You are on page 1of 4

Chaos and Gravitational Waves

Neil J. Cornish
Department of Physics, Montana State University, Bozeman, MT 59717

The gravitational waveforms of a chaotic system will exhibit sensitive dependence on initial con-
ditions. The waveforms of nearby orbits decohere on a timescale fixed by the largest Lyapunov
exponent of the orbit. The loss of coherence has important observational consequences for systems
where the Lyapunov timescale is short compared to the chirp timescale. Detectors that rely on
matched filtering techniques will be unable to detect gravitational waves from these systems.

The evolution of two compact objects under their mu- The phase space coordinates we have in mind are the
tual gravitational attraction is a major unsolved problem position and momentum of the masses. These coordi-
in general relativity. A vast array of analytic and numeri- nates are well defined in the test particle, post-Newtonian
cal approximations have been used to attack the problem, and post-Minkowski approximations, but their meaning
arXiv:gr-qc/0106062v1 20 Jun 2001

but much remains to be understood. Amongst the many is less clear in the full theory. Indeed, the full evolution
outstanding questions is the degree to which the dynam- equations are a set of coupled, non-linear partial differ-
ics may be chaotic[1, 2, 3, 4] or effectively chaotic[5, 6]. ential equations, so the real issue is gravitational turbu-
For example, when one or more of the masses is spin- lence, not chaos. We will avoid this complication by as-
ning it has been shown in the test-particle[2] and post- suming that the evolution equations can be approximated
Newtonian[3] limits that certain classes of orbits are by some set of non-linear ordinary differential equations
chaotic, at least when radiation reaction is turned off[4]. for the phase space variables Xi (t). Here t is taken to be
Here we consider the physical and observational con- the time measured by a distant observer, but the choice
sequences wrought by chaotic behaviour in the orbital is unimportant. The evolution equations can be written:
dynamics. A key feature of chaotic systems is their sen-
sitive dependence on initial conditions - the so-called dXi
= Hi (Xj ) . (1)
“butterfly effect”. It is intuitively obvious that sensi- dt
tive dependence in the dynamics will be reflected in the
Linearizing about a solution to this equation yields an
gravitational waveforms, but the precise connection has
equation for the perturbation:
not been established until now. We find that the wave-
forms of nearby orbits decohere on a timescale fixed by d δXi (t)
the largest Lyapunov exponent. The observational conse- = Kij (t) δXj (t) , (2)
dt
quences of this result depend on two timescales: the chirp
timescale Tc ; and the Lyapunov timescale Tλ . A third where
timescale of interest is the gravitational wave period Tw .
The wave period and chirp timescale for a system with ∂Hi
Kij (t) = (3)
reduced mass µ, total mass M and orbital separation R ∂Xj Xi (t)
are given by
 3/2  4 is the infinitesimal evolution matrix. The solution to the
R 5 M2 R linearized equations of motion can be expressed in terms
Tw ≃ πM and Tc ≃ .
M 256 µ M of the evolution matrix Lij (t):
In the limit that one of the masses is very much smaller δXi (t) = Lij (t) δXj (0) . (4)
than its companion we have µ ≪ M and Tc ≫ Tw . The
Lyapunov timescale has to be calculated on a case-by- The eigenvalues and eigenvectors of Kij (t) contain infor-
case basis, but values as short as Tλ ∼ Tw are possi- mation about the local stability of the phase space trajec-
ble for very unstable orbits. Orbits whose Lyapunov tory. In a fixed coordinate system the eigenvectors may
timescale is short compared to the chirp timescale will vary with time. To avoid this, a non-constant basis can
produces highly unpredictable waveforms. The number be defined that stays aligned with the eigenvectors[8].
of templates required to detect these waveforms is expo- This is done by parallel transporting the basis vectors
nentially large, making them impossible to detect[3, 7] along the trajectory Xi (t). Barring degeneracies, the in-
with the Laser Interferometer Gravitational Observatory finitesimal evolution matrix is diagonal in the eigenvector
(LIGO). basis: Kij (t) = δij λi (t) (no summation on the i), where
The connection between orbital instability and wave- the λi (t) are the eigenvalues of Kij (t). The evolution
form decoherence is easily established. It relies on the matrix is also diagonal in the eigenvalue basis and has
fact that the gravitational waveform hµν (t) can be ex- components
pressed in terms of the phase space coordinates of the Z t 
system, so the divergence of nearby trajectories in phase ′ ′
Lij (t) = δij exp λi (t )dt . (5)
space is reflected in the divergence of nearby waveforms. 0
2

The principal Lyapunov exponent for the trajectory The quantity


Xi (t) is defined:
δΦ(t)
1 λh = lim lim ln (13)
ln L∗ji (t)Lij (t) . δΦ(0)

λ = lim (6) t→∞ δΦ(0)→0
t→∞ 2t

The multiplicative ergodic theorems establish the exis- is equal to the principal Lyapunov exponent of the tra-
tence of this limit for a large range of situations. Tra- jectory. In other words, the waveforms of nearby trajec-
jectories with positive Lyapunov exponents are unstable tories will decohere exponentially in time, on a timescale
against small perturbations. When the unstable orbits set by Tλ = λ−1 .
have non-zero measure in phase space, the system is said A couple of comments are in order. Firstly, the limit
to be chaotic. A positive principal Lyapunov exponent t → ∞ used in equations (6) and (13) will return
will dominate the long term dynamics. Setting λ1 to be λ = λh = 0 if we include the effects of radiation re-
the principle eigenvalue, the asymptotic form of the evo- action. This reflects the fact that the dynamics is not
lution matrix can be written: chaotic in a formal sense[4]. The limit t → ∞ has to
  be replaced by the limit t → T , where Tλ ≪ T ≪ Tc .
f (t) 0 0 · · · This is done by replacing the Lyapunov exponent by a
 0 0 0 ···  finite-time or local Lyapunov exponent[11], and making a
Lij (t) ≃ eλ t  0 0 0 · · ·  (7)
 
similar change to (13). Clearly, the whole notion of wave-
.. .. .. . .
 
form decoherence only makes sense if the Lyapunov time
. . . .
scale Tλ is very much shorter than the chirp timescale
Tc . It remains to be seen if this condition is satisfied
where |f (t)| = 1. If X1 (t) is real, then f (t) = 1.
for any realistic systems. The second comment concerns
A gravitational wave can be decomposed into two po-
the gauge non-invariance of the Lyapunov exponents and
larizations with amplitudes h+ (t) and h× (t). A gravi-
the choice of time variable. Lyapunov exponents gained a
tational wave detector responds to the wave according
bad reputation when they were used to study the Bianchi
to
IX dynamics. Different time slices yielded principal Lya-
s(t) = h+ (t)F+ + h× (t)F× , (8) punov exponents that were positive or zero, making it
difficult to decide if the dynamics was chaotic. However,
where F+ and F× are the antenna patterns for each po- this difficulty can be avoided by making relative, rather
larization. The antenna patterns for LIGO and LISA can than absolute comparisons. For example, the ratio Tλ /Tc
be found in Refs.[9] and [10] respectively. The gravita- is the same in any coordinate system, so the choice of
tional wave amplitude h(t) measured at the detector can time variable is irrelevant. The confusion surrounding
be expressed in terms of the phase space variables Xi (t). the Bianchi IX dynamics could have been avoided if the
Nearby orbits have gravitational waveforms that differ by Lyapunov timescale had been compared to the average
time between bounces[12].
∂h(Xm ) We can illustrate the connection between orbital in-
δh(t) = (t) δXi (t)
∂Xi stability and waveform decoherence with a simple exam-
ple. Consider the Lagrangian for a test particle in the
∂h(Xm ) Schwarzschild spacetime (in units where G = c = M = 1)
= (t) Lij (t) δXj (0) . (9)
∂Xi
 2  2
Employing an eigenvector basis and using the asymptotic 1 r − 2 dt r dr
L = +
form (7) for the evolution matrix yields 2 r dλ r − 2 dλ
 2  2 !
δh(t) ≃ eλ t g(t) , (10) 2 dθ 2 2 dφ
+r + r sin θ . (14)
dλ dλ
where
∂h(Xm ) The system is completely integrable (non-chaotic) as it
g(t) = (t) f (t) δX1 (0) (11) has four generalized coordinates t, r, θ, φ and four con-
∂X1
stants of motion, the mass µ, energy E, z-component
is an oscillatory factor with bounded amplitude. The of angular momentum Lz and total angular momentum
divergence of the waveforms can be better expressed in squared L2 . Despite the lack of chaos, the dynamics does
terms of their relative phase. Writing the amplitude of admit isolated unstable orbits that serve to illustrate the
the reference trajectory as h(Φ), and defining Φ0 (t) to be connection between orbital instability and waveform de-
the phase for which h(Φ0 ) = 0, we find the relative phase coherence. Restricting our attention to equatorial orbits
is given by with fixed energy E allows us to make the replacement
δh(Φ0 (t)) d r−2 d
δΦ(t) = . (12) = (15)
h′ (Φ0 (t)) dλ rE dt
3

in the equations of motion: In terms of the transformed phase space coordinates the
infinitesimal evolution matrix has components
E3r Ep2r Ep2φ
ṗr = − − +  1
√ 0 0 0

(r − 2)3 r(r − 2) r2 (r − 2) 2 2
ṗφ = 0
 
 
 0 − √ 1
0 0
ṙ = Epr Kij = 
 2 2 
, (22)
Epφ 
 0

φ̇ = . (16) 0 0 1 
r(r − 2)
 
 
Our reference trajectories are circular orbits with 0 0 0 0
and the evolution matrix has components
pr = 0
 √ 
pφ = L z = L et/2 2 0 0 0
1 p  
r = r± = L(L ± L2 − 12)  √ 
2
 0
 e−t/2 2 0 0
θ = ωt . (17) Lij = 
.
 (23)
 0 0 1 t
 
Perturbing the equations of motion about the reference  
trajectory yields the infinitesimal evolution matrix 0 0 0 1
2E 3 (r+1) EL2 (3r−4)
 
0 r22EL
(r−2) (r−2)4 − r 3 (r−2)2 0 Taking the limit defined in √ (6) yields the principal Lya-
 0

0 0 0
 punov exponent λ = 1/2 2. √ The Lyapunov timescale
Kij =   . (18)
E 0 0 0 for this orbit, Tλ = λ−1 = 2 2, is actually shorter than
E
0 r(r−2) − 2EL(r−1) 0 the gravitational wave period Tw = 2π.
r 2 (r−2)2
Next we need an expression for the gravitational wave
The eigenvalues are: amplitude in terms of the phase space variables. The cor-
p rect approach would be to solve the Teukolsky equation
2E 2 r3 (r + 1) − L2 (3r − 4)(r − 2)2 using the method described by Hughes[5], but no ana-
λ1 = −λ2 = lytic solutions exist in the strong field limit. Instead we
r2 (r − 2)2
will use the quadrupole approximation, knowing full well
λ3 = λ4 = 0 . (19)
that it is being applied outside of its domain of validity.
The well known innermost stable circular orbit at r = 6 Since our goal is to illustrate a general phenomenon - not
√ p
to produce accurate templates for LIGO data analysis - a
has L = 2 3, E = 8/9 and λ1 = λ2 = 0. Circular
orbits with larger values of L come in two varieties, those qualitative description of the waveform’s dependence on
with r = r+ and those with r = r− , e.g. setting E = 1 the phase space variables will suffice.
and L = 4 yields orbits with r+ = 12√and r− = 4. The The gravitational wave amplitude can be written as
outer orbit is stable, λ1 = −λ2 = i √ 6/108, while the hµν (t) = h+ (t)e+ × ×
µν + h (t)eµν , where the polarization
inner orbit is unstable, λ1 = −λ2 = 1/2 2. The unstable tensors have non-zero components
orbit at r = 4 has eigenvectors e+ e+
xx = 1 yy = −1
e× e×
 
1 2 1 xy = 1 yx = 1 . (24)
e1 = √ , 0, , − √
3 2 3 2
  A detector situated at r = R on the z axis will encounter
1 2 1 a wave with plus polarization given by
e2 = √ , 0, − , − √
3 2 3 2
2µE 2 2pφ pr (r − 3) E 2 r2
 
e3 = e4 = (1, 0, 0, 0) . (20) +
h (t) = − sin 2φ +
R (r − 2)2 (r − 2)3
Since the zero eigenvalues have degenerate eigenvectors, 2
pφ (r + 2) p2r (r − 3)
! #
Kij can not be diagonalized, but it is enough to put Kij + − cos 2φ , (25)
into Jordan normal form with the transformation matrix r(r − 2)2 r−2
 √ √
and similarly for the cross polarization. Specializing to

7 2
24 − 7242 0 0
  the orbit with E = 1, L = 4 and r = 4 we have
 
 0 0 0 1  8µ
h+ = − cos 2φ
 
Pij =  . (21)
 7 7
 R

 6 6 0 −2 


√ √
 ∂h+ 2µ
= − sin 2φ
−782 7 2
8
7
8 0 ∂pr R
4

so that
∂h+ 3µ
= − cos 2φ √
∂pφ R δΦ(t) ≃ et/2 2
δΦ(0) . (28)
+
∂h 17µ
= cos 2φ As promised, the phase difference grows exponentially on
∂r 2R √
a timescale equal to Tλ = 2 2.
∂h+ 16µ
= sin 2φ . (26)
∂φ R

Putting everything together in equation (9) yields Acknowledgements


µ t/2√2 √
δh+ (t) ≃ e (119 cos 2φ − 175 2 sin 2φ)
84R √ I would like to that Scott Hughes and Janna Levin for
×(3 δr(0) + 6 δpφ (0) + 6 2 δpr (0)) , (27) several interesting discussions.

[1] N. J. Cornish & N. E. Frankel, Phys. Rev. D56, 1903 [8] C. P. Dettmann & G. P. Morriss, Phys. Rev. E53, R5545
(1997). (1996).
[2] S. Suzuki & K. Maeda, Phys. Rev. D55, 4848 (1997); [9] E. E. Flanagan, Phys. Rev. D48, 2389 (1993).
Phys. Rev. D61, 024005 (2000). [10] N. J. Cornish & S. L. Larson, Preprint gr-qc/0103075,
[3] J. J. Levin, Phys. Rev. Lett. 84, 3515 (2000). (2001).
[4] N. J. Cornish, Phys. Rev. Lett. 85, 3980 (2000). [11] P. Grassberger, R. Badii & A. Politi, J. Stat. Phys. 51,
[5] S. A. Hughes, Phys. Rev. D61, 084004 (2000). 135 (1988); H. E. Kandrup, B. L. Eckstein & B. O.
[6] B. Schutz, Talk given at the 3rd International LISA sym- Bradley, Astron. Astrophys. 320, 65 (1997).
posium (2000). [12] N. J. Cornish & J. J. Levin, Phys. Rev. D55, 7489 (1997).
[7] N. J. Cornish, talk given at the Center for Gravitational
Physics and Geometry, Penn. State, October 20th (1997).

You might also like