a uniformly rotating coordinate system.
Let R(t) denote the linear operator on R ~+2 which rotates each pair of
coordinates counterclockwise by angle t. Setting x = R(t)q and y = R(t) p,
one finds that x and y satisfy Hamilton's equations,

= M - l y -- Kx
~=VU(x)-Ky

(1)

with Hamiltonian function,

H(x, y) = 89y r M - l y + xrKy - U(x)


where K is the (2n + 2) x (2n + 2) block diagonal matrix with 2 x 2 blocks
[o -~]. Introducing z = (x, y ) ~ R 4n+4 one can write Hamilton's equations
as

= JVH(z)

(2)

where J is the (4n + 4) x (4n + 4) matrix

,[o o']
and I denotes an identity matrix.
A restpoint of Eqs. (1) satisfies y = MKx and
VU(x) + M x = 0

(3)

40

~oeekel

A configuration, x, satisfying Eq. (3) is called a relative equilibrium of the


(n + 1)-body problem. Because of the complexity of the equation, very few
relative equilibria are known explicitly.
G. R. Hall studied the limiting case when all but one of the masses
tend to zero (3). Suppose mo= 1 and mi=81z~, 1 <~i<~n, where the /t I are
fixed constants and e > 0 is a small parameter. If x ~ is a relative equilibrium
for a sequence of positive e's tending to 0 and if x ' ~ 2 a s B --*0, then 2 is
called a relative equilibrium of the (1 +n)-body problem. Usually, the
dependence on 8 will be suppressed.
Assume that the center of mass is at the origin so that
Xo = - 5 ~. iztxi

(4)

i~l

Hall showed that for such a limiting configuration, 20 = (0, 0) and ~ lies
on the unit circle for 1 ~<i ~<n. It is possible that several of the small masses
coalesce in the limit but it is assumed here that ~ ~-2j for i # j.
To prove Hall's results in this special case, consider the ith pair of
components of Eq. (3) divided by ms:

j,,i

mj(xj~- xi) -I- x I ~- 0


r

(5)

If i = 0, then using mj--- e/aj together with the fact that the limiting positions
are distinct gives
Xo = O ( t ) ,

20 = 0

If i # 0, then taking the inner product with x~ gives

Since 2~# (0, 0) and r~o= Ix,I + O(B), it follows that


Ix,I = 1 + O(~),

12,1 = 1

One can specify a relative equilibrium of the (1 + n)-body problem by


the n angles 0t such that $~= (cos 0~, sin 0~), 1 <~i<~n. Taking the inner
product of Eq. (5) with ( - s i n 0~, cos 0~), dividing by 8, and taking the limit
yields
/~j sin(0j- 0,) [ ~ j#i

I] = 0

(6)

Linear

Stability of Relative Equilibria

41

for 1 <~i<<.n. Hall observed that this equation can be viewed as the
equation for critical points of the function:

v(o)= 5".

+ - 5".

i<jre(O) 2i<j

(7)

where the ranges of the summation indices do not include 0 and r~(O)=
2(1-cos(01-0j)). Thus the relative equilibria of the (1 + n)-body problem
are exactly the critical points of V(O).
The Morse index of a critical point of V(O) is determined by the
Hessian quadratic forrn Voo. Because of the symmetry Of the problem with
respect to simultaneous rotation of all masses, this quadratic form always
has nullity at least 1. It is appropriate to call a relative equilibrium of the
(1 + n)-body problem nondegenerate if ~i # ~y for i # j and if the nullity of
V00 is equal to 1. In particular, 2 is a nondegenerate local minimum of V(O)
provided that Voo is positive semidefinite with nullity 1. It follows from the
implicit function theorem that any nondegenerate critical point, 2, is a
relative equilibrium of the (1 + n)-body problem, that is, a family of relative
equilibria, x *, converging to ~ exists.

3. LINEAR STABILITY
A relative equilibrium of the (n + 1)-body problem determines a
periodic orbit for which the configuration rotates rigidly. The main goal of
this paper is to study the linear stability of the periodic orbits associated
to relative equilibria in a family x" converging to a relative equilibrium,
~, of the (l+n)-body problem. Equivalently, one can study the linear
stability of the associated restpoints of Eq. (1).
Using the notation of Eq. (2), the linearized equations at a restpoint
z = (x, y) are
9 =Aw

(8)

where

Here A =JS, where S is the symmetric matrix SfDVH(z).


Normally, a restpoint is called linearly stable if 0 is a stable restpoint
of the linearized Eq. (8). However, the symmetries and integrals of the
present problem make it impossible to satisfy this condition.

42

Moeckel

Consider the four-dimensional subspace, W~, of C ~' + 4 spanned b y the


vectors
(x, 0)

(0, Mx)

(Kx, 0)

(0, KMx)

Then using the homogeneity of the Newtonian potential and the fact that
x is a relative equilibrium, one finds that W~ is an invariant subspace for A.
The eigenvalues of A I ,,1 are 0, 0, + i and there is a nontrivial Jordan block
associated to the repeated 0. Thus the restpoint is not linearly stable in the
conventional sense. However, this instability arises from the fact that the
given periodic orbit is part of a family of rigidly rotating periodic orbits
with different rotation frequencies; the angular positions of nearby
solutions in this family drift away from each other and this is reflected in
the nontrivial Jordan block.
A similar drift occurs in the four-dimensional subspace, W2, spanned
by
(r 0)

(0, Me)

(r/, 0)

(0, Mr/)

where ~ = (I, 0, I, 0,...) e C 2n+2 and !/= (0, 1, 0, 1,...) e C z~+2. This subspace
is also invariant and the eigenvalues of A [ ,,2 are + i, + i with a nontrivial
Jordan block. Clearly this is associated to a drift in the center of mass.
It is traditional in celestial mechanics to view the drifts in these two
subspaces as harmless. Indeed, they can be eliminated by fixing the
momentum, angular momentum and center of mass and passing to a
quotient manifold under the action of the rotational symmetry group. Thus
it is reasonable to formulate a definition of linear stability based o n the
behavior of A in a complementary subspace. To define such a subspace, it
is necessary to introduce the skew inner product of two complex vectors,
0, W ~ C4n+4:

~2(v, w) = vrJw
A matrix of the form A =JS with S r = S is called Hamiltonian. The
Hamiltonian property is equivalent to

Aw)= --~(Av, w)

(10)

Using this it is easy to show that the skew-orthogonal complement o f an


invariant subspace of a Hamiltonian matrix is again invariant. Let W
denote the skew orthogonal complement in C ~+4 of W~ ~ W2. Then W is
an A invariant subspace of dimension 4 n - 4. A relative equilibrium, x, is
called linearly stable if 0 is a stable restpoint of the restriction o f the
lincarized equation (8) to W. Furthermore, x is called nondegenerate if

Linear Stability of Relative Equilibria

43

A I w is nonsingular. With these definitions, the main result o f this paper


can be stated as follows.
Theorem 1. Let x ~ be a f a m i l y o f relative equilibria o f the (n + 1)body problem with masses m o = 1 and m~ = 8#t, 1,<~i <<.n. Suppose that as
8--*0, x ~ converges to a nondegenerate relative equilibrium, ~, o f the
(1 + n)-body problem. Then x ~ is nondegenerate f o r 8 sufficiently small. In
this case, x ~ is linearly stable f o r 8 sufficiently small if and only if R is a local
minimum o f V(O).

The proof is carried out in the rest of this section. The first step is to
establish the behavior of the eigenvalues of A ] ,, as 8 --* 0. Throughout the
proof, 8 is suppressed in the notation.
Let v E W denote an eigenvector of A with eigenvalue 2 = ~ + ifl.
Writing v = (w, w') with w, w ' ~ C 2"+2 and using (9) gives
w' = M ( K + 2 I ) w

(11)
Bw = 0

where
B = M - tD VU(x) - ( K + 21) 2
This reduces the problem from 4n + 4 to 2n + 2 dimensions. In what
follows, the quantity 12(v, ~), where ~ denotes the complex conjugate, is
significant. A short computation shows that
f2(v, ~) = 2 i f l w r M ~ - 2wrKM~:

(12)

Since v is skew-orthogonal to the subspace W2 defined above,


WO"~'8~IW l'Jt- "'"

dt-814nWn=O

where wi denotes the ith pair of coordinates of w. This equation can be


used to eliminate wo. If one thinks of B as a block matrix with 2 x 2 blocks
B u, O<~i,j<~n, then B w = O becomes /~v~=0, where ~ ' = ( w l ..... w , ) ~ C 2"
and B c = B o . - s # j B o i . Note that if the normalization /~1 [Wll2+ . . - +
~. Iw.l~= 1 is imposed, then wo = 0(8).
Recall that in the limiting configuration, ~, the n small masses lie on
the unit circle at angular positions 0i. This motivates replacing w~ by its
radial and angular components. Set
w t = pi(cos

Oi, sin 0~)+ zi(--sin 0i, cos Oi)

44

Moeekd

and let p = (Pi ..... p,) and z = (zt ..... x,). Then w=R(p, ~), where R is an
orthogonal 2nx2n matrix. Setting C f R - ~ B R , one finds, after some
computation, that
C=[

(3 - ~2)i+ 0(8)
- 2 M + O(s)

-22I+22/+O(8)s/z' Vse + O(e2)]

(13)

where/~ = diag(#~ ..... #,). Moreover, writing Eq. (12) in these variables and
dividing by ~ gives
1

-~2(v,O)=2ifl(pTl~#+Zrl~{)-2zr#:+2prlJs
8

(14)

where, the O(e) arises from the terms involving We. The normalization
condition adopted above can be written

prlx:+zrUg= 1

(15)

The eigenvector Eq. (11) becomes C(p, z)= 0. The determinant of C is


a polynomial, P(2), of degree 4n in A, which reduces to 22"(1 +22) n when
8 = 0. The roots are the 4n eigenvalues of the restriction of A to W ~ W1.
It follows that 2n of these eigenvalues converge to 0 and the other 2n
converge to + i as e-} 0. This splitting of the eigenvalues determines a
factorization P(2) = Pt(2) P2(2), where Pt(2) and P2(2) are polynomials of
degree 2n converging to 2 ~ and (I + 22)n, respectively.
To get more precise estimates of the eigenvalues, note that it follows
from C(p, z)=O together with Eqs. (13) and (15) that
(3 - 22)p + 2A~ = 0(8)
- 2 2 p - 22T = 0(8)

(16)

Elimination of z from these two equations gives 2(1 +22)p = 0(8), while
elimination of p gives 22(1-I-,~2)'c=O(8). It follows from these together
with (15) that
,~2(1 + ,~2) = 0 ( 8 )

Thus the eigenvalues of A are of the form 2 = +_i+_0(8) or 2 = O(x/r~).


To investigate further the latter case, set A= ~/8 ( in (13) and take the
determinant. After dividing by e', the result is a polynomial of degree 4n in
( which reduces to the polynomial det[3/z-lVe8 + (2I] of degree 2n when
8=0. This same polynomial can be obtained as the limit of 8 - ' P t ( x / ~ ()
since the limit of P2(x/~ () is 1. It follows that the eigenvalu~ of the form
A= ~ ( have the property that (2= ~ + o(1), where ~ is 'an eigenvalue of

Linear Stability of Relative Equilibria

45

- 3#-~ Voo. Moreover, every eigenvalue of - 3#-~ Voo of multiplicity k


occurs as the limit as 8 ~ 0 of ~2 for 2k eigenvalues of A. Note that the
signs of the eigenvalues of # - ~Voo are the same as those of the Hession Voo.
These observations about eigenvalues can be used to prove part of the
theorem. First note that by the nondegeneracy hypothesis on $, 0 is a
simple eigenvalue of Voo. Hence the multiplicity of 0 as an eigenvalue of A
on W ~ W~ is 2 for e sufficiently small. Since 0 already has multiplicity 2
as an eigenvalue of A on W~, it follows that A Iw is nonsingular and x" is
nondegenerate for e sufficiently small. Also, if $ is not a local minimum,
- 3 # -1Voo has a positive eigenvalue. This implies that for 8 sufficiently
small, there is an eigenvalue of A of the form 2 = x/~ (, where ( has a
nonzero real part. Hence x ~ is not linearly stable.
The rest of the proof relies on a lemma about stability of linear
Hamiltonian systems. Consider a linear differential equation of the form
(8), where A is a 2 m x 2 m Hamiltonian matrix. It follows from Eq. (10)
that

o(v, (A -hi)w)= -O((A + 2I)v, w)

(17)

for any 2 ~ C. Using this, one can prove the following well-known fact [7]:
if v and w are generalized eigenvectors of A corresponding to eigenvalues
2 and # and if 2 +/z # 0, then f2(v, w) = 0. The following lemma uses these
properties of Hamiltonian matrices to derive a simple stability criterion.
Lemma 1. Suppose A is a Hamiltonian matrix such that every eigenvector v ~ C ~ of A satisfies O(v, ~ ) ~ 0 (where ~ denotes the complex
conjugate of v). Then

all of the eigenvalues of A are imaginary;

every generalized eigenvector of A is an eigenvector; and

0 is a stable restpoint for the differential equation 9 = A w.

Proof. To prove the first claim, suppose 2 = ~ + i/~ is an eigenvalue.


If 2 were real, there would be a real eigenvector v and then 12(v, g ) =
[2(v, v) --- 0 by skew-symmetry. Hence/~ # 0. If ~ were not zero, then taking
# = 2 , one has 2 + / ~ = 2 ~ : ~ 0 . Applying the observation above about
generalized eigenvectors to v and w = ~ leads to a contradiction. Thus ~ = 0
and/~ # 0 and so 2 is imaginary.
Next assume that w is a generalized eigenvector which is not an eigenvector. Without loss of generality, one may assume that (A - 2 1 ) 2 w = 0 but
( A - 2I)w = v ~ O. Then v is an eigenvector for 2. Using Eq. (17),

~(v, ~)= -t~(~, v)= -~(~, (A -2I)w)= o((.4 + 2x)~, w)

46

Moeckel

Since )7 is an eigenveetor for ~ = - 2 , O(v, )7)-0, a contradiction. This


establishes the second claim. The third follows from the first two. II
Note that the lemma applies to A I w because A I w is Hamiltonian in
the sense that Eq. (10) holds for v, wv W. To complete the proof of the
theorem, it is now shown that if .~ is a local minimum of V(O), then for 8
sufficiently small, every eigenvector v of A Iw satisfies the hypothesis
I2(v, ~) # 0 of the lemma. The eige.nvectors associated to eigenvalues of the
types 2 = + i + 0(8) and 2 ffi O(~/8) are treated separately.
If 2 --- -J-i+ O(8), Eqs. (16) show that T -- +2ip + 0(8). Substitution of
this expression into Eq. (14) gives
1
- I2(v, g)---

+2iprlz~+ 0(8)

Moreover, Eq. (15) shows that 5prize= 1 +O(8). It follows that for 8
sufficiently small, t2(v, g) # 0 as required.
If 2=O(q/~), then 2 = q / ~ , where ( 2 = ~ + o ( I ) and r is a nonzero
eigenvalue of -3/~-1V~. Since $ is a local minimum, all of the nonzero
eigenvalues of the matrix -31z-lVoo are negative. Hence ~ - i v + o(1),
where ? is real and bounded away from zero. The imaginary part of ,t is
therefore of the form

(18)
Now Eqs, (16) show that p = O(V~ ). Setting p = ~/~ (7 in the equation

C(p, ~)=0 and using (13) gives


=

o(,/7)

Using this together with Eq. (18) in Eq. (14) gives

18ate,

= -

i vq

+ o(vq)

Since p = O(v/~), zr/zf ffi 1 + O(e) and so t2(v, ~ ) # 0 for 8 sufficiently small.
This completes the proof of the theorem.

4. EXAMPLES
This section contains several applications of the theorem.

47

Linear Stability of Relative Equilibria

4.1. Lagrange's Triangle


The equilateral triangle is the only noncoUinear relative equilibrium of
three masses. It was shown by Routh that it is linearly stable if and only
if
27(morn1 + morn2 + rnt m2) < (rno+ ml + m2) 2
It is easy to see that this is satisfied only if one mass, say mo, dominates
the other two. The calculation of this stability criterion is fairly involved.
On the other hand, it is easy to check that the equilateral triangle is a
nondegenerate minimum of the (1 + 2)-body problem. Taking 01 = 0 and
02 = (7r/3), one finds

89 "/3 - 89
J
which has eigenvalues 0 and #1/a2(89+ ,,//3)> 0 as required.

4.2. Maxwell's Rings


If #1 . . . . .
/~,, the regular n-gon with a central mass is a relative
equilibrium of the (n + l)-body problem for all e. For the sake of brevity,
this configuration is called simply the ring. Clearly the ring converges to
itself as e ~ 0 so it is also a relative equilibrium of the (1 +n)-body
problem. Hall observes that it is not a local minimum of V(O) when n ~<6
but that it is a local minimum when n = 7. It will now be shown that the
regular n-gon with a central mass is a nondegenerate local minimum of
V(O) if and only if n >17. If n ~<6 it is still nondegenerate, so it follows from
the theorem that in this case, the ring is not linearly stable no matter how
small e is chosen. This represents a correction to Maxwell's study.
Taking gi = 1, the first partial derivatives of V(O) are

Vo,= ~ sin(O,--Oj) [1
j= t
j#i

r#

where r~2 = 2(1 - c o s ( 0 , - 0j)). Differentiating again gives the entries of the
Hessian Voo

Vo,oj=

cos(0~- 0j)

2r3.

2r 3

Vo~o~=-- ~ Vo~oj
j=l
j#i

- cos(0,-- 0j)

for

i# j

48

Moeckel

This is a circnlant matrix, that is, Vo,oj-- Vo,+~o~+kfor all/, j, and k with
indices interpreted modulo n. It is possible simply to write down the eigenvalues of such a matrix. The eigenvectors take the form u~ffi (p, p2,..., pn)
where p -- e (2~u/n) is an nth root of unity. Since the nth coordinate of ut is
1, the corresponding eigenvalue is just the nth coordinate of Voou~. Since
the eigenvalues are real, one finds

n-I
ln-ll_c/cjl
3 ~zl l _ c j l
~.~= ~ c j ( 1 - c j t ) + ~ j ~
r3 + ~ L
~

j~t

jffil

r~

(19)

where cj=cos(21rj/n), %=cos(21rjl/n), and r~ffir~nf2(1-cj). Clearly


21= ).n - i.
Setting 1= 0 gives 2o = 0. The ring is nondegenerate provided 2~ # 0,
l= 1..... n/2 and is a nondegenerate local minimum provided 2~>0,
1= 1..... n/2. Consider the first sum in (19).
n~' cy(1-c/t)= ~ Cj(1--Cjl)=- ~ CjCjt
jffil
j=l
j~l
This Vanishes unkss 1 ffi 1, in which case it gives -n/2. The other two sums
in (19) give
,-1 (3 + cl)(1
Z
2r] - cJ') > 0
j~!

Hence A~>0 for 1=2 ..... n/2 and it remains only to consider ;h (Maxwell
did not recognize that the case 1= 1 is special). The formula for r] gives

~l(3+c,)(1-cj)
2r]

~[l=jffil

n "~'3+c,
--]ffij=l 4rj

n
2

Writing the numerator in the second sum as 4 - ( 1 - c t ) - - 4 - 89 gives


~-1 1

1 n-t

Using the formula Q=2 sin(nj/n), one finds that the second sum can be
computed explicitly. Its value is 88cot(n/2n). Hence ,11 > 0 if and only if
ln-tl

zc

An = n j ~ , ;>~nn c~ ~n + ~
and A~< 0 ff the inequality is reversed. The crucial quantity An is just an
average reciprocal distance between the points of a regular n-gon.

49

Linear Stability of Relative Equilibria

For small values of n one can evaluate both sides of the inequality
with a computer. In this way one sees that the inequality is true for n - - 7
but the reverse inequality holds for 2 ~<n ~< 6. It is now shown that the
inequality is true for n >/8. In fact the stronger inequality
1

A~>G+~

is proved. It seems to be the case that A. is monotonically increasing with


n. If this were true, then it would suffice to verify the inequality for n -- 8.
Unfortunately, the author has been unable to prove this monotonicity.
Instead, A. is replaced by a lower bound which is provably monotonic.
Note that

nA~-- Y~ f(j)
j~l

@
Q
0

"

o
0

"

9
9

0
9

"

O
O

Fig. 1. Some linearly stable relative equilibria.

86516/1-4

50

Meeekel

where f ( x ) = 89csc(xx/n). Consider the integral of f ( x ) over the interval


I'1, n - 1]. Since f ( x ) is convex, using the trapezoidal rule gives an upper
bound for the integral, that is,

fT

- l f ( x ) dx < 89

f ( 2 ) + ... + f ( n - 2 ) + 8 9

which leads to the lower bound,

An>A~ =lln

ll=~+cos(x/n) ~-~
~ csc
1 7t

One can verify that A~- is monotonically increasing and also that A~->
(I/2x) + 89 which completes the proof.

4.3. Other Minima


For n = 3 ..... 8, Hall numerically found relative equilibria which are
analogous to the equilateral triangle in that the n small masses are
clustered together. The n small masses are all equal. These numerical
results were verified by the author using a method which searches for minima of V(O). They are depicted in Fig. 1. Nondegeneracy was also checked
numerically. Because of the nondegeneracy, these configurations are limits
of sequences of relative equilibria of the (n + 1)-body with masses m o ffi 1
and m i = 8, 1 ~<i~< n. The theorem shows that the corresponding periodic
orbits are linearly stable for 8 sufficiently small.

ACKNOWLEDGMENTS
This research was supported by the NSF and the Sloan Foundation.

REFERENCES
1. Andoyer, M. H. (1906). Sur les solutiones p~riodiques voisines des position d'b:luilibre
relatif dans le probleme d~ n corps. Bull. Astron. 23, 129-146.
2. Brumberg,V. A. (1957). Permanentconfigurationsin the problem of four bodies and their
stability. Soviet Astron. 1(1), 57-79.
3. Hall, G. R. Central configurationsin the planar 1 + n body problem. Boston University,
preprint.
4. Lagsange,J. L. (1873). Essai sur le probl6me des trois corps. In Ouvres, Voi.6, GauthierVlllars, Paris.
5. Maxwell,J. C. (1890). Stabilityof the motion of Saturn's rings. In Niven, W. D. (ed.), The
Scientific PaperJ of James Clerk Maxwell, Cambridge University Press, Cambridge.

Linear Stability of Relative Equifibria

51

6. Maxwell, J. C. (1983). Stability of the motion of Saturn's rings. In Brush, S., Everitt,
C. W. F., and Garber, E. (eds.), Maxwell on Saturn's Rings, MIT Press, Cambridge, MA.
7. Meyer, K., and Hall, G. R. (1992). Introduction to Hamiltonian Dynamical Systems and the
N-Body Problem, Fol. 90, Applied Mathematical Sciences, Springer, New York.
8. Pedersen, P. (1952). Stabilitatsuntersuchung im restringierten Vierk6rperproblem. Dan.
Mat. Fys. Medd. 26, 16.
9. Routh, E. J. 0875). On Lapace's three particles with a supplement on the stability of their
motion. Proc. Lond. Math. Soc. 6, 86-97.
I0. Sim6, C. (1978). Relative equuilibria of the four body problem. Cel. Mech. 18:165-184.

Das könnte Ihnen auch gefallen