Добавил:
Upload Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:
FirstCourseGR.pdf
Скачиваний:
38
Добавлен:
07.03.2016
Размер:
7.41 Mб
Скачать

227

9.3 The generation of gravitational waves

Communicating with 1 W lasers, LISA achieves a remarkable sensitivity, and will be able to see the strongest sources in its band anywhere in the universe. We will return to the kinds of sources that might be detected by LISA and the ground-based detectors after studying the way that gravitational waves are generated in general relativity.

9.3 T h e g e n e ra t i o n o f g ra v i t a t i o n a l wa v e s

Simple estimates

It is easy to see that the amplitude of any gravitational waves incident on Earth should be small. A ‘strong’ gravitational wave would have hμν = 0(1), and we should expect amplitudes like this only very near a highly relativistic source, where the Newtonian potential (if it had any meaning) would be of order one. For a source of mass M, this would be at distances of order M from it. As with all radiation fields, the amplitude of the gravitational waves falls off as r1 far from the source. (Readers who are not familiar with solutions of the wave equation will find demonstrations of this in the next sections.) So if Earth is a distance R from a source of mass M, the largest amplitude waves we should expect are of order M/R. For the formation of a 10 M black hole in a supernova explosion in a nearby galaxy 1023 m away, this is about 1017. This is in fact an upper limit in this case, and less-violent events will lead to very much smaller amplitudes.

Slow motion wave generation

Our object is to solve Eq. (8.42):

 

¯

μν = −

 

 

t2 +

μν.

 

 

2

2

h

16π T

 

(9.64)

 

 

 

 

We will find the exact solution in a later section. Here we will make some simplifying – but realistic – assumptions. We assume that the time-dependent part of Tμν is in sinusoidal oscillation with frequency , i.e. that it is the real part of

Tμν = Sμν (xi) ei t,

(9.65)

and that the region of space in which Sμν =0 is small compared with 2π/ , the wavelength of a gravitational wave of frequency . The first assumption is not much of a restriction, since a general time dependence can be reduced to a sum over sinusoidal motions by Fourier analysis. Besides, many interesting astrophysical sources are roughly periodic: pulsating stars, pulsars, binary systems. The second assumption is called the slow-motion assumption, since it implies that the typical velocity inside the source region, which is times the size of that region, should be much less than one. All but the most powerful sources of gravitational waves probably satisfy this condition.

228

Gravitational radiation

¯

Let us look for a solution for hμν of the form

h

B

μν

(xi) ei t.

(9.66)

¯μν =

 

 

 

(In the end we must take the real part of this for our answer.) Putting this and Eq. (9.65) into Eq. (9.64) gives

( 2 + 2)Bμν = −16π Sμν.

(9.67)

¯

It is important to bear in mind as we proceed that the indices on hμν in Eq. (9.64) play

¯

almost no role. We shall regard each component hμν as simply a function on Minkowski space, satisfying the wave equation. All our steps will be the same as for solving the scalar equation (2/∂t2 + 2)f = g, until we come to Eq. (9.75).

Outside the source (i.e. where Sμν = 0) we want a solution Bμν of Eq. (9.67) that represents outgoing radiation far away; and of all such possible solutions we want the one that dominates in the slow-motion limit. Let us define r to be the spherical polar radial coordinate, where the origin is chosen inside the source. We show in Exer. 29, § 9.7 that the solution we seek is spherical outside the source,

Bμν =

Aμν

ei r +

Zμν

ei r,

(9.68)

r

r

where Aμν and Zμν are constants. The term in ei r represents a wave traveling toward the origin r = 0 (called an ingoing wave), while the other term is outgoing (see Exer. 28, § 9.7). We want waves emitted by the source, so we choose Zμν = 0.

Our problem is to determine Aμν in terms of the source. Here we make our approxima-

tion that the source is nonzero only inside a sphere of radius ε

2π/ . Let us integrate

Eq. (9.67) over the interior of this sphere. One term we get is

 

'

2Bμν d3x 2|Bμν |max4π ε3/3,

(9.69)

where |Bμν |max is the maximum value Bμν takes inside the source. We will see that this term is negligible. The other term from integrating the left-hand side of Eq. (9.67) is

'

(

 

2Bμν d3x =

n · Bμν dS,

(9.70)

by Gauss’ theorem. But the surface integral is outside the source, where only the spherical part of Bμν given by Eq. (9.68) survives:

(

 

d

r=ε

 

 

n · Bμν dS = 4π ε2

dr Bμν

≈ −4π Aμν,

(9.71)

again with the approximation ε

2π/ . Finally, we define the integral of the right-hand

side of Eq. (9.67) to be

Jμν = '

 

 

 

Sμν d3x.

(9.72)

Combining these results in the limit ε 0 gives

 

Aμν =

4Jμν ,

(9.73)

¯

μν =

μν

ei (rt)/r.

 

h

 

4J

(9.74)

229

 

9.3 The generation of gravitational waves

 

 

 

 

 

 

 

 

 

 

 

 

These are the expressions for the gravitational waves generated by the source, neglecting

 

 

terms of order r2 and any r1 terms that are higher order in ε.

 

 

 

 

 

 

 

 

h

μν }

are

 

 

It is possible to simplify these considerably. Here we begin to use the fact that {¯

 

 

components of a single tensor, not the unrelated functions we have solved for in Eq. (9.74).

 

From Eq. (9.72) we learn

 

 

 

 

 

 

 

 

Jμν ei t = '

Tμν d3x,

(9.75)

 

which has as one consequence:

 

 

 

 

 

 

 

 

i Jμ0 ei t = '

Tμ0 ,0 d3x.

(9.76)

 

Now, the conservation law for Tμν ,

 

 

 

 

 

 

 

 

Tμν ,ν = 0,

 

 

(9.77)

 

implies that

 

 

 

 

 

 

 

 

Tμ0

,0 = −Tμk ,k

 

(9.78)

 

and hence that

 

d3x = (

 

 

 

 

 

i Jμ0 ei t = '

Tμk ,k

Tμk nk dS,

(9.79)

the last step being the application of Gauss’ theorem to any volume completely containing the source. This means that Tμν = 0 on the surface bounding this volume, so that the right-hand side of Eq. (9.79) vanishes. This means that if =0, we have

Jμ0

= 0,

¯

=

0.

(9.80)

 

hμ0

 

These conditions basically embody the laws of conservation of total energy and momen-

¯μ0

tum for the oscillating source. The neglected higher-order parts of h are gauge terms (Exer. 32, § 9.7).

The expression for Jij can also be rewritten in an instructive way by using the result of Exer. 23, § 4.10.

dt2

'

T00xlxm d3x = 2

'

Tlm d3x.

(9.81)

d2

 

 

 

 

 

For a source in slow motion, we have seen in Ch. 7 that T00 ρ, the Newtonian mass density. It follows that the integral on the left-hand side of Eq. (9.81) is what is often referred to as the quadrupole moment tensor of the mass distribution,

Ilm := '

T00xlxm d3x

(9.82a)

= Dlm ei t

(9.82b)

(Conventions for defining the quadrupole moment vary from one text to another. We follow Misner et al. (1973).) In terms of this we have

h

2 2D

jk

ei (rt)/r.

(9.83)

¯jk = −

 

 

 

It is important to remember that Eq. (9.83) is an approximation which neglects not merely all terms of order r2 but also r1 terms that are not dominant in the slow-motion

230

 

 

 

 

Gravitational radiation

 

 

 

 

 

 

 

 

 

 

h

,k is of higher order, and this guarantees that the gauge

 

 

 

 

 

approximation. In particular, ¯jk

¯

=

0 is satisfied by Eqs. (9.83) and (9.80) at the lowest order in r1 and .

 

condition hμν ,ν

 

Because of Eq. (9.83), this approximation is often called the quadrupole approximation for gravitational radiation.

As for the plane waves we studied earlier, we have here the freedom to make a further restriction of the gauge. The obvious choice is to try to find a TT gauge, transverse to the direction of motion of the wave (the radial direction), which has the unit vector nj = xj/r. Exer. 32, § 9.7, shows that this is possible, so that in the TT gauge we have the simplest form of the wave. If we choose our axes so that at the point where we measure the wave it is traveling in the z direction, then we can supplement Eq. (9.80) by

hTT

=

0,

 

 

 

 

 

 

 

(9.84)

¯zi

 

 

 

 

 

 

 

 

 

hTT

 

hTT

 

 

2

i r

/r,

(9.85)

¯xx

= −¯yy

= −

(I xx I yy) e

 

 

hTT

 

2 2I

xy

ei r

/r,

 

 

 

(9.86)

¯xy

= −

 

 

 

 

 

 

 

where

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

1

l

 

 

(9.87)

 

 

I jk := Ijk 3

δjkIl

 

 

is called the trace-free or reduced quadrupole moment tensor.

Examples

Let us consider the waves emitted by a simple oscillator like the one we used as a detector in § 9.2. If both masses oscillate with angular frequency ω and amplitude A about mean equilibrium positions a distance l0, apart, then, by Exer. 30, § 9.7, the quadrupole tensor has only one nonzero component,

Ixx = m[(x1)2 + (x2)2]

 

= [(21 l0 A cos ω t)2 + ( 21 l0 A cos ω t)2]

 

= const. + mA2 cos 2 ω t + 2 ml0A cos ω t.

(9.88)

Recall that only the sinusoidal part of Ixx should be used in the formulae developed in the previous paragraph. In this case, there are two such pieces, with frequencies ω and 2ω. Since the wave equation Eq. (9.64) is linear, we shall treat each term separately and simply add the results later. The ω term in Ixx is the real part of 2ml0A exp(iωt). The trace-free quadrupole tensor has components

I yy

= I zz

1 Ixx

2 ml0A eiωt,

 

I xx

Ixx

 

31 Ijj

=

32 Ixx =

34 ml0A eiωt,

(9.89)

 

=

= − 3

 

= − 3

 

 

all off-diagonal components vanishing. If we consider the radiation traveling in the z direction, we get, from Eqs. (9.84)–(9.86),

¯xx

= −¯yy = −

0

A eiω(rt)/r,

¯xy

=

 

 

hTT

hTT

2 mω2l

hTT

 

0.

(9.90)

The radiation is linearly polarized, with an orientation such that the ellipse in Fig. 9.1 is aligned with the line joining the two masses. The same is true for the radiation going in the

231

9.3 The generation of gravitational waves

y direction, by symmetry. But for the radiation traveling in the x direction (i.e. along the line joining the masses), we need to make the substitutions z x, x y, y z in Eqs. (9.85)–(9.86), and we find

hTT

= 0.

(9.91)

¯ij

There is no radiation in the x direction. In Exer. 36, § 9.7, we will fill in this radiation pattern by calculating the amount of radiation and its polarization in arbitrary directions.

A similar calculation for the 2ω piece of Ixx gives the same radiation pattern, replacing Eq. (9.90) by

hTT

hTT

= −

4 mω2A2

e2iω(rt)/r,

hTT

=

0.

(9.92)

¯xx

= −¯yy

 

 

¯xy

 

 

The total radiation field is the real part of the sum of Eqs. (9.90) and (9.92),

hTT

= −

[2 m ω2l

A cos ω(r

t)

+

4 m ω2A2 cos 2 ω(r

t)]/r.

(9.93)

¯xx

0

 

 

 

 

 

Let us estimate the radiation from a laboratory-sized generator of this type. If we take m = 103 kg = 7 × 1024 m, l0 = 1 m, A = 104 m, and ω = 104 s1 = 3 × 104 m1,

then the 2ω contribution is negligible and we find that the amplitude is about 1034/r, where r is measured in meters. This shows that laboratory generators are unlikely to produce useful gravitational waves in the near future!

A more interesting example of a gravitational wave source is a binary star system. Strictly speaking, our derivation applies only to sources where motions result from nongravitational forces (this is the content of Eq. (9.77)), but our final result, Eqs. (9.84)– (9.87), makes use only of the motions produced, not of the forces. It is perhaps not so surprising, then, that Eqs. (9.84)–(9.87) are a good first approximation even for systems dominated by Newtonian gravitational forces (see further reading for references). Let us suppose, then, that we have two stars of mass m, idealized as points in circular orbit about one another, separated a distance l0 (i.e. moving on a circle of radius 12 l0). Their orbit equation (gravitational force = ‘centrifugal force’) is

l02

=

 

 

2

 

 

 

=

 

 

 

m2

mω2

 

l0

 

 

ω

 

(2 m/l3)1/2

,

(9.94)

 

 

 

 

 

 

 

 

 

0

 

 

where ω is the angular velocity of the orbit. Then, with an appropriate choice of coordinates, the masses move on the curves

x2

(t) = −x1(t),

y2

(t) = −y1(t),

 

x1

(t) = 21 l0 cos ω t,

y1

(t) = 21 l0 sin ω t,

 

where the subscripts 1 and 2 refer to the respective stars. These equations give

Ixx =

41 ml02 cos 2 ωt + const.,

 

Ixy = −1 ml2 sin 2 ωt.

+

 

Iyy

41 ml02 cos 2 ωt

 

const.,

=

4 0

 

 

 

(9.95)

(9.96)

The only nonvanishing components of the reduced quadrupole tensor are, in complex notation and omitting time-independent terms,

I xy

= 41 iml02

e2iωt.

 

I xx

= −

I yy

=

1 ml2

e2iωt,

 

 

4

0

(9.97)

232

Gravitational radiation

All the radiation comes out (perpendicular to the plane of

with frequency = 2 ω. The radiation along the z direction the orbit) is, by Eqs. (9.84)–(9.86),

¯

hxx

¯

hxy

= −2 i ml02 ω2

e2iω(rt)/r.

 

hyy

= −

2 ml2

ω2

e2iω(rt)/r,

(9.98)

= −¯

 

0

 

 

This is circularly polarized radiation (see Exer. 15, § 9.7). The radiation in the plane of the orbit, say in the x direction, is found in the same manner used to derive Eq. (9.91). This gives

hTT

hTT

=

ml2

ω2

e2iω(rt)/r,

(9.99)

¯yy

= −¯zz

0

 

 

 

all others vanishing. This shows linear polarization aligned with the orbital plane. The antenna pattern and polarization are examined in greater detail in Exer. 38, § 9.7, and the calculation is generalized to unequal masses in elliptical orbits in Exer. 39.

The amplitude of the radiation is of order ml02ω2/r, which, by Eq. (9.94), is of order (mω)2/3m/r. Binary systems are likely to be very important sources of gravitational waves, indeed they may well be the first sources detected. Binary systems containing pulsars have already provided strong indirect evidence for gravitational radiation (Stairs 2003, Lorimer 2008). The first, and still the most important of these systems is the one containing the pulsar PSR B1913+16, the discovery of which by Hulse and Taylor (1975) led to their being awarded the Nobel Prize for Physics in 1993. This system consists of two neutron stars (see Ch. 10) orbiting each other closely. The orbital period, inferred from the Doppler shift of the pulsar’s period, is 7 h 45 min 7 s (27907 s or 8.3721 × 1012 m), and both stars have masses approximately equal to 1.4M (2.07 km) (Taylor and Weisberg 1982). If the system is 8 kpc = 2.4 × 1020 m away, then its radiation will have the approximate amplitude 1023 at Earth. We will calculate the effect of this radiation on the binary orbit itself later in this chapter. In Ch. 10 we will discuss the dynamics of the system, including how the masses are measured.

Order- of-magnitude estimates

¯

Although our simple approach does not enable us to write down solutions for hμν generated by more complicated, nonperiodic motions, we can use Eq. (9.83) to obtain some order- of-magnitude estimates. Since Djk is of order MR2, for a system of mass M and size R, the radiation will have amplitude about M( R)2/r v2(M/r), where v is a typical internal velocity in the source. This applies directly to Eq. (9.99); note that in Eq. (9.93) the first term uses, instead of R2, the product l0A of the two characteristic lengths in the problem. If we are dealing with, say, a collapsing mass moving under its own gravitational forces, then by the virial theorem v2 φ0, the typical Newtonian potential in the source, while M/r φr, the Newtonian potential of the source at the observer’s distance r. Then we get the simple upper limit

h φ0φr.

(9.100)

Соседние файлы в предмете [НЕСОРТИРОВАННОЕ]