This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
0 .
2. If
1 and 9^ = arctan./._ 8 ^." / 5 _ 1 3.
0 , 87 + 1 > 0, then signU(80) = signA34 . ii) If P > 0, 87 + 1 < 0, ipiOy) > 0, then sign U(90)
sign\34 .
in) If 0 > 0 , 87 + 1 < 0, ifiiO^) = 0, then U{60) = 0 iv) If P > 0 , 87 4-1 < 0,
, in order to determine the other function. 2.2 oo \ liri Jtz\ = R * J\t\ 1, a_l ' - z ^o + T/2), then ip(t) is a t—symmetric periodic solution of period T. A periodic solution of the reduced isosceles problem is said to be a doublesymmetric periodic solution if it is simultaneously a r—symmetric and a t—symmetric periodic solution. This kind of periodic solutions are charac terized by the next result. Proposition 5 Let tp(t) = (r(t),f(t),z(t),z(t)) isosceles problem (3). n/y/2 (see for instance Belbruno et al.1). Moreover all these periodic orbits are double-symmetric periodic orbits. We note that the periodic orbits of the reduced circular Sitnikov problem correspond to periodic orbits of the infinitesimal mass, but in general they are not periodic orbits involving the three masses; that is they are not periodic orbits of the circular restricted isosceles problem. More precisely, each periodic orbit of the reduced circular Sitnikov problem gives a 2—dimensional invariant torus of the restricted isosceles problem filled with periodic orbits with period r = 27rp if T = 2np/q for some p, q G N coprime and p > q/(2y/2), and it gives a 2-dimensional invariant torus filled with quasiperiodic orbits if T = 2itw for some irrational w > l / ( 2 v ^ ) (see Theorem 2). 6.2 0 sufficiently small. (a) Case T = q/{2\/2), can be continued to: + A ) . (I, V>, s), l-periodic in (p and es, which is e2-close to the identity in the CT~2~m topology, such that transforms the Hamiltonian system of Hamil tonian ks(J,tp,es) into a Hamiltonian system of Hamiltonian Ke(I,ip,es). This new Hamiltonian is a C r _ 2 _ m function of the form: K^I^es) be one of the frequencies allowed in Theorem 7. Then, in the coordinate system (He, ,¥>) + aN) + 2\S N + l \ cos ((fc<" +1) , imply that for |/i| small enough there exists a generating function 0 (S,E,ip,J), defined in some neighborhood of S and global in the angles rl> € T", such that the map 6 : ($, E, \j), J) ■-» (5, E, V», J) is given by (y) = (0, z ) . This implies y € K. So, we have proved that is null, see Figure 5.1. The lower branch is asymptotic to the c axis on the 6 < 0 side. [E) = Sup{||T a 0| £ || | a G V}, f+(M) can also be dealt with without any information about the central configurations. The case v > v+{M) is a , the same argument applies to any set {cp — = g mod 7r, £ = 0 mod n, R = 0. = 9 + v, and expand with respect to 6 - -2(506$
Shapere-Wilczek's Connection and Curvature Forms
Appealing to the Stokes paradox, a form of which states that the only solution to the Stokes equation corresponding to a rigid translation of a cylinder is a rigid translation of the fluid as a whole, Shapere and Wilczek asserted that the rotation and translation of the organism associated with a vector field V(a) on the boundary of the organism that generates a velocity field v(z, z) can be determined from the asymptotics of that velocity field. Specifically Arot ■^SW
dz
— Im d> Jlim\z\ !^oo
27r
.
z),
(7)
rz)
(8)
*
and Atr
dz , 2nizViz<
-f
Jlim\z\—»oo One of the purposes of our paper is to provide some corrections and clar ifications to SW. It turns out that while formula for Atr is correct, Arot is not (unless C is a circle). We provide the corrections in section 4. By the general theory of connections, the curvatures are the infinitesimal rigid motions 1Z associated with traversing infinitesimal rectangles centered at C in shape space, spanned by the vectors nV^ and eV£. More precisely,
n{T)v*,tv*)c =e^C
+b
(9)
where w and b are written as U) -=
Arot([Vv*,ev?n}) cnFrot +
« n
+ « n+Wfl.
(10)
and
6 == Atr{[nvl «&]) = «?^mn + « » + « l + « * • (11) The reason for choosing this particular decomposition of the Stokes curvature form, as used by Shapere and Wilczek, will become evident in the examples. For the calculations, one makes the change of variable z = w(f), in which Atr becomes Atr
-y
»M0, "(0)-
(12)
37
The expression for the Lie bracket pulled back to the £ plane is [V„,Vm] = (Vm ■ V)vn -- (Vn ■ V)vm
-{-dzTz^Vm ,dvm 9 ( dz^
l
3
+
{
lzTz^Vm~ _ ( a ^ dz]^Vn
y„.
(13)
T h e Elliptical Swimmer (following t h e S W recipe)
We now compute the SW curvatures for the elliptical cylinder and apply the results to study swimming strategies for a nearly elliptical swimmer (we call this kind of swimmer a self deforming ellipse). It turns out that the rotational component according to SW's recipe is not correct. We present it here for the sake of comparison with the corrected result in section 4). The ellipse is given by the simplest conformal mapping, mi = M, all the others m.i = 0. In fact, the transformation 2=
W
(0=fl«+y)
(14)
maps the region exterior to the unit circle 7 in the f-plane to the region exterior to the ellipse C with semi-axis R(l + M) and R{\ — M). The hodographed basis vn(z(o-)) =
\on+1.
has a complicated explicit expression in the z-plane, namely, for s = R(exp(i6) +
Mexp(-i6)),
we have vn(s)=
expWn + 1)*)
+
=
x(Lt^EEj
\
Fortunately, the calculations in the £ plane are feasible. The boundary condition for the ellipse corresponding to formula (4) takes the form 0(CT)
1 a2 + M -r 0>)+
~ ar ^
^ = ACT"+1'
{15)
38
Because the details of determining the solution to this equation parallel those found in Muskhelishvili in his computation of the stresses on a plate with an elliptical hole, will only state the results:
0 n> 1 0 n= -1 A£ n+1 n < - 1 ,
(16)
and n > -1 n = -l
X
V>(0 = {
(17)
e^$(n + lWn<-l-
Substituting these into the formula for the v n (£) we obtain
Av„tf,0 = * ( f l -
z (e - M) n
(18)
4>' (0 + ^(0
n > -1 n = -1 = < A, Ae+1-(n + l ) ^ i | ^ A r + ^ ^ ^ ( n + l)Ar,n<-l
xt ~\
Expressions (7) and (8) for the connection may now be used to decompose the boundary vector fields into horizontal and vertical parts. Boundary Value n+1
Horizontal Projection Xan+1 0 0 ACT-1 - ImMXa-*
Xvn(a) = A<x ,n>0 Xvo(a) = ACT AU_I (cr) = A Xv-2(cr) = ACT-1 Xvn(a) = Xan+\n< - 2 ACT n + 1
Vertical Projection 0 Arot = ImX Atr = X Arot = ImMX 0
(19)
Caveat: the results for the rotational part are not correct, see Proposion 3 in section 4. To compute the Stokes curvature we compute the Lie bracket of horizontal vector fields, re-expand the resulting vector field in terms of the Fourier basis, then use the above table to determine the corresponding rigid rotation and translation. As an example if m,n > — 1, [VVn,evm} = (er)(m + 1)CT m—n-f 1
oo
erj(n + 1)CT'n - m + l \) ]V"* >JMCTY, fc=0
(20)
39 comparing this with the above table to determine the vertical components we find that for m,n > —1
{
(m + VjM^r0
n-m=j>0 and j is odd otherwise. (21)
(n + 1)M V f%n = ptr
u
_ ptr
- n = k>0 and k is odd otherwise.
m
.= 0
(22)
(23)
For the circle M = 0 and the result of Shapere and Wilczek is recovered, the only modes that couple are those that differ by one. If M ^ 0 all modes that differ by an odd number couple, although the strongest coupling takes place between ones that differ by just one. The effects of coupling between distant modes becomes more pronounced the further the shape is from a circle. 3.1
Examples and discussion
Two examples will be presented, using only the translational part of the con nection (so that the previous calculations are correct). The first, an analog of the example of Blake's swimming circular cylinders 10 , under symmetrical deformations. The second example will be that of a long thin organism that swims utilizing the undulatory mode. The results of this model will be com pared with observations of the swimming motions of nematodes by Gray and Lissman. For the first example consider an ellipse with semi-axis 1 — M and 1 + M, and swimming stroke parameterized by: M S{a,t) = (a+ —) + (.025 cos 2irt)vu + (.025sin27rt)ui 5 (7
+(.015 cos 2wt)vls + (.015 sin 2irt)vig
(24)
To calculate the net translation of the swimmer the curvature components corresponding to the coupling of modes 14 and 15,14 and 19,15 and 18, and 18 and 19 are needed. After one swimming stroke the swimmer's net translation
40
4
5
6
Figure 1. Ellipse (M = 3) at t=0, 0.25, 0.5, and 0.75
is given by: / •7ri4j5ai4
+ / Jo
Fi5Tga,i5ai8dt
(25)
^14 15014019^
Jo = .043 + .017M 2 - .019M
(26)
See Figure 1. Next, the undulatory mode of swimming will be studied by taking the ec centricity of the ellipse to be large and the deformations to have purely imag inary coefficients. Because this strategy is used by swimmers of all Reynolds numbers, this two dimensional model may provide a starting point for build ing models where the inertial forces are not negligible. J. Gray and H.W. Lissmann7 conducted an extensive study of the various locomotion modes of Nematodes. They made observations and measurements of nematodes creep ing on top of gelatin and damp glass, gliding through densely packed suspen sions, and swimming through fluids. The present model will be applied to the latter class of nematodes, the Turbatrix Acetii. While swimming in an open fluid is clearly a three dimensional problem, the quantitative similarly
41
between the two and three dimensional cases suggests that we might obtain a reasonable model for the Tubatrix whose body motions were observed to be essentially planar by Gray and Lissman. The Turbatrix Aceti is a nematode which appears occasionally in domestic vinegar. It has a body length of about 840/im and a diameter of 28/im. Its ratio length/width ranges 13 about 45. It swims by passing undulatory waves down its body. The observed wavelength was 712/im, the frequency is 5.2 s e c - 1 or about one complete stroke every 1.9 seconds. The amplitude was measured to be approximately 107^m and observed speeds were 718/xm per second, or 138/xm per stroke. Thus the foreword progress was was about 16.5% of the worms length per stroke. To model the Tubatrix pick as the base shape an ellipse with semi-axis .06 and 1.94. Frequently the amplitude of the undulatory wave was observed to increase as it passed down the body of the nematode, the tail moving nearly four times as much as the head. This characteristic is related to the ability of the nematode to swim without yawing. The yawing motion associated to a particular swimming stroke is not detected by the present model because the side to side motion is zero on average. As a simple model, consider undulatory swimming motions with a constant amplitude. The specific modes are chosen to best represent the observations of Gray and Lissman (1963), see Figure 2. The amplitude and ellipse size were chosen based on measurements of Gray and Lissman. Parametrize the swimmer by:
S(a,t) = a+^1
+ .17icos2irt(v3(cr) + V-5(
dvnd£
- HtLo(Ma2)k(n 0,
+ l)tf}Mam-n-\
n<
-1
n > -1
42
t-
i
r-
i
k
• •
k—
Figure 2. Locomotion of Turbatrix aceti at t = 0,0.2,0.4,0.6,0.8 and 1.0
' [Er= 0 (^ 2 ) fc ] 2 dvnd[. d£ dzVm
I ~)
((n + l)lf)Ma-n~m-1
- (n + l ) e ^ o - n - m + 1 ) , n < - 1
£7j(m + l ) ( j m - , l + 1 ,
n>-l.
The resulting curvature coefficients are: F3<% = 4,F_ 6 i _§ = 4,F_5i4- = 4M, F_5_4 = 4, all other non-zero components being obtained by antisymme try. The net translation is then found by computing:
L
F3ia3(t)d.4(t)dt
+ I F_6_-5a-6(t)d.-5(t)dt+ Jo + f Jo
f Jo
F-5<4a-5(t)a4(t)dt
F_sAa-5{t)d4(t)dt. (28)
The result of this integral is -0.71. After scaling, the predicted translation per stroke is 154pm, or 18% of its length. Recall that the observed translations were approximately 16.5% of the length. The result seems quite good, in spite of the simplifications, namely: the model is two-dimensional; only four
43
modes were excited, with constant amplitude; the correct Reynolds number is of order 1, in the limit of applicability of Stokes approximation; effects of order > 3 in the amplitude were not taken into account. 4 4-1
Mathematical details and corrections to SW The canonical vs. the hodographed Fourier basis
Suppose that in Goursat's representation for the solutions of planar Stokes equations:
v(z) = W) + (
(29)
we take ip and 4> holomorphic in the exterior of the closed curve C. Without loss of generality, a_i a_ 2 4> — a0-\ 1 7T + ... • z z* * = *=i + ^ + ... (30) z z2The presence of a 0 is a "confession of debt" to Stokes' paradox, since there is no solution vanishing at infinity for the translation of a cylinder. Write
/KO
\e=i
)
with (1.2)
V
*
= MZ)
jvrkeal=^
T T T
- Z(P^Z) = 1 4
m
- Hz" ~ k*zk-X)
Thus for each k < 0 we have four terms, taking for xp and <$> either zk or izk. This is called the canonical basis. Observe that Vk have good symmetry properties with respect to conjugation z —> z and reflection z —> —z. However, Wk do not have these symmetries. This is already one good reason to suspect that the canonical basis is not the best choice. In Appendices A and B we obtain the curvatures for the circle, using the canonical basis, in Appendix C we show that in the case of the ellipse, the canonical basis does not allow a practical computation.
44
Let z = z(£) the conformed map taking the exterior of 7 = S 1 to the exterior of C. We transport "hodographly" the Fourier basis {cr n + 1 } in 7 to C, that is: Vn(s) = ( £ ( s ) ) n + \
n€Z.
We adopt the indexing n —» n + 1 just to maintain the convention used by SW. In contradistinction with the canonical basis, the symmetry properties are preserved by the hodographed Fourier basis. Suppose, for instance that C is symmetric with respect to the x-axis. Taking real coefficients, we pro duce symmetric deformations for all n. With imaginary coefficients, we get ondulating deformations. The Fourier basis has other advantages. For any conformal map, half of the problem to find the Stokes extensions is trivial: for n > — 1 we take always ip = fn+1 and 4> = 0. As we will see in section 5, there are also remarkably simple properties for the power expenditure operator. 4-2
Total force and total torque
A simplified version of Stokes' paradox is given by the following Lemma 1 If the potentials ij) and <j> have no singularities at 00, as in (30) then the total force vanishes:
I fds = 0 Proof. This follows from the identity fds — —2nidU where
U = 4>(z) + zp-xp . In Appendix E we present a derivation of the force field / = 4fiRe(
45
For the rotational component of the connection, we now present a correction1 regarding the rotational part. Criterion 2 In dimension two the rotational part (as in dimension 3) is de termined by the condition "total torque T = 0 ". In fact, we have Lemma 2 Using the canonical basis, T = 47r/j/m[6_i] Proof. A simple calculation gives tds = Im[z/]ds = -2fj.Re[zdU] = -fidF where F = zz(& +4>')-z7p-zlp
+^ + l
e Z=
I il>{z)dz JZo
Observe that £ is multivalued if b-\ ^ 0. Remark 1 The reader may feel unconfortable with the presence of the con stant rigid translations of the fluid as a whole, which do not decay at infinity: 1 and i as allowable vectorfields. Can one do differently? the answer is yes. Translations of the shape can also be induced by stokeslets. However, these have even worse behaviour, logarithmic at infinity14. Thus, in two-dimensions, one can still define translation-horizontality us ing the criterion "total force = 0", but this requires replacing the constant vectors 1 and i by the two stokeslets, on the x and y directions. In a future work we plan to check if the criterion using stokeslets yields the same results for the connection. 4-3
The 1-form of the connection
Given a velocity field V along C, we denote (a, b) the translation and u>k the infinitesimal rotation such that V — uiz — (a, b) is horizontal. First, we introduce terminology. Definition 1 If(a,b) = (0,0) we say V is translationally-horizontal. 0 we say V is rotationally-horizontal.
Ifu> =
46
Along C, we expand V in terms of canonical components oo
4
«w=^+EE«* • k
SW argue that vh(s) = v(s) — a(iz~l)\c
dZ 2 n i z
v(z,z)
(32)
with
a = Im
7r~.v(z,z)
(33)
2 n l
is rotationally-horizontal. In other words, the rotational component of the connection, would be
Ars
^-Mz,z),
(34)
ll{%
This is not correct. Note that an infinitesimal rotation of C is (except for the circle), a combination of i/z with other elements of the canonical basis. Hence, for the rotational part, it is not sufficient to consider the single element i/~z in the basis, corresponding to xj>-\ with imaginary coefficient. In other words, the flaw stems from the fact that, unless C is a circle, i/z\c does not represent an infinitesimal rotation Re = iz\c ■ 4-4
Infinitesimal rotations: corrected formula
Consider a conformal transformation
z = u(t) = R{£+Y + --- + ^ ) from 7 to a curve C (we can also take n = oo). The following lemma is obvious.
47 Lemma 3 The infinitesimal rotation vAth angular velocity w = 1, is given by i z ( 0 = iR(v0 + miV-2 +
h mn u_( n + i))
(35)
where, in the right-hand side, v^k = exp(i(—k + 1)6) are elements of the hodographed Fourier basis. For each k = 0, k = —2, k = - ( n + 1 ) (k = - 1 is absent), we must compute the Stokes extensions of the hodographed Vk(s). Actually, for each k we need to find "only" the coefficient 6* j of ipk in the expansion of v\ in terms of the canonical basis. In the proposition below we omit the subscript - 1 in the 6fc's. Proposition 1 For indices k = 0,k = -2,---,k
= - ( n + 1),
the rotational part of the connection is given by: Arot,
A
x _ {Vk)
1
Mb*)
,, fi x W
~ R Re{b" + m, 6-2 ■ • • +mn &-(»+D) '
In fact, the vectorfields along C h Vk = t,
Imbk
*-flite(6P + m l 6 - » - - H n B 6 - ( ^ ) )
n, I
*(l''+mi
^ w
- » + ~ + m « "-(»">) (37)
are rotationally-horizontal. Conjecture 1 Following Ken Meyer's talk in this Conference, our bet is 95%: All the remaining elements of the hodographed Fourier basis are rotationallyhorizontal (that is, for k > 0 and for k < — (n 4-1) )■ This involves finding the asymptotics of the Stokes extensions for the corresponding elements of the hodographed Fourier basis. We now give a more intrinsic version for (36). In order to obtain Arot(v) = a we require that (v — aiz)\c produces zero total torque. De note T the operator yielding the total torque. From T(v — ctiz) — 0 we get a = T(v)/T(iz) . Theorem 2 rot _ -^Ep lim Im — & Airot(, (v)A = v{z)dz u(x :
where Tc = T(iz) is the torque associated to the infinitesimal rotation iz(s) ofC.
48 4-5
The ellipse. Rotational components.
We saw, in section 3 the expressions for the Stokes extensions
<M0 =--4 +
+ •••,
bki
MO --• ^
+
-
We are interested in finding which A; have nonvanishing coefficients 6iLj. We claim that only UQ A and U_2,A do have them. In fact, since ( A _
Afi
,n = 0
• ,n= -2 £ € 2 -M (omitting the superscripts k) it follows that the only nonvanishing 6*i occur for k = 0 and A; = —2. Hence L e m m a 4 The following expression for the total torque holds I
Ann T(v-2,x)
= Im{R\) =
-RIm(\),
■MRIm{\),
in particular the rotational components of the connection are
A^(Xv0) = - ^ M A ) ,
A-(A,_ 2 ) = - ^ M A ) .
lc J-c To compute Tc, we consider the infinitesimal rotation iz\C = iz(C). The Stokes extension is given by: iz(C) = iR{a + Mo—1) = iRa + iRMa~l
= v0,iR(o-) +
V-2}IRM((T).
Thus the infinitesimal rotation of the ellipse requires two elements from the hodographed Fourier basis, precisely those having coefficient s 6* x ^ 0. Now, iR
m =
iRM2i2
+ M~l
-T—^-^TAT iR2 iR?M2 +■ RZ R£ iR2(l + M2) + ■•
49 Lemma 5 The total torque associated to the infinitesimal rotation T(iz) of the ellipse is T c = -47r/ifl(l + M 2 ). substituting the above expression in the expression for the rotational compo nents of the connection yields Proposition 2 The rotational components of the connection at the basis el ements VQ,\ and i>-2,A are
Arot(vo,x) = Arot(v.
-2,A) =
Im(X) = U)Q, R{1 + M 2 ) MJm(A) = u.-2R{l + M2)
in particular the following decomposition into rotationally-horizontal purely rotational components holds
. Im(X) \
/m(A)
,
MJm(A)
«* J .XW = ( A --l
l + M2
) a --%1 +
and
j
M^
Summarizing: T h e o r e m 3 Given
V(
Im(X0) M/m(A_2) Arot{v) = " R(l + M2) ' R{l + M2)' The rotational components of the curvatures can be computed in a similar fashion as in section 3, finding the 6_i residues in the Lie brackets. 5 5.1
Power expenditure The differential form fds = gdO
Forces can be represented by the same space Vc of velocity vectorfields along C, with the L2 norm. Intrinsically, however, it is an abuse to consider 2fif = 2/x (2Re(
(38)
50
as an element of Vc- Actually, the stress force is a differential form fds, where s is arclength parameter along the curve C. It belongs to the dual space V*. In what follows we use the notation / = f(v(s)) for the vector of forces associated to the vector of velocities v(s) along C. We now explore (to our benefit) the possibility of not identifying these spaces. We use the pull-backs of the forces as forms to 7, under the conformal mapping z = z(£) sending the exterior of the 7 to the exterior of C. Proposition 3 Let s be the arc length parameter along the curve C, and 6 the polar angle in the unit circle corresponding under the conformal mapping z = z(£). Let f = f(v{s)) be the vector of forces associated to the vector of velocities with complex potentials
(39)
where
i = d4o-' d
in=m j y
_
jma
\dtj dz/d£ z{£)a-1 dP4>ldt? dz/d£
z{0a-^_m£zMe 2 (dz/dO
(40)
The proof is a long but straightforward calculation, using the chain rule and taking into account that £ = a = exp{i9) along 7. Since we are using hodographed velocity fields v(a) from 7 to C, the power expenditure is given by the usual inner product V = 2fi I {v,g)i where now we identify gd0 with the vector field g 6 V 7 , the space of vector fields along 7.
51
Lorentz reciprocity 14 carries over to this representation, in other words, the operator ^:V7->V7,
v^g.
(41)
is self-adjoint and positive. If we could determine the spectrum of V1, then the task of computing power expenditures would be completely solved. We assert that one of the advantages of the hodographed basis is that half of the spectral problem is a free lunch! In fact, for n > - 1 we know that ip = £-( n+1 > , 0 = 0 . hence V = a - ( n + 1 > = <7 n + 1
and drj)
9=~(J-1
= - ( n + l ^ - ^ + ^ a - 1 = - ( n + 1) an+2 a'1 = - ( n + l)c7 n + 1
Theorem 4 For every conformal mapping z = z(£)> ifn+1 > 0, then <jn+1 is an eigenvalue ofV"1, with eigenvector — (n + 1). It is interesting to look what happens geometrically in the physical bi plane. The force field / G Vc> is parallel (with opposing sense) to the velocity field v(z(a)) = c n + 1 . The scale factor depends on the conformal mapping, ds/dd To compute the power expenditure V, the scale factor disappears when we change the integration to 7. The difficult part of the spectral problem involves only the indices n + 1 < 0. The following proposition shows that they form an invariant subspace. Teorem 5 The operator Vy(v) = g leaves invariant the subspace of nega tive Fourier modes an+1. Proof. Use Lorentz reciprocity. Let m > 0; then (V(a " + 1 ) , am) = (V(am),
an+1) = f -maman+ld9
= 0
52
5.2
Computing g for the ellipse; negative Fourier modes
Recall that for p < —1
4"-H'*',. *
= A (
P +
1 ) ^ - ^
We insert in (39-40). In the denominators appear factors 1 — Ma2, coming from the derivatives of z(£) and £ 2 — M, followed by conjugation (remember a = o-1). Well, for 0 < M < 1
n=0
We know, however, that in the final result for g, only negative powers of C do appear. The term /J
=
^
( 7
= A(p + l K + 1
is in accordance, but the terms. I, 777, IV, V are problematic. Nonetheless, a "miracle" occurs: Lemma 6 For the conformal mapping z = R(£ + M/£) of the exterior of unit circle onto the exterior of the ellipse C, let f{Xap+1) denote the vector of forces associated to the hodographed vector field Xvp+\(o-) = Xap+1, let gd6 denote the pullback of the differential form f ds to the unit circle, so f ds = gdO (see Proposition 3), then I + III + IV + V = 0, in particular for p + 1 < 0, f{Xap+i)ds
= X{p+ l ) ^
1
d6.
Proof. A direct calculation. Theorem 6 For the ellipse, the hodographed vector fields v(a) = ap+1 for p + 1 < 0 are also eigenvectors ofVy : v —> g, with eigenvalue p + 1. We think that this is indeed remarkable: when we vary the eccentricity 0 < M < 1, using the hodographed Fourier basis, the power expenditure functional does not change! 5.3
Hypotrochoids
The degree of algebraic difficulty increases, but we can still calcultate, g(Xap+1)d6 for the hypotrochoids 11
—(«♦£).
(42)
53 with m = 1,2,..., and M a positive constant satisfying 0 < mM < 1. We present the results: Theorem 7 Consider the hodographed Fourier basis vn(a) =
g(Xo-') = (-/? + Mfta-1, /? =
^ ^
so the eigenvalue is different from the others. 4- For m — 4 we found g(Xa-1) = -Xa-1+
ACT"2
Here the hodographed Fourier basis is not an orthogonal basis for the power expenditure operator. 6
Efficiency of an elliptic microswimmer
There are basically two competing efficiency notions for microswimming. Con sider a swimming stroke of period r, in shape space S and the ratio X E
X/T E/T
mean velocity mean power
.
Froude 's efficiency is an non-dimensional quantity which results from mul tiplying this ratio by a "characteristic force" T, and in the Stokesian realm it is equivalent (up to a shape dependent factor) to Lighthill's efficiency (mean velocity) .„,. — . (44) mean power The other notion, which we call SW efficiency (because Shapere and Wilczek essentially use it) is given by efz, =
etsw = —£ ,
(45)
54
where the presence r in the denominator is equivalent to consider all swimming strokes with period T = 1. In this paper we will consider the latter notion, and consider a variational problem first introduced by Shapere and Wilczek leading to the most efficient strokes. For a comparison of the two notions, see 15 . 6.1
The variational problem
We expand the time-dependent, small shape deformations in terms of a basis, so that the problem is linearized to a Lagrangian £ = ^{K4A) ~ v-(Tq,q)
.
(46)
where the matrix K encodes the power expenditure, F the curvatures, v is a Lagrange multiplier. The Euler-Lagrange equations are Kq = vTq .
(47)
In other words, C describes the problem of minimizing the energy expenditure E =\ j\Kq,q)dt
,
X = \ f
.
for a prescribed holonomy (Fq,q)dt
Inserting q = eVeint
,
(48)
, A= — v
(49)
, S = AT * .FAT * skewsymmetric,
(50)
where V is an eigenvector, gives TV = \KV Define W = fCW so that: BW = \W Taking into account that
T
=
2TT/Q,
ef-—
.
(51)
the efficiency can be written as ,
(52)
55
where we omit the factor 27r. Some simple manipulations with E and X give E = ^Q2€2(fCV, V)T X =-l-ne2(FV,V)T
.
Since TV = XK.V, we get e f = ^ - = -iA,
(53)
Recall that A = iQ./v is an eigenvalue of (51), and as expected, it is purely imaginary. In short: Finding SW's efficiency is equivalent to find the supremum of the purely imaginary spectrum. 6.2
Numerical results
For the elliptical swimmer, undergoing symmetric swimming strokes, matrices K and F are given in Appendix F. We found the spectrum using MathLab, for M = 0 (circle), M — 0.1 e M = 0.2 (ellipses with small eccentricity) and M = 0.8 e M = 0.9 (high eccentricity). Regarding the number of Fourier modes, we solved the problem exciting the first five, the first ten, the first 30, and finally the ten modes between 30 and 39. Tables of results are available upon request. Our observations are as follows. As expected, the absolute values increase as more modes are considered. Interestingly, they do not diminish significantly as we exclude the lower Fourier modes. This indicates, as shown by SW for the circle and the sphere, that the more efficient strokes involve essentially the high geometric modes. In the table below we list the highest eigenvalues (51), for small and for high eccentricities. First 5 modes First 10 modes First 30 modes 10 modes, from 30 to 39
M = 0 1.9906i 2.4008i 2.7016i
M = 0.1 1.9664i 2.2845i 2.4884i
M = 0.2 1.9431i 2.1826i 2.3139i
A* = 0.8 2.4541i 3.31621 4.91681
M = 0.9 2.7878i 4.2483i 7.2448i
2.6529i
2.4717i
2.3134i
4.0299i
5.6556i
We have also considered very high order modes, between 30000 e 30009, and we verified that there seems to exist a higher bound for all efficiencies, as conjectured by SW 2 .
56 The advantage of high order geometric modes supports Nature's choice for ciliary motion rather than deformations of a membrane. The mechanical stress would be too strong for a membrane undergoing high order deformations. On the other hand, it is quite easy to a ciliary envelope to emulate these motions. Acknowledgments Thanks to LNCC/CNPq for supporting a scientific visit of one of the authors (J.D.) to Rio de Janeiro, February 1998. Appendix A. The Lie bracket Let z = x + iy,z = x — iy. Denote d_ _ 1 fd_ _ .d_\ % dz ~ 2 \dx dy)
d__l(d_ ' dz~ 2\dx+
.d_\ dy) '
l
A vector field v = (a(x, y), b(x, y)) = a— + b— is represented as v = f
dz
+ f
a=z
where / = a(x, y) + ib(x, y) = f(z, z). Lemma 7 The Lie bracket w = [u, v\ can be computed as follows:
w = [u, v] = 9z £;
9z 9z 9z.
or as
7" 7.
fz fz Jz fz.
-
9 9 .
(54)
L d Td w = h— + h—, az oz
with h = 9zf + 9z-f ~ fz9 - h9Consider, for n, m < 0, the vectorfields vn = zn , um = zm-
razz™-1
(55)
57 written as differential operators as fi
ft
Vn=7l—+Zn az
— az
u m = tzm - mzz™-1)—- + (zm - rrizz™-1) — . az az We obtain [vn, Vp] = pznzp~1
- nzPz"-1.
(56)
Curvatures for the circle For the circle Sa : r = a, the translational component of the curvature is the constant term in the Fourier expansion of these brackets alongo Sa- For n < p < 0 we get Ftr(vn,vp)
= an+pFtr(a^,aM) 1
=
=
2
f*
^ L f
2 n
(pe(n-^Dtf _
ne
dg
Just the first integral can contribute, and only if p = n + 1, so that F"VnUM) = — a
.
(57)
B. Curvatures for the ellipse using the canonical basis The ellipse x2 v2 2 + a b2 is described in polar coordinates by r = r(0), with /cos :2 8
sin20\"1/2
- +-£r)
•
( 58 )
It is possible, but not practical, to compute directly the curvature coef ficients using the canonical basis. Let's see why. To compute Ftran3(vn,vp), we substitute (58) in (56). The two terms have the same absolute value, /cos 2 fl V a2
+
an^\"^b2 )
1 ) / J
58
but different phases exp(i(n - p + 1)0) and exp(i(p - n + 1)0). As a working hypothesis, suppose we will have only p + n odd (recall p + n is < 0 ). The situation is not completely desperate, since the absolute values as well as the phases are trigonometric polynomials. We must now expand the bracket in the basis {vn>wn}. Assuming that we need just the even negative n's, we mould get finite trigonometric expressions on both sides. However, finding the coefficients of the expansion becomes a very complicated combinatorial-matricial problem. C. Ellipse using the Fourier basis. More details. Lie brackets In the text we only gave the case m,n < — 1. For m < —1 and n > — 1 we have: [A„t/„, Amum] = ^T, " ^ T ^ " 1 ~
n Mk(T n-m+1+2*; bliYJ
)
\J
k
_ J2 ^Hm
+
1)Mkan+m+1-2k
k
+ J2 ~^.(m + l)(jfc + l)Jlf
fc
<7-n-m+1+2fc -
k _ J2
*!^»l( m
+
l)(fc
+
1)M*+lCT-n-m-l+2fc
H
k
For m,n < — 1 one gets - m)Mkan+™+l-2k
[Kvn, Xmvm] = £ ^r(n k
+
K
h
J2 ^-(n
- m)(k +
i)Mk+1a~n-m-1+2k
k
+ J2 ^-(m
- n)(k + l)Mka~n-m+1+2k
K
k
+2k k
+
59
A„A
£ ^ ( n + l)MVn+m+l+2fc D. Formula for the Force f
Given a closed curve C, we denote the exterior normal by dz as where s is the arclength, traversed counterclockwise. Proposition 4 n = —i —
1. The pressure is given by p = -4/xfle [<*>']
(59)
f = 4nRe(4>' n) - 2/x( z W - W) n .
(60)
2. The stress force at z € C is
3. The differential form fds can be written as fds = -2t/i {(
{zW -W)dz)
(61)
Proof. We start with %l> = / -rig. The velocity components are u — f and v = —g. Since / and g are harmonic, Au = Av = 0, Stokes equations are satisfied with pressure p = 0. We compute the stress tensor: en =fx =
^ifx+9y)
ei2 = o ( / y - 9x) = fy = ~9x
e22 = ~9y = - g C / x + S y )
We write the normal vector as n = p + iq. Then
fenei2\fp\
\e21en)
\q)
=
Hxfy
\fy-fx)
'fx-fy\
\.(P\ \q)
[P
Jv fx J \-q, = (en + ien)(p - iq) = V n
60
Now we compute with
- yGxx
Uy = Fy — lFXy — GX — J/GXy vx = 2GX + xGxx - yFxx =
Vy
X(jTXy
y+Xy
using one the the Cauchy-Riemann equations, namely Fx = Gy. Using the other one, Fv = —Gx we obtain: Au = - 4 F « , Av = ~4Fxy . Hence, the pressure is P=-^Fx
= -2fi(<j>' + W),
and the components of the stress tensor are: e n = -xFxx
- yGxx
e22 = xGxy - yFxy
(en =
ei2 = 2 ("x + uy) = -xFxy
-eu)
- yFxx
In complex variables notation e n + ie\2 = —
61 E. Matrices K and F for the elliptic swimmer We consider symmetric deformations, so all A„ = 1.
Curvatures n\m
0 0 0 1
I R
-2 0 2 0 -3 0 3
M R
-4 0 4 0
1 1 R
0 M+l R 2 R
0 0 0 2M R
2
-2
2 R
0
0
0 oM±i
1 R 2M*+M-1
1 R o M+l z R
0
0
0
2M R
0
0
2M'+M-1
n
3 R
0
0
3 R
0
0
0
0
0
0
2 R 2 o2M +M-l
R
4
0
0
0
-4
M R
0
0
0 M+l R
3
-3
R
0 1 0 0 0 0 0 0 0
1 0 2 0 0 0 0 0 0
-2 0 0 1 0 0 0 0 0
2 0 0 0 3 0 0 0 0
-3 0 0 0 0 2 0 0 0
3 0 0 0 0 0 4 0 0
R oM+1
R 4 R
Power expenditures ■n\m 0 1 -2 2 -3 3 -4 4
0
2 •* R
QM + 1
"^
0
-4 0 0 0 0 0 0 3 0
4 0 0 0 0 0 0 0 5
4 R
0
0
0
0
62
References 1. A. Shapere and F. Wilczek, Geometry of self-propulsion at low Reynolds number, J.Fluid Mech. 198, 557-585, (1989). 2. A. Shapere and F. Wilczek, Efficiencies of self-propulsion at low Reynolds number, J.Fluid Mech. 198, 587-599, (1989). 3. J. Koiller, R. Montgomery and K. Ehlers, Problems and progress in Microswimming, J. Nonlinear Science 6, 507-541, (1996). 4. K.M. Ehlers, The Geometry of Swimming and Pumping at Low Reynolds number, PhD Thesis, Univ. of Calif., Santa Cruz (1995). 5. A. Cherman, M.Sc. thesis, Centro Brasileiro de Pesquisas Fisicas, (1998). 6. K. Ehlers, in preparation. 7. J. Gray and H.W. Lissmann, The Locomotion of Nematodes, J. Explt. Biol., 4 1 , 135-154, (1964). 8. J. Lighthill, On the squirming motion of nearly spherical deformable bod ies through liquids at very small Reynolds number, Commun. Pure Appl. Math. 5, 109-118, (1952). 9. J. R. Blake, A spherical envelope approach to ciliary propulsion, J.Fluid Mech. 46, 199-208, (1971). 10. J.R. Blake, Self propulsion due to oscillations on the surface of a cylinder at low Reynolds number, Bull.Austral.Math.Soc.3, 255-264, (1971). 11. N.I. Muskhelishvili, Some basic problems of the mathematical theory of elasticity, P.Noordhoff, Groningen-Holland, (1953). 12. G.K. Batchelor, An introduction to Fluid Mechanics, Cambridge Univer sity Press, (1970). 13. D.B. Dusenbery, Life at Small Scale, Scientific American library, (1996). 14. C. Pozrikids, Boundary Integral and Singularity Methods for Linearized Viscous flow, Cambridge Texts in Applied Mathematics, 1992. 15. J. Koiller and J. Delgado, On efficiency calculations for nonholonomic locomotion systems, Reports Math. Phys., 42, 165-183, (1998).
2 - D I M E N S I O N A L INVARIANT TORI FOR T H E SPATIAL ISOSCELES 3 - B O D Y PROBLEM MONTSERRAT CORBERA Departament de Fisica i Matematica Aplicades, Universitat de Vic, Sagrada Famtlia 7, 08500-Vic, Barcelona, Spain E-mail: [email protected] JAUME LLIBRE Departament de Matematiques, Universitat Autdnoma de Barcelona, 08193 - Bellaterra, Barcelona, Spain E-mail: [email protected] We consider the circular Sitnikov problem as a special case of the restricted spa tial isosceles 3—body problem. In appropriate coordinates we show the existence of 2-dimensional invariant tori that are formed by union of either periodic or quasiperiodic orbits of the circular Sitnikov problem, these tori are not KAM tori. We prove that such invariant tori persist when we consider the spatial isosceles 3—body problem for sufficiently small values of one of the masses. The main tool for proving these results is the analytic continuation method of periodic orbits.
1
Introduction
The main objective of this work is to prove the existence of 2—dimensional invariant tori filled of periodic or quasiperiodic orbits for the spatial isosceles 3—body problem. We note that in particular these tori are 2—dimensional invariant tori for the general spatial 3—body problem. We start reducing, with the help of appropriate coordinates, the dimension of the phase space of the isosceles problem, obtaining in this way the reduced isosceles problem. We see (in Theorem 2) that our tori filled of periodic or quasiperiodic orbits come from periodic orbits of the reduced isosceles problem. Using the analytic continuation method, we prove (in Theorem 6) the existence of symmetric periodic orbits of the reduced isosceles problem, for sufficiently small values of one of the masses, near the known periodic orbits of the reduced circular Sitnikov problem (a particular reduced restricted isosceles problem). Finally we analyze the 2—dimensional invariant tori of the isosceles problem that come from those periodic orbits. In this paper we present results without proofs. The proofs can be found in Corbera and Llibre2, and they constitute the main results of the Ph. D. of the first author.
63
64
2
Equations of motion of the isosceles problem
We consider three particles P\, Pi and P 3 of masses mi, 777.2 and 7713 respectively, such that mi = m.2- Their initial positions and velocities in the inertial coordinate system (X,Y,Z,X,Y,Z) are (X, Y, Z 2 , X^Y^Zi), {-X,-Y,Z2,-X,-Y,Z2) and (0,0,Zi,0,0,Zi), respectively. Of course the dot denotes the derivative with respect to the time t. The spatial isosceles 3-body problem, or simply in this work the isosceles problem, consists of de scribing the motion of these three particles under their mutual Newtonian gravitational attraction. We note that the name of isosceles comes from the fact that the particles P\, P2 and P3 at any instant form an isosceles triangle, eventually degenerated to a segment. We introduce cylindrical coordinates {r,6,z} 6 R + x S 1 x R as follows. First we put the origin of the coordinate system at the center of mass of mi, 77i2 a n a " "13. Then we define a new variable z = Z\ — Z2 G R which denotes the distance between the third particle P3 and the plane that contains the particles Pi and Pi with the convenient sign (positive in the upper position and negative in the down position). Finally we consider polar coordinates, (r, 0) € R + x S 1 , in the plane that contains P\ and P2. We choose the unit of mass in such a way that TTZI = 7712 = 1/2 and T7i3 = fx, and the unit of length is chosen so that the gravitational constant to be one. Then the kinetic energy in the coordinate system (r, 6, z, r, $, z) is given by
r =■K" + rH2
+
1+M " ) '
and the potential energy goes over to U =
1
fi
(z 2 + r 2 ) 1 / 2 '
8r
Therefore the Lagrangean eqxiations of motion for the isosceles problem are
d (dL\ 3L for q € {r, 6, * } . dt\dq) = ~~ dq where the Lagrangean is L = T — U; that is 1 | ( r ) = # - 72
Hr
" I T " " (Z1 4. r 2\3/2 '
2
d
It{T
9)-- = 0 ,
( " iU
dt V l + M J
(1) /iz
(22
+ r 2)3/2
•
65
We note that the second equation of system (1) leads to the angular momentum first integral C = r26 .
(2)
Of course, system (1) also has the first integral given by the energy H = T+U. 3
Equations of motion of the reduced isosceles problem
In this work we only consider solutions of the isosceles problem (1) having nonzero angular momentum (i.e. we do not consider solutions with collision between the masses). Under this assumption it is sufficient to consider solu tions of (1) having a fixed value of the angular momentum C = c for some c ^ 0, because the phase portrait of the isosceles problem on each angular momentum level c ^ 0 is the same, as it is shown in the following proposition. Proposition 1 Let i(t) — (r(t),r(t), z(t), z(t),8(t),8(t)) be a solution of the isosceles problem (1) with angular momentum C = c for some c ^ 0. 7/ we take a = c^/c2, then 7(t) = (ar{a3/2t),a-1/2f{a3/2t),az{a3/2t), 1 2 3 2 3 2 _3/2 3 2 a - / i ( a / t ) , 0(a / t.),a 0(a / *)) is a solution of (1) with angular mo mentum c. Assuming that the value of the angular momentum of the isosceles prob lem is fixed to C = c for some c ^ 0, we reduce in two units the dimension of the phase space obtaining the reduced isosceles problem .. _ c 2 r
1
~:i*~8r1~'
nr (22 + r 2)3/2 • ( )
(1 + /0* {z2 + r 2 ) 3 / 2 ' 4
Relationships between the reduced isosceles problem and the isosceles problem
Any solution j(t) — {r(t),r(t),z(t),z(t),0{t),9(t)) of the isosceles prob lem (1) with angular momentum c ^ 0 is given by the solution (p{t) = (r(t),f(t),z(t),z(t)) of the reduced isosceles problem (3) by taking 6{t) = ^fy
and
6(t) = ^ - ^ d r + e0
(mod 2TT) , (4)
where 9Q — 6(0) is an arbitrary integration constant. On the other hand given a solution tp(t) of the reduced isosceles problem (3), we obtain infinitely many
66 solutions 7e 0 (t) of the isosceles problem (1) with angular momentum c, one for each choosing of the integration constant 9Q € S 1 . Moreover these solutions are given by -ye0(t) = (
°(t)=
I -^dr
+ e0 = F(t) + e0,
Jo r*(T) 1 for a fixed 9Q & S . Then the set {Ugoesi7e0} " diffeomorphic to one of the following manifolds. (a) A circle S 1 C Sc formed by a periodic orbit of (1) with period S/J,)2, if'
128TTC 3 /(1 +
(b) A 2-dimensional torus S 1 x S 1 C £c, if if is a T-periodic orbit. Moreover this torus is formed by union of (i) periodic orbits of period mT, if F(T) — 2nl/m and I, m coprime;
with I G Z, m € N
(ii) quasiperiodic orbits, if F(T) = u2ir with u> an irrational number. (c) A cylinder S ' x R c ^ , if(fis neither the equilibrium point nor a periodic orbit. 5
Symmetric periodic orbits of the reduced isosceles problem
It is easy to check that equations (3) are invariant under the symmetry (r, f, z, z, t) —► (r, - r , - z , z, -t) . This means that if
67
This symmetry can be used to find r—symmetric periodic solutions of the reduced isosceles problem through the following well known result. Proposition 3 Let
(5)
i.e. the time reversibility symmetry, that will be denoted in what follows by t—symmetry. This symmetry is also used to find t—symmetric periodic solutions. Proposition 4 Let
be a solution of the reduced
(a) Iff(t) andz(t) are zero att = to andr(t) andz(t) are zero att = t0+T/4 but they are not simultaneously zero at any value of t S (to, to + T/4), then f(t) is a double-symmetric periodic solution of period T. (b) Ifr(t) and z(t) are zero att = to and r(t) and z(t) are zero att = to+T/4 but they are not simultaneously zero at any value of t € (to, to + T/A), then (fi(t) is a double-symmetric periodic solution of period T. 6
Restricted isosceles problems
To obtain the restricted isosceles problems we assume that the value of the mass 7713 is infinitesimally small (i.e. u = 0). Then the equations of motion
68
of the restricted isosceles problem become 1 r -= r62~8r2 ' 2 Jt{r o)- = 0 ,
(6)
z z — {z2 + r 2 ) 3 / 2 '
We observe that the first two equations of (6) are the equations of motion of a 2—body problem in polar coordinates. Then the particles P\ and P? (the primaries) move on the plane 2 = 0 describing a solution of this 2—body problem. Moreover the particle P3 that lies on the straight line orthogonal to the plane that contains Pi and P2, moves under the gravitational attraction of the previous two, but does not influence their motion. Thus for every value of the angular momentum c ^ 0 and every solution (r(t),6(t)) of that 2—body problem, system (6) defines a different restricted isosceles problem; it can be a circular, elliptic, parabolic or hyperbolic restricted isosceles problem depending on the nature of the solution (r(t),6(t)). As above if we assume that the value of the angular momentum is fixed to C = c for some c ^ O , then we can reduce in two units the dimension of the phase space obtaining the reduced restricted isosceles problem
c2 _ J_
r = r° z =
8r 2 ' z
(7)
(22+r2)3/2
In this work we only are interested in solutions of the reduced restricted isosceles problem (7) for c ± 0 that are periodic. So we will consider only the reduced circular and elliptic restricted isosceles problems, that we will call reduced circular Sitnikov problem and reduced elliptic Sitnikov problem respectively. 6.1
The reduced circular Sitnikov problem
Without loss of generality we can assume that the primaries are describing a circular orbit of radius 1/2 (or equivalently, a circular orbit of period 27r). This corresponds to fix the value of the angular momentum to c = 1/4. Then the equation of motion for the infinitesimal mass becomes Z =
~ ( * 2 + 1/4)3/2 '
W
69 which is the equation of the known circular Sitnikov problem. This equation can be integrated and we know analytic expressions of all its solutions (see for instance Belbruno et al.1 or Szebehely6). We note that the knowledge of an analytic expression for the solutions of the circular Sitnikov problem plays a key role in our analysis, because it allows us to prove our results analytically. Assume that (z(t), z(t)) is a solution of (8) with arbitrary initial conditions z(0) = ZQ and i(0) = z 0 . Then it is clear that
1_
~~ ToT3" ~ 8 r 2 '
Z =
~(z2+
z r 2 )3/ 2
;
having initial conditions r(0) = 1/2 ,
r(0) = 0 ,
2(0) = z0 ,
i(0) = i 0 .
That is,
Reduced elliptic Sitnikov problem
We assume that the primaries are describing an elliptic orbit with period 2ir and eccentricity e. This corresponds to fix the value of the angular momentum to c = \ / l — e 2 /4. Then, choosing conveniently the origin of time, a solution of the reduced elliptic Sitnikov problem is a solution ip(t) = (r(t),r(t),z(t),z(t)) of the reduced restricted isosceles problem
•■ = 1~ei3
r
~
Z =
16r
- J 2_
8r ' z 2 3 2 ~(z*+ r ) / '
70
having initial conditions r(0) = (1 - e)/2 ,
f(0) = 0 ,
z(0) = zo ,
i(0) = i 0 ,
for some zo. A) € R. Since r(t) is 27r-periodic, the periodic solutions of the reduced elliptic Sitnikov problem must have period multiple of 27r. Therefore if
Continuation of periodic orbits
We use the analytic continuation method of Poincare 4 (see for instance Siegel and Moser5 or Meyer and Hall3) to continue periodic orbits of the reduced circular Sitnikov problem (which are double-symmetric periodic orbits) to symmetric periodic orbits of the reduced isosceles problem for /x > 0 suffi ciently small. These periodic orbits are continued in two different ways. The first way goes directly from the reduced circular Sitnikov problem to the re duced isosceles problem. The second way uses two steps, first we continue the periodic orbits from the reduced circular Sitnikov problem to symmetric periodic orbits of the reduced elliptic Sitnikov problem for small values of the eccentricity e, and after we continue those symmetric periodic orbits of the reduced elliptic Sitnikov problem to the reduced isosceles problem for fj, > 0 sufficiently small. Using direct continuation (the first way) we can continue all periodic orbits of the reduced circular Sitnikov problem except the ones that have period multiple of An. In particular, we can continue the periodic orbits with period 27ru; and u> irrational, of the reduced circular Sitnikov problem, which are quasiperiodic orbits of the circular restricted isosceles problem. The continuation in two steps (the second way) allows us to continue only periodic orbits of the reduced circular Sitnikov problem with period T = 2np/q for all p, q € N coprime and p > q/(2\/2), which are also periodic orbits of the circular restricted isosceles problem. We note that the periodic orbits of the reduced circular Sitnikov problem that cannot be continued directly can be continued in two steps. Moreover the rest of the periodic orbits with period T = 2-Kp/q can be continued in both ways obtaining different periodic orbits for the reduced isosceles problem.
71 The main results about continuation of periodic orbits of the reduced circular Sitnikov problem to symmetric periodic orbits of the reduced isosceles problem for /x > 0 sufficiently small are summarized in the following theorem. Theorem 6 Let if be a periodic orbit of the reduced circular Sitnikov problem with period T > ir/\/2. Then
2TTUJ
with w > 1/(2\/2) an irrational number.
(i) ip can be continued to an 1-parameter family on /j, of doublesymmetric periodic orbits with period near T. (b) Case T = 2-npjq for some p, q € N coprime with p > q/{2\/2). (i) p odd: (1) ip can be continued to an 1-parameter family on p. of doublesymmetric periodic orbits with period near T. (2)
72
(a) two I—parameter families on e of r—symmetric periodic orbits and two 1—parameter families on e of t—symmetric periodic orbits (that are not double-symmetric) of the reduced elliptic Sitnikov problem, with period 2wp = qT, for e > 0 sufficiently small, when p is odd; (b) two 1—parameter families on e of double-symmetric periodic orbits of the reduced elliptic Sitnikov problem, with period 2irp = qT, for e > 0 sufficiently small, when p is even. 8
2-dimensional invariant tori for isosceles problem
Let YT denote the 2—dimensional invariant torus of the circular restricted isosceles problem that comes from a periodic orbit of the reduced circular Sitnikov problem with period T (see Section 6.1). Then we have the following result. Theorem 8 The torus YT can be continued to the following families of 2—dimensional tori of the isosceles problem with fi > 0 sufficiently small. (a) Case T = 2iru) with LJ > l/(2\/2) an irrational number. (i) YT can be continued to an 1-parameter family on fi of sional tori. (b) Case T = 2ivp/q for some p,q € N coprime with p >
2-dimen
q/(2\/2).
(i) p odd: (1) YT can be continued to an 1-parameter family on fj. of 2—dimen sional tori. (2) YT can be continued to four 2-^parameter families on n and r of 2—dimensional tori. (ii) p even and q ^ 1: (1) YT can be continued to an 1-parameter family onu of 2—dimen sional tori. (2) YT can be continued to two 2-parameter families on n and r of 2—dimensional tori. (Hi) p even and q = 1: (1) YT can be continued to two 2-parameter families on n and r of 2—dimensional tori.
73
Remember that the phase portrait of the isosceles problem on each angu lar momentum level c with c ^ 0 is the same (see Proposition 1). Therefore we have obtained 2—dimensional invariant periodic and quasiperiodic tori on each angular momentum level c ^ 0. Acknowledgments The authors are partially supported by a DGES grant number PB96-1153. References 1. E. Belbruno, J. Llibre and M. Olle, On the families of periodic orbits which bifurcate from the circular Sitnikov motions, Celestial Mechanics and Dynamical Astronomy 60, 99-129, (1994). 2. M. Corbera and J. LLibre, Periodic and quasi-periodic motions for the spatial isosceles 3-body problem, preprint, (1999). 3. K. R. Meyer and G. R. Hall, An Introduction to Hamiltonian Dynamical Systems, Springer-Verlag, New York, (1991). 4. H. Poincare, Les Methodes Nouvelles de la Mecanique Celeste, 3 Vols. Gauthier-Villars, Paris, 1892-1899, reprinted by Dover, New York, (1957). 5. C. L. Siegel and J. K. Moser, Lectures on Celestial Mechanics, SpringerVerlag, Berlin, (1971). 6. V. Szebehely, Theory of orbits, Academic Press, New York, (1967).
T H E GLOBAL F L O W F O R T H E SYNODICAL SPATIAL K E P L E R P R O B L E M MARCIA P. DANTAS Departamento de Matemdtica, Universidade Federal de Alagoas, 57032-160 - Maceid, AL, Brasil E-mail: [email protected] JAUME LLIBRE Departament de Matematiques, Universitat Autdnoma de Barcelona, 08193 - Bellaterra, Barcelona, Spain E-mail: [email protected]
1
Introduction
The knowledge of the global flow of the synodical spatial circular restricted three-body problem for fi = 0 (i.e. the synodical spatial Kepler problem) allows to describe the flow for fj, > 0 sufficiently small during a finite time. The description of this flow for n = 0 uses the integrability of the system as a Hamiltonian system. 2
Topological study of Hamiltonian systems
It is well known that a Hamiltonian system has at least one first integral, the Hamiltonian H. Every first integral F defines, through their inverse image F~l (/) = / / , sets which are invariant under the flow of the system. The topological study of a Hamiltonian system which has only one inde pendent first integral is the topological study of the map H; more precisely, the topological classification of the sets //, and of the points h € R for which the topology of //, changes. If the Hamiltonian system has more than one independent first integral that are in involution, then we study the topology of the intersections of the sets defined by the inverse image of the first integrals and the foliation of the phase space by these sets. In other words, for a Hamiltonian system with n degrees of freedom, we study the topology of the subsets //,, Ihjlt •••, Ih,fi,...,fr-i of the phase space where H,F\,„.,Fr-i axe r independent first integrals in involution. In short, we describe how the phase space is foliated by the energy levels //, and also by the subsequent levels defined by the other first integrals Fi, i = l,...,r — 1. Of course, r = n if the Hamiltonian system
74
75
is integrable. Furthermore we are interested in the change of topology of the subsets //,, //;,/,, •-., Ihji fr-i, more precisely, in the bifurcation sets of the maps H, ( # , > , ) , ..., (Jf ) Fi J ...,F r _ 1 ). During this last century several mathematicians have used the topological study for understanding the dynamics of Hamiltonian systems, integrable or not. Thus Poincare was the intellectual mentor of this idea which was a kind of new approach to study the flow of Hamiltonian systems. This topological method was used by Birkhoff5, in the study of the three-body problem and some general dynamical systems, and by Kaplan 14 , in the two-body problem. However a great interest in it appears after the publication of Smale 22 ' 23 on the topology of mechanical systems. These works of Smale were followed by many authors who extended the results of the topology of the n-body problem, for n > 2, i.e. the motion of n bodies which are moving in the space under the action of their gravitational forces. See for instance the papers [2], [6-7], [11-12] and [15-20] due to the authors Albouy, Cabral, Casasayas, Easton, Fomenko, Llibre, Martinez Alfaro, McCord, Meyer, Nunes, Pinol, Soler and Wang. The main interest of the topological study is that the intersections of the invariant sets are invariant by the flow of the Hamiltonian system; i.e. if an orbit has a point in that intersection, the whole orbit is contained in the intersection. In other words, an orbit is contained in some Ihj1,...,/r_1 and therefore, from the global point of view, the topology of 7/,,/j j r _ 1 is a very important invariant of the orbit (see Smale 22 ). Thus the topological study of a Hamiltonian system allows us to give a first topological classification of its orbits. More precisely, for integrable Hamiltonian systems under good conditions (see the Liouville-Arnold Theorem) the topological study allows us to know how the flow evolves and where the orbits are moving. This is a way to understand the global flow of integrable Hamiltonian systems. Liouville-Arnold Theorem (Arnold, 1963) Consider an integrable Hamil tonian system X # with n degrees of freedom defined by the C2 Hamiltonian H -.UxW1 - » R , where UxRn is an open subset of R2n andletH,Fi,...,Fn_i be n independent first integrals in involution. If //,,/, / n _, ^ 0 and (h, f\,..., fn-1) is a regular value of (H, Fi,.. .F n _ i ) , then the following state ments hold. (a) /*»,/,,...,/„_! is a n-dimensional the flow ofX-H-
submanifold of U x R" invariant under
(b) If the flow on a connected component J£ , ,_ of I/lj1,...,fn_1 is com plete, then Ihji,...,/„_! " diffeomorphic to Tk x R n ~* for some k £
76
{0, ...,n}. We note that i///J/i .../„ i ^ compact, then the flow on it is always complete. (c) Under the hypothesis (b) the flow on //J,/,,...,/„_, ** conjugated to a linear flowonTkxRn~k. For more details about Hamiltonian systems and the proof of the previous theorem see Abraham and Marsden 1 and Arnold 3 ' 4 . The Liouville-Arnold Theorem shows that, for integrable Hamiltonian systems, the invariant sets associated with the intersections of all indepen dent first integrals in involution are generically submanifolds of the phase space. Moreover, if the flow on such submanifolds is complete, then these submanifolds are diffeomorphic to the union of generalized cylinders and the flow on them is conjugated to a linear flow. We note that the invariant sets are not always manifolds. The topological study is important for Hamiltonian systems whether they are integrable or not. The non-integrable Hamiltonian systems with symme try (as the n-body problem for n > 2) have more than one independent first integral in involution. Thus for such systems many authors consider the foliation of the phase space by the known first integrals. For integrable Hamiltonian systems (as the two-body problem) we consider the foliation of the phase space by all the independent first integrals in involution. Once we know the global flow of an integrable Hamiltonian system, we can understand the flow of a perturbed Hamiltonian system which is sufficiently close to the integrable one during a finite time. This is achieved by using the continuous dependence of the solutions of differential equations with respect to initial conditions and parameters. In the last fifteen years there has been an intense development in the topological study of integrable Hamiltonian systems. We can divide these studies into three types. The first one is Fomenko's study of foliations of the phase space of arbitrary integrable Hamiltonian systems which satisfy some additional assumptions. He constructed a Morse type theory for Boot integrals defined on isoenergy surfaces of integrable Hamiltonian systems. By developing these ideas, he obtained general results about some aspects of the foliations stating topological invariants which, under additional assumptions, allow to conclude if the system is integrable or not (see Fomenko 12,13 and Casasayas et al.8). The second way to study the foliations of the phase space of integrable Hamiltonian systems is to study the foliation for a particular integrable Hamil tonian system without using any additional assumption (see the papers [7] and [15-19] due to the authors Casasayas, Llibre, Martinez Alfaro, Nunes, Piiiol
77 and Soler). The Liouville-Arnold Theorem helps in the study of this foliation but it is very far from solving it. In fact, from the point of view of obtain ing the foliation of the phase space by the invariant sets Ih and /A,/,,...,/,. for r = l,...,n — 1, this theorem only establishes the possible topology of a connected n-dimensional submanifold Jj!,/, /„ i when (h,/i,...,/n-i) is a regular value of the map {H,F\, ...,Fn-i) and the flow on them is complete. Thus, for a given value (h, f\, . . . , / n - i ) € R n , the Liouville-Arnold Theorem does not give the number of connected components and their topology. It only says that, for regular values under assumptions which are in general difficult to verify, as in the case of the completitude of the flow on a connected compo nent, its topology is one of the following generalized cylinders Rk x (S 1 ) , for some k = 0,1, ...,n; but it does not give the exact A;. Also the LiouvilleArnold Theorem does not give any information about the foliation of /;, by 7/ij,, or about the foliation of 7/,,/j by Ih,fi,h an( ^ s o o n It is known that the set of regular values (h, f \ , . . . , / n - i ) has full measure in R", but in general it is different from R n . In order to study the foliation of the phase space by the invariant sets -f/i,/,,...,/n_,, we must consider the critical and the regular values (h,/i,...,/n-i) • Thus when ( / i , / i , . . . , / n - i ) is a critical value, the invariant set /fc,/ ll ...,/ n _ l is called a separatrix level. Of course on a separatrix level the flow does not need to be conjugated to a linear flow. Finally, the third type of studies is about foliations of the phase space for Hamiltonian systems which are not integrable, but they have some inde pendent first integrals, as for instance the n-body problem with n > 2, see Albouy 2 , Cabral 6 , Easton 11 , McCord et al.20 and Smale 22 - 23 . In this work, we study the topology of a particular integrable Hamiltonian system, i.e. the synodical spatial Kepler problem which plays a main role in Celestial Mechanics. 3
Synodical spatial Kepler problem
It is well known that the two-body problem can be reduced to the Kepler problem. We can consider two kinds of referential (the inertial and the rotat ing) and two kinds of motion according to the dimension n of the configuration space, the planar when n = 2, and the spatial when n = 3. Of course each spatial motion takes place in a plane, but it continues to be a spatial motion once the phase space has dimension 3. Thus there are four different models of the Kepler problem: two with two degrees of freedom, the sidereal (synodical) planar Kepler problem, and two with three degrees of freedom, the sidereal (synodical) spatial Kepler problem.
78
These Kepler problems have been studied by several authors (see the papers [7] and [15-19] ). In this work we present the topological study of the synodical spatial Kepler problem. In fact a topological study of the four Kepler problems including the regularizations of the singularities can be found in [9] (where there are the proofs of the results here presented). By considering a coordinate system that rotates with frequency 1 around the X3-axis, the sidereal spatial Kepler problem becomes the synodical spatial Kepler problem. Thus, the synodical spatial Kepler problem is the integrable Hamiltonian system with three degrees of freedom associated to the Hamiltonian H
=
IYI2 1 2 " jx[ "
XlY2
+ X Yl
(1)
* '
with phase space {(X,Y) € (R 3 \ {(JJ.,0,0) ,(fi-1,0,0)}) independent first integrals in involution
x K 3 }, and three
C = —2H (Jacobi integral), r, lY|2 1 / -J . E = '—- ■—- (sidereal energy), M = |M| = | X A Y | (modulus of the sidereal or synodical angular
momentum).
This model of the Kepler problem is a limiting case of the circular re stricted three-body problem in rotating coordinates when the mass parameter of one of the primaries tends to zero. Thus, knowing the global flow of the synodical spatial Kepler problem we can obtain information about the flow of the circular restricted three-body problem during a finite time when the mass parameter of one of the primaries is sufficiently small. The first work on the topology of the Kepler problem was carried out by Kaplan 14 , in 1942. There he studies the foliation of the planar sidereal Kepler problem for negative values of the energy. In 1982 Llibre 15 studied the synodical planar Kepler problem describing the foliation of the phase space by //,, and the foliation of //, by Ihe, where h and c are the values of the Hamiltonian and the angular momentum, re spectively. He used Devaney's regularization 10 of the binary collision and McGehee's regularization at infinity21. In 1984 Casasayas and Llibre 7 , in their study of the anisotropic Kepler problem, considered the sidereal planar Kepler problem. They obtained the foliation of the phase space by //,, and the foliation of //, by 7/,c, where h and c are the values of the sidereal energy and the angular momentum, respectively.
79 Their study also includes the separatrix levels and the regularizations of the origin and infinity. In 1993 Llibre and Soler19 improved the foliation of the phase space by //,, and the foliation of Ik by Ikc for the synodical planar Kepler problem. In 1994 Llibre and Nunes 17 did a general study of integrable Hamiltonian systems with two degrees of freedom defined by a central force. Using the topological method, the synodical spatial Kepler problem was studied by Llibre and Martinez Alfaro16 in 1985. There the regularization of the collision singularity was considered. Subsequently in 1990 Llibre and Pinol 18 , in their study of the elliptical restricted three-body problem, included the regularizations of the collision and infinity singularities and so improved the results on the synodical spatial Kepler problem. In both papers the au thors stated the topology of the invariant sets Icm3f> i-e- they considered the independent first integrals in involution C, M$ = E + C/2 — XiY? — X2Y\ and F = M2 + Mf = M2 - Mi where M = (Mi, Af2, M 3 ). In this paper we study the synodical spatial Kepler problem taking as independent first integrals in involution C, E and M. We extend the previous results by stating the foliation of the phase space by Ic, the foliation of Ic by J ce , and the foliation of by Icem- In fact, in [9] we also give the topology of ■»m, -*c,mi -*e a n d
4
Jew
Topology of Ic, Ice and Icem
For presenting our results we need some notation. We define R + = (0, oo) and denote by Sm the m-dimensional sphere in R m + 1 of radius 1 with center at the origin 0. We note that S° is formed by two points. Throughout this paper manifold means differentiable manifold. If A and B are topological spaces the product space Ax B has the product topology. We use the symbol « to denote a diffeomorphism between two manifolds. Thus we present without proof the following diffeomorphisms: S ' x R « S2 \ {one point}, 5 1 x R 2 « S 1 x (S2 \ {one point}) 52 x R 3 « S2 x [S3 \ {one point}) S'xS'xR 2
«53\51, » S5 \ S2,
« S 1 x (R 2 s {0}) « S 1 x (S2 \ 5°) « 5 3 N { S 1 U 5 1 } ,
2
S x S x R « S2 x (R 3 \ {0}) « S2 x (S3 x 5°) » 5 5 x (S2 U S2). The results about the synodical spatial Kepler problem are stated in the following theorems. From them we obtain the foliation of the Jacobi levels Ic by the first integrals E and M. The first theorem concerns the topology of the Jacobi levels Ic.
80
Theorem 1 The Ic level of the synodical spatial Kepler problem is diffeomorphic to S5 \ S1) R i ^ x R ^ S 1 x R 4 « 5 2 X R 3 N S ' S2US1' 2 2 «S2xS2xR S US )
i/c>3, i/0
If c = 3, Ic is obtained by identifying a circle of S5 \ S2 with another circle of S 5 \ S 1 . This circle is formed by equilibrium points of the system. The next result is about the topology of the /«, levels. In this case we need to distinguish between the motion on the plane that does not contains the X3-axis (e ^ —c/2) and the motion on the plane that contains the X3-axis (e = - c / 2 ) . Theorem 2 The following statements hold for the Ice levels of the synodical spatial Kepler problem. (a) Ife^
- c / 2 , then S1 ife < 0 and |e + c/2| = m , , S1 x S 3 if e < 0 and je + c/2| < m», S 1 x R 3 « S 1 x (S3 x Poo) ife > 0.
(b) Ife=
- c / 2 , then c e
/ Sx x S 1 x R 2 « 5 1 x S 1 x ( 5 2 \ P 0 ) i/e < 0, * \ 5 l xS1 x S ' x R i/e>0.
Here PQ and P ^ denotes points of S2 and S3, respectively and m, =
l/J=Te. We remark that the circular spatial orbits are contained in S1 x S3 if m = m . > |e + c/2| > 0, and in S 1 x S 1 x R 2 if m = m . > |e + c/2| = 0. Theorem 3 77ie following statements hold for the I^m levels. (a) Ifrnj^O,
then
Icxm. *
' S1 5 1 x S1 S1 x S1 x S1 ' S1xSl S'xR S1 x R x S1
i/e<0 and0< ife < 0 a n d 0 < i/e < 0 and 0 < ife<0and0< ife>0and0< i/e > 0 and 0 <
|e + |e + |e + |e + |e + |e +
c/2| c/2| c/2| c/2| c/2| c/2|
= m = m„, = m <m„, <m<m., <m = m„ = m, < m.
81 (b) Ifm = Q, then
'52xR S2 xRxS0
i/e<0, ife>0.
Here m» = \fsj-2e. For finishing we state the results about the change of topology of the sets hem, i-e. for the n i p G = (C, E, M) . For this we introduce the bifurcation set E G formed by the points (c, e, m) in R 3 for which the map G fails to be locally trivial (see Llibre and Nunes 17 , and Smale 22 ). For an integrable Hamiltonian system with 3 degrees of freedom, we can characterize E G Proposition 4 For an integrable Hamiltonian system with 3 degrees of free dom, the bifurcation set is given by Ef = {y : y is a critical value off}
U {y : Iy changes its topology} ,
where f = ( / i , / 2 . / 3 ) , and fi, for i = 1,2,3 are three independent first inte grals in involution. For the synodical spatial Kepler problem we have the following results. Proposition 5 The set of critical values of the map G = (C, E, M) is given by Uc,e,m)
£ R 3 : - 2 e f e + ^\
< 1 and m=
, e + ^ I.
It is well known that, for the Kepler problem, the parabolic orbits are those for which the energy is zero (e = 0) and the collision or ejection orbits those for which the total angular momentum is zero (m = 0). Thus we have the next two results. Corollary 6 The parabolic orbits of the synodical spatial Kepler problem that are not collision or ejection are associated to regular values of the map G. Corollary 7 The orbits of the vertical plane containing the X3~axis of the synodical spatial Kepler problem that are not collision or ejection are associ ated to regular values of the map G = (C,E,M). From Theorem 3, and by Corollaries 6 and 7, we have the following results for the synodical spatial Kepler problem. Corollary 8 There is a change of topology in the Icem levels associated to regular values in the following cases. (a) When m = 0, i.e. at collision or ejection orbits. (b) When e = 0, i.e. at parabolic orbits.
82
Finally, by Theorem 4 we obtain the next result. Theorem 9 The bifurcation set of the synodical spatial Kepler problem is given by E G equal to { ( c , e , m ) G R 3 : c = 3} U {(c,e,m) G R 3 : m = 0 , j e + f | } u {(c, e, m) G R3 : e = 0, - 1 / (2m 2 )} . 5
Foliation of the phase space
From Theorems 1, 2 and 3 we state the foliation of the synodical spatial Kepler problem by the first integrals C, E and M. We note that the domain of definition of c, e and m is delimited by the bifurcations of the circular orbits e = —l/2m 2 , the planar motion, m = |e + c/2|, the collision orbits m = 0 (and consequently e = — c/2), and the parabolic orbits e — 0 (see Theorem 9). Corollary 10 The foliation of the phase space of the synodical spatial Kepler problem by the first integrals C, E and M is given in Tables 1, 2, S, 4 and 5. There m 3 = e + c/2, m» = l/\J—2e means the maximal value of the angular momentum that exists for e < 0, and e* satisfies \e + c/2\ = 1/y/—2e. From now on we give the description of the flow. We call direct space (DS) and retrograde space (RS) to the region of the Jacobi levels with e > —c/2 (m-3 > 0) and e < —c/2 (7713 < 0), respectively. If c > 3, then e £ [ei,e2] U [e3,oo) and each of the Jacobi levels Ic has two components I\c and /2c, see Table 1: (1) In Iic we have that e G [ei,e2] and 0 < |e + c/2| < m < m»: (a) If e = Ci, then m = m* and there is one retrograde circular planar orbit in the level Ice « S1. (b) If e\ < e < —c/2, then formed by the levels Ice planar orbits when m = when — (e + c/2) < m 771 =
the orbits are in the retrograde space RS « S1 x S3. There is a set of: 5 1 elliptic — (e + c/2); S1 x S 1 elliptic spatial orbits < m,\ and S 1 circular spatial orbits when
771,.
(c) If e = - c / 2 , then the orbits are in the level Ice « S 1 x S 1 x R 2 that contains the collision or ejection orbits. There is a set of: S2 ejection-collision elliptic orbits when m = 0; S1 x S 1 elliptic spatial orbits when 0 < m < m«; and S 1 circular spatial orbits when m = m».
83 (d) If —c/2 < e < e2, then the orbits are in the direct space DS formed by the levels Ice « S 1 x S 3 . There is a set of: S1 elliptic planar orbits when m = e + c/2; S 1 x S 1 elliptic spatial orbits when e + c/2 < m < m»; and S1 circular spatial orbits when m = m». (e) If e = e2, then m = m, and there is one direct circular planar orbit in the level Ice « S1. (2) In he we have that e € [e3,oo), 0 < e + c/2 < m < m» if e < 0, and 0 < e 4- c/2 < m if e > 0. The orbits are in the direct space DS : (a) If e = e3, then m — m, and there is one direct circular planar orbit in the level 7 ce « 5 1 . (b) If e3 < e < 0, then the direct space DS is formed by the levels Ice « S 1 x S 3 . There is a set of: S 1 elliptic planar orbits when m = e + c/2; S1 x S 1 elliptic spatial orbits when e + c/2 < m < m»; and S1 circular spatial orbits when m = mt. (c) If e = 0, then the orbits are in the direct space DS formed by the levels Ice w Sl x R 3 . There is a set of: S 1 parabolic planar orbits when m = c/2, and S 1 x S1 parabolic spatial orbits when m > c/2. (d) If e > 0, then the orbits are in the direct space DS formed by the levels Ice « S 1 x R 3 . There is a set of: S 1 hyperbolic planar orbits when m = e + c/2, and S 1 x S 1 hyperbolic spatial orbits when m > e + c/2. If c = 3, Ic is foliated as in the case c > 3 except for the value e = e2 = e3 at which J c e m is a circle of equilibrium points that are common to both components I\c and he- See Table 2. If 0 < c < 3, we have that e € [ei, co), 0 < |e + c/2| < m < m, if e < 0, and 0 < e + c/2 < m if e > 0. Each of the Jacobi levels Ic has one component, see Table 3: (1) If e = ei, then m = m, and there is one retrograde circular planar orbit in the level Ice ^ S1 which coincides with the retrograde space RS. (2) If ei < e < —c/2, then the orbits are in the retrograde space RS formed by the levels Ice « S 1 x S3. There is a set of: S 1 elliptic planar orbits when m = - (e 4- c/2); Sl x S 1 elliptic spatial orbits when - (e + c/2) < m <m»; and S 1 circular spatial orbits when m = m„.
84
h
Ice
1
s
5ax5
3
e € [ei,e2)U[e3,oo)
*cem
|e + c / 2 | < m
ei = e
1 s 1 S x5' S ' x 5 ' x Sl Sl x S 1 S'xR S ' x 5 ' x 51 S'xS1 S1 x S 1 S ' x S ' x Sl S1 x 5 '
—7713 = TTl = 771* —777,3 = m < 771, —77i3 < m < m« —m.3 < m = m« 0 = m < m, 0 < m < m, 0 < m = m, 7Ti3 = m < m , 7713 < 771 < 771, 7713 < m = m , 77i3 = m = m . T713 = m — m« 7713 = m < 77i« 7713 < m < m* 7713 < 771 = 771, 7713 = "^ 7713 < m
ei < e < - c / 2
he
s2
S1
xS'x R2
X R3
S^S3
sll s
he S1 X R4
s3
5
x
Sl
x5
e = -c/2
- c / 2 < e < e2 e = e2 e3 = e
3
xRJ
e3 < e < 0 0<e
S1 Sl 1 S'xS1 S " L l 1 S xS S " x 5 1 S1 x 5 ' S1 x R S'xRxS1
Table 1. The foliation of Ic for c > 3.
(3) If e = - c / 2 , then the orbits are in the level Ice « 5 1 x S1 x R 2 that contains the collision or ejection orbits. There is a set of: S2 ejectioncollision elliptic orbits when m = 0; 5 1 x 5 1 elliptic spatial orbits when 0 < m < m«; and 5 1 circular spatial orbits when m = m,. (4) If —c/2 < e < 0, then the orbits are in the direct space DS formed by the levels Ice f s S ' x S 3 . There is a set of: S1 elliptic planar orbits when m = e + c/2; S1 x S 1 elliptic spatial orbits when e + c/2 < m < m,; and S 1 circular spatial orbits when m = m,. (5) If e = 0, then the orbits are in the direct space DS formed by the levels Ice w 5 1 x R 3 . There is a set of: S 1 parabolic planar orbits when m = c/2, and S 1 x S 1 parabolic spatial orbits when m > c/2. (6) If e > 0, then the orbits are in the direct space DS formed by the levels Ice « S1 x R 3 . There is a set of: S1 hyperbolic planar orbits when m = e + c/2, and S1 x S 1 hyperbolic spatial orbits when m> e + c/2. If c = 0, we have that e G [ei, 00), 0 < |e + c/2| < m < m , if e < 0, and 0 < e + c/2 < m if e > 0. Each of the Jacobi levels Ic has one component, see
85
h
e € (ei.oo) ei = e
i-ce
S1
h
SxxS3 s3
ei < e < - c / 2
S 1 xS1 x R 2
e = -c/2
3
S'xS3 s
- c / 2 < e < e2
Sl
e = e2 = e 3 3
S'xS3 s
e2 < e < 0
Sl x R J
0<e
^cem
S1 1 S xS1 l S x 5 ' x S1 51 x S 1 S2xR 1 5 x S 1 x Sl 1 SlxSl S " l l 1 S xS S " Sl x S 1 x S 1 S'xS1 S1 1 S xS1 1 5 x 51 x 51 51 x S 1 5'xR S'xRxS1
|e + c/2| < m -7713 = m = m , — 7713 = 771 < 771, —7773 < 777 < 777, —7713 < 771 = 771,
0 = m < m» 0 < m < m» 0 < m — m, 7Ti3 = m < m « 7773 < 771 < 771, 7773 < 777 = 777* 7713 = 7 7 1 = 771, 7713 = m < m , 7713 < m < m , 7713 < 771 = 771* 77I3 = 771 77I3 < 771
Table 2. T h e foliation of Ic for c = 3.
/c
e £ [ei,oo) ei = e
■"ce
S
1
S1 x 5
3
ei < e < - c / 2
S2 X
S1 xS1 x R 2
e = -c/2
R3 S1
S'x53
-c/2 < e < 0
S1 x R J
0<e
*cem
1 s l S xS1 l l S xS x 51 1 S xS' S'^xR S ' x S ' x 51 SIx51 51 x S 1 1 5 x S1 x 5 1 5'xS1 S'xR S'xRxS1
|e + c / 2 | < 771 —7713 = m = 771. —7713 = m < m » —7713 < m
< m ,
—7713 < m = m . 0 = 771 < 771, 0 < 771 < 771, 0 < m = 771, 7713 = m < m , 7713 < 771 < m , 7773 < m = 771, 7713 =
m
7713 < 771
Table 3. The foliation of Ic for 0 < c < 3.
Table 4: (1) If e = ei, then m = m, and there is one retrograde circular planar orbit
86
h
s2
eG [e ls oo) e\ = e
Ice l
s
1
3
X
S xS
X R
Sl xS'x 5'xR Sl x R J
s2
ei < e < 0 e= 0 0<e
■•cem
S1
S'xS1 S ' x S ' x 51 Sl xSl SzxRx5u 5'xRxS1 S'xR S'xRxS1
|e + c/2| < —m3 = m = —m3 = m < -7Ti3 < m < —T7i3 <m = 0= m 0< m T7l3 = m
m m» m, m» m,
7Tl3 < 771
Table 4. The foliation of Ic for c = 0.
in the level Ice « S 1 which coincides with the retrograde space RS. (2) If ex < e < 0, then the orbits are in the retrograde space RS formed by the levels Ice K S1 X S3. There is a set of: S1 elliptic planar orbits when m = - e ; Sl x S 1 elliptic spatial orbits when —e < m < m»; and S 1 circular spatial orbits when m — m,. (3) If e = 0, then the orbits are in the level I*e % Sl x S 1 x S1 x R that contains the collision or ejection orbits. There is a set of: S 2 ejectionparabolic orbits and S2 parabolic-collision orbits when m = 0, and S1 x S1 parabolic spatial orbits when m > 0. (4) If e > 0, then the orbits are in the direct space DS formed by the levels Ice ss S1 x R 3 . There is a set of: S 1 hyperbolic planar orbits when m = e + c/2, and S1 x S1 hyperbolic spatial orbits when m > e + c/2. If c < 0, we have that e € [ei, oo), 0 < |e + c/2| < m < m , if e < 0, and 0 < e + c/2 < m if e > 0. Each of the Jacobi levels Ic has one component, see Table 5: (1) If e = ei, then m = m* and there is one retrograde circular planar orbit in the level Ice ss S 1 which coincides with the retrograde space RS. (2) If ei < e < 0, then the orbits are in the retrograde space RS formed by the levels Ice ss S 1 x S3. There is a set of: S1 elliptic planar orbits when m = —e; S 1 x S 1 elliptic spatial orbits when —e<m<m.\ and S 1 circular spatial orbits when m = m«. (3) If e = 0, then the orbits are in the retrograde space RS formed by the levels Ice «s S 1 x R 3 . There is a set of: S 1 parabolic planar orbits when m = —c/2, and S 1 x S1 parabolic spatial orbits when m > - c / 2 .
87
h
he l
s 2
S1 x S 3
s
X
s2
S1 x R J
X
R
S1 x S ' x S'xl S1 x R J
e e [ei,oo) e\ = e
-*cem
1 1s S x5' ex < e < 0 S ' x S ' x 5 1 SlxSl 0 < e < -c/2 S'xR 5'xRxS1 e = -c/2 S'xRxS1' S'xRxS1 S'xR -c/2 < e 5'xRxS1
|e 4- c/2| < m —77i3 = TTI = m » - m 3 = 771 < 771, —7713 < 771 < 771. —7713 < TTl = m « —777,3 = 771 —77I3 < 771 0 = 777 0<77l 7773 = 777 7773 < 771
Table 5. The foliation of Ic for c < 0.
(4) If 0 < e < —c/2, then the orbits are in the retrograde space RS formed by the levels Ice « S 1 x R 3 . There is a set of: S 1 hyperbolic planar orbits when m = — (e + c/2), and 5 1 x S1 hyperbolic spatial orbits when m > - ( e + c/2). (5) If e = - c / 2 > 0, then the orbits are in the level Ice w S 1 x S 1 x S1 x R that contains the collision or ejection orbits. There is a set of: S2 ejectionhyperbolic orbits and S2 hyperbolic-collision orbits when m = 0, and Sl x S 1 hyperbolic spatial orbits if 0 < m. (6) If e > —c/2, then the orbits are in the direct space DS formed by the levels Ice « 5 ' x R 3 . There is a set of: S1 hyperbolic planar orbits when 777 = e + c/2, and S1 x S 1 hyperbolic spatial orbits when m > e + c/2.
References 1. R. Abraham and J.E. Marsden, Foundations of Mechanics, Benjamin, Reading, Massachussets, (1978). 2. A. Albouy, Integral manifolds of the n-body problem, Inventiones Math. 114, 463-488, (1993). 3. V.I. Arnold, Mathematical Methods of Classical Mechanics,, SpringerVerlag, (1978). 4. V.I. Arnold, V.V. Kozlov, A.I. Neishtadt, Dynamical Systems III, Ency clopaedia of mathematical sciences, Springer-Verlag, Berlin, (1978). 5. G.D. Birkhoff, Dynamical Systems, New York, (1927).
88
6. H. Cabral, On the integral manifolds of the n-body problem, Inventiones Math. 20 ,59-72, (1973). 7. J. Casasayas and J. Llibre, Qualitative analysis of the anisotropic Kepler problem, Memoirs of the Amer. Math. Soc, Vol. 52, n° 312, (1984). 8. J. Casasayas, J. Martinez Alfaro and A. Nunes, Knots and links in integrable Hamiltonian systems, J. Knot Theory Ramification 7,123-153, (1998). 9. M. P. Dantas and J. Llibre, The foliation of the phase space for the 3-dimensional Kepler problem, preprint, (1999). 10. R. Devaney, Singularities in Classical Mechanics Systems, in Ergodic The ory and Dynamical Systems I, Proceedings Special Year, Maryland 197980, A. Katok (ed.), Birkhauser, Basel, 211-333, (1981). 11. R. W. Easton, Some topology of n-body problems, J. Differential Equa tions 19, 258-269 (1975). 12. A. T. Fomenko, On typical topological properties of integrable Hamilto nian systems, Izv. Akad. Nauk (SSSR) Ser. Mat. 52 ,378-407, (1988); English Transl. in Math. USSR Izv. 32 (1989). 13. A. T. Fomenko, Differential equations and applications to problems in Physics and Mechanics, Mechanics Analysis and Geometry: 200 years after Lagrange, M. Francaviglia (editor), Elsevier Science Publishers B. V., (1991). 14. W. Kaplan, Topology of the two-body problem, American Mathematical Monthly 49, 316-326,(1942). 15. J. Llibre, On the restricted three-body problem when the mass parameter is small, Celest. Mech. 28, 83-105, (1982). 16. J. Llibre and J. Martinez Alfaro, Ejection and collision orbits of the spa tial restricted three-body problem, Celest. Mech. 35, 113-128, (1985). 17. J. Llibre and A. Nunes, Separatrix Surfaces and Invariant Manifolds of a Class of Integrable Hamiltonian Systems and Their Perturbations, Mem. of the Amer. Math. Soc, 107, n° 513, (1994). 18. J. Llibre and C. Pinol, On the elliptic restricted three-body problem, Celest. Mech. 48 , 319-345, (1990). 19. J. Llibre and J. Soler, Global flow of the rotating Kepler problem, in Hamiltonian System and Celestial Mechanics, Advanced Series in Non linear Dynamics, Vol. 4, World Scientific, Singapore, 125-140, (1993). 20. C. K. McCord, K. R. Meyer and Q. Wang, The integral manifolds of the three body problem, Mem. Amer. Mat. Soc, 132, n° 628, (1998). 21. R. McGehee, A stable manifold theorem for degenerate fixed points with applications to celestial mechanics, J. Differential Equations 14, 70-88, (1973).
89
22. S. Smale, Topology and mechanics, I, Inventiones Math. 10, 305-331, (1970). 23. S. Smale, Topology and mechanics, II. The planar n-body problem, In ventiones Math. 11, 45-64, (1970).
U N B O U N D E D G R O W T H OF E N E R G Y I N P E R I O D I C P E R T U R B A T I O N S OF GEODESIC FLOWS OF T H E T O R U S AMADEU DELSHAMS Departament de Matematica Aplicada I, Universitat Politecnica de Catalunya, Diagonal 647, 08028 Barcelona, Spain E-mail: [email protected] RAFAEL DE LA LLAVE Department of Mathematics, University of Texas at Austin, Austin, TX, 78712, USA E-mail: [email protected] TERE M. SEARA Departament de Matematica Aplicada I, Universitat Politecnica de Catalunya, Diagonal 647, 08028 Barcelona, Spain E-mail: [email protected]
1
Introduction
The goal of these notes is to summarize the main ideas of a paper by the authors. l We will study the effects of a periodic external potential on the dynamics of the geodesic flow in the (2-dimensional) torus. This is a Hamiltonian system with two and a half degrees of freedom with Hamiltonian H{p,q,t) = ^gq{p,p) + U(q,t). Here g is the metric on the torus, and U is the external periodic potential. If U = 0, the system is autonomous and, therefore, the energy of the geodesic flow Ho(p,q) = \gq(p,p) is preserved. Note also that when the metric is flat, the geodesic flow is integrable. We will establish, using geometric methods, a result that had been es tablished by J. Mather using variational methods. 2 Namely, that for generic metrics and potentials—in particular for arbitrarily small potentials and for metrics arbitrarily close to integrable—, one can find orbits whose energy grows to infinity. More precisely, we will indicate the proof of: Theorem 1 For every r = 1 5 , . . . , oo, u, there exists a Cr-residual set in the set of metrics g on T 2 , and of periodic potentials U : T 2 x T 1 —♦ R, such that the Hamiltonian H(p,q,t) = ^gq(p,p) + U(q,t) has an orbit of unbounded energy.
90
91
The result is somewhat more surprising if we realize that, for high en ergy, the potential is a small perturbation of the geodesic flow and that since there are two time scales involved (see subsection 2.2), the frequency of the perturbation is smaller than the frequency of the geodesic flow. In this sense this result can be understood as an analogue to Arnold diffusion. The difference between this work and other approaches to Arnold diffusion in the literature comes from the fact that here, the unperturbed system, which is the geodesic flow, is a two degrees of freedom Hamiltonian system that, in general, is not integrable and it is assumed to have some hyperbolicity properties. Indeed, the hyperbolicity properties involve that the system contains hyperbolic sets with transversal intersection in an energy surface. This is somewhat stronger hyperbolicity that the so-called a-priori unstable systems in the notation of Chierchia and Gallavotti 3 . We propose the name a-priori chaotic for this kind of systems, where there are some conserved quantities, but there are hyperbolic orbits with transverse heteroclinic intersections in the manifolds corresponding to the conserved quantities. This result on instability has been proved by J. Mather using a variational approach (the differentiability requirements were much smaller). The proof presented here will rely on standard tools in the study of dynamical systems. Most of them have been used traditionally in Arnold diffusion. We also use very strongly the theory of hyperbolic dynamical systems. We hope that this can serve as a clarification of the relation between variational meth ods and more geometric methods. We also point out that S. Bolotin and D. Treschev have presented another geometric proof 4 . The main difference with the method presented here is that they rely more on KAM theory and they do not use the theory of normally hyperbolic manifolds. The generic conditions that we need can be described rather explicitly. We require for the metric that the geodesic flow possesses a hyperbolic peri odic orbit with a homoclinic orbit. Once we fix this periodic orbit and the homoclinic connection, we require on the potential that the Poincare function £ ( r ) computed on it is not identically constant (see (24)). The proof can be divided in several steps that are largely independent. As we mentioned before, each step only requires almost standard tools. Let us go over a thumbnail sketch of all these steps and they will be fleshed out in the following sections. The same method of proof allows to prove somewhat more general results. If we assume the consequences of one step, then all the subsequent ones remain valid. Moreover, we can use the same method to obtain more consequences.
92 Indeed, the proof will be established by the somewhat more general Theorem 18. Riemannian
Geometry
It was shown already by Morse—with some improvements by Mather—that a generic metric possesses a hyperbolic closed geodesic with a homoclinic orbit. Note that this shows that the geodesic flow is not integrable and is a-priori unstable in the sense of Chierchia and Gallavotti. 3 It is a crucial observation that, when we consider the geodesic flow as a 4-dimensional dynamical system, a closed geodesic lifts to a cylinder A (one periodic orbit for each energy surface). This cylinder is invariant and normally hyperbolic. Scaled variables As can be seen by elementary dimensional analysis, if we rescale the momenta and the time, we can transform the perturbation introduced by U for high energy as an small and slow perturbation of the geodesic flow. Intuitively, if the geodesies are moving very fast, in their time scale, the potential is changing slowly. For the analysis that will follow, the slowness is much more important than the smallness. Another elementary remark is that we can consider the system as au tonomous by considering one extra angle variable. We can come back from this extended phase space to the regular phase space by means of the Poincard return map. In particular, the cylinder A is invariant under the action of the unperturbed Poincar6 map, and it lifts to a 3-dimensional manifold A, invari ant under the unperturbed extended flow. In this sketch, we will be somewhat cavalier about distinguishing between the objects in the extended phase space and in the regular phase space. The theory of normally hyperbolic manifolds Since A is uniformly hyperbolic in the extended phase space, we can find a manifold close to it that is invariant for the perturbed system. The stable and unstable manifolds of this invariant manifold still intersect as transversally as possible. The importance of this invariant manifold is that it will serve as a template for many types of orbits.
93 Averaging theory Taking advantage of the fact that the perturbation is slow, we can study the motion on the perturbed invariant manifold using the method of averaging. Assuming that the system is differentiable enough, we can perform several steps of averaging so that the system on the 3-dimensional invariant manifold is a very small periodic perturbation of an integrable Hamiltonian system. KAM theory Once we know that a system is close to an integrable Hamiltonian system, we can find out that it is covered very densely by invariant tori, emerging from invariant circles of the Poincare map. Poincare-Melnikov theory It turns out to be possible to compute perturbatively the gain in energy for a homoclinic excursion. Under some generic hypotheses, it is possible to show that a homoclinic excursion leads to a jump from one KAM torus to another, giving rise to heteroclinic orbits between them. This idea can be formulated precisely by introducing what we call the scat tering map which relates the asymptotics in the future with the asymptotics in the past. This map can be constructed precisely using hyperbolicity theory and it can be computed perturbatively using the Poincare-Melnikov method (the fundamental theorem of calculus and a—somewhat delicate—passage to the limit). Under the assumption that the scattering map moves enough to jump over the gaps between KAM tori—which we have already shown are very small—we can construct pseudo-orbits whose energy grows unboundedly. These pseudorbits can be described as follows. When the potential is in such a condition that produces a jump of energy in the direction that we want, we execute the jump, following a heteroclinic orbit, and in the situations when it is not favorable, we just bid our time near a KAM torus. Shadowing lemma We need to construct orbits that follow these pseudo-orbits closely. This can be done by a topological argument very close to the original one invoking the obstruction property. 5 ' 6 Now, we turn to the task of fleshing out this sketch.
94 2 2.1
The m a i n s t e p s of t h e proof Riemannian geometry and dynamics of the geodesic flow
Let us begin by describing the dynamics of the geodesic flow. The phase space of the geodesic flow is T*T 2 = R 2 x T 2 . We will denote the coordinates in T 2 by q and the cotangent directions by p. The geodesic flow is a two degrees of freedom Hamiltonian, with Hamiltonian Ho(p,q) = ^gq(p,p), with respect to the canonical symplectic form Q, which, moreover, is exact: Q = dd. In local coordinates, 9 — J^Ptdgt, Cl — Y^i^Pi A dqi. We will denote by $(*;Pi9) = $t(Pi<7) this geodesic flow, and we remark that, even though this flow is not integrable, the energy E = HQ is preserved. For each E, we will denote £ E = {(p,q) : Ho(p,q) = E}, which is a 3dimensional manifold diffeomorphic to T 1 x T 2 , invariant under the geodesic flow. We observe that, for any £0 > 0, tBo = UE>E0^B - [E0, 00) x T 1 x T 2 , that is, the phase space is foliated by the energy surfaces. Given an arbitrary geodesic "A" : R —» T 2 , parameterized by arc length, we will denote by Ae(t) — (XE(t), X%(t)) the orbit of the geodesic flow that lies in the energy surface E#, and whose projection over q turns along the range of "A". Moreover, we fix the origin of time in XE so that it corresponds to the origin of the parameterization in "A". (Formally Ho(XE(t)) = E, and Range("A") = Range(A|), "A"(0) = A«,(0).) Note that
(A|(*).A^(t)) = (V2EXp1/2 (yZEt) ,A«/2 (V2Et)) .
(1)
So that, in this situation, the role of E is just a rescaling of time. The following theorem is due to Morse and Mather: Theorem 2 For a Cr generic metric g,r>2, and for any value of the Hamil tonian Ho(p,q) = E > 0, there exists a periodic orbit Ag(t), as in (1), of the Hamiltonian flow associated to Ho whose range AE is a normally hyperbolic invariant manifold in the energy surface. Its stable and unstable manifolds W? ,u are 2-dimensional, and there exists a homoclinic orbit 7B(<) such that its range 7 B satisfies *
Vc(w^\A*)n(wx a \A*). Moreover, this intersection is transverse as intersection of invariant manifolds in the energy surface along 7£. For E — 1/2, we have that dist(Ai/ 2 (* + a±),7i/2M) —► 0
as t —> ±00.
(2)
95 We suppose that the length of the unit geodesic "A" on the metric g is 1. Note that the speed of a unit geodesic is 1 and, therefore, its energy is 1/2. Then "A" as an orbit of the geodesic flow has period 1 and energy 1/2. Because of this we use the notation Ai/2The quantity A = a + - a_ is called phase shift. In this situation the energy is preserved and this makes the periodic orbit non-hyperbolic in the whole space. In order to obtain some hyperbolic object it is natural to discuss what happens for all energy surfaces considering A = U B > E ^-B for all values of the energy. Lemma 3 Define A = U B > B &■£• This is a 2-dimensional manifold with boundary which is diffeomorphic to [EQ, CO) X T 1 , and the canonical symplectic form on T*T 2 Q restricted to A is non-degenerate and invariant under the geodesic flow $ t . The stable and unstable manifolds to A, W\, W%, are Z-dimensional man ifolds diffeomorphic to [EQ, co)xT 1 xR, and7 = U E > £ 0 1E is 0,2-dimensional manifold diffeomorphic to [£0,00) x R that verifies 7c(WX\A)n(wx\A).
Let us note that the intersection of the stable and unstable manifolds along 7 is transverse, because TxW\nTxWl
=
Txl.
This meaning of transversal intersection has several important consequences. For example, 7 will be a locally unique intersection. Since the manifold A is a cylinder, we are going to take a coordinate sys tem in A, given by one real coordinate (momentum) and one angle (position). The real coordinate will be J — \/2Ho > T/2EQ. For the angle coordinate, we will take (f € T 1 , which is determined by dJ A dtp = fi|A and ip = 0 corre sponds to the origin of the parameterization in "A". We can identify a point in A with the values (J, (p) taken by these coordinates. The coordinate repre sentation of the periodic orbit AE(<) in these coordinates is given by (J, Jt), where J = \/2E. 2.2
The scaled problem
In order to make the perturbative structure of the problem more apparent we will scale the variables p and the time ( in the Hamiltonian H(p, q, t) = \gq{p,p) + u{q,t). We pick a (large) number Em, and call e = l/y/E^. Since e2H(p,q) = ^gq(ep,ep) + e2U(q,t), we introduce p = ep and consider the symplectic form
96 fi = dp Adq = eQ., noting that q,p are conjugate variables in fi. Next, we introduce a new time t = t/e, and we obtain a Hamiltonian system with Hamiltonian H€(p, Q, eT) = ±gq(p, p) + s2U(q, ^) = #o(p,
(3)
We also introduce E = E/Et, and we are going to see that, for sufficiently large E,, there exist hyperbolic whiskered tori with heteroclinic orbits between them whose energies are between E = 1/2 + 6 and E = 2 — 6 for some small 6 > 0. Since the result is for all sufficiently large Em, we will be able to take a larger E* and continue the transition chain. From now on and until further notice, we will drop the bar from the problem. We will refer to the bar variables as the rescaled variables and the original ones as the physical variables. Then the Hamiltonian He and all the functions derived from it will be 1/e periodic in time. In order to do this more apparent we will use the notation given in (3). 2.3
The inner and outer maps
We will define two maps in A which describe, respectively, the inner dynamics in A and the outer dynamics given by the homoclinic connections that lie in 7. In the unperturbed case the dynamics is known, and we will be able to give explicit expressions for these maps. Later on, when we consider the problem with external potential, the perturbative computation of these maps will be the main tool to control the behavior of the orbits. The inner map FQ will be the time l/e map of the geodesic flow, also called Poincar£ map, restricted to A. As we are dealing with the autonomous case, the phase portrait of this map is identical to that of the geodesic flow. For simplicity we will consider the Poincare map associated to the initial time 0, that is Fo(x) = $ i / e ( i ) , where $t(x) denotes the geodesic flow on the torus. We note that the behavior of this unperturbed map is exactly the same as the unperturbed flow, and then all the possible choices of the initial time give rise to the same map. Even though this map is defined in all the phase space, we are interested only on the dynamics in A, which, in the coordinates (J, (p), is given by: F0{J,
(4)
We call attention to the fact that Fo is a twist map and this will be useful to study the effects of the perturbation on it. We also note that the periodic orbits of the geodesic flow are invariant curves for this map, and they are given in these coordinates by J = const.
97
The outer map or "scattering map" So : A —» A, associated to 7, will transform the asymptotic point at —00 of an orbit homoclinic to A into the asymptotic point at +00. Thus we define x+ = S 0 (x_) if there exists z 6 7 C T*T 2 , such that dist ($t(x±), $t(z)) —» 0 ,
as
t —► ±00.
From the results of Morse and Mather of Theorem 2: dist(A 1/2 (t + a ± ) , 7 1 / 2 ( 0 ) -» 0,
as
t -» ±00,
(5)
and the rescaling properties give
dist (AE (t/y/W + a±/VwYjE
(t/VzE))
Hence, expressing the map So in the coordinates
^>0 as t -► ±00.
(6)
(J,f),
5 0 (J,a_ + tp) = (J,a+ +
(7)
Now, in the unperturbed case, the dynamics is clear: if we take a periodic orbit KB, we have: FO(AE)
= AE,
(8)
SO(AE)
= AE.
(9)
These two conditions have a geometrical meaning: (8) means that A^ is an invariant curve of the Poincare map, and therefore, its orbits under the continuous flow constitute a 2-torus, which is invariant under the flow. Con dition (9) means that the stable and unstable manifolds of this curve only have homoclinic connections. After the scaling done in section 2.2, we have that for high energy, or what is the same, for small e, the external potential is an small and slow perturbation of the geodesic flow. Therefore, all the geometric structures that we constructed based on normal hyperbolicity and transversality persist for high energy. In particular, the manifold A will persist as well as the transversality of the intersection of its stable and unstable manifolds along 7. On the other hand this perturbation is 1/e-periodic in time. This will allow us to define F, S analogues of the maps FQ, SO, a n ( l to compute them perturbatively. The idea now is to use the maps F, S when the external forcing is acting. As we will look for orbits with growing energy, we will look for invariant curves
98
of F, which have heteroclinic connections between them. Then we have to look for closed curves Aj, A2 such that F(A i ) = Ai>
t = l,2,
S(A!)nA2^0. Therefore, the main technical goal will be to compute perturbatively, for high energy, the inner and the outer maps F and S, show that applying them alternatively we can construct heteroclinic orbits between several invariant curves, and then, show that these orbits can be shadowed by real orbits. Moreover, as we characterize the objects in A by the angle ? and the mo mentum J, which is the square root of the energy, to compute the outer map 5 , we will follow the classical Poincare' method, to measure the increment of energy during a homoclinic excursion. 2.4
The perturbed invariant manifold. Normal hyperbolicity
In order to study the dynamics with external potential, we consider the au tonomous flow associated to the Hamiltonian (3). dH0 dU P = —dT( p > q )~' 6 ~d a ~( q > £s '>
6 = e
2
q=^(p,q),
> (10)
5 = 1,
defined on the extended phase space T*T 2 x ^T 1 . The reason why we introduce the notation 6 = e2 above is so that we can study more conveniently the limit e —► 0. The normally hyperbolic pertur bation theory can make statements for the above system for all 6 < 60 and one can choose 60 independently of e. This allows us to make statements for £ small enough, even if the scaled flow is meaningless in this limit. We note that for the unperturbed case (5 = 0), we have that A := A x i f 1 ~ [Jo, 00) x T 1 x -T1 £
£
is a 3-dimensional manifold locally invariant for this flow. This manifold is also normally hyperbolic. The next theorem assures the persistence of a hyperbolic manifold for e > 0 close to this one, and it relies in results of hyperbolicity theory. x
99
Theorem 4 Assume that we have a system of equations as in (10), and of class Cr, 2 < r < oo. Then, there exists em > 0 such that for \e\ < e», there is a C r _ 1 function T: [Jo + # e 2 , o o ) x T ' x ^ T 1 x (0,e 2 ) -» T*T 2 x -T1 £
£
such that A£ = T ({Jo + Ke2,oo) x T 1 x ^ T 1 x {e 2 })
(11)
is locally invariant for the flow (10) and is £2-close to A in the Cr~2 sense. Moreover, A£ is a hyperbolic manifold. We can find a C r _ 1 function T* : [Jo + A ^ . o o ) x T 1 x -T1 x [0,oo) x [0,£2) - T*T 2 x ^ T 1 such that its (local) stable invariant manifold takes the form Ws'loc(Ae)
= F ([Jo + Ks2, oo) x T 1 x i f 1 x [0, oo) x {e 2 }^ .
(12)
Ifx = F(J, ip, s,e2) e A,, then W^s-loc(i) = . P ( { J } x {^} x {s} x [0, oo) x {e2}). Therefore W^8,loc(A£) is e2-close to W s - loc (A) in the Cr~2 sense. Analogous results hold for the unstable manifold. Let us observe that Theorem 4 only guarantees local invariance for A£. But, in a moment, we will show that this manifold has KAM 2-dimensional tori, that will provide with invariant boundaries for it. Therefore, it is possible to take A£ invariant. This remark leads to some extra uniqueness, etc. which could be useful for future developments. However, we will not use it in the present paper. 2.5
The perturbed inner map. Averaging theory
Once we have the 3-dimensional perturbed invariant manifold A£ given by (11), we can consider the flow restricted to it, and, after some symplectic changes of variables, we can write it as a Cr~2 Hamiltonian system with one and a half degrees of freedom, and 1/e periodic in the time s. More precisely, this Hamiltonian can be written in the local coordinates (J,
(13)
The Hamiltonian ke( J, y>, es) is non-autonomous, but it has an slow, periodic, dependence on the time s. We can apply the Adiabatic Invariance Theorem, eliminating the fast angle (f by means of an averaging procedure up to any order £ m . (More details on the averaging procedure can be found in 7 or in '.)
100
Theorem 5 Let ke(J,(p,es) be the Cr~2 Hamiltonian, l-periodic in
= K°e(Les) +
£m+1Kl{I,tP,es),
with K®(I,es) — ^P + Oc1^2)) where the notation O c ( £ 2 ) means a function whose C1 norm is 0(£ 2 )Now, we consider the perturbed inner map in these new variables. First of all, let us note that the Poincare return maps are defined in the intersection of the 3-dimensional perturbed invariant manifold Ae given by (11) with the Poincare sections given by fixing the time s, that we denote A* C T*T 2 : (Als)
= T{[J + Ke2,oo]xT1
x{s},e2).
By Theorem 4, for every s, A| is e2-close to the unperturbed manifold A in the C~2 sense. In the variables (I,ip), we have that the perturbed inner map F, defined on A, is £m-close to an integrable map: Lemma 6 In the conditions of Theorem 5, the map F : A —» A, which is exact symplectic, can be written in the coordinates (I, xf)) as F(I,4>) = (l,iP+±A(I,e))
+emRE(I,iP),
(14)
where A(I,e) = e f*/£ D1K°(I,es)ds = I + 0(e2), and R£ is a Cr-m-4 func tion. Now, in these new coordinates, the KAM theorem provides the existence of invariant curves for the perturbed inner map F, which fill out the manifold A except for gaps which are of the order of e m / 2 + 1 : Theorem 7 Assume thatr — m — 4 > 6. Then, ifeis small enough, for a set of w of Diophantine numbers of exponent 5/4, we can find invariant curves Au for F which are the graph of C~m~7 functions uw, with \\uw\\Cr-m-7 < const e m / 2 , and the motion on them is Cr~m~7 conjugate to the rotation by w. The set of these invariant curves cover the whole annulus except for a set of measure smaller than const e m / 2 + 1 . Moreover, we can find expansions uu — u2>+£ m u^+r w , with ||r||c,—m-s < const e2m, and ||u 1 || Cr _ m _ g < const.
101
Remark 8 Note that these KAM curves with frequency w that we have pro duced for the map F are really whiskered tori for the extended flow with frequencies (u>,e). They could have been produced also by appealing to the Graff-Zehnder Theorem, but it is not direct from this theorem that the higher is the differentiability of the flow the smaller are the gaps between the tori. In any case, proceeding as in Zehnder 8 we can obtain a normal form for the Hamiltonian H£(p,q,es) in a neighborhood of these KAM tori: G{I, a, if, s, zB, z u ) = w/ + a + ^ / 2 + {z\ Q{
(15)
Such normal forms are useful later in the study of A-lemmas in subsection 2.7. We also call attention to a more general normal form due to Fontich and Martin 9 which does not require that the motion on the tori satisfies Diophantine conditions and which is also the basis of the proof of a A-lemma. R e m a r k 9 Note that KAM tori produced by Theorem 7 are of codimension 1 inside A£. If we choose a submanifold whose boundary consist on two KAM tori, this submanifold will be an invariant manifold for the extended flow. In this case the results of hyperbolicity perturbation theory can be extended to include uniqueness, as observed before. Once we have the existence of the invariant tori of system (10) it is worth while to obtain some explicit approximations in the phases {
(17)
is a I-periodic in ((p,r) function which verifies
wD1g(
+ Oc-"-«(£3),
with U given in (16), and \\g(-, -;£)||Cr-™-8 is bounded uniformly in e.
(18)
102
Furthermore, we can choose g in such a way that g = D2J1. This h satisfies (obviously)
and
uDihfa
T; e) + eD2~h{
H;-\e)
Qr — m — S
fc3).
(19)
is bounded uniformly in e.
We call attention to the fact that, due to the slow dynamics, the functions g(ip, r;e) and h(
The perturbed outer map. Poincare-Melnikov theory
Due to the fact that A£ is a 3-dimensional hyperbolic manifold and that its 4-dimensional stable and unstable manifolds intersect in a 3-dimensional "homoclinic" manifold % close to 7 = 7 x ^T, we know that any point on A£ is connected to another point through some orbit, and then it has sense to define the perturbed outer map. On the other hand, we know that A£ contains other invariant submanifolds that are the tori produced by Theorem 7. Then, this homoclinic manifold 7 £ will contain, among others, homoclinic trajectories to each of these torus as well as heteroclinic trajectories between some of them (those that axe close enough). The fact that x+ = S ( x - ) for two points x± € A£ is, by definition, equivalent to the existence of a point z 6 T*T 2 x AT1, such that the trajectory %(t) starting on it converges for t —* ±00 exponentially fast to the trajectories A*(£) started in x± respectively, i.e. A^O) = x±, 7 e (0) = z, and dist(A £ t (*),7 £ (t))
as
t -» ±00.
(20)
We want to give criteria to decide whether x + and x_ belong to diflFerent tori given by the inner map and then 7 £ (t) is a heteroclinic connection between these tori. As the points x± are in the perturbed manifold A£, they can be char acterized by the angles s,tp and the value of the perturbed Hamiltonian Hc. Then, first we will compute x + = 5(x_) in these local coordinates. Prom (7) we already know that
103
Figure 1. Illustration of the perturbed tori and the outer map: y lies in the torus of £ _ , but with the angle coordinates of x+ = S ( x _ ) .
computing the increment of the energy, which in the unperturbed system is preserved. Since the energy He is not constant in the perturbed tori, it is not enough to compare Hc(i+) with He(x-) to decide whether x + = S(x_) and x + lie, or not, in the same torus. Instead, we will look for the point y which lies in the torus of x_, but with the angle coordinates of x + = <S(x_), which are (tp_ + A,s) + 0(e 2 ) (see Fig. 1). If x + lies in the same torus as x_, then x+ = 5(x_) = y. Then the function He(S(i-)) — Hc(y), which we will explicitly compute to first order, has to be not zero if the tori have a heteroclinic connection. In general, this function is a measure of how much the homoclinic excursion gains over just staying on the torus. It gives us a quantitative measurement of the width of the heteroclinic connections. We can compute HE(y) using the local expression of the torus that con tains x_, given in Lemma 10 in terms of the energy and angle coordinates. All this is summarized in the following lemma. Lemma 11 Let x+ and x_ two points on A£, such that i+ = 5 ( x _ ) . Let y be a point on A e , belonging to the torus that contains x_, with the phases of
104
x+.
Then H€{x+) - He(y) = e3M(
lim
+ 0(e5)
/
V(T„T 4 )-OO./_ T I
dtIhU(i%(t+-22=),e8 WBV
-g(
+ 0(e2),s),
+ et)
sflEh
+ sT2)
(21)
- sTi)J + 0(£ 5 ), E = J$/2, and where g is
the function given in Lemma 10 verifying (18), associated to this torus. Then, it is clear that, if the function M((fo,£S,E;e), usually called Mel nikov function, has non-degenerate zeros, the image of the torus of frequencies (y/2E, e) by the outer map S intersects itself, or, what is the same, there exists some homoclinic orbit to it. Relatedly, this function will allow us to quantify the width of the jumps in heteroclinic connections. It so happens that the Melnikov function is the gradient of a potential—often called the Melnikov potential— 10 defined by: L(^es,E;e) = (^
hm_£
dtU (,% (t + J L )
-~h(ip0 + a+ + V2ET2,es
,£S
+ eT2)
+
et) (22)
+h(
(23)
105
0(e'm/2+1
Figure 2. Illustration of the action of the map S on a torus r: S(T) intersects transversally tori close to r.
where +T2 C(T)
=
lim
+T3+a+
dt
U(^/2(t),T)
m,T 3 )->oo
dt
U(A1/2(t),;
-Ti+a-
(24)
Proposition 13 Given a metric that satisfies the genericity conditions of Theorem 2, and orbits h-\/2-,1\/2 of the geodesic flow as in the conclusion of Theorem 2, the set of periodic potentials for which the PoincarS function £(T) is identically constant is aCl closed subspace of infinite codimension for I > 0. 2.7
Transition chains and shadowing orbits
In this section we will see that, if the Poincare function £ ( r ) is non-constant, the perturbed invariant tori have heteroclinic intersections if l + m / 2 > 3, that is, if the measure of the gaps between them is smaller than the jumps produced by the homoclinic excursions (see Figure 2). This give us the existence of a so called transition chain. After this, some results on the literature and some
106
point set topology, will give us the existence of orbits that follow an infinite chain of these tori. Lemma 14 Assume that r > 15. / / the function C(T) is not constant, we can find K > 0 such that for e sufficiently small, given a KAM torus T, we can find other KAM tori T+, T~ such that W£fhWf + , W?fhWf_, xoia He{T) > maxH£(T~) + Ke3, maxH e (T) < minH e (T + ) - Ke3. Remark 15 The lemma above does not assert the existence of transverse homoclinic orbits to any of the tori T, T~ and 7" + . The existence of transverse homoclinic orbits is related to the existence of nondegenerate critical points of the Poincare function. We emphasize that, for our purposes, what we need are transverse heteroclinic intersections. Our next goal is to see that the heteroclinic orbits that can be obtained by Lemma 14 can be shadowed by true orbits of the system. As it is usual in the literature for Arnold diffusion, the key step is to find an appropriate inclination lemma (also called sometimes A-lemma). In the literature, one can find very sharp inclination lemmas—including even some estimates of the times needed to do the shadowing—for analytic maps, when the rotation is Diophantine. n - 1 2 - 1 3 . The result that we have found best adapted to our purposes is that of Fontich and Martin 9 for whiskered tori with one dimensional strong (un)stable directions—as is the case in the problem we are considering—, which works for C1 maps and only requires that the torus has an irrational rotation. A particular case of these results 9 is: Lemma 16 Let f be aCl symplectic mapping in a 2(d + 1) symplectic man ifold. Assume that the map leaves invariant a C1 d-dimensional torus T and that the motion on the torus is an irrational rotation. Let F be a d+1 manifold intersecting Wq- transversally. Then
Wf ell'" ■*(r). i>0
An immediate consequence of this is that any finite transition chain can be shadowed by a true orbit. The argument for infinite chains requires some elementary point set topology. Lemma 17 Let { T ^ } ^ be a sequence of transition tori. Given {ei}^ a sequence of strictly positive numbers, we can find a point P and a increasing
107
sequence of numbers Ti such that
*rt(P)e N£.«(*), where N£i(Ti) is a neighborhood of size £i of the torus %. From Lemma 14 and Lemma 17, we obtain the following result, which clearly implies Theorem 1. Theorem 18 Assume that the metric g satisfies the assumptions of Theo rem 2 and that U is such that the Poincai function C is not constant. Assume moreover that both g and U are C15. Then, there exist M > 0, a > 0, such that if Ii = [Bi_,Ei+]
i = l,...
is any sequence of intervals such that EL
>M,
{E\-Ei)>M(E\)-a. Then, we can find an orbit p{t),q(t) of the Hamiltonian flow and an in creasing sequence of times t\ < t
and (p(ti), q(U)) is in a neighborhood of size M(EX_)~2 of the periodic orbit Remark 19 By assuming more differentiability in the hypothesis of the the orem, we can get a to be arbitrarily large. This is because, if the system is sufficiently differentiable, we can perform more steps of averaging theory so that the KAM tori are much closer. Remark 20 Observe that the existence of the symbolic dynamics above es tablishes that the set of orbits that have unbounded energy is uncountable. We have uncountably many choices of energy ranges to hit. By going over the argument with somewhat more care we can see that there are different orbits that realize these choices. 3
Some final remarks
We conclude with some final questions and remarks about two problems that have been asked us and that are relevant from the physical interpretation of the problem. The most important question is whether the mechanism presented here is the only one or even the dominant one. All the questions that we present
108 here have a version for the orbits constructed here and for the set of orbits whose energy grows to infinity. Q l : What is the measure of the set of orbits whose energy goes to infinity (resp. those constructed here)? We know nothing about this question except to point out that using the sym bolic dynamics, the set of orbits constructed here is uncountable. Presumably, the set of orbits constructed here is of zero measure. Q2: What is the fastest rate of growth of the energy one can achieve? The following upper bound is elementary. Proposition 21 If q{t) is a solution of the Hamiltonian flow of H(p,q,t) \9q(P,P) + U{q,t), then
=
\H(p(t),q(t))\
109 It seems to be unknown whether this is really necessary or whether it is just an artifact of the method of proof. We point out that, in a very similar problem, Chirikov conjectured in a related case that indeed one would not need to increase the time spent near the tori for a long shadowing chain. Hence, if the Chirikov conjecture were true for our case (it seems to us that in our case the conjecture is more plausible than in the case originally stated for), then we could get a rate of escape that agrees with the upper bound up to a finite constant. Q 3 : Is there a quantum mechanical analogue of the phenomena discovered here? Of course, we consider as the quantum mechanical model the Hamiltonian given by the Laplacian and a time periodic potential. We note that an analogue of Proposition 21 can be proved in quantum mechanics, from
! < * | f f | ¥ ) = ( ¥ dv * dt
One is tempted to argue that the construction presented in the previous sections is too delicate to survive the quantum mechanical effects of uncer tainty etc. We would like to decide whether there are analogues of the existence of orbits whose energy goes to infinity and also an analogue of the symbolic dynamics result. Note that in quantum mechanics, as soon as we construct one solution whose energy grows to infinity, we have an open and dense set of solutions whose energy grows to infinity.
Acknowledgements We thank J.N. Mather for communicating his results and for encouragement. This work has been partially supported by the NATO grant CRG950273. Research by A.D. and T.M.S. is also supported by the Spanish grant DGICYT PB94-0215, the Catalan grant CIRIT 1996SGR-00105, and the INTAS project 97-10771. Research by R.L. is also supported by NSF grants. We also thank TICAM, UPC and IMA for invitations that made possible these collaborations.
110
References 1. A. Delshams, R. de la Llave and T.M. Seara, A geometric approach to the existence of orbits with unbounded energy in generic periodic pertur bations by a potential of generic geodesic flows of T 2 , preprint 98-591, mp_arcOmath.utexas.edu,(1998). 2. J.N. Mather, Graduate course at Princeton, 95-96, and Lectures at Penn State, Spring 96, Paris, Summer 96, Austin, Fall 96. 3. L. Chierchia and G. Gallavotti, Drift and diffusion in phase space, Ann. Inst. H. Poincare" Phys. Theor., 60, 1-144, (1994). 4. S. Bolotin and D. Treschev, Unbounded growth of energy in nonautonomous Hamiltonian systems, Nonlinearity, 12, 365-388, (1999). 5. V.I. Arnold, Instability of dynamical systems with several degrees of freedom, Sov. Math. Doklady, 5, 581-585, (1964). 6. V.I. Arnold and A. Avez, Ergodic problems of classical mechanics, Ben jamin, New York, (1967). 7. V.I. Arnold, V.V. Kozlov and A.I. Neishtadt, Dynamical Systems III, volume 3 of Encyclopaedia Math. Sci. Springer, Berlin, (1988). 8. E. Zehnder, Generalized implicit function theorems with applications to some small divisor problems, I and II, Comm. Pure Appl. Math., 28, 91-140, (1975), and 29, 49-111, (1976). 9. E. Fontich and P. Martin, Arnold diffusion in perturbations of analytic integrable hamiltonian systems, preprint 98-319, mp_arc<8 math.utexas.edu, (1998). 10. A. Delshams and R. Ramirez-Ros, Melnikov potential for exact symplectic maps, Comm. Math. Phys., 190, 213-245, (1997). 11. J.P. Marco, Transition le long des chaines de tores invariants pour les systemes hamiltoniens analytiques, Ann. Inst. H. Poincare Phys. Thior., 64, 205-252, (1996). 12. Jacky Cresson, A A-lemma for partially hyperbolic tori and the obstruc tion property, Lett. Math. Phys., 42, 363-377, (1997). 13. Enrico Valdinoci, Whiskered transistion tori for a priori stable and un stable hamiltonian systems, preprint 98-200, mp_arcfflmath.utexas.edu, (1998).
SPLITTING A N D MELNIKOV P O T E N T I A L S I N HAMILTONIAN SYSTEMS AMADEU DELSHAMS AND PERE GUTIERREZ Departament
de Matematica Aplicada I, Universitat Politecnica de Diagonal 647, 08028 Barcelona E-mail: [email protected], [email protected]
Catalunya,
We consider a perturbation of an integrable Hamiltonian system, possessing hyper bolic invariant tori with coincident whiskers. Following an idea due to Eliasson, we introduce a splitting potential whose gradient gives the splitting distance between the perturbed stable and unstable whiskers. The homoclinic orbits t o the perturbed whiskered tori are the critical points of the splitting potential, and therefore their existence is ensured in both the regular (or strongly hyperbolic, or a-priori unsta ble) and the singular (or weakly hyperbolic, or a-priori stable) case. The singular case is a model of a nearly-integrable Hamiltonian near a single resonance. In the regular case, the Melnikov potential is a first order approximation of the splitting potential, and the standard Melnikov (vector) function is simply the gradient of the Melnikov potential. Non-degenerate critical points of the Melnikov potential give rise to transverse homoclinic orbits. Explicit computations are carried out for some examples.
1
Introduction
For more than 2 degrees of freedom, the problem of measuring the splitting of the whiskers of hyperbolic invariant tori is closely related with the existence of instability in nearly-integrable Hamiltonian systems, i.e. with the Arnold diffusion. In this lecture, the splitting is studied in a wide setting, and a general Poincare-Melnikov theory is developed.
Setup We start with a perturbation of a hyperbolic integrable Hamiltonian, with n + 1 > 3 degrees of freedom. In canonical variables z = (x, y, ip, I) £ D C T x R x T" x R n , with the symplectic form dx A dy + dip A dl, consider a Hamiltonian of the form H(x,y,ip,I;n)
= H0{x,y,J)
+ /i#i(x,y,v?,I),
H0(x, y, I) = <«, I) + 1 (A/, /> + ^ + V(x) + (A, /) y,
111
(1) (2)
112
where p is a perturbation parameter. The Hamiltonian equations associated to H are: ± = y+(X,I)
+
y = -V'{x)
- fidxHi (x, y, tp, I),
(p = uJ + AI + Xy 1=
iidyHi(x,y,
It will shown in section 2 that, under weak assumptions, the unperturbed Hamiltonian Ho has n-dimensional whiskered tori (hyperbolic invariant tori) with coincident (n 4- l)-dimensional whiskers (invariant manifolds). For a given whiskered torus of Ho, its (unique) whisker is filled by homoclinic orbits (biasymptotic to the torus). Our aim is to study the splitting of the whiskers, and the persistence of some homoclinic orbits, for fj, / 0. Main achievements To deal with this problem, the tools used are Poincare-Melnikov theory, and a geometric method based on Eliasson's approach. Our contributions can be summarized as follows: • A general Poincare^-Melnikov theory for Hamiltonian systems is devel oped, defining a scalar function L (Melnikov potential) whose gradient M (Melnikov function) gives the splitting distance at first order in (i. • There exists a scalar function C (splitting potential) such that, in suitable variables, its gradient M. (splitting function) gives exactly the splitting distance. Besides, the splitting potential £ is approximated at first order in n by the Melnikov potential L. • The results are significant for more than 2 degrees of freedom. Motivation The study of the splitting in the Hamiltonian (1-2) is closely related to the problem of Arnold diffusion in a general nearly-integrable Hamiltonian system: H(4>,J) = h(J) + ef(4>,J),
(3)
in angle-action variables (4>, J) £ T n + 1 x R n + 1 . Here, the small perturbation parameter is e. Near single resonances, it is known l i 2 ' 3 that one step of (resonant) normal form procedure can be performed and leads, under some generic hypotheses
113
and after a scaling, to a Hamiltonian of the type (1-2), taking as HQ the truncated normal form. To make this clearer, consider a selected action J* = 0, and assume that its associated frequency vector djh(0) G R n + 1 has a single resonance (this means (k\djh(0)) = 0 for a certain k' G Z n + 1 \ {0} and (k,djh(Q)) ^ 0 for n+1 any k G Z not co-linear to k*). It can be assumed that djh(0) = (0,w*), with u* G R n nonresonant. Near J*, the unperturbed Hamiltonian h in (3) can be written as:
h(J) =-- (djh(o),j) + i (djh(o)j,j)
+ o3{j).
We write
=cr
d]h(o) =
>!)■
where we have put /3 2 > 0 in order to fix ideas, A G R n , and A is an (n x n)matrix. With some scaling, we can assume 0 — 1, and our Hamiltonian written in the form H (x, y, (p, I) = h(y, I) + ef(x, y,
= + i (A/,/) + ^ - + e V ( i ) + (A,/)y,
if, (x, y, p, 7; e) = £i?(x, y,
f
f(x, 0, y?, 0)dV>
x G T.
In the normalized expression for H, note that Ho (the truncated normal form) is an integrable Hamiltonian, and then Hi can be considered as a perturbation of some size E\L where \i can be determined in terms of e. In this sense, the expression obtained generalizes the Lochak's example 4 (which, in its turn, generalizes the famous Arnold's example, 5 designed to describe the diffusion). Under generic hypotheses, it can be shown that the Hamiltonian Ho has whiskered tori with coincident whiskers associated to this hyperbolic point (see
114
section 2). Therefore, although there is no hyperbolicity in h, the perturbation / provides some weak hyperbolicity, which appears in the truncated normal form HQ- This hyperbolicity disappears for e —► 0, because the Lyapunov exponents of the whiskered tori of HQ are of the form ±i/ect. To have fixed exponents, we replace y, I by \/ey, \fil (a non-canonical linear change), and divide the Hamiltonian by e. Then the new system is still Hamiltonian, and we obtain obtain for H = Ho + Hi an expression of the form (1-2), with
-$•
" = °(£"2)'
It has to be pointed out that, after this procedure, in general the truncated normal form HQ is a coupled Hamiltonian: A ^ 0 in (2). So the motivation for the coupling term {X,I)y is that this term appears in a natural way when one studies a nearly-integrable Hamiltonian, in a region close to a single res onance. As a particular case, note that if A = 0 in (2), then the unperturbed Hamiltonian HQ is somewhat simpler because it is formed by a pendulum and n rotors: we then say that Ho is uncoupled. We will show in section 4 that the formulation of Poincare-Melnikov theory is simpler in this special case. Although the (homoclinic) splitting between the whiskers of hyperbolic tori in single resonances is very important in the detection of Arnold diffu sion (through the construction of transition chains), we point out that there are other important difficulties related with this problem. These difficulties are the study of the transition properties of the tori, the detection of heteroclinic intersections between whiskers of different tori, and jumping the gaps associated to double resonances. Regular and singular cases According to the motivation above, it is convenient in (1-2) to allow u to depend on an additional parameter e, considering fast frequencies LJ = u*/y/e. The parameters e and (i can be whether independent or linked by a relation of the type \i = ep with some p > 0; these two cases will be called, respectively, regular and singular. We have shown that, in the study of a general nearlyintegrable Hamiltonian, the actually relevant case is the singular one (with p — 1/2), and that this feature is directly related to the weak hyperbolicity of the truncated normal form. Concerning the regular situation, we recall that the strategy of keeping e > 0 fixed and letting pi —► 0 (having in this way a regular system) was introduced by Arnold 5 in order to avoid dealing with a singular perturbation problem. In this case, Poincare-Melnikov theory can be applied directly to the
115 detection of the splitting, but only if the parameter fi is taken exponentially small with respect to e. This is due to that the Melnikov integrals involved are exponentially small in e, as in the second example shown in section 5 (for the first example shown, the integrals are not exponentially small, because the perturbation is not analytic in this case). In the singular case, one assumes that the parameters e and fi satisfy a power-like relation of the type /i = e9 (the smaller p the better), and one lets e —> 0. In this case, the problem of detecting the splitting from the Melnikov integrals is much more intricate, because of the exponentially small character of the integrals involved. However, some recent works 6 ' 7 , 8 suggest that, under some weak conditions, the Melnikov integrals give the right predictions for the splitting. Nevertheless, the existence of homoclinic orbits has been established in several works. 1 - 9 ' 10 This result is valid for regular and singular systems, and we recall it in section 6. 2
T h e u n p e r t u r b e d Hamiltonian
Assumptions In this section, we take \L = 0 and study the unperturbed Hamiltonian H0 denned in (2). Note that the given ingredients of Ho are the vectors o>, A e R n , the symmetric (n x n)-matrix A, and the function V(x) of x £ T. We require the following assumptions: • The function V(x) has a unique and nondegenerate global maximum. To fix ideas, we require V(0) = 0,
V'(0) = 0,
V(x) < 0 Vx ^ 0
V"(0)<0,
(mod 2TT).
(4)
• The following nondegeneracy condition holds: det (A - AAT) = det ( * A ^ ^ 0.
(5)
• The vector w is assumed to satisfy a Diophantine condition: for some T > n — 1 and 7 > 0,
|(fc,w)|>7|itrT
VfceZn\{o}.
(6)
116
The unperturbed torus and its homoclinic whisker The integrable Hamiltonian Ho can easily be studied. Let us introduce P(x,y)
= £+V{x),
P(x,y,/) = P(x,y+(A,J));
(7)
then H0 can be rewritten as
H0 = (u,I) + ±((A-MT)l,l)
+
P(x,y,I).
We see that, on every plane / = const, the Hamiltonian Ho reduces to a 1degree-of-freedom Hamiltonian: a generalized pendulum (the standard pendu lum being given by V(x) = c o s x - 1 ) . This pendulum has (x,y) = (0, - (A,/)) as a hyperbolic equilibrium point, with (homoclinic) separatrices given by y + (A, J) = ±1/—2V(x). The Lyapunov exponents of the hyperbolic point are ± a , where we define a = y/-V"(0). Therefore, the Hamiltonian Ho possesses an n-parameter family of ndimensional whiskered tori given by the equations / = const, y = — (A, / ) , x = 0, with (n+l)-dimensional whiskers. The stable and unstable whiskers of each torus coincide, and hence all orbits on this (unique) whisker are homoclinic, i.e. biasymptotic to the torus. We will focus our attention on a concrete hyperbolic torus, that we assume located at the origin: / — 0, x = y = 0. Note that the vector u>, assumed Diophantine, consists of the frequencies of this torus: (p = u>. In view of the nondegeneracy condition (5), the neighboring tori have different frequencies. Parametrizations for the unperturbed Hamiltonian We denote % the whiskered torus of Ho having frequency vector UJ. This torus can obviously be parameterized by To-
z'o{
As mentioned above, the stable and unstable whiskers of the torus To coincide; this homoclinic whisker is given by the equations I — 0, P(x,y) = 0. We denote Wo the positive part (y > 0) of the homoclinic whisker (it is often called separatrix). To give a suitable parametrization for Wo, we consider the 1-degree-of-freedom Hamiltonian P(x,y), and denote (xo(s),yo(s)) the associated homoclinic trajectory, with xo(0) = n, yo(0) > 0. Note that Xo(s) goes from 0 to 2TT when s goes from —oo to oo. It is clear that we can give the whisker Wo the parametrization W0:
z0(s,V5) = (xo(s),y 0 (s), ¥ ) + (xo(s)-7r)A,0),
s € R,
117
where the term (io(s) — TT)A expresses the phase drift undergone by any tra jectory when traveling along Wo- This drift is associated to the coupling term. Note that, with our definition, the dynamics on Wo is given by the equations s = 1, ip = u. One has lim [ZQ(S + t,tp + wt) — ZQ(IP±7rA + ut)] = 0, t—>ioo
and this implies that that every trajectory on Wo is biasymptotic to two different trajectories on the invariant torus %. If A is an integer (a very special case) then these two trajectories on % coincide. 3
Preservation of the whiskered torus and its whiskers
The local normal form Before studying the splitting, we have to establish the surviving under per turbations of our Diophantine whiskered torus, as well as its local whiskers. Then we have to extend them to global whiskers in order to compare the sta ble and the unstable ones. The surviving of the torus and its local whiskers under a small perturbation can be ensured by means of the hyperbolic KAM theorem, a version of the KAM theorem adapted to this problem. Roughly speaking, the hyperbolic KAM theorem provides a symplectic transformation $ taking our Hamiltonian into a local normal form H = H o $ (in some domain), having a simpler expression in which the perturbed torus becomes transparent, as well as its whiskers. This kind of result follows from a convergent KAM-like iterative scheme. We are interested in a normal form defined in a whole neighborhood of our concrete torus, 1 , n according to the "Kolmogorov's approach" to KAM theory. This approach allows us to control a neighborhood of the local stable whisker, which can be ensured in this way to contain also a piece of the global stable whisker (this feature is used in section 6). On the contrary, in the "Arnold's approach" (used in other papers) the normal form only holds on a Cantor set, although a large family of surviving tori is obtained. Some more comments and references to papers following both approaches are given in a recent paper of the authors. 10 In most papers (like for instance n ) , the hyperbolic KAM theorem is dealt in terms of some local variables in a neighborhood of the torus, in such a way that the whiskers become coordinate planes. A significantly new approach was introduced by Eliasson, who rewrote the hyperbolic KAM theorem and expressed it directly in the "original variables". l This is more suitable to our
118
purpose of carrying out a global control of the whiskers in order to study their splitting (see section 6). Another key fact is the use of exact symplectic transformations to normal form in the hyperbolic KAM theorem. To recall what an exact symplectic transformation is, consider the l-form 77 = -(ydx + Idip), whose differential is the standard symplectic 2-form: dr? = dxAdy+dipAdl. Then a transformation $ is symplectic if the l-form <1?*77 — 77 is closed, and it is exact symplectic if this l-form is exact (= dS, globally, for some scalar primitive S). Eliasson J used the exactness of the normalizing transformation as a cru cial tool in order to detect homoclinic intersections between the whiskers, in both regular and singular systems (although he did not compute the split ting). A similar result was also obtained by Bolotin. 9 In a further step, in the present lecture the exactness allows us to put the splitting function as the gradient of a splitting potential (see section 6). Another paper that has influenced our version of Eliasson's theorem is a recent one by Niederman. n This paper deals with a similar framework (using the Kolmogorov's approach but not working in the original variables), and obtains more accurate estimates for the normal form. Let us introduce first some notations. Concerning the domain, we define for r > 0 the complex set Br = {(x,y, if, I) : |x|, |y|, | / | , |Im p | < r} . For a function f(x, y, y>, I) analytic on some domain D (and continuous on its closure), we denote | / | D its supremum norm. Theorem 1 (Eliasson's theorem) Let H = HQ + HH\ as described in (12) and in the assumptions (4~6), with r > n — 1. Assume H analytic on Br (r < ro). Then for |/J| small enough, there exists an exact symplectic transformation $ = $ ( - ; / J ) : Bur —► Br (analytic with respect to (x,y,
(8)
Besides, one has $ = id + 0(M)> a = 0(M)> b = 1 + O(M)The most important point about this result is that, thanks to the use of the original variables x, 7/ u the local normal form H can be put in terms of the generalized pendulum P(x,y,I). By using this feature, a "global" control of the whiskers, very useful in order to compare them and study the splitting, can be carried out. 1,1D In is not hard 10 to establish the validity of theorem 1 in the singular case, with fi — ep and w = w*/y/e, for |e| small enough.
119
Paratnetrization of the perturbed torus It is clear that the normal form H given in (8) has a whiskered torus of fre quency vector ijj. We denote this torus as T, and its associated local whiskers as W^ (stable) and W,",. (unstable). The torus T has the following obvious parametrization: f:
i »
= (0,-(A,a),^a),
^ T .
This torus can be translated to a whiskered torus T of the original perturbed Hamiltonian H:
T:
* » = * ( r M),
In section 4, it will be useful to give a first order approximation in /u for the shift suffered by the perturbed torus T with respect to the unperturbed torus To, along the /-direction. We will denote I*(f) the /-component of z*((f). To describe this approximation, we consider the (zero average) scalar function x{
(9)
where the notation / denotes the (^-average of a function / . The existence of X is ensured by the Diophantine condition (6). The function x, introduced by Treschev, 12 provides a first order approx imation 10 for the perturbed torus: /*(
(io)
where we define C = - (A - AA T ) _ 1 (3/J/i - \dyHx) (0,0,0). Parametrizations of the perturbed whiskers As in section 2, we can also take parameters on the perturbed local whiskers of the normal form H: W£ c :
z{s,(f) = (xo{bs),y0{bs) - (A,a) ,
for ±s > so, > € T n , with some SQ — so(r). For the original Hamiltonian H, the local whiskers can be parametrized as follows: W,oC:
*toc(*,¥>)=*(5(*,V>)),
±s>s0,
f€Tn.
In the parameters s, tp, the dynamics of H on W^c is given by s = 1, ip = u). We need to extend these local whiskers to global whiskers, in order to measure the splitting between them. The parametrizations of the whiskers
120
Wioc> vaJid for ±s > SQ, can easily be extended to further values of s in a natural way, since the whiskers are formed by trajectories associated to our Hamiltonian H. We denote W * the extended or global whiskers. These global whiskers remain at distance 0 ( M ) t o t n e unperturbed whisker Wo, for an interval of real values of the parameter s. This interval can be chosen large enough in order to make it possible to compare the whiskers W + and W~ far from the torus T. 4
Poincar6—Melnikov theory
In this section, we develop Poincare-Melnikov theory in order to give a first order approximation for the splitting of the separatrix Wo into the perturbed whiskers W * associated to the perturbed torus T . Besides, we want to de scribe the set W + n W + , i.e. the homoclinic orbits to T. Melnikov potential and Melnikov function In order to provide a first order approximation for the splitting, we introduce the (scalar) Melnikov potential L(ip) and its gradient, the Melnikov function M(ip) = d^Lfa), by means of improper integrals, /^-independent and periodic in i/3 € F . These integrals are always absolutely convergent, thanks to the fact that the phase drift along the separatrix (due to the coupling term in (2)) and the first order deformation of the perturbed hyperbolic tori are taken into account. We stress that our use of absolutely convergent integrals in the formu lation of the general Poincare-Melnikov theory for whiskered tori makes a difference with respect to some previous works, 13>14>15 where conditionally convergent integrals are used and the integration limits have to be carefully chosen. Next we define these functions in several cases, in increasing order of complexity. • The simplest case is that of a perturbation vanishing on the whiskered torus, Hi = 02{x,y,I). In this case, the whiskered torus remains un changed (T = To). We define the Melnikov potential through the follow ing integral: oo
(Hi-H^){zo{tif
/
+ ut))dt
■oo / o o
Hx(z0{t,cp + ut))dt + const. •oo
(11)
121
Note that the additive constant is such that L = 0. For the Melnikov function, it is clear that oo
/
dvHi(zo{t,
ut))dt.
(12)
-oo
The absolute convergence of the integral (12) for the Melnikov function was already pointed out by Robinson, 15 stressing that in other cases this integral is only conditionally convergent. We also recall that, for n = 1, the Melnikov potential (11) coincides with a formula given by Delshams and Ramirez-Ros. 16 • Now, we consider the uncoupled case A = 0 (note that this case intersects but does not include the previous one). In this case, we define the Melnikov potential through the following integral, also absolutely convergent: r°° . L(
/
[Hi (zo(t,
• Finally, in the general case (which includes the two previous ones), we define (tfi - ST - {*, Ho}) (zd(«, V + ut)) dr..
(13)
■oo
Recall that H\(x,y,I) denotes the >average of H\, and that the function xi.xi 2/> f< -0 — xif) is t n e (zero average) function solving the small divisors equation (9). Notice that L = 0, because the function inside the integral has zero average. The absolute convergence of the Melnikov integral (13) can be ensured using that the function H\ — H\ — {x, Ho} vanishes on To, together with the fact that Wo tends to To with exponentially decreasing bounds. We remark that the formula (13) is useful in both the coupled and the uncoupled cases (in (2), A ^ 0 and A = 0 respectively). An example illustrating the uncoupled case was given by the authors. 10 Related expressions, also valid in both cases, were previously obtained by Treschev. 12 In that paper, although the Melnikov potential was not introduced, the Melnikov function was expressed with the help of some correcting terms giving rise to the absolute convergence. We have improved that expression, including the correcting terms in the integral and providing a more compact formula.
122
First order approximation for the splitting distance The following standard result 1 0 shows that a first order approximation for the splitting between the global whiskers W±, measured along the /-direction, is given in terms of the Melnikov function M = dvL. Since both whiskers are (n+l)-dimensional manifolds contained in the same (2n+l)-dimensional level of energy, it is enough to express its distance by an rc-dimensional measure. We take the difference I~ — I+ as the measure for the splitting (we denote 7 ± (s,^)) the /-component of the parametrizations z±(s,(p)). Theorem 2 Assuming |/x| small enough, one has for any \s\ < so and
(14)
We stress that the Hamiltonian character of the equations implies the fact that the first order approximation of the splitting is simply the gradient of a scalar function L. An important fact in this theorem is that I\,..., In are first integrals of HQ. Following Treschev, 12 it is not difficult to generalize theorem 2 by considering any given first integral F of the unperturbed Hamiltonian HoThus, it can be given an analogous first order approximation for F~(s,
ip) - ?+{s,
(n2),
where we define L(s,(p) = L(tp - us), which can be considered a function defined on the separatrix Wo- Using this fact^an alternative measure for the splitting could be J f - JJ 1 ",..., I~_1 — l£_1,P~ — P+ (if w„ ^ 0), instead of
/;--/+,...,/--/+. Transverse homoclinic orbits As a simple corollary of theorem 2, we see that in the regular case the simple zeros of the Melnikov function M give rise, for |/i| small enough, to transverse homoclinic intersections between the perturbed whiskers. As is well-known, if a point belongs to the homoclinic intersection, then its whole orbit is also contained in this intersection. Thus, it is enough to find the zeros of M(
123
(defined o n F ) is a Morse function (its critical points are all nondegenerate: a generic property), we deduce from Morse theory that for |/i| small enough there exist at least 2 n transverse homoclinic orbits. It is well-known that this argument does not apply in the singular case, w = w'/i/I and /x = e p , because the Melnikov function M is typically ex ponentially small in e (see the second example in section 5). To ensure that /j.M(tp — uis) dominates the O (/i 2 )-term, one has to assume \x exponentially small with respect to e. For larger values of fi, the existence of intersections cannot follow directly from (14). In fact, the study of the splitting in the singular case requires a more careful analysis, 6 ' 7 ' 8 which is not carried out here. Nevertheless, the effective existence of a number of homoclinic intersections, for both the regular and singular cases, will be established in section 6. 5
Some examples with small divisors
To illustrate the properties of the Melnikov potential L, in this section we consider some examples, showing that L has nondegenerate critical points. As a measure of the transversality of these points as zeros of M = dvL, we also estimate the determinant of the symmetric matrix dvM = d2L at the critical points. Consider the Hamiltonian H = H0 + \xH\, with 1 v2 HQ(x, y, I) = (w, /> + - (A/, /) + y + cosx - 1, Hx{x,
f{
fke*^>.
The integrable Hamiltonian HQ is uncoupled (A = 0 in (2)), and consists of a pendulum and n rotors (the standard pendulum is given by V(x) — cos x — 1). Note that the perturbation H\ depends only on the angles x,
(15)
then this is an analytic function, and p is its width of analyticity in the angles if. On the contrary, if the coefficients are polynomially decreasing:
IM ~ n^.
( 16 )
124
then the function f(tp) is not analytic but only differentiable (it is Cp for any p
2 yo(t) — ±o(t) = — — .
We have: (cosxo(t) - l ) / ( y + u / t ) d t + c o n s t = 2 / -oo
J-oo
J
^
2
'dt+const.
COsh t
Taking into account that the additive constant is such that L = 0, and writing L(f) = Z)fc=4o Lkex(k'lfi\ the Fourier coefficients Lk can be computed explicitly using residue theory: J-oo coslr t
sinh (f (k,Lj))
(note that the mean value / = /o does not influence the Melnikov potential). For the Melnikov function, it is clear that M^ = ikLkUpper bounds In the analytic case (15), an upper bound for the Melnikov potential L can be given 10 from the expressions of the coefficients. The upper bound obtained holds for e > 0 small enough, and is exponentially small in e. Its size depends strongly on the small divisors properties of the frequencies:
^l^)l^^)-P(-c£-/(-+2)), where C = C(T, 7*, p) is a constant. It is an important point in this estimate to assume a perturbation with an infinite number of harmonics. As stressed by Lochak, 4 one is then forced to take into account the small divisors as sociated to the frequencies, and this leads to the exponent l / ( 2 r + 2) inside the exponential. Notice that this exponent in the upper bound is reminiscent
125
of the Nekhoroshev-like estimates. Instead, if one assumes a finite number of harmonics (like in the Arnold's example 5 ), then one obtains the exponent 1/2, but this case is highly nongeneric. In an analogous way, we can obtain an upper bound for the differentiable case (16), but then the bound becomes a power of e: max \LUp)\ < const • e r / 2 T . "" v eT"
(18)
The golden mean and the Fibonacci numbers To establish the effective existence of splitting, one has to obtain also lower bounds, giving a more precise description of the asymptotic behavior of the Melnikov potential. This requires a more careful analysis of the small divisors associated to the frequency vector w*. This analysis can easily be carried out for the golden mean, a very simple case with 2 frequencies (i.e. with 3 degrees of freedom):
«* = (i,n),
^^y11-
(is)
This case was first considered by Simo 17 and, later on, 6 lower bounds for the Melnikov function and for the splitting, in the analytic case (15), were obtained. Recently, 10 lower bounds have been obtained for the Melnikov potential L and for the determinant of d£L at the critical points (ensuring also that L has nondegenerate critical points). These lower bounds (recalled in the second example below) are exponentially small with respect to e. This implies that, in order to deduce the existence of splitting as a consequence of theorem 2, the parameter \i has to be taken exponentially small in e. The differentiable case (16) is substantially different. A concrete exam ple 18 shows that the maximum of the Melnikov function has a lower bound of finite order in £. Then taking \x as a suitable power of e is enough in order to establish the existence of splitting from theorem 2. We stress that an essential point in dealing with the singular case is to assume that, in the perturbation, at least the harmonics fk corresponding to the small divisors associated to u* are nonvanishing, because the dominant harmonic is found among these ones. Under this assumption, one can obtain 6 the largest lower bounds in the Melnikov approximation, in order to ensure that this approximation dominates the O (M2)-remainder. For the golden mean (19), the associated small divisors are directly related to the Fibonacci numbers: F0 = F1 = 1,
Fn = F n _i + F n _ 2 ,
n > 2.
126
We recall some basic facts concerning these numbers, that will be used below. Defining CF =
1
i
n + fi-1 " > / § '
we have Fn = CF ( f i n + 1 _ (-i)»+in-<«+»>),
n>0.
The best rational approximations of fi are given by the convergents Fn/Fn-\. In other words, the indexes k^ — (Fn,—Fn_i) (and also (—F n ,F n -\)) are the ones that give the dominant behavior among the small divisors (k,u*). More precisely, one has:
<*v>-*-*-.»- ^ - ^ ° ( £ ) , and also the following inequality: not a Fibonacci number, \{k,u')\
6
n > 1,
for any k = (fci, — k2) such that k2 > 0 is
= \k1-k2n\>^.
(20)
Note that the frequency vector (19) satisfies the Diophantine condition (6) with T = 1. This frequency vector is considered in the two examples that we next study. Lower bounds: An example with finite-order splitting Now we consider a concrete example in the differentiable case (16), analogous to the one of Delshams et al., 18 and obtain lower bounds for the maximum of the Melnikov potential L, and for the determinant of d%L at a critical point. For the perturbation, we consider the following function:
n>l
In this function, the only nonvanishing Fourier coefficients are the ones as sociated to the Fibonacci indexes k^n\ Since |fc(")| = Fn+i ~ fin+2, the coefficients decrease as in (16). Applying (17), the Melnikov potential L(ip) is given by the series
i(y) = E 5 « cos ( A;(n) ^)> n>l
127 with
sn
1
2TT
yTe nn(r+i ) 8 i n h (_a 7 .y
Note that all the coefficients are positive. The main part of this expression is given by CO
_
_bo
47T :
P
n
ffi — ffl(r} — n(r 1 n i n i r O 1 "" °n °n(c) — nV i i J l O g " 1 „Qn /7'
To determine the dominant behavior, we look for the minimum exponent 6°, for n > 1. This is reached for No — No(e), with ■K
nNn
2(r + 1 ) ^ ' and this gives the largest coefficient: SN0 > 5^ o ~ £ r / 2 . This coefficient itself constitutes a lower bound for the maximum of the Melnikov potential, because all the coefficients are positive and
max |L( V )| = 1(0) =
(21) n>l
We can also get an upper bound, which coincides with the one of (18). Let us break the series (21) in two parts. For n < No, note that
sS sS-i
=
<
7T
\
fir+ie*P|\2£P+lJe)
n ^ l CXP( \ '- Qr+i l 2 0 " ° + ^ ) "■ ( - 5 - )
> L
Using also Sn < §S£ (from the fact that sinhx > e x /3 for x > 1), the sum 12n
max I^Mi ~ er/2. ve'r-1 The Melnikov potential L(ip) has
128 eigenvalues of 3 2 L(0) as a measure for the transversality. We have 3^(0) =
-£sn^)(jfcW)T, n>l
and then det^L(O) =
(Z^n)
V«>i — 2_,
fee^y w y
-
/ FnFm-\(FnFm-\
_
(X^nFn-xSn) \n>l
-fn-1
/
Fm)SnSm
n,m>l
(Wrn-l — Fn-iFm)
= Y,
SnSm — /__,
Fn_m_lSnSm,
l<m
l<m
where we have used the formula FnFm-i -Fn_xFm = ( - l ) m + 1 F n _ m _ i . Note that all terms in this series are also positive. To estimate the determinant, note that Fn_m_lSnSm
~ ft n 5 n •
ft
m
Sm,
and hence the indexes n and m can be separated. This allows us to find the indexes N\{e) and Mi(e) that give the dominant term in the series of the determinant, in the same way as before. In this way, we easily obtain an upper bound and a lower bound for the determinant: det^L(0)~er. In fact, we should estimate the minimum eigenvalue of d£L(Q). This eigenvalue can be put in terms of r = trdjL(O) and 6 = detd£L(0). Again, note that r ~ £ r / 2 (applying the same method). The minimum eigenvalue is given by T -- v V
2
2
- -46 . * r
and this has clearly a lower bound of order e r / 2 . Then it is a consequence of theorem 2 that, for (j. = o (e r / ' 2 ), the critical point
129
Lower bounds: A singular example For the sake of completeness, we also include an example 10 in the analytic case (16). For the perturbation, we consider a "full" Fourier series, with the coefficients Vfc€Z2\{0}.
\fk\=e-W<>
Note that a non-even function f(tp) is allowed, so we are not assuming that the perturbation H\(x,y) is reversible, unlike other papers. 19 ' 7 - 8 Following the method by Delshams et al. 6 (though the context is some what different), it is shown 10 that the dominant harmonics in the Fourier series of the Melnikov potential L(
2ir
(,—fn+ip
n > 1.
° ^ **(«!*)'
The main part of this expression is given by ,„n,
47T
:e-»°,
£ = £(£) = C
F
n^p+_JL_.
(22)
For a fixed e > 0, to find the dominant harmonic among the Fibonacci ones, one has to look for the minimum exponent 6°, n > 1. Let us define
" n ~ U n ° V ~ n4°B'
Do_
2CFp V2C/
The minimum exponent among the 6° is reached for an only integer NQ = No(e), such that log£Ar0 is the closest to loge, among the loge n . Then the coefficient S%0 is the dominant one among the 5 ° , and it is not hard to check that the "whole" coefficient SN0 is also dominant among the Sn- One can also check from (20) that the non-Fibonacci coefficients Lk, with k ^ k^n\ do not dominate. In terms of e, the value of the minimum exponent depends on e in the following way: _ c(loge) 0 °No ~ £ l/4 ' where C(TJ) is a continuous function, defined as the (4 log Q)-periodic extension of c(n) = C0 cosh C^^-J
.
\V-Va\<2
log Q,
130
with Co = ^ly/2nCpP,
77o = log£o-
The extreme values of this function are given by Q3/2/-Y
C 0 < c(r)) < — — - = (1.029085.. .)C 0 . In this way, the maximum value of the Melnikov potential can be approx imated by its dominant Fibonacci harmonic, and one obtains the following upper and lower bound:
™™\L^)\~^z™v\—7T74-JIt is also shown 10 that the Melnikov potential L(ip) has nondegenerate critical points. In order to detect these points, one has to consider an ap proximation given by at least the 2 dominant harmonics, because with only 1 harmonic the approximation to the matrix d^L would be degenerate. In the discussion above, it can also be considered the integer Ni(e) reaching the "second" minimum among the 6°; this integer satisfies \N\ — NQ\ = 1. Calling TV = N(e) — mm(N0,Ni), it turns out that £^+i < £ < £N, and the Fibonacci coefficients with indexes N and N + 1 give the 2 dominant harmonics in the Fourier expansion of the Melnikov potential. The two dominant harmonics give the main part of the Melnikov potential L(ip). In the trigonometric form, this main part can be written as L(
,
where a^, 07/+1 are some phases. The number of critical points is given by the determinant A „ = det (*<">, A<™)) = F „ _ , F „ + 1 - F* = ( - l ) " + \ which implies that, for e small enough, L(tp) has exactly 4 critical points. At every critical point if*, one has |detajL(v9*)|~4|5w5N+1|^0. Then L((f) is a Morse function, because all its critical points are nondegen erate. Note also that 4 = 2 2 is the minimum number of critical points for a Morse function on T 2 . To estimate the size of the determinant, we use (22) again: 4|<7°oO
I -
647f2
r-(b°„+b°„^)
ho
, ho
_ ci(loge)
131
where c\ (r?) is another (4 log Q)-periodic function, defined from Cfa) = fi3/2C0cosh (j^jk)
,
\r) - Vo\ < 21o g n,
with 7]'0 — log y/eo^T. The extreme values of this function are given by n3/2C0
fi3C0
Thus, one obtains an upper bound and a lower bound for the determinant at the 4 critical points:
de,a^-)~^e X p(-2fl2|£)) Proceeding as in the previous example, one also finds an estimate for the minimum eigenvalue. Since, in this case, 8 <; r, the minimum eigenvalue can be approximated by 6/T. This leads to an estimate of the type 1
ex
/
c2(\oge)\
iiTT P ( — i l T T - J ' where 02(77) = 01(77) — c(r]). This is also a positive periodic function, with
2
-
'w/ - V
2,
Then it is a direct consequence of theorem 2 that there exist 4 transverse homoclinic intersections, for fi = o (exp{— C2(log£)e -1 / 4 }). The estimate obtained gives a measure for the transversality of the splitting. However, it has to be recalled again that this is actually a regular situation, and a justification in the singular case /x = e p , for some p > 0, does not follow directly from theorem 2. 6
Flow-box variables and splitting potential
The aim of this section is to sketch the proof of the result 10 that, using suitable variables, the "whole" splitting distance (and not only its first order approximation) is the gradient of some function, in order to establish the existence of homoclinic orbits even in the singular case.
132
Flow-box variables In order to provide a clearer formulation for the problem of measuring the splitting, it is convenient to introduce new symplectic variables in which the Hamiltonian equations are very simple. The flow-box variables W = (S, E, ip, J) are constructed 10 with the help of the flow associated to the normal form H given in (8), from a suitable Poincare section containing the set S = z ( s i , T n ) C W,QC, with some fixed s^ > so- The new variables are then given by an exact symplectic transformation (x, y, ip, I) = ^ ( 5 , E,ip, J), defined on a real neighborhood of S = * _ 1 (5) = (0,0, T \ a). Thanks to the use of the Kolmogorov's approach to the hyperbolic KAM theory, the neighborhood where the flow-box variables are defined contains a piece of both whiskers. In the construction of the variables, one can make the local stable whisker become a coordinate plane (see (24)), and then the global unstable whisker can be seen as a graphic over the local stable one. In this way, the splitting distance and the homoclinic intersections between the two whiskers appear much more transparently. Our Hamiltonian takes, in the flow-box variables, a very simple form: H = HoV
=
E+(cj,J-a),
and hence the associated Hamiltonian equations are 5 = 1,
E = 0,
ip = u,
J = 0.
(23)
We recall that analogous flow-box variables have already been used 20,6 in some case where the symplectic change can be defined explicitly from the expression of the normal form, which is integrable. In our case, the normal form H is, in general, not integrable, and the construction of the flow-box variables is more involved (it uses implicit functions). Parametrizations in the flow-box variables Let us describe more precisely how the whiskers can be parametrized in the flow-box variables. Let us denote W^. = \I/ _1 (VVj^c) the local stable whisker (or more precisely a piece of it). This whisker becomes a coordinate plane, given by E = 0, J = a, and can be parametrized as follows: W^:
^ ( S ,
V
) = *-1(Z(
5 I
+ 5 , ^ ) ) = (S,0,VP,O),
sel,
^ F ,
(24) where I is some interval containing s = 0 (we have replaced s — si by s for a clearer notation: in this way we have s = 0 on S).
133
Now, we define W~ = # _ 1 o $ - 1 (VW~) as an invariant manifold of H, which is the equivalent in the flow-box variables for (a piece of) the global unstable whisker. Let us parametrize: W- :
W-{s,
SGI,
In components, we write W~(s,(p) =
(S-(s,
There is splitting of the whiskers when J~(s,tp) ^ o o r E~(stp) ^= 0. Never theless, it suffices to control the ./-component because the whisker is contained in the zero energy level: E~ + (w, J~ - a) = 0. The approximation given in theorem 2, expressed there in the original variables, remains true after changing to the flow-box variables. So the Melnikov function M = d^L also provides a first order approximation in /z for the splitting distance J~ (s,
S = $-d&8,
E = E-ds6,
ip = tp-dj6,
J = J-d^6.
(25)
In order to compare the whiskers W ^ and W~, it will be useful to express the splitting distance J~(s,f) — a as a gradient. This cannot be deduced
134
directly from (25), but this obstruction is easily overcome, 20 introducing new parameters that substitute the initial ones s, ip on the whiskers: S = S~(s,
tp = ip-(s,ip).
In terms of the new parameters S, rp, the unstable whisker W - appears nicely as a graphic over the stable whisker Wj£c, through the parametrization: W":
W-{S,4>)=(s,E-(S,1>),il>,J-(S,rl>j),
S el,
i> G T" (26)
(the interval I can have undergone a reduction). It is then natural to introduce the splitting potential as the following scalar function, periodic in tp: C(S,il>) = 6(S,0,rl>,a),
S el,
V € T".
(27)
This function also depends on /x, and is determined up to an additive constant. The (vector) splitting function can then be denned as the gradient of £ with respect to the angles: M(S,ip) =
d^£(S,rp).
The next theorem is easily deduced from the equations (25). Theorem 3 The functions £ and M. only depend on ip — uS: £(S, V) = £ ( 0 , 4 > - u S ) ,
M(S, rp) = M(0,tp-
uS).
Besides, these functions are related with (24) and (26) in the following way: E-(S,W
= ds£(S,i>),
J~(S,4>) - a = M{S,1>).
According to this theorem, the function M. gives the splitting distance (expressed in the parameters S, rp). It is important to stress that the fact that the splitting distance can be put as the gradient of some potential is a reflection of the Lagrangian properties of the whiskers. As a corollary of theorem 3, one can recover a result due to Eliasson: J there exist at least n + 1 homoclinic orbits (not necessarily transverse), biasymptotic to the whiskered torus T. This result, valid for both the regular case and the singular case, comes from the fact that a function on T n has at least n + 1 critical points (not necessarily nondegenerate), according to the Lyusternik-Schnirelman theory. 21 Then for a fixed 5, the splitting potential £ ( 5 , •) has at least n + 1 critical points, which give rise to respective homoclinic intersections between the whiskers W±, and hence to homoclinic orbits, contained in both whiskers.
135
First order approximation for the splitting potential Finally, using Poincare-Melnikov theory, we can obtain first order approxi mations for the splitting potential C, introduced in (27), and for the splitting function M- At first order in /i, these approximations are given, respectively, by the Melnikov potential L and the Melnikov function M denned in section 4, but they are good enough only for the regular case. Theorem 4 For S e l and ip € T", one has C{S, ip) = nL{xp - UJS) + O (M 2 ) ,
M(S, if) = /iAf (tf - uS) + O (/i 2 ) •
We finish with some remarks about the additional difficulties of the sin gular case. Note that theorem 4 provides an O (M 2 ) error term that is not small enough in the singular case fj. = ep with p > 0, due to the fact that the functions L and M are exponentially small with respect to e. (This is illustrated in the second example of section 5). Nevertheless, one can expect that, under some weak hypotheses on the perturbation, the predictions of the splitting given by the Melnikov potential L are also valid in the singular case, for some p > 0. To get better bounds of the O (/i 2 )-term for real values of the variables S, ip, one should bound this term on a complex strip of these variables. This requires some improvements of the results presented here. First, one needs a more precise version of theorem 1, carrying out a careful control on the loss 6 of complex domain in the angular variables. Such an improvement of the normal form theorem has already been performed by the authors, 10 and in fact analogous results had previously been obtained 2 ' 6 for somewhat different contexts. On the other hand, one needs an extension theorem and the flow-box vari ables extended to a suitable complex domain, which would lead to a significant refinement of theorem 4, of the type £ ( 5 , T/0 = fxL ty - Su>'/y/e) + O ( / x V " ) , for 5, %l> on a complex strip |ImS| < 7r/2 — e 1 / 4 , |Im ip\ < p — e 1 / 4 . Then one could obtain, for real values of S, i/>, exponentially small upper bounds for the error term, which would be dominated by the first order approximation provided by Poincare-Melnikov theory, under some general hypotheses on the perturbation. If this is true, then the Poincare-Melnikov theory gives the right predictions for the splitting even in the singular case. The problem of giving asymptotics for the exponentially small splitting of separatrices is now being researched by the authors. In fact, the strategy de scribed above has been followed 20 ' 6 in simpler situations in which the normal
136 form is integrable and the flow-box variables can be defined explicitly. Rudnev and Wiggins 8 announced an important generalization, but their proof contains essential errors. Therefore, the problem of giving asymptotics for the splitting in the Hamiltonian (1-2), in the singular case, remains still open. Acknowledgments This work has been supported in part by the Catalan grant CIRIT 1998SGR00042. Research by A.D. is also supported by the Spanish grant DGICYT PB94-0215. References 1. L.H. Eliasson, Biasymptotic solutions of perturbed integrable Hamilto nian systems, Bol. Soc. Bras. Mat, 25, 57-76, (1994). 2. M. Rudnev and S. Wiggins, KAM theory near multiplicity one reso nant surfaces in perturbations of a-priori stable Hamiltonian systems, J. Nonlinear Sci., 7, 177-209, (1997). 3. A. Delshams and P. Gutierrez, Homoclinic orbits to invariant tori in Hamiltonian systems, in C. Jones, S. Wiggins, A. Khibnik, F. Dumortier, and D. Terman, editors, Multiple-Time-Scale Dynamical Systems, IMA Volumes in Mathematics and Its Applications, Springer-Verlag, to appear (1999). 4. P. Lochak, Canonical perturbation theory via simultaneous approxima tion, Russian Math. Surveys, 47, 57-133, (1992). 5. V.I. Arnold, Instability of dynamical systems with several degrees of freedom, Soviet Math. Dokl., 5, 581-585, (1964). 6. A. Delshams, V.G. Gelfreich, A. Jorba, and T.M. Seara, Exponentially small splitting of separatrices under fast quasiperiodic forcing, Comtn. Math. Phys., 189, 35-71, (1997). 7. G. Gallavotti, G. Gentile, and V. Mastropietro, Pendulum: separatrix splitting, preprint, 97-472, mp_arc<&math.utexas.edu, (1997). 8. M. Rudnev and S. Wiggins, Existence of exponentially small separatrix splitting and homoclinic connections between whiskered tori in weakly hyperbolic near-integrable Hamiltonian systems, Phys. D, 114, 3-80, (1998). 9. S.V. Bolotin, Homoclinic orbits in invariant tori of Hamiltonian systems, in V.V. Kozlov, editor, Dynamical systems in classical mechanics, 168 Amer. Math. Soc. Transl. Ser. 2, 21-90. Amer. Math. Soc. Adv. Math. Sci., 25, Providence, RI, (1995). .
137
10. A. Delshams and P. Gutierrez, Splitting potential and Poincare-Melnikov theory for whiskered tori in Hamiltonian systems, preprint, submitted to J. Nonlinear Sci., (1998). 11. L. Niederman, Dynamics around simple resonant tori in nearly integrable Hamiltonian systems, preprint, 97-142, mp_arcamath.utexas.edu, to appear in J. Differential Equations, (1999). 12. D.V. Treschev, Hyperbolic tori and asymptotic surfaces in Hamiltonian systems, Riissian J. Math. Phys., 2, 93-110, (1994). 13. P.J. Holmes and J.E. Marsden, Melnikov's method and Arnold diffusion for perturbations of integrable Hamiltonian systems, J. Math. Phys., 23, 69-675, (1982). 14. S. Wiggins, Global bifurcations and chaos: analytical methods, 73 Appl. Math. Sci. Springer, New York, (1988). 15. C. Robinson, Horseshoes for autonomous Hamiltonian systems using the Melnikov integral, Ergodic Theory Dynam. Systems, 8, 395-409, (1988). 16. A. Delshams and R. Ramfrez-Ros, Melnikov potential for exact symplectic maps, Comm. Math. Phys., 190, 213-245, (1997). 17. C. Simo, Averaging under fast quasiperiodic forcing, in J. Seimenis, editor, Hamiltonian Mechanics: Integrability and Chaotic Behavior, 331 NATO ASI Ser. B: Phys., 13-34. Plenum, New York, (1994). 18. A. Delshams, V. Gelfreich, A. Jorba and T.M. Seara, Splitting of sepa ratrices for (fast) quasiperiodic forcing, in C. Simo, editor, Hamiltonian systems with three or more degrees of freedom, 533 NATO ASI Ser. C: Math. Phys. Sci., 367-371. Kluwer Acad. Publ., Dordrecht, Holland, (1999). 19. G. Gallavotti, Twistless KAM tori, quasi flat homoclinic intersections, and other cancellations in the perturbation series of certain completely integrable Hamiltonian systems. A review, Rev. Math. Phys., 6, 343411, (1994). 20. A. Delshams and T.M. Seara, Splitting of separatrices in Hamiltonian systems with one and a half degrees of freedom, Math. Phys. Electron. J., 3, 1-40, (1997). 21. S.-N. Chow and J.K. Hale, Methods of bifurcation theory, 251 Grundlehren Math. Wiss., Springer-Verlag, New York, (1982). Internet access: The authors' quoted preprints are available in the P r e p r i n t s pages of http://www-mal.upc.es, or in the p u b / p r e p r i n t s di rectory of f t p : I Ittp-mal.upc.es.
I N F I N I T Y MANIFOLDS OF C U B I C POLYNOMIAL HAMILTONIAN V E C T O R FIELDS W I T H 2 D E G R E E S OF FREEDOM MANUEL FALCONI Departamento de Matemdticas, Fac. de Ciencias, UN AM, Mixico, D.F. 04510, Mixico, E-mail:[email protected] ERNESTO A. LACOMBA Departamento de Matemdticas, UAM-I, P.O. Box 55-534, Mixico, D. F. 09340, Mexico, E-mail: [email protected] JAUME LLIBRE Departament de Matematiques, Universitat Autonoma de Barcelona, 08193 Bellaterra, Barcelona, Spain, E-mail: [email protected] Let X be the Hamiltonian vector field with two degrees of freedom associated to the cubic polynomial Hamiltonian H (i,y, z,w). Using the Poincar6 com pactification we show that all the energy levels of X in R4 reach the infin ity in a surface topologically equivalent to the intersection of the 3-dimensional sphere S3 = {(x, y,z,w) 6 R4 : x2 -I- y2 + z2 + w2 = 1} with {(x,y,z,w) G R4 : H3 (x, y, z,w) = 0}, where Hz denotes the homogeneous part of degree 3 of H. Such a surface is called the Infinity Manifold associated to H. In this paper we describe all possible infinity manifolds of cubic polynomial Hamiltonian vector fields with 2 degrees of freedom. Our method is general, but since actual computations can become very cumbersome, we work out in detail only three out of ten possible cases.
1
Introduction
We consider the Hamiltonian vector field X with 2 degrees of freedom gener ated by a cubic polynomial in four variables H (x, y, z, w). This vector field can be extended to infinity in the following way. Poincare compactification allows to project the vector field onto the north and the south hemispheres of the sphere S 4 C R 5 . The extended vector field X on the equator S 3 C S 4 represents the asymptotic behavior at infinity of our original Hamiltonian vector field. The vector field in the open hemispheres is not Hamiltonian any more but it keeps having nevertheless a first integral. All the corresponding energy levels intersect the equator S 3 in the same (perhaps singular) 2 di mensional infinity surface W, invariant by X. This surface turns out to be exactly the intersection with the sphere S 3 of the set of zeroes of the cubic homogeneous part H3 (x, y, z, w) of H (x, y, z, w). We see that W is compact
138
139
and it is generically smooth and with isolated equilibrium points. In this work we describe the topological classification of surfaces W, through the computation of the index at equilibrium points of the vector field X restricted to W. This vector field Y on the cubic surface W represent the asymptotic behaviour at infinity of the vector field X. One can compute the index of the vector field Y on any connected component U of W as the sum ot its indices at the isolated equilibrium points in U. By the PoincareHopf Theorem, the Euler characteristic x of £/ is equal to the computed index. Finally, the genus g of U is related to its characteristic through the formula X = 2 — 2g. The genus of any compact connected smooth surface determines its topology, as required. In Section 2 we describe in detail the Poincare compactification of poly nomial Hamiltonian vector fields as in Delgado et al 3 . We see how the infinity manifold and its induced flow are defined. Not only the infinity manifold but its flow depend only on the highest degree homogeneous part of the poly nomial Hamiltonian. Hence, we may assume without loss of generality that the Hamiltonian is a homogeneous polynomial. Then we specialize to cubic homogeneous polynomial Hamiltonians with two degrees of freedom. In Section 3 we study general topological considerations. The most im portant conclusion when applied to cubic surfaces W is that the genus g of any connected component must be even. In Section 4 we consider the simple case of cubic reducible homogeneous polynomial Hamiltonians with two degrees of freedom. Since any such polyno mial is the product of a quadratic polynomial times a linear one, we begin by studying the quadratic case. The conclusion for cubic reducible Hamiltonians is that the corresponding cubic surface W is generically the union of three disjoint spheres. In Section 5 we begin the study of the topological classification in the cubic irreducible case. We take a normal form proposed by Sylvester and Arnold, which has only 4 coefficients. Via a change of variables we can see that the set p of parameters where the corresponding W is singular, is the intersection of 15 hyperplanes with the closed positive cone. We study the topology of the infinity manifold W — E°° in each of the connected components in the complement of p. There are too many com ponents, but one can take a quotient by the equivalence relation obtained by permutation of coordinates and other obvious symmetries. This way the number of components is reduced to 10. This is sketched in Subsection 5.1 by following a combinatorial procedure, whose details are given in the com panion paper Falconi, Lacomba and Llibre6. In Subsection 5.2 we apply the Poincare-Hopf Index Theorem to compute index and then genus for surfaces
140
W in the 3 components appearing along the diagonal in the coefficient space. This can be done for the remaining 7 components, only that the compu tation of the equilibrium points and indices although straightforward becomes very cumbersome. 2
Poincare" Compactification of a Hamiltonian Polynomial System
Let X be a polynomial vector field of degree m in R n . The Poincari. compactified vector field X corresponding to X is an analytic vector field in duced in Sn as follows (see, for instance Cima and Llibre 2 ). Let Sn = {V = (yu ■ ■ ■ 2/n+i) € R n + 1 | y\ + ■ ■ ■ + yl+1 = 1} (the Poincar6 Sphere) and TySn be the tangent space to 5 n at point y. Consider the central projec tion 0 * :T(0>...,o,i)5'n —♦ 5 n . These maps define two copies of X, one in the northern hemisphere and the other one in the southern hemisphere. De note by X' the vector field D^ o X defined on 5 n except on its equator n_1 is identified with the infinity of R". Sn-i = Sn n { y n + i = o} . Clearly S In order to extend X' to a vector field on the whole Sn it is necessary that X satisfies suitable conditions. In the case that X has degree m, X is the only analytic extension of y™+iX' to 5 " . On S n \ S n _ 1 there are two symmetric copies of X, and knowing the behaviour of X about 5 n _ 1 , we know the be haviour of X at infinity. The Poincare compactification has the property that Sn~x is invariant under the flow of X. When this procedure is applied to a Hamiltonian vector field XH = fe' ••••& - ^ ■ ■ ■' " £ £ ) w h e r e H ■■ R 2 d — > R * a Polynomial of degree m -f 1, a new compactified vector field is obtained on the sphere 5 2 d , which is denoted by XH and is given by ~ , dH* dH* dH* XH = (~ + -M/i, • • •, -x— + Aj/d, - — oyd+i oy2d ay i dH* + Ay 2 d,Ay„ +1 ), dyd where n = 2d and H'(yi,--yn+1)
x
h \yd+i, • • •,
(2)
= yZ£H(-p-,...-^)
(3)
\J/n+l
and J^f
dH'
(1)
8H*\
J/n+1
141
ir1
Figure 1. Figure 2.1: Poincari compactification.
Computing the derivatives of H* along solutions curves of XH one gets 4- PH' u' + dy„+i y "+i
*£ d
[ff («£ + **) + *£(-«£ + **«)] + *Er*^ (5)
-*(£[*»
= A(m +
, a//* 2/«+d] + a w yn+ij dyn+i
!)#*.
From this equality it follows that the level energy H* — 0 is invariant under the flow of XH . In what follows for any h € R we consider our Hamiltonian H of degree m in the form H = h-\-H\A Hm, where Hi is the homogeneous part of H of degree i for i = 1,2, • • •, m. The vector field XH satisfies the following theorem Theorem 1 Let H:R2d —► R be a polynomial Hamiltonian of degree m de pending on n=2d variables (j/i, • • • yn), and let XH be the Hamiltonian vector field associated to H. Let XH be the Poincari compactification of XH. Then the following statements hold:
142
(a) The energy level Eh = H * (0) is mapped diffeomorphically by the cen tral projection onto a subset E^ (E^) of the northern (southern) hemisphere. These sets are invariant under the flow of XH . __ (b)Let S"" 1 = Sn n {y n + 1 = 0} . Then the restriction XH \ 5 n _ 1 is given by f dHm
+
Xy
\dyd+\ where A = i=l
v
■-,
dHm a
oyid
+Ayd,
dHm
;-ddHy:+^)
(6)
W|fc)"
(c)The flow of XH leaves the subset EP° = {y £ S"" 1 \Hm{yu--yn) = 0} invariant. (d) The boundary of E^ is contained in EP°. (e) If 0 is a regular value of Hm | 5 n _ 1 , then E°°is an (n — 2) —dimensio nal smooth manifold. _ (f) the critical points of XH | 5 n _ 1 are contained in E °°. For a proof of Theorem (1) see Delgado et al 3 . In this paper we are interested in describing the topology of E°° for the case when dimension n = 4 and degree m = 3, corresponding to a cubic Hamiltonian system with 2 degrees of freedom. In this case H3 (x, y, z, w) is a homogeneous cubic polynomial and W = E°° = S 3 n {H3 = 0} is 2 dimensional. In order to get a classification of the topology using the index of the vector field (6 ) defined on E°°, we need the following very well known result. Remark 2 Since W is a compact, orientable surface, then each connected component U ofWis characterized by its genus, or equivalently by its Euler characteristic. In fact, we know \ = 2 — 2g, where x is the Euler characteristic of U and g is its genus. On the other hand, by the Poincare-Hopf Theorem, the characteristic is equal to the index of our vector field in W. 3
General Topological Considerations
First we prove by using results from duality theory some general properties about E°° = Sn n {Hm = 0} . As a corollary we will obtain that each cubic surface has even genus. We write Sn = G _ 1 (1), where G (yi, • • •,y n + i) = y\ + y\ + ■ • • yl+1 and define V+ = {yGRn+1\Hm(y)>0}
,
V. = {y \ Hm (y) < 0} ,
(7)
143
and
w+ = v+nsn, V = {y\Hm(y)=0},
w_ = v_nS",
,.
W = VnSn.
{0)
Let M be a compact orientable manifold without boundary of dimension 2r. Each homology group with coefficients in Z (the set of integers) is of the form H,(M) = L , ( M ) © T , ( A f ) ,
(9)
where Lq is a free group and Tq is a torsion group. The cohomology group of M is denoted by H' (M). From the Poincare Duality Theorem there is an isomorphism T:H r (M) -» H r (M). We define 7 : H r x H r — Z by 7{x,y) = (T(x) ,y). This bilinear form is symmetric or skewsymmetric if r is even or odd respectively, see Dold4. In a natural way T induces a unimodular form T in H r /T r (M) x H r / T r (M) which is defined in the usual way by taking the class representative. Clearly, the unimodular form F(x,y)=x.y,
(10)
is symmetric or skewsymmetric if r is even or odd, respectively. The following form of the Poincare dualitiy theorem is very well known. Theorem 3 Let M be a compact, orientable manifold without boundary of dimension 2r. Then Lq(M)^L2r_q(M):
r,(M)sr 2r - g (M).
U1
>
The symbol = denotes isomorphism between the spaces. Let i: W —> W+ be the canonical inclusion, where A denotes the closure of A, and W, W+ are given by (8). This map defines a homomorphism i„q : Hq (W) —► H9 (W+). The kernel of i*q is denoted by Kq. Unless it is not relevant, we omit the subindex q. Lemma 4 Assume that the degree of Hm is odd. Let ^»q:Hq (W) —* H, (W) be the homomorphism induced by the antipodal mapping *$?. IfO < q + 1 < n, then Hq (W) = Kq® * . , {Kq). Proof: Since m is odd, we have that S ^ W + U * ^ )
and W = W + n 4 ' ( W + ) .
(12)
On the other hand Hi(S !
'
| Z , i€{0,n} [ 0 , otherwise.
(13) v
'
144 We take q such that 0 < q + 1 < n. Since the Mayer-Vietoris sequence
.—fH^1(5B)-H,(W0->H,(i^)eH9(»(wg)-H,(S")-.(14) is exact and H,+i (5 n ) = H g ( 5 n ) = {0}, it follows that
1*
(15)
induces in homology the following commutative diagram K C H, (W) - i * H, ( H ^ ) * . | =*
S| * .
(16)
H,(W)-^H,(*TO). Since AT is the kernel of z. we get that <£, (if) is the kernel of j , . It follows that
(17)
p(tf.(JO)cH,ro©0.
(18)
and
Set (0,z) € {0} © H, (\& (W+)). Since
(19)
¥>(*.(*)) = H , ( w g © { 0 } .
(20)
Analogously
Therefore, we have H, (W) = K © * , (A").
■
Proposition 5 Let n=4k+3, with k in NU{0} . It is assumed that T^k (W) 0. Then there exists r such that i/2fc+i (W) = Z 4 r . Proof: Since the dimension of W is 4fc + 2, it follows from Theorem 3 that T 2fc+1 (W) = T2k (W) = 0.
=
(21)
145
Then H2k+i (W) is free of torsion and using Lemma 4 we get that H2fc+i (WO = Z 2 '.
(22)
The antipodal homeomorphism ty induces the bilinear form B(x,y)
= J " ( s , * , y ) = a:.*.y,
(23)
for (x, y) e K x K. Since ^ reverses orientation we get that x.y = —^»x.^»y. Thus x.*«y = - * , x . * , * , y = - * , x . y = y.*»x,
(24)
where the last equality is due to the fact that T is skewsymmetric. So B is a symmetric bilinear form. From the equality H2jt+i (W) = K®$!*K, it follows that there exists a basis a\, ■ ■ ■ Qj, ^ , Q I , • • • \t*aj for which the matrix (
Oti.Ctj
V * . Cti-CXj
,K
0 b\ -bO)'
corresponding to T is unimodular. Therefore B is unimodular; on the other hand B takes even values, see Lopez de Medrano 7 , then we conclude that b has a even number of rows. So, dim K = I is even. Finally H2fc+1 (W) = Z 4 r . This proves our assertion.
(26) ■
Corollary 6 If n-3, then W is not a connected surface of odd genus. Proof: Since n = 3, we have k = 0. Therefore, Hi = Z 4 r . Moreover, Hi is the fundamental group abelianazed, then we get that Hi = I?9. So, we have that 2g = 4r. It follows that g is an even number. ■ 4
Topological Classification in the Cubic Reducible Case
In this Section we study the topological classification of the infinity set E°° in the case when H3 (x, y, z, w) is homogeneous of degree 3, but also reducible. Since any reducible cubic polynomial is factorized into a quadratic times a lin ear hamiltonian, we start by studying the case of homogeneous Hamiltonians H? (x,, y, 2, w) of degree 2. Let H2 (x, y, z, w) = a n x 2 + a223/2 + 0.33Z2 + a^w2 + 2 (ai2xy + a ^ x z + auxw + 0232/2 + a24j/u> + 034.01;) the general expression of a quadratic form in 4 variables. It can be diagonalized by means of an orthogonal transformation taking it to the form
146 #2 (x,y, z,w) — p\x2 + P2J/2 + p3z2 + p4iu2,where pi, i = 1,2,3,4 are the eigenvalues of the matrix associated to H2. Generically they are all different from zero. According to Theorem 1, the corresponding E°° is invariant under the flow defined by the following system of differential equations x = p3z + (pi - p3) x2z 4- (p2 - Pi) xyw, y = PAW + (pi - p3) xyz + {p2 - pi) y2w, z = -p\x + (pi - p3) xz2 + {p2 - Pi) yzw,
. ^
W = ~P2V + (Pi ~ Pi) XZW + (p 2 - Pi) yW2.
Prom Remark 2, we need to compute the index of this vector field at each equilibrium point. This gives the Poincare characteristic, and hence the genus of the compact orientable surface E00. After some algebraic manipulations, the equilibrium points of (27) are obtained as solutions of the system of equations Pzyz - Pixw -pixw + p2yz p3z2 +pix2 PiW2 + p2y2
0, 0, 0, 0.
(28)
In order to solve (28), it is convenient to distiguish the following possibilities, according to the signs of the eigenvalues: I) All eigenvalues have the same sign. Therefore, E°° = ill II) Three eigenvalues have the same sign and one has a different sign, i.e. the signature of H<x is odd. Because of the symmetry and homogeneity of H2, it is enough with ana lyzing the case pi, P2, Pi positive and p3 negative III) There are two positive eigenvalues and two negative ones. It is enough to consider the case p\, p2 positive and p3, p^ negative. Theorem 7 Let XH the vector field induced by a quadratic polynomial hamiltonian. If E°° ^ 0, then generically we have _ a) if H2 has odd signature, the vector field XH has exactly four equilibrium points in E°° with index 1. __ b) if H2 has even signature, the vector field XH has 8 equilibrium points in E°°. Four of them have index 1 and the others have index -1. Proof: Recall that we have to intersect with the sphere S 3 , which is the level 1 hypersurface for the function G (x, y, z, w) = x2 + y2 + z2 + w2. a) This is case II). Then w = y = 0. We are left with just one equation
147 from ( 28) plus the condition for being in the unit sphere p3z2 + pix2 = 0, 2 _,_ ,2 _ ,
(29)
We get 4 equilibrium points (fc,0,fci,0), where k = ±W
p
J , fci
± w ~^* are real numbers, since p3 < 0. Computing the jacobian we get d(H2,G) d(x,z)
2pxx 2p3z = 4 (p! - p 3 ) xz, 2x 2z
(30)
so that y, u; can be taken as independent variables at equilibria. At these points y - p4w + (pi - pz) xyz + (p2 - p*) y2w, w = -p2y + (pi - Pi) xzw + (p2 - PA) yw22,
(31)
where x, z are written as functions of y, w. Hence , the jacobian matrix of the linear approximation of this system at (fci,0, k,0) is '(Pi-P3)M -P-Z
P \ u u ) . (P\-pz)kxk)
(32)
Its trace and determinant are Tr = ±2^/—P~\pl t^ 0 and Ai = p2p4 —pip3 > 0. Therefore, Index = 1 at these points. b) This is case III). Notice from the fourth equation in (28) that w = 0 is equivalent to y = 0 and we have again the 4 points from case a). Similarly, x = 0 is equivalent to z = 0. Then the equations P2y2 + PAW2 = 0, y2a +l «w.2 a=- i1, give 4 new equilibrium points ( 0 , r , 0 , r i ) , where r =
( 33 ) _£4_ ± « / P4-P2
'
ri
=
± w ~ ^ . A computation as in the case of the other 4 points shows that x, z can be taken as independent variables. The jacobian matrix of the linear approximation of the vector field at (0,r,0, r\) is (P2-P4)nr -P2
\ P3 KP2-P4)r^rJ,
with trace Tr — ±2s/—p2p4 ^ 0 and determinant A2 = p\pz — p2Pi = — A i . Now we investigate the index of these 8 points. The trace is always nonzero, but the determinant Ai at the points (fc, 0, A:i, 0) and A2 = —Ai at the points (0,r,0,r\) can have any sign. Excluding the nongeneric case
148
PiPz — P2P4, we see that when Ai > 0, the points (k,0,fci,0) have index 1, while the points (0,r,0,ri) have index - 1 ; and viceversa when Ai < 0. Hence, there are 4 points od index 1 and 4 points of index —1 in any case. Finally, from the first two equations in (28). we see that in order to have an equilibrium point with all nonzero coordinates we need that A i = 0, obtaining a nongeneric case with a closed curve of equilibria. This is a bifur cating case, where we find that the above 8 equilibrium points are contained in the closed curve of equilibria ■ Theorem 8 a) If the signature of the quadratic form H? is odd, then E°° has two connected components, each one homeomorphic to a sphere S 2 . b)If the signature is 2, then E°° is homeomorphic to a torus. c) If fyhas signature 0 or 4, then £°°=0. This means that there are no escape solutions. Proof: a) Each compact connected component of E°° has an even Poinca re characteristic \- The value 4 is imposible, since the genus g is greater than zero and x = 2 — 2g. Hence, there are 2 components with x — 2, i.e. g — 0. So, they are spheres S2. b) By a long but straightforward verification one proves that in this case E°° is connected. Since there are 4 equilibrium points with index 1 and 4 points with index —1, we have x — 0 a n d g = 1. This is topologically a torus Sl x S1. Case c) is trivial. ■ For example, if H2 = x2 +y2 - z2 — w2, under the condition x2 + y2 + z + w2 = 1, we get x2 + y2 = 1/2 and z2 + w2 = 1/2, which defines a torus S1 x S1. This belongs to case b) in the above theorem. On the other hand, if Hi = x2 + y2 + z2 - w2, we can write equivalently x 2 + y2 + z2 — 1/2 and w2 = 1/2, which define two spheres S2 x 5°. This example corresponds to case a) in the above result. Let us consider the case where i/3 (x, y, z, w) is a reducible cubic homo geneous polynomial. Any cubic homogeneous polynomial contains in general 20 coefficients as follows 2
H3 (x, y, z, to) = ax3 + a3x2w cixz 2 eixyz
by3 + cz3 + dw3 + a i i 2 y + a.2X2z+ + bixy2 + b3y2w+ + C2yz2 + czz2w + d\xw2 + d2yw2 + dzzw2+ + e2xyw + e^xzw + e^yzw.
. . * '
If Hz is reducible, the following result describes generically the topology of£°°. Proposition 9 If Hz is a reducible polynomial then EP° is generically a union of three disjoint spheres Proof: If H3 is reducible, we can write H3 = LQ, where L is linear and
149
Q is quadratic. The equation LQ = 0 is equivalent to i) L = 0: a hyperplane. whose intersection with S3 is S 2 , and ii) Q = 0: whose intersection with S 3 is S2 x S°. This gives 3 spheres for E°°. The case where Q = 0 gives a torus S 1 x S 1 is not possible generically. Indeed, if E°° is the union of a torus and a sphere, being simultaneously smooth, we would have H\ {E°°) = I? since it is the abelianized of the fun damental group. This contradicts Proposition 5 , since the range of H\ (E°°) must be a multiple of 4. ■ The example L = 0, Q = 0, given by the equations w = 0; x2 + y2 + z2 — w = 0, when intersected with S3 produces 3 two-spheres 2
w = 0,x2 + y2 + z2 = l,
,
w2 = lx2 + y2+z2
(6b
=l
>
illustrating the proposition. On the other hand, let L = 0: ax+by+cz+dw = 0 any hyperplane and Q = 0: x2 + y2 — z2 — w2 — 0. Since L = 0 contains the 2-dimensional subspace defined by ax + by = 0 and cz + dw = 0, we see that it must have points in common with the torus x2 + y2 = ^, z2 + w2 = \ obtained by intersection of Q = 0 with S 3 . Hence, the torus has always points in common with the sphere defined by any hyperplane through the origin. 5
Topological Classification in the Cubic Irreducible Case
The case H3 irreducible is considerably harder. For analyzing this case we will take the normal form proposed by Sylvester ( Segre8 and Arnold et al 1 ) H3 (x, y, z, w) = x3 + y 3 + z3 + wz + (ax + by + cz + dw)3 .
(37)
This normal form produces only irreducible cubic polynomials. Application of linear isomorphisms of R4 generate new irreducible polynomials. Any such endomorphisms can be identified to a 4x4 real matrix, requiring thus 16 entries. With the 4 coefficients of the normal form, we get the 20 coefficients that the general form (35) has. We expect that the normal form (37) generates a big open subspace of the 20 dimensional space of coefficients of the cubics. So our topological classification of surfaces E°° = {H3 = 0} n S3 will depend on 4 parameters. Notice that E°° is a smooth surface if and only if VH3 (p) is not parallel to p, for any p € E°°. By homogeneity of H3, this is equivalent to the condition V//3(p)^0, for any p G E°°. Indeed, assume that V # 3 (po) = Ap0
150
for some po € E°°. But Euler formula for homogeneous functions gives P0.VH3 (po) = 3H3 (po) = 0- Prom the above formula, we get A||pol|2 = Po.ViJ 3 (Po)=0.
(38)
Hence A = 0 and VH3 (po) = 0, as required. Let A (a, 6, c, d) = 0 be the equation which defines the set of parame ters (a, 6, c, d) € R4 such that there is a point x 6 E°° (a, 6, c, d) satisfying Vi/3 (x) = 0. We have that V//3 (x) = 0 if and only if the following equalities hold a (ax + by 4- cz + dw) =—x2, b(ax + by + cz + dw)2 = -y2, c(ax + by + cz + dw) = -z , d (ax + by + cz + dw) = — w2.
.
.
This means that the parameters a, 6, c, d are non-positive, i.e. (a, 6, c, d) be longs to (R~) with R _ = (—oo,0]. By extracting square roots in equations (39), we obtain a homogeneous linear system in x,y,z,w. Only that ± 1 fac tors have to be introduced in each of the 4 equations. We verify at once that the condition that the determinant of the system annihilates, can be writ ten as the set of zeroes of the following polynomials in the new non-negative variables A = ( - a ) 3 / 2 , B = ( - 6 ) 3 / 2 , C = ( - c ) 3 / 2 , D = {-d)3/2 . (A + e j S + ^ C + eaD) 2 = 1,
(40)
with e< G {-1,1}. The change of variables F : ( R - ) 4 -» ( R + ) 4 is the homeomorphism given by F (a, 6, c, d) = ( ( - a ) 3 / 2 , ( - 6 ) 3 / 2 , ( - c ) 3 / 2 , (-df2)
= (A, B, C, D),
(41)
transforming {A = 0} C (R~) into the subset p of ( R + ) , defined where the equations (40) are satisfied. Here R + = [0,00). The complement of p induces a partition of the parameter space R4 into open connected components. In a companion paper Falconi, Lacomba and Llibre 6 , we have proved that all surfaces corresponding to parameters in the same component are diffeomorphically equivalent. Consequently our classification problem has been reduced to determine all connected components and for each component, to identify the topological surface W.
151
5.1
Counting the Connected Components
To determine all the connected components of the partition induced by p , we use a combinatorial method, which we briefly describe below. We consider the system A + B + C + D = a0,A-B-C-D = a'0, A-B + C + D = ai,A + B-C-D = a\, A + B-C + D = a2,A-B + C-D=:o-'r A + B + C - D
= O-3,A-B-C
. . '
[
+ D =
The above hyperplanes (40) are defined by taking the a'S equal to ± 1. Each 4-tuple (A,B,C,D) £ R4 determines an 8-tuple (<J"O,<7I,<72,<73,CTO> a1,a2,o-3). Let J_ = (—oo, —1), Jo = ( - l i l ) ) J+ — (I) 0 0 )- We say that two 8-tuples (xi,...,xs) and (t/i,...,ys) of R 8 with Xi , y» ^ ± 1 , are equivalent it we have Xi £ J^ if and only if j/i 6 J^ for each i — 1,..., 8 and U S {—, 0, + } . Proposition 10 Two points (A,B,C,D) and (A\,Bi,Ci,Di) belong to the same connected component of (R + ) \ p if and only if their corresponding 8-tuples are equivalent. Proposition 10 is proved in Falconi, Lacomba and Llibre 6 . It gives a com binatorial procedure to count connected components, but we have to identify equivalent components modulo permutation of coordinates. This is due to the fact that the equations which define W are invariant under a permutation of coordinates (see Remark 11). We proved that two points (A,B,C,D) and (A\,B\,C\,D\) belong to the same connected component of (R + ) \ p if and only if their correspond ing 8-tuples are equivalent. That is, each component is identified by a word of 8 letters in the alphabet {-, 0, + } . In principle we have 3 8 words, but cer tain constraints allow us to reduce the number of possibilities. The following inequalities are easy to prove, using equations (42) °0 ^ °'l)^2,0'3iffii o 2) o 3 ^ °0,
(43)
cr0 > 0.
(44)
The sum of c?i and a\ is equal to 2A, for all i. Therefore we get 0 < (o-o + o-'0)/2 = (a, + a[)/2 = (a2 + a'2)/2 = (a 3 + a'3)/2.
(45)
By comparison of the equations which define a we see that the inequalities Ci >
hold for i = l,2,3, and k ^ i.
ff'k,
152
Let n = (CTO + 0o) /2- The entries of a have the following properties. (i) Since fi is non-negative then UQ G JO o r °~o €E J + . (ii) If (To € Jo then each entry of a belongs to Jo- This follows from (43) and \i > 0. (iii) From (46) we obtain that the entries of a given 8-tuple satisfy either the inequalities O'Q
< cr'i < a'j < a'k < fi < ak < o-j < ai <
CTO,
(46)
or ff0 <
CT
» ^
CT
j ^
CT
* ^ M < ^fc < ^
< cr< < °o-
(47)
As long as we are not interested in the value of the entries of any octect, but in their sign we will identify an octect with its corresponding array of signs (see Proposition 10); i.e., if the i-th entry of a belongs to Jk, we replace it by k in a, where k = - , 0 or +. Although the following remark is obvious, it is useful in order to identify all the cubic surfaces. Remark 11 Notice that the topological type of the cubic does not depend on the ordering of the 4-tuple (A,B,C,D). Hence, the octect generated by a per mutation of (A, B, C, D) corresponds to a component in the parameter space where the cubic has the same topological type as the component corresponding to (A, B, C, D). So, applying permutations to the parameters (A,B,C,D) in equations (42), we prove the following result Lemma 12 Suppose we have u,v,s,t 6 {+,0,—}. Then the octects of the form 1) (CTO,U,u,v,a' 0 ,s,s,t), 2) (ao,v,u,u,o-'Q,t,s,s) and 3) (CTO,U,v,u,o-' 0 ,s,t,s) are equivalent. Also the octects of the form 4) {o-0,u,u,v,a'0,t,s,s), 5) (o-0,v,u,u,a'0,s,t,s), 6) (o-0,u,v,u,o-'Q,t,s,s), 7) (o-0,u,v,u,cr'0,s,s,t), 8) (o-0,u,u,v,a'0,s,t,s) and 9) (a0,v,u,u,a'0, s,s,t) are equivalent. An 8-tuple <7 is trivial if it is equivalent to an octect generated by a set of parameters (A, B, C, D) which has al least one entry equal to 0. All these 4-tuples correspond to the connected component containing the (0,0,0,0) in R4\PBefore stating the conditions which imply that an octect a is a trivial one, we introduce some notation to simplify. We say that an octet
153 It is not hard to prove that an 8-tuple a is trivial if any of the following 7 groups of conditions is satisfied i)a'0cr3i ^ O ^
,a[a2,o'2ai,
,(ToO'l ,a'sa2,o'2(T3,
"K^i ,OQ02 ,a[(73,o' ai,
iii)a0a'2 iv)a'0(-,
v)a3a0,(T2(TU
vi)a2{-aQ), vii)o\o0,
3
\(-(Tx),a2(-
-0-2), tf3("-0-3),
(48)
cr 1 CT 3 ,CT3(7 0 ,
(--°'s){-°\),°W
Because of condition (44), we have that o"o = 0 or <JQ = +, for all a. To count all the possible octects, we will first consider the case OQ = 0, and then the case ao = +• To identify equivalent octects, we apply the set of constraints (43)-( 47), the conditions for trivial octects given by (48) and the Lemma 12. This way we get 20 equivalence classes of octects. It is important to remark that two non equivalent classes can repre sent surfaces of the same topological type. Let (A, B, C, D) be a point in a component represented by an octect of the form (si,S2, • • •,Sg) with s< in {+,0, —} , and P be a permutation in the set {A, B,C, D} . When we sub stitute the set (P(A), P(B), P{C), P{D)) into the Equations (42), an octect (s'j, s'2,- - • i s'g) with sj in {+, 0, - } is obtained. Notice that since (A, B, C, D) and (P(A), P(B), P(C), P(D)) determine surfaces of the same topological type, then we can consider that (s\, s2, ■ ■ ■, Sg,) and (s[, s 2 , • • -, s'&) are equiv alent. Using this method for the 20 octects identified above, they can be put together into 10 classes, as shown in the following table 0 1 2 3 4 5 6 7 8 9
[0,0,0,0,0,0,0,0], {+,+,+,+,0,0,+,+], [ + , + , + , + , - , - - - ] , [+,0,0,0,0,0,0,0], [+,0,0,+,-,0,0,0 , [+,+,+,0,0,0,0,0], [+,+,+,0,-0,0,0], [+,+,+,+,0,0,0,0 ) [+,+,+,0,-,0,0,+],[+,+,+,0,-,-,0,0 , [+,+,+,+,0,0,0,+], [+,+,+,0,-,-,-,+], [+,+,+,+,0,+,+,+], [+,+,+,+,-,0,0,0], [+,+,+,+,-,0,0,+], [+,+,+,+,-,-,0,0], [+,+,+,+,-,-,0,+] , +,+,+,+ -,-,-,0],
[+,+,+,+,-,+,+,+], [+,+,+,+,-,-,-,+]•
so, there are at most 10 different topological types.
(49)
154 5.2
Poincari Index Theorem and the topology of E°°
In this subsection we prove that E°° corresponding to regions 0, 1 and 6, which are found along the diagonal of the parameter space, is topologically a sphere, three disjoint spheres and a connected surface of genus 6, respectively. We begin by region 0. P r o p o s i t i o n 13 / / H3 = x3 + y3 + z3 + w3, then E°° = (H3 = 0) D S3 is a sphere S 2 . Proof: Indeed, the equations to be satisfied are x3 + y3 + z3 +w3 = 0, x2 + y2 + z2 + w2 = 1.
(50)
We will prove that to any point (x, y, z, 0) in the equator of 5 3 C R 4 , i.e. such that x2 + y2 + z2 = 1, there corresponds exactly one point of E°°, and conversely. In fact, this point generates the vector (x, y, z, — f/x3 + y3 + z3 1 satisfying Hz = 0. By homogeneity, the whole positive ray through that point belongs to {Hz = 0} . Intersection with S3 provides a unique point (x, y, z, -y/x3
+ y3 + z3) L - 1 1*1 2 2 2 3 3 3 2/3 (x + y + z + (x + y + z ) ) V
(51)
in E°°. The converse is clear. 3
■ 3
3
3
P r o p o s i t i o n 14 If H3 = x + y + z + w - {x+ y + z+ wf , then E°° = (Hz = 0) n S 3 is a connected surface Proof: We define the sets D+ = {{x,y, z, 1) | Hz(x,y,z,l) = 0} and D~ = {(x, y, z, — 1) | Hz (x, y, z, — 1) = 0} as the intersections of (Hz = 0) with the hyperplanes (w = 1) and (w = —1), respectively. The radial projec tion n + (x) : R 4 \ { 0 } -> S3 is defined by 11+ = x / ||x||, while we define II~ = —II + . Notice that II + projects D+ onto the north hemisphere (w > 0) while II~ (D+) is its projection onto the south hemisphere (w < 0). One can verify that
E°° = n + (D+)urir (D+). If we let 6 = y + z, c = y — z, we obtain D+ = (l + b)x2 + (l + b)2x + b(l + b) + ^b(b2-c2)
=0.
(52)
Since (1,0,0,0) G E°°, we can see that n + (D+) nUr (£>+) ^ 0. So to prove that E°° is connected it is enough to prove that n + (D+) is connected and this is true if D+ is connected. This fact is proved below.
155
Figure 2. Figure 5.1: The open regions A > 0.
If b 7^ — 1, we solve Equation (52) as
- (1 + 6)2 ± ^/(l + 6)4 - 46 (1 + 6)2- -( l + 6)6(6 2 -c 2 ) 2 (1 + b)
(53)
It gives real roots whenever the discriminant A = (1 + b) ((1 + 6) 3 - 46(1 + b) - b (62 - c 2 ))
(54)
is non negative. We need to locate the curves where A = 0 and then the regions where A > 0 in the plane 6-c. We first notice that A ^ 0 when 6 = 0. We can write A = 6(1 + 6) (c2 -
156
By looking at the three factors 6,1+6 and c 2 - ^ (6) in the above expression for A, we verify that there are two open regions where A > 0, which are shaded in the Figure 5.1. These shaded regions are the projection of the set D+ along the x-axis.Then D+ is the union of three subsets A, B, C defined by A= Ux,y,z)\xeR,l + b = 0,b2 = c2}, B= \{x,y,z) | l + 6 > 0 , x = - ( l + 6 ) / 2 ± V S / 2 ( l + 6 ) } , C = {(x,y,z) | 1 + 6 < 0,s = - (1 + 6) /2 ± V S / 2 ( 1 + b)} .
(55)
We see that B projects on the upper shaded region, C projects on the lower shaded region, while A projects into the 2 points where they touch each other. For each point in the subset A > 0 there are exactly 2 points in D+ projecting into that one. Those 2 points are joined together as we approach the subset A = 0. There is an exception at the line b = — 1 because of the singularity in the expression for x. As we approach it from the shaded regions the solutions satisfy x —> ±oo. So, we conclude that D+ is connected. ■ According to Theorem 1 and considering the normal form (37) for if3, the surface E°° is invariant under the flow of the vector field given by the system of ordinary differential equations x = Z(z2 + ce2) + xX y = 3(w2 + de2) + yX z = 3 ( - x 2 - ae2) + zX w = 3 ( - j / 2 - be2) +wX
= Fj (x), = F2 (x), = F3 (x), = F4 (x),
, , ^°>
where e = ax + by + cz + dw and A = 3z (x 2 + ae 2 ) — 3x (z 2 + ce 2 ) + 3w (j/ 2 + 6e2) — 3y (w2 + de2), and x = (x, y, z, w). The equilibrium points of the vector field (56), are the solutions of the equations 3(z 2 + ce 2 ) + xA = 3(w2 + de2) + yX = - 3 ( x 2 + a e 2 ) + *A = - 3 ( y 2 + fee2) + wX =
0, 0, 0, 0,
.
. ^'>
with the extra condition that they must belong to the sphere S 3 . Lemma 15 The number of equilibria and total index of System (56) are as follows a) 6 equilibrium points with the global index 2 if all the parameters are null b) 10 equilibrium points with global index 6 if the parameters are -(1/2,1/2,1/2,1/2) c) 26 equilibrium points with global index -10 if the parameters are -(1,1,1,1).
157
We summarize the equilibrium points as follows. We first list the equilibrium points for the parameters a = b = c = d = — 1. They are separated in groups, acoording to the number of null coordinates. There are 4 points with 2 null coordinates and index 1 Pl:x
= 0 = z,y = ±,w=-^;P2 P 3 : y = 0 = w,x=±,z=-±;Pi
= -Pu = -P3.
^>
There are 16 points with one null coordinate and index - 1 . Instead of listing all of them, we give the list corresponding to the points with x — 0 or y = 0. The remaining points are obtained using the symmetries x <—► z, y <—* w and x <—> —x, which are made more precise below. P5:x
= 0,y = - ( ^ y \ z
P7:x
= 0,y = - ( ^ )
P9:x = - ( ^ ) Pu:x
1 / 2
=~ { ^ y
/
1 /
\ z
,y^0, \ y
={ ^ ) ( ^ y \ =e - ^ ) ( ^ y
2
/
\
W W
= -x, U , = ( l ^ ) ( ^ )
= 0,z = -x,W
=
= -l = -y-,
(59)
1 / 2
( ^ ) ( l ^ y
/ 2
The points Pg, Ps, Pio and P\i are the negatives of P5, P7, Pg and Pg, respectively. There are 6 points with no null coordinate. The first two points have index - 1 and the remaining four points have index 1 P21 : x = - 5 , y = - \ , z=\,w= i P22 : x = i , y = | , z = - i , w = - \ Pn:x
= -l/2(^)1/2;z
P25: x = - l / 2 ( ^ )
= l/2(^)1/*;x 1 / 2
;
2
=l/2(^)
= y,z = w 1
' ; x = y, z = w
P26 = - P 2 5
I f a = 6 = c = d = 0, then the equilibrium points are:
° i : (°'75' 0 '-^)' <>2 = -Ou 03:( ^ , 0 , - ^ , 0 ) , 04 = - 0 3 , O5 = (31 2' ~ 2 ' ""2)' ^
6
5
~ ~® '
The points O j , 0 2 , 0 3 and O4 have index 1, while 0 5 , 0 6 have index - 1 .
(61)
158
If a = b = c — d = —1/2, we get Qi = ( 0 , 1 / A 0 , - l / > / 2 ) ; Q 2 = - Q i , Q3 = (l/v/2, 0, - 1 / >/§, 0 ) ; Q 4 = - Q 3 > Q5= Q7
=
^
,
^
(
( f i ^ ! ,
^
,
ffi^!,
^
)
;
fi#l,
Q 9 = (1/2,1/2, - 1 / 2 , - 1 / 2 ) ; Q 1 0 -
Q
fi#!)
6
= -Q5,
(62)
, Q8 = _ Q 7 ,
-Q9.
All these points but Qg and Qio whose index is - 1 , have index 1. Notice that there two symmetries for the vector field (56). One of them is the symmetry <Si sending x <—► -x. The other one S2 ■ x <—► z, y *—► w, t 1—» —t is valid only when the parameters satisfy a = c and b = d. These two symmetries explain why the equilibrium points of the vector field appear in groups of 2 and 4 points. To compute the index for each equilibrium point, we give a chart of E°° about that point, which is equivalent to a choice of two independent variables, so that (56) becomes a system of two equations in the plane. Then one has to compute the eigenvalues of the linear approximation of the resultant planar system. We remark that g) a 3 j.( is a homogeneous function of degree 3, since it is a determinant where all the terms of the first row are homogeneous of degree 2, while the terms of the second row are linear. Let x*, xs denote the other 2 complementary variables. If $\x3'xl / O w e can take x^ and xs as independent variables, so that we can solve x< = Xi (x*, xa) and Xj = Xj (x*, x3). Computation of the partial derivatives J^-, | ^ , g^i and JEi show that they are quotient of homogeneous polynomials of degree three, and therefore they are even functions. On the other hand, the right hand sides Fi of system (56) are polynomials containing terms of degrees two and four. Hence, they are even functions. Then their partial derivatives of the first order are odd functions. In a chart about an equilibrium point where x^, xa can be taken as inde pendent variables, system (56) has the form ±k = Fk(xk,xs,Xi(xk,x3) X, = F, (xk,Xa,Xi
,Xj(xk,xa))
= $ f c (x f c ,x s ),
(Xfc,X,) ,Xj {xk,Xa))
= $a (Xfc,X s ) .
,g3.
The entries of the matrix of its linear approximation are clearly odd functions, since they have the form
^^dFk dxa dxa
+
dFkdx1 dxi dx,
+
dFkdx1 dxj dxa'
159
etc. The index of the equilibrium point is obtained from the sign of the determinant _ d$k d $ 3 dxk dxs
d$k d$3
dx, dxk '
(65)
which is an even function. When we apply symmetry <Si to an equilibrium point, we obtain a new equilibrium point with the same index, since V is invariant under S\. Let us apply symmetry 52 to one of the equilibrium points, from the cases that we are considering. By straightforward verification we see that V is invariant as well, so that, the new equilibrium point has the same index. Hence, for the component with 26 equilibrium points, is was enough to analyze only 9 of them. We conclude that the topology in the three components can be described as follows. T h e o r e m 16 Let H3 (x, y, z, w) = x 3 + y3 + z3 + w3 + (ax + by + cz + dw)3 a homogeneous cubic polynomial and W=(H$ — 0) n S3. Then a) W is diffeomorphic to a sphere if a=b=c=d=0. b) W is diffeomorphic to the disjoint union of three spheres if a=b=c=d= -1/2. c) W is a connected surface of genus 6 if a=b=c=d=-l. Proof: According to propositions 13 and 14, we have connected surfaces in cases a) and c). In case a) we have a sphere because the global index is 2, so that the genus is zero. In case c) the global index is -10, hence the genus is 6 as asserted. Finally, in case b) the global index is 6 and recalling that the genus must to be even (see Corollary 6), the only possibility is that the surface has three connected components of index two, i,e. spheres. ■
References 1. V. Arnold, A. Varchencko and S. Goussein-Zade\ Singularites des appli cations differentiates, MIR, (1986). 2. A. Cima and J. Llibre, Bounded polynomial vector fields, Trans. Amer. Math. Soc. 318, 557-579, (1990). 3. J. Delgado, E. Lacomba, J. Llibre and E. Perez, Poincare compactification of Hamiltonian polynomial vector field, in Hamiltonian Dynamical Systems, History and Applications The IMA vol. in Math, and its appli cations 63 , 99-115, Springer-Verlag, (1994). 4. A. Dold, Lectures on algebraic topology, Springer-Verlag, (1972).
160
5. M.Falconi, Estudio asintdtico de escapes en sistemas hamiltonianos polinomiales. Ph.D. Dissertation, UNAM, (1996). 6. M. Falconi, E.A. Lacomba and J. Llibre, Topological Classification of Real Cubic Surfaces in the 3-Dimensional Sphere. To appear. 7. S. Lopez de Medrano, Involutions on manifolds Springer-Verlag, (1971). 8. B. Segre, The non-singular cubic surfaces, Oxford Press, (1942).
RELATIVISTIC CORRECTIONS TO ELEMENTARY GALILEAN D Y N A M I C S A N D DEFORMATIONS OF POISSON BRACKETS RUBEN FLORES-ESPINOZA Departamento de Matemdticas, Universidad de Sonora, Rosales y Blvd. Luis Encinas, Hermosillo, Sonora 83000, Mexico. E-mail: [email protected] YU M. VOROBJEV Department of Applied Mathematics, Moscow Institute of Electronics and Mathematics, B. Vuzovsky per. 3/12 Moscow 109028, Russia. E-mail: [email protected] Deformations of Poisson brackets generated by the contraction of the Poincar6 algebra into the Galilei algebra are studied in the framework of the Poisson cohomology theory. We show that relativistic deformations of the Galilean Poisson structure are trivial, in particular, the cohomology class of the relativistic 2-cocycle is zero. This fact allows us to interprete an elementary relativistic Hamiltonian dynamical system as a Hamiltonian system on a homogeneous symplectic space of the Galilei group with a corrected Hamiltonian.
1
Introduction
The elementary relativistic (nonrelativistic) mechanics as the theory of dy namical systems on the homogeneous symplectic spaces of the Poincare group (the Galilei group) was developed in the fundamental work of Souriau 1 . Vari ous applications of this approach can be found, for example, in2-3>4>5>6. We are interested in relativistic effects coming from the contraction of the Poincare algebra into the Galilei algebra from the viewpoint of Poisson geometry. The Poincare group is generated by the Lorentz transformations and translations of the Minkowskii space. Its Lie algebra g is the semidirect sum of the Lie algebra of the Lorentz group and the 4-dimensional abelian algebra, j = o(3,l)jR4. So, the Poincare algebra g can be denned as a 10-dimensional vector space V consisting of all vectors of the form X = (A,v,x,t), where A € o(3), v,x, G R 3 , t € R, and the Lie bracket is given as 3,5 [Xi,X 2 ] = nAi,A 2 ] +vi ®v 2 -v2®vi,
Aiv2 -
A2v\,
Aix2 - A2xi + t2v\ - t\v2, vi ■ x2 - v2 • xA.
161
162
For each e > 0, define the vector space automorphism U€ : V —> V by (A, v, x, t) - - (J4, £ V V e 1 / 2 x, t).
(1)
Since C/£ is a vector space isomorphsim, we can defined a "new" Lie bracket on V as \X1,X2}e
U£-1[Ue(X1),Ue(X2)}.
=
The bracket [, ] e continuously depends on e including e = 0. For £ > 0, the Lie algebra ge = (V, [, }e) is isomorphic to the Poincare algebra, whereas for £ = 0, go is the Galilei algebra. This procedure is just the contraction of the Poincare^ algebra into the Galilei algebra 5 . Physically, the parameter e is equal to c - 2 , where c is the speed of light, and £ - > 0 corresponds to the nonrelativistic limit. Let V* be the dual of V. For each e > 0, consider the Lie-Poisson space g* = (V*,{, }£) equipped with the standard linear Poisson bracket {, }£ : C°°(V*) x C°°(V) — C°°(V) associated with the Lie algebra gE. We can modify the family of linear Poisson brackets { , } £ into a 2-parameter family of affine Poisson brackets { , }£<m on V in the following way. It is well known 1,3,5 that the antisymmteric 2-form on V \{Xl,X2)=vl-x2-v2-xl
(XUX2GV)
(2)
is a 2-cocycle of the Lie algebra g£ for each e > 0. Moreover, the second cohomology space H2(go) of the Galilei algebra go is one-dimensional and is generated by the cohomology class [A] of A. Thus the general second co homology class in H2(go) can be written as m[\], where, from the physical viewpoint, the parameter m is identified with the "mass" of the Galilei group (for details, see 3,5 ). For e > 0 (i.e., for the Poincare group), [A] = 0 in H2(ge). Now taking into account the general fact 5,8,10 : any 2-cocycle of a Lie algebra generates an affine Poisson bracket which is a deformation of the linear Poisson bracket, we obtain the family of affine Poisson brackets { , }CyTn on V* associated with the 2-cocycle m[A]. Using coordinates (A,x,v,t) on V in (1), we can identify the dual space V* with the Eclidean space R 1 0 = {£ = (J, K, P,H)}, where J,K,P G R 3 and H G R. In terms of these dynamical variables, the affine Poisson brackets { , } £ ] m are given by the relations 4 : {H,Ji)t,m
= {H,Pi)£<m
= 0,
=
\Ji,Jjfe,m GjjA: Jfe, {Ji, Pj}e,m = 6»jfc Pk, {#i, Pj}c,m = (eH + m)«y,
where Gij* is the Levi-Civita tensor.
{H, Ki}c<m = —Pi, \JiiK-i)e,m = €ijk "fc> {-^i, Kj}£<m = —£ G^fc Jfc, {Pi, Pj}c,m = 0,
(3)
163
Our goal is to investigate the family of Poisson brackets context of the deformation theory and Poisson cohomology7'8 applications 11 ' 12 ' 13 . For a fixed m > 0, we iterpret the Poisson brackets {, , } £ , m mation of the bracket {, , }o,m, where the parameter e plays the deformation parameter. In invariant terms, the Poisson bracket represented via the Poisson tensor r £ , m = r + mA + eTl.
(3) in the and their as a defor role of the (3) can be (4)
Here the bivector field T on V* represents the "limiting" bracket {, , }o,o> A corresponds to the cocycle A in (2), and the bivector field TZ defines the relativistic deformation of the Poisson structure ro, m . We have the following coordinate representations for 72. and A:
n
= -\^k]kwiAw-l + m>>wiAwl'
A=
<5ii »j
d Wi
d 8P/
(5) (6)
A
Since the brackets (3) satisfy the Jacobi identity for all £, the bivector field 72. satisfiew, the linearized Jacobi identity7 relative to ro,m- In the lan guage of the Poisson cohomology, this means that 71 is a 2-cocycle of the Poisson manifold (V'*,ro, m ). For convenience, the Poisson structure To,m and the bivector field 71 will be called the Galilean Poisson structure and the relativistic 2-cocycle, respectively. Our main result is that the Poisson cohomology class of the relativistic 2cocylce 71 is zero, or in other words, the infinitesimal relativistic deformation of the Galilean structure r 0 , m is trivial. For the case m = 0, this fact is not evident a priori by the following reasons: the Poisson structures T and r £ i o are not isomorphic and the second Poisson cohomology space of the Galilean structure T is not trivial (see §2). This result is applied to the investigation of elementary relativistic (nonrelativistic) dynamics1, that is, a dynamical system on a connected symplectic manifold, where the Poincare group (the Galilei group) acts transitively by symplectomorphisms. We consider the following family of Hamiltonian systems on V* associated with affine Poisson brackets (3):
j = JxVjf K = -$Lp
+ Kx VKf + Px + KxVjf-eJx
VPf, VJC/ - (eH + m)V P /,
(7)
164
P = PxV,f H = P-
+ (£H +
m)VKf,
VKf,
where the symbol x denotes the cross-product in R 3 and / = f£(0 G C°°(V*) is the Hamiltonian function smoothly depending on the parameters e. For e = 0, equations (7) represent a collection of the elementary Galilean (nonrelativistic) systems: the trajectories of (7) lie on the homogeneous symplectic spaces (symplectic leaves) O of the Galilean group. If m = 0, then each submanifold O C V* is a coadjoint orbit of the Galilei group. Under m > 0, O is an orbit of the affine representation of the Galilei group on V*. For e > 0, equations (7) describe an elementary relativistic dynamics whose spaces of motions coincide with the coadjoint orbits of the Poincare group. Following the general thesis: (deformtations of Poisson brackets) =>• (corrections to Hamiltonian functions) and, using the resutls 7 ' 8 ' 11 , in §2 we construct an infinitesimal transformation which transforms the relativistic system (7) into a nonrelativistic Hamiltonian system, up to 0 ( e 2 ) , with a corrected Hamiltonian. The generator of this infinitesimal transformation is determined by the primitive of the relativistic 2-cocycle TZ, for which we give an explicit formula. Moreover, in §3, for m > 0, from the Sudarshan-Makunda transforma tion 4 we deduce an isomorphism between relativistic dynamics and the ele mentary Galilean dynamics defined in sufficiently large domains of V*. This fact may be useful for a Hamiltonian description of a classical motion of a relativistic spinning particle in an external electromagnetic field ] . A Hamil tonian approach for systems of such kind based on the twistor technique was suggestied in 6 . 2 2.1
Deformations of the Galilean Poisson structure Basic notions
Here we recall some necessary facts from the Poisson geometry. A more ex tensive information can be found in 7 ' 8 ' 9,10 . Let M be a C°°-manifold. A skew-symmetric contravariant 2-tensor field (a bivector field) $ on M is called a Poisson tensor if the formula {f,g} =
(8)
defines a Lie bracket on the space C°°(M) of smooth funcitons on M. In this case, the bracket (8) is said to be a Poisson bracket. The pair (M, ty) is called a Poisson manifold urith the Poisson structure ^ .
165 Let (M, SP) be a Poisson manifold. The Hamiltonian vector field of / € C°°(M) is vj = -\P(d/). Dynamical systems generated by Hamiltonian vector fields on M are called Hamiltonian systems on the Poisson manifold. The symplectic leaf of a Poisson structure ^ through a point m € M is the set Cm of all points m! e M such that m and m' can be joinded by a piecewise smooth curve each segment of which is a trajectory of a Hamiltonian vector field. So, £ m is a connected submanifold of dimension d i m £ m = r a n k ^ ( m ) . Each symplectic leaf carries the symplectic structure u; with the property: w(w/,» 9 ) = {/,} ^ ^1 /, € C°°(M). A point m € M is said to be regular if the rank of a Poisson structure ^ is locally constant, otherwise, m is singular. A symplectic leaf through a regular point is also called regular. So, M is a disjoint union of symplectic leaves. A functon k on M Poisson commuting with any function (that is, $(dk) = 0) is said to be a Casimir function. Casimir functions are constant along symplectic leaves. Let V* (M) be the space of skew-symmetric contravariant A;-tensor fields on M. The Jacobi identity for the bracket (8) is equivalent to the following equation for V:
I*,*1=0,
(9)
where [, ] : Vk(M) x V m (M) -» Vk+m~1(M) is the Schouten bracket7**10. This operation is a natural extension of the usual commutator between vector fields on M. In particular, for any f,g € C°°(M), [*, fj = vf, [[ [*, /],$]] = —{f,g}, and the Schouten bracket between # and a vector field w — wxd/d^ on M is given by the formula [*,wY> = * " 5 t t ^ - V'diW* -
w'ds*ij,
where tf = $<J'ft A d,- and ft = 3 / 9 ^ . It follows from (9) that a Poisson tensor ^ induces the coboundary op erator Z? = [*,•] : Vfc(M) -> V fc+1 (M), £>2 = 0, which is called the Lichnerowicz-Poisson cochain of the Poisson manifold (M, ty). The space of fc-cocycles Z f c (M,*) consists of all A e Vfc(M) such that DA = 0. A cocycle .A is a coboudary if DB = A for a certain i? € V f c _ 1 (M). The space of coboundaries is denoted by Bk(M, \t). Then W*(Af, * ) = Zk(M, *)/Bk(M,
*)
are called the Poisson cohomology spaces of (M, \I>). We have W°(M, ¥ ) = {Casimir functions}. A vector field z on M is called a Poisson vector field of its flow preserves the
166 Poisson bracket, that is, if Jtf,*! = 0. So, Hl(M,V) = {Poisson vector fields/Hamiltonian vector fields}. A bivector $ € V 2 (M) is called an infinitesimal deformstion of a Poisson structure \P if the bracket {f,9h
= {f,g} + et>(df,dg)
(eeflU])
2
satisfies the Jacobi identity up to 0{e ). An infinitesimal deformation $ is said to be trivial if there is an infinitesimal transformation £ >—> £+ew + 0(e2) generated by a vector field w on M, which transforms {, } into {, } £ up to 0(£2). L e m m a 1 A bivector field $ on M is an infinitesimal deformation of # if and only i / $ is a 2-cocycle, $ € Z2(M, # ) , and the triviality of$ is equivalent to the condition: $ is a coboundary, $ € B2{M,^), whose primitive is just the generator w of the above infinitesimal transformation. So, we have H2(M, ty) = {infinitesimal deformations/trivial infinitesimal deformations}. A submanifold N of (M, \P) is called a Poisson submanifold if any Hamiltonian vector field on M is tangent to N. In this case, the Poisson tensor on N is given by the restriction of ^ on N. We say that a Poisson submanifold N in M is regular if N is a union of regular symplectic leaves of (M, $ ) . A mapping 7 : M —» M between two Poisson manifolds (M,\P) and (M, * ) is called a Poisson mapping if ^*{^!{df,dg)) = ^{^'df^'dg) for all f,g € C°°(M). If / is also a diffeomotphism, then / is said to be a Poisson automorphism or equivalence. Two Poisson manifolds (M, ^ ) and (M, \P) are globally equivalent if there is^a Poisson automorphism between some regular Poisson submanifolds N C M and N C M. Note that if the topological struc ture of symplectic leaves of \& and # is different, then this is an obstruction to the global equivalence. A smooth deformation of a Poisson structure f on M is an e-family of Poisson tensors Hf£ on M smoothly depending on a parameter e G [0,1] and such that \I>J „ = * . In this case, $ = 4sH „ is an infinitesimal deformation of \P. 2.2
Symplectic leaves ofT£<m
Consider the Poisson tensor r £ ) m (4) of the affine Poisson bracket (3). There are two independent Casimiar functons of r e > m smoothly depending on pa rameters e > 0 and m > 0: CiJJ, = -p2
+ eH2 + 2mH,
(10)
167 Table 1. Topological structure of regular symplectic leaves
values of parameters
symplectic leaves
e > 0, m > 0; a > - m 2 / e , 0 > 0
e = 0, m > 0; a € R, 0>O
£ > 0, m = 0 ; a < 0 ,
0>O
e = 0, m = 0; a < 0, 0>O
Ci% = {(eH + m)J~Kx
topological type
o±
T*R 3 x S 2
Oo,m
T*R 3 x S 2
e,0
R 5 x S2 x S1
00,0
R 5 x 50(3)
u
P)2 - e(J • F ) 2 .
(11)
2
Here the terms such as P and J ■ P mean the norm and the inner product of vectors in the 3-dimensional Eucledian space. Funcitons (10) and (11) are related to the physical invariants of the Poincare group by the formulas Q,m = (M 2 — m2)/e and Ci,Jn = M2s2, where e~2M2 is the mass squared and s2 is the spin squared (see 1,2,5 ). The 8-dimensional regular symplectic leaves of the Poincare structuer Tem are given by the connected components of the regular level sets of the Casimir functions (10) and (11):
0 £ , m (a, 0) = {Ci'X = a, Ci% = 0>O}cV'^
R 10 .
(12)
For e — 0, submanifolds (12) are connected. For e > 0, each level set (12) has two connected components Ofm = Ofm(a,0): Otm = Oc,m n {sign {eH + m) = ± } .
(13)
The topological structure of the leaves depends on the values of parame ters e and m. We give the corresponding classification in Table 1. For 77i > 0, the topological type of the symplectic leaves does not change as e —♦ 0. Moreover, each Ofm or £>o,m is isomnorphic, as a symplectic manifold, to the direct product T*R 3 x S 2 , where T*R 3 is equipped with the canonical cotangent symplectic structure, and the symplectic structure on S 2 is defined by the area 2-form (see 2 ' 4 ' 5 ). The limit 0 —* 0 corresponds to
168
the singular points of the Poisson structure ro, m - If e > 0, then the singular points of r e i 7 n correspond to the levels: a = -m2/e (the massless case) or (3 = 0 (the spinless case). For m = 0, there is another picture: the symplectic leaves 0£ 0 and Oofi are not diffeomorphic. In particular, this implies that there is no global equivalence between the Poisson structures re>o and F. The singular points of T£io and T correspond to the critical values of the Casimir functions: a = 0 and (3 = 0. 2.3
Relativistic 2-cocycle
(i) Case m = 0. Let (V*, T) be the Poisson manifold, where T is the Galilean Poisson tensor in (4). Consider the regular Poisson submanifold (NQ, T) of V*, where N0 is the union of the 8-dimensional regular symplectic leaves (12),
N0=
(J
0o,o (a,/?).
(14)
a
First, note that the vector fields
z
* = K-m+p-w>
Z2 = (Px(Px
K)) ■ A
(15) (16)
are linear independent Poisson vector fields of (iV~o,r) transversal to symplec tic leaves C?o,o- So, 7,13 n\N0,T)
w C°°(R 2 ) x C°°(R 2 ).
(17)
Using this fact and the general results on computing Poisson cohomology 7 ' 8 ' 10 ' 12 , we prove that the Poisson cohomology space Ti2(No,T) is not trivial, W 2 (iVo,r)wC 0 0 (R 2 ).
(18)
Theorem 1 The cohomology class of the relativistic 2-cocycle TZ (5) in H2(N0,r) is trivial: TZ is a coboundary, IT,W\=1Z, where the primitive W is the following vector field on N0: 2P2\
dK
DP)
(19)
169
P2\ +
P2
+iLf£^pnyPx{PxK)).».
p
P2{IfxP)2(-
V-K)-r(P-K)(P.J))(KxP).^.
(20)
Proof. We are seeking a solution of equation (19) in the form: W = a\Z\ + a2Z2 + b(K x P) • -J^, where Z\ and Z2 from (15) and (16), a\,a2,b G C°°(NQ) are some functions. Taking into account the relations
pi,p2} = ir,H2j, pi, (K x P)2fl = [r, 2HJ -(KxP)
+ (J. P)%
and evaluating equation (19) on the Casimir functions of T, we obtain a\ = H2/2P2 and a2 = -^{^ - 2\JKSy be found by direct computations.
- "(K^pp)-
Finally, the function b can
Corollary 1 The relativistic cocycle TZ defines a trivial infinitesimal defor mation of the Galilean Poisson structure T. The most general primitive of TZ has the form w = W + k\Z\ + k2Z2 + (Hamiltonian vector
field),
(21)
where k\ and k2 are arbitrary Casimiar functions oJT. Note that the primitive w (21) of T cannot be smoothly extended to the whole space V*. Indeed, if a point £ = (J, K, P, H) 6 NQ is moving to the set of singular points So = {K x P = 0}, then the vector field W becomes singular. The same effect takes place in several other examples 11 ' 12,13 . (ii) Case m > 0. Let To.m = T 4- mA be of the modified Galilean Poisson structure on V* defined in (4), (6). First, we observe that Z = 0/dH
(22)
is a Poisson vector field of Toim transversal to the symplectic leaves C?o,mProposition 1 For any m > 0, W1(^*,r0,m)«Coo(R2)
and n2(V*,T0,m)
Moreover, the cohomology class of 11 in H2(V*,ro,m) 11, where
« C°°(R 2 ). is trivial, [To,™, Wl =
V.-H«+ZK»+itPx(J+LPxK)).» m on
m
oK
2m V
p., V
m
//
oK
170
This fact can be derived from the resutls of §3. It is interesting to note that, by comparing with (20), formula (23) gives the primitive of r 0 ) m , which is well defined in the whole space V. But if m - » 0 , then W has no limit. So, adding the "mass" factor A to the Galilean structure T leads to qualitative changes in the Poisson structure. 2.4
Infinitesimal relativistic corrections
Now we are going to apply Theorem 1 for the investigation of the family of dynamical systems (7) as e —► 0 and m = 0. For a given function / = / £ (£) on V*, we consider system (7) as a perturbation of the limiting Hamiltonian system for e = 0. In general, the corresponding perturbation vector field is not Hamiltonian with respect to the Galilean Poisson structure T (see also 7,11 ). Denote by (V*, r £ ) o, / ) the dynamical system corresponding to system (7) with e > 0 and m = 0 and by (NQ,T, f) the dynamical system corresponding to (7) with e = 0 and m = 0, where the domain NQ is defined in (14). Proposition 2 Let fe € C°°(V*) be a function smoothly depending on e € [0,1]. Then the infinitesimal mapping £~Z
+ ew(t) + 0(e2)
(ZeN0),
(24)
where the vector field w is determined by (21), transforms the nonrelativistic Hamiltonian system (No, I\ fo+£f^), up to 0(e2), into the relativistic system (V, r ffi o, f c ) , where the "relativistic correction" to the Hamiltonian /o is given by the formula
/"> = Wo + f
(25) £=0
The proof of this fact follows from Corollary 1 and Lemma 1. The same fact is valid in the case m > 0. But, as we will see in §3, in this case there is a global equivalence between relativistic and nonrelativistic dynamics. 3
Equivalence between elementary relativistic and nonrelativistic Hamiltonian dynamics. Case m > 0
Let us consider the e-parameter family of affine Poisson brackets (3) for a fixed m > 0. Let (V*,ro,m) be the Poisson manifold associated with the "limiting" bracket (3) with e = 0, and (V*,TSim) be the Poisson manifold associated with the "deformed" bracket (3) with e > 0.
171
For any e > 0 and m > 0, using symplectic leaves (12) and (13), we introduce • the regular Poisson submanifolds G m and G £ , m of (V, To.m) as Gm=
U
O0,m(a,p)cV%
(26)
Q6R,/3>0
G£,m =
(J
O0,m(a,/9)cGm;
(27)
a>-m2/e,/3>0
• the regular Poisson submanifolds Vfm ^ m =
U
^
m
and V~m of (V*,r £ i m ) as
(a^),
P + m n P ~ m = 0.
(28)
a>-m2/e,/3>0
Note that G m , G £ , m , and Vfm are open 10-dimesional submanifolds in V*. In the limit as e —► 0 the submanifolds G £ , m and P ^ coincide with Gm. However, each point of Vfm goes to infinity as e —» 0. ^From the viewpoint of the dynamical system (7), these submanifolds have the following physi cal interpretation. The domain Gm (or P^m) can be viewed as a union of the symplectic phase spaces of nonrelativistic (or relativistic) systems corre sponding to various nonzero values of the spin (that is, for /3 > 0). For small e > 0, the invariant domains G£iTn and V+m approximate Gm and V~ is not essential. Now let us consider the map 7 £ : (J,K,P,H) i-» (J', K', P', H') defined by the formula H' = - ( - m + yj2m.eE + m 2 ), J' = J, K,
=
(29)
V2meH + m*K m
+
eP x (J - ±K x P) V-eP2 + 2meH + m2 + \f2meH + m2'
P ' = P. It is easy to see that 7 £ gives a C°°-map from G £ i m onto V+m. Theorem 2 For any e > 0 and m > 0, formula (29) defines a diffeomorphism le : G £ , m - V+m such that 70 = idGm and ^€ is a Poisson automorphism between the Poisson manifolds (G £ , m ,r 0 | m) and (P£ m ,r c>TO )>
(7e)»ro,m = r £>m . The transformation 7 £ symplectomorphically maps each symplectic leaf Oo,m(ot,(3) onto the leaf 0 £ > m (a, \/a2 + m2(3/m).
172
Formula (29) can be derived from the Sudarshan-Mukunda realization 4 . On the other hand, (29) is verified by direct computations. Note that as e —» 0 the infinitesimal part of the transformation (29) is just given by the vector field W in (23). Now Theorem 2 allows us to establish the following relation between the relativistic and nonrelativistic Hamiltonian dynamics associated to sys tem (7). For a given f£ e C°°(V*) smoothly depending on e S [0, oo), by (P+ ffl , T£ym, f£) we denote the restriction of the Hamiltonian system (7) with e > 0 to the invariant Poisson submanifold P + m C V*. The restriction of system (7) with e = 0 to the invariant Poisson submanifold G £ , m is denoted by (G e ,m,ro lT n,.F £ ), where Ft is a certain Hamiltonian function. Proposition 3 For every e > 0 and m > 0 the relativistic Hamiltonian sys tem ( P + ^ r ^ n n / e ) is isomorphic to the nonrelativistic Hamiltonian system (G£,m^o,rmFc), where the "corrected" Hamiltonian function Fe is given by
F£ = -y€of£ = f0 + e(Lwf0 + ^
)+0(e2).
(30)
Here 7 £ is defined by (29) and W is the vector field in (23). So, as £ —♦ 0 a nonrelativistic dynamical system can be interpreted as the Hamiltonian perturbation (30) of the nonrelativistic system ( G m , r o , m , / o ) This makes it possible to apply the pertubation theory of Hamiltonian sys tems, like the KAM theory, to the investigation of relativistic dynamical effects (see also 11 ). Acknowledgments The authors wish to thank R. de la Llave and H. Cendra for helpful conver sations and suggestions concerning this work. One of the authors (Yu. M.) was partially supported by the Consejo Nacional de Ciencia y Tecnologia de Mexico through the Catedra Patrimonial de Excelencia II, No. 940036. References 1. J.-M. Souriau, Structure of dynamical systems. A symplectic view of Physics, Progress in Mathematics, Vol. 149, Birkhauser, Boston, (1997); Structure des systemes dynamiques, Dunod, Paris, (1970). 2. N.M. J. Woodhouse, Geometric Quantization, Second Edition, Clarendon Press, Oxford, (1992).
173
3. V. Guillemin and S. Sternberg, Geometric Asymptotics, Mathematical Surveys, Vol. 14, Amer. Math. Soc, Providence, RI, (1977). 4. E.C.G. Sudarshan and N. Mukunda, Classical Dynamics: A modern per spective, R.E. Krieger Publishing Co., Malabar, Florida, (1983). 5. V. Guillemin and S. Sternberg, Symplectic Thecniques in Physics, Cambdrige Univ. Press., Cambdrige, (1984). 6. A. Bette, J. Math. Phys., 22 (6), 2158-2163, (1992). 7. Yu. M. Vorobjev and M. V. Karasev, Functional Anal. Appl., 22, 1-9 (1988). 8. V.P. Maslov and M.V. Karasev, Nonlinear Poisson Brackets. Geometry and Quantization, Translations of Mathematical Monographs, V. 110, Amer. Math. Soc, Providence, RI, (1993). 9. A. Weinstein, J. Differential Geometry, 18, 523-557, (1983). 10. I. Vaismann, Lectures on the geometry of Poisson manifolds, Progress in Mathematics, V. 118, Birkhauser, Berlin, (1994). 11. R. Flores-Espinoza and Yu.M. Vorobjev, in New Trends for Hamiltonian Systems and Celestial Mechanics, Eds. E. Lacomba and J. Llibre, World Scientific, London, 357-374, (1995). 12. R. Flores-Espinoza, Yu.M. Vorobjev, M.V. Karasev, in Quantization, Co herent States and Poisson structures, Eds. A. Strasburger et al., Polish Scientific Publishers PWN, Warsaw, 233-240, (1998). 13. R. Flores-Espinoza, Yu.M. Vorobjev, Russian Mathematical Surveys, 49, 6, 223-224, (1994).
HETEROCLINIC P H E N O M E N A IN T H E SITNIKOV PROBLEM ANTONIO GARCIA AND ERNESTO PEREZ-CHAVELA Departamento de Matemdticas, UAM-Iztapalapa, Apdo. Postal 55-5S4 Mexico DF. 09340, Mixico. E-mail: [email protected] and [email protected] We give the deduction of a Melnikov function for the Sitnikov problem. Using a perturbation method introduced by Melnikov and a thorough analysis of the geometry of certain auxiliary functions that we introduce, we prove analytically the existence of transverse heteroclinic orbits. As a consequence, we can embed a Bernoulli shift near these orbits. This shows that the Sitnikov problem possesses chaotic dynamics.
1
Introduction
In 1929, J. Chazy 2 predicted the existence of oscillatory motions in the threebody problem, that is, orbits whose behavior are unpredictable for long time intervals. K. A. Sitnikov11 was the first to answer positively this question giving a particular model. This kind of motions was studied widely by V. Alekseev1. However his work is very difficult to read. In 1973, J. K. Moser8 retook the Sitnikov model and was able to put the ideas of V. Alekseev in a more tractable way. The Sitnikov's model consists of two equal masses describing elliptic or bits in a fixed plane denoted by II; and a third massless particle moving on the orthogonal line to the plane through the center of mass of the former masses. J. K. Moser studied the Poincare map denned assigning the position of the primaries and the velocity of the third particle when it crosses II to the corresponding ones in the next crossing of the third particle with II. Taking the eccentricity of the ellipses as a parameter e, J. K. Moser showed that the problem is integrable for e = 0; and that for small e > 0 it is possible to embed a Bernoulli shift near the parabolic orbit of the massless particle, for the Poincare map associated to the flow. The basic fact used to prove this result is the existence of transversal homoclinic points. In this paper we prefer to use the time 27r-map as a Poincare map, and write the system modelling the Sitnikov problem as a perturbed Hamiltonian system, with the eccentricity e as perturbation parameter. For e = 0, the system is integrable and the level curves of the unperturbed Hamiltonian become the orbits of solutions. These curves are divided in two families, the
174
175 first one consists of bounded curves which correspond to the periodic orbits, and the second one consists of unbounded curves. Their mutual boundary consists of two special orbits, called the parabolic orbits, which escape to plus infinity and minus infinity with zero limiting velocity. Using the ideas of McGehee6, we prove that for e > 0, the points corre sponding to escape to ±00 can be seen as degenerate critical points for the flow. In order to clarify the panorama, and to distinguish the escapes above the plane with the escapes below, we have done our analysis in three different coordinate charts, one containing the origin in the phase space and the other two containing +00 and —00. In this framework we prove that the parabolic orbits still exist for e > 0 small enough, and that they act as heteroclinic orbits joining the two critical points at ±00. In fact McGehee proved, using a suitable change of coordinates, that the degenerate critical points are trans formed into periodic orbits, and that the stable and unstable sets of these periodic orbits are real analytic submanifolds. The aim of this paper is to prove, using the Melnikov techniques that the above analytic submanifolds intersect transversally along the heteroclinic or bits for e > 0 small enough. A similar result has been obtained by Dankowicz and Holmes3, however our approach is different. We never use elliptic func tions in order to compute the Melnikov function. But above all, the geometry of some auxiliary functions that we introduce and some basic computations allows us to prove analytically that the zeros of the Melnikov function are sim ple; this is our main result. We also give a detailed analysis of the deduction of the Melnikov function. 2
The Sitnikov Problem
In this paper we reanalize the classical problem studied by K. A. Sitnikov11 in 1960. We consider two non zero equal masses mi and TT12 moving around each other on ellipses of eccentricity e in a plane II, and a massless particle 7713 moving on the axis E perpendicular to the plane II passing by the center of mass. As in Moser8, we normalize the time so that the period T of the primaries is 27r and the gravitational constant G = 1. Then from the Kepler's third law, n2a2 = G(rn\ + m?) where n = 2it/T = 1. Hence if a = 1 then we must have mi = mi = 1/2. Now since the center of mass is at the origin, thus when the eccentricity e = 0, the primaries are moving in circles of radius 1/2 so that its mutual distance is one. With these hypotheses the motion of the first two particles is not afected by the motion of the massless particle which always remains on the axis E (see figure 1).
176
Figure 1. The Sitnikov Problem
Let x be the coordinate on the axis E of the massless particule, its motion satisfies the equation X
X
=
|*3
(2.1)
+ r2 , 3 / 2 '
where r (t) is the distance of each of the primaries to the center of mass located at the origin. Assuming that at time t = 0 the primaries are in the pericenter, we have r(t) = r(—t). The expansion of r(t) with respect to the eccentricity is given by (2.2)
r(t) = - (1 + ecost + h.o.t.) Taking x = y, the equation (2.1) becomes
GW-T
(2.3) 0
y ■ ( -
X
.
(.'+J)''
)
■
3x cos (
:-+»••» )
+
0 xB(x,i i.oy' («"
177
where B(x, t, e) is smooth and 27r-periodic with respect to t. In addition, (2.3) is a Hamiltonian system with Hamiltonian function given by H{x,y,t)
-y2-
=
1
(2.4)
2
(x + r 2 ) 1 / 2
= HQ(x,y) + eHi(x,y,t)
+ h.o.t.,
where H0{x,y) = -y2 Hi(x,
1
-
2
(* + i ) 1 / 2 >
cost 4(x ,
(2.5b)
1)3/2-
+
(2.5a)
Of course, if e — 0 the system (2.3) is integrable and therefore we may obtain explicitly the global flow. There are two possibilities for 7713 to escape: by above, when x —► +00; and by below, when x —> —00. To study the flow near these regions, we follow McGehee6 and Moser8 and make one of the following changes of coordinates: 1 X =
r
V = -P,
when x > 0,
y = p,
when x < 0.
1 X =
9 >
(2.6)
The equation (2.3) is transformed in: ■
q
1 3
~ 2q
P
'
n P
^
|l + rV! 3 / 2 '
(2.7)
The equation (2.7) can be extended to the points q = 0, these points can be thought as points at plus or minus infinity, each one associated with a different limit velocity. Once this extension is done, the phase space is covered with three different coordinate charts. In the first one containing the origin, we study the bounded motions governed by the equation (2.1). In the other two charts corresponding to the escapes to +oo and - c o respectively, the dynamics are given by the equation (2.7). We remark that the line x = 0 and the line y = 0 (or p = 0) are axes of symmetry; moreover, they are orthogonal to the flow except at the equilibrium solution. In (2.7), the point q = 0, p = 0 is a degenerate fixed point due to the factor q3 and it is contained in the line of fixed points q = 0.
178
Figure 2. Flow for e = 0 in the three local charts.
3
The case t = 0
In this case, H = H0 and the system (2.3) is integrable. The global flow is given by the level curves of (2.5a). See figure 2. There is one orbit that leaves +oo with zero velocity and reaches — oo with zero velocity and one orbit that does the opposite. Both are in the level set Ho = 0; we call them the heteroclinic orbits of the unperturbed problem, they are marked in bold lines in the figure (2). In order to give a parametrization of these orbits, we start by defining the following functions
/(*) =
V
J>
,
g(x) = J /(«) du.
(3.1)
H. Dankowicz and P. Holmes3 worked with these functions and obtained analytic expresions of the heteroclinic orbits using elliptic integrals. We prefer to use their geometric properties. The function g(x) is smooth, increasing and odd. It is convex if x > 0, and concave if x < 0. Let 0o(£) be the inverse of g(x), and MV = TtM*)- Then ^°W
=
77X77v>'
. " S 1 tf>o(0 = ±°o.
, l i m ^0(0 = 0.
/ (00 W )
t-»±oo
t->±oo
179 These equations show that the heteroclinic orbits are parametrized by
r 0 (t) = (*>(*), ifoW) and rs(t) = (-to(0,-iM*))-
(3-2)
We are going to deal with the first orbit, To(t). The analysis of the other orbit is similar. The functions 4>o(t) and ij>o(t) are analytic since they are the components of a solution of the analytic differential equation (2.3). 4
Melnikov function for the heteroclinic orbits
When e 7^ 0, the degenerate fixed point in the equation (2.7) corresponding to the escape to plus infinity with zero limit velocity is transformed into a 27r-periodic orbit 71. The parabolic orbits approach this periodic orbit when t —► +00. Let
* M,to) - {rAt)) - {Mt
+ to))
where the functions $i(i, to) and ^\(t,to) tions of the variational equation of (2,1).
+ e ^i{tto))
+ h.o.t.,
(4.1)
are smooth, in fact they are solu
180
Since to acts as a parameter, it is clear from our previous analysis that by varying to or ZQ = To (to) we get the same effect . So from here on, we will identify to with its image under T. Then for each z 0 6 To, let E*° be a cross section at ZQ in the phase space, and let zu(e,zo) and zs(e,zo) be the points in W" (—00) and W° (+00) respectively, that lie in the intersection with E Zo . In order to measure the splitting of W" (-00) and W° (+00) near the point zo = To(io) we define the function M* («, z0) = Ho (za(e, z0)) - H0 (z"(e, z0)),
(4.2)
for zo € To and 0 < e. We remark that using the equation (4.1) and appro priate E 2 °, the functions z"(e, zo) and zu(e,z0), M* are smooth. Using Taylor's expansion of the above function at the point ZQ, with re spect to the parameter t we get M* (e, zo) = cM(zo) + h. o. t., where the last function is the Melnikov function defined by u M(z0) = ±Ho(z°(e,z0))-^Ho(z (e,zo))
(4.3) t=o
Our next objective is to compute M{zo). First, define z3(e,t,to) as the solution of the equation (2.3), such that at time t = 0 it is at za(e, zo)- Thus z s (0,t,t o ) = r ( t + t o ), and pi
j t (z*(e, t, t0)) = JVH0 (za(e, t, t 0 )) + eJVtfi (za{e, t, t 0 )) + h.o.t. The following computation uses the previous equation and the fact that {Ho, Ho} = 0, where {Ho, Ho} = VH0JVH0 is the Poisson bracket and J = (_?i 0) is the canonical symplectic matrix.
H0(z°(e,A,t0))
- ^H0(z'(e,zo))
=J
-£-HQ(z°(e,t,t0))dt
= JQ §-e (vffo (zs(e, t, to)) ^ {z\c, t, to))) dt =J
§; (eVtfo (za(e, t, t0)) (JVff, (z'(e, t, to))) + h.o.t.) dt
= 1
^(t{H0,H1}{za(e,t,to))
= f Jo
+ h.o.t.) dt
{Ho,Hl}{za(e,t,to))+er1(e,t,to)dt,
181 where r)(e,t,to) is a smooth function. Therefore at e = 0 we get (4.4)
= -^(a,«o))
-lH0(z'(e,zo)) £=0
£=0
{#o,ffi} {T{t +10)) dt.
/ + Jto
Taking the limits when A —> oo we get --H0(z°(e,z0))
= -
lim
A->OO
e=0
at
(4.5)
—H0(z'(€,A,to)) e=0
/»O0
+ /
{ff 0 ,ffi}(r(t + to))
Ahead we shall prove the existence of the former limit. Using the same arguments, we also get (4.6)
-H0(z"(e,z0)) £=0
/
J-oo
{H0,Hl}(T{t
+ tQ))dt+
lim
A-+-00 d e
^-H0(zs(t,A,t0)) £=0
Let us observe that, in the local chart containing +oo ^H0(z'(e,A,t0))
VH0(z°(e,A,t0))^(za(t,A,t0)).
=
(4.7)
Then by (4.1), J^(z'(e, A, to)) is bounded in a compact neighborhood of +oo, and linu^+oo VH0(zs(e, A, to)) = 0. Therefore rA
M(z0)=
lim
/
A-+ooJ_A -A +oo
/
{ff 0 ,ffi}(r°(t + io))dt
{H0,H1}(r0(t + t0))dt.
■OO
Remember the identification of to with T(io) = -Zo a n d let us observe that M(ZQ) does not depend on the particular choice of the functions za(e, ZQ) and zu(c, zo), hence the dependence of M* (e, zo) on the particular choice of those functions occurs only in the terms of bigger order in e. With the previous analysis we have proved the following result.
182
Theorem 4.1. The Melnikov function M(-) for the Sitnikov problem is given by r+oo
{ff0,ffi}(r°(t + to)))dt ■oo
r+
-£ 5
°° 3<j>o{t + to)Mt
+ to) cos{t) dt.
(4.8)
Transversality for e > 0
In this section we study a perturbation method developed by Melnikov7 in 1963, in order to prove the existence of transverse heteroclinic orbits. As a consequence, using the ideas of J. K. Moser 8 , we can embed a Bernoulli shift near the parabolic orbits in the Sitnikov problem, concluding that the system possesses chaotic dynamics. We start stating the Melnikov theorem in this case. Theorem 5.1 (Melnikov). Let ZQ = r°(io) o. point on the heteroclinic orbit for the unperturbed system (3.2) such that M{z0,to)
=0
and
dtQ
* 0.
(5.1)
(20,to)
Then W* (+00) and W? (—00) intersect transversally at (zo,to) on the cross section E 2 0 . Now we have to check that the Melnikov function given in the theorem (4.1) satisfies the hypothesis of the theorem (5.1) for to = kit, k £ Z. First, we need to simplify the form of M(ZQ) given by the equation (4.8). Lemma 5.2. The Melnikov function is given by: M(z0) = Asm(t0),
3 r A^irmM^^du. 2 Jo
where
Proof. Let u =■ t + to in (4.8), thus the Melnikov function becomes:
M(zo)=J
+
°° 3<j>o(u)ip0{u) cos(u - to) du
*(<%(«) +\)S/2
3 ,, , [+°° A>(uM(u)co8(ti) =-cos(to) j ,^IS , l v , / 9 du 00 (*§(«)+ * ) 6 ' 2 +00
3 -i«w* )^ /f+°° + T sin(t i ( *0 o ) /
, . „Mu)Mu)s\n{u) , . . ,^r,5/a du. n
4 - ^ y - c o "wg(«) + D
(5.2)
183 The first integral is zero because the integrand is an odd function, and since the function in the second integral is even we get the result. □ L e m m a 5.3. The integral /0°° ^fflff+1 °.°[u) du is positive. Proof. We integrate by parts, using that rpo(t) = -ji4>o(t), [°° 4>o(u)ip0(u)sin(u)
oo
sin(i)
(5.3) H
Jo
5 2
dU
($(«) +l) '
~
+
f
1
F(u) =
3 ($(«) +
(5.4)
l\3/24)
By the Implicit Function Theorem and the definition of F'(u) =
-v/20o(u)
($(«) +i)"'
4
and
1
F"(u) =
4>Q(U)
we have
- (Wo(«) -
"($(«) + i) 4
-!)•
Using that <j>o(u) is increasing and that its limit when t —► oo is inifinity we have that F(u) is even, decreasing if u > 0, lim„_oo F(u) = 0 and it has a unique inflection point on the interval [0, +oo) at uo = ff("7ff )• That is, F(u) is bell shaped. Two consequences of theses facts are that if 0 < u < UQ then F(u) is above the line L\{u) joining the points (0,F(0)) and {UQ,F[UQ))\ and if 7r/2 < u < 37r/2 then F(u) is below the line L,2{u) joining the points (■7r/2,ir(7r/2)) with (37r/2,F(37r/2)). Using Mathematica we compute the following values, where we use four significant digits: u0 = 0.1199, <j>Q{uQ) = 0.2357, F(u0) = 1.974, <j>0(0) = 0, F(0) = 2.667, c£o(7r/2) = 2.079, F ( J T / 2 ) = 0.341, 0O(3TT/2) = 4.524, F(37r/2) = 3.536 x 10" 3 . Using this values we obtain: Li(u) = -5.779u + 2.667
L2(u) = -9.724 x 10 _ 3 u + 4.936 x 10" 2 (5.5)
Let A(u) = F(u)cos(u), thus the second term in the equation (5.3) can be written as /•OO
/•7T/2
/•37T/2
/•OO
/ A(u) du = / A(u) du -1- / A(u)du4- / A(u)du4■ / \{u)du. Jo JZ-IT/2 Jo Juo J-x/2
184
We denote the above integrals as A\, A2, A3 and A4 respectively. First, /•5w/2
A4 — I
,.7w/2
\{u) du +
J3n/2
/-9w/2
X(u) du+
/.llir/2
X(u) du+
J5x/2
J7n/2
X(u) du+ ■ ■■ . J9TT/2
Using the fact that F(u) is decreasing, and the symmetries of the cosine function, it is easy to check that this series is alternating; the terms of the series are decreasing, converging to zero and the first one is positive. These facts imply that the series converges to a positive number, proving that A4 is positive. Clearly A2 is positive. Finally, using the relation between L\(u) and L2{u) with F{u) we have: /•«o
Ay. = / Jo
/"«o
F(u) cos(u) du >
L1{u)cos{u)du
= 0.2776
(5.6)
Jo 3TT/2
\As\
j F(u)cos(u)du T/2 Jn/2
< /
L2(u) |cos(u)| du
■K/2 J-K/2
/•3TT/2 ron/1
< - / K/2
L2(u)cos{u)du
= 0.03762
(5.7)
+ A3 + A4> 0.2776 + A2 - 0.03762 + A4 > 0.
(5.8)
Summarizing, we have that Ai+A2
□ Using the above lemmas we get our main result. Theorem 5.4. Let to = kn, where k 6 Z. Thus the Melnikov function M(to) of the Sitnikov problem satisfies M(to) = 0 and dt' "' ^ 0. Moreover, these are the only points with the above properties. The fact that W* (+00) and Wu (—00) intersect transversally give us im portant consequences. The results of Smale and Birkhoff for homoclinic points as given in J. K. Moser8 and restated by Dankowicz and Holmes 3 , for degener ate fixed points allows us to guarantee the existence of a hyperbolic invariant set where the Poincare" map is topologically equivalent to a Bernoulli shift. That is, in a neighborhood of the heteroclinic points we can introduce symbolic dynamics and with this, we assure the existence of infinitely many periodic orbits with large period and infinitely many oscillatory orbits. Recalling that an orbit is said to be oscillatory if limsupx(t) = 00 and liminf x(t) < 00 as t —♦ ±00 (See Z. Xia 12 ). This also shows that the Sitnikov problem is very sensitive with respect to initial conditions and therefore chaotic and nonintegrable.
185
Acknowledgments The authors are partially supported by by Conacyt-Mexico, projects 4002005-1406 PE and 400200-1 Exp. 980108. References 1. V. Alekseev, Quasirandom dynamical systems I, II, III, Math. USSR-Sb. 5, 73-128 (1968); 6, 505-560, (1968); 7, 1-43, (1969). 2. J. Chazy, Sur l'allure finale du mouvement dans le probleme des trois corps I, II, III, Ann. Sci. Ecole Norm. Sup. 3 e Ser. 39, (1922); J. Math. Pures Appl. 8, (1929); Bull. Astron. 8, (1932). 3. H. Dankowicz and P. Holmes, The existence of transversal homoclinic points in the Sitnikov problem, J. Diff. Equations 116, 468-483, (1995). 4. A. Garcia and E. Perez-Chavela, Melnikov techniques in the Sitnikov problem. To appear. 5. J. Guckenheimer and P. Holmes, Nonlinear Oscillations, Dynamical Sys tems, and Bifurcations of Vector Fields, Springer-Verlag, New York, (1983). 6. R. McGehee, A stable manifold theorem for degenerate fixed points with applications to celestial mechanics, J. Diff. Equations 14, 70-88, (1973). 7. V. K. Melnikov, On the stability of the center for time periodic pertur bation, Trans. Moscow. Math. Soc. 12, 1-57, (1963). 8. J. K. Moser, Stable And Random Motion In Dynamical Systems, Annals Math. Studies 77, Princeton University Press, (1973). 9. T. J. Morrisey, A degenerate Hartman theorem, Israel Journal of Math ematics 95, 157-167, (1996). 10. C. Robinson, Dynamical systems; stability, symbolic dynamics and chaos, CRC Press, (1994). 11. K. A. Sitnikov, Existence of oscillatory motions in the three-body prob lem, Dokl. Akad. Nauk. USSR 133, 303-306, (1960). 12. Z. Xia, Melnikov method and transversal homoclinic orbits in the re stricted three body problem, J. Diff. Equations 96, 170-184, (1992).
D O U B L Y - S Y M M E T R I C P E R I O D I C SOLUTIONS OF HILL'S L U N A R P R O B L E M R. CLARISSA HOWISON 619 Washington Blvd. Baltimore, MD 21230, USA. E-mail: [email protected] KENNETH R. MEYER Department of Mathematical Sciences, University of Cincinnati Cincinnati, Ohio 45221-0025, USA E-mail: [email protected] The existence of a new family of periodic solutions to the spatial Hill's lunar problem is established. These solutions have large inclinations and are symmetric with respect to two coordinate planes. In this family the infinitesimal particle is very close to the primary.
1
Introduction
This paper establishes the existence of a new family of periodic solutions to the spatial Hill's lunar problem. The periodic solutions of this family have large inclinations and are symmetric with respect to two coordinate planes — hence the name doubly-symmetric periodic solutions. In this family the infinitesimal particle (the moon) is very close to the primary (the earth). These periodic solutions are perturbations of circular solutions of the Kepler problem. By the Kepler problem we mean the spatial central force problem with the inverse square law of attraction. A related paper of the authors 7 established the existence of two new families of periodic solutions to the spatial restricted three-body problem by Poincare's continuation method. These families exist for all values of the mass ratio parameter \i and have large inclinations. In one of the family the infinitesimal particle is far from the primaries and in the other case the infinitesimal is very close to a primary. In this note we will indicate that the latter family exists in Hill's lunar problem also. This is reasonable since Hill's lunar problem is a limit of restricted problem developed to study the motion of the moon 5 . The small parameter e will be introduced as a scale parameter in such a way that s small means the infinitesimal is close to the primary. The per turbation problem is very degenerate. First of all, even to the second ap proximation the characteristic multipliers are all + 1 , and second, the periodic solutions that we establish are undefined when e = 0. These difficulties are
186
187 overcome by exploiting the symmetries of the problem and using the implicit function theorem of Arenstorfx. In 1965, Jeffreys8 showed that there exist doubly symmetric, periodic so lutions to the three dimensional restricted three-body problem. His method of the proof depends heavily on a symmetry argument, together with a standard perturbation method applied to the mass ratio n of the restricted problem There is no natural parameter in Hill's lunar problem corresponding to (J. and so a scaling parameter is introduced. The problem becomes considerable more difficult. 2
Hill's lunar p r o b l e m
One of Hill's major contributions to celestial mechanics was his reformula tion of the main problem of lunar theory: he gave a new definition for the equations of the first approximation for the motion of the moon 5 . Since his equations of the first approximation contained more terms than the older first approximations, the perturbations were smaller and he was able to obtain series representations for the position of the moon that converge more rapidly than the previously obtained series. Indeed, for many years lunar ephemerides were computed from the series developed by Brown, who used the main prob lem as defined by Hill. Even today, most of the searchers for more accurate series solutions for the motion of the moon use Hill's definition of the main problem 4 . Before Hill, the main problem consisted of two Kepler problems — one describing the motion of the earth and moon about their center of mass, and the other describing the motion of the sun and the center of mass of the earth-moon system. The coupling terms between the two Kepler problems are neglected at the first approximation. Delaunay used this definition of the main problem for his solution of the lunar problem, but after twenty years of computation was unable to meet the observational accuracy of his time. In Hill's definition of the main problem, the sun and the center of mass of the earth-moon system still satisfy a Kepler problem, but the motion of the moon is described by a different system of equations known as Hill's lunar equations. Using heuristic arguments about the relative sizes of various physical constants, he concluded that certain other terms were sufficiently large that they should be incorporated into the main problem. In a popular description of Hill's lunar equations, one is asked to consider the motion of an infinitesimal body (the moon) which is attracted to a body (the earth) fixed at the origin. The infinitesimal body moves in a coordinate system rotating so that the positive x axis points to an infinite body (the sun)
188
infinitely far away. The ratio of the two infinite quantities is taken so that the gravitational attraction of the sun on the moon is finite. The Hamiltonian of the three-dimensional Hill's lunar problem is H = o (l/i + vl + »I) - x iy2 + Z23/1 - = (x? - x\ - x\) -
,
=^ (i)
9
(see Llibre, Meyer and Soler ). 3
a
Symmetries and Special Coordinates
The Hamiltonian (1) is invariant under the two anti-symplectic reflections: 111 : (xj,22,2:3,2/1,2/2,2/3) —*
{xi,~x2,-x3,-yi,y2,y3), (2)
TZ2 : (xi,x2,x3,yuy2,y3)
—► (xi, - x 2 , x 3 , -yuy2,
-y3).
These are time-reversing symmetries, so if (xi(t),x2(t),x3(t), 2/1 (*),2/2(*), 2/3(*)) ^ a solution, then so are (xi(—t), — x2(—t),±x3(— t), —2/1 ( — 0,2/2( — 0,-F2/3(—t)). The fixed set of these two symmetries are Lagrangian subplanes, i.e. £1 = {(si, 0,0,0,1/2,1/3)},
£2 = {(xi,0,x 3 ,0,1/2,0)},
are fixed by the symmetries TZi,TZ2. If a solution starts in one of these La grangian planes at time t = 0 and hits the other at a later time t = T then the solution is 4T-periodic and the orbit of this solution is carried into itself by both symmetries. We shall call such a periodic solution doubly-symmetric. Geometrically, an orbit intersects £1 if it hits the Xi-axis perpendicularly and it intersects £2 if it hits the xi,X3-plane perpendicularly. To be more specific, let (Xi(t,a,/3,7),X 2 (<,a,/9,7),X3(t,Q,i9,7).Vi(*,a,i9,7),V2(*,a l /3,7),l'3(*,a,/3, (3) be a solution which starts at (a, 0,0,0,/?,7) 6 £1 when t = 0, i.e. X 1 (0,a,/3,7) = a,
X 2 (0,a,/3, 7 ) = 0,
X 3 ( 0 , a , / ? , 7 ) = 0,
yi(0,a,/?,7) = 0,
y2(O,a,/J,7) = 0,
Y3(0,a,M)
(4)
=T
"These is a typographical error in the Hamiltonian of Hill's lunar problem found in Meyer and Hall10.
189
The solution with a = Q 0 , 0 = (3Q, 7 = 70 will be doubly-symmetric periodic with period AT if it hits the £2 plane after a time T, i.e. *2(T,ao,/3o,7o)=0,
^ (T,a 0 , A), 7o) = 0,
Y3(T,ao,0ano)
= 0.
(5)
This solution will be a nondegenerate doubly-symmetric periodic solution if the Jacobian d(X2,Yi,Y3) -£— ^ - { T , a 0 , 0 o , 70) (6) d(t,a,0,-r) has rank three. It follows from the Implicit Function Theorem that nondegenerate doublysymmetric periodic solutions can be continued under a small conservative perturbation which preserves the symmetries. In general, a nondegenerate doubly-symmetric periodic solution may not be nondegenerate in the classical sense, i.e. a nondegenerate doubly-symmetric periodic solution may have all its multipliers equal to one. Jefferys8 proved the existence of nondegenerate doubly-symmetric peri odic solutions of the spatial restricted three-body problem by first setting the mass ratio parameter \i equal to zero to get the Kepler problem in rotating coordinates. He then showed that some of the circular solutions of the Kepler problem where nondegenerate doubly symmetric periodic solutions. Thus, by the above remarks these solutions can be continued into the restricted prob lem for small \i. Since there is no natural parameter like [i we will introduce a scale parameter e. This makes the analysis much more delicate. We follow Jefferys by using a variation of the Poincare-Delaunay elements. First, the Delaunay elements (£, g, k, L, G, K) are a coordinates on the elliptic domain of the Kepler problem. The elliptic domain is the open set in K6 which is filled with the elliptic solutions of the Kepler problem. The elements are: £ the mean anomaly measured from perigee, g the argument of the perigee measured from the ascending node, k the longitude of the ascending node measured from the x\ axis, L semi-major axis of the ellipse, G total angular momentum, K the component of angular momentum about the x 3 -axis. £,g, and k are angular variables defined modulo 2ir, and L,G and K are radial variables. If i is the inclination of the orbital plane to the X\,X2 reference plane, then K = ±Gcosi, and so an orbit is in the xi,X2-plane when K = G. (Often, k and K are denoted by h and H, but we are Hamiltonophiles.) An orbit hits £1 at time t = 0 if it is perpendicular to the Xi-axis. So its orbital plane must be through the ii-axis or k = 0 mod n, its perigee must be on the xi-axis or g = 0 mod TT, and it must be at perigee (apogee) or £ = 0 mod TT. Thus, C\ in Delaunay elements is defined hy £ = g = k = 0 mod -K.
190
An orbit hits £2 at time t = T if it is perpendicular to the xi,X3-plane. So its orbital plane must be perpendicular to the xi,X3-plane or k = n/2 mod 7r, its perigee must be in the Xi,X3-plane or g = n/2 mod ir, and it must be at perigee (apogee) or t = 0 mod 7r. Thus, £2 in Delaunay elements is defined by t = 0, g = k = TT/2 mod TT. Since these coordinates are not valid in a neighborhood of the circular orbits of the Kepler problem, we change to Poincar£ elements as follows: first make the symplectic linear change of variables 9i = I + 9 + k, q2 = -k -g, Q3 = l + 9,
p! = L - G + K, p2 = L-G, p3=G-K,
and now apply the symplectic change of variables defined by the generating function P2 W(q, P)=qiPx + -2- tanq 2 + P3q3 so that P2 = v^2p2 cos q2 and Q2 = y/2p2sinq2. changes gives the new variables:
This combination of variable
Q\ = q\ = I + 9 + k,
Pi=Pi=L-G
+ K,
Q2 = - V 2 ( L - G ) s i n ( f c + g),
P2 = y/2(L - G) cos(k + g),
(7)
Q3 = q3 = l + g, P3=p3 = G-K. These variables are valid on circular orbits which occur at L = G (see Howison6 or Szebhely 12 ). The circular orbits with L = G correspond to Q2 = P2 = 0. Thus, C\ in Poincare elements is defined by Q2 = 0, Q\ = Q3 = 0 mod 7r, and £ 2 in Poincar6 elements is defined by Q2 = 0, Q\ = 0 mod n, Q3 = TT/2 mod 7r. 4
Approximate Solutions
Move the infinitesimal mass close to the origin by scaling the variables: x —» e2x, y —> e~1y, which is symplectic with multiplier e - 1 . Letting H —> e~lH, expanding the potential in e, and by dropping the constant terms, the Hamiltonian becomes
H = £_3
M "Rf
{xiV2
~ X2Vl) + £3i/t(x'y'e)'
(8)
191 where H^ is analytic and order 1 in e. The solutions that will be establish will have the new x, y coordinates of order 1 in £ and so the original x will be order £ 2 and the original y will be order £ _ 1 . We will not scale time and the solutions we establish will have periods which are order 1 in s. Note that as £ —» 0 the Hamiltonian tends to infinity, thus we can not just set £ = 0. We will need approximate solutions to the equations and good estimates. (Scaling time does not remove the difficulties of the problem. If we scale time so that the Hamiltonian becomes order 1 in £ then the new periods will tend to infinity.) In Delaunay elements, the Hamiltonian becomes H = ^
r
- K
+ e3HHe,9,k,L,G,K,e).
(9)
Since these coordinates are not valid in a neighborhood of the circular orbits, we change to Poincare" elements (7) and the Hamiltonian becomes H
= , w " £ „v2 ~Pl + \ {P2 +Ql) 2 (Pi + Pz) *
+S3Ht(QUQ2,Qz,Pl,P2,P3,£)(10)
Thus the equations of motion are Q = Hp, P = —HQ or £~3
^1==77^ ^ 3 \P\ + P3)
1 + £3
/i>
- P i = 0 + £3/4,
Q 2 = Pi + £ 3 /2,
P2 = - O a + £ 3 /s,
^3
-^3=0 + £ 3 / 6 ,
=
. p l p ^ 3 + £ 3 /3,
(11)
(Pi + P3) where the /j are the appropriate partials of W. First let us consider the approximate equations in order to find the correct approximate periodic solutions. Consider the approximate equations Wl
Qi
(Pi
+ P3)3
= P2,
(Pi
Pi = 0, P2 = - Q 2 ,
e- 3
Qz-
1,
+ P3) 3 '
P 3 = 0.
(12)
192
These are of course, the equations of motion for the Kepler problem in scaled, rotating Poincare elements. The solution of equation (12) are O lt\ Ql(i)_
/ e-3 -l)t V(Pl+P3)3 1
+ qu
Q2(t) ~-= qi cos t + P2 sin t,
Q3(t) = ■
(
£
~
W) = —q2sint
\ t 4- /jo
3
Pi(t)=pu
V(Pl+P3> 3 1 t-l-93.
P*{t) -=
+p2Cost,
(13)
P3,
for initial conditions {qi,q2,Q3,Pi,P2,P3) at t = 0. The periodicity conditions are the same as those in the Section 3. That is, at t = 0; Qi = in, Q2 = 0, Q3 = jn and at t = T; Qx = {i + k)n, Q2 = 0, Qs — 0' + m + l/2)7r where i and j are 0 or 1, and k, and m are arbitrary integers. To satisfy these symmetry condition at t = 0 and at t = T we have so solve the equations
«^-(ra^- 1 ) T + i , r = ( i + i f c ) '' Q 2 ( r ) = p 2 s i n T = 0,
Q3(T) =
(14)
( ( ^ F ) T + J7r = {j + m + 1/2)7r>
The second equation is solved by taking p? = 0, thus selecting a circular orbit of the Kepler problem. The difference between the first and third equa tion has a solution with T = (m — k + l/2)w. It remains to solve the third equation. With this choice of T it becomes .
,,
£ - 3 (m - k + i )
Recall that Pi + P3 = L which is the semi-major axis in the Kepler problem and that we want solutions which are order 1 in L. Let n be a fixed small integer. To solve (15) set m+l/2 = £ - 3 , k = m — n, and (pi +P3) 3 = (n + 1 / 2 ) . With this choice the period becomes T = n +1/2 We shall indicate in Section 5 that these solutions of the approximate equations (12) are actually approximations of actual doubly-symmetric peri odic solutions of the true equations (11). Thus, our main theorem is
193 Theorem 4 . 1 . There exist doubly-symmetric periodic solutions of the spatial Hill's lunar problems with large inclination which are arbitrarily close to the primary.
5
Outline of the proof
Here we will outline the proof and refer the reader to Howison and Meyer7 for a more detailed account of a similar result. Consider the equations (11). The periodicity conditions remain: at t = 0; Q\ = in, Q
Qi(*)= ( (
£ +
^ - M t + gi+g^i.
Q2(t) = 92 cos t + pi sin t + e3g2, Q&)
=
^i(*)=Pi+e394, ^M*) = p? cos t - q
it 1 \a) f + 93 + £3$3, \(Pl +P3) J
Pz(t) = P3 + £396,
for initial conditions (qi,q2,q3,Pi,P2,P3) and for time ie[0,7], where ft = 3i(*,9l.92,93,Pl,P2,P3)To satisfy the symmetry condition at t = 0 we have the solution
Q i ( ' ) = \T^ - s - l l i + Mr + e ^ i , \(Pl+P3) J Q2W = p 2 s i n ( + £ 3 g 2 ,
Q*W = I 7-^
^ 1 * + i* + £333-
\(Pi+Ps) / Next we need to solve this for the symmetry condition at t — T which are now:
194
GiW = I T~^ <& - 1 I l + i7r + £39i = (* + *) T, \(Pl+P3) / Q2{t) = p2sint
+ e3g2 = 0,
<&(*) = (7—^ ^ ) * + ^ + e3Sa = ( i + m + i ) 7 r , V(Pi + Ps) / or =• - 1 I t - kit + e3g! = 0,
(P1+P3)3
i
p2 sin t + eJp2 = 0, 3 3 - 1 h - ( m + 2i)7r + £ 5 3 = 0.
V(A+ft)
)
This is done by applying the Arenstorf implicit function theorem twice. Roughly speaking Arenstorf's theorem applies to situations where the prob lem is undefined when e = 0, all that is required is that the perturbation and the derivatives of the perturbation are sufficiently small. See Arenstorf1 or Howison and Myer7 for details. First we consider the difference of the first and third equation, together with the second equation. This is the system of equations t + (k - m - i ) 7r + e3g3 - £35i = 0, p2 sin t + e3g2 = 0. This has solution t = (m + 1/2 — k)n, p2 = 0 at e = 0. Along this solution, the determinant of the derivative of the system with respect to t and P2 is given by 1 0 0 sin(m + 1/2 - k)ir)
±1^0.
Thus by the Arensdorf implicit function theorem, there exists a neighbor hood N of 0 and functions T(pi,p3,e) near (m + 1/2 — k)n and p2(p\,P3,e) near 0 for e G M and (pi,P3) arbitrary such that
195 T(puP3,e)
~ (m + \ - fc)* + e 3 5 3 - £39i = 0,
P2(pi,P3,£)sinT(p 1 ,p 3 ) e)+£ 3 p2
=0.
Putting this solution for T into the third equation, we need to solve Qz{T) (m + \)ir — 0, or £-3
(m+--fcW-£353+£35i
(pi +P3) 3
- (m + - j w + e3g3 = 0,
which is equivalent to solving m + - - fc J TT - £ 3 5 3 + e 3 3i - f £3 f m + - J TT - e 6 5 3 J ( P l + p 3 ) 3 = 0 for (pi +P3) 3 whenever e &M — {0}. Now the solution for T left both m and fc arbitrary, so for the moment regard m and k as free variables. Then letting m + k = e~3 and letting fc = m - n for n a small integer, we need to solve R =
v
+
2) * ~£3fl3+£3si ~(*~e6g^ ^Pl+P3^3 = °'
By this choice of m and A;, T becomes T(pi, p 3 , e) = (n + 1/2)TT + £ 3 g 3 - e3g\ which is uniformly bounded as e approaches zero. dT/dpi at e = 0 is given by -Ad{e3gs—£392)/dp\ where the partialsof the & are evaluated along solutions; t = (n + \)ir, P2 = 0, £ = 0, (pi,p 3 ) arbitrary. By the another Grownwall argument the partials of the
196
References 1. R. F. Arenstorf, A new method of perturbation theory and its application to the satellite problem of celestial mechanics, J. Reine Angew. Math., 221, 113-145, (1966). 2. E.A. Belbruno, A new regularization of the restricted three-body problem and an application, Celestial Mech., 25, 397-415, (1981). 3. P. Guillaume, New periodic solutions of the three dimensional restricted problem, Celestial Mech., 10, 475-495, (1974). 4. M. C. Gutzwiller and D. S. Schmidt, The motion of the moon as computed by the method of Hill, Brown, and Eckert, Astronomical Papers XXIII(I), U.S. Naval Observatory, Washington, DC, (1986). 5. G. W. Hill, Researches in the lunar theory, Chapter I, Amer. J. Math., 1, 5-26, (1878). 6. R.C. Howison, Doubly-symmetric Periodic Solutions to the Three Di mensional Restricted Problem, Ph.D. Dissertation, University of Cincin nati, (1997). 7. R.C. Howison and K.R. Meyer, Doubly-symmetric periodic solutions of the spatial restricted three-body problem, to appear in J. Diff. Equations. 8. W. H. Jefferys, A new class of periodic solutions of the three-dimensional restricted problem, Astron. J., 7 1 , 99-102, (1965). 9. J. Llibre, K. R. Meyer, and J. Soler, Bridges Between the Generalized Sitnikov Family and the Lynpunov Family of Periodic Orbits, to appear in J. Diff. Equations. 10. K.R. Meyer and G.R. Hall, An Introduction to Hamiltonian Dynamical Systems, Springer-Verlag, New York, (1991). 11. D.S. Schmidt, Families of Periodic Orbits in the Restricted Problem of Three Bodies, Ph.D. Dissertation, University of Minnesota, (1969). 12. V. Szebehely, Theory of Orbits, Academic Press, New York, (1967).
O N P R A C T I C A L STABILITY REGIONS FOR T H E M O T I O N OF A SMALL PARTICLE CLOSE TO THE EQUILATERAL P O I N T S O F T H E REAL E A R T H - M O O N S Y S T E M ANGEL JORBA Departament de Matematica Aplicada i Analisi, Universitat de Barcelona, Gran Via 585, 08007 Barcelona, Spain. E-mail: [email protected] This is a summary of a preliminary study on the existence of stability regions near the equilateral libration points of the real Earth-Moon system (here, by "real" Solar system we refer to the model defined by the ephemeris DE406 from JPL). We start from a model that takes into account the main effects coming from Earth, Moon and Sun to locate regions where trajectories are confined for long time intervals. Then, these regions are checked against a numerical integration that takes into account the real motion of Earth, Sun and the planets. This shows the existence of regions such that any initial condition inside them produces a trajectory that remains there for at least 1000 years.
1
Introduction
Let us start first with the simplest model for this problem, the well known Restricted Three Body Problem (from now on, RTBP). The RTBP models the motion of a particle under the gravitational attraction of the Earth and Moon (also called primaries), under the following assumptions: i) the particle is so small that it does not affect the motion of Earth and Moon; and ii) Earth and Moon are point masses that revolve in circular orbits around their common centre of mass. It is usual to take a rotating frame with the origin at the centre of mass, and such that Earth and Moon are fixed on the X axis, the (X, Y) plane is the plane of motion of the primaries, and the Z is orthogonal to the (X, Y) plane. The units of distance, mass and time are defined such that the Earth-Moon distance, the total mass of the system (Earth+Moon) and the gravitational constant are all one. With these units, the period of the motion of Earth and Moon around the origin equals 2n. Defining momenta as p^ = x — y, py = y + x and pz = z, the equations for the motion of the particle can be written as an autonomous three degrees of freedom Hamiltonian system. The corresponding Hamiltonian function is HRTBP = M
+ Pi + Pz) + VPx ~ XPv ~ ~T~~ ~ ~T~,
(1)
being rpE = (x - fi)2 + y2 + z2 and r2PM = (x - /i + l ) 2 + y2 + z2. It is well known that this Hamiltonian has five equilibrium points, that
197
198 l
1
1
1
i
-
0.8 0.6
_
0.4 0.2
-
L,
U
X,
0
-0.2
-
.
M
E
-
-0.4 -0.6 -0.8
-
-1 -1.5
'
-
XL« i
-0.5
•
0.5
1.5
Figure 1. The five equilibrium points of the RTBP.
are displayed in Figure 1: three of them lay on the X axis (they are called collinear points, Eulerian points, or simply 1-1,2,3), and the other two form an equilateral triangle (in the (X, Y) plane) with Earth and Moon (they are called triangular points, Lagrangian points or simply £-4,5). For more details see, for instance, [1] or [2]. It is also well known that, for the value of the masses of Earth and Moon, the triangular points are linearly stable. It is known that, under very general conditions, there exists a neighbourhood of these points such that any initial condition inside this neighbourhood produces a trajectory that remains close to the point for huge intervals of time (see [3], [4], [5], [6] and references therein). On the other hand, let us consider the motion of a particle in the real Solar system, close to the triangular points of Earth and Moon. Here, by real system we refer to the one obtained from the well known JPL ephemeris [7]: the ephemeris contain the position of the main bodies of the Solar system (Earth, Moon, Sun and planets) so we can easily write the vector field acting on the particle. This is a numerically defined vector field, that we can integrate for the time span for which the ephemeris are available. Note that, for the real system, the triangular points are not longer equilibrium points since the hypotheses used to derive the RTBP are not satisfied. It is know that arbitrary trajectories starting near the (geometrically defined) triangular points go away after a short time (see [8] and references therein) and hence, the RTBP is not a good model to study this problem because it displays a quite different qualitative behavior. In this talk we will first use the so called Bicircular Problem (from now
199 on, BCP). This model can be seen as a time dependent periodic perturbation of the KTBP, that includes the main effect coming from the Sun. As far as we know, this model is first introduced in [1]. As we will see, the BCP model has some of the main features of the real model (for instance, the neighbourhood of the equilateral points is unstable). Our purpose is to use the BCP model to describe some properties of the real model. In particular, we will compute, for the BCP, families of 2-D in variant tori near the equilateral points, as well as their normal behaviour. We will show (by computing the normal modes of the tori) that some of them are hyperbolic, while some others are elliptic. Those elliptic tori are found at some distance from the equilateral points. It is known that lower-dimensional normally elliptic tori give rise, under suitable hypotheses, to regions of effec tive stability around them (see [9], [10]). Finally, we will show, by means of numerical simulations, that some of these stability regions found in the BCP model subsist in the real system, at least for time spans of 1000 years. This talk is based on results contained in [11], [12] and [13]. They can be downloaded from the web server http:Wwww.maia.ub.es/dsg, in the p r e p r i n t s section. 2
The Bicircular Problem
Let us assume that Earth and Moon are point masses that revolve in circular orbits around their common centre of mass (as in the RTBP), and let us also assume that this centre of mass revolves in a circular orbit around the Sun (that is also considered as a point mass) that is fixed at the origin. We are also assuming that the motion of the three bodies is coplanar. The purpose of the Bicircular Problem is the study of the motion of an infinitesimal particle under the attraction of these three bodies. It is usual to take the same reference frame and normalized units as in the RTBP. So, in this system of reference, Earth and Moon are fixed on the x axis while the Sun is turning around the origin. Defining momenta as px = x - y, py — y 4- x and pz = z, then the motion of the infinitesimal particle can be described by a Hamiltonian system that depends on time in a periodic way: HBCP
= 2 (Px + P\ + Pi) + yPx ~ xpv
-
2" (ysmO — xcosw), (2) as where rPE = (x - /x)2 + y2 + z2, r2PM = (x - fj. + l ) 2 + y2 + z2, rPS = (x - xs)2 + (y - ys)2 + z2, *s = as cos6, ys = -as sin0, and 6 = u>st. The TPE
TPM
Tps
200 /i LJS
= =
0.012150581623433623 0.925195985518289646
ms as
= =
328900.54999999906 388.81114302335106
Table 1. Floating point values used for the different constants of the BCP. Not all the digits are significative (they appear because the definition of these quantities involves quotients).
values used for these constants are listed in Table 1. We stress that the BCP can be written as a periodic time dependent perturbation of the RTBP:
ire "BCP
=
ZJ fi "RTBP + £nBCP,
G I L "BCP = -"*S I \rpS
1
ysmu -xcoso o as
\ . ) ' \°) )
and it is clear that HB^9p = HRTBP, and that H££QP — HBCPNow, let us focus on the L5 point (due to the symmetries of the RTBP and the BCP, the same results will be valid for L 4 ; of course,J;his is not true for the real system). As the vector field of the perturbation EHBCP does not vanish on L5, this point is not longer an equilibrium solution. An straightforward application of the Implicit Function Theorem shows that, under a generic non-resonance condition (satisfied in this case) and assuming e small enough, L 5 is replaced by a periodic orbit with the same period as the perturbation. This orbit tends to L5 when e tends to zero. Let us consider now the Poincare section defined by t = 0 mod Ts, where Ts = 2n/u)s is the period of the perturbation. Let us call PE the corresponding Poincar6 map. Note that the fixed points of this map correspond to periodic orbits (of period Ts) for the flow, and vice versa. Now, by means of a continuation process, let us compute the fixed points of PE for e ranging between 0 and 1. The results are displayed in Figure 2 (left), where the horizontal axis shows the x coordinate of the fixed point and the y coordinate refers to the value of e. Note that, for e = 1, there are three fixed points of Pe that are close to L$. The (x,y) projection of the corresponding periodic orbits for the flow are displayed in Figure 2 (right). By computing the eigenvalues of the differential of Pc at the fixed points, one can see that orbit P O l is unstable, and that orbits P 0 2 and P 0 3 are linearly stable (see Figure 2). Moreover, the three periodic orbits have an elliptical mode very close to the {z,pz) plane. In what follows, we will refer to this mode as the vertical mode of the corresponding periodic orbit. For more details about the BCP, see, for instance, [14] or [15].
201
Figure 2. Left: Continuation of the periodic orbit that replaces i 5 in the BCP. Right: The three periodic orbits that appear in the left plot for e = 1.
2.1
The vertical families of 2-D tori of the BCP
Let us now consider the vertical mode of one of the periodic orbits. It is known that, under generic conditions (see [9], [10]), there exists a Cantor family of 2-D invariant tori that extend this linear mode into the complete (i.e., nonlinear) system. This is very similar to the well-known Lyapunov centre theorem: instead of obtaining a (smooth) family of periodic orbits, one obtains a (Cantor) family of 2-D tori. Next, we will focus on the computation of these vertical families of twodimensional tori. Their interest will be clear later on. 3
Computation of invariant curves
In this section we will briefly explain a numerical method for computing fam ilies of invariant curves of high dimensional maps. The method does not require the knowledge of an algebraic expression for the map, it only requires a procedure to numerically evaluate the map (and its differential) at a given point. So, it is a suitable method to be applied to the Poincar£ map of an ODE. More details can be found in [11]. 3.1
A Newton method
Let us consider a map / from an open set of R n into itself, with an invariant curve with (irrational) rotation number u\ this is, we assume that there exist a map x : T 1 —♦ R n such that f(x{6)) = x{6 + w),
for all 0 € T 1 .
(4)
202
To simplify the discussion, let us assume that we know the value u , so we only need to determine the Fourier coefficients of the function x(8), x(0) = OQ + ^2 ak cos(fc0) + bk sin(fc0),
k € N.
k>0
Let us now fix a truncation value N (the selection of N will be discussed in Section 3.3), and let us try to compute an approximation to the 2N + 1 unknown coefficients UQ, ak and bkj 0 < k < N. First, let us define a mesh of 2iV + l points on T 1 , 0J = » y . . 0<j<2N. 3 2N + 1 -J Imposing condition (4) on the points of this mesh we obtain the system of 2N + 1 equations, f(x{0j)) - x{9j + u) = 0,
0 < j < 2N,
(5)
that we will solve, by means of the Newton method, to determine the 2N + 1 unknown Fourier coefficients. The differential of this equation can be easily obtained by using the chain rule: Let F be the function that maps each sequence of 2N + 1 Fourier coefficients (ao, a\, b\, 02,62,..., aw, 6AT) on the vector of 2N + 1 components given by
(f(x(0j))-x(0j+uj))oiJS2N. Hence, denoting by Df(p) the Jacobian of / at a given point p G K n , we have
dF -DfWOi)) dak Note that Df(x(9j))
^l(0j)
^M)
^
(9,).
can be easily computed, and that
= Incos(k9J),
dX
^
+
U})
(9j)-Incos(k9J+kcj),
(6)
where /„ denotes the n x n identity matrix. Similar formulas hold for the derivatives with respect to bk- From these formulas one can easily assemble the Jacobian matrix of F: it consist of (2N +1) x (2N + 1) boxes, being each box a n x n matrix (we recall that ak and bk are arrays of n components).
203
3.2
On the uniqueness
If x(9) denotes a Fourier series corresponding to an invariant curve then, for any
Error estimation
To select the truncation value N used in the discretization (see Section 3.1), we will use the following estimation for the error: E(x,u)
= sup \f{x{6)) - x{9 + w)|. 0ST>
Clearly, x is an invariant curve with rotation number u iff E(X,LJ) is zero. For each curve, we have estimated the quantity E(x,w) by simply tab ulating its value on a mesh of points on T 1 . Of course, this mesh must be finer (typically, we have used a 100 times finer mesh) than the one used to obtain (5) from (4). When this estimate is bigger than a prescribed threshold (here we have used 10 - 1 2 ), the program automatically increases the number of Fourier coefficients and the number of discretizing points (i.e., it increases the truncation value N) and recomputes the invariant curve. This process is repeated until the estimated error E(X,UJ) is lower than 1 0 - 1 2 . 3.4
Results
By means of a standard continuation procedure (see details in [11]), we have computed the vertical families of 2-D tori for the three periodic orbits POl,
204 Ep6ilon«0-69
EptlkxM).91
Figure 3. Vertical families of invariant tori for several values of e. The horizontal axis shows the value of the coordinate z of the invariant curve when z = 0, and the vertical axis shows the corresponding rotation number. The gaps that conform the Cantor structure are too narrow to be seen at this scale.
P 0 2 and P 0 3 , for several values of t (see Figure 3). For definiteness, let us call Fj the family that is born in P O j , j = 1,2,3. Note that the families F l and F2 are connected, as it is suggested by the diagram in Figure 2 (left). As P O l is hyperbolic, we expect that the tori in F l are also hyperbolic, at least for moderate values of the vertical amplitude. P 0 2 and P 0 3 are elliptic orbits so, under generic conditions, it is expected that most of the tori in families F2 and F3 with a small vertical amplitude are also normally elliptic. 0 The study of the normal stability of these families is the topic of Section 5.
"Although most of them are going to be elliptic tori, a small measure set of them is ex pected to be normally hyperbolic, due to some resonances involving internal and normal frequencies.
205 4
Stability regions
It is also known that, under generic conditions of analyticity, non-degeneracy and non-resonance, a lower dimensional and normally elliptic invariant tori is surrounded by a region of effective stability. Here, by "region of effective sta bility" we mean that trajectories starting inside this region need a extremely long time to escape of a neighbourhood of the torus [10]. Here we will try to estimate this region by means of purely numerical methods (see [16] for a semi-analytical approach in the case of a periodic orbit). So, given a 2-D torus on one of these families, we will first select the unique point a^ on it satisfying t = 0 (i.e., we are selecting the point on the Poincare section), z = 0 and z = ZQ > 0 (io will be used as a parameter). Then, we will take a 2-D mesh of points around the (x, y) coordinates of ao (and inside the (x,y) plane), keeping fixed all the remaining coordinates. Then, we will use each point on the mesh as initial condition for a nu merical integration of the vector field. If the trajectory does not cross the y = -0.5 plane, and it does not approach collision with Earth and Moon dur ing the integration time (see Section 4.1), we admit that the initial condition belongs to the stability region. Next, in Section 4.1, we will summarize the results from the simulations for the BCP and for the real system. 4-1
Some numerical simulations
Let us first select the 2-D torus of the family F3 such that, in the Poincare section t = 0, z is zero and z is 0.58 (the reasons for this selection will be clear later). Let us first make a numerical simulation around this torus, as explained above. Figure 4 (left) displays the set of initial conditions (x, y) such that the corresponding trajectory does not escape for a time span of 5000 periods of the Moon. Then, let us take this mesh to make a numerical simulation for the real system, by using the JPL ephemeris (file DE406). Here, the initial integration time is the first full Moon of year 2000 (this is to have a relative position of Earth, Moon and Sun similar to the one of the BCP when t = 0). Figure 4 (centre) shows the initial conditions that do not escape after 500 years, and Figure 4 (right) contains the ones that do not escape after 1000 years. The unit of distance in the scales of the plots is the Earth-Moon distance. Note that the size of these regions is quite big. Let us now consider the torus (of the family F3) corresponding to z = 0 and i = 0.83. Doing the same simulations as before, one obtains the results displayed in Figure 5. As before, the left graphic refers to the BCP (the
206
Figure 4. Left: stability region for the BCP, for i = 0.58. Centre: stability region for the real model (500 years). Right: stability region for the real model (1000 years). The box corresponds to the explored region. The position of Ls is marked with a big "+" sign.
-1
«f
<0«
«.4
-0.2
0
92
-I
-0J
-Ol
-0.4
42
9
02
-1
410
-Of
-0.4
«2
0
02
Figure 5. As Figure 4, but for the case z = 0.83.
integration time is again 5000 revolutions of the Moon), and the centre and right graphics refer to the real system (for 500 and 1000 years respectively). This kind of stability was already observed in [14], but with smaller regions and much shorter time intervals (around 60 years). The main improvements in this work come from the computation of the vertical families of 2-D tori for the BCP and to use initial conditions around them. A natural question is whether this is the typical behaviour around the vertical family. To this end, we have repeated the above simulations around several tori in family F3. The results are summarized in Figure 6 (more com plete results and explanations can be found in [12]). The horizontal axis refers to the z coordinate of the torus in the Poincare section, and the vertical axis contains the number of points of the initial mesh that do not escape for 1000 years (as the meshes used have the same density, the number of surviving points can be considered an estimate of the size of the stable region). Appar ently, the stability region has two big components, centered (approximately) around the values z = 0.58 and z — 0.83. Simulations for longer time intervals on the BCP model (as well as a discussion on the relevance of the total simulation time) can be found in [15] and [12]. For longer simulations on the JPL model, see [12].
207
0.2
0.3
0.4
0.S
0.6
0.7
0.8
0.9
Figure 6. Horizontal axis: value of the i coordinate. Vertical axis: an estimate of the size of the stability region (see text).
5
Normal behaviour of lower dimensional tori
Note that the numerical simulation is not a good method to detect if a given object is normally stable. For instance, an hyperbolic object can have almost coincident stable and unstable manifolds (i.e., with a very small splitting), such that the trajectories look like confined in a neighbourhood of moderate size (in fact, they keep escaping and coming back). To detect this situation, it is very convenient to have methods to compute the normal behaviour of invariant tori. Note that the linear approximation to the normal behaviour of a torus is given by a linear differential equation that depends on time in a quasi-periodic way. The study of linear systems that depend on time in a quasi-periodic way is an old topic of the theory of ODE. Here we will focus on the numerical computation of the normal behaviour of these systems. This problem has been addressed in [17] and [18] (see also references therein). Here we will explain a method for numerically compute the normal modes of an invariant curve. This method can also be applied to tori of higher dimensions (see [13] for the details). Let x(6) be an invariant curve of a n-dimensional map / , and let UJ be the corresponding rotation number. Hence, they satisfy equation (4). Then, it is not difficult to see that the linear behaviour around this invariant curve
208
is given by the dynamical system xi+i = A(6i)xi, 1
,?s
where Xi € R n . To keep the notation simple, we have used again the letter x to denote an infinitesimal displacement w.r.t. the invariant curve. 5.1
On the reducibiUty of linear quasi-periodic systems
The system (7) is said to be reducible if there exists a linear change of vari ables, Xi = C{0i)yi, such that the transformed system, Vi+i =
!+!=of+%'}
m=c l{e+w
~
w*)cW'
satisfies that the matrix B does not depend on the angles 9, i.e., that the matrix B is constant. Note that the normal behaviour of a reducible system like (7) is completely determined by the eigenvalues of the reduced matrix B. It is known that not all the linear quasi-periodic systems like (7) are reducible, but is also known that reducibiUty appears in many cases. In the concrete case of elliptic lower dimensional tori of Hamiltonian systems, it turns out that the usual proofs of existence produce reducible tori. Hence, the possible non-reducibiUty must be confined into a set of very small measure. As far as we know, the only paper that proves the existence of lower dimensional tori without using reducibiUty conditions is [19], although it does not show the existence of non-reducible tori. Let us start by considering reducible linear systems like (7). This is, we are assuming that there exists a matrix C{6), 9 G T 1 , such that the matrix B = C~1(0 + ui)A(9)C(9) does not depend on 9. Let V be the vector space {V>: T 1 i-» R n } , and let us define the translation 7^ as Tux{9) = x(9 + u>). Let us consider the following generalized eigenvalue problem: to look for couples (A, V) € C x V such that A(0)1>{0) = XT^{9).
(8)
It is not difficult to show that (see [13] for more details): 1. If A is an eigenvalue of B, then A is an eigenvalue of (8). 2. If A is an eigenvalue of (8), then there exists k € Z such that Aexp(v / -Tfcw) is an eigenvalue of B.
209 The numerical method we will use to compute the normal behavior of the invariant curves is based on computing the spectrum of a discretization of (8), as explained in Section 5.2. It is interesting to note that the reducibility hypothesis introduces strong restrictions on the spectrum of (8). Hence, it seems possible to detect nonreducibility by computing the spectrum of a discretized version of (8) and showing that it is not compatible with the reducibility condition (see [13]). It is also worth noting that all the examples included in this paper have a discretized spectrum compatible with reducibility. 5.2
Methodology
The numerical method for computing the normal behaviour of the invariant curves is then the following: once the invariant curve has been obtained (by means of the Newton method explained in Section 3), we can easily obtain the matrix corresponding to the discretization of the operator that maps ip(8) to A(6)ip(6): note that this matrix is precisely the first term that appear in the differential of F in (6), Df ■ $*-, but composed with the linear operator that, given a set of points, produces the Fourier coefficients. Let us denote this first part of the differential by A. Then, the initial eigenvalue problem is equivalent to the (standard) eigenvalue problem T^Aip
= Xrp.
(9)
Note that, as the matrix of 71„ is block diagonal (in the usual basis of sinus and cosinus, these blocks are 2 x 2 ) and very easy to write, we obtain the discretized matrix of T-UA by simply multiplying the matrices of these two discretizations. Then, it is not difficult to obtain the corresponding eigenval ues. Once these eigenvalues have been obtained, the next problem is to classify them. We recall that, if the initial system is reducible, these eigenvalues can be expressed in the form exp(i/--!kw)X, where A is an eigenvalue of the reduced matrix. Note that u is known here, and we are only interested in identifying the values A (note that, in the elliptic case, these values are not uniquely defined, see [13]). 5.3
Some results
Figure 7 displays, in the complex plane, the eigenvalues of (7) for some tori in the families F l and F2. More concretely, we display the transition from F l to F2 (see also the bottom-right plot in Figure 3). In the left plot, we observe three circles, two of them correspond to hyperbolic behaviour (those
210 ■h0.S34217767* dz-0.1915664043
»-OS350333394 dbO.OMSOMMS ' 1
. J . ■ • - »■*» a ° .
•*•■ -v -v. •tf ■*.
0.5
*
„ 0
-05
-1
1
■
-
$
fc
->v. ......*: a "
■««><> Q a o
B
a
/
«y-#,, ■
05 ■
°
*
^
w-0 5370035099 tlz-0 1810663949
■
0
■05
<*>
• **»»
' ' , «ll-»* x '
•1 -1
-0.5
0
0.5
1
Figure 7. Eigenvalues in the transition from family Fl to F2. The corresponding rotation number (w) and the value of i (dz) are included at the top of each plot.
are the ones of radius different from 1), while the circle of radius 1 correspond to neutral and elliptic behaviour. Looking at the (numerical values of) the eigenvalues of modulus one, it is easily observed that there are only two of them that are real: 0.915799019153 and 1.091942641438, while all the others can be found as a rotation (of an angle multiple of u>) of them. In the set of eigenvalues of modulus 1, one finds a couple of eigenvalues equal to one (they come from the symplectic structure of this problem), their rotations by a multiple of u, the normal frequencies and their corresponding rotations. They are studied with more detail in [13]. The sequence of the three plots in Figure 7 shows how the change of stability takes place. The study of the bifurcations that appear here is left for a future work. Figure 8 (left) shows the eigenvalues for a torus in the family F3. In the right plot, we display the modulus of the eigenvalues against an index labelling them. It is clearly seen the mild instability of this torus. In this case, there are no eigenvalues of modulus different than one that are real. This corresponds to a complex saddle, where the modulus of the eigenvalues of the saddle are 0.999897464349 and 1.000102546166. Due to the errors introduced in the truncation of the linear operator (9), we expect that high order modes will have a rather big error. Those modes are easily detected since there are no other eigenvalues obtained from them by rotations of angle leu. Moreover, they do not persist if the truncation order of the operator is increased.
211 w.O 5055149066 Oz-0 8958352596
w-0.5055149066 dz>0.8S58352S96
0. 0.99965
1 ,;.-.. 1.00005 1.0001 1.00015 ■ 1.0002 ■1
-0.5
0
0.5
1
0
100
200
300
400
500
Figure 8. A mild instability in family F3.
6
Conclusions
The main purpose of this work has been to show the existence of regions of practical stability close to the equilateral points of the real Earth-Moon system. We have focused the discussion on the L5 point, and similar results hold for Li ( [12]). Although this result is based on a numerical integration of the real system, it has been shown how some standard techniques from dynamical systems theory allow to guess, approximately, where these regions should be. So, the final numerical exploration has only been done in a small part of phase space. An interesting point is the persistence of these regions for longer time intervals because, in this case, they could contain some kind of dust or even some asteroids. Note that an asteroid inside this region must oscillate up and down across the ecliptic plane. They can reach a distance between 0.5 and 0.9 (we recall that the unit of distance of the RTBP is the Earth-Moon distance) from the ecliptic, and it seems reasonable (we recall Figure 6) that they either reach a distance close to 0.6 or to 0.8 adimensional units. In any case, note that they must spend most of their time away from the ecliptic plane and, when they cross that plane, they do it with their highest speed. Hence, we should not expect to find them very close to the triangular points. Finally, let us mention that these regions could also be interesting for some space missions, since no control is necessary to keep an spacecraft there.
212
Acknowledgements The author wants to thank interesting discussions and remarks by C. Simo. This research has been supported by the Spanish grant DGICYT PB94-0215 and the Catalan grant CIRIT 1998S0GR-00042. References [1] V. Szebehely, Theory of Orbits, Academic Press, (1967). [2] K.R. Meyer and G.R. Hall, Introduction to Hamiltonian Dynamical Sys tems and the N-Body Problem, Springer-Verlag, (1992). [3] A. Giorgilli, A. Delshams, E. Fontich, L. Galgani and C. Simo, Effective stability for a Hamiltonian system near an elliptic equilibrium point, with an application to the Restricted Three Body Problem, J. Diff. Equations 77, 167-198, (1989). [4] C. Simo, Estabilitat de sistemes hamiltonians, Mem. de la Real Acad. de Ciencias y Artes de Barcelona, Vol. XLVIII, no. 7, (1989). [5] A. Celletti and A. Giorgilli, On the stability of the Lagrangian points in the spatial Restricted Three Body Problem, Cel. Mech. 50,31-58,(1991). [6] G. Benettin, F. Fasso and M. Guzzo, Nekhoroshev-stability of L4 and L5 in the spatial Restricted Three-Body Problem, Regular and Chaotic Dynamics 3, 56-72, (1998). [7] h t t p : / / s s d . j p l . n a s a . g o v / h o r i z o n s . h t m l [8] C. Diez, A. Jorba and C. Sim6, A dynamical equivalent to the equilateral libration points of the Earth-Moon system, Cel. Mech. 50, 13-29, (1991). [9] A. Jorba and J. Villanueva, On the Persistence of Lower Dimensional Invariant Tori under Quasi-Periodic Perturbations, J. of Nonlinear Science 7, 427-473, (1997). [10] A. Jorba and J. Villanueva, On the Normal Behaviour of Partially Elliptic Lower Dimensional Tori of Hamiltonian Systems, Nonlinearity 10, 783822, (1997). [11] E. Castella and A. Jorba, Numerical computation of a family of twodimensional invariant tori in the Bicircular problem, preprint, (1999). [12] A. Jorba, A preliminary numerical study on the existence of stable motions near the triangular points of the real Earth-Moon system, preprint, (1999). [13] A. Jorba, Numerical computation of normal frequencies of invariant curves of higher dimensional maps, in progress. [14] G. Gomez, A. Jorba, J. Masdemont and C. Simo, Study of Poincar6 Maps for Orbits near Lagrangian Points, ESOC Contract 9711/91/D/IM(SC), Final Report, (1993).
213
[15] C. Sim6, G. G6mez, A. Jorba and J. Masdemont, The bicircular model near the triangular libration points of the RTBP, in From Newton to Chaos, ed. A.E. Roy and B.A. Steves, 343-370, Plenum Press, (1995). [16] A. Jorba and J. Villanueva, Numerical computation of normal forms around some periodic orbits of the Restricted Three Body Problem, Phys. D 114, 197-229, (1998). [17] A. Jorba, R. Ramirez-Ros and J. Villanueva, Effective Reducibility of Quasiperiodic Linear Equations close to Constant Coefficients, SIAM J. on Math. Anal. 28, 178-188, (1997). [18] H. Broer and C. Simo, Hill's equation with quasi-periodic forcing: reso nance tongues, instability pockets and global phenomena. Bol. Soc. Bras. Mat. 29, 253-293, (1998). [19] J. Bourgain, On Melnikov's persistency problem. Math. Res. Lett. 4, 445-458, (1997).
VARIATIONAL M E T H O D S FOR Q U A S I - P E R I O D I C SOLUTIONS OF PARTIAL D I F F E R E N T I A L E Q U A T I O N S RAFAEL DE LA LLAVE Mathematics Department, The University of Texas at Austin Austin, Texas 78712-1082, USA. E-mail: [email protected] We show how to use variational methods to prove two different results: Existence of periodic solutions with irrational periods of some hyperbolic equations in one dimension and existence of spatially quasi-periodic solutions of some elliptic equa tions.
1
Introduction
In this note, we will present two results. The first is a remark about a question of ( ! ) on whether one can use variational methods to produce solutions with irrational period of
&
du a ? + V(u) == 0 «(0,t) = «(!,*)
We will show that for some particular numbers there is an extremely simple answer. The second result is to show how one can use variational methods to produce quasiperiodic solutions to some difference equations as well as par tial differential equations and to equations involving some pseudo-differential operators. The method of proof is related to some of the results of AubryMather theory in dynamical systems and we recover the corresponding results in dynamical systems as particular results. Of course, the Aubry-Mather the ory for dynamical systems is much more developed and it includes not only results about existence of minimizing periodic orbits, but also construction of connecting orbits, etc. Many of these results do not have an analogue yet for PDE's. The connection of Aubry-Mather theory with PDE's was pointed out in ( 2 ). Results for difference equations were obtained in ( 3 ). Here we will present two proofs for partial difference equations that are based respectively in ( 4 ) and ( 5 ). These strategies can be adapted to the more complicated case of partial differential equations. As it will appear in the proof, the arguments rely just on translation invariance, comparison principles, periodicity, and a variational structure (for differential and pseudodifferential equations we will also need some compactness properties, which amount to regularity results) and, hence they work in great generality. For
214
215
the purposes of these notes we will just present the simplest cases. We point out that an important element in our proof is the consideration of the heat flow as a tool. The relevance of the heat flow for Aubry-Mather theory seems to have originated in (6) and was used in ( 7 ). 2 2.1
Periodic solutions of irrational periods for wave equations Statement of results
We recall that a number T € K is called of constant type if \Tk — l\> C|fc| _1 V k, I £ Z. It is well known (Liouville's theorem) that all irrational numbers which solve quadratic equations with rational coefficients are of constant type. We sketch a proof of the delightful classical argument. In effect, if £ is an irrational number and P(f) = a£ 2 + b£ + c = 0 with a,b,c e Z, we have P'(£) ^ 0 because, otherwise P(x) = a(x - £) 2 = ax2 - 2£ax + a£ 2 , and, then £ = —b/2a would be rational. If n, m G Z and \m/n — £| is sufficiently small, by the mean value theorem \P{n/m)\
= \P(n/m) - P(()\ < 2\P'{0\ \n/m - f|
On the other hand 2 = —Trlan + brim + cm2\ > —zi m2 m because the numerator is an integer and different from zero by hypothesis. Putting together the two inequalities above, we have:
\P(n/m)\
n/m -- £ l >
l
2|P'(0|
1 m2
Quadratic irrationals are not the only constant type numbers. It is easy to show that a number is constant type if and only if it has a bounded con tinued fraction expansion. In particular, the set of constant type numbers is uncountable (but it has measure zero). As usual, given a function / : R x R —> R satisfying f(x, t) = f(x + l,t) = f(x,t + T). We will write
= £/*,
ie2*i[kx+l(t/T)}
f(x ,t) =
fc,/€Z
and denote
H3
= {/HI/112 == E i/wif(i fc.iez
+ k2 + i2y < oo> .
216
We will also denote by D the operator ^ j — gfrTheorem 2.1 Let T be a number of constant type. Let V : R —► R satisfy i) 0 < a < V < 0, where a is any number bigger than zero and 0 depends on T in a way that will be made eocplicit ii) \V"{x)\ < K Let f : R x R -> R satisfy Hi) f{x + l,t)= f(x,t-rT) iv)
= f{x,t)
feL2 Then, there exist a u € L 2 such that:
«;(&-
-&)u + V(u) = f
b) u{x + l,t) =u{x,t
+ T)
=u{x,t)
As we will see later, the smallness hypothesis on 0 are optimal. Of course, u satisfies a) only in the weak sense at this stage (as it is typical of variational results). For rational T, ( 1,s ) contain a beautiful argument that shows that weak solutions are strong solutions. This argument does not seem to be available in our case. Nevertheless, we can obtain smooth solutions for constant type periods imposing smoothness in V and / and smallness in / . Theorem 2.2 Let T be an irrational constant type number as before. Assume that f e Hr and V € Cr+2 and that its Cr+2 norm is sufficiently small. Then, there is one u e HT which satisfies Du + V(u) = f The solution is unique in a ball in Hr around the origin. Note that, in particular, when V(0) = 0, / = 0, the only solution with period T is u = 0 contained in a sufficiently small ball around the origin. This should be compared with the results of ( 9 ) which construct solutions of D u 4- V(u) for many frequencies. 2.2
Proof of Theorems 2.1, 2.2
The main observation is that, under the hypothesis that T is a constant type number D has closed range in H'^. In effect, the range is just the closed subspace of functions / such that /o,o = 0.
217 Note that D is diagonal on trigonometric functions and that if / is such that /o,o = 0 if we denote by
□ _ 1 / = E fk,i(27ri)2(_fc2
+
p / T 2 ) Ekll{x,t)
where Ek,i{x,t) = exp[2ni(kx + l(t/T)], we have that on functions / with zero average
nn" 1 /-n _1 a/ = /. (Even if the first can only be defined for functions with zero average, we can define the second for any / and have D D / = / - /o,o-) The crucial remark is that when T is constant type we have |A;2 - (Z/T) 2 | > C> 0 V(fc, 0 € Z 2 - (0,0) Since k
2
-^
= T~2(Tk-l)(Tk
+ l)
we see that when k,l are of the same sign, we can bound \Tk — l\ > C|fc| -1 and \Tk + l\ > T\k\, therefore the product can be bounded from below by CT. When k, I are of opposite sign, \Tk -l\> T\k\ and \Tk + l\ > C]^-1. I This shows clearly that D can be defined as a bounded operator on the space of functions with zero average. Since it is clear that the average of a function in the range is zero, R a n D is precisely the set of functions in Hr with zero average. Clearly, it is a closed subspace. Note also that Ran^)- 1 - = {constant functions}. The rest of the proof of Theorem 2.1 is very similar to parts of the cor responding result in ( 1 , s ), which we reproduce for the sake of completeness. Unfortunately, the bootstrap of regularity in these references, does not go through in our case, indeed, we suspect that the conclusions of the bootstrap may not be true in our case. Note that D u + V{u) = f implies that V(u) - / G Ran(D) and that, since 0 < a < V < 0 we can find V~l and that / H < ( V - 1 ) ' < a - 1 . If we write V(u) - f = w,u = V~l(w - f) the original problem becomes
v-^w-f)-^ fv-i(w-f)+nr*w = o
218
where w € Ran(D), that is V-1(io-/)+D
i-i
toGRanO"1
This problem has a variational structure. If V is a primitive of V - 1 , then (2.2) is the equation for the critical points of
+ (l/2)(w,n _ 1 w)
S(w) = fv(w~f)
considered as a functional on the space Y — Ran(D). Since Y is closed, it is a Banach space. Now we turn to computing the second derivative of the functional. We leave to the reader the easy verification that the functional admits second Gateaux derivatives and that they are what one would expect through a formal calculation. The only non-trivial part is to check that w —► JV(w — f) is differentiable, which follows easily from the Moser estimates for Sobolev spaces (See ( 10 ) 13 §3.) We have that:
D2S{wh = fv"{w - /) 7 2 + ( 7 ,D _1 7 ) Since the second derivative of V is the first derivative of V - 1 we see that V > - / ) 7 2 > / r 'Thw2^
I
-i-i
Since ( 7 , D
7) < T - ^ h l l ^ if fTxT
>
T-1C4TT2
, the second derivative of the functional S is strictly positive definite and, therefore the functional is strictly convex. The condition above is the condition on /? that we alluded to in the statement of the theorem. Note that it involves T and its number theoretic properties. If this condition holds S is a convex functional, moreover, we have S(w)>a\\w\\l0-b\\w\\Ho-c +
(1)
for some a, 6, c € R . The functional is bounded from below and we can find a sequence wn such that S(wn) —► inf r i 5 ( w ) . The bounds (1) show that ||w n ||H 0 is bounded, hence we can extract a weakly convergent subsequence by BanachAlaoglu theorem.
219
As it is well known, convex differentiable functionate in Hilbert spaces are lower semi-continuous in the weak topology. Hence, the infimum is reached by the limit point and this limit point satisfies the variational equations. This finishes the proof of Theorem 2.1. ■ Note that the conditions in 0 we obtained are optimal for theorems of the form indicated. If we consider V(u) = Xu, and A is an eigenvalue of D , we do not expect any solutions. The eigenvalues of D get as close as desired to the value we claimed. The proof of Theorem 2.2 is quite elementary. If we write u = u\ + 6, where u G Ran(D), 6 G R = Ran(D) x , the equation D u + V(u) = / is equivalent to
Ul =
-n~y(Ul+s)-/)
6
=±JlV(ul+6)-f}
We only have to check that the right hand side of the equation is a contra diction if we consider it as an operator acting on (u\, S) with a contradiction constant that can be controlled by ||V||c"-+a. This follows from the Moser estimates standard in Sobolev spaces. See e.g. (10) 13 §3) I
3
Aubry-Mather theory for configurations on lattices
The classical Aubry-Mather theory is concerned with variational problems for functions defined on the integers. A prototype example is the variational principle
L(u) = 5Z( 1 /2)(«. - "i+i) 2 - 5(Ui) i€Z
w i t h S ( u + l) = S(u). The Euler equations become - u i + i - t i i _ i +2ui-S'(ui)
= 0
(2)
which is the well known "standard map" of dynamical systems. Notice that, besides being a variational principle for the standard map, we could consider L(u) as the energy of a chain of penduli coupled by springs. The Ui would be the angle between the iilx pendulum and the rest position, S(ui) the potential energy due to the position of the pendulum. Another alternative physical interpretation (which appeared in solid state physics) is a chain of atoms in a one dimensional periodic potential coupled by a harmonic
220
interaction. With the latter physical interpretation, this is called the FrenkelKontorova model, which is a qualitative model of deposition on the surface of (one-dimensional) crystals. The goal in this lecture is to investigate the existence of quasi-periodic solutions of similar equations when the independent variables are in Z d not just in Z. This could be interpreted physically as arrays of penduli coupled to their nearest neighbors. That is, we want to consider the variational principle for maps u : Zd -» R
L(u)= £ (l/2)( Ui - Uj ) 2 -£S(u<) and its variational equation (-At*)* - S'(Ui) = 0
(3)
where A is the discrete Laplacian (Au)* = £)ij-ii=i Uj — 2dui and, again S(u + 1) = S(u) which we will assume is a C2 function corresponding to the position and (1/2)(UJ — Ui+i) 2 the elastic energy of the spring between two consecutive penduli. One of the first results of Aubry-Mather theory is that, for every u € R, (2) has a solution Uj such that supi |UJ — wi\ < oo. This result is somewhat surprising since for the apparently very similar equation Ui+\ +Uj_j — 2u< = a with o ^ O , the solutions grow parabolically. We refer to ( n ) for an up-to-date review of Aubry-Mather theory that, nowadays, includes not only the existence of solutions as indicated above, but much deeper geometric characterizations of the quasi-periodic invariant sets, their stable manifolds and many other geometric and dynamical properties of these sets. We will present two proofs of the following result. The proofs are based on ( 4 ' 5 ). There, the reader will find more details as well as generalizations. Another proof can be found in ( 12 ). T h e o r e m 3.1 With the notations above, for every u; G Rd there exists a solution of (3) such that sup|uj - (w,i)| < oo t
The proofs we will present make use of the heat flow, whose importance was noted in ( 6 - 7 ). Our goal here is to present the argument in its simplest form so that it is clear that it is a consequence of only:
221
• Variational structure • The variational structure is somewhat local • Periodicity in space of the variational problem • Periodicity in the u's of the variational problem • Twist conditions (comparison theorems for the gradient flow) • Some mild regularity assumptions that guarantee the existence of a flow for all times as well as differentiability of functional As detailed in (4'5) there are several other contexts where we have the properties above and, hence we have an analogue of Theorem 3.1. As for the locality of the variational principle, we point out that it is very mild. It will be apparent in the details of the proof. The locality requirements in the proof in (4) are slightly stronger than those in ( 5 ). We again note that the variational structure of the problem with period icity conditions is crucial since the equation —Au + a = 0 leads to quadratic growth at oo. From the point of view of variational theory these are some what non-standard problems since the spaces of trial functions are modeled in the spaces £°°(Zd) and, the functional are only formal — the sums are not meant to converge. If we try to regularize the problem by well defined ones, we see that we have to get some control at oo to guarantee the linear growth. This is conveniently achieved by introducing the customarily called Birkhoff configurations. Definition 3.1 We say that u is a Birkhoff configuration if for every k € Zd, l e Z i o e have either ui+k + / > in V i S Zd
or
ui+k +1 < Ui V i £ Zd
That is, if we translate the graph horizontally and vertically by integers, the graph does not cross itself. Given u) 6 Rd we denote Bu, = {u | u is Birkhoff Ui - (u>, i) € t°°] . We note that Bu is not empty since (a», i) belongs to it. We also note that if u € Bu, then Ufc - UQ is restricted to lie in an interval of length 1. In effect, if Ufc — uo + ' > 0 then because of the Birkhoff property unk+k — unk + I > 0 and adding we obtain U(„+1)fc - UQ + nl > 0 (similarly for <). Since "(n+i)fc — (wi (ft + l)k) is to remain bounded independently of n we see that
222
the inequality we have to pick in Definition 3.1 is the same as that when we compare (w, k) with I. For k e Z d , I € R we introduce (Cku)i = ui+k; 1llUi = u* -f1. Note that k j C C = Ck+>, nlTln = Kl+n and Ck1ll = ft'C*. A more compact notation can be obtained as follows: We say that u > v <=> Ui > Vi for all i (similarly for <) [note that this is not a total order, there are many non-comparable elements] Using this notation, we can say that a configuration u is Birkhoff if and only if, for every k € Z d , I € Z CkTZlu >u
or
CkTllu < u .
This makes it clear that CkBu = KlBM = S w We also note that Bu is closed under pointwise limits. It will be important to note that the fact that a configuration is Birkhoff implies a priori bounds for the components. If u G Bw, then Uk+i + I — u< has to have the same sign as (w, k) + I. (Otherwise, applying the Birkhoff property, we would get a contradiction with u^ — (u, k) G £°° ) Therefore, evaluating for i = 0 we obtain that when u & Bu, Uk — UQ has to lie on an interval of length two (which we can write explicitly as a function of*). Hence, if we consider the subset of u e Bu with UQ € [a, b], applying Tychonov theorem, we conclude it is compact under the pointwise limit topology and, since it is closed, we obtain it is compact. Relatedly, if we consider two configurations as equivalent if they differ by an integer — this is natural since this does not affect the variational principle — we can choose a representation with UQ € T 1 = R/Z. Hence, BU/TZ is compact. The next ingredient in the proofs is the heat flow. This is the solution of the equation
du
= Au-- S'(u) ~dt~~ Note that, formally, this is the gradient flow for the (strictly speaking, mean ingless) variational principle. There is no problem in showing that the heat flow is defined for all time since the R.H.S. of the equation is Lipschitz on C°°(Zd). We denote by $t(u) the solution at time t of initial condition u. An important result for the proof is the comparison principle
223
Lemma 3.1 Ifu > v then $ t ( u ) > $t(*>). The proof of this lemma is not difficult. The crucial observation is that the twist condition implies that the off-diagonal terms in the variational equations of the flow are non-negative. A simple calculation shows that the elements of the matrix of the varia tional equation are: — gu.g„\ • The non-negativity of this quantity is the twist hypothesis. Now we observe that comparison is true for linear equations with diago nal coefficients (they become exponentials) and for those who have positive entries — use the Taylor expansion of the exponential. Using Trotter product formula, it is true for sums of diagonal and positive entries, hence for the variational equations of heat flow. Once the result is true for the variational equations it is true for the full equations by interpolation. We refer to ( 4 ) for more details if needed. I We also note that nlCk$T = $TCk1ll. Hence, the heat flow of a Birkhoff configuration is Birkhoff. Note also that if u € (w,t) +l°° then A(u) + 5(u) € *°°, therefore $ T B W C Bu and, moreover, the heat flow projects to BwfR,. As remarked before, the heat flow is the gradient flow of the variational principle. But recall that since L(u) = L(lZlu) the variational principle is "defined" on the compact set B^/H. Hence, the gradient should vanish at one point in the set. Unfortunately, the previous argument does not work since the variational principle L is not "defined" in anything except a formal sense. So we need to regularize somehow and study the convergence of the regularizations. The first solution that we present is that in ( 4 ). For a cube A we define:
LA(«)=
(i/2)k-^| 2 -£ 5 ( u ')
£ i€A
i€A
We note that L\(u) is a well defined function in B^/71 continuous when BU/TZ is given the pointwise convergence topology. Moreover we have: |LA(*CU) = - ] T ( - A - S ' ( U ) ) 2 + t6A
where the 7i are "gradients."
Y, t€9A
J
>(U)
(4)
224
The intuition is that for large enough cubes the boundary terms should be ignorable. A moment's reflection shows that this should be the case, except when the bulk terms EigA are very small, but in this case, we should have a critical point! More precisely, Lemma 3.2 Assume that there is no solution of (3) in Bu then, we can find e > 0, N such that ifutBu |A| > N then £ i 6 A ( - A u + S'(u)) 2 > e. We show that the negation of the conclusion implies the negation of the assumptions. The negation of the conclusions is: • 3{uW}~ = 0 ,{A n }~ = 0 such that u<»> e Bu, |A n | -> oo,
• rAn(«(»)) = E
r ck „A„(c fc "uWHr A >W) we see that the sequence Cknu^ — consisting of configurations in Bw — makes arbitrarily small. Hence TA (u n ) —► 0 where A„ C A n + i . By passing to a subsequence (the Tychonov topology is metrizable in this case) we can assume that u^ -* u*. Since TA_ (u) > TA (u) for all u, we have r A „(ii ( n + m ) ) —» o and therefore T An (u*) = 0 for all n. Hence u* satisfies (3)
I
Corollary 3.1 In the assumptions of Lemma 3.2 we can find C > 0 such that r A (tt) > C(|A| - 1) V u € Bu Given a large enough cube, divide it in cubes of size N and apply to each of the cubes Lemma 3.2. ■ It is very easy to show that the terms J* are uniformly bounded, hence if there was no critical point, applying Corollary 1 to estimate the right hand side of (4) we have: ^A«T(«)
< - C [ ( | A | - 1) + lAI 1 -' 1 /")] V u € Bu
Hence, for big enough |A|, the right hand side would be negative. But this is impossible since L A is a continuous function in the compact set B^/TZ. We, therefore conclude that there is a critical point.
225
A second proof of the result of Theorem 3.1 can be obtained by observing that if w = (1/N)(ni,..., n^) with AT G N, r i i , . . . , n j € Z we can define LN(U)=
(l/2)|«i-«j|a- £S(u«)
Y,
(5)
\i\oo
In this case, the Birkhoff orbits are periodic and satisfy: U t i , . . . , t j + N , . . . , i d = W.,,...,»j,...,t d +
1j
This implies that
jtLN{u) = - J2 (-Au + ^tt))? Hence, we can conclude that the heat flow leads to a solution of 3 when
w eQ d . To prove the result for any u we observe that if we find a sequence uA") —♦ u with u^ € Bu* solving 3, (we can assume without loss of generality \UQ \ < 1). By the fact that u' n ) in in Bu for every j e Z d we know that Uj lies in an interval of size \j\ max n (w' n ',u) that is, it remains bounded. Hence, by passing to a subsequence, we can obtain that the u^ —► u pointwise. Since the u( n )'s satisfy (3) so does u. Remark 3.2 The proofs above do not need that the interaction is nearest neighbor. They go just through without changes assuming 1.
L(u) = £
HB{u)
(6)
BCZd
where HB(U) is a C 2 function which depends only on u\B and the sum is formal. 2. The twist condition: ^ " g ^ > 0 when i ^ j . 3. Translation invariance: H-RkB(1lku) 4. Periodicity: HB{du)
= HB(u)
= HB(u)
Vu,VB C Zd,fc € Z d .
Vu,VS C Z d ,/ € Z.
5. Some conditions of decay for HB when B is large.
226
Note that (6 ) is quite common in Statistical Mechanics. HB describes the interaction between the bodies in B. The conditions of translation invariance appear very frequently in Statistical Mechanics since it is a reflection that the interactions do not depend on the translations. The twist condition appears in statistical mechanics as ferromagnetism. As for the condition on decay, we will just say that it is easy to check that the conditions are satisfied when the interaction is finite range (i.e. HB = 0 when the diameter of B is large enough. ) Once we prove the result for LA(U) = 5Zscz d diamB
227 the equilibrium configurations are equicontinuous in small enough scales. (The equicontinuity in the large is still obtained from considering Birkhoff config urations as before.) For PDE's this can be obtained using elliptic regularity theory. Some equations to which the the above conditions apply are
Au + V'(x,u) = 0 where V(x + e,u + l) = V(x,u) Vi € R d ,u € R,e € Z d , / € Z k
Y,L2i+V'(x,u)=0 t=i d
where Li are Z periodic vector fields satisfying Hormander's hypoellipticity conditions and V is in the previous case. (~A)1/2u
+ V'(x,u)
(7)
with V as above. We note that the last equation is not a PDE, since (— A) 1 ^ 2 is a non-local operator. Nevertheless all the argument can be carried though. The comparison principle for the heat flow follows because e'^ - A ^ is convolution with the Poisson kernel, as it is well known in harmonic analysis, which is positive. Then, one can use Trotter product formula as indicated in the proof. (An alternative, more general, argument, suggested by C. Gutierrez is to use the subordination identity.) Problems involving (— A) 1 / 2 similar to (7) appear in several applications. Let us just note that they appear in the study of geostrophic flows and in Levy diffusions. One can also combine this treatment of PDE's with the previous remarks and obtain quasiperiodic solutions in other homogeneous spaces besides the torus. We need to assume that the fundamental group satisfies the condition of being residually finite.
■
4
Acknowledgements
The research of the author has been supported by NSF grants. Section 3 is based on joint work with A. Candel, H. Koch and C. Radin. Both D. Bambusi and an anonymous referee made suggestions which improved the exposition.
228
References 1. H. Brezis, Periodic solutions of nonlinear vibrating strings and duality principles, Bull. Amer. Math. Soc. (N.S.), 8, 409-426, (1983). 2. J. Moser, Minimal solutions of variational problems on a torus, Ann. Inst. H. Poincare Anal. Non Liniaire, 3, 229-272, (1986). 3. M. L. Blank, Metric properties of minimal solutions of discrete periodical variational problems, Nonlinearity, 2, 1-22, (1989). 4. H. Koch, R. de la Llave, and C. Radin, Aubry-Mather theory for functions on lattices, Discrete Contin. Dynam. Systems, 3, 135-151, (1997). 5. A. Candel and R. de la Llave, On the Aubry-Mather theory in statistical mechanics, Comm. Math. Phys., 192, 649-669, (1998). 6. S. B. Angenent, Monotone recurrence relations, their Birkhoff orbits and topological entropy, Ergodic Theory Dynamical Systems, 10, 15-41, (1990). 7. C. Gole, A new proof of the Aubry-Mather's theorem, Math. Z., 210, 441-448, (1992). 8. H. Brezis and L. Nirenberg, Forced vibrations for a nonlinear wave equa tion, Comm. Pure Appl. Math., 3 1 , 1-30, (1978). 9. W. Craig and C. E. Wayne, Newton's method and periodic solutions of nonlinear wave equations, Comm. Pure Appl. Math., 46, 1409-1498, (1993). 10. M. E. Taylor, Partial differential equations. Ill, Springer-Verlag, New York, (1997), Nonlinear equations, Corrected reprint of the 1996 original. 11. J. N. Mather and G. Forni, Action minimizing orbits in Hamiltonian systems, In Transition to chaos in classical and quantum mechanics (Montecatini Terme, 1991), pp 92-186. Springer, Berlin, (1994). 12. M. L. Blank, Chaos and order in the multidimensional FrenkelKontorova model, Teoret. Mat. Fiz., 85, 349-367, (1990).
T H E SPLITTING OF INVARIANT L A G R A N G I A N SUBMANIFOLDS: GEOMETRY A N D D Y N A M I C S JEAN-PIERRE MARCO Institut de Mathematiques de Jussieu, Universiti Paris 6, 175 rue du Chevaleret-75013 Paris, Prance. B-mail: [email protected] We give precise definitions for the splitting of invariant manifolds, in an abstract setting, both from the point of view of Euclidean and symplectic geometry. We then discuss the relevance of these notions in the problem of heteroclinic connections between nearby tori, and in the estimation of transition times along chains of hyperbolic fixed points.
1
Introduction
This paper is intended to clarify the basic definitions and properties, in higher dimensions, of the so-called "splitting of separatrices" of a hyperbolic fixed point for a symplectic map in the plane. It may be viewed as an introduction to the paper 4 , where the properties of the splitting of stable and unstable manifolds of a hyperbolic invariant torus are investigated both by symplectic and analytical methods, and then applied to the problem of estimating the transition times along a chain of such tori. The content of the present paper is rather elementary but the geometric features of the splitting do not seem to have been yet investigated in any detailed way, and we find it necessary to establish firmly the grounds of this theory before getting into the quite involved and unavoidable calculations. We give in the two first paragraphs the definitions and basic properties of the "angles" between two Lagrangian submanifolds of a symplectic manifold. Paragraph 1 is devoted to the linear case: we examine first the natural and primitive notion of symplectic angle of two Lagrangian subspaces in a sym plectic vector space: being quite easily computable it proves to be the fun damental notion for our purposes, in spite of its weak geometric significance. We introduce then the refined notion of Euclidean angle, which necessitates a scalar product in addition to the initial symplectic form and encapsulates the entire geometrical information. We refer to 4 for a presentation of the transformation formulas for these two angles. The globalization of these angular notions is given in the second para graph, we study first the case of submanifolds of general symplectic manifolds (with an almost-complex structure for the Euclidean notions), and then limit the discussion to the case of cotangent bundles, the most important one being
229
230
of course the cotangent bundle T*Te of the ^-dimensional torus. In the third paragraph we specialize the preceding notions to the case of the stable and unstable manifolds of a hyperbolic torus invariant under a Hamiltonian flow. We need first some definitions borrowed from the hy perbolic theory of dynamical systems: oddly enough, the classical definitions of (partially) hyperbolic tori usually involve special normalizations in their neighborhoods; we find it unsatisfactory and take here the opportunity to fill this gap and define the hyperbolicity of an invariant torus by means of invariant decompositions of its tangent space, adapted to the hyperbolic fea tures of the linearized system. The existence of invariant stable and unstable manifolds for hyperbolic tori can then be deduced from the general theory of hyperbolic systems. One can prove furthermore that, under the natural assumption that the hyperbolic torus be isotropic, these invariant manifolds are Lagrangian (at least if the necessary condition on their dimension is met); their splitting is thus naturally defined following our preceding general study. The last paragraph illustrates two dynamical interpretations of the split ting quantities, mainly in the problem of transition chains (see 1). The first question is to pass from the homoclinic data to heteroclinic ones, the easiest problem being to find the maximal distance between two heteroclinically con nected nearby tori along a simple resonance surface. We show that, due to the existence of a natural horizontal and vertical decomposition of the space, the relevant quantity in this problem is the symplectic angle of the invariant man ifolds. We address finally the question of transition times along a given chain of hyperbolic objects. In the case of chains of hyperbolic tori in Hamiltonian systems, this problem is closely related with the question of the optimality of the so-called Nekhoroshev exponents, we refer to 2 and to forthcoming work for a general study. Here we limit ourselves to the problem of transfer around a fixed point for a symplectic map, i.e. the number of iterates needed for a given submanifold, transverse to the stable manifold of the fixed point, in or der to intersect a given ball centered on the unstable manifold, with controlled C 1 distance to the unstable manifold. This is thus a quantitative version of the usual A-lemma, which proves to be very useful for the computation of transition times along chains of hyperbolic objects.
2
Angles of Lagrangian planes
Let (W, B) be a vector space endowed with a nondegenerate bilinear form, and denote by 0(W, B) the orthogonal group relative to the form B, i.e. the
231
space of all linear isomorphisms
V(u, v) E W2.
The d-Grassmannian G<j = Gd{W) is the set of all d-dimensional linear subspaces of W. Given a pair (Vi, V2) of elements of G<j, we define the angle AB{V\, V2) of (Vi, Vb) relative to B as the orbit of the pair under the diagonal action of 0(W, B) on G j x G
= {(4(Vi),*(K 2 )) |
0(W,B)}.
Suppose now that (W, fi) is a symplectic vector space of dimension 2£, and recall that a subspace V of W is said to be : - isotropic when the induced two-form iyQ vanishes, or equivalently when V is contained in its symplectic orthogonal, - Lagrangian when it is isotropic and ^-dimensional, or equivalently when V is equal to its symplectic orthogonal. We examine in this paragraph the properties of the angles of Lagrangian subspaces of W, relative first to the symplectic form fi and then to an addi tional Euclidean form on W, satisfying some compatibility conditions w.r.t. the symplectic form ft. To begin with, we recall the basic definitions con cerning the variety of all Lagrangian subspaces, together with the classical construction of coordinates over particular open sets. 2.1
The Lagrangian- Grassmannian.
It is well-known that one can identify W with R2*, with coordinates (m,Vj), in such a way that the symplectic form fi reads ft = J2i=i dviAdui. We shall explicitely do this identification below and denote by Ho (horizontal) and Vo (vertical) the Lagrangian subspaces R' x {0} and {0} x R* respectively. The set of all Lagrangian subspaces L of R2e will be denoted by A(R2*, ft), or simply A. The symplectic group Sp(2£, R) acts transitively on A, and the stabilizer of Ho is the subgroup C C Sp(2£, R) of all symplectic lifts of linear isomorphisms of H 0 . So C ~ Gl{£,M.) and the Lagrangian-Grassmannian A is thus an algebraic variety of dimension £(£ + l ) / 2 which is isomorphic to the quotient Sp(2£,R)/Gl(£,R); we will see another identification below. The construction of local coordinates on the manifold A is classical. For getting first about the Lagrangian character of the spaces, consider a pair of ^-dimensional transverse spaces H and V of R 2 ' and denote by pn and py the projections on H and V associated with the direct sum R 2 / = H © V. Let then L be an £-dimensional subspace transverse to V. The restriction to L of
232
the projection pa is an isomorphism of L onto H, by transversality of L and V; denote its inverse by q : H —> L. Then L is the graph of the linear map JL = Pv °Q from H to V (identifying H© V with H x V); we say that the linear map JL is the equation o/L relative to H and V. Since L = {a+jt(a) \ a G H}, where a and jh(o) are considered as elements of R2t, one may also consider the map lL : H —♦ R 2 ' defined by li,(a) = a+ji,(a), giving a linear embedding of H onto L which we call the natural embedding for L, relative to H and V. Assume now that H and V are Lagrangian subspaces of R2e. Note that the dual space H* of H can be naturally identified with V, by means of the symplectic form: for v G V, simply define v*(u) = Q(u,v) for u G H, and use the fact that V is Lagrangian to prove that this gives rise to an isomorphism. Given an ^-dimensional subspace L (not necessarily Lagrangian), one can now define a bilinear form on H through the relation 7 L (a,6)
= n(a, jL(&)),
a G H, 6 G H;
the form TL is the cooordinate of L relative to H and V. Using the identification of H* with V, this is the classical identification of a linear map from H to H* to a bilinear form on H; so the correspondence L — i > TL is one-to-one. The crucial property is that 7L is symmetric if and only i/L is Lagrangian. Indeed note that n{a + jL{a),b + jL(b)) = n(a,j L (&)) + n(j L (a),&) since H and V are Lagrangian, and the symmetry of 7L is thus equivalent to the vanishing of fi on L. Let us denote by Ay the set of all Lagrangian spaces which are transverse to the space V; clearly Av is an open subset of A (classically refered to as a "Schubert cell"). The correspondence L t-> TL allows us to identify Ay with the space 5(H) of symmetric bilinear forms on H. Since «S(H) is isomorphic to R'( / + 1 )/ 2 , this identification can be considered as a coordinate map over the open set Av- Varying V and H, we obtain in this way an atlas for the Lagrangian-Grassmanian A. The following remark will be useful in the sequel. We have seen that V is naturally identified with the dual space H*; on the other hand one can consider the quadratic form Q L on H associated with TL (i.e. QL{O) = 571,(0,0)), and its derivative dQ^ from H to H* = V. Then one finds that
dQh(a)(b) = T L M ) = 7 L ( M - 0(6, jh(a)),
233
so that dQ^a) is identified with JL{O)- In other words the Lagrangian space L is also the graph of the derivative of the quadratic form QL- The form Q L is said to be a generating function for L. 2.2
The symplectic viewpoint.
The first remark concerning the possibility of denning symplectic angles is disappointing: the action of the symplectic group on A2, which would be the most natural thing to consider in order to define symplectic angles, does not lead to an interesting notion. Indeed, given two pairs (Li,L2) and (Lj,L 2 ) of Lagrangian spaces, a necessary and sufficient condition for the existence of a symplectic isomorphism ip of R2e such that ip(L\) = L'x and ^(L2) = L 2 is simply that the intersections Li n L2 and L\ n L'2 have the same dimension. In this respect this is no different from the action of the full linear group Gl(2£, R): the symplectic angle of two Lagrangian spaces is just the dimension of their intersection. One thus has to refine the primitive notion, which we do by considering linear Lagrangian embeddings, instead of abstract Lagrangian subspaces. This leads to the Definition 1 Consider two linear Lagrangian embeddings li and I2 of two (.-dimensional vector spaces E\,E2 into R2*. The symplectic angular form of the pair (li,h) is the bilinear form A, on E\ x E2, defined by At(a,b) = n(li(a),l2(b)),
aeEu
b e E2.
The geometric significance of the angular form A3 is quite weak and does not go much beyond the following transversality result, whose proof is straight forward. Proposition 2 Let Lt = li(Ei), for i = 1,2. Then KenA.
:= {a € Ei | A.{a,b) = 0,V6 € E2) = ff'fL, n L 2 )
(with an analogue for Ket2At). The spaces L\ and L2 are thus transverse in R2< if and only if the symplectic angular form is non-degenerate. The form As cannot be defined for a pair of Lagrangian subspaces (work ing with embeddings amounts to fixing bases for the two spaces), but one encounters frequently situations where such embeddings are "natural". The most important one is given by the previous construction of coordinates on A, when a decomposition R 2 ' = H © V is given from the start. The corre spondence L H-» ZL identifies Ay with the set of linear Lagrangian embeddings of H into R 2 ' whose ranges are transverse to V. As a consequence, if a pair
234
(Li,L2) in Ay is given, one can define as above the symplectic angular form of the associated pair of embeddings (l^ ,IL3)Definition 3 Let (Li,L/2) be a pair of elements of Ay. The splitting form of the pair (Li,!^) relative to the subspaces H and V is by defi nition the symplectic angular form As of the two natural embeddings l^ and ZL3 relative to H and V (here of course E\ = E2 = H^. Given a basis B H of H, the matrix S in B H of the bilinear form CTHV(LI, L2) is the splitting matrix of the pair (Li,L2), relative to H, V and B H . CHV(LI,L2)
As one would expect, the splitting form is directly related to the coordi nates relative to H and V: P r o p o s i t i o n 4 Let (Li,L2) be a pair of elements of Ay. The splitting form of the pair (Li,L2) is given by the difference of the coordinates ofh\ and L2 relative to H and V: CTHV(LI,L2)
=7L3 - 7 L l -
It is thus a symmetric form on H. The proof is a simple calculation. One can also consider the splitting form as a quadratic form on H and the preceding proposition shows that it is also (twice) the difference QL 3 — Qhi of the generating functions of Lx and L 2 relative to H and V. <7HV(LI,L2)
One would like now to interpret the splitting form in a more quantitative way. As far as transversality properties alone are concerned, one finds as before that Li and L2 are transverse in R2e if and only if the splitting form is non-degenerate, with dim(Li n L2) = dimKer ( O " H V ( L I , L 2 ) ) - But even in R 2 , where each line is a Lagrangian subspace, there is no relation between the splitting form of two lines and (for instance) their usual angle. An element L € Ay0 is given by an equation J'L(°) = 7fli ^vith 7 6 R and a € Ho = 1 (so its usual equation reads y = 7 1 ) , and the form 7L associated with L is simply given by its matrix [7] relative to the basis e\ = 1 of R. The splitting matrix of a pair (Lj, L 2 ) of A Vo , with equations j ^ k (a) = 7* a (k — 1,2), is thus just the difference [72—71]- The angle is 6 = t a n - 1 (72)—tan - '(71) € ] — n/2,w/2], (mod 7r) and it is for instance possible to make 6 go to infinity while keeping the matrix [72 - 71] constant. By contrast, the knowledge of both coordinates TLJ and TL 2 completely determines the relative positions of the two subspaces. The best way to see this will be to introduce first the notion of Euclidean angles for Lagrangian
235
subspaces, and then to show how to compute their determinations from these data. 2.3
The Euclidean viewpoint.
We now assume that a compatible complex structure is given on the ambient symplectic vector space (W, Q), meaning that W is identified with Ce, with coordinates IU, = Uj +ivj. The usual Hermitian form 2 is given by i
2(tu, w') — > J uij w'j = n(to, to') + i £l(w, w') i=i
for w and w' in C*, where II is the usual scalar product of R 2 '. The identifi cation W s; R 2 ' ~ C* is explicitely used below. The three groups relative to the forms 2, II and fi are of interest in the study: they are respectively the unitary group U(£,C), the Euclidean orthogonal group 0(2£, R) and the symplectic group Sp(2*,R). We recall the following useful relations: U(£, C) = 0(2*, R) n Sp(2£, R) = Gl(£, C) n 0(2*, R) = Gl(i, C) n Sp(2£, R). Using the expression of 2, one sees that an ^-dimensional subspace L is Lagrangian if and only if every R-basis of L, orthonormal for II, is a unitary C-basis of C ' for the hermitian complex structure. In this case, the union B U iB is a symplectic II-orthonormal basis of R2*. This is also equivalent to the equality i L = Orthn(L), where Orthn(L) is the ^-dimensional subspace II-orthogonal to L. As a consequence, the unitary group U(£, C) acts transi tively on A, and the stabilizer of the subspace Ho is 0(£, R). The manifold A is thus also isomorphic to the quotient U(£,C)/0(£, R). Let Gf denote the Grassmannian of all ^-dimensional subspaces of R 2 / . Given a pair (Li, L2) in G 2 , one defines the Euclidean angle of (Li, L2) as the orbit of (Li,L 2 ) under the diagonal action of the orthogonal group 0(2*, R) on G 2 , so the space of Euclidean angles is the quotient space G 2 / 0 ( 2 * , R ) . This notion is valid in particular when (Li, L2) is a pair of Lagrangian spaces, but in that case the intersection of the orbit of the pair with A2 is exactly the orbit of the pair under the diagonal action of the unitary group U(£, C) on A2: if ip € 0(2*, R) transforms a Lagrangian space into another one, tp is indeed unitary. As we are only concerned with the study of Lagrangian objects, the following definition is thus natural: Definition 5 Let (Li,Lj) be a pair of Lagrangian subspaces ofCe. The Eu clidean angle A(Li,L2) o/(Li,L2) is the orbit of the pair under the diagonal
236
action ofU(6,C) on A 2 . The space of Euclidean angles of Lagrangian spaces in R2t is thus the quotient A2/U(t,C). The central result of this paragraph is the following Proposition 6 Let (Li, L2) be a pair of Lagrangian subspaces of R 2< . There exists a unique unitary isomorphism VL1L3 G U(C, C), diagonalizable relatively to a unitary basis contained in Li, with eigenvalues of the form ete, 9 € ] — 7r/2,7r/2], such that ¥>LIL 2 (LI) = L2. If
237
As for uniqueness assume (p is a unitary map as in the proposition. Then consider a unitary basis JE?H0 = (a\,...,ae) of eigenvectors of ip, contained in Ho, and let pk be the (complex) eigenvalue corresponding to a*. One sees that pk ak is in the intersection of the complex line C a* with L2, which is thus at least one-dimensional, and exactly one-dimensional since L is Lagrangian. This condition determines pk in the required form and this proves the unique ness of f. The lemma is proved. One can easily extend the proof to the case L € A and non transverse to Vo (see 4 ) . Finally to prove the proposition, note that if <po is a unitary isomorphism mapping Ho onto Li, the set U(Li, L2) of unitary isomorphisms sending L\ onto L2 is given by U(U, L 2 ) = {ifo o ip o v~l
I
1
where U(HQ, ip^ (L2)) is the set of unitary isomorphisms mapping Ho onto yj0"1(L2). It is then clear how to deduce the proposition from the lemma. D One can thus identify the Euclidean angle denned by the pair (Li,L2) with the conjugacy class in U(n, C) of the unitary morphism
The splitting of regular pairs of Lagrangian submanifolds.
In this paragraph we consider pairs of Lagrangian submanifolds of a given symplectic manifold (M, fi) and globalize the preceding definitions of "split ting". We need first assume some intersection property and some regularity for the intersection: so we say that a pair of Lagrangian submanifolds is regular
238
when the intersection is a nonempty submanifold of M. The case of pairs of Lagrangian immersions is more involved: if JC* : L< —► M, i = 1,2, are two injective Lagrangian immersions, we say that the pair ( £ j , £ 2 ) is regular when the intersection / — £ i ( L ) n C^(L) is nonempty, the sets h = C~l{I) are submanifolds of L i ; and the restrictions of £» to Ij are immersions. In that case, we consider the set I obtained from the union Ii UI2 by identifying the pairs of points having the same images under the £ j , so one can consider I as a submanifold of both Li and L2, and we define naturally the map X : I —+ M as the restriction of £ j to Ii-, we say (a bit improperly) that X is the intersection of C\ and £2. Of course two Lagrangian immersions or submanifolds transverse in M and with nonempty intersection define a regular pair, but due to dimensions the intersection is then simply a union of isolated points. A more interesting case is the one of submanifolds of the same (regular) level H of a Hamiltonian function on M; the transversality relative to Ji entails that the intersections are one-dimensional, being thus the union of isolated orbits of the Hamiltonian vector field (recall that the Lagrangian submanifolds of ~H are automatically invariant under the Hamiltonian flow). But non-transverse cases will also occur naturally in nearly-integrable Hamiltonian systems, in connection with the invariant manifolds of invariant tori near multiple resonances. The first paragraph is devoted to straightforward but necessary transla tions from the linear case in a coordinate free setting. In the second one we specialize to the case when M is a cotangent bundle, which is both the most important example and makes it possible to write down -some- formulas. 3.1
Splitting forms and Euclidean angles of pairs Lagrangian submanifolds.
We begin quite generally with Lagrangian immersions: let (C\, £2) be a reg ular pair of injective Lagrangian immersions Lj —+ M, and denote by X their intersection. Since I C Lj for i = 1,2, we can define the two vector bun dles TiL», and consider their product X over I, whose fiber over x = (2:1,12) is the product TXlL\ x T I a L2. We consider finally the bundle B(X) of bi linear forms of X, whose fiber over x is Bil(T X l Li,T l 3 L2). Note that given x = (xi,X2) € I, the maps TXiCi are Lagrangian linear embeddings of the spaces TXiLn into Tj^M. Definition 8 The symplectic angular form of the pair ( £ 1 , £2) is the section A of the bundle B(X) whose value over x = (xj,X2) € I is the symplectic angular form A associated with the pair (TXlC\,TX2C2), i-e. for a G T X l L!
239 andb€TX2L2: A*(a,b) =
nx{x)(TXlC1(a),TX2C2(b)y
One has the same invariance property as in the linear case, to wit: P r o p o s i t i o n 9 If $ is a symplectic diffeomorphism of M, the maps C\ = $ o Ci define a regular pair of Lagrangian immersions with the same intersection set I, and the symplectic angular form of the pair (C\,C'2) is equal to the one of (Ci,C2), i.e. the symplectic angular form is invariant under symplectic diffeomorphisms. Let us move on to global splitting forms, assuming that we are given in the tangent bundle TM two supplementary Lagrangian subbundles 7i and V (i.e. such that the Whitney sum H. @ V is TM); denote by pu, py the associated projections from TM onto 7i and V. Let now A be the set of all Lagrangian submanifolds of M and Ay the set of Lagrangian submanifolds everywhere transverse to V. Proceeding as in the linear case again, we denote by S(7i) the bundle of symmetric bilinear forms of the subbundle 7i, and for every submanifold I of M, Sj(H) the restriction to T of S(7i). If £ is in Ay and x 6 £, the tangent space TXC is a Lagrangian subspace of the symplectic vector space (TXM,QX), transverse to Va; and one can define as in the linear case the natural linear embedding l£ of TxC associated with the decomposition (7ix,Vx) of TXM. We are thus led to the following Definition 10 Let (C\,C2) a regular pair of Lagrangian submanifolds of M, with Ci € Ay for i = 1,2, and I = ^ fl £ 2 . The splitting form of the pair (C\,C2), relative to H,V, is the section S of Sx(7i) whose value over x G l is the splitting form onx y of the pair of linear Lagrangian embeddings An analogous definition can be formulated for pairs of Lagrangian immer sions. Clearly the same equivariance property as in the linear case holds: P r o p o s i t i o n 11 Let $ be a symplectic diffeomorphism of M, and (£i,C2) a regular pair of Lagrangian submanifolds of M, with Ci £ Ay. Let us denote by primed letters the images by $. Then for x 6 J and x' = <3>(x) 6 1', the splitting form H'x, of the images (£^,£2), relative to the decomposition (?i',V'), is the image under T x $ of the splitting form S z relative to the decomposition (?i,V):
E't(s)(rs*(a)>rs*(6))=EI(a,6). The symplectic angular form is equivariant under symplectic
diffeomorphisms.
240
Let us finally state the definitions for Euclidean angles. We assume that the symplectic manifold (M, fi) is endowed with a compatible almost-complex structure, i.e. an operator J of TM such that J 2 = —I and such that the bilinear map (a, 6) i-» U(a, b) = fi(J(a),6) is positive definite in each fiber of TM, and defines thus a Riemannian metric on M. In that case TM has a Hermitian structure given by H(a, b) = II(a, b) + iQ(a, b), and we can speak of the orthogonal group bundles U and O whose fibers over x £ M are U(TXM, E) and 0(TxM,U). For x £ M, it is thus possible to define the space £x of Euclidean angles of Lagrangian subspaces of TXM, we write £ for the bundle (over M) of Euclidean angles and £x for its restriction to a submanifold I. This being said we can phrase the Definition. Let (£1, £2) be a regular pair of Lagrangian submanifolds of M, with I = £ j D £2. The Euclidean angle of (£1, £2) is the section of £x whose value at x £ T is the Euclidean angle of (TxC\,TxCi). One defines in a natural way the representative of the pair, as the section ofUj whose value overx £l is the representative of(TxC\,TxC2), and the associated sections given by the measures and the determinations. 3.2
Lagrangian submanifolds of cotangent bundles.
We now study more concretely the case of cotangent bundles (M = T*V, IT, V), endowed with its canonical Liouville exact symplectic structure (recall that the Liouville one form A on M reads ^2 Pi dqi m cotangent coordinates). One can keep in mind the primary example of the ^-dimensional torus V = Te = R ' / Z ' (or an open subset of it). This example is however not quite generic because then the bundle T*V (as well as TV), is canonically identified with V x Re. As a consequence T(T*V) is just the product (V x R') x (Rl x R'), and we have a canonical horizontal/vertical decomposition T(T*V) = 7io © Vo- For general V a vertical subbundle can always be defined intrinsically as the subspace of all vectors tangent to the fibers T*V, but the horizontal one cannot and depends on (in fact is equivalent to) the choice of a connection on V. We restrict ourselves to submanifolds defined as images of one-forms on V. In the case of a trivial bundle, these are graphs of maps from V to R* so that in all cases we say that such submanifolds have the graph property. If £ = Im/3, the form 0 defines an embedding of V into T*V with image £, hence an identification between V and £), and it makes sense to speak of the pull-back /3*(A) which is a one-form on V and in fact is just equal to 0 (this is indeed the characterization of the Liouville form A). Thus £ is Lagrangian if and only if /3 is closed, and exact if and only if 0 is. In this last case, the manifold is the graph of the derivative of a function S : V —* R which is
241
called a generating function for £. More generally, when £ is Lagrangian, we say that the cohomology class 6 € Hl(V,R) of f3 is the defect of exactness of £. In the case V = Te, the vector space H1 (Te, R) is ^-dimensional. In fact using the canonical coordinate system (fit,...,61) on R' (the universal cover of T*), the forms dfy, i = 1 , . . . , £ (or rather their cohomology classes) form a natural basis for Hl (Te, R) so that 6 € Re is an ^-vector. If we lift (canonically) the coordinate system (6i) to a coordinate system {6i,Zi) of T*Te = T ' x R ( (of the universal cover rather), then Zi is just the component on d6i, and the defect of exactness appears as (minus) the translation in the fibers one has to operate in order to pass from the initial Lagrangian manifold to an exact one. With this in mind we return to the effective "computation" of the various objects. Let (x*) be a coordinate system over an open subset U C V and let (xj,j/i) be the associated coordinate system on U* = 7r_1(J7) C T*V, the coordinates (j/i) being thus the components on the basis of one-forms (dxi). Let w denote the natural projection T(T*V) —> T*V. We can lift again the preceding coordinates and obtain a coordinate system (xi,yi, £t,f?i) on U = w~l(U*), the coordinate & (resp. 7ft) being the component on the vec tor field d/dxi (resp. d/dy,). Note that the basis (d/dxud/dyi) ofT{XtV)T*V is symplectic for the canonical symplectic form fi(I]y). Note furthermore that over the open set U the vector fields (d/dxi) and {d/dyi) generate two supple mentary horizontal and vertical subbundles Ti and V, so that one can speak of the coordinates and splitting forms relative to that decomposition. Finally V is always the restriction to U* of the vertical subbundle of T(T*V), whereas 7i depends on the initial choice of local coordinates. Given £ = Im/3 C T*V a Lagrangian submanifold x a point of V we first determine the tangent space Tp^C. As a first step, since /3 is an embedding of V onto £, one gets directly T]g(i)£ = Txji{TxV). Consider now a local coordinate system (XJ) with lift (xi,t/i,fi,77j) to T(T*V) and write /? = (ft) in these coordinates. The tangent map T(3 : TV -► T(T'V) reads:
(xi,Zi) •-» (xi,
ft(x),&,
^Q^WSJ),
and the subspace Tp^C appears as the graph of the linear map $ >-* E " = i {dPi/dXj{x)) Zj, from TXV to the vertical fiber of T0{x)(T*V). Let H and V be as above; since the basis (d/dx^d/dyi) is symplectic, we see that in the basis (d/dxi), the matrix Tx of the coordinate ^x of the Lagrangian
242
subspace T /3 ( x j£ relative to H and V is just:
rx = Particularizing to the case of an exact submanifold £s — Im dS, we say (im properly) that the map TdS : TV -» T(T"V) is the Hessian of 5 (not to be confused with the second differential ddS, from TV to TTV). The tangent space T
rx =
r d2s (x) , j dxidxj
Given now two Lagrangian subraanifolds £» = Im/?j, i = 1,2„ an intersection point x € A n£2> and a coordinate system around x = n(x), we immediately get the splitting matrix of the pair of tangent spaces (TxCi, TXC2) in the lifted coordinate system: just replace /? by /?2 — 0i in the formula above. In the case of exact submanifolds the same is true, replacing S by S 2 — Si. 4
The splitting of the invariant manifolds of hyperbolic tori
In this paragraph, (M, fi) will be a given 2£-dimensional smooth symplectic manifold, H G Cr(M, R), 2 < r < oo, a Hamiltonian on M, X # its vector field, $ its flow and <j> the time-one associated map. We say simply "invariant" for "invariant under the flow <S>". If V is an invariant submanifold of M, given a Riemannian metric || || on M, and a subbundle E of TyM invariant under T
m{Tcj>^) = M ( ( T 0 | £ ) - 1 )
where, as usual, ||Ta<£|£;|| = Sup{||Ta<£(ti)|| | u € Ba(E)}. concerning invariant tori is the following:
Our basic definition
Definition 12 Let k £ N, k > 1. An invariant torus T C M is said to be k-hyperbolic if there exists a continuous decomposition TTM = E+ © E°®E~ and a Riemannian metric \\ || on M such that : (i) dimE+ = d i m £ - = k, (ii) the tangent bundle ofTis contained in E°; (Hi) M{T4>\E+) < 1, m(T4>\B-) > 1; (iv) M{T4>lEo) = m(T 1, a hyperbolic torus is not normally hyperbolic (whereas the case k = 1 corresponds to the usual normally hyperbolic periodic
243
orbits). Nevertheless the existence of invariants manifolds for both the points of the torus and the torus itself may be proved quite easily following the (excellent) text 3 . More precisely, denoting by ET(Dk, M) the space of Cr embeddings of the unit ball Dk into M (endowed with the Cr topology), one shows that there exists a continous fonction V from V to Er{Dk,M) such that, if we denote by A 0 the embedded ball V(a){Dk): (i) a € A a for all a G V; (it) the ball A a is tangent to the fiber E+ at the point a £ V; (Hi) there exists a constant C > 0 such that, for all a £ V and b € A a , d(fn(a)Jn(b))
W+(V) = |J A„ is a C" manifold, tangent to the fiber TaV © E+ at each point a of V. There exists a neighborhood N oi V such that if p G]0,1[, there exists a constant C > 0 satisfying: W+{V)
= {beN\
d{fn(b),V)
< Cpn}.
Let us consider now the symplectic geometry of the problem. The first result about invariant tori is general and does not require any hyperbolicity assumption. Proposition 13 If T c M is a d-dimensional torus, invariant under $ , and if the flow on T is conjugated to a minimal rotation. If there exists a neighborhood O of T in M such that Q p is exact, then T is necessarily isotropic (and d < (.). See 4 for a proof. We limit ourselves in the sequel to the case of isotropic hyperbolic tori, this is indeed enough for all the classical examples, where the
244
tori are detected by means of KAM theory, and thus satisfy the assumption of the lemma. The main result is now that the asymptotic manifolds of such tori inherit their isotropic character. Proposition 14 Let T C M be an isotropic k-hyperbolic invariant torus. Then the stable and unstable balls A+ and A j associated with its points are isotropic, and the invariant manifolds W+(T) and W~[T) are isotropic. Note first that if the
Dynamical interpretations The symplectic splitting and the distance between connected tori
We briefly adress in this paragraph the question of the significance of the various preceding notions for the construction of transition chains in general perturbed systems. To put it in the context of dynamical instability, the problem is to convert homoclinic data (splitting form) for a given torus in estimates for the maximal distance at which another torus will necessarily be connected to the first; so the question is to convert homoclinic into heteroclinic properties. One can simplify and abstract the situation one step more and assume that the second torus, together with its invariant manifolds, is just a shift of the first by a translation. If one further assumes that "suitable" sections have been choosen, the problem finally reduces to the following . Problem (Persistence of transverse intersections). Given two Lagrangian submanifolds C\ and L 0 as large as possible such such that the intersection rru(C2) n C\ is non-empty when 0 < r < D(u). Here TV denotes the translation of vector v. The vector u is introduced in order to detect the possible anisotropy of the maximal distance D{u). The main point now is that one can take advantage of the natural de composition of the tangent space to T*Tl as a sum of two Lagangian spaces
245
(horizontal and vertical) and remark that only vertical translations will be of interest in the instability problem, since one tries to detect orbits drifting in the action space. The striking fact is that the use of the splitting form is par ticularly convenient for this specific choice of displacements. To see this, let us formalize the problem as follows: we consider a local symplectic chart of T ' T ' near the intersection point p, identify a neighborhood of p with R?e = H 0 ©Vo, and suppose that the two Lagrangian manifolds are, locally near p, given as the graphs of two functions f\, /2, from Ho to Vo- Note that this does not yet make use of the Lagrangian character of the manifolds. We rephrase now the problem as follows: Problem (Persistence of transverse intersections for graphs). Let f\ and fi be two C2 fonctions from Re to Rl, such that /i(0) = /2(0) = 0, and denote their graphs by G\ and G?. Assuming that the intersection of G\ and G2 is transverse and given a unitary vertical vector u € {0} x Re, find D(u) > 0 such that the intersection r r u (G2) n Gi is non-empty for 0 < r < D(u). Note that the vertical vector u may be considered as a vector in the target space Re, and that the translated graph r r u (G2) is simply the graph of the function fi + ru. The condition for the intersection of the graphs is thus equivalent to the vanishing of the difference g = (fo + ru) — f\. Here we just give the solution at a local level, as an easy application of the Picard fixed point theorem, which we recall in a suitable version: Proposition 15 Let h where R is k-Lipschitz exists a neighborhood U U onto the ball B(h(0),
be a map of a Banach space E such that h = Id + R, in a given ball B(0,p), with k 6 [0, l[. Then there of 0 in £?(0, p) such that h is a homeomorphism from (1 - k)p).
Returning to our problem, the transversality condition is equivalent to the invertibility of the derivative Dg(0) = D/ 2 (0) - £>/i(0). Let A = Dg(0) and view Re with the usual Euclidean norm. We apply the lemma to the function h = A~l g. The vanishing of g is equivalent to that of h and the lemma thus asserts that a sufficient condition for the vanishing of h is that ||/i(0)|| < (1 - k)p. Since h{0) = rA~l u, one gets that if r\\ A'1 u\\ < (1 - k)p, the intersection we are interested in is non-empty. This will give a lower bound for D(u), provided we can find an estimate for p (note that A is independent of r). One thus has to estimate the best possible radius p. Remark first that if the functions f\ and f^ are linear, the upper bound is clearly infinite. One is thus led to introduce a parameter of nonlinear distorsion, which is no other but an upper bound for the second derivatives of f\ and /2- So let 6k =
246
Sup x€R «(||I> 2 /fc(x)||) for A; = 1,2, and consider 6 = M a x ^ i , ^ ) Obviously Dh(0) = Id, and a second order Taylor expansion shows that given p > 0, the difference R = h - Id is &-Lipschitz over the ball .8(0, p), with k — p£||.A - 1 ||. One easily sees that the best choice for p is given by p - 1 = 2<5||A~1||, and one can thus ensure the existence of an intersection when r satisfies the inequality: r||A_1u|| < 46\\A
TT. l \\
We have thus shown the following: Proposition 16 With the notation as above, set A = .0/2(0) — Dfi(0), 6k = Sup l6R «(||D 2 /fc(x)||), (k = \,2), and 6 = Max((5i,(52)- Then the intersection persists at least when r < D(u) with D(U)-1=46||A-1u||||A-1||. Note that we also get an isotropic lower bound for intersection, i.e. D{u) > D with D _ 1 = 46 (||A -1 1|) 2 . Note finally that in the case when the submanifolds are Lagrangian, the matrix A is (directly related with) the splitting matrix of the pair at the point of intersection. 5.2
Angles and transfer times.
We address now briefly the question of the time of transition along a chain of heteroclinically connected hyperbolic invariant objects, i.e. the time needed for a point to move from a neighborhood of the first object to a neighborhood of the last one. For the sake of simplicity, we limit ourselves here to the case of chains of hyperbolic fixed points of a symplectic map in R2<, for which we analyze only the notion of transfer time around a single point. Given a symplectic diffeomorphim
247
with their usual Sup norm; we denote by B(a,S), Bh(ah,6), Bv(av,6) the 2 h balls of radius 6 in R ', Ho and Vo respectively, so if a = (a , av), B(a, 6) = bh(ah,6)xBv(av,6); - the (local) stable and unstable manifolds of 0 may be straightened, so W~{0) = H 0 and W+{0) = V 0) - A is transverse to Vo at some intersection point a, so in a small enough neighborhood of a it may be considered as the graph of a well-defined C 1 map / , from a ball B% = B(0,<5) in H 0 to V 0 (since 0 = ah). Definition 17 Fix now a ball B~ = B(b,6') centered on b 6 W~(0) and a positive e. We define the transfer time associated with B_ and e as the lower bound of the n'0s such that for n > no, the intersection / " ( A ) n B- contains the graph of a map g from Bt = Bh{bh,8') to V0, with furthermore \\Dg\\ < e. Note that one cannot of course require that the intersection / n ( A ) D S_ be equal to the graph of the map g, it is indeed much more complicated in general. We now show how to estimate the transfer time in a very simple, but quite general case. Consider a symplectic diffeomorphism tp of R2e, and suppose that the origin 0 is a hyperbolic fixed point for tp. Since the notion of transfer time is a purely local one, we can assume the existence of local C 1 diffeomorphism C, defined in a neighborhood of 0, such that the conjugate ip = C~lTp~C is linear and diagonalizable (provided that the necessary nonresonance conditions on the spectrum of the fixed point are satisfied). We introduce thus the matrix of xl> in a symplectic basis adapted to the unstable and stable decomposition associated with tj), namely M
'A-1 0
0 A
and we denote by A < 1 the maximum of the moduli of the eigenvalues of A. Again, when the submanifolds are Lagrangian, the derivative of / is directly related to the splitting matrix of the pair at the point of intersection Let p = (x, /(x)) be a given point of A over the ball B+. The image of p under rp is il>(p) =
(tA-1x,Af(x))
from which we deduce that the first iterate ip(A) contains the graph of the map
f\ : u (-»' Af Au
248
defined on the ball Bh(Q,8/X), with values in Vo- We have thus an immediate estimate on the derivative of f\: IP/ill < A2 ||D/||. It is also very easy to localize the intersection point a\ = ip(a) C A ] ( l Vo: since Vo is invariant under ip, one gets immediately
IM
= A(||a|| + 6 \\Df\\).
One can now iterate the operation, and assert the following P r o p o s i t i o n 18 The iterate A n = ipn(A) contains the graph of a map from the ball Bh(Q,6/Xn) to V 0 , with | | / n | | < A " ( | H | + 6||D/||),
fn,
||D/n||
The estimate on the transfer time is now straightforward: one has simply to check that the ball B*L is contained in the domain of the map / „ , together with the conditions ||/„||<<5',
\\Dfn\\<e.
One gets easily the following: P r o p o s i t i o n 19 With all the previous assumptions, the transition time no satisfies the inequality n 0 < Sup ( ^ l o g — ,
zlog
,
^ l o g - ^ J
where L — —logA. Of course this is just a foretaste of the general study, which is much more cumbersome since one has to generalize the notion of transfer time to the case of hyperbolic tori, which necessitates the introduction of the notion of ergodization time, and then to take care of the cumulative effects along the chain.
249 References 1. V.I. Arnold, Instability of dynamical systems with several degrees of free dom, Soviet Math. Doklady 5, 581-585, (1964). 2. J. Cresson, Symbolic dynamics and Arnold diffusion, Preprint, (1998). 3. M.W. Hirsch, C.C. Pugh and M. Shub, Invariant manifolds, Lecture Notes in Math. 583, Springer Verlag, (1977). 4. P. Lochak, J.P. Marco and D. Sauzin, On the splitting of the invari ant manifolds in multidimensional near-integrable systems, Inst. Math. Jussieu, preprint 220, (1999).
CROSS-SECTIONS I N THE P L A N A R AT-BODY P R O B L E M CHRISTOPHER MCCORD University of Cincinnati, Cincinnati, Ohio 45221-0025, E-mail: [email protected]
USA.
In the planar TV-body problem, N point masses move in the plane under their mutual gravitational attraction. It is classical that the dynamics of this motion conserves the intgrals of motion: center of mass, linear momentum, angular mo mentum c, and energy h. Further, the motion has a rotational symmetry. The dynamics thus takes place on a (4N — 7)-dimensional open manifold, known as the reduced integral manifold mji(M, i/). The topology of this manifold depends only on the masses M = ( m i , . . . ,mN) and the quantity v = h\c\2. In spite of the central importance of this manifold in a classical dynamical problem, very little is known about the topology of m/j(M, v). In this note, we build on the topological analysis of Smale to describe the homology of mji(M, u). A variety of homological results are presented, including the computation of the homology groups for v very large for all M; and for all v for three masses, and for four equal masses.
1
Introduction
This work is part of a recent series of studies of the homology of the integral manifolds of the AF-body problem. We build on the results of the author and Ken Meyer 6,5 to investigate the existence of global cross-sections in the planar AT-body problem. In so doing, we bring together three themes that lie at the origins of dynamical systems: celestial mechanics; cross sections and the methods of homology. Both cross sections and homology (or analysis situs) were introduced by Poincare to further his studies of the three body problem 9 ' 10 . In McCord and Meyer6, these themes were woven together to produce a homological test for the existence of a global cross section in the planar three body problem. There, the homology calculations were extended from the planar three body problem to the planar AT-body problem. We now combine those results with the homological test to investigate the existence of a global cross section in the planar AT-body problem. The AT-body problem refers to the motion of N point masses, moving under their mutual gravitational attraction. In the planar N-body problem, the particles are all constrained to a plane. As a Hamiltonian system, the dynamics conserves the quantities of center of mass, linear and angular momentum, and energy. A level set of these first integrals is the so-called integral manifold m. The system also admits a rotational symmetry about the angular momentum vector, and the resulting quotient manifold is the reduced integral manifold rns = m/SC>2.
250
251 The (reduced) integral manifold is an open algebraic set of dimension 4N — 6 (resp. 4N — 7), with the precise structure depending on the values m i , . . . TUN of the masses, the angular momentum c and the energy h. It is a classical result that, for fixed masses and non-zero angular momentum, the topology of the integral manifold depends only on the quantity v — he2 . To indicate this dependence, we will denote the integral manifolds and reduced integral manifolds by m(M, v) and mn(M, v), where M = ( m i , . . . ms) is used to display the dependence on the masses. Let <J>M,V denote the gravitational flow on H I R ( M , V). It is this parameterized family of flows that we investigate for global cross sections. In the next two sections, we review the results of McCord and Meyer6 and McCord 5 on the homology of cross sections and the homology of integral manifolds respectively. In the last section, by combining these results we see to what extent the homology of the planar integral manifolds permits or forbids the existence of a global cross section. 2
Cross-Sections and Homology
Given a manifold M without boundary and a flow (j>, a global cross section to <(> on M is a codimension-1 submanifold C such that: 1. For each point p € M there is a t(p) > 0 such that $(£(p),p) 6 C. 2. There is a continuous positive function r : C —* R such that (a) $(4,p) $ C for all p € C and 0 < t < r(p). (b) $ ( r ( p ) , p ) e C for all p G C . 3. There is an open neighborhood U of C x {0} in C x R such that $ \u is a homeomorphism from U to an open neighborhood of C in X. The function r is called the return time. The function P : C —► C defined by P(c) = $(T(C), C) is a homeomorphism and is called the Poincari map or first return map. The existence of a global cross section imposes restrictions on both the manifold and the flow. Some of the necessary conditions for the existence of a global cross section were formulated in McCord and Meyer6 as: Theorem 2.1. If the flow <j>: M xR —► M on the manifold M admit a global cross section C, then • M is a fiber bundle over S1 with fiber C.
252
• There is a long exact homology sequence - Hk+1(M)
- Hk(C)id^-
Hk{C) -> Hk(M)
-
• If M is of finite type, there exists an integer polynomial Q(t) with 0 < Q(t) < Pcit) such that PM{t) = (1 + t)Q(t). • If M is of finite type, then x(M) — 0, (the Euler characteristic of M is zero). • Hi(M;Z)
has a factor Z.
• The flow has no equilibrium points. Remark. Of finite type means that the homology of M is finitely generated. In that setting, PM (t) is the Poincar6 polynomial of M: a formal polynomial whose nth coefficient is the nth Betti number of M. The Euler characteristic is the alternating sum of the Betti numbers, and is also the value of the Poincare polynomial evaluated at t = —1. The inequality 0 < Q(t) < Pc{t) should be interpreted term-by-term: each coefficient qn is non-negative, and less than or equal to the nth Betti number of C. In this work, we apply these homological conditions to the planar integral manifolds mn(M, u). The homology of the integral manifolds was investigated by the author 5 . By combining those results with Theorem 2.1, we can draw some conclusions on the existence or non-existence of global cross sections. 3
The Homology of the Integral Manifolds
The starting point for this analysis is Smale's decomposition of the integral manifolds, via a projection onto the position coordinate 12 . 3.1
Smale's Decomposition
Smale showed that the topology of the integral manifolds is completely deter mined by the level sets of the gravitational potential VM(qi,-
■ ■qw) =
on the mass ellipsoid P(M) =
j(9i,.
N
N
1
i=\
i=l
J
253
The spaces m(M, v) and V{M) are related by the projection <£ : m(M, f) —> V{M) denned by ${q,p) = (53mi|<7i|2) 9- Since m(M, i/), V and VM are all invariant under the natural S02 action on R2N, there is a well-defined projection cp : mR(M,v) —» VR{M) and a well-defined potential function f/jw defined on the reduced mass ellipsoid or configuration space VR(M) = V{M)/S02. The level sets of UM on VR(M) will be of central importance to us. Let t(M,i/) = {q <EP(M)\VM(q) < -y/^} and let tR(M,v) = t(M,i/)/S02. Similarly, define a(M,v) = {q &V{M)\VM{q) >-y/^} and a f i ( M » = a(M,v)/S02. The boundary b ( M » in V{M) is the level set V^^-y/^v); the boundary b/j(M, v) in VR(M) is the level set Uj^^y/^v). The relation between these sets and the integral manifolds is: T h e o r e m 3.1 (Smale, T h e o r e m E 1 2 ) . There are singular orientable S2N-3 -bundles $ : m(M, v) -+ t(M, v) <j>:mR(M,v) ->xR(M,v). That is, $ and <j> map onto t(M, u) and xR(M, v) respectively, and under either $ or <j>, the pre-image of a point in the interior is a (2N — 3) -sphere, while the pre-image of a point in the boundary is a single point. One implication of this theorem is that the topology of mR(M, u) can change only at the critical values of UM, the so-called central configurations. Because of their importance, the central configurations have been intensively studied 8 , but even with this study, there are still very few values of M for which all of the central configurations are known. Of the results that we do have available to us, the most important for our present purposes is that for each M, there exist values - c o < i/_(M) < v+{M) < 0 such that all critical values of VM and UM he in the interval [— y/—v-(M), — y/—i/+(M)\X1. As we will see, for understanding both the homology of the integral manifolds in general and the existence of cross sections in particular, these values partition the parameter set into three regions. In the regions v > u+{M) and v < i/-{M), we have complete homological information; but in the region v~(M) < p < v+{M), we have only partial results. 3.2
The Homology Decomposition
Smale's description of the integral manifold in terms of the level sets of UM , when interpreted homologically, gives us a description of the homology of mR(M, v). This was carried out by the author 5 , with the results summarized here.
254
Theorem 3.2. For all values of M and v > 0, Hp(mR(M,v))*Hp(VR(M)). For all values of M and v < 0, Hp (mR(M,„))
^ Hp (vR(M, u)) © Hp_2N+3 (vR(M, v), bR(M, u)) * Hp (VR(M)) 0 # p + 1 (aR(M, u), bR(M, u)) © Hp-2N+3
(VR(M,
V), bR{M,
i/)).
These formulae will be used in two ways. When all of the central config urations of UM are known, we can compute the homology of mR(M, v) for all v. Even when the central configurations are not all known, these formulae can be exploited to provide partial information about the homological structure of mR(M, v). Both of the homological formulae are useful in these. For example, from the first formula we obtain: Theorem 3.3. If —\f^v is a regular value of UM, then x (WR(M, V)) = 0. On the other hand, it follows immediately from the second formula that: Corollary 3 . 1 . For all values of M and u,
(VR(M))).
That is, for all masses and energies, the homology of the integral manifold is at least as complicated as that of the configuration space, which is well known. It is torsion-free, and the Betti numbers are most easily expressed in terms of the Poincare polynomial (i.e. the polynomial whose ith coefficient is the ith Betti number). The Poincar6 polynomial is AT-l
Pv»(M)(t)= Y[(l + it). t=2
We know that the homology of the integral manifold is at least this rich (for negative energy, twice as rich). In some cases, we can say more, identifying ranges of values for which 0 . is an isomorphism. Corollary 3.2. IJUM is a Morse function and —\f^v is a regular value, then Hp (m R (M, u)) =* Hp (P(M)) for allp except in the ranges N-3
255
straightforward consequence of Smale's Theorem E, but the case v < v-{M) appears to be new. Corollary 3.3. For u+{M) < v < 0, Hp (m fi (M, i/)) 2 Hp (VR(M, u)) © Hp-2N+3
(PR(M,
v)).
Fort/ < u-{M), Hp (m R (M, i/)) 2 tfp (PR(M, v)) © /r aJV _5-p (P«(M, v)). 5.5
T/iree Bodies & Four Bodies mth Equal Masses
To move beyond these homological generalities, it is necessary to know all of the central configurations for a given set of masses. Unfortunately, there are very few instances in which the full set of central configurations is known. For three masses, the complete set of central configurations has been known since Lagrange, and the topology of the integral manifolds has been thoroughly investigated 2,4 . The only central configurations are the three collinear config urations (one for each ordering of the masses), and the Lagrange configuration, with the masses at the vertices of an equilateral triangle. Generically, these central configurations occur at different potential levels v\ > v2 > v$ > 1/4, which, together with v = 0, divide the parameter space into six intervals i = (0,oo), iv = (1/3, v2),
ii = ( J / I , 0 ) , m = (f 2 ,^i), v = (1/4, J/ 3 ), vi = (-00, f 4 ).
The homology groups of mR(M,i/) H.(mR{M,v)) i ii iii iv V
vi
0 Z Z
z z zy3 z
in those intervals are given in Table 1. 2 z*1 0 ti4 0 z3 0 z3 0 z 0 z* 0 1
3 0
z
0 0 0 0
4 0 t1s z1
t z
0
5 0 0 0 0 0 0
x(™«) -1 0 0 0 0 0
Table 1. H. (mR(M, v)) for Three Masses
For four bodies with equal masses, the complete set of central configu rations has only recently been established1. More precisely, a set of central configurations has been known for some time, and had been conjectured to be the complete set. Albouy verified the conjecture, establishing that there are
256
50 central configurations, occuring in four classes: 6 square configurations; 8 equilateral configurations; 24 isosceles configurations and 12 collinear configu rations. All are non-degenerate, and all configurations in the same class have the same potential value. The potential levels for the isosceles and collinear configurations are only known numerically. The information about the central configurations is given in Table 2. Configuration
Potential
Multiplicity
Index
Square Equilateral Isosceles Collinear
-7.6568Gmt -8.196lGm$ -8.2827Gmf -9.6790Gm*
6 8 24 12
0 0 1 2
Table 2. Central Configurations for Four Equal Masses
There are four critical values v$ > v\ > v-i > 1/3, which, together with v = 0, create six ranges: i = (0, 00) ii = (-7.6568Gm$,0) iii = (-8.196lGmi , -7.6568Gmi) iv = (-8.2827Gm§,-8.196lGmf) v = (-9.6790Gm§, -8.2827Gm§) vi = (-00, -9.6790Gm3) From this information, the homology of the pairs (TR(M, I/),XR(M, U)) and (afl(M, I/),XR(M, V)) can be computed (see §4 of McCord 5 for details). The table of homology groups is given in Table 3. H, i ii iii iv V
vi
0 Z Z
1
2
z5 z 55 z5 z z 5u
zB z b6 z6 zlv z 11
z z z z z
z
3 0 0
zti14
z z z
4 0 0 0 0 0 0
5 0
z
0 0 0 0
6 0 1> lu z 18
z
0 0
7 0
z 88 z6 z z1'2 0
8 0 0 0 0 0 0
X 2 0 0 0 0 0
Table 3. H. = H, (mn(M, u)) and x = X (mJt(M>")) f° r Four Equal Masses.
With the homology of the integral manifolds completely determined, it
257
will be a simple matter to determine whether or not the homological conditions for a global cross section are satisfied. 4
Cross-Sections for the Integral Manifolds
By combining the results of the previous two sections, we can determine when the homology of the integral manifolds permits or forbids the existence of a global cross section. First, because the homology of the integral manifolds is known exactly for u > i>+(M) and for v < i/_(M), we can demonstrate the following: Theorem 4.1. For all N and M, • Ifv > v+(M),
then there cannot be a global cross section.
• If v < V-{M), there is no homological obstruction to the existence of a global cross section. Proof. For v > 0, x (m R (M, v)) = x (VR(M, V)) = (~l)N(N - 2)!. In partic ular, the Euler characteristic is nonzero, so no global cross section can exist. For v+(M) < v < 0, (1 + t) occurs as a factor of (1 + t2N+3), with quotient (l-t + ... + t2N). Thus Q{t) = (1 - t + ... + t2N){l + 2t)(l + 3t) •.. (1 + {N - l)t) has negative TV - 1 coefficient of -(N - 1)!. For v
258
What happens in the range v- (M) < u < i/ + (M)? In that range, the only general result we have is the Euler characteristic statement x (WR{M, V)) = 0 of Corollary 3.3. This is consistent with the existence of a global cross section, but as this applies even to v > u+(M), where no global cross section can exist, it should not be taken too seriously as evidence for existence. In fact, in the few cases we know completely, global cross sections do not exist for v > U-(M). Namely, it was shown by McCord and Meyer6 that for all choices of masses for three bodies, there cannot be a global cross section for v > i/_(M). Similarly, we can use the computations for four equal masses to show: Proposition 4.1. For four equal masses, there exists no global cross section for v > v-(M). Proof. For v < 0, the Poincare polynomials derived from Table 3 are: 1+ 1+ 1+ i + 1+
5t + 6t2 + t 5 + 5t6 + 6t7 5t + 6*2 + 6<3 + 10i6 + 6? 5t + 6t 2 + 14t3 + 18*6 + 6t 7 u + m 2 + 1 3 + i2t 7 llf. + H i 2 + r 3 .
All are divisible by (1 + 1 ) , and the quotients are 1 + At + 2t 2 - 2t3 + 2t4 -t5 + 6t6 1 + At + 2t2 + At3 - At4 + At5 4- 6t6 1 + At + 2t2 + 12*3 - 12*4 + 12i 5 + 6f6 1 + At + 13t2 - 12t3 + 12t4 - 12t5 + 12t6 1 + lOt + 1 2 . Clearly, only the last, corresponding to v < i/_(M), has non-negative coeffi cients. □ 5
Questions and Conjectures
Drawing on those examples, it is reasonable to conjecture that for all masses, there exists no global cross section for u > i/-(M). For v < i/_(M), we have observed that there is no homological obstruction to the existence of a cross section. For N — 3 and v < v-{M), mR,{M,v) = B4 x S1. Some preliminary investigations suggest that for N = A and v < j/_(M), there may be an Sl -bundle structure. Based on these rather meager findings, we can also conjecture that for all N, there is no topological obstruction to the existence of a global cross section for v < f_(M). That is, for all masses and all v < i/_(M), m/j(M, v) has a fiber bundle structure over S 1 .
259 But even in that case, there remains the very sustantial dynamical obsta cles to the existence of a cross section. Even for N = 3, it is an open question whether or not a global cross section to the flow exists for v < v-(M). In that case, the integral manifold has three components. In each, two of the masses are close to each other, and the third is widely separated. For convenience, suppose mi and 7712 form the binary, and 7713 is far removed. If qo is the center of mass of the binary, then the angle <j> between q\ — qo and 93 — qo is a global angular variable (i.e. the S 1 factor in product B4 x S 1 ). The most natural candidate for a global cross section is the set of syzygies C = {<j> = 0}. To be a global cross section, a necessary condition is that <j> ^ 0 on C. To investigate this, a suitable choice of coordinates must be made. One such choice is to write the system in Jacobi coordinates, convert to polar coordinates, and then eliminate variables to write 7 ff(ri,ra>fli>fl2,0,*)
=
TTli 77*2
a b r ^ + ^ l + afcf^ + S } T7l2"l3
1"1
ffiima
\/rJ+Q\r\-r-2airir 3 cos(0)
y/rl+alrJ+2airir3
cos(^)'
where M, =
iral
m m i 8 Mo = mi+mj' l J 2
m3(mi
+ma) m1+m■2+m3,
c is the angular momentum, r\,r2,Ri,R2 are all radial components (i.e. are all non-negative) and
260
4. A. Iacob, Metode Topologice i Mecanica Clasica, Editura Academiei Socialiste Romania, Bucuresti, (1973). 5. C. McCord, On the homology of the integral manifolds in the planar N-body problem, To appear in Ergod. Th. & Dynam. Sys. 6. C. McCord and K. Meyer, Cross sections in the three-body problem, To appear in J. Dynam. Diff. Equations. 7. K. Meyer, Periodic Solutions of the AT-Body Problem, To appear in Lec ture Notes in Mathematics. 8. R. Moeckel, On central configurations, Math. Z. 205, 499-517, (1990). 9. H. Poincar6, New methods of celestial mechanics, Vol. 1, Periodic and asymptotic solutions. History of Modern Physics and Astronomy 13, American Institute of Physics, New York, (1993). 10. H. Poincare, Analysis Situs, J. de L'Ecole Polytechnique II Sirie 1, 1121, (1895). 11. M. Shub, Appendix to Smale's paper: "Diagonals and relative equilibria", Manifolds-Amsterdam 1970 (Proc. NufHc Summer School) Lecture Notes in Mathematics 197, 199-201, (1971). 12. S. Smale, Topology and mechanics II, Invent. Math. 11, 45-64, (1970).
E X I S T E N C E OF A N ADDITIONAL FIRST I N T E G R A L A N D COMPLETENESS OF THE FLOW FOR H A M I L T O N I A N VECTOR FIELDS JESUS MUCINO-RAYMUNDO Institute de Matemdticas UNAM, U. Morelia, N. Romero 150, Col. Centro Morelia 58000, Michoacdn, Mixico. E-mail: [email protected] Pairs of real analytic Hamiltonian vector fields X^, Xg in Poisson involution over (not necessarily compact) symplectic manifolds are considered. We address the following problem: describe how a two-dimensional orbit C of the induced (R , + ) action falls to an isolated common zero of X^ and Xg. A generalization of the Poincar6-Hopf index is introduced to describe the dynamics of Xh on C. PoincareHopf index at least three on some £, implies that X), has incomplete flow (i.e. is not well defined for all time). Completeness of the X^-flow implies Poincare-Hopf index one or two on C, and a full description of C and Xh is provided. Explicit examples of the index computation are given.
1
Introduction
Let (M,u>) be a 2n-dimensional real analytic symplectic manifold, not nec essarily compact, In > 4. Given a real analytic Hamiltonian function h : M —> JR., two desirable and nice properties of the associated Hamilto nian vector field Xh are: Existence of an additional first integral g for Xh', i-e. the existence of a real analytic function g : M —* R, functionally independent with h (almost everywhere) and in Poisson involution {h,g} = 0. Completeness of Xh\ i.e. for all initial condition the corresponding trajectory of Xh is well defined for all real time. It is elementary that if M is compact then Xh is complete. However many interesting phase spaces M are non-compact and the problem of de ciding whether a Hamiltonian vector field is complete, remains nontrivial and interesting. Recall, for example the study of escapes and non-collision sin gularities on higher dimensional manifolds in celestial mechanics 14 . On the other hand, under the assumption of complete integrability inside nonsingular compact energy levels {/i -1 (c)} many dynamical aspects of Xh are very well understood. This study began with the celebrated Liouville-Arnold the ory, but we remark the more recent work of Fomenko for dim(M) = 4 and {/i _1 (c)} a three-dimensional smooth manifold, see 6 chapter 2.
261
262 The goal of this paper is to show the existence of a link between com pleteness and existence of a second first integral near the singular points of Xh in M. We assume the following hypothesis (HI), (H2). (Hi) The existence of an additional first integral g for X/,, with p S M a common zero of X^ and Xg. Since h and g Poisson commute they define a local ( R , +)-action, giving rise to a singular analytic foliation ^ o n M , having as leaves two-dimensional orbits of the action (the orbits can be zero, one or two-dimensional). Let £ 2 be a two-dimensional orbit, where by definition the orbits are connected. We address the following questions: How can a two-dimensional orbit £ 2 "fall" to a singularity p ? How is the dynamics of Xh. inside of £ 2 ? Recall that an analytic set in M is the common zero locus of a finite collection of real analytic functions. (H2) T has a separatix £ by p, i.e. there exists a two-dimensional connected analytic set £ C M such that; p G £ is an_isolated singularity of £, and £ = U(£ 2 is the closure of two-dimensional orbits, containing zerodimensional orbits but not one-dimensional orbits. The last condition says that on £ the vector fields X^, Xg are R-linearly independent (whenever one of them is non zero), and otherwise they have common isolated zeros. For example, in a four-dimensional M, the ansatz for a separatrix is £ = {/ l - 1 (0)}n{ 5 - 1 (0)} (where we may assume without loss of generality that h(p) = 0 = (p)), whenever we can verify that Xh and Xg are R-linearly independent as above. Now we describe the shape of a separatrix £ near a singular point p. We intersect the separatrix with a small ball Be(p) C M, centered at p and of radius e > 0 in some Riemannian metric in M. It is known that for small enough ball, every £ n S £ ( p ) is homeomorphic to a finite union of copies of a cone, where the vertex correspond to p, see Section 2 and 4 . Here by a cone£j we mean any surface homeomorphic to {(x,j/,z) € R 3 | x 2 + t/2 — z2 — 0, z > 0}. Note that since the (R 2 ,+)-action is not necessarily complete, £ can assume intricate topological patterns, in particular there may exist several
263
cones {Ci} C C. Fixing a cone £< C C, there exists a resolution of Ci at p, by this we mean a continuous map ^:B(1(0)cR2-»A , (for B t a(0) an open ball centered at 0 of radius e 2 ), which is a smooth embed ding of the punctured ball B«a(0) — {0} over its image in d and Vi(0) = PThe existence of resolutions is shown in Section 2. Note that at p the cone Ci can be smoothly embedded in M or be singular (see examples on Section 7). The dynamics of Xh inside Ci is described by the pulled back smooth non-singular vector field ip*Xh on the punctured ball Bea(0) - {0}. In section 2 we will show that (H2) implies that the Poincare-Hopf index of ip*Xh at 0; PH(1>rXh,0) is a well defined positive integer number. For example, if Ci is smoothly embedded at p, then PH(ip*Xh,0) is the usual Poincare-Hopf index of Xh restricted to the submanifold d. Our main result is as follows: 1.1 Theorem. Let C C M be a separatrix by p € M coming from the closure of a two-dimensional orbits of the (R 2 , +)-action induced by Xh, Xg, satisfying (Hi) and (HZ). Given a cone Ci C C and ipi its resolution. 1.- If PH{-4>'Xh,Q) > 3, then Xh (or Xg) is incomplete. 2.- If PH(ip*Xh,0) — 2 and Xh, Xg are complete, then C is homeomorphic to a two-sphere (having p as unique singularity of Xh in C). 3.- If PH(xl>*Xh,0) = 1 and Xh, Xg are complete, then C is homeomorphic to a union of pieces of the following types: a plane (having p as a unique singularity of Xh in C), or a two-dimensional sphere (having two singularities of Xh in C), or a singular surface obtained from a two-dimensional sphere identifying two different points to one (this last one will correspond to p, the unique singularity of Xh in C). Note that parts (2) and (3) are closely related with Liouville-Arnold Theo rem in four-dimensional manifolds. Completeness is a hypothesis in LiouvilleArnold theory, see 2 p. 6 (the section called: What the Liouville-Arnold does not say). If completeness is assumed as in (2) and (3), the two-dimensional
264
orbits falling to a singularity are; a plane or a cylinder. Hence (2) and (3) describe how they are immersed in M. If we pick exactly one two-dimensional orbit in M a new version of 1.1 parts (2) and (3) is as follows. 1.2 Corollary. Let £ be a separatrix that is the closure of a two-dimensional orbit C2 of the (R 2 ,+)-action induced by Xh, Xg, whose singular points {pj} C C satisfy (HI) and (H2). If Xh, Xg are complete then Y,PH(4>;Xfl,0)
=
lor2,
here the sum runs over all the cones {d} C C having resolutions {&}. Note that compactness in M or £ is not required, only completeness for Xh and Xg on C2 is essential. For incompleteness we have: 1.3 Corollary. The existence of separatrix that is the closure of a twodimensional' orbit and with three or more singular points and hence three or more cones, implies that Xh (or Xg) is incomplete on this orbit. The basic ideas of this paper come from holomorphic dynamical systems. We follow the seminal ideas of J. C. Rebelo on the study of complete holo morphic (
265
2
Conic structure, resolutions and Hamiltonian vector fields on separatrices
Let Xh be a real analytic Hamiltonian vector field on (M,w) satisfying (HI) and (H2). We start by reviewing the description of the cone structure for a separatix £ at singular points p as in Section 1. This is due to H. Whitney, J. Milnor et al. Consider some Riemannian metric in M and let d be the induced distance function, Br{p) C M denotes the open ball centered at p with radius r. By (H2), the point p is an isolated singularity of £ as a real analytic set. We work locally in some Br(p) such that Br(p) n £ has p as unique singular point, and where obviously Br(p) n (£ - {p}) is a smooth submanifold. A key fact is: 2.1 Lemma. The distance function d : Br(p) D (£ — {p}) —* R + given by q •—> d(p,q), can have at most a finite number of critical values. Proof. See 9 p. 16 in the case when £ is an algebraic set and 4 p. 58-59 for the analytic case. □ Hence we can find a number e € (0,r) such that for all q € Bsijp) n (£ {p})> where 6 S (0, e), the distance function d is free of critical values. In consequence the tangent plane TqC <£ TqSs, where 6 = d(p,q) and Ss is the sphere of d-radius 6 centered at p. It follows that every sufficiently small sphere Ss intersects £ in a onedimensional smooth compact manifold (not necessarily connected), i.e. a finite collection of disjoint circles. For example if Ss D £ is only one circle, then £ must be a topological manifold near p. Fixing a circle in Ss f~l £, when e goes to 0, the associated family of circles describes a topological cone £i C £ . We want to show that £ 4 can be parametrized, the classical ideas are as follows (see also 9 p. 16-22 and 4 p. 58-59). There exists a smooth vector field V on the punctured ball (5 £ (p) — {p}) C M with the following properties: * The vector V(q) will point away from p for all q € Bf(p) — {p}; assume without loss of generality the existence of normal Riemannian coordinates in Be(p), let W(q) £ TqM be the tangent vector of the unique geodesic between p and q in the normal Riemannian coordinates, then the Riemannian inner product < V(q), W(q) > will be strictly positive. * The vector V(q) will be tangent to £< in all q € Be(p) n (£< - {p}).
266
* The local trajectories q(t) of V are such that <%(*)) = d(q(t),p) =t + constant , where d is the Riemannian distance. In particular for each initial condition a 6 S e n Ci the corresponding trajectory solution q(t) : (0, e2] —» M with initial condition q{e2) = a are well defined, and q(t) tends uniformly to p as t goes to zero. Hence, the restriction of the flow of V to the invariant submanifold B e (p)n (£j - {p}) maps the product (Se n£j) x (0, e2] diffeomorphically onto the twodimensional punctured ball Be2(0) - {0} C R 2 . We summarize all the above in the: 2.2 Corollary. [Existence of resolutions.] Let Ci C C be a cone in a separatrix, then there exists a resolution of Ci at p. This is a continuous map ipi:B^{0)cB.2-*Ci
,
(for an open ball B€2 (0) centered at 0 of radius e2 > 0), which is a smooth embedding of the punctured ball B€i(0) — {0} over its image in d and i>i{0) = PD Now, we want to study the induced Hamiltonian vector fields. 2.3 L e m m a . The vector fields rp*Xh and tp*Xg on the punctured ball (Bei(0) — {0}) C R 2 , have the following properties. 1.- They are smooth vector fields. 2.- They are free of zeros, everywhere tl-linearly
independent and commute.
3.- Any trajectory ofip'Xh or ii>*Xg having the origin 0 G B(i{Q) as a-limit (or u-limit) is well defined for all values of time in some interval (—oo, a) (respectively (a, + o o ) / Proof Part (1) follows since ^ is an embedding outside of the origin and part (2) follows from (HI), (H2). For (3), a trajectory of ip'Xh having the origin as a-limit corresponds to a trajectory of X/» in M having p € M as a-limit. Since X^ is real analytic at p 6 M, the existence of trajectories for time (—oo, a) follows. For trajectories having 0 as w-limit the idea is similar. D To get a classification of pairs if)\XH, i>*Xg on the punctured ball B^(0) —
267 {0}, we introduce in it a smooth Riemannian metric g defined by
gWxh,i>;xh) = 1 = gwxa,\i>;xg),
g{^xh^*xg)
= o.
Some main features of this metric are: i) Since the vector fields commute and they are orthonormal the metric g is flat (see 15 p. 261). ii) Condition (3) in Lemma 2.3 says that the metric g is geodesically complete around the origin 0 6 Be7(0). 2.4 Definition. A flat punctured ball with a frame {(S r (0) — {0}, g), F, G} is an oriented smooth punctured ball (5 r (0) — {0}) C R provided with a flat Riemannian metric g and a positively oriented orthonormal frame F , G of smooth vector fields, satisfying (i) and (ii) above. Obviously pairs of Hamiltonian vector fields Xh and Xg, satisfying (HI) and (H2) on each cone d give origin to flat punctured balls with frame
{(B€2(0) - {0}, g), Wto
tfXg}.
We say that two flat punctured balls with a frame are isometric if there exists a Riemannian isometry between them sending an orthonormal frame on the other. We want to classify the flat punctured balls with a frame up to isometry. To perform this, we introduce two isometry invariants; the Poincare-Hopf index and the residue (for simplicity to describe them we con tinue with the above notation using Hamiltonian vector fields). The Poincare-Hopf index: Consider rp'Xh as a smooth vector field on fl£a(0) - {0}. Let 7 C Bea(0) be a clockwise circle (in the usual metric of R 2 ) of radius e 2 /2 centered at 0. The topological degree of the map PfT : 7 -» q
S1
" U\xM\\ '
is well defined, where || || is the usual Euclidean norm in R 2 , and S1 C R 2 is the clockwise unitary circle. This degree is called the Poincare-Hopf index and we denote it by PH(ip*Xh,0) € Til. Note that in our case ip*Xh is not necessarily well defined at 0, hence a priori the above number is not the classical Poincar6-Hopf index of an isolated smooth zero. However, let us recall explicitly that: The number PH(ip*Xh, 0) is independent on the choice of the closed path 7.
268
Let 7i be any path homotopic to 7 in Be2(0) - {0}. The difference of the index computed with 7 or 71 is the usual Poincare-Hopf index of i/>*Xh in the interior of the region bounded by 7 U 71, that is always zero since rp'Xh. is free of zeros in Bea(0) - {0}, and the assertion follows.
PH{ti>;xh,o) = PH(ii>;xg,o). Note that on each tangent plane of the Riemannian surface (B£a(0) — {0}, g), the vector fields ip'Xh and tyXg differ by a (7r/2)-rotation (with respect to g). It is follows that corresponding PH maps differ by a (7r/2)-translation in S1, and they have the same degree. The residue: Define in B£a (0) — {0} two closed smooth one-forms a and /? as follows: a(#*h) = 1, o ( p , ) = 0,
0WXh) = 0 , /?(V**9) = 1 • Let 6 be the usual flat Riemannian metric in 1R . Using the flatness of the metric g, there exist local isometry maps 4>: V C (R 2 ,6) —» (Be2(0) - {0},g), where <j>*a = dx and >*/? = dy, hence in fact they are closed forms. Given a punctured flat ball we define its residue at 0 by
6
(H/) *where 7 is a trajectory as above. It measures the lR 2 -time required to travel along 7 under the (R 2 ,+)-action defined by Vi-X/i and i>ZXg. A pictorial description is as follows. Consider a closed path 71 in the punctured ball, constructed using small pieces of tp'Xh and/or ip*Xg trajec tories, such that is homotopic to 7 in the punctured ball E>\ (0) — {0}. The integration of a along 7! measures the time required in the dynamical system defined by ip*Xh to travel along its ip*Xh trajectory pieces in 71. The inte gration of /3 measures the same thing for ip*Xg. Note that since a and 0 are closed one-forms, the residue is the same for 71 as for 7 (since 71 is homotopic to the original 7). The Poincare-Hopf index and the residue are well defined for any flat punctured balls with frame and are isometry invariants. To show that they characterize up to isometry flat punctured balls with a frame, we require complex analysis.
269 3
Holomorphic vector fields; flat structures and classification of zeros
We start by reviewing how a holomorphic vector field gives rise to a flat Riemannian metric. This is a basic idea coming from the theory of meromorphic quadratic differentials, see 16 for the general theory and 10 , u for the relation with meromorphic vector fields. Let / = u + v7—Tw : fi C C -» C be a holomorphic function, in some domain fl. We have the following associated objects: * A holomorphic vector field
X--
-
*
>
!
■
* A meromorphic differential form bJ
=
dz f{z) •
* A pair of real smooth vector fields on Q. 5 Re(X)^(X +
X) = U |
+
, |
, , Sm(X) = J(X
8 -X)---V— ox
d oy
+ U—
,
here X means the conjugate vector field and J : TTR2 —> T H 2 is the usual complex structure in R 2 =
1
■?.)■
gx = |
I 0
« a +v a /
Some features and relations between the above objects are as follows: The vector field X has u) as time-form, namely u(X) = 1, and for any smooth trajectory 7 from ZQ to z\ in fi - { zeros of / } , the definite integral
dz /,
W)
£<E
is the complex time required to travel from ZQ to z\ along 7 under the field X. For ZQ e fi - { zeros of / } the function
-f
F(z) =
dw
/(«0
: B(z0) C Q-*
•c,
270
here z in some ball B(ZQ) around zo where / ^ 0, is well denned, holomorphic and its differential Fm satisfies:
F
-<*> - 1 ', F.(Se(JQ) = ~
, F.^m(X)) - A .
The first equality says that F is a holomorphic flow box for X, and the last two describe smooth flow boxes for $te(X) and $Sm(X). From the expressions above the commutator gives
[»e(X),9m(X)] = , 9
9
9
a,
n
- u — + u— = 0 . ox ay
Since the Riemannian metric gx has 9te(X), 5Sm(X) as orthonormal frame, it is well known that the curvature of gx is identically zero, 15 p. 261. Moreover, every map F(z) : B(z0) C (fl - { zeros of / } , gx) -» (
a ,
He(<*
» ■
a
dx
a
dy
having as trajectories straight lines of slope arg(c). 3.2 E x a m p l e . For X = 2 ^ , let u — dz/z be the meromorphic form in fi =
a, *>
s
a ax +
2/
a a^'
having as trajectories straight lines through 0 in C. The associated imaginary vector field is rs.
,
9
Sm{z-)
x
a
= -y^
+ x
a a^'
having as trajectories closed circles around zero. They correspond to closed geodesies in the cylinder S^ x R.
271 3.3 Proposition. [Analytic classification of zeros.] Let X = f{z)-§^ be a holomorphic vector field on a ball Br(0) C € centered at 0. Up to a holomor phic change of coordinates it has the following form. 1.- If X has a zero of order one in 0, then it is
Xz
lz'
where A = / ' ( 0 ) . 2.- If X has a zero of order s>2in0,
then it is
z" d 1 + \z°-->a*' for A = / -4^y the complex residue of the associated differential form where 7 is a simple clockwise path enclosing 0. Proof. See 3 p. 312.
D
Given a holomorphic vector field A" on a ball Br(0) C C, having a unique zero at 0, obviously it defines a flat punctured ball with a frame {(Br(0) —
{0},gx),Xe(X),Sm(X)}. A key fact is: 3.4 Corollary. [Equivalence of the analytic and geometric classifications.] Let X\, X2 be two holomorphic vector fields on balls BTi (0) C C centered at 0, having unique zeros at 0. Then X\ and X2 are holomorphically equivalent if and only if their associated flat punctured balls with frame are isometrically equivalent. Proof. Assume that there exists a smooth isometry v : (B r i (0) - {0},g X l ) — (B r ,(0) - {0},g* 2 ) such that v.3te(Xi) = 9te{X7) ,
i/,9m(X 1 ) = 9 m ( X 2 )
holds in the punctured ball Bri (0) - {0}, we want to show that v is holomor phic everywhere in S r i ( 0 ) . Since gxx, gx a > a n d 8 (the usual Riemannian metric in <E) are conformal and 0 is a conformal puncture in the usual Riemann surface structures of Bri (0) (see J p. 40 for the concept of conformal puncture). It follows that v is a conformal map from BTx (0)-{0} onto Br^ (0)—{0}, hence v is holomorphic and
272
bounded. Finally by the Riemann continuation Theorem, v is holomorphically extendable as v>(0) — 0. The converse is immediate.
□ Now we describe explicitly the geometry of the metric gx and the singular foliation by die(X) trajectories at any zero. Consider the Riemann sphere C P 1 = C U {00} provided with the natural flat metric in C, and where 00 is a "singular point" of this metric. Also introduce in C P 1 the singular real foliation by geodesic trajectories of g j in C, this foliation is singular at 00 (this type of singularity is called a dipole). Define a half sphere as the subset U = {z G C I Im{z) > 0} U {00} c C P 1 . A flat elliptic sector is an open neighborhood of 00 S "H (which does not contains 0). The above sector is a Riemannian surface with boundary, having a foliation by unitary geodesies. 3.5 L e m m a . [Geometry and dynamics of zeros.] Let X be a holomorphic vector field on a ball B(0) C C centered at 0. 1.- If X has a zero of order s = 1 at 0, then (P(0) — {0},gx) is locally isometric to an Euclidean cylinder 5j- x (0,00) (where 0 G B(0) corresponds to the extreme having 00 as second coordinate). The trajectories of the vector field 9te(X) assume one of the following models: center, source or sink. £.- If X has a zero of order s > 2 at 0, then (B(0) — {0},gx) is locally isometric to a suitable glue of 2s — 2 flat elliptic sectors (see the proof for full details). The trajectories of the vector field 5Re(X) define 2s — 2 elliptic sectors. 3.- The real associated vector fields satisfy PH(tRe(X),0)
= PH{^tm(X),0)
= order
ofXatO.
Proof. For case (1), using Proposition 3.3 (1) it follows that the vector field is X = A z j j . We consider two sub cases. If 5Re(A) = 0, then the linear part of 3?e(X) has pure imaginary eigenval ues. Its trajectories are circles, i.e. closed geodesies in the metric g*, giving
273 rise to a center. The flow of the orthonormal vector field Qm(X), sends closed geodesies to closed geodesies. Moving a fixed closed geodesic with the flow of ^■m(X) in the direction of the zero of X, we get the description of a cylinder. If Ke(A) ^ 0, then we consider the rotated vector field ey/~rieX, where 8 € [0,2?r), such that 5Re(ev/riflA) = 0 as in the above sub case. The Riemannian metrics coming from X and e^~^eX are isometric (see n for rotated vector fields). Finally note that the trajectories of $ie(X) correspond to open geodesies in the cylinder, describing a source or a sink. For case (2), if the order of the zero is s > 2, there are two isometric invariants of (B(0) — {0},gx) : the order s 6 IN, and A 6
€ < C | 6 > y > 0 , x>0]
.
Remove the bands from (DP1. Now we glue the boundaries, using isometrics: In A, glue X to x 4- \f^\b for x > a. In B, glue y/^-ly to a + \/^ly for y >b. Then, an open neighborhood of the point coming from oo € (DP1, in the new flat surface, is the local model for [z2/(l + Az)]g^ having s = 2 and A = a + y/^lb. The case X — a + \f^\b ^ 0 and s > 3 is now easy, following the same ideas. Assertion (3) follows by simple inspection. D Note that the complex residue A € C in 3.3 and in the proof of 3.5, is the same that the residue defined in Section 2, when we consider the associated flat punctured ball of X and the natural identification R 2 =
Hamiltonian and holomorphic vector fields
Now we are ready to describe how a pair of Hamiltonian vector fields in Poisson involution produce holomorphic vector fields on Riemann surfaces.
274
4.1 Proposition. [Real description of holomorphic vector fields.] Let L be an oriented paracompact smooth two-dimensional manifold. There exists a correspondence between: 1.- Pairs of non vanishing smooth vector fields F and G in L, everywhere TR.-linearly independent and commuting [F, G] = 0. 2.- Pairs g, F, where g is aflat Riemannian metric and F is a unitary geodesic vector field on (L,g). 3.- Pairs J, X, where J :TL —> TL, J 2 = -Id, is a smooth complex struc ture making (L, J) a Riemann surface and X is a holomorphic non vanishing holomorphic vector field. Proof. Given (1), define g using F and G as orthonormal frame. In the other direction given (2), define G as the rotated vector field of F by a positive oriented (7r/2)-angle (using the metric g). Given (3), consider g coming from X as in Section 2, and define F = 3?e(X). For the converse given g, define the complex structure J : TL —> TL as the rotation by a positive oriented (7r/2)-angle. The explicit formula for the holomorphic vector field is X = F + y/^JF, see 12 p. 116 for full details.
□ A key observation in the paper is the following: 4.2 Corollary. Given two Hamiltonian vector fields X/,, Xg on a symplectic manifold (M, w) in Poisson involution. For a two-dimensional orbit C2 C M of their associated (R 2 ,+)-action, there exists a Riemann surface structure L = (C2, J) and a holomorphic vector field X on L satisfying Xh = »e(Jf)
and Xg = *Sm(X) .
Proof. Recall that by definition £ 2 is a connected smooth manifold where X^ and Xg are always R-linearly independent (and in particular never zero). 0 A very useful result is the extension of 4.1 to cover F and G with isolated common zeros, as follows. 4.3 Corollary [A compactification procedure.] Given {(Br(0) — {0},g), F, G} any flat punctured ball with a frame (here F, G is a frame of real smooth vector fields). There exists a Riemann surface structure in the full ball Br(0) = (Br(0),J) coming from the flat structure of g, and a holomorphic vector field X on Br(0) with a zero in 0, satisfying F = fte(X)
and
G = 9m(X) .
275 Proof. Applying results 4.1 and 4.2, the punctured ball B r (0) — {0} has a Riemann surface structure and a holomorphic vector field X, such that F = Re(X) and G = 5tm{X). By results 3.4 and 3.5 the flat structure (arising from F and G) can be recognized as the conformal structure of the holomorphic vector field X. Then the Riemann surface structure J extends to 0 in a unique way, since it is a conformal puncture. Moreover, the holomorphic vector field extends to 0 (applying Riemann continuation Theorem). □ We summarize the work of Sections 2-4 in the following scheme: { Pairs of Hamiltonian vector fields tp'Xh and ijj'Xg coming from the resolution of a separatrix, satisfying (HI) and (H2). }
n { Holomorphic vector fields X = f(z) Jfe on some ball B(Q) c
Complete holomorphic vector fields on Riemann surfaces
By Section 4, the classification of complete holomorphic vector fields on Rie mann surfaces is equivalent to the classification of pairs of smooth complete and commuting vector fields F and G on two-dimensional manifolds (where F and G are R-linearly independent or have common isolated zeros). A geometric version is the following: 5.1 Corollary. Equivalence between flow completeness and geodesic com pleteness. Given X a holomorphic vector field on a Riemann surface L the following assertions are equivalent. 1.- The Riemannian surface {L — { zeros of X},gx)
is geodesically complete.
2.- The real vector fields $le(X) and^m(X) onL—{ zeros of X] are complete (i.e. for all initial conditions the corresponding trajectories are well defined for all real time).
a
276 5.2 Lemma. Let L — (L, J) be a connected Riemann surface, and a nonidentically zero complete holomorphic vector field X on L. Then, up to biholomorphism, X and L are as follow: Case: 1
Vector field X: \z£ in CP 1
2 3
z 2 £ in CP 1 £mC
4
\z%-z in C
5 6
\z£ in
Topology of the real vector field SRe(X): a source-sink in the sphere, or two centers in the sphere a dipole in the sphere parallel lines in the plane a linear; center, source or sink in the plane parallel geodesies in the cylinder parallel geodesies in the torus
Here A 6
277
6
Proof of the main results
Proof of Theorem 1.1. Let £ C M be a separatrix by p € M coming from the closure of a twodimensional orbit of the (R 2 , +)-action induced by Xh, Xg, satisfying (HI) and (H2). Basically the problem is that £ C M is not necessarily a topological manifold. Remove from the separatrix its singular points £ - {pj} obtaining a two-dimensional manifold £°. Thinking in £° as an abstract manifold, we can compactify the punctured cones (£{ - {pj}) C £° by adding one point in each punctured cone (see Corollary 4.3). Obtaining a new two-dimensional manifold L (probably disconnected), having smooth commuting vector fields, moreover by results 4.1 and 4.2 we get a Riemann surface L with a holomorphic vector field X having as real and imaginary parts the original Hamiltonian vector fields Xh and Xg. The key point is: completeness hypothesis for Xh on £ C M is equivalent to completeness for X on L, by Lemma 5.1, hence the third column in Lemma 5.2 describes real complete Hamiltonian vector fields. To show 1.1 part (1) we note that Poincare-Hopf indices coming from complete holomorphic vector fields in Lemma 5.2 are 1 or 2, since index 3 or more in a cone of £ implies incompleteness. For 1.1 part (2), note that the existence of a singularity of Poincare-Hopf index 2 corresponds to case (2) in Lemma 5.2. For 1.1 part (3), the existence of a cone having Poincare-Hopf index 1 corresponds to cases (1) or (4) in Lemma 5.2. Hence if £ a two-dimensional manifold, then is homeomorphic to a plane or a sphere. In particular if £ is compact but not a manifold, it has two cones by the same singular point p, we get that £ is homeomorphic to the singular surface obtained from the sphere identifying two different points to one. □ Proof of Corollary 1.2. In presence of zeros of Xh on £ we are speaking of cases (1), (2) or (4) in Lemma 5.2. Where the sum of the Poincare-Hopf indices is 1 or 2. □ 6.1 Remark. On the analyticity assumption: Note the results 1.1 and 1.2 can be stated for smooth Hamiltonian vector fields Xh, Xg on smooth symplectic manifolds (M,u>) satisfying (HI) and replacing (H2) as follows. (H2') the singular foliation T associated to Xh, Xg has a smooth separatrix £ by p, i.e.
278
£ is the common zero locus of a finite collection of smooth functions, £ = UjC? is the closure of two-dimensional orbits, containing zerodimensional orbits but not one-dimensional orbits. conclusion in Lemma 2.1 remains true for £ as above. 7
E x a m p l e s of t h e index c o m p u t a t i o n
Our first example uses normal forms theory for four-dimensional integrable Hamiltonian systems at simple singular points, due to J. Moser, H. Riissmann (for the analytic case), and L. H. Eliasson (for the smooth case). Assume that X/, and Xg form an integrable Hamiltonian system in a four-dimensional manifold M. Let p £ M be a simple singular point of the Hamiltonian vector field X^. Locally we can identify p with 0 € R by a symplectic change of coordinates, where R 4 = {(xi,x 2 ,2/1,1/2)} and u = dx\ Adj/i +dx2 Adj/2 describe the canonical symplectic four-dimensional manifold. 7.1 Theorem. In a neighborhood of 0 € ( R 4 , u ) a simple singular point of a real analytic integrable Hamiltonian vector field Xh with two degrees of freedom there exist real analytic (smooth) coordinates in which h = Ai£i + A2& + h(£i, &) > 9 = M1C1 + M262 + 9(£1,62) , where the functions h, g are real analytic (smooth) and their initial terms are of second degree at 0. The quadratic functions £1 and £2 depend on the type of the point and have the form: 1.- £i = 5(1? + yf), £2 = 5(^2 + 2/f) (elliptic point). 2- i\ — \(?\ + 2/i)> £2 = £22/2
(saddle-centerpoint).
3- £1 = X1J/1, £2 = Z22/2 (saddle point). 4.- £1 = xiyi + Z22/2, 62 = zii/2 - 12I/1 (saddle-focus
point).
Where Ai/i 2 - A2M1 ^ 0. Proof. See L. M. Lerman and Ya. L. Umanskiy expositions in p. 27.
7
p. 515 and
8
D
Our assertion is: 7.2 Corollary. For saddle-focus points there are two smooth separatrices by
279
0 € 1R , and the Hamiltonian vector field X/, satisfies PH(ii>;xh,o)
= i
on both separatrices. Proof. Let us begin by sketch the computation in the simplest case. Consider the Hamiltonian functions: h = ii - xij/i + x2y2 , 9 = £2 = x\y2 - x2y\ , that is the particular case Ai = /i 2 = 1, A2 = /ii = 0 and h = 0 = g in Theorem 7.1. The Hamiltonian vector fields are:
Xg =x2-
d
d
d
d
dx 1
0X2
oy\
Oy2
d ox 1
d d Xi— + y20x2 ay 1
d yi^— oy2
There are two separatrices by 0 6 R 4 given by A = {2/1 = 0 = 2/2} and £ 2 = {xi = 0 = x 2 } . The Hamiltonian vector fields assume the expressions
y\Ah
=
d -xi-OX\
d
d
d
OX2
OXi
0X2
in C\, and W2xh
= y\^—+y2^—
,
ip2xh
= y2x
vi^r
oy\ oy2 oyi Oy2 in £ 2 . They have Poincare-Hopf index 1. In particular, each separatrix £i — {0} C E is isometric to a flat cylinder. Note that ^JX/, and tl>\Xg on C\, are the real and imaginary vector fields associated to the holomorphic vector field
making z = x\ + \/—1^2 see Example 3.2, and similarly for £2.
280
Assume Ai/x2 - A 2 ^i i=- 0 and the vanishing of the higher order terms h, p for the next case in the proof. Use that the above separatrices persist, and note that ip*Xh, ip*Xg are linear combinations of the linear Hamiltonian vector fields in the above case. Hence Poincare-Hopf indices jire also 1. We leave the general case (when the higher order terms h and g are non zero), as an exercise for the interested reader. D
7.3 Example. An incomplete Hamiltonian vector field X^ having a singular separatrix £ and PH(rXh,0)
= 2.
Let be the Hamiltonian functions h = x\ - x\ - y\ 4- 3yiy| and g = -2xxx2
- 3y 2 y 2 + y2
in (R 4 ,u;). The Hamiltonian vector field is Xh = (3y2 - 3 y | ) ^ — - 6yiy2^+ 2xx— - 2x2 — . dx 1 dx2 dyt dy2 The vanishing of the Poisson bracket is an easy computation 0 = {h,g} = - ( - 6 y l 2 / 2 ) ( 2 i i ) + (-3y 2 + 3 j £ ) ( - 2 s 2 ) - ( - 3 y ? + 3 T / | ) ( - 2 X 2 ) + (6y 1 y 2 )(-2x 1 ) . There exists a separatrix by 0 G R 4 , given by £ = {/i-1(0)}n{5-1(0)}. The verification of (HI) and (H2) for the singular point 0 € £ is very simple. Find an explicit resolution for a singular analytic set, usually is a very difficult (and unpleasant) task, fortunately for this separatrix the resolution is given by i> : R 2 - . £ C R4 3 2 (*, s) >-> {-t + 3ts ,3t2s - s 3 , t2 - s 2 ,2ts) . Note that rp is an embedding of R 2 — {0} over £ - {0}. Consider an auxiliary vector field F = A(t, s) J^ 4- B(t, s)J^ in R 2 , satis fying rp,F = Xh on £, straightforward computation shows that
F = rXh = (-t2 + s2)^-2ts-^-
.
281 This vector field can be recognized as the real vector field 3?e(X) associated to the holomorphic vector field
X= -
2
# . dz '
introducing coordinates z — t + \ / - l s . Hence ip*Xh has Poincare-Hopf index 2 at 0 € R 2 . If we note that C is unbounded in R 4 , it follows using Theorem 1.1 part (2) that Xh is an incomplete Hamiltonian vector field on R 4 . 7.4 E x a m p l e s . Six-dimensional systems. From 7.2 higher dimensional examples are easy to see. Let (R6,u>) be the canonical symplectic six-dimensional manifold, consider the Hamiltonian functions: h = xm+
Z22/2 + x3h(x3, y3) , g = x-^y2 - x2y\ + y3g(x3,y3)
,
where {x3/i(x3,y3),2/3<7(x3,y3)} = 0 . There are two analytic separatrices by 0 € R 6 given by £ i = {j/i = 2/2 = x 3 = y3 = 0} and C2 = {xi = x2 = x3 = y3 = 0} . The hypothesis (HI) and (H2) hold on both separatrices and the PoincareHopf index is 1 (see proof of 7.2). A more interesting example is in the work of L. H. Eliasson 5 p. 33, let us follow him word for word. Consider the Lagrangian spinning top, having principal moments of inertia h = I2^ h, it is rotational invariant around the third principal axis of inertia and the gravitational field is invariant around the vertical. It can be described by a Hamiltonian system on T*SO(3), having two additional first integrals Q3 and Q3 from the rotational invariance. The vertical positions is a circle T in the configuration space. In local coordinates (R 6 ,w) the Hamiltonian functions are: _.
H{xi,...,y3)
.
l . o
o.
+^-y3
Qi(xi,...,y3)
. l o
TYl.
TTi o
o
= — ( y j + y$) + ( — y 3 - -^)x1 - —x2 + + 03(xi,x2,yuy2)
1
jy3x^y2
,
- xiy2 - x2yi + 7,yi(A ~ xl) + y3 + 03(xux2,yi,y2) Q3(xu...,y3)
= y3 ■
,
282
If we fix the value Q f (xi,...,i/3) = 2/3 = C2, then if and Q§ become func tions only on the (xi,X2,2/i,i/2)-space. For the point (xi,£2,£3,^1,1/2,1/3) = (0,0, c\, 0,0,C2), a separatrix is given by the equations £ = {X3 = CU 2/3 = C2, ^ _ 1 ( C 3 ) , (Qf) _1 (C4)} , where C3, C4 are suitable constants. If 2/3 > 4m/i, then L. H. Eliasson asserts that under suitable symplectic coordinates the Hamiltonians are generated by h = xiyi + x2y2 , 9 = xi2/2 - x2y\ . From 7.2 the Poincare—Hopf index on the separatrix is 1. Acknowled gments The author thanks Professors D. del Castillo, J. Seade and the referee by very useful comments. Partially supported by DGAPA-UNAM and Conacyt 28492-E. References 1. W. Abikoff, The Real Analytic Theory of Teichmuller Space, Lecture Notes in Mathematics 820, Springer Verlag, (1980). 2. M. Audin, Spinning Tops, Cambridge Studies in Advanced Mathematics 51, Cambridge, (1996). 3. L. Brickman and E. S. Thomas, Conformal equivalence of analytic flows, J. Diff. Equations 25, 310-324, (1977). 4. D. Burghelea and A. Verona, Local homological properties of analytic sets, Manuscripta Math., 7, 55-66, 1972. 5. L. H. Eliasson. Normal forms for Hamiltonian systems with Poisson commuting integrals - elliptic case, Comment. Math. Helvetici, 65, 435, (1990). 6. A. T. Fomenko, Integrability and Non-integrability in Geometry and ! lechanics, Kluwer, (1988). 7. L. M. Lerman and Ya. L. Umanskiy, Classification of four-dimensional integrable Hamiltonian systems and Poisson actions of I t in extended neighborhoods of simple singular points, Russian Acad. Sci. Sb. Math., 77, 511-542, (1994). 8. L. M. Lerman, and Ya. L. Umanskiy, Four-Dimensional Integrable Hamiltonian Systems with Simple Singular Points (Topological Aspects), Translations of Math. Monographs, Vol. 176 Amer. Math. Soc. (1998).
283
9. J. Milnor, Singular Points of Complex Hypersurfaces, Annals of Mathe matical Studies 69, Princeton University Press, (1968). 10. J. Mucino-Raymundo, Complex structures adapted to smooth vector fields, Preprint (1997). 11. J. Mucino-Raymundo and C. Valero-Valdes, Bifurcations of meromorphic vector fields on the Riemann sphere, Ergodic Theory and Dynamical Systems, 15, 1211-1222, (1995). 12. K. Nomizu and S. Kobayashi, Foundations of Differential Geometry Vol. II, Wiley-Interscience, (1969). 13. J. C. Rebelo, Singularites des flots holomorphes, Ann. Inst. Fourier, Grenoble, 46, 411-428, (1996). 14. D. G. Saari and Z. Xia, Off to infinity in finite time, Notices of the Amer. Math. Soc, 42, 538-546, (1995). 15. M. Spivak, A Comprehensive Introduction to Differential Geometry Vol. II, Publish or Perish, (1979). 16. K. Strebel, Quadratic Differentials, Springer Verlag, (1984).
SIMPLIFICATION OF P E R T U R B E D H A M I L T O N I A N S T H R O U G H LIE T R A N S F O R M A T I O N S j . P A L A C I A N A N D P. Y A N G U A S
Departamento de Matemdtica e Informdtica, Universidad Publica de Navarra, 31006 Pamplona, Spain E-mail: [email protected], [email protected] This paper deals with the reduction of perturbed Hamiltonian systems. For this purpose we use a technique based on Lie transformations. Under certain conditions, the number of degrees of freedom of such a system is reduced, by extending an integral of the unperturbed part to the whole transformed system, up to a certain order of approximation. The procedure is illustrated for polynomial, as well as for Keplerian-type systems.
1
Introduction
The idea of reducing a perturbed system by the extension of an integral of the unperturbed part to the whole system is valid not only for Hamiltonians, but also for any system of differential equations. Nevertheless, here we focus on systems of Hamiltonian type, that is, scalar functions of the form H(x) = 7io(x) + HP(x).
(1)
Reducing H. consists in providing a symplectic change of variables x = X(y) which transforms (1) into fC, where K(y) = K.a{y) + fCp(y) with £n = HQ and fCp simpler than Tip. Poincare can be considered as a pioneer in developing a method to simplify systems of ordinary differential equations, but not necessarily of Hamiltonian nature. His studies were followed by Birkhoff, who concentrated on the Hamil tonian case and then by Whittaker. They developed a theory for polynomial systems. Recently, there has been a resurgence of this subject. Among oth ers, we emphasize the studies of Kummer 11 , Cushman et al.3 and Meyer12 in the field of polynomial-type Hamiltonians. Some authors have also dealt with reduction in the frame of perturbed Keplerian systems and we name, for instance, the works by Deprit 5 and Cushman 2 . The usual way of approaching the problem of constructing K. from H is by calculating the corresponding normal form. The tools used for this purpose are Lie transformations. Indeed, after developing Ti up to a certain order, one executes a Lie transformation to calculate K, step by step, from K.\ to K-i for a certain positive integer L. First of all one has to scale the system defined by Ti so as to introduce a dimensionless small parameter e. The way 284
285
of doing this depends strongly on the type of system one deals with. Some interesting examples of different scaling strategies can be looked up in the paper by Meyer 13 . After scaling and dropping the primes to avoid a tedious notation, H can be written as
n(x;e) = J2-iHi(*),
(2)
«=o zwhere Hamiltonians 7i»(x) are analytic functions in x. In this way, the prob lem can be placed in the context of a General Perturbation Theory for Hamiltonian systems, since H is composed of two distinct parts, the main Hamiltonian HQ and the small perturbation eH\ 4-e 2 W2/2! + The problem of building formal integrals for Hamiltonian systems has received a wide treatment in the last thirty years and the results have been applied in fields such as Molecular Physics or Astrodynamics. For instance, in galactic models of three degrees of freedom, the search for the third inte gral —the first and second integrals being, respectively, the total energy of the system and the angular momentum— becomes very important to analyse the onset of chaos. A numerical treatment of this problem appeared in the paper by Henon and Heiles9. The book by Gutzwiller8 contains some chap ters devoted to analytical and numerical studies of chaos in many dynamical systems. Recently, Celletti and Giorgilli1 have provided an algorithm to build the n — 1 formal integrals (apart from the Hamiltonian itself) for polynomial Hamiltonians. They proceed constructively by calculating the integrals de gree by degree, starting at degree two and treating separately the resonant and non-resonant situations. However, the construction of formal integrals does not permit to reduce, at least directly, the original Hamiltonian so as to extract useful conclusions from the reduced one. Our approach consists in generalizing the concept of normal forms by selecting a function Q(x) and proposing thereafter a symplectic change of variables. This transformation is built in such a way that the usual Poisson bracket of each term of the transformed system, £ j , and the function Q be zero for all i = 1 , . . . , L. Moreover, if in addition to the above, one picks Q as an integral of 7io> t n e n Q becomes an integral of K, independent of it, that is, we compute formal integrals by reducing the initial Hamiltonian through one or several Lie transformations. We have divided the paper into six sections. Sec. 2 contains a description of the Lie transformations techniques for Hamilton systems and the applica tion to the calculation of normal forms. In Sec. 3 we propose the method, showing that it is a generalization of the usual procedures based on normal
286
forms. Sec. 4 is devoted to the polynomial case, whereas Sec. 5 deals with perturbed Keplerian systems. We have coded the algorithms for the Lie trans formations, as well as for the construction of integrals, with MATHEMATICA 3.0 22 . Finally, Sec. 6 is dedicated to the conclusions. 2 2.1
Normalization by Lie Transformations Lie transformations
In Hamiltonian Theory, the aim of Lie transformations is to generate a symplectic change of variables depending on a small parameter as the general solution of a Hamiltonian system of differential equations. The method was first proposed by Deprit 4 and can be stated as follows. An analytic Hamiltonian function (2) depending on a small parameter e is transformed into another Hamiltonian
£(y; 0
=
id(v)
=
t=0
•2-' 75" n^(y),
through a generating function
W(x;e)
= £-rW, +1(*), l »=o -
following the recursive formula
= w8.i,)+EoQ{w?-V).^+i}.
7#> =
(3)
with i > 0, j > 1 and li\ ' = Hi. Note that W(x;e) is conserved under the transformation and thus, it can also be expressed as VV(y;e). The recursion process can be summarized in the so-called Lie triangle, see Figure 1. Hence, Eq. (3) yields now the partial identity £« 0 (W,) + >Ci = W i)
(4)
where Hi collects all the terms known from the previous order and £ « 0 denotes the Lie operator (linear operator) associated to the Poisson bracket of two functions. More concretely, for two scalar fields P and Q defined over an open domain of R 2 n , the Poisson bracket of P and Q is given by the relation CP(Q)
== {P ,Q}-
fdPV =
\dx)
dx'
287
K-o
Kz
K2 II
(0)
(1)
(0)
W
n
H
(2)
K
(3)
/
/
Hi
/
/
/
H
(0)
(3)
(2)
1 /' I
(4)
W,
H\
n\
/ n2
(i)
H
y (0)
w;
Figure 1. At each order i > 1 of the process, the diagonal H, with j + k = i is built starting with WJi\ and finishing with H^'l). Note that HQ cannot be determined unless Wi be previously known.
where J denotes the skew-symplectic matrix of dimension 2 n x 2 n . In iden tity (4), called the homology equation, Wj and K-i must be determined ac cording to the specific requirements of the Lie transformation one performs. Now, the transformation X(y;e), which relates the "old" variables, x, to the "new" ones, y, is a near-identity symplectic change of variables. Explic itly, it is given by X
e*
= y + Y^i\cw(y),
(5)
where the Lie operator applied to a vector y means that it is applied to each component of y. Besides, the notation £yy refers to the application of £>v i times, that is, £yv(y) ~ ^-w(£yv1(y))> if t > 2. Consequently, Eq. (5) gives the set of variables x in terms of y with the use of the generating function W. Similar formulae can be used to obtain the inverse transformation. The above method is formal in the sense that the convergence of the vari ous series is not discussed. Moreover, the series diverge in many applications. However, the first orders of the transformed system can give interesting in formation and the process can be stopped at a certain order L. This means that these terms of the series are useful to construct both the transformed Hamiltonian and the generating function, since they are unaffected by the
288
divergent character of the whole process. In those circumstances, one can use the General Perturbation Theorem. T h e o r e m 2.1 (Meyer) Let L > 1 be given, and let {Pi}$L0, {Qi)i=\ and {Ri}i=i be sequences of linear spaces of smooth functions defined on a common domain fi in R 2 n with the following properties: i) QiCVi!i
=
l,...,L;
ii) Hi£Vi,i
=
Q,l,...,L;
Hi) {Vi,Kj}CPi+j,i
+j =
l,...,L;
iv) for any D 6 Vi, i = 1 , . . . , L, one can find B € Q, and C € 7£j such that B =D+
{H0,C}.
Then, there exists an analytic function, W,
W(z;£) = £ - W i + 1 ( s ) , i=o l-
with Wi e TZi, i — 1 , . . . , L, such that the change of variables x = X{y\ e) is the general solution of the initial value problem dx
„dW,
.
x(0) = y, and transforms the convergent Hamiltonian (2) to the convergent Hamiltonian L
IC{y;e) --
ei
ICi(v) +
0{eL+1),
i=0 *'
with K.i 6 Qi, i = 1 , . . . , L. The corresponding version for non-autonomous Hamiltonians can be found in the paper by Deprit 4 and the book by Meyer and Hall 15 . There are some generalizations for differential equations systems and for tensor fields (see the paper by Kamel 10 for the adaptation of Deprit's method to differ ential equations and the paper by Meyer14, where Theorem 2.1 appears in the context of tensor fields and includes as particular cases those dealing with Hamiltonian systems and differential equations).
289
2.2
Normal forms
As our purpose is to extend the theory of normal forms, in the following we make a brief summary of the basic results for analytic Hamiltonians in this respect. For more details on the topic we refer the reader to the book by Sanders and Verhulst 20 . The book by Wiggins21 contains the treatment of normal forms for vector fields as well as some illustrative examples. We consider a Hamiltonian H(x;e) which admits an expansion in powers of the small parameter e as in (2). Furthermore, we restrict ourselves to the semisimple case, that is, the Lie operator £«„ is such that for two linear spaces of smooth functions (or algebras of functions) V and 1Z, any function F in V may be decomposed into a sum F = F\ + F^ where Cn0 (•fi) = 0> F2 — CH0{G) and G € 72.. The general situation (semisimple plus nilpotent case) has not been considered in the literature up to our knowledge except for polynomial-like Hamiltonians as we shall see in Sec. 4. Hamiltonian 7i(x;e) is said to be normal if the Poisson bracket { H, Ho } = 0. If the system is not normal, a symplectic change of variables x = X(y;s) is said to normalize it if the transformed Hamiltonian fC(y;e) is normal, that is, it satisfies { K., HQ } = 0 . Then, K is called the normal form of H, or the normalized Hamiltonian. In terms of Lie operators, K, is the normal form of 7i when £ « 0 (JC) = 0. If V denotes an algebra of functions containing the terms Tii then, the normalization produces an effect of decomposition of the algebra V as a direct sum of two subspaces V =
ker(CHo\V)®im(Cno\V),
where ker (£ W o ) is the kernel of £ « 0 and im (£n0) i s t n e image of Cn0 ■ Clearly, both ker (Cn0\'P) anc ^ i m (£w 0 l^) n a v e structures of algebra of functions and they are, more specifically, subalgebras of V. Note that, in terms of Theo rem 2.1, ker(£ W o |P) is Q. Therefore, normalizing H means finding a Lie transformation that projects H onto an element K, in ker(£« 0 |'P). Thus, a normalization can be also understood as a simplification. Moreover, it is a geometrical opera tion, in the sense that its purpose is extending a symmetry of the Hamiltonian of zeroth order to the whole perturbation. Thus, the normalization reduces the number of degrees of freedom by one, because it makes an integral appear. In this sense, the normalization extends the purpose of the averaging method to higher orders. In practice, the normalization consists in calculating at each order of per turbation the terms /C» of the Hamiltonian /C, together with their correspond-
290
ing generating functions, VV<, following the rules of the Lie transformation described in the previous subsection. Let us note that the generating function coming from the transformation X(y;e) belongs to an algebra Q of functions, not necessarily V, but such that for any F £ V and G £ 71 the property { F , G} £ V is satisfied. Then, Hamiltonian 7i is decomposed as a direct sum of its normal form K. and a term which is the image of the generating function. The reason is that the Lie operator C-H0 associated with the normalization and restricted to the algebra V is semisimple. As we have noted before, a general treatment of the computation of normal forms for systems with any type of Lie operator associated to the unperturbed part of the Hamiltonian is an open question. These are the basic concepts we need to make an extension of the theory of normal forms. We have also described roughly in this section, the basic tools and methods which are used to simplify perturbed Hamiltonian systems. 3
Formal Integrals
The central idea of this section is to simplify a Hamiltonian system of the form (2) so that the transformed Hamiltonian K be a system with one degree of freedom less than Ti. Given an analytic Hamiltonian of the type (2), with W* £ Vi for i = 0 , 1 , . . . ,L for some linear spaces of functions, Vi, and a function G(x) 6 Vj, for some j > 0, we want to build a symplectic change of variables x = X{y; e) through a Lie transformation, such that the transformed Hamiltonian L
K{y\e)
Ei
ICi(y) + 0(eL+x)
satisfies the condition { Ki, Q } = 0, for each i — 1 , . . . , L. Let "H and Q be given. Essentially one might ask about the way of con structing K. such that { fCi, Q } = 0, i = 1 , . . . , L. If in addition to that we get that the Poisson bracket { Wo, G } also vanishes, we shall obtain an integral of the truncated system K. independent of it, and therefore, the number of degrees of freedom of K. would be one less than the one of 7i. The construction of K. must be done order by order, i.e., one has to proceed in an ascendent way from i = 1 to i = L. For that, the partial differential identity, that is, the homology equation (4) has to be solved. Note that the terms Hi are known and the solution of (4) is the pair (Wi, tCi) with the extra condition { £<, Q } = 0 for each i = 1 , . . . , L. In order to calculate this solution we have to split "Hi = H* + H? , where
291 H* e ker (Cg) and H* = Hi - H*, for each i - 1 , . . . , L. In this way, we choose Ki — H' and Wj as a solution of £«o(W0 = H*.
(6)
Proceeding with the scheme stated before, two problems arise. The first concerns the way of splitting Tii conveniently, i.e., the question is if it is always possible to determine ker(£g). The second question is related to the solvability of Eq. (6). In other words, if Tif is given, in which situations can we find Wj? In the following we adapt Theorem 2.1 to the general situation of con structing formal integrals and this is the basis of the theory we develop in this paper. Theorem 3.1 Let L > 1 be given, let {Pi}f=0, {Qi},L=i and {ftjf = 1 be sequences of linear spaces of smooth functions defined on a common domain Q in R 2 n and let Q be a function in Vj, for some j > 0, with the following properties: i) QiQVi,i ii) HieVi,i
=
l,...,L;
=
0,l,...,L;
in) {Pi,Kj}CT>i+j,i
+j =
l,...,L;
iv) for any D G Vi, i = l,...,L,
one can find B G Qi and C QlZi such that
B = D + {H0,C}
and
{Q,B}=0.
Then, there exists an analytic function W, L 1 i ~e W(sc;;e) == 1^ 7? W i + i ( x ) ,
»=0
with Wi € TZi, i — 1 , . . . ,L, such that the change of variables x = X(y;e) the general solution of dx dW, ~de = J -dx~^ x(0) = y,
is
■ e),
and transforms the convergent Hamiltonian (2) to the convergent Hamiltonian
1
£(V,e)
z
-t t=0
Ki(v)
+ 0(eL+1),
292
with fCi 6 Qi and { Ki, Q } = 0, i = 1 , . . . , L. Besides, if { Ho , Q } = 0, then Q is an integral (formal integral) of K. Proof Note that the difference between this result and Theorem 2.1 is that here we add the function Q. Then, condition iv) of Theorem 2.1 is slightly modified in the sense that we also require that functions B € Qi satisfy {Q, B} = 0. According to Theorem 2.1, Ki £ Qi, then the additional re quirement { Ki, Q } = 0 is readily satisfied. (See also the reference by the authors 19 ). □ A crucial remark is that in the cases where Q is an integral of Ho, the effect of constructing K, with Ki € kei (Cg) for i = 1 , . . . ,L, is to extend (formally) the integral of the unperturbed system to the whole transformed Hamiltonian K. It means that the choice of Q can be done adequately if one knows previously the integrals of Ho- In this way, the number of degrees of freedom is reduced by one after the transformation. An outstanding feature is that the construction of formal integrals al lows one to obtain periodic orbits. Indeed, as Q becomes an integral (formal integral) of K, the fact of transforming Hamiltonian H into K implies the in troduction of a symmetry group. The existence of this symmetry group gives rise to "generalized" relative equilibria on K. The analysis related to the sta bility and bifurcations of these equilibria leads to determining the qualitative behaviour of the corresponding periodic orbits on the Hamiltonian H, as it was demonstrated by Moser 17 . We can advance that the type of the Lie operators C-n0 a n d Cg determines each reduction process. In some occasions, the partial differential identity £wo(Wi) = H* cannot be solved exactly. Thus, one could give an asymptotic approximation of Wi in terms of a recursive algorithm. This is the Relegation algorithm and has been applied in some problems in the context of Keplerian systems. For more details the reader is referred to the work by Deprit et a]. 7 . 4 4-1
Polynomial Hamiltonians The theory
In the following we shall concentrate on the application of the calculation of formal integrals to polynomial Hamiltonians of the type (2), x being a vector in R 2 n . The unperturbed part, Ho, is a quadratic polynomial in x, whereas for each i > 0, Hamiltonians Hi stand for homogeneous polynomials of degree i + 2 in x. We start by recalling the Normal Form Theorem for the general equilib rium, which is the basic result in this field.
293 Theorem 4.1 (Meyer) Let A = J B, with B a symmetric matrix of dimen sion 2n. Then, there exists a symplectic change of variables, x = X(y;e) — y + ..., which transforms Hamiltonian (2) to L
£i
/C(y; e) ~ -- 2.*, 7J Ki(y) +
0{eL+1),
(7)
i=0
through a generating function W. In particular, K-a(y) = Ho(x). Besides, K-i is a homogeneous polynomial of degree i + 2 such that, if /CQ(J/) = \ yl Ry, where R = J B J', then {iCi, /Co ) — 0,
(8)
for alii = 1 , . . . ,L. In addition to that, W is defined as L-1 e* W(x; e)- = 2_^ 77w i + i(s).
i=o *•
■with Wi a homogeneous polynomial of degree i + 2. The above theorem says that the terms of the normal form, K, are invari ant under the flow denned by exp(s A1). In that respect one can say that the normal form is simpler than the original Hamiltonian. Indeed, Hamiltonian K}0 is an integral of eK-i + e2 IC2/2 + ... + eL K.i/L\, but not necessarily an integral of K. This implies that in the polynomial case the concept of nor malization has been generalized with respect to the one stated in Subsec. 2.2, since now {K, Ho} is not necessarily zero. Let us clarify now when the number of degrees of freedom is reduced with the calculation of the normal form. Theorem 4.2 Let H(x; e) be defined as in (2), a polynomial Hamiltonian of n degrees of freedom, with quadratic part Ho = \ xl B x, where B corresponds to a symmetric In x 2n-matrix. Let A = JB be the matrix associated to the linear system of differential equations defined by Ho- Let x = X{y\e) = y + ... be the symplectic change of variables which transforms Ji into its normal form, the convergent Hamiltonian K. Let A = S + N represent the Jordan decomposition of A, where S corresponds to the semisimple part and N to the nilpotent one. Then, the quadratic polynomial Ts(y) = %s{x) = — ^xlJSx is an integral of Tio, provided that 5 ^ 0 . Moreover, by means of the normal form transformation, Is(y) becomes an integral (formal integral) of K. independent of it and the number of degrees of freedom of K, is n — 1.
294 For the proof of this theorem, see the work by the authors 19 . Now, let us focus on the semisimple case, that is, when A = S. Then, A is diagonalizable and the above result implies that {AC*,ACo} = 0 for i = 0,l,...,£«. The reason is that K,Q = K,Q. It is the classical result of normal form for a Hamiltonian near an equilibrium point with a simple linear part. One should notice that in this particular case, /Co is a formal integral of AC (or an actual integral if we truncate K,). It means that the effect of normalizing Ti is to build a new integral independent of K.. This feature allows one to reduce the number of degrees of freedom of the original system by one. Consider now the situation where A = N. Then, { £<, ACo } ^ 0, i = l,...,L. The reason is that, as S = 0, the function J s ( y ) = — \yi J Sy = Q and we cannot have a new integral of the reduced system. Then, although the normal form AC is a reduction of H, it is not so drastic, in the sense that one does not introduce a new integral for the transformed system. Thus, the number of degrees of freedom of AC coincides with the number of degrees of freedom of H. Finally, we treat the general case, that is, A = S + N with S ^ 0 ^ N. One should notice that the semisimple case is a particular case of this. In both situations the function 2$(y) becomes an integral of K. and the number of degrees freedom of the initial Hamiltonian is reduced by one. Once we have recalled what is known for polynomial systems, let us adapt the approach described in last section to the polynomial context, in order to go one step further. On the one hand, we intend to provide a result which can be used to enlarge the thesis of the Normal Form Theorem. It means that we are interested in applying it to calculate formal integrals of Hamiltonians whose unperturbed parts are related to nilpotent matrices A. On the other hand, this approach will be used to calculate different types of reduced Hamiltonians or to make successive reductions so as to lower the number of degrees of freedom by two or more units. In order to adapt Theorem 3.1 to the polynomial context, we have to identify the sets Vi with the linear spaces of homogeneous polynomials in x of degree i + 2, take Qi as some subsets of Vi and 7£j as the linear spaces of rational functions, where the subtractions of the degrees of the numerators by the degrees of the denominators are i + 2. Then, W is a rational function and the change of variables x = X(y; e) transforms the convergent Hamilto nian (2) (of polynomial type) into the Hamiltonian AC, such that each Ki is a polynomial in y of degree i + 2 with { AC», Q } — 0, for i = 1 , . . . , L. Besides, if { Wo , Q } = 0, then Q is a formal integral of KL. We have to emphasize the rational character of the function W. The reason is that the solution of Eq. (6) can be either a polynomial or a rational
295 function of the variable x. In this respect, properties ii) and iii) of Theorem 3.1 are strong hypotheses. However, when they are satisfied, one can assure that the Poisson brackets of the Hamiltonians H™ with the generators W n , for some positive integers I, m and n, are always polynomials of the required degrees. If this is not the case, one could obtain for some i, a rational term fd. Thus, we shall stop the process of the Lie transformation at that order, so as not to introduce logarithmic terms. A particular situation appears when the solutions of all partial differential equations (6) are always polynomials. In this case, the sets Hi would be subsets of Vi and we would never go out of polynomial domains. For more details about this subject the reader is referred to the work by the authors 19 . 4-2
Application: the semisimple 1-1 resonance
Let us take an example in order to illustrate the theory explained before: W=
TIQ
+ £ H\.
As the unperturbed part, we have chosen the Hamiltonian corresponding to two harmonic oscillators with the same frequencies, i.e., 2
2 Ho == i ( X + Y
) + £(x2! + 2 / 2 ) ,
where u represents the frequency of the oscillators with physical dimensions 1/time, (x, y) stands for the position of the particle in Cartesians and (X, Y) are the corresponding momenta. We have added an arbitrary cubic polynomial perturbation, which has twenty terms. Hamiltonian Tio is semisimple, thus the classical normalization allows us to reduce by one the number of degrees of freedom of system H, but as we have already shown, this is not the unique way of proceeding. Studies of the 1-1 resonance via the calculation of the normal form can be found in the papers by Kummer 11 or Miller16, to name a few. First of all, we know that Tio has three functionally independent integrals. We choose the following: Ii=H0,
J2 = i ( w 2 x 2 + X 2 ) ,
I3 = xY-yX,
(9)
which correspond, respectively, to the total energy of the unperturbed system, to the energy of the oscillator along the Ox axis and to the modulus of the angular momentum vector. Other integrals associated to system Wo are J\ = j (w 2 y 2 + y 2 ) and J2 = UJ2 xy + XY. It is not difficult to check that the relations among the latter and the former ones are Jy = I1—12 and 4 J2 Ji = a; 2 / 2 4- J 2 . The I's and the J's correspond to the invariants of HQ. The
296
version of the invariants for the isotropic oscillator in n degrees of freedom, with n > 2, can be looked up in the thesis by Yanguas 23 . Then, we take as the function Q appearing in Theorem 3.1, these three integrals. Taking I\ corresponds to the classical normalization. The first order in the Lie process vanishes and it is from the second order that we get information about the relative equilibria. We do not study this here, but a good reference is the book by Sanders and Verhulst 20 . Since I\ becomes a formal integral of K,, then the identity 4 / 2 (C - ^2) = w2 7 2 + J 2 holds, C being the constant I\. It implies that, in the reference frame OI2I3J2, the reduced phase space is a sphere. By taking Q = I2 we obtain a non-vanishing first order in the Lie process. Since with the usual approach we have to go to second order to retain the resonant terms in the normal form, now we clearly obtain a different reduced Hamiltonian. In this occasion, after making C = I2 one obtains the identity AC Ji = w2 Jf + Jf, which corresponds to a paraboloid of revolution in the frame defined by OI3 J\J2In the third case, with Q = I3 we get not only a first order vanishing, but also the second order. Moreover, the generator at this step is not polynomial and logarithmic functions appear. Thus, the process should be stopped there and we obtain no information in this way. Now, since C = w2 /f, then identity 4J2 J i = C + Jf holds. That is, the reduced phase space defined by I2, J\ and J2 is a hyperboloid of two sheets when C > 0 and it is an elliptic cone when C = 0. According to the last paragraphs, we have obtained two different (useful) reductions of the same problem. What remains is to analyse whether there is a connection between the two reductions or not. In principle, different reduced Hamiltonians could provide different information about the original problem. 5
Keplerian Systems
Perturbed Keplerian systems are present in many problems of Celestial Me chanics. The analysis of such systems is usually carried out by means of simplifications which convert the original Hamiltonian into an equivalent one, but easier to be studied. The simplifications performed in most of the theo ries are based on normal forms. As the Hamiltonians representing this type of problems are not polynomial, a Taylor expansion of them must be done so as to apply the Normal Form Theorem. The reader is referred to the book by Meyer and Hall 15 for some examples. However, other techniques can be used to avoid the problems originated mainly by the lack of convergence of some series expansions or by the large
297 amount of terms in the calculations. Delaunay can be considered as a pio neer of this. Indeed, he introduced a set of action-angle variables specifically designed to deal with the Lunar main problem. The ideas of Delaunay have been adapted by Deprit and his co-workers to a more modern context of Lie transformations, and they have proved to be rather useful in the treatment of some problems, like the Lunar motion, n-body problems or artificial satellite theories. A usual way of modeling Keplerian systems is by means of the twobody problem Hamiltonian. Let x = (x, y, z) and X = (X, Y, Z) represent the position and velocity of a particle attracted by an inhomogeneous sphere (primary) in a reference frame whose origin is placed at the centre of mass of the primary, and such that one of its axes is in the direction of the rotation of the primary. Then, the Hamiltonian associated to the motion of that particle is given by
no =
\\\Xf-^-.(XY-yX),
where n is the gravity constant (after normalizing all the masses to 1), and u) represents the frequency of rotation of the primary. Now, one should add to Ho a perturbation, i.e., an analytic Hamiltonian eH\. This perturbation normally comes from a development of Legendre polynomials in powers of the parallax quotient {ct/r), where a measures the mean equatorial radius of the primary and r 2 = x2 + y2 + z2. However, this is not always the case. Indeed, there are cases where the perturbation is a polynomial in Cartesian variables. This is the situation, for instance, of some models of dust particles, or, in the context of Physics, van der Waals potentials or Zeeman and Stark effects. The book by Gutzwiller8 contains many examples of perturbed Keplerian systems appearing in Celestial and Classical Mechanics. The most usual way of dealing with Keplerian systems is by using Delau nay variables. They are formed by three angles: i) (., the mean anomaly; ii) g, the argument of periaxis and iii) h, the argument of the node. We have to add the three momenta associated to the coordinates, that is, i) L, which is related to HQ by means of the identity _Jf2_
ii?
=
1||Y|,2
M
2""
iixir
ii) G, which stands for the modulus of the angular momentum vector and iii) H, which is the third component of this vector. Now, a Delaunay transformation 5,6 (symplectic change of variables) permits to express a per-
298
turbed Kepler system H = Ho + e Hi in Delaunay variables. In this way, H0 =
-CJH.
2L 2
Therefore, the Lie operator associated to Ho is ■^Wo
= n
d
W
-
d
ijj—.
where n refers to the mean motion of the particle and is related to L by the identity n — /i 2 /L 3 . One should note the semisimple character of £ « 0 . Unfortunately, it does not mean that the manner in which the simplifications are carried out is as easy as in the case of polynomials, as we shall see in the next paragraphs. Normally, the perturbation can be expressed as a Fourier series in the angles (, g and h. However, in many occasions the initial Hamiltonian is not given in terms of the Delaunay angles. Actually, it can appear as a function containing positive or negative powers of r combined with trigonometric sums in various angles: the argument of perigee g, the argument of the node h, the true anomaly / or the eccentric anomaly E (these two latter variables being functions of £, L and G, see references5,6 for details). Thus, if one tries to express r, / and E explicitly in terms of the mean anomaly, then transcendental functions are introduced. Therefore, the usual way of approaching these problems is by doing a double expansion of the perturbation: first, a Taylor series in powers of the orbital eccentricity, e = y^l -G2/L2, and second, a Fourier development in I must be performed so that it appears explicitly in the equations. Besides, the coefficients of the trigonometric terms involve algebraic functions (rational and square roots) of the momenta L, G and H. Nevertheless, the procedure already mentioned presents some serious ob jections. First of all, if the particle under study revolves around the primary with a large eccentricity, say, 0.3, the convergence of the Taylor expansions is very poor. Besides, the number of terms needed to compute the normal form grows very fast and it becomes too big to make realistic computations. An alternative to bypass the previous drawbacks consists in avoiding the Taylor and Fourier expansions of the equations and make exact normalizations valid for any value of the eccentricity (for perturbed elliptic trajectories). According to Sec. 3, several reductions can be performed in order to simplify Hamilton systems with unperturbed part Ho- Indeed, what one needs is to pick Q as one of the integrals of Ho and apply the procedures exposed in Sec. 3 to extend that integral to the truncated transformed system K. Since the two-body problem is maximally superintegrable, five independent integrals
299
can be found. In terms of Delaunay elements we have: g, h, L, G and H. Among these integrals, a "clever" choice would be to take any of the momenta. The reason is that, if one tries to take Q as either the angle g or h, very difficult quadratures in terms of L, G and H should be performed. Thus, either L, G or H can become an integral of K.. At this stage one should notice that forcing L to be an integral of K is equivalent to "eliminate £ from the perturbation", and similarly with G and g and with H and h. This is what astronomers call "elimination of variables", which is, in essence, an average over the specific angle. However, if the Taylor and Fourier expansions are not used, the problem of solving exactly the homology equation (4) arises. For instance, take the case where Q = G. Then, suppose that Hi, for i > 1, has been already calculated. At this step, the homology equation that must be solved corresponds to the partial differential equation n
~df-U}-dh+fCi
=
nu
where the unknowns are the functions AC, and Wj.^ Following^ the method exposed in Sec. 3, terms Hi are split as the sum H*+Hf, where H* € ker (CG) and Hf — Hi—H*. Hence, K,i is chosen as H* and Wi is taken as the solution of
dWi dWi -df-UJ-dh=Hi-
n
~#
(10)
The problem is that, in general, Eq. (10) cannot be solved in closed form in terms of known functions (see18 for details on how to solve the correspond ing homology equation), due to the form of the right member of Eq. (10). The reader should have in mind that H? is not a Fourier series in I and that it depends on the mean anomaly through either the eccentric or the true anomaly. A way of circumventing this trouble is solving the equation in pow ers of the quotients n/w or tu/n. If any of the two ratios is small, Eq. (10) can be approximated with few terms. This occurs for slow-rotating orbits (| - n2/(2 L2)\ > | - u H|) or fast-rotating orbits (| - /z 2 /(21?)\ < | - w H\). This method yields better convergence than the one provided by the classical expansions in e, the eccentricity of the orbit, and £. In fact, this is the relega tion method 7 applied to perturbed Keplerian systems. For moderate rotating Keplerian systems | — fi2/(2 L 2 ) | « | — w H\ the corresponding homology equa tion can still be solved in terms of a generalization of the incomplete gamma function, as it is shown in 18 . However, the previous can be simplified in some cases. For instance, if the perturbation is axially symmetric, that is, H is an integral of H, then
300
h is an ignorable coordinate and the Lie operator can be written simply as £ H 0 = nd/dL In these circumstances, Eq. (10) can be solved in closed form. An analysis of the three possible choices of Q is an open question and could provide alternatives to the classical studies. 6
Conclusions
Theorem 3.1 is the main result of this work. It represents a generalization of the theory of calculation of normal forms to simplify perturbed Hamiltonians. This result allows one to reduce the number of degrees of freedom of these sys tems in all the situations, at least up to a certain order of approximation. The reason is that it gives a method of calculating formal integrals for perturbed Hamiltonians like (2). In addition to that, we show that the classical normal form is not the only possibility for reducing such systems. Given H, we can construct as many re duced Hamiltonians K with one degree of freedom less than H, as integrals Q of the unperturbed part Wo we know. In this way, the corresponding reduced phase spaces will be different and it will allow to study problems from differ ent points of view. For instance, the equilibria in each reduced space would represent different periodic orbits (in principle). Finally, Theorem 3.1 allows one to make more than one reduction, which can be interesting when dealing with systems of more than two degrees of freedom. Acknowledgments Research has been partially supported by CICYT P B 95-0795 (Spain), by a Project of Departamento de Education y Cultura, Gobierno de Navarra, Orden Foral 508/1997 (Spain), by a grant for J.P. from Universidad Publica de Navarra and by a grant for P.Y. from Departamento de Educaci6n y Cultura, Gobierno de Navarra, Orden Foral 143/1998 (Spain). J.P. thanks Prof. L. Peletier (Dept. of Mathematics) and P.Y. thanks Prof. T. de Zeeuw (Dept. of Astronomy), both from Leiden University (Holland), for the facilities they provided during the stay of the authors there. References 1. A. Celletti and A. Giorgilli. On the Stability of the Lagrangian Points in the Spatial Restricted Problem of Three Bodies, Celestial Mechanics & Dynamical Astronomy 50, 31-58 (1991).
301
2. R. Cushman. A Survey of Normalization Techniques Applied to Per turbed Keplerian Systems, in Expositions in Dynamical Systems, eds. C. K. R. T. Jones, U. Kirchgraber and H. O. Walther, Dynamics re ported: new series 1, Springer (1992). 3. R. Cushman, A. Deprit and R. Mosak. Normal Form and Representation Theory, Journal of Mathematical Physics 24, 2102-2117 (1983). 4. A. Deprit. Canonical Transformations Depending on a Small Parameter, Celestial Mechanics 1, 12-30 (1969). 5. A. Deprit. The Elimination of the Parallax in Satellite Theory, Celestial Mechanics 24, 111-153 (1981). 6. A. Deprit. Delaunay Normalisations, Celestial Mechanics 26, 9-21 (1982). 7. A. Deprit, J. Palacian, E. Deprit and J. F. San Juan. The Relegation Algorithm, submitted for publication in Celestial Mechanics & Dynamicai Astronomy (2000). 8. M. C. Gutzwiller. Chaos in Classical and Quantum Mechanics. Interdis ciplinary Applied Mathematics 1, Springer-Verlag, New York (1990). 9. M. Henon and C. Heiles. The Applicability of the Third Integral of Motion: Some Numerical Experiments, The Astronomical Journal 69, 73-79 (1964). 10. A. A. Kamel. Expansion Formulae in Canonical Transformations De pending on a Small Parameter, Celestial Mechanics 1, 190-199 (1969). 11. M. Kummer. On Resonant Classical Hamiltonians with Two Equal Fre quencies. Communications in Mathematical Physics 58, 85-112 (1978). 12. K. R. Meyer. Normal Forms for the General Equilibrium, Funkcialaj Ekvacioj 27,. 261-271 (1984). 13. K. R. Meyer. Scaling Hamiltonian Systems, SIAM Journal of Mathemat ical Analysis 15, 877-889 (1984). 14. K. R. Meyer. A Lie Transform Tutorial II, in Computer Aided Proofs in Analysis, eds. K. R. Meyer and D. S. Schmidt, The IMA Volumes in Mathematics and its Applications 28, Springer-Verlag, New York, 190210 (1991). 15. K. R. Meyer and G. R. Hall. Introduction to Hamiltonian Dynamical Systems and the N-Body Problem, Applied Mathematical Sciences 90, Springer-Verlag, New York (1992). 16. B. R. Miller. The Lissajous Transformation III. Parametric Bifurcations, Celestial Mechanics & Dynamical Astronomy 51, 251-270 (1991). 17. J. Moser. Regularization of Kepler's Problem and the Averaging Method on a Manifold, Communications on Pure and Applied Mathematics 23, 609-636 (1970).
302
18. J. Palacian. Normal Forms and Reduced Phase Spaces for Threedimensional Perturbed Keplerian Systems, submitted for publication in Nonlinearity (1999). 19. J. Palacian and P. Yanguas. Reduction of Polynomial Hamiltonians by the Construction of Formal Integrals, submitted for publication in Nonlinearity (1999). 20. J. A. Sanders and F. Verhulst. Averaging Methods in Nonlinear Dynam ical Systems, Applied Mathematical Sciences 59, Springer-Verlag, New York (1985). 21. S. Wiggins. Introduction to Applied Nonlinear Dynamical Systems and Chaos, Texts in Applied Mathematics 2, Springer-Verlag, New York (1990). 22. S. Wolfram. The Mathematica Book, 3rd ed, Wolfram Media/Cambridge University Press, Cambridge (1996). 23. P. Yanguas. Integrability, Normalization and Symmetries of Hamiltonian Systems in 1-1-1 Resonance, Ph.D. Thesis, Universidad Piiblica de Navarra, Pamplona (1998).
LINEAR STABILITY IN THE 1 + iV-GON RELATIVE EQUILIBRIUM GARETH E. ROBERTS Applied Mathematics Department, Box 526, University of Colorado at Boulder, Boulder, CO 80309-0526. E-mail: [email protected] We study the linear stability of the relative equilibrium in the n-body problem consisting of n equal masses at the vertices of a regular n-gon with an additional body of mass m at the center. This configuration is shown to be linearly unstable when n < 6. For n > 7, a value hn is found such that the configuration is linearly stable if and only if m > hn. This value is shown to increase proportionately to n3.
1
Introduction
A relative equilibrium is a special solution of the n-body problem which rotates rigidly about its center of mass if given the correct initial momentum. In rotating coordinates these special solutions become fixed points, hence the name relative equilibria. For example, when n = 3 the only noncollinear three-body relative equilibrium consists of each body, regardless of its mass, located at the vertex of an equilateral triangle. This configuration is linearly stable only when one of the masses dominates the other two. 2 ' 10 This situation occurs in our galaxy with Jupiter, the sun, and the Trojan asteroids, where the sun is the dominant mass. Moeckel has conjectured that a relative equilibrium can be linearly stable only if it contains a mass significantly larger than the others. 8 This is the case for the equilateral triangle. It is easy to see that the regular n-gon with a central mass is a relative equilibrium for any value m of the central mass (including m = 0) and thus it is natural to treat m as a parameter. Maxwell carried out such an analysis in his study of Saturn's rings and concluded that for sufficiently large values of m, the ring is linearly stable. 3 ' 4 Moeckel provides a thorough analysis of the stability problem in general, making use of the special properties of the linearized matrix to obtain a factorization of the characteristic polynomial. 7 He then proceeds to use his techniques to analyze some specific, highly symmetric relative equilibria including the regular n-gon with a central mass, hereafter referred to as the 1 + n-gon. (As a correction to Maxwell's study, Moeckel reveals that the 1 + n-gon is linearly stable for sufficiently large m only when n > 7.) We carry the analysis a step further. We investigate the cases n — 3,4,5,6 to determine if they are linearly stable for any value of m, and for n > 7, we
•jm
304
ask how large the central mass m must be relative to the mass of the bodies in the n-gon in order for the configuration to be linearly stable. It turns out that for the cases n = 3,4,5,6, the 1 + n-gon is not linearly stable for any values of m. For n > 7, the configuration is linearly stable if and only if m is greater than a certain value hn, and as n increases, hn increases like rn3, where r « 0.435. In other words, the 1 +n-gon relative equilibrium undergoes a bifurcation in stability at a value of m proportional to n 3 . For a simple application to our results, consider the nearly circular rota tion of the moon about the Earth. The Earth's mass is approximately 81 times that of the moon's. Suppose that the moon was suddenly split into n equal pieces, with each "new" moon landing close to a vertex of a regular n-gon with the Earth at its center. Then the Earth would be 81n times heavier than any one of its moons. The question then becomes is the Earth heavy enough to maintain this configuration? The answer is no, unless n G { 7 , 8 , . . . , 13}, as r n 3 grows faster than 81n. We would like to point out that similar results were obtained using differ ent methods by Elmabsout. 1 At the time this paper was being completed, we were unaware of his work. The results here are more precise and presented in more detail. 2 2.1
Determining Linear Stability of Relative Equilibria Relative equilibria in the n-body problem
We let the mass and position of the n bodies be given by m* and q t e R 2 , i = 1 , . . . ,n, respectively. Let r^ = ||qj — q j | | be the distance between the ith and j t h bodies and let q = ( q i , . . . , q n ) € K 2n . Using Newton's law of motion and the inverse square law for attraction due to gravity, the secondorder equation for the ith body is given by miq'i
Y ^ rniTrij^
-qj)
dU
where t/(q) is the Newtonian potential function:
=E
t/(q):
miTTij
i<j
Tij
We let the momentum of each body be Pi = m.i
— dp
305 „rr/ , p = W(q) =
dH - ^
(1)
where M is the diagonal mass matrix with diagonal mi,mi,m2,JTi2, • • • ,mn,mn and i / ( q , p ) is the Hamiltonian function:
#(q.p) = £
il|2 -^(q) 2?Tlj
= ipTM_1p-^(q)-
01 coswt sinwt . To introduce so that euJt = -10 - sin wt cos cot coordinates that uniformly rotate with period 2n/co, we let x< — euJtqi and yi = euJtpi. This is a symplectic change of variables preserving the Hamil tonian structure of system (l). 5 The new system becomes Next, let J
dH_ dy dH y = Vtf(x)+«;ffy (2) dx where K is a 2n x 2n block diagonal matrix with J on the diagonal and H(x, y) is the Hamiltonian function: wA'x + M - 1 y
H(x,y)
= i y T M - 1 y - U(x) -
coxTKy.
While system (1) has no equilibria, system (2) does. An equilibrium point in this new system will correspond to a periodic solution in the n-body problem consisting of a configuration of masses which rotates rigidly about its center of mass. Using the fact that KM = MK and K2 = —I, an equilibrium (x, y) of system (2) must satisfy y = —uMKx and W ( x ) + w 2 Mx = 0 A relative equilibrium of the n-body problem is a configuration x € R satisfies the algebraic equations in (3) for some value of co. 2.2
(3) 2n
which
Linear stability
Linearizing system (2) about a relative equilibrium (x, y) yields the matrix A =
'wK S
M~l uK
(4)
where 5 = DVU(x) is a 2n x In symmetric matrix. The characteristic poly nomial of A, P(\), is of degree An and is an even polynomial, since A is a
306
Hamiltonian matrix. 5 Suppose that v is an eigenvector of A with eigenvalue A, and write v = (vi, V2) with vi,V2 £ C 2 n . The eigenvector equation Av = Av then reduces to v 2 = M(XI — wK)vi Bvx
=0
where B = M-lS
+ (to2-X2)I
+ 2\u;K.
(5)
Therefore, to obtain the eigenvalues of A, one need only take the determinant of B and find the roots. In other words, P(A) = Det(5). We will call two vectors v and w M-orthogonal if v T M w = 0. Direct calculation shows that M~lS and K are symmetric and skew-symmetric, re spectively, with respect to an M-orthonormal basis. Writing B with respect to an M-orthonormal basis and taking the transpose will not change its de terminant. This gives an alternative argument for showing P(A) = P(—A). Throughout this work, P(A) will denote the characteristic polynomial for A: P(A) = det [M~1S + (u2 - A 2 )/ + 2\uK] .
(6)
Moeckel's idea is to obtain a factorization for P(A) by finding a subspace of R2™ which is invariant for both M~lS and K.7 Suppose that W is a subspace of R 2 n such that M~lSW = KW = W. Letting WL = {v e R 2 n : v T M w = 0 Vw G W) be the orthogonal complement of W with respect to M, direct calculation shows that M~1SW± = KWX = Wx. If Tj and T2 are M1 orthonormal basis for W and W- respectively, then the matrix B written with respect to the basis T = T\ U T2 is "Bi 0 ' 0 B2 where B\ and B2 are the restrictions of B to the subspaces W and Wx with respect to the basis T\ and T2, respectively. It follows that det(S) = det(.Bi) • det(S 2 ) or P(A) = Pi(A)P2(A). Moreover, since writing M _ 1 5 and K with respect to an M-orthonormal basis yields a symmetric and a skew-symmetric matrix respectively, the same argument used above to show P was even applies here to Pj and P 2 .
307 Proposition 2.1 Suppose that W C R is an invariant subspace for both M~1S and K. Then the stability polynomial can be factored into two even polynomials in A, P(X) = Pi(X)P2(X), each given by equation (6) with the matrices involved restricted to the subspaces W and W x , respectively. This proposition is not very useful if we can't "guess" invariant subspaces for both M~1S and K simultaneously. Moeckel successfully does this with the collinear, regular polygon, and 1 + n-gon relative equilibria. 7 Building on the work of Palmore 9 , he utilizes the special structure of the matrix S in these cases to find two and four-dimensional invariant subspaces. He finds enough subspaces to calculate all of the eigenvalues and then performs a thorough analysis to classify their stability. In the next section, we will describe these subspaces for the 1 + n-gon and obtain a factorization of P(X). In Section 4, we locate the precise value for which the 1 + n-gon becomes linearly stable. In Section 5, we provide the estimates on this bifurcation value to show it is asymptotic to m 3 . 3 3.1
The 1 + n-Gon Invariant subspaces and factoring P(X)
The 1 + n-gon is the relative equilibrium consisting of n equal masses at the vertices of a regular n-gon with an additional mass at the center. We set mjt = 1, for k € { 1 , . . . ,n} and let mo = m represent the mass of the body at the center. The position of the A;th body is given by x* = (cos 0*) sin 0*) where &k = 2ixk/n for k € { 1 , . . . ,n} and by xo = (0,0). It is easy to check that x = ( x o , x i , . . . , x n ) satisfies equation (3) for any value of the central mass m, so it is natural to treat m as a parameter. Recall that r^ represents the distance between the z-th and j - t h bodies. Two frequently used formulas are r„k = 2(1 - cos0fc)
and
rnk = 2sin(7rA;/n).
As mentioned in Section 2, this configuration begets a periodic solution in which the bodies rotate uniformly about the central mass with rotation speed u) = w(m), which in this case is given by 2
1
n—1 K-"*
LJ" = m+±an, CT„ = 2^ k=i
1 r
"fc
n—1 1 V~*
= i 2^ fc=i
7r
csc
. ";
—• n
This formula follows directly from any component of equation (3). Note that as the mass of the central body increases, the period of the circular orbit decreases.
308
Recall that S — DVt/(x). Direct computation reveals that Soo • ■ • Son
S=
:
:
_ "nO ' ' " &nn ,
where Sy is the 2 x 2 matrix given by mirrij
[7-axijxT.]
if
i^j
Sjj — - 2_^ Sij and Xjj = x, r ~ x< (I is the 2 x 2 identity matrix). Note that each block Sij is symmetric and that S^ — Sji. Using the fact that the diagonal blocks of 5 are the negative of the sum of the blocks in the corresponding rows, it is clear that both v = ( 1 , 0 , . . . , 1,0) and Kv = - ( 0 , 1 , . . . , 0 , 1 ) are in the kernel of S. Therefore, W = span{v, ATv} is a two-dimensional invariant subspace for both M-1S and K. Taking the matrices in equation (6) restricted to W yields the 2 x 2 matrix
[ 2wA
w2-\2
Taking the determinant of this matrix and applying Proposition 2.1 yields the quartic factor (A 2 +w 2 ) 2 and the repeated eigenvalues ±wi, ±wi. These values are a result of a drift in the center of mass and are evident in any relative equilibrium. Another two-dimensional invariant subspace which is also evident in any relative equilibrium comes from the configuration itself. Letting WQ = span{x, Kx} and applying Proposition 2.1 yields the 2 x 2 matrix r3w2-A2-2wA 2wA -A2
j
Here we make use of the fact that x and Kx are eigenvectors of M~1S with eigenvalues 2u>2 and — u>2, respectively.7 Taking the determinant of the ma trix above yields the quartic factor A2(A2 + w 2 ) with eigenvalues 0,0, ±ui. These values arise because the equilibrium point (x, —wMKx) of system (2) is not isolated and has two degenerate directions corresponding to rotation or scaling.
309 Since the factorizations we will be considering lead to even polynomials (Proposition 2.1), we will make the substitution z = A2 and call Q(z) = det [M-*S + (u2 - z)I + 2-Jz'uK} the stability polynomial. It is standard to call a relative equilibrium nondegenerate if the remaining 2n — 2 roots of the stability polynomial are nonzero (or the remaining An —A eigenvalues are nonzero). We will call a relative equilib rium spectrally stable if the remaining 2n — 2 roots of the stability polynomial are real and negative, and linearly stable if in addition to being spectrally stable, the linearized matrix A in (4) for the remaining An — 4 eigenvalues is diagonalizable. Due to the rotational symmetry of the 1 + n-gon configuration, S has a particularly nice form. After some computation, one can check that the space Ui = span{u, Ku}, where u = (uo, u i , . . . , u„) and uk = e'^'x*;, {Oki = 2nkl/n), form a two-dimensional complex M~ ^-invariant subspace in C 2 n + 2 for each I € { 2 , 3 , . . . , [n/2]}. When 1 = 1, this perturbation does not leave the central mass fixed at the origin. (This point was overlooked by Maxwell.3) Instead of (0,0), the first component of u must beuo = (—n/(2ra), —in/(2m)). Note that choosing / = 0 yields the two real vectors {x, Kx) already accounted for above. For I > 2, the restriction of the operator M _ 1 S to the subspace Ui is Pi - 3Qi + 2m -Mi iRi Pi + 3Qi-m
(7)
where n-l
V^ Pl
= k=\ 2s
*• —
cos
^k c o s &kl r\
2^r3
nk
n-l
V
C0S
. Qi = fc=l 2^-
^
- C0S
n-l
&M
Kk
.* = £ sin 8k sin 6ki /t=i
When I = 1, we obtain the matrix P1+2m + n -i(Ri - n) i{Ri + n/2) Pi - m - n / 2
(8)
These calculations rely on the fact that a given row of 2 x 2 blocks of S can be expressed in terms of one block at the end of the row using a rotation matrix. The reader is encouraged to see Moeckel's work for the details. 7 Taking real and imaginary parts of u and Ku yields a four-dimensional real invariant subspace for M~lS and K except when n is even and / = n/2. In this exceptional case, the vectors u and KM are each real, so that we obtain a two-dimensional M~^-invariant subspace. We ignore this case for
310
the moment. Setting u = v + i w , yields the four real vectors {v,w, where
Kv,Kv/}
vk = cos0fcj(cos0fc,sin0fc) wfc = sin0ju(cos0fc,sin0fc) (Kv)k
= cos0 fc i(sin0 fc ,-cos0 fc )
(K-w)k =
sin6ki(sm9k,-cos0k)
are the components for k € { 1 , 2 , . . . , n } with (0,0) as the first two en tries when I ^ 1 and (—n/(2m),0) and (0, —n/(2m)) as the first two en tries for v and w, respectively, when 1 = 1. It is easy to check that these vectors are linearly independent. Letting W; = sp&n{v,Kw,Kv,w} for I G { 1 , 2 , . . . , [n/2]}, one can check that the invariant subspaces W, Wb, W\,... ,W[n/2] are all M-orthogonal. When n is odd, the union of these spaces has dimension 2 + 2 + 4 x (n - l)/2 = 2n + 2, and for n even, we obtain a dimension of2 + 2 + 4 x ( n - 2)/2 + 2 = 2n + 2. Thus we have completely decomposed R 2 n + 2 into M~XS- and AT-invariant subspaces from which we can determine all of the eigenvalues for the 1 + n-gon. Remark: 1. For a given /, the perturbation corresponding to the invariant subspace Wi pushes the fcth body in a direction given by the position of the k{l + l)th (mod n) body. For example, when I = 0, the perturbation stretches or shrinks the configuration, giving us one of the degenerate directions. As / increases, the perturbation involves more and more twisting requiring a larger and larger central mass for linear stability. We prove this fact in Section 4. 2. The fact that the perturbation for I = 1 does not leave the central mass fixed is a result of Proposition 2.1. Since the space W is M - 1 5-invariant, so too is its M-orthogonal complement. Thus, if v is a vector in a different invariant subspace, it must belong to the M-orthogonal complement of W. This means that both the sum of the odd entries of Mv and the even entries of Mv must vanish. If the first two entries for the case I = 1 were zero, then these sums would not vanish. Hence the need for the extra factor —n/(2m). We see from (7) and (8) that the restriction of Wj = span{v, A"w, Kv, w}
311 to the matrices in equation (6) is given by the 4 x 4 matrix ai+u2-X2 -Ri 2Xu 0
-Ri -2Aw 0 6j+u;2-A2 0 2Au> 0 &i+u> 2 -A 2 Ri -2Aw Rt ai + w2 - A2
(9)
for I > 2 where ai = Pi - 3Qi + 2m and 6j = Pj + 3Qj - m, and ai+w2-A2 -(fli-n) -2Aw 0 - ( / * ! + n / 2 ) 6i + u> 2 -A 2 0 2Au> 2Aw 0 6i + u>2 - A2 i?i + n/2 i?i - n a\ + u2 - A2 -2Aw 0
(10)
when I = 1 where a\ = Pj + 2m + n and bi = Pi — m — n / 2 . Taking the determinant of each matrix yields the quartic factor (z = A2) Gi(«) = (*2 + ai* + Pi)2 + *u2c2z where ~(m\ a (m) ' a ( m
mm)
Q
=
jm + crn-2P1-%ifl \ m + an-2Pl
=l iil^l
\ - l 3 ( P i - § + ^ ) m + nP 1 + ^ ( 4 P 1 + n + a n ) if I = 1 - | 3 ( P ( + 3Q[ + ^ ) m + ( p / + «^ ) 2 _ 9 Q 2 _ /ja. if | ^ l
( u (U)
, , W
_ j 2Rx - ^ if / = 1 ~ \ 2Rt if / ^ 1
Here we have used the fact that u2 = m+
-2Aw - -J-" b% + w2 2 - A2
h
(13)
Taking the determinant of this matrix yields the quadratic factor (z = A2) F%(z) = z2 + anz + 0~ where a j ( m ) and /?jj(m) are given by (11) and (12), respectively, with I — n/2.
312
3.2
Properties of Pi,Qi and Ri
Before we analyze the polynomials Gi and F « , we state some important facts and identities about the quantities Pi,Qi and J2j. (i) flj = ( Q i + i - Q i _ i ) / 2 (ii) Pl+1 -Pi + Ql+1 -Qi
= Ri + Ri+i
(iii) Pi and Qi are strictly increasing in I for 1 < I < [n/2]. (iv) Pj = Ri > 0, Qi = 0 and i?« = 0. (v) Pi,Qi and Ri are positive for 1 < I < [n/2] except for Qi and R%. (vi) Pj > Qi and Pi > Ri for 1 < I < [n/2]. The first two items follow from the formulas for the sum and difference of cosine and sine while the third is proved by Moeckel.7 The fourth item follows straight from the definitions of Pi,Qi and Ri. The fact that Pj and Qi are positive (except for Qi) follows directly from items three and four. When n is odd, we have Rn-i = \(Q*±i — Qn-a) = ^(Qr>-1 — <2->-3) by symmetry. Thus, identity (i) and the fact that Qi is strictly increasing for 1 < I < [n/2] implies that Ri is positive for 1 < I < [(n —1)/2]. Finally, the last item follows from the calculations: ^
(1 -cos0 f c )(l -cos6kl)
^
^
= 2w
Pi-Qi = 2L.
lr
k=\
_R
p 1
"*
4r 4r
fc=l
_ y - 1 - cosgfc cos flu - sm6k6ki ' r^ 2rjL
= _
nk
k=\
l-cos^fci
.
> °
"fc
y ^ 1 -cosg f c ( t _i) ^ 2rjL
>
~
'
nk
k=\
Some important summation formulas which can be derived using standard complex analysis and the formula for summing a finite geometric series are n-l
n-l
rnk
= 2 cot — , n
fc=i
2 ^ cos ekrnk fc=i
= cot
57T
37T
3TT
n
cot — n
and
E
l .
cos0 fc r nfc
1
1
= - c o t - - - c o t - +
7T
cot-.
313
These formulas in turn provide convenient expressions for Pi,Qi and Ri. For example, note that
_ ^
I
5
COS' e Ok - cos^ k
V^ ^
l + COS 0*
£-5 2 + (cosg t -l) _ 1 V" J - _ i V" fc=i <7n
2
1
n * . 7T
k=i
nK
*=i .
4 cot-—. 2n
.
(14)
Similar calculations reveal that IT
,,_.
^ = 4-n-icot---cot-,
_
(15)
g2 = ^ n - i c o t ^ ,
(16)
_
5
1
1
37T
37T
3
3
7T
,
it2 = a n - - cot cot — and 4 2n 4 2n _ 5 1 57T 3 37T 7 7T P3 = r " - 4 C O t ^ - 4 C O t ^ - 4 C O t 2 ^ " 3.3
.,_.
(17) .„„. (18)
Conditions on Gi and F» for linear stability
If wefixan n value, the coefficients of the polynomials Gi and F " vary with m. We are interested in what values of m make the 1 + n-gon linearly stable. Rather than calculate the roots specifically, it is easier to find necessary and sufficient conditions on the coefficients which yield real and negative roots. To simplify notation we will let 71 = 71 (m) = a2 — 4 $ . When n is even and I = n/2, 7; is the discriminant for F » . Note that the slope of ai(m) is always one, so that 71 (m) is a parabola opening upward. It is easy to show that F^ has real, negative and distinct roots if and only if: a » > 0,
/3n > 0
and 7» > 0.
Recall that Gt(z) = {z2 + atz + ft)2 + AtJtfz. If q = 0, then this polynomial would always have repeated roots. Fortunately, this is never the case. L e m m a 3.1 The coefficient ci is nonzero for 1 < I < [(n — l)/2]. Proof: For / ^ 1, cj = 2Ri which is positive by item (v) of the last section. (When n is even and / = n/2 we are only interested in the quadratic
314 .F" which does not involve c?».) When I = 1, we use the fact that Ri = P\ and equation (14) to obtain 1
C\
=On
7T
- - cot —- --n/2. 2 2n
We numerically verified that c\ is negative for 3 < n < 11 and positive for 12 < n < 14. We can show that c\ > 0 for n > 15 using some basic estimates. Using the symmetry of the n-gon, and the fact that c s c i > 1/x and cotx < 1/x, we have CXn > CSC
n
n
2n
h CSC
n
1
1- CSC
7n
n
> - ( ! + - + ■- + -)
>2(l + f) 7T
I
> ctib +n 2
\ £ '
as desired. A proof similar to that of Moeckel's7 shows that Gj has real, negative and distinct roots if and only if ai > 0,
ft^O,
7J
> 0
and
<5, > 0
where Si = 6i(m) is the discriminant of Gi divided by the positive term cfw4: 6i(m) = Ptiaf - 4/?,)2 + cfu2at(af
- 36/3,) - 27cfw4.
(19)
Note that <5{(m) is a fifth degree polynomial in m with leading coefficient identical to the slope of Pi(m). Recall that Gi(z) has repeated roots if and only if 5, = 0 and F%(z) has repeated roots if and only if 7, = 0. Suppose that one of these polynomials has a real, negative repeated root and thus the 1+n-gon has a pair of repeated pure imaginary eigenvalues. In order to have linear stability, we would need a basis of eigenvectors for the linearized matrix A in (4). However, this is never the case. Lemma 3.2 Suppose that F»(z) or Gi(z) has a real, negative repeated root ZQ corresponding to one of the nondegenerate invariant subspaces Wi described above. Let B' be the restriction ofM^S + (w2 - z)I + 2y/zwK to Wi. Then the geometric multiplicity of ZQ for B' is always one and consequently, the linearized matrix A in (4) is not diagonalizable.
315 Proof: Recall that an eigenvector v = (vi,V2) of A with eigenvalue Ao = \/ZQ satisfies the equations v 2 = M(XQI 5VJ
— LJK)V\
=0
with B given by (5) evaluated at A = Ao- If the geometric multiplicity of ZQ for B' is only one, then there is only a single vector vi in the kernel of B, and hence A will have only one eigenvector for the repeated eigenvalue A0 = ,/ZQ. For F»(z), the 2 x 2 matrix in (13) can never have nullity 2 since the off diagonal terms are always nonzero (u> ^ 0 and Ao / 0). For Gi(z), the 4 x 4 matrices in (9) and (10) will have nullity greater than one if and only if every 3 x 3 sub-matrix has zero determinant. For I j^ I, taking the determinant of the matrix in (9) with the first row and fourth column deleted gives -2A0WCJ(6J4-O;2-A^).
By Lemma 3.1, this quantity vanishes only if A2, = 6/ + u>2. However, A2, = ZQ < 0 by assumption and k+u/2
= Pi + ZQi + CT„/2 > 0
by item (v) of the last section. A similar argument works for I = 1. This completes the proof. The importance of this lemma is that the roots of the polynomials F% (z) and Gi(z) must be real, negative and distinct in order to have linear stability. In other words, if all the roots of these polynomials are real and negative, but the discriminant vanishes for one of them, then the 1 + n-gon is spectrally stable, but not linearly stable. We will call a polynomial stable, when its roots are real and negative, and call it linearly stable, when its roots are real, negative and distinct. 4
Finding the Bifurcation in Stability
For a fixed n, we will show that as the mass m of the central body increases, the polynomials Gi become linearly stable in succession. In other words, as m increases first G\ becomes linearly stable, then G2 becomes linearly stable, and so on, until m passes through a bifurcation value hn where G—1 or F^ become linearly stable depending upon the parity of n. Since we are interested in the values of the central mass m for which Gi and F» are linearly stable, it is important to know when the relevant coefficients aj,/?j,7i, and 61 are positive. Recall that aj,/?j,7i, and 61 are
316 functions of m of degree 1,1,2 and 5 respectively. We denote m Ql and mpl as the lone roots of on(m) and 0i{m) respectively. Therefore, we have f 2Pi -
m," 1
nPi + ^ ( 4 P i + n + g w ) . /„
mA = <
3(P1
.,
^ x
if
- f + ^)
(Pi + *f)2 - 9Q 2 - fl2 iff ^ 1 . 3(P, + 3Qj + *f)
(20)
To simplify the notation, we define the following coefficients: . _ f 2
36Q 2 + 4Rf - 8anPi
if f ^ 1.
2
Then, 7j(m) = m — 2Ajm 4- Bi and r,± = A, ± ^ A 2 - B,
(21)
are the two roots of the quadratic 7i(m). Since 6i(m) is a fifth degree polyno mial, it has at least one real root. We will let Aj denote the largest real root of 6i(m). The first logical question is whether the roots of 71 (m) are real or not. A short computation shows that 2 1
6OP12 + 40Pi(j n - 38Pm + 4(
_ J 2
2
~ ' ~ \ 64P, + 40P,crn - 4Pf + 4a +
if f = 1
+ 288Q,(P, + Qi) if I £ 1 (22) For I ^ 1, items (v) and (vi) from Section 3.2 show that A2 — Bi is positive. For the case f = 1, we can numerically verify that A 2 — Bi > 0 for 3 < n < 10. To show that A\ — B\ > 0 for n > 11, we have 72CT„Q,
A 2 - S i > 60Pj2 + 40Pi(
317
two real roots m = Vi±. Note that since the leading coefficient of 7i(m) is one, 7z(m) > 0 if and only if m < Tj_ or m > T/+. Figure 1 gives a rough sketch of the lines aj(m),/?j(m), the parabola ji(m) and their corresponding roots. Later in this section we explain the relationship between these roots. 7i(m)/
Pl(m)/
/
\
m.0,/
rnai
Q((m)
____ m
" ^ \
/ '
+
Figure 1. Graph of the coefficients Q((m),ft(m) and 71 (m).
n/2 + a n / 2 ) , is positive except for
Lemma 4.1 The slope o//3i(m), 3(Pi 3 < n < 6.
Proof: We can numerically verify the claim for 3 < n < 8. Using (14), it suffices to show that an > § + \ cot ^ . An argument similar to the one used in Lemma 3.1 yields 7T 0~n > CSC
27T h CSC
n >
n., 1 *(1+2 + n.7r + 1,
37T h CSC
n 1 1, 3 + 4)
n 1 7r > T: + T c o f c TT"
4n h CSC
n
for n > 9
n
2 4 2n as desired. This result is extremely significant in its own right since /3; and Si share the same leading coefficient. When 1 = 1, the discriminant <5i(m) is positive for sufficiently large m only when n > 7. Thus, for 3 < n < 6, in order to obtain stability we have to hope that in the intervals where 6\ (m) is positive, the other necessary coefficients are also positive. However, this is never the case.
318
Theorem 4.2 For n = 3,4,5,6, the 1 + n-gon is not spectrally stable for any value of the central mass m. Proof: The relevant constants mai}mpl,ri± and Aj for each value of n are given in Table 1. (Note that the fourth column gives the mass values for which the 1 + n-gon is degenerate, as /3j = 0 implies that zero is an eigenvalue. These values agree with those listed in Appendix B of Meyer and Schmidt's 6 paper on the configurations which bifurcate from the 1 4- n-gon.) We handle each case individually. For n = 3, it is necessary that m > mQi « .6334 in order for a i ( m ) to be positive. However, by Lemma 4.1 and from Table 1, we see that 6i(m) is negative for m > Ai « .2744. Thus, the polynomial G\ is never stable for n = 3. For n = 4, since Ti_ is negative, it is necessary that m > T i + « 2.1553 in order for 71 (m) to be positive. However, by Lemma 4.1 and from Table 1, we see that ^i(m) is negative for m > A i « 1.0072. Thus, the polynomial G^ is never stable for n = 4. For n = 5, G\ is in fact linearly stable in the interval 7.558 < m < 7.9804 because all the necessary coefficients are positive (including the discriminant). However, since T2- is negative, and T^+ » 45.3904, it is necessary that m > 45.3904 for 72(771) to be positive. Thus, the polynomials G\ and G2 are never stable concurrently. For n = 6, G\ is in fact linearly stable in the interval 11.5168 < m < 22.9103. However, since T2_ is negative, and T2+ « 70.6461, it is necessary that m > 70.6461 for 72(m) to be positive. Thus, the polynomials G\ and G2 are never stable concurrently.
n 3 4 5 6
I 1 1 2 1 2 1 2 3
ma,
m0l
r,_
0.6334 0.7929 -0.9571 0.9612 -0.8507 1.1340 -0.6160 -0.0774
0.7705 2.3797 -0.2500 6.4782 -0.2442 20.9068 -0.2201 0.0060
-8.8403 -8.8416 -0.2292 -8.7580 -0.2361 -8.8320 -0.2179 0.0061
ri+ 0.7685 2.1553 23.7708 4.4805 45.3904 7.7602 70.6461 86.8392
A, 0.2744 1.0072 7.9804 48.4064 22.9103 77.5479
Table 1. The approximate roots of aj(m),/3j(m),7i(m), and the largest root of 5|(m) for the cases n = 3,4,5, and 6.
319
The remainder of the paper will be concerned with the case n > 7. Note that the slope of A(m) for I ^ 1 is 3(Pj + ZQi 4- cr n /2) which is always positive. This fact, along with Lemma 4.1, implies that the leading coefficients of the terms aj(m),/3j(m),7j(m) and 6i(m) are all positive for all I. Hence for sufficiently large m, the 1 + n-gon will be linearly stable. We now show that there is a unique mass value hn such that m > hn is a necessary and sufficient condition for linear stability. The next lemma explains the order in which the important coefficients become positive as m is increased. The sequence goes as follows: We start out with both a/ and Pi negative and yi positive. As m increases, /3j becomes positive first, but by the time c*j becomes positive, 71 has become negative. (So we don't have stability yet.) Then once 7/ becomes positive again, 61 is now negative. Finally, once m is larger than Ai, all four coefficients are positive and will remain that way for all m > Aj (see Figure 2).
Figure 2. Graph of the coefficients Q|(m),/?|(m), ji(m) and 5j(m) illustrating the order of their roots.
Lemma 4.3 For n > 8, mgl < I V < mai < Tl+ < Aj for 1 < I < [n/2]. Proof: We first show that m^, < mar mai = 2Pi + - - an
For I = 1, we have n 1 7r - - - cot — 2 2 2n
nPi + ^ ( 4 P i + n + a n )
m0l = —
3(A-f + ^)
> 0
<0
where the last inequality follows from Lemma 4.1. For / > 2, we can verify that mp, < mai numerically for n = 8 , . . . , 13. For n > 14, we can rewrite
320
equation (20) as i„ 1_ m0l = Qi-- 3 P ' " 6 n
1
Rjt 3Pt + 9Qi + |CT„ '
Using the fact that Pi > Ri and Qi and an are nonnegative, we have m0l
-Pi - -an + -Ri-
Since m Q , = 2P/ — crn, it suffices to show that \Pi >Qi + | * n + IR,.
(23)
For the case I = 2, we can use the expressions from (15), (16) and (17) to obtain 7 „ ^ 5 1„ 13?r T T „ ^P2-Q2-rn--Ri >0 = C T n - - c o t - - c o t by our usual estimates. To prove inequality (23) for the cases I > 3, it suffices to show that Pi > 5<7n/6, since Pi > Qi and Pi > Ri. Since for any fixed n, Pi is increasing, it suffices to show that F3 > 5
(24)
Recall that the leading coefficients of a/(m), 0i(m) and 71 (m) are positive for all I. Simply stated, ai(m) > 0
iff
m > mai
(25)
0i{m) > 0
iff
m>m0l
(26)
ji(m) > 0
iff
m < T,_
or
m > Tl+
(27)
(see Figure 1). Therefore, we have li(rnai)
= a?(m Q l ) - 4/? i (m Ql )
= -m(mai) <0 where the inequality follows from (24) and (26). Inequality (27) then implies that Tj_ < m a , < T j + .
321
Next, we have that 7i("»ft) = of (m f l ) - 401 ("»/»,) = af(m0l) >0 with the strict inequality following from (24) and (25). Therefore, either m/3, < Tj_ or mpi > Fi+. However, the latter inequality implies that m^, > m Q , which contradicts (24). Finally, we note that S,(r,+) = f3n? + c?w 2 ai(7, - 32ft) - 27c4w4 = -32cfw 2 a,/3; - 27c4w4 <0 where aj,/3j and 7/ are evaluated at m = Ti + . Thus since 6i(m) is of odd degree with leading coefficient positive, its largest root, Aj, must be greater than Fi+. This completes the proof. Lemma 4.4 For each n > 7 and for each integer I, 1 < I < n/2, the poly nomial Gi is linearly stable if and only if m > Aj. If n is even, then Fn is linearly stable if and only ifm > r « + . Proof: The second statement follows directly from Lemma 4.3. In order for F<* to be linearly stable, a«,/J» and 75 must all be positive. This occurs if and only if m > T^ + . For the quartic Gi, we first handle the case n — 7 separately. (This is necessary because Lemma 4.3 is not valid when n = 1 and / = 2.) We list the relevant data in Table 2. It is clear from the data in all three cases that G/ is linearly stable when m > A^. To argue this is also a necessary condition, one can calculate that ^ ( I V ) < 0 and that Aj is the only root of 61 (m) larger than Tj + . Thus the only way to have all four coefficients positive is when m > A;. This finishes the case n = 7. Note that the last column of Table 2 is increasing, and that the 1 + 7-gon is linearly stable if and only if m > 139.8523. / 1 2 3
mai 1.3094 -0.2846 0.9819
mpt -643.2843 -0.1814 0.3242
Ti_
r<+
-9.0892 -0.1813 0.3274
11.8784 98.8129 135.5674
A, 15.7260 110.2791 139.8523
Table 2. The (approximate) roots of ai(m),/?i(m),7j(m), and the largest root of Si(m) for the case n = 7.
322 To do the cases n > 8, Lemma 4.3 shows that it is sufficient for m > Aj as then all the relevant coefficients will be positive. For the other direction, we know that aj,/3j, and % are positive concurrently if and only if m > T j + . To show it is necessary that m > Aj, we need only show that Aj is the only root of 6i(m) larger than r ; + . Recall that (19) 6i = 0Hi + c?(m + ^-)aai
- 32c2(m + ^ ) a , f t - 27cf(m + ^ ) 2 .
Differentiating this expression with respect to m and evaluating at m = Tj + yields Si{Ti+) = -24c2/x/?j - 36c2/xctj/?,' - 32c2a,/?j - 54cffi where QJ and /?j are evaluated at m ~ Tj + and (JL = Tj + 4- crn/2. (Here we use the fact that a2(Ti+) = 40i(Ti+).) Lemma 4.3 gives aj(ri+) > 0 and 0i{Ti+) > 0 so that the above quantity is negative. Thus, not only is the discriminant negative at m = rj+, but it is also decreasing. Further computation and estimation reveals that the third derivative of 71 with respect to m is positive for m > Ti+. This in turn implies that fy'(m) is concave up for all m > T j + and therefore 6i(m) has exactly one root (Aj) larger than T j + . This completes the proof of the lemma. As the largest root of the fifth-degree polynomial 6j(m), it is unlikely that an explicit formula for Aj exists, not to mention the fact that it would probably be incomprehensible. Thus it is reasonable to seek an upper bound for Aj which has a useful explicit formula. Recall that Gi(z) = (z2 + ajz + A ) 2 + 4w 2 c 2 z. £,From Lemma 4.3 we know that when m > Tj+, aj,/3j and 7J are positive. This means that the polynomial z 2 + auz + /3j has two real and distinct negative roots r\ and r-i with r\ < -ai/2 < r 2 < 0. Since G/(r 1]2 ) = 4u>2cfri,2 < 0 and Gt(z) > 0 for z > 0, the Intermediate Value Theorem tells us that Gj has two negative real roots si and s 2 satisfying si < r\ < —ai/2 < r 2 < s 2 < 0. When Ti+ < m < Aj, we know from Lemma 4.3 and from the proof of Lemma 4.4, that the discriminant of Gi{z) is negative. This means that s\ and s 2 are the only real roots of Gi(z). When m = A;, the discriminant of Gi vanishes, and a negative real, repeated root is born in addition to s\ and s 2 . Finally, as m increases past Aj, Gi{—oci/2) eventually becomes positive and Gj has four distinct negative real roots. That is, Gj will be linearly stable (see Figure 3). Finding a sufficient condition for Gi(—oti/2) > 0 leads to the value m = Aj given by = At + 2y/2\ci\ + A« =
\&+ 2V2|c t |) 2 --Bi
+
2V2\ci\(rn.
323
Figure 3. Graph of the quartic Gi(z) for specific m-values from the regimes Tj + < m < Aj, m = Ai and m > A|, respectively.
It is a straight-forward argument to show that Ai is positive for all I from which it follows that Aj is real and Ai > Tj + . For the rest of the paper we will only be concerned with the larger root of 7j(m) so that Tj is understood to represent Fi+. Lemma 4.5 For n>7,
ifm>Ai,
Ti
then Gi is linearly stable. Moreover, for
1 < I < [(n - l)/2].
Proof: Since A/ > Tj, the coefficients aj,/?j, and 7J are all positive for m > Aj. So by our arguments above, it suffices to show that m > Aj implies that G j ( - a / 2 ) > 0. We compute that Gi(-«//2) = ^lo7 j 2 - 2 a ; 2 C j % which is positive if and only if -v* W^J 7, - i2ct u
•> / 32c?u,V„/2 - 2P1 - n/2) if / = 1 > I if/^i. 3 2 c 2 w 2 ( ( 7 n / 2 _ 2Pl)
It is not difficult to see that the quantities on the right-hand side of the above inequality are negative. Therefore, to show that Gj(—aj/2) > 0, it suffices to show that 7J2 — 32c2w4 > 0 or equivalently, m2-2(At
+ 2\/2|cj|)m + £j - 2v/2|cj|crn > 0.
The largest root of the quadratic in this inequality is m = Aj. Thus we have shown that for m > Aj, Gj(—Qj/2) > 0 which implies that Gj is linearly stable. This proves the first statement of the lemma and shows that Aj > Aj. The only inequality left unproven is Aj < Tj+i. This is significant, as it relates the stability of Gj with that of Gj + i. To prove this, we make use of identity (ii) from Section 3.2: Pi+i - Pi + Qi+i - Qi = Ri + Ri+i.
(28)
324
We explain the case I > 2 (the case I = 1 is similar). Note that |cj| = 2i?j since Ri is nonnegative. We have from equation (28) that Ai+i = 2<7n + 8Pi + 10Qj + i + 8Qi + 8(Ri + Rt+1) > 2an + &Pi + 18Qi + 4\/2i?i = Ai + 4V2Rt where the middle inequality follows from the fact that i?j is nonnegative and that Qi is strictly increasing in I. Next, equation (28) also gives us Af+1 — Bi+i > radicand of A{ after some computation. Since Ai+\ > Ai + 2\/2|cj| and the radicand of Tj + i is greater than the radicand of A;, it follows that Tj+i > Aj. Lemma 4.5 not only provides us with an upper bound on the seemingly elusive quantity Aj, it also provides us with the bifurcation value for the 1 + n-gon relative equilibrium. Since Aj < Tl+i < A ; + 1 , it is clear that as m increases, Gi will become linearly stable before Gi+\ does. This proves that the polynomials Gi become linearly stable in succession. In some sense, this means that the more the configuration is twisted, the more unstable it becomes, or conversely, the most stable perturbations involve less rotation of the n-gon (including the one that moves the central mass). If n is even, then since A n -a < Tn, F« is the last polynomial to become linearly stable and this happens when m > T«. If n is odd, then G»-i is the last polynomial to become linearly stable and this happens when m > A « - i . For n even, we have a precise formula for the bifurcation value while for n odd, we have explicit expressions bounding the bifurcation value above and below. This completes the proof of our main result. Theorem 4.6 Forn even, the 1 + n-gon relative equilibrium is linearly stable if and only ifm > r » . F o r n odd, the 1 + n-gon relative equilibrium is linearly stable if and only ifm > A » - i . 3
5
Estimating the Bifurcation Value
As explained in the last section, we have an explicit formula for the bifurcation value given by m = T« for n even. When n is odd, the best we can do is bound the bifurcation value below by r « - i and above by Ar»-i. Fortunately, these two bounds are asymptotic to each other as n approaches infinity. Table 3 lists the bifurcation value hn for several different values of n. In this section
325
hn/rn6 0.9372 0.9530 0.9590 0.9677 0.9709 0.9763 0.9782 0.9819
n 7
hn 139.8523 8 212.2611 9 304.1366 10 420.9930 11 562.1966 12 733.9608 13 934.9493 14 1,172.075
K
n 15 20 50 100 500 1,000 10,000 100,000
1,443.300 3,446.298 54,274.97 434,797.6 5.43780323 x l O 7 4.350332026 xlO 8 4.350365372 xlO 1 1 4.350365691 xlO 1 4
/in/rn3 0.9830 0.9902 0.9981 0.9995 1.0000 1.0000 1.0000 1.0000
Table 3. The bifurcation value hn, where the 1 + n-gon becomes linearly stable, for various values of n and its comparison with the estimate rn3.
we prove that hn increases asymptotically to rn3, where r is given by the formula: 13 + 4>/l0 y , 2TT
1
3
Using a computer, one can calculate that r = 0.435036581297 is accurate to twelve decimal places. Remarkably, Maxwell calculated r to be 0.4352 in his work in 1859.4 He achieves this by correctly "guessing" the force (only in the case with n even) which will produce the most instability, although he provides no proof for why this is the case. Still, it is a great tribute to his insight and ability that Maxwell was able to arrive at this result so well ahead of his time. We handle the case when n is even first. Calculations show that n-l
n-l
1 - COS#fcCOS0k"
2 ^
k=l
1 - COS 0k
Kk
Jt=l
n-l
,
-
n-l
— fill 4- V^ —— _ _ V ~ 4 ^—' r 3 . 2 ^ fcodd
n-l
+ E Ac odd
2cos0 fc 2rnfc
_
— r..'
To ease on notation, we define the sums (for n even) n—1
£n
E x ^ * = kodd E -■
kodd i
n—1
1
IJ
nk
nk
i
IJ
n
*
A similar calculation yields Q» = -cr n /4 + ^ n . Plugging these expressions for
326
P^ and Q%
mto
formulas (21) and (22) yields
T» = 26£n - 4n„ - ^
+ 4V(5^„ - 7jn)(8£n - T7„).
We will show that £„ is order n 3 while T?„ and a n are order n l n n . Hence, r n ~ (26 + 8v/l0)£ n . L e m m a 5.1 £„ is asymptotic to fn3 where
- = -L 3 V — 1 —3
T
47r 'f-; ( 2 f c - l ) " Proof: We will prove the case when n/2 is even. The argument for n / 2 odd is virtually identical. Using the symmetry of the n-gon and the fact that rnk = 2 s i n ( ^ ) ( we have £„ =
1 / 3 7r , 3TT a ( n / 2 - 1W\ csc - + csc3 — + ••• + csc 3 i - i ^4 \ n n n J
(29)
T
We will show that lim f n M 3 = f. Using the fact that cscx > 1/x, it is easy n-«oc
to see that n3
>
4*3 V1+33
+
(30)
(n/2-l)3;
Using the Taylor expansion of sin 3 x yields x 3 — | x 5 < sin 3 x and therefore 3
CSCJ X < —
1
1
1
r-r = -3 + —
xr
(31)
x3 - | x 5 x3 x(2 - x 2 ) provided x < \/2. In order to make use of this inequality, we will split the terms in £ n into two groups. Let fcn be the largest odd integer less than n/6. For the terms with k > kn, csc ^ < 2. For any k < fc„, we may apply inequality (31) to obtain c
sc 3 (*f) .. 1 + n3 " fcM '
1 fc7r(2n2-fc27r2)'
Letting f(u) = l/(u7r(2n 2 - u27r2)), it is clear that / ( u ) is positive and de creasing for 1 < u < kn. Since there are exactly n/4 terms in (29), we have that
^<4^(1 + P+ - + ( M 3 J + 4 - 4 ( / ( 1 ) <
i ^ 3 V1 + 35 + " ■ + (n/2 - 1)3 J
+
+
^J
167r(2n2 - TT2)
+
2tf'
(32)
327
Letting rn-
47r 3 V 1 + 3 3 + ' " + ( n / 2 - l ) 3 J '
inequalities (30) and (32) give Tn
fn
.
" ri3 "
Tn
1
n
' 167r(2n 2 - -TT 2 ) ' 2 n 2 '
Letting n go to infinity yields the desired result. Lemma 5.2 an and r]n are asymptotic to ^ Inn and ^ Inn respectively. Proof: We verify the result for an first. Again using the symmetry of the n-gon, we have n
0~n
1
=£
CSC
7rA; n
1 2
+
when n is even. Since the function g(u) = csc ^ is decreasing for 1 < u < n/2, we have /•n/2
g{2) + p(3) + • • • + g(n/2)
< I
g(u) du < g(l) + g{2) + ■■■+ g(n/2 - 1)
using lower and upper approximating sums. It follows that
r
/2 g(u)
i du+- <
/•n/2 On
< 5(1) + /
which evaluates to ■K \ i n In fcot < ^n < 2 n ^From this inequality it follows that
7T
CSC
n
+
p(u) du
n In (cot 7T
?.n)
Urn -—;—r—— = 1 n-.oo (nlnn)/7r The same argument works for nn by using h(u) = csc ^(2u — 1) in place of g(u) and carrying out the same approximations. The 1/2 factor arises from the integration of h(u) and is expected since r]n contains half of the terms of an. This completes the proof of the lemma. The case when n is odd is slightly more complicated since we do not have an explicit expression for the bifurcation value and the formulas we do have
328
are more involved. We will let c denote the crucial index (n — l ) / 2 . Recall that for n odd, Tc < hn < A c . Since 6kc — i^k- nk/n, we have f
cos
zJs. kt
cos0fcc=H1 - _„«„* cos ^ k, Q d d \ n
and
.
s m n0 f c c H
sin ^7 *£ k even ,f - o».» .S_wf
fcQdd_
Using these formulas, the symmetry of the n-gon, and rnk = 2sin ^ , we see that c
c
1
k=i r « *
. ,. c o s ^ c o s ^ 'nfc
fc=i
n 1 v^ ■> nk 1 v-> , ,.1.4.1 , 7rfc o 7rfc 1 v~* / , \fc4-i k = - > c s c 3 — + - > (-l) f c + 1 cot — c s c 2 — - T > ( - l ) f c + 1 c o t — fc=i fc=i fc=i
An argument similar to the one used in the proof of Lemma 5.2 shows that the last term in the sum above is order n l n n . Likewise, an argument similar to the one used in the proof of Lemma 5.1 shows that
E csc
o irk1
nA•? v°^°
1
— ~ -=•3 > TTT and 3 n TT ^-^ k 3 nk n ^ (-i)*+i 27rfc V(-l)^cot^csc2^ ~ ^ V ,3 k=l fc=l It follows that Pc is asymptotic to fn 3 . Further computation reveals that
Pc-Qc^
A
4
k+1 +' A /^.^ \t(-D 'I 4
nk C O t
•
fctl
Since the right-hand side is order n l n n , it follows that Qc is asymptotic to Pc. Finally, we have that sin ^ n
sin ^ n_
7rfc = I y-(_i)*+i cot ![* 4 fc=i ^
v
n
329 which means that Rc is lower order than Pc and Qc- Since Ac = Ac + 4V2RC + yj {Ac + 4\/2Rc)2 - Bc +
4V2Rcan,
we see that T c ~ Ac. Using the fact that Pc ~ Qc, equations (21) and (22) imply that hn ~ Tc ~ (26 + 8v / 10)fn 3 . We have proven the following theorem: Theorem 5.3 The 1 + n-gon relative equilibrium becomes linearly stable at a central mass value which is asymptotic to rn3 where T
~
3 27T
' ^
(2Jfc - l ) 3 '
Acknowledgments: I would like to thank Glen Hall for his advice and encouragement regarding this work. References 1. B. Elmabsout, Stability of some degenerate positions of relative equi librium in the n-body problem, Dynamics and Stability of Systems 9, 305-319, (1994). 2. M. Gascheau, Examen d'une classe d'equations differentielles et applica tion a un cas particulier du probleme des trois corps, Comptes Rendus 16, 393-394, (1843). 3. J. C. Maxwell, Stability of the motion of Saturn's rings, W. D. Niven, edi tor, The Scientific Papers of James Clerk Maxwell, Cambridge University Press, Cambridge, (1890). 4. J. C. Maxwell, Stability of the motion of Saturn's rings, S. Brush, C. W. F. Everitt, and E. Garber, editors, Maxwell on Saturn's rings, MIT Press, Cambridge, (1983). 5. K. Meyer and G. R. Hall, Introduction to Hamiltonian Dynamical Sys tems and the TV-Body Problem, Applied Mathematical Sciences 90, Springer, New York, (1992). 6. K. Meyer and D. S. Schmidt, Bifurcations of Relative Equilibria in the N-Body and Kirchoff Problems, SIAM J. Math. Anal. 19, 1295-1313, (1988).
330
7. R. Moeckel, Linear Stability Analysis of Some Symmetrical Classes of Rel ative Equilibria, Hamiltonian dynamical systems (Cincinnati, OH, 1992), 291-317, IMA Vol. Math Appl. 6 3 , Springer, New York, (1995). 8. R. Moeckel, Linear Stability of relative equilibria with a dominant mass, J. Dynam. Diff. Equations 6, 37-51, (1994). 9. J. Palmore, Measure of degenerative relative equilibria, I, Annals of Math. 104, 421-429, (1976). 10. E. J. Routh, On Laplace's three particles with a supplement on the sta bility of their motion, Proc. London Math. Soc. 6, 86-97, (1875). 11. D. S. Schmidt, Spectral stability of relative equilibria in the n + 1-body problem, New Trends for Hamiltonian Systems and Celestial Mechan ics 8, 321-341, E. A. Lacomba and J. Llibre, editors, World Scientific Publishing Co., (1995).
ANALYTIC CONTINUATION OF CIRCULAR A N D ELLIPTIC KEPLER MOTION TO THE GENERAL 3 - B O D Y P R O B L E M JAUME SOLER Departament
d'Informatica i Matemdtica Universitat de Girona C. Lluis Santald s/n, 17071, Girona, E-mail: [email protected]
Aplicada Spain
The existence of families of periodic solutions of the general planetary 3-body problem in which one of the orbits is ellipic is shown. Each family derives from uncoupled circular and elliptic Kepler motion in a given resonance, with the eccen tricity of the elliptic orbit as the parameter. The standard method of Poincar^'s analytic continuation cannot be directly applied due to the vanishing of a deter minant, so a strong form of the implicit function theorem is used.
1
Introduction
Let PQ, PI and Pi be three material points of mass mo, m\ and 7712, re spectively, moving in a plane under their mutual Newtonian gravitational attraction. We are particularly interested in the planetary problem, in which one of the bodies is much larger than the other ones. As mo will be asumed to be the large mass then we will call PQ the Sun and Pi and Pi the planets. We will set mo = 1, mi = i/i/i,m2 = i^p, where fi is a small parameter. We are interested in periodic solutions which, in the limit \x —► 0, become circular and elliptic resonant orbits of two uncoupled Kepler problems. These peri odic solutions will be shown to exist, when fi ^ 0 provided a certain integral does not vanish. This integral is computed numerically in a few instances and shown to be different from zero.
2
The equations of motion
We consider an inertial frame with origin at the center of mass of PQ, PI and P2, denote by
m
° l l - 112 ,
m
l | | • ||2 ,
m
2 i | . i|2
4> = -g-llftll +—iigiir + - y I momi
mom.2
m\m2
ko-gill
II90-92II
Iki-ftI
ti\
toll
(i)
332
We will write the Lagrangian (1) in Jacobi coordinates (see 3 for a definition in the general n-body problem). In the three-body problem, they are «i = 9i - 9o> «2 = 2
; {m0Qo + m\Qi) mo + mi and, in this new reference system, the Lagrangian of the problem becomes Mi
|2
M2
I|2
, m0mi
C = -ylluill + -2-IIU2H + j
mQm2
mim2
^ +i i ^ ^ i i +i i ^ ^ l j '
, .
W
where UQ has been eliminated by means of the center of mass integral moqo + mi<7i + m2q2 = 0, and where a = mi/(mo + mi), y9 = mo/(m 0 + mi), Mi = m0mi/(mo + mi), M 2 = m2(m0 + m i ) / ( m 0 + mi 4- m 2 ). We write u* = (Xi,!/t), g = (xi,t/i,£2,2/2), apply a Legendre transformation to the the Lagrangian (2)and obtain the Hamiltonian
«-5Bi<*? +tf>+ d5<*! + morrii
mom2 ||U2 + OUl||
tf>
<3)
mim2 ||U2-/3«l||'
where Q = (Xi,Yi,X2,Y2) is the vector of momenta associated to coordi nates q, with Xi = dC/dxi and Yi = dC/diji. In order to normalize the Hamiltonian, we perform a canonical transformation of multiplier l/n on (3) Q = O/MI 9 = 9, which gives a new Hamiltonian n(Q,q)
= ^H(fiQ,q).
(4)
We introduce <TI, 02 by letting Mi = &in, M2 = a2n, write everything in terms of
(TiH 1 + mi
1 + m i - CT2/X
a2\i,
333 so that, dropping the bars, (4) can be written as + /iV + C(/x2),
H = Hi+H2 where, for i = 1,2, H --lM
Ox
+ Yh-
s/xl + vi
and V = V 1 + V 2 + V3 + V4, the V, given by Vi =
v2 =
°\
(6)
°l
(7)
(*? + y 2 ) 1 / 2 ' (zl + y 2 2 ) 1 / 2 '
v3 = a\o2{xix2
+ yiy2)
{xl + vl?'2
(8)
'
o\ 02 (9) ((*! - X2) 2 + (2/1 - 2 / 2 ) 2 ) 1 / 2 ' Notice that if we think of /i as a small parameter, the Hamiltonian (5) is the sum of two independent Kepler problems plus a weak coupling term. In what follows, we will use polar coordinates and their asssociated mo menta (r, 0, R, $) for P\ and Delaunay elements (t,g, L, G) for P 2 (see 3 and the Appendix for the definition of Delaunay elements). These coordinates will be referred to as polar-Delaunay variables. In polar-Delaunay variables, the Hamiltonian (5) of the problem becomes
v4 =
H=
£(*♦$-
r
-^S+nV
+ Otf),
(10)
and the corresponding equations of motion are r=
<j\
(11)
on
1 $
dV
<j\ r 2
d$
1 $2 (7i r 3
~ r2
dV
„,
„ , ,,
- ^ + 0(A 2x
(12) (13) (14)
334
e = a2L-* + n^+0(n2), S = / i ^ + 0(/i 2 ), L = -
M
G = - ^
^ + 0(M 2 ), + 0{n2).
(15) (16) (17) (18)
The equations of the three-body problem have a well-known symme try which has been widely used to show the existence of periodic solutions. Poincare defined a symmetric conjunction of the three-body problem as a con figuration in which the three bodies are in a line, with velocities perpendicular to this line. Then the following result holds: Theorem 1 If att = 0 the three bodies are in a symmetric conjuntion and at t — T/2 they are again in symmetric conjunction, then the motion is periodic of period T. For a proof see 4 As is well-known, the motion is actually periodic only in a rotating frame (x, y) in which one of the small bodies remains always on the x-axis. The trajectories given by the theorem are periodic in the rotating system but they need not be so in the non-rotating system. If the Hamiltonian is written in polar coordinates, it is easily seen that it depends on the angular positions
(19) (20) (21)
335
In Poincare's method of analytic continuation, (19), (20) and (21) are thought of as a system of equations, to be solved for T and some other variables as functions of (i and some parameters by means of the implicit function theorem. 3
T h e flow for /i ^ 0
It is clear that, by rescaling the variables, we can take <j\ = 1 with no loss of generality. We will take <72 = 1 as well: this is not so restrictive as it could seem at first thought because, as we are dealing with analytic functions, if a certain system of equations can be solved for 02 = 1, they can also be solved for neighbouring values of o-iIn order to shorten the written formulas, let's introduce some notation. Define x = (x1,x2,x3,x4), y = (2/1,2/2,2/3,2/4) and z = (zi, ■ ■ ■ ,z 8 ) by x= (r,
(e,g,L,G),
z = {x,y). With these conventions, the system of differential equations (11)-(18) becomes
x\
y)
=(Ua(x)\
(Ux(x,y)\ +fi
2
+U{lih
{wo(y)J {w1(x,y))
or, in terms of the z variables, z = F0{z) + fiF1{z) + O(ji2)
(22)
Let zo = (zo,2/o) be a vector of initial conditions, and let z(zo,t) be the flow of the system (22) with initial value ZQ at t — 0. The flow can be expanded as a power series in /j. +00
z(z 0 ,0 = £ / 2 ( f c ) ( z o , < )
(23)
Jt=0
where z^\zo,t) is the flow of the problem when fi = 0 (the unperturbed problem) with initial conditions 20, z^ is the solution of the first variational equations, given by Lagrange's formula zW{zo,t)
= Z(z0,t)
f 2TV*b,T)F,[><0>(zb,T))dT, Jo
and the matrix Z(ZQ, t) is a fundamental solution of the linear system Z(t) =
^(z^(z0,t))Z(t).
(24)
336 Such a fundamental solution can be easily obtained provided the flow of the unperturbed system is explicitly known, in which case we have
dz^ict)
Z(z0,t) =
dC
(25) <=*0
As is well-known, for initial conditions ZQ + 6z in a neighbourhood of ZQ, each term in the series expansion (23) can be expanded as a power series in the 6's and coefficients functions of the time t
i\6z .*)
/ zW(zQ + 6z,t)-j=0
where z^k'^(6z, t) is a vector whose components will be denoted by a subscript i ranging from 1 to 8, each component z\ '3'(6z,t) is of the form /xfc times a homogeneous polynomial of degree j in bzi and coefficients functions of t, and the dependence on ZQ is not explicitly shown (notice that z^k'^ is of degree k + j jointly in p, 6r). The terms z^0'^ are those of the unperturbed solution, and they are given by the direct sum of two independent flows
z^\6z
,*) =
"\y^\Sy,t)J
By the same reason, the fundamental matrix (25) splits as
= rz&X(i){6x 0
Z(t) =
y(i)(Sy,t)J-
(26)
where
X(j){8x ,t) =
t) ddx
yu )(Sy,t) =
dSy
are matrices whose entries are homogeneous polynomials of degree j in &Cj, <5j/», respectively. Any term z^\6x, 6y, t) can be easily obtained through Lagrange's for mula (24), using (26) and expanding the factor F\{z^Q\zo + 6z, r)) as a power series in 6z as well. The zeroth order term in the 6's of z^ will then be given by *
Jo
Z{o](r)Fx(z <°'°)(T))dT.
337
4
Existence of periodic solutions
We consider families of periodic solutions of the unperturbed problem with the planets Pj and P2 moving, respectively, in a circular and an elliptic orbit around the origin, in such a way that the semiperiod of Pi is a multiple of the semiperiod of P\, and the eccentricity e of P2 is the parameter along the family. It is clear that, by rescaling, we can restrict ourselves to the case in which Pi moves in a circle of radius 1 and period 2n. It will be shown that these families can be continued to solutions of the full three-body problem, periodic in a rotating frame. Let k be any positive integer > 2, let Lk = vk and let's consider the solution with initial conditions r 0 = 1 + 6r, 4>o = 0, Re = 0, $0 = 1, together with to = 0, 9o = 0, L0 = Lk + 6L, G0 = G, which, for \i = 0 are those of two uncoupled Kepler motions: a circular orbit of radius unity and period 2ir and an elliptic orbit of semimajor axis and mean motion 1/fc (semiperiod 7* = fcir). The value of G is taken as the parameter defining the family: as L is kept fixed in each family and the relation (40) holds, G can be taken as the parameter defining the family instead of e. We will show that the conditions of symmetry (19)-(21) can be solved for \i ^ 0. Specifically, we want to solve the system (f>(Sr, 6L, G,T + 6T, /1) - g(6r, 6L, G,T + ST, /1) = kir, e(6r,6L,G,T + 6T,n) = n, R{6r,6L,G,T
+ ST,fi) = 0,
(27) (28) (29)
for 6r, SL, 6T as functions of fi and the parameter G. System (27)-(29), as it stands, cannot be solved by means of the implicit function theorem because the determinant d(
338
into equation (29), yielding an equation in 6r and \i which has a branching point at /i = 0 and can be solved provided the integral
C{k,e) = (-l)k
f
(-V( 0> (T)cosTdT + 2Vj, 0) (T)sinr) dr
(30)
does not vanish, where Vr (r) and Vi (T) stand for the partial derivatives of V evaluated on the unperturbed solution z^°'°\t). The problem is analogous to the behaviour of the first kind periodic solutions of the restricted threebody problem near the (n + l ) / n resonances, where they cannot be continued but are shown to connect with second kind solutions. In that case, there are two conditions of symmetry and the determinant vanishes in the mentioned resonances. Then one of the equations can be solved and the result substituted into the other, thus showing the connection between the families; see 5 or 2 for details. Of course, the integral in (30) cannot be represented by a closed expression in terms of elementary functions and has to be computed numerically. i,Prom (8) and (9), using polar coordinates (r, <j>) for Pi and orbital elements for P2, we get the following expressions for the relevant terms of the potential (Vi and V2 are easily seen to make no contribution to the integral) V3 = rd~2 cos(4> - v), V4 = - ( r 2 + d2 - 2drcos(<j> -
v))~1'2,
where d = a{\ — ecosi?), and a,e,v,E are, respectively, the semimajor axis, eccentricity, true anomaly and eccentric anomaly of Pi- Now on the unper turbed solution we have r(t) = 1, <j>(t) = t, a = fc2^3, and if we write cost; and sinu in terms of e, E, cosw = (cosE — e)/(l — ecosi?), 2
sinu = v T ^ e s i n £ / ( l - e c o s £ ) ,
(31) (32)
then the integrand in (30) depends only on the variables k,e,E,r and the integral can be easily computed taking the eccentric anomaly as the variable of integration. We quote some numerical values of C to show that it does not
339
vanish in general e = 0.1 e = 0.2 e = 0.3 e = 0.4 A: = 3-1.14522 -2.78782 -6.14385 -16.84 k = A 0.21531 0.80801 1.99539 4.74640 fc = 5-0.05375 -0.27200 -0.82452 -2.08195 1.06536 k = 6 0.02245 0.10459 0.37528 k = 7 -0.01422 -0.04897 -0.18312 -0.58683 fc = 8 0.01070 0.02882 0.09682 0.33832 The result is formally stated in the following theorem, in which we restrict the possible values of e in order to avoid collision with P j . Theorem 2 Let k be any integer > 2, eo € (0,1 — k~2^) and GQ — Lky/T eZ. If C(k, eo) 7^ 0 then there exist MO > 0, 7 > 0 and functions 6r(G, M, t), 6L(G, M, t), 6T(G, M, t), defined for n G [0, Mo), G G (G 0 - 7, G 0 + 7) C (0, Lk), analytic in G, (x1^ and t which satisfy the equations (27)-(29). Proof: in all the calculations that follow, d stands for an analytic func tion of its variables beginning with terms of order at least i and c\ and c-i are constants whose value is not relevant to the final result. ^From relations (66), (67) and (68) in the Appendix we get the following expression for equations (27) and (28) 3fc7T
0 = ST + CIM - - y - < ^ 2 + STO{ST, SL) + fj.O(6T, SL, Sr, M), 1
0 = -ST-
(33)
37T
—j^SL + C2M + 02{ST, SL, ST) + »0{ST, SL, Sr, fi).
(34)
If we write ST and SL as formal series in Sr and /J,, the method of in determinate coefficients gives the following expansions for the solution of the system (33), (34), ST = -clfi SL
+ -knSr2 + n
- C i + kc2
, „ ,,
+
03(Sr,n),
-
^ 2 7 3 - M + <%(*•, M).
(35) (36)
^From (69) we get the following expansion for equation (29) R = Cn-
SrST + nO(6T, SL, Sr, M) + 04(Sr)
(37)
and substitute (35) and (36) in (37); we then obtain 3Jfc7T
0 - CM + ^-Sr3
+ iiSrfa + tffc + Sr403,
(38)
340
where ft = ft(5r, /i) are analytic and C is given by expression (30). In order to solve equation (38), we rescale by denning Sr = /J 1/3 Ar, divide by a common factor fi and let £ = /x 1 / 3 . We get 0 = C + ^ - A r 3 + eArd(e, Ar) + eAr4C2(e, Ar) + e3Cs(e, Ar). where Ci are analytic functions. For t = 0 this equation has a real solution given by
Ar =
J-2C
V 3fc7T and, when C ^ 0, the implicit function theorem applies and we get eventually an expansion of 6r as a power series in /x 1 / 3 fr
5
= &
/ 3 + 0
>("'
/ 3
)-
Appendix
In what follows some expansions relative to the transformation between polar coordinates and momenta and Poincare variables will be given. As we will use the classical orbital elements, we first recall some formulas that we will need. Two of these orbital elements are the semimajor axis a and the eccentricity e of the elliptic orbit; they are related to Delaunay variables L, G by a = L2,
(39) 2
2
e = y/\ - G /L .
(40)
A third orbital element gives the position of the orbit in the plane: it is the argument of the pericentre, that is, Delaunay coordinate g. A second angle is needed to give the position of the body along its orbit: here a choice is possible among Delaunay variable I, classically known as the mean anomaly, the true anomaly v (angular position measured from the pericentre), and the eccentric anomaly E, an auxiliary angle defined by the relation E tan
1-e
"VTTi
v tan
2-
(41)
The mean and true anomalies are related to each other through Kepler's equation ^=£-esin.E,
(42)
341
which can be shown to be just an expression of the conservation of the angular momentum. In terms of the orbital elements, position and velocity can be calculated. The distance to the origin is given by r = a(l-ecos£),
(43)
from where the radial velocity r can be found r = aeE sin E. Now E can be obtained by differentiating Kepler's equation (42), bearing in mind that the mean anomaly is given by £ = a _ 3 / / 2 t; this gives finally the following expression for R R = ea-1/2(l-ecosE)-1sinE.
(44)
The position angle
(45)
and the angular velocity <\> can be found, if necessary, through the angular momentum G = <^>r2. Neither the classical orbital elements nor Delaunay elements are well de fined on circular orbits because the argument of the pericenter is not denned. If we give initial conditions (in rectangular coordinates) for a circular orbit, then there exist arbitrarily small variations in these initial values which will cause the argument of the pericenter g to vary from — n to 7r. The same applies to any angular variable having the pericenter of the orbit as origin, i.e. any of the anomalies. In contrast to that, the magnitudes e sin v and e cos v are well denned and depend smoothly on the variations of the initial conditions. The same can be said about e sin £, e cos I, e sin E, e cos E. In the transformation from polar coordinates and momenta to Poincare elements, direct use of the angular variables £, v, E will be avoided, and pairs such as (e sin E, e cos E) will be used instead. Notice that in each one of the pairs (esin^,ecos^), (e sin u,e cost;), (esinE,ecos.E), each component can be expanded as a power series in the components of any other pair. The differences v — E, E — £ and v — £ can be expanded in the same way as well. We will quote a few of these expansions that will prove useful in what follows. A standard reference for the subject is 1. ^From Kepler's equation (42) the following expansions can be derived in a variety of ways esin^ = esini? — esini? ecos.E + 03(esinE,
ecosE),
e cos £ = e cos E + (e sin E)2 + 03(e sin E, e cos E),
(46) (47)
342
e s i n £ = esin^ + esin^ ecos^ + ( ^ ( e s i n ^ e c o s ^ ) , 2
e c o s £ = e c o s £ - (esin£) + 0 3 (esin^,ecos£).
(48) (49)
The difference v — E can be expanded from (41) as v — E = -esinE
+ - e s i n £ ecosi? + C>3(esmE,ecosE),
(50)
and the difference E — i is just given by Kepler's equation (42). We will give explicitly the change from Poincare variables to polar co ordinates and momenta. Poincare variables A, A, £, 77 are defined in terms of Delaunay elements by the following set of formulas (see 6 ) A = L,
(51)
A = * + ,
(52)
i = ^2(L-G)cosg,
(53)
r) = -y/2(L-G)sing.
(54)
We look now for the formulas that change from polar coordinates and momenta to Poincare variables. We have in mind a particular case: that of a circular orbit of radius unity and angular momentum unity; the trajectory is given by r = 1,
1 + fir, 8R, 1 + <5$, 1 + <5A,
and work out the formulas for the transformation. i,From the relation of the energy 2A2
2 V
rf)
r'
we readily obtain A as A=(-J?2+2/r-$2/r2r1/2. i,From equation (43), we obtain the following expression for ecosE ecosE = - l + r.R2 + $ 2 / r ,
343
and from (44) esinE = ifrA - 1 so that we can write e sin E and e cos E as power series in 6r, 6R, 6$. Now, according to (52) and (45), we have A = (P + £-E
+
E-v
whence we obtain, using (42) and (50), \ =
Si{6r,6R,6$),
where S\ is a series in the indicated variables with no constant term. Finally we will find expressions for £ and T) as functions of the polar coordinates and momenta. Notice that, according to the definitions (53) and (54), those variables are not essentially different from e cos g and —e sin g. As g — A — £, we can now write e cos g = cos A e cos £ + sin A e sin £
(55)
e sin g = sin A e cos I — cos A e sin I
(56)
and using (55), (56), (46), (47), we can expand the right hand terms as power series in Sr, 6R, 6$ whose coefficients are trigonometric expressions in
e+r?=l{L-G), whence * = A-\(S2 + V2).
(57)
344
From the same expressions we can write ecos<7 =
-LsA-\2A-\(e + W, l
2
2
1 2
esing = --L 7? A- (2A-i(C + ?3 )) / , and the right hand sides can be expanded as power series in 6 A, £, 77. Invert ing (55) and (56) we get esin^ and ecos^ as power series in the mentioned variables and coefficients trigonometric polynomials in A, and from (48) and (49) we eventually find esini? and ecosE, again as series of the same type. From (43) and (44) and what we already know, we get the needed expansions for r and R. The expansion of <j> follows a similar reasoning using 4> = X + v-E
+
E-£,
and expanding v — E and E — I as before. The above procedures will yield the series up to any order if one takes care to expand to the required order in each step. The result for polar coordinates and momenta as functions of Poincare variables is, up to second order and expanding with respect to 6X, r = 1 + 2<5A — cos A £ + sin A 77 3 3 + 6A 2 - - cos A 6A£ + - sin A 6A77 + sin A 6A£ + cos A 6Xr) + sin 2A £77 + sin 2 A f2 4- cos2 A T72, 0 = A + <5A + 2sinAf + 2cosA77 — sin A <5Af - cos A 6Arj + 2 cos A 6X£ 5 5 5 — 2 sin A <5A?7--- sin2A 772 + - cos2A £77 + - sin2A £ 2 , 4 2 4
3
3
R = sin A £ + cos A 77 — - sin A 8A£ - -- cos A 0 A77 + cos A 6X£ ~ sin A 6XT] + 2 cos 2A f 77 + sin 2A(f2 - 772)
$ = 1 + 6 A - U2 Poincare variables in terms of polar coordinates and momenta are given by A = 1 + 6$ + ~6r2 -- 2<5r6$ + i<5i?2 + 25$ 2 , X = (p + 64>-- 26R - hrSR + 6R6$, £ = cos 4>(-6r + 2<5$) + sin^> 6R
345
+ sin 4>{8r8(j)--26cf>8$ +UR6*) + cos <j>(6r2 + 26$2 --6r6^ -
+ 6(f>6R),
r} = sin4>(6r - 2<5$) + cos
+ cos
+ pR6$)
- sin
+ 8
We finally give some of the terms z^k'^(6z,t) in the expansion of the flow near a circular orbit as a power series in the small variations of the initial conditions. Let V be the transformation from polar coordinates and momenta to Poincare variables as computed above, that is, (A,A,C,r?)=P(r,^fl,$), and let .^(Ao, A0,fo,77o) be the flow of the Kepler problem in Poincare vari ables, given by A(t) = Ao, \(t) = Ao +
AQ
3
t,
Z(t) = &, v(t) = %• We are interested in the flow in a neighbourhood of the circular orbit r = l,<j> = t,R = Q,$ = 1. If we take initial values 1 + Sr, 6
= V-1 TtV{\ + 6r,6cp,6R, 1 + «*).
The zeroth and first order terms in the £'s of this flow are r(o,o) = 1 ;
0
tf
^(0,0)
= Qt $(o,o) =
1(
and r (o,i)
=6rcost
^(o.i) = -26rsmt RW 0
= -6rsmt
& 'V = 6$.
+ 6Rsint + 6$(2-2cost), + 6
(58) + 4smt),
(59) (60) (61)
346 We will use a subscript star to indicate that the particular initial values 6(f> = SR = (5$ = 0 are taken, so that only 8r ^ 0. In this case we have ri°' 2 ) = 6r2(l - cos* + sin 2 t),
(62)
4 ° ' 2 ) = 6 r 2 ( - - t + 2 sin t + - sint cost),
(63)
R[°'2)
$i°'
2)
= Sr2 {sin t + 2 sin t cost),
(64)
= 0.
(65)
The flow j/°)(2/o,t) which describes the unperturbed motion of P^ with initial values j/o — (^A &9, L + 6L, G + 5G) is t(t) = 6e + (L +
6L)-%
9(t) = 69, L{t) = L + 6L, G(t) = G + SG The only term in /x which will be relevant to the problem is R^'°\ matter of fact, the terms x^1'0^ are given by
x^{t)
= M*(o)(i) j T
As a
X^irWiiz^ir)), 1 0
and from this formula we get the following expression for itf - ) i? (1 ' 0) (t) = iisint f + fxcostf
(-VP(T)<°>
sinr - 2vj,0) (r) cosT) dr
(-VW(r)cosr
+ 2Vj, 0 ) (r)sinr) dr,
where Vr (T) and Vi (r) stand for the partial derivatives of V evaluated on the unperturbed solution z^0,0^(t). In order to solve the equations (27)-(29) we write the above expansions with initial values 8<j> = 0, 6R = 0, 6$ = 0, 81. = 0, 8g = 0, 8G = 0, set t = Tk+8T and expand in 8T. We indicate these operations by a superscript star and quote the results, using generic constants a* throughout to stand for quantities whose value is not relevant to Theorem 2. The expansion for the angular variable
0*(6T,6L,6r,M) = kit + ST + aui - — Sr2 + ST 0(6T, 8L, Sr) + /10{8T, 6L, Sr, /x),
(66)
347
where Ok stands for a series in the specified variables beginning with terms of degree at least k. We remark that in equation (66) there are no linear terms in either SL or Sr and there are no quadratic terms in either SL2 or 6L6r. Maybe a few comments on the nature of these terms will be helpful to get some insight into the physics of the problem. Terms in 6T are natural to appear just because an expansion around Tk has been made. The absence of terms in SL (both linear and quadratic) is clear because any change in the orbit of P2 could possibly affect the motion of Pi only through the gravitational interaction, that is, they must appear multiplied by p.. The absence of a linear term in Sr is due to the fact that the expansion of the energy (the value of Hamiltonian (5) when c\ = a2 = 1 and \i = 0) just happens to have no term linear in Sr and so neither has the period. The expansion for g, the other term in equation (27), is g*(ST, SL, Sr, p) = a2p + p. 0(ST, SL, Sr, p).
(67)
As for equation (28), we can write £*{6T, SL, Sr, p) = L^ST
- ZL^TkSL
(68)
+ a3p + 02(ST, SL) + p 0(6T, SL, Sr, n) Finally, we give the expansion of equation (29) R*{6T, SL, Sr, n)=C/i-
SrST + nO(ST, SL, Sr, ft) + 04(Sr)
(69)
where C is given by the following expression C = (-1)* /
(-VJ:0){T)cosrdr
+ 2V(J?)(T) s i n r j dr.
(70)
It is easily seen, even though they haven't been calculated, that there are no terms in Sr3 in the expansion of R*. First and second order terms are known from (60) and (64) so that, if Sr ^ 0, we can write Ri°\t)
= -Srs\nt
+ 6r2{sint + sm2t) +Sr3p(t) + 04(Sr),
where p{t) is an analytic function; now if we take t = kn we get R(Q)*(kn) - Sr3p{kw) + 04(Sr) but on the other hand we know that fi(°)*(fc7r) = 0 for any value of Sr because at time t = k-n the planet P2 is at an apsis of its orbit and so its radial velocity R is 0, which implies p(kn) = 0 and justifies the absence of 0$(8r) in equation (69).
348
Acknowledgments The author is partially supported by a DGES grant number PB96-1153. References 1. Brouwer and Clemence, Methods of Celestial Mechanics (Academic Press, New York, 1961). 2. P. Guillaume, Families of Symmetric Periodic Orbits of the Restricted Three Body Problem, when the Perturbing Mass is Small, Astron. Astrophys. 3, 3-57 (1969). 3. K. R. Meyer and G. R. Hall, An Introduction to Hamiltonian Dynamical Systems (Springer-Verlag, New York, 1991). 4. H. Poincare, Les Methodes Nouvelles de la Micanique Celeste, 3 Vols. (Gauthier-Villars, Paris, 1892-1899, reprinted by Dover, New York, 1957). 5. D. Schmidt, The family of direct periodic orbits of the first kind in the restricted problem of three bodies, Ordinary Differential Equations (Aca demic Press, 1972). 6. V. Szebehely, Theory of orbits (Academic Press, New York, 1967).
T H E P H A S E SPACE OF F I N I T E S Y S T E M S
KURT BERNARDO WOLF Centro de Ciencias Fisicas Universidad National Autdnoma de Mexico Apartado Postal 48-3, 62251 Cuernavaca, Morelos E-mail: [email protected] NATIG M. ATAKISHIYEV Institute de Matemdticas Universidad National Autdnoma de Mexico Av. Universidad s/n, Col. Chamilpa, 62191 Cuernavaca, Morelos E-mail: [email protected] SERGEY M. CHUMAKOV Departamento de Fisica, Universidad de Guadalajara, Mexico Czda. M. Garcia Barragdn y Olimpica, Sector Reforma, 44281 Guadalajara, Jalisco E-mail: [email protected] The Newton equation of a shallow optical waveguide (a harmonic oscillator) is compatible with the algebraic structure of SU(2). Wavefunctions satisfy orthog onality relations with respect to a discrete measure of 2( + 1 points, of position, momentum, or mode number. The Wigner function for SU(2) in the representation t devolves the classical 'c-number' variables of position, momentum, and energy, picturing the state —a (21 + l)-dimensional vector of sensor data— on the surface of a sphere of radius t+ \- The three coordinates are canonically conjugate in this SU(2) sense. Linear SU(2)-optical systems rotate the sphere rigidly, while others deform it nonlinearly.
1
I n t r o d u c t i o n : discrete a n d finite s y s t e m s
Parallel processing of finite data sets by optical means can achieved doping a strip on a transparent substratum with channels that are planar multimodal waveguides. If a finite set of coherent light-emitting devices modulate a com mon field by factors {/m}m=o, propagation along the waveguide to an array of field sensors will correspond to the fractional Fourier transformation on this data set. This is because the waveguide is the (first-order) optical analogue of the quantum harmonic oscillator, which is well known to rotate phase space. When T is the axial oscillation wavelength of the guide, the Fourier transform occurs at | r , inverted imaging occurs at | r , and the inverse Fourier transform for \T = —\T. These considerations led us in Ref. l to propose a Fourier-
349
350
Kravchuk transform based on an oscillator model 2 with an su(2) Lie algebra of difference operators; the Kravchuk 3 functions were further investigated in Ref. 4 . Since the SU(2) group is compact, the system is discrete and finite. The quantum harmonic oscillator is a very well-known Hamiltonian sys tem (indeed!); the phase space representation of its wavefunctions can be clearly depicted with the Wigner quasiprobability distribution function 5 . One of us had proposed a generalized definition of the Wigner function on the full Heisenberg-Weyl group for polychromatic optics 6 , and the group SU(2) became a natural testing ground for the proposed generalization 7 . It has also provided a phase space picture for discrete, finite systems. In Sec tion 2, we base our construction on the oscillator Newton equation; perhaps surprisingly, several algebraic structures are consistent with it 8 , among them the finite oscillator model of the shallow waveguide. In Section 3 we turn the finite oscillator into a Hamiltonian system through identifying the dynamical variables with the generators of SU(2) in any fixed (and well-known) finite-dimensional irreducible representation space, and in Section 4 we characterize the eigenfunction basis of the Hamil tonian through Kravchuk functions. In Section 5 we present a digest of the generalized Wigner function for SU(2) introduced in Ref. 7 , which gives opera tional meaning to a so-defined 5ft3 'meta-phase' space, two of whose 'c-number' coordinates are those of 'ordinary' phase space, and the third is the mode number of the oscillator. Section 6 recounts the reduction of this 5R3 function to a function on the sphere; this sphere we regard as a proper phase space for finite systems. On it we can plot the face of a finite signal and follow its evolution along a waveguide, and through prisms and lenses of various kinds. The concluding Section 7 condenses the authors' present lines of research on SU(2) optical systems. 2
Newton's equation for the finite waveguide model
In the classical one-dimensional harmonic oscillator potential V(q) = \q2, a particle of unit mass obeys Newton's equation q = —q (the dots are time derivatives); geometric optics uses the same equation for the transverse ray co ordinate across a waveguide, and where the evolution parameter is the length along the optical axis. As does quantum mechanics, paraxial wave optics assumes that there exists an operator Q whose eigenvalues q represent position, and replaces time derivatives by Lie commutators with a Hamiltonian evolution operator H (times i). With these assumptions, Newton's equation becomes the Lie-
351 Newton equation [H,[H,Q]) = Q.
(1)
The commutator of the Hamiltonian and position operators is by definition (—i times) the momentum operator (whose eigenvalues provide the momen tum coordinate), and so the Lie-Newton equation (1) becomes the simultane ous pair of Hamilton equations [H,Q] = -iP,
[H,P} = iQ.
(2)
So far, we have not specified what [P, Q] is. And indeed, more than one algebraic structure is consistent with the Newton equation (1). To close into a Lie algebra, the three operators Q, P and H must satisfy the Jacobi identity, from which it follows that [H,[P,Q]] = 0
=»
[P,Q] = if(H),
(3)
allowing any function f of H (including constants and/or central operators). The requirements that [P, Q] be within the algebra and also preserved under Hamiltonian evolution, are satisfied by the following two distinct Lie algebras: oscillator algebra [P, Q) = iA, SU(2) algebra [P, Q] = iH,
Eq. (2) and [A, Q) = [A, P] = [A, H] = 0, [Q, H] = iP, [H, P] = iQ.
Also, two distinct q-algebras are consistent with the oscillator Newton equa tion 8 , but we shall not address these here. 3
The finite oscillator on the SU(2) group
In the familiar oscillator algebra, Q and P generate the Heisenberg-Weyl subalgebra, with A; their spectra are 3? and serve as classical phase space coordinates. The eigenvalue of A labels irreducible representations: ft for quantum mechanics, X/2n =/= 0 for (polychromatic) paraxial optics. In the new finite oscillator, the algebra is su(2). The spectra of Q, P and H are discrete (the integers and half-integers) and, within an irreducible representation of spin £, they are finite: m = —£,—£+ 1,...,£. This we interpret as the (normalized) positions qm of 2£ + 1 equidistant sensors in the wavefield 7 . The set of measured values is f = {fm \ m — -£, -£ + 1 , . . . ,£}, fm = f(qm) & C, and form a complex vector in the $t2e+1 space of signals. We are in effect representing the generators of SU(2) by means of difference operators 2 ' 7 ; denoting by q € 3? the (continuous) coordinate and using the
352
unit-shift operators e±a"f(q) = f(q ± 1), we can write position Ji = Q = q-, momentum J2 — -P = -i^[ae(-q)
(4) dq
e
dq
- a((q) e~ ],
atti) = ^/{e + q){£-q + l), mode J3 = H-(l+$)l = -\[ctt(q)e-d<
+at(-q)e9*].
(5) (6) (7)
These operators satisfy the commutation relations of the Lie algebra su(2): [Ji, JI|(/|J3l/)|2. 4
(8)
Kravchuk's 'discrete' wavefunctions
The solutions to the eigenvalue equation of the finite oscillator or waveguide (7) with 2 ^ + 1 points and levels, H(q)4>n{q,2l) = {n + \)4>n(q,2Z),
n = 0 , 1 , . . . ,2*.
(9)
are the normal modes of the wavefield in the finite waveguide. The wavefunc tions satisfy the discrete orthogonality and completeness relations on this set of 2£ + 1 points qj, £M«*>2*)4>n{qj,2t) = «„.,„,
qj =j-e,
(10)
j = 0,1,...,21
(11)
j=o it
2 > „ t e , 2 « ) 0 „ t o ' , 2 O = 6j'j,
They are given in terms of the symmetric Kravchuk polynomials kn(q, 2£) of degrees n in q 3 , and as usual normalized and multiplied by the root of the
353
binomial distribution:
*.(..«) - fm+1, v) Jnt+ffg - ;j| g + I .
W h a t is phase space? — T h e generalized W i g n e r function
Given a function ip{q), q € Sft, its (traditional 5 ) Wigner function is a sesquilinear functional of ip, and a function of phase space (q,p) G 3?2, of the form W+(q,p) = ^jdxxP{q-\xYe-ixi,xl>{q+\x).
(13)
The plots of this function show intuitively the one-dimensional objects of the theory (particles, coherent states, light beams) as bumps and oscillations of a two-dimensional surface; its level curves draw the partitura of the signal on a pentagram of canonically conjugate variables. The Wigner function is not strictly a probability distribution because in small regions it can have negative values; but there is agreement that its properties merit the place it has as one of the foundations of quantum mechanics, on par with the Heisenberg and Schrodinger interpretations. Equation (13) can be seen to define phase space as the domain of the Wigner function, which here is associated to the Heisenberg-Weyl group of translations of phase space in position and momentum (multiplication by linear phases). To define a 'phase space' associated to the finite oscillator built on SU(2), we have generalized the definition of the Wigner function (13) to d-parameter
354
Lie groups Q with generators J = {«/fc}fc=n whose elements (in polar coordi nates) are g[y] = exp(—iy ■ J), where y ■ J := X2jt=i VkJk, and square brackets indicate that arguments are in this coordinate set. The generalized Wigner function between two wavefunctions
(14)
of the Wigner operator given by W ( x ) = / dg[y} exp[iy ■ (f - / ) ] Jg = f dg[y] exp(ix-y)g[y} Jg
(15)
~ (2n)dY[6(xk
- Jk),
(16)
k
which is a function of £ 6 SR^. The last expression involving Dirac 6's is meant to suggest that, were the Jk some c-numbers, the arguments Xjt of the Wigner operator would have support on that point only; but since the Jk are operators, the xk in (16) serve as the continuous classical coordinates for the d generally non-commuting observables Jk, associated to a set of physical quantities. When
(17) (18)
where dy = dyidy^dy^ is the Cartesian measure in 3R3, sincx = x _ 1 s i n x , and dg is the invariant Haar measure on SU(2). The angle TJ = \y \ is counted modulo 47r and distinguishes between the conjugation classes of the group. The integral over the group is over the manifold of a 3-sphere. Observe that while the range of the polar parameters y is the compact SU(2) manifold, the space x in the argument of the Wigner operator is SR3. The Wigner operator (16) acts linearly on column vectors f G 9J 2 ' +1 (the wavefield values), and can be thus represented by a (2£+ 1) x (2^+ 1) matrix We(x): W(x) f = f
dg\y] exp(ix • y)T3?\jf] f = We(x)
f.
(19)
JsU(2)
This we call the Wigner matrix of spin £, and is essentially the Fourier trans form of the unitary irreducible ('Wigner-D') matrix representations D f [y ] of
355
SU(2) in polar coordinates, We(x)=
[
±sinc±|j7|dyexp(ix.y)D<[y],
(20)
./SU(3)
&W]=
, - 1 i,~, / dxexp{-ix-y)W'(x). 7i-> sine 51 j/1 7*3
(21)
In the 5R2<+1 space of vectors f and g, the SU(2) Wigner function (14) is the sesquilinear form We(f,g\x)
e £
= fW(x)g=
[fm}mW^m.(S)gm.,
(22)
m,m'=—(
The elements of the Wigner matrix W ( x ) — {W^m,(x)}m,m' = —V have been computed in analytic and numerical form. In particular, from the unitarity of the Wigner D-matrices, D < ([y]) t = D*[—y] (and choosing the in tegration range to be invariant under inversions), follows that the Wigner W^-matrices are self-adjoint: W £ ( f )* = We(x). 6
The Wigner function in 3-space
The asignments (4)-(7) establish the correspondence between the 'c-number' arguments of the Wigner function Wt(f\x) and the observables of a wavefield in a shallow waveguide; explicitly,
(
q —p
\
/ position \ = I -momentum I ,
(23)
E - £ - \ ) \ mode / where E is the energy. Because (x\,xi) = (q, —p) are coordinates of ordinary phase space, but there is now one more, X3 = mode number, we may call x e SR3 the meta-phase space of the SU(2) model. For large £ and small £3, the sphere \x\2 = q2+p2 + (E-£-^)2 « £2 can be approximated by its osculating paraboloid, the classical energy of the (infinite) oscillator, E « (p2 + q2)/2£. We expect that, within this 5R3 meta-phase space for 2^ + 1 points, the values of the Wigner function should peak around radii x = 1^1 w ^- Indeed, when we integrate over the sphere angles, we compute the radial marginal distribution: MLia.(fIX) = I du W<(f |*u) = ||f||2 Rl{X), Js2
(24)
356
where ||f||2 = (f,f) = X)l/il 2 ' s t n e squared norm of the signal, and the dependence on the radius is contained entirely in the factor Re(x)> which is expressible as x _ 1 times a sum of four sine-integral functions, which peaks strongly i n ^ < x < ^ + l - Picturing the Wigner function in 5ft3, we find it concentrated on a sphere, and we can choose between plotting spherical slices of it, or integrating over the radius x to obtain a function over the unit sphere. For the latter we want yet an analytic expression, but the former has been be handled numerically and plotted through level curves on a sphere 7>12>13. The plots for different slices of x (arithmetic or geometric mean, or elsewhere between I and (. + 1) are not too different and stand out starkest for £ + ~. The Wigner function (14)-(15) defined over the group SU(2) has all the desirable properties of the classical Wigner function (13), such as being real for <}> = x, being SU(2)-covariant: We(Be[y) f, T>e[y ] g| x) = We(f, g| R [ - y ] x ) ,
(25)
with the 3 x 3 'rotation' matrices R[y] € SU(2), and with total marginal distribution M'(f,g)=/
d £ W ' ( f , g | x ) = 7r 3 (f,g).
(26)
Thus the Hilbert-Schmidt norm of the SU(2) Wigner operators is w3. Among the many formulae we would like to derive analytically (which are at present computed only numerically) are the marginal distributions on arbitrary planes and axes (through one and two integrations). 7
Outlook
In more mathematical terms, the domain of a Wigner function are HilbertSchmidt operators —above we considered only those operators that are sepa rable into a ket-bra pair of wavefunctions p = \ip)(4>\— and on the co-adjoint orbits of the group (classified by irreducible representations). The multipli cation of the Haar measure (18) by a radial function of the classes r\ = \y\ leaves its covariance property (25) invariant. This freedom allowed in Ref. 14 the definition of the Wigner function on Lie groups whose right- and leftinvariant Haar measures are distinct, as they are in the two-parameter affine group. There, the Wigner function and the wavelet transform are essentially the same. The Wigner functions for the compact and noncompact one-parameter groups (and all abelian groups) are given in Ref. 15 , and the non-abelian two-parameter group in Ref. 14 . Among the three-parameter Lie groups, the
357
Heisenberg Weyl 6 , SU(2) 7 , and the Euclidean IS0(2) 15 groups have been visited. In the latter, the Hilbert spaces of wavefunctions can be the oscilla tory Helmholtz wavefields in two dimensions, or the discrete, infinite sets of data in a discrete version of two-dimensional optics. A challenge is contained in Sp(2,3?) = SU(1,1) = S1(2,5R), because the group is not of exponential type and the structure of its irreducible representations is arduous; yet it is the pro totypical dynamical algebra for several one-dimensional mechanical systems and models of half-infinite discrete data 16 . Also the Poincare and conformal groups are of interest. The partitura of a signal on phase space is its Wigner function. When we place in the waveguide a flat prism of angle a, 'optical rays' are deflected by the group element exp(iuaQ) {v is a fixed coefficient). With free flight and prisms we can generate all SU(2) transformations of data sets that rotate rigidly the two-sphere of phase space. Data sets derived from an original signal by SU(2)-optical elements can be recognized for having the same, albeit rotated, Wigner function. On the other hand lenses, which act on the wavefield by exp(ry<22), are in the enveloping algebra of SU(2); their action results in deformations —aberrations— of the Wigner function on the sphere. In Ref. 12 we studied in this way the quadratic nonlinearity of the optical Kerr medium Hamiltonian HK„t = H0,c + 0H^c. The Wigner function, being sesquilinear in two wavefields (one is the standard beam and the other is the object beam) serves for holographic encoding and decoding 17 of finite signals. 8
Acknowledgements
This work was performed under the support of Project DGAPA-UNAM 104198 Optica Matemdtica. At various stages in the program we have ben efited from the collaboration, computer and numerical support of Guillermo Krotzseh, Ana Leonor Rivera and Luis Edgar Vicent. References 1. N.M. Atakishiyev and K.B. Wolf, Fractional Fourier-Kravchuk transform. J. Opt. Soc. Am. 14, 1467-1477 (1997). 2. N.M. Atakishiyev and S.K. Suslov, Difference analogs of the harmonic oscillator. Theor. Math. Phys. 85, 1055-1062 (1991). 3. M. Krawtchouk, Sur une generalization des polinomes d'Hermite. C. R. Acad. Sci. Paris 189, 620-622 (1929); A. Erdelyi, W. Magnus, F. Oberhettinger, and F.G. Tricomi, Higher Transcendental Functions, Vol. 2 (McGraw-Hill, New York, 1953).
358
4. N.M. Atakishiyev and K.B. Wolf, Approximation on a finite set of points through Kravchuk functions. Rev. Mex. Fis. 40, 366-377 (1994). 5. E. Wigner, On the quantum correction for thermodynamic equilibrium, Phys. Rev. 40, 749-759 (1932); M. Hillery, R.F. O'Connell, M.O. Scully, and E.P. Wigner, Distribution functions in physics: fundamentals, Phys. Reports 106, 121-167 (1984); H.-W. Lee, Theory and application of the quantum phase-space distribution functions, Phys. Reports 259, 147-211 (1995). 6. K.B. Wolf, Wigner distribution function for paraxial polychromatic op tics. Opt. Commun. 132, 343-352 (1996). 7. N.M. Atakishiyev, S.M. Chumakov and K.B. Wolf, Wigner distribution function for finite systems, J. Math. Phys. 39, 6247-6261 (1998). 8. M. Ank, N.M. Atakishiyev and K.B. Wolf, Quantum algebraic structures compatible with the harmonic oscillator Newton equation, J. Phys. A 32, L371-L376 (1999). 9. P. Feinsilver and R. Schott, Operator calculus approach to orthogonal polynomial expansions, J. Comp. Appl. Math. 66, 185-199 (1996). 10. N.M. Atakishiyev, L.E. Vicent and K.B. Wolf, Continuous vs. discrete fractional Fourier transforms. J. Comp. Appl. Math. 107, 73-95 (1999). 11. N.Ja. Vilenkin and A.U. Klimyk, Representation of Lie Group and Spe cial Functions Vol.1, (Kluver Academic Publ., Dordrecht, 1991). G. Gasper and M. Rahman, Basic Hypergeometric Series (Cambridge Uni versity Press, Cambridge, UK, 1990); A.F. Nikiforov, S.K. Suslov and V.B. Uvarov, Classical Orthogonal Polynomials of a Discrete Variable (Springer-Verlag, Berlin, 1991). 12. S.M. Chumakov, A. Frank and K.B. Wolf, Finite Kerr medium: Schrodinger cats and Wigner functions on the sphere. Phys. Rev. A 60, 1817-1823 (1999). 13. K.B. Wolf, N.M. Atakishiyev and S.M. Chumakov, Wigner distribution function for finite signals. SPIE Proceedings Vol. 3076 (Photonic Quan tum Computing, Orlando, Fla., 23-24 April, 1997). The International Society for Optical Engineering, 1997, pp. 196-206; N.M. Atakishiyev, S.M. Chumakov, L.M. Nieto and K.B. Wolf, Wigner function on groups for various optical models. Proceedings of the Fifth International Confer ence on Squeezed States and Uncertainty Relations (Balatonfured, Hun gary, May 27-31, 1997). Ed. by D. Han, J. Janszky, Y.S. Kim and V.I. Man'ko, NASA/CP-1998-206855 (May 1998), pp. 301-308; K.B. Wolf, N.M. Atakishiyev and S.M. Chumakov, Wigner distribution function for some dynamical models. Proceedings of the V Wigner Symposium (Vi enna, Aug. 25-29, 1997). World Scientific Publ. Co. (1998), pp. 370-372.
359
14. S.T. Ali, N.M. Atakishiyev, S.M. Chumakov and K.B. Wolf, The Wigner function for general Lie groups and the wavelet transform. Ann. Inst. Henri Poincare / Physique Theorique , Submitted (2000). 15. L.M. Nieto, N.M. Atakishiyev, S.M. Chumakov and K.B. Wolf, Wigner distribution function for Euclidean systems. J. Phys. A 3 1 , 3875-3895 (1998). 16. N.M. Atakishiyev, E.I. Jafarov, Sh.M. Nagiyev and K.B. Wolf, Meixner oscillators. Rev. Mex. Fis. 44, 235-244 (1998). 17. K.B. Wolf and A.L. Rivera, Holographic information in the Wigner func tion. Opt. Commun. 144, 36-42 (1997).