7 Gravitational-Wave Tests of Gravitational Theory
7.1 Gravitational-wave observatories
Soon after the publication of this update, a new method of testing relativistic gravity will be realized, when
a worldwide network of upgraded laser interferometric gravitational-wave observatories in the U.S. (LIGO
Hanford and LIGO Livingston) and Europe (VIRGO and GEO600) begins regular detection and analysis of
gravitational-wave signals from astrophysical sources. Within a few years, they may be joined by an
underground cryogenic interferometer (KAGRA) in Japan, and around 2022, by a LIGO-type interferometer
in India. These broad-band antennas will have the capability of detecting and measuring the gravitational
waveforms from astronomical sources in a frequency band between about 10 Hz (the seismic
noise cutoff) and 500 Hz (the photon counting noise cutoff), with a maximum sensitivity to
strain at around 100 Hz of (rms), for the kilometer-scale LIGO/VIRGO
projects. The most promising source for detection and study of the gravitational wave signal is
the “inspiralling compact binary” – a binary system of neutron stars or black holes (or one of
each) in the final minutes of a death spiral leading to a violent merger. Such is the fate, for
example, of the Hulse–Taylor binary pulsar B1913+16 in about 300 Myr, or the double pulsar
J0737–3039 in about 85 Myr. Given the expected sensitivity of the advanced LIGO-Virgo detectors,
which could see such sources out to many hundreds of megaparsecs, it has been estimated that
from 40 to several hundred annual inspiral events could be detectable. Other sources, such as
supernova core collapse events, instabilities in rapidly rotating newborn neutron stars, signals from
non-axisymmetric pulsars, and a stochastic background of waves, may be detectable (see [352] for a
review).
In addition, plans are being developed for orbiting laser interferometer space antennae, such as DECIGO
in Japan and eLISA in Europe. The eLISA system would consist of three spacecraft orbiting the sun in a
triangular formation separated from each other by a million kilometers, and would be sensitive primarily in
the very low-frequency band between and
, with peak strain sensitivity of order
.
A third approach that focuses on the ultra low-frequency band (nanohertz) is that of Pulsar Timing Arrays (PTA), whereby a network of highly stable millisecond pulsars is monitored in a coherent way using radio telescopes, in hopes of detecting the fluctuations in arrival times induced by passing gravitational waves.
For recent reviews of the status of all these approaches to gravitational-wave detection, see the Proceedings of the 8th Edoardo Amaldi Conference on Gravitational Waves [272].
In addition to opening a new astronomical window, the detailed observation of gravitational waves by such observatories may provide the means to test general relativistic predictions for the polarization and speed of the waves, for gravitational radiation damping and for strong-field gravity. These topics have been thoroughly covered in two recent Living Reviews by Gair et al. [170] for space-based detectors, and by Yunes and Siemens [452] for ground-based detectors. Here we present a brief overview.
7.2 Gravitational-wave amplitude and polarization
7.2.1 General relativity
A generic gravitational wave detector can be modelled as a body of mass at a distance
from a
fiducial laboratory point, connected to the point by a spring of resonant frequency
and quality
factor
. From the equation of geodesic deviation, the infinitesimal displacement
of the
mass along the line of separation from its equilibrium position satisfies the equation of motion























The waveforms and
depend on the nature and evolution of the source. For example, for
a binary system in a circular orbit, with an inclination
relative to the plane of the sky, and the
-axis
oriented along the major axis of the projected orbit, the quadrupole approximation of Eq. (86*) gives

7.2.2 Alternative theories of gravity
A generic gravitational wave detector whose scale is small compared to the gravitational wavelength
measures the local components of a symmetric tensor which is composed of the “electric”
components of the Riemann curvature tensor,
, via the equation of geodesic deviation, given, for a
pair of freely falling particles by
, where
denotes the spatial separation. In general
there are six independent components, which can be expressed in terms of polarizations (modes with
specific transformation properties under rotations and boosts); for a wave propagating in the
-direction,
they can be displayed by the matrix











Massless scalar–tensor gravitational waves can in addition contain the transverse breathing mode (c).
This can be obtained from the physical waveform , which is related to
and
to the required
order by






More general metric theories predict additional longitudinal modes, up to the full complement of six (TEGP 10.2 [420*]). For example, Einstein-Æther theory generically predicts all six modes [205].
A suitable array of gravitational antennas could delineate or limit the number of modes present
in a given wave. The strategy depends on whether or not the source direction is known. In
general there are eight unknowns (six polarizations and two direction cosines), but only six
measurables (). If the direction can be established by either association of the waves with
optical or other observations, or by time-of-flight measurements between separated detectors,
then six suitably oriented detectors suffice to determine all six components. If the direction
cannot be established, then the system is underdetermined, and no unique solution can be
found. However, if one assumes that only transverse waves are present, then there are only three
unknowns if the source direction is known, or five unknowns otherwise. Then the corresponding
number (three or five) of detectors can determine the polarization. If distinct evidence were found
of any mode other than the two transverse quadrupolar modes of GR, the result would be
disastrous for GR. On the other hand, the absence of a breathing mode would not necessarily
rule out scalar–tensor gravity, because the strength of that mode depends on the nature of the
source.
For laser interferometers, the signal controlling the laser phase output can be written in the form
where


Some of the details of implementing such polarization observations have been worked out for arrays of
resonant cylindrical, disk-shaped, spherical, and truncated icosahedral detectors (TEGP 10.2 [420*], for
recent reviews see [256, 403]). Early work to assess whether the ground-based or space-based
laser interferometers (or combinations of the two types) could perform interesting polarization
measurements was carried out in [404, 70, 267, 171, 411]; for a recent detailed analysis see [301].
Unfortunately for this purpose, the two LIGO observatories (in Washington and Louisiana
states, respectively) have been constructed to have their respective arms as parallel as possible,
apart from the curvature of the Earth; while this maximizes the joint sensitivity of the two
detectors to gravitational waves, it minimizes their ability to detect two modes of polarization.
In this regard the addition of Virgo, and the future KAGRA and LIGO-India systems will
be crucial to polarization measurements. By combining signals from various interferometers
into a kind of “null channel” one can test for the existence of modes beyond the and
modes in a model independent manner [78]. The capability of space-based interferometers to
measure the polarization modes was assessed in detail in [388, 302]. For pulsar timing arrays,
see [245, 14, 74].
7.3 Gravitational-wave phase evolution
7.3.1 General relativity
In the binary pulsar, a test of GR was made possible by measuring at least three relativistic effects that
depended upon only two unknown masses. The evolution of the orbital phase under the damping effect of
gravitational radiation played a crucial role. Another situation in which measurement of orbital phase can
lead to tests of GR is that of the inspiralling compact binary system. The key differences are that here
gravitational radiation itself is the detected signal, rather than radio pulses, and the phase
evolution alone carries all the information. In the binary pulsar, the first derivative of the binary
frequency was measured; here the full nonlinear variation of
as a function of time is
measured.
Broad-band laser interferometers are especially sensitive to the phase evolution of the gravitational
waves, which carry the information about the orbital phase evolution. The analysis of gravitational wave
data from such sources will involve some form of matched filtering of the noisy detector output against an
ensemble of theoretical “template” waveforms which depend on the intrinsic parameters of the inspiralling
binary, such as the component masses, spins, and so on, and on its inspiral evolution. How accurate must a
template be in order to “match” the waveform from a given source (where by a match we mean maximizing
the cross-correlation or the signal-to-noise ratio)? In the total accumulated phase of the wave detected in
the sensitive bandwidth, the template must match the signal to a fraction of a cycle. For two
inspiralling neutron stars detected by the advanced LIGO/Virgo systems, around 16 000 cycles
should be detected during the final few minutes of inspiral; this implies a phasing accuracy of
or better. Since
during the late inspiral, this means that correction terms
in the phasing at the level of
or higher are needed. More formal analyses confirm this
intuition [99, 153, 97, 323].
Because it is a slow-motion system (), the binary pulsar is sensitive only to the lowest-order
effects of gravitational radiation as predicted by the quadrupole formula. Nevertheless, the first correction
terms of order
and
to the quadrupole formula were calculated as early as 1976 [405] (see
TEGP 10.3 [420*]).
But for laser interferometric observations of gravitational waves, the bottom line is that, in order to measure the astrophysical parameters of the source and to test the properties of the gravitational waves, it is necessary to derive the gravitational waveform and the resulting radiation back-reaction on the orbit phasing at least to 3PN order beyond the quadrupole approximation.
For the special case of non-spinning bodies moving on quasi-circular orbits (i.e., circular apart from a
slow inspiral), the evolution of the gravitational wave frequency through 2PN order has the form


Similar expressions can be derived for the loss of angular momentum and linear momentum. Expressions
for non-circular orbits have also been derived [175, 107]. These losses react back on the orbit to circularize
it and cause it to inspiral. The result is that the orbital phase (and consequently the gravitational wave
phase) evolves non-linearly with time. It is the sensitivity of the broad-band laser interferometric
detectors to phase that makes the higher-order contributions to so observationally
relevant.
If the coefficients of each of the powers of in Eq. (135*) can be measured, then one again obtains
more than two constraints on the two unknowns
and
, leading to the possibility to test GR. For
example, Blanchet and Sathyaprakash [59, 60] have shown that, by observing a source with a sufficiently
strong signal, an interesting test of the
coefficient of the “tail” term could be performed (but see [22]
for a more sophisticated analysis).
Another possibility involves gravitational waves from a small mass orbiting and inspiralling into a (possibly supermassive) spinning black hole. A general non-circular, non-equatorial orbit will precess around the hole, both in periastron and in orbital plane, leading to a complex gravitational waveform that carries information about the non-spherical, strong-field spacetime around the hole. According to GR, this spacetime must be the Kerr spacetime of a rotating black hole, uniquely specified by its mass and angular momentum, and consequently, observation of the waves could test this fundamental hypothesis of GR [345, 322].
7.3.2 Alternative theories of gravity
In general, alternative theories of gravity will predict rather different phase evolution from that of GR, notably via the addition of dipole gravitational radiation. For example, the dipole gravitational radiation predicted by scalar–tensor theories modifies the gravitational radiation back-reaction, and thereby the phase evolution. Including only the leading 0PN and –1PN (dipole) contributions, one obtains,
where









These considerations suggest that it might be fruitful to attempt to parametrize the phasing formulae in a manner reminiscent of the PPN framework for post-Newtonian gravity. A number of approaches along this line have been developed, including the parametrized post-Einsteinian (PPE) framework [451, 347], a Bayesian parametrized approach [250], and a parametrization based on the post-Newtonian expansions discussed above [288]. The discovery of relationships between the moment of inertia, the gravitational Love number, and the quadrupole moment of neutron stars (“I-Love-Q” relations) in general relativity has opened the possibility of testing theories using gravitational waves in a manner that is relatively free of contamination from the neutron-star equation of state [448, 447].
7.4 Speed of gravitational waves
According to GR, in the limit in which the wavelength of gravitational waves is small compared to the
radius of curvature of the background spacetime, the waves propagate along null geodesics of the
background spacetime, i.e., they have the same speed as light (in this section, we do not set
). In
other theories, the speed could differ from
because of coupling of gravitation to “background”
gravitational fields. For example, in the Rosen bimetric theory with a flat background metric
, gravitational waves follow null geodesics of
, while light follows null geodesics of
(TEGP 10.1 [420*]).
Another way in which the speed of gravitational waves could differ from is if gravitation were
propagated by a massive field (a massive graviton), in which case
would be given by, in a local inertial
frame,


The most obvious way to test this is to compare the arrival times of a gravitational wave and an
electromagnetic wave from the same event, e.g., a supernova or a prompt gamma-ray burst. For a source at
a distance , the resulting value of the difference
is







For a massive graviton, if the frequency of the gravitational waves is such that , where
is Planck’s constant, then
, where
is the graviton
Compton wavelength, and the bound on
can be converted to a bound on
, given by
The foregoing discussion assumes that the source emits both gravitational and electromagnetic radiation in detectable amounts, and that the relative time of emission can be established to sufficient accuracy, or can be shown to be sufficiently small.
However, there is a situation in which a bound on the graviton mass can be set using gravitational
radiation alone [423*]. That is the case of the inspiralling compact binary. Because the frequency of
the gravitational radiation sweeps from low frequency at the initial moment of observation
to higher frequency at the final moment, the speed of the gravitons emitted will vary, from
lower speeds initially to higher speeds (closer to ) at the end. This will cause a distortion of
the observed phasing of the waves and result in a shorter than expected overall time
of
passage of a given number of cycles. Furthermore, through the technique of matched filtering, the
parameters of the compact binary can be measured accurately (assuming that GR is a good
approximation to the orbital evolution, even in the presence of a massive graviton), and thereby the
emission time
can be determined accurately. Roughly speaking, the “phase interval”
in Eq. (139*) can be measured to an accuracy
, where
is the signal-to-noise
ratio.
Thus one can estimate the bounds on achievable for various compact inspiral systems, and for
various detectors. For stellar-mass inspiral (neutron stars or black holes) observed by the LIGO/VIRGO
class of ground-based interferometers,
,
, and
. The result
is
. For supermassive binary black holes (
to
) observed by the proposed laser
interferometer space antenna (LISA),
,
, and
. The result
is
.
A full noise analysis using proposed noise curves for the advanced LIGO and for LISA weakens these
crude bounds by factors between two and 10 [423, 433, 41*, 42, 23, 377, 445]. For example, for the inspiral
of two black holes with aligned spins at a distance of
observed by LISA, a bound of
could be placed [41]. Other possibilities include using binary pulsar data to bound
modifications of gravitational radiation damping by a massive graviton [154], using LISA observations of
the phasing of waves from compact white-dwarf binaries, eccentric galactic binaries, and eccentric inspiral
binaries [98, 209], using pulsar timing arrays [244], and using DECIGO/BBO to observe neutron-star
intermediate-mass black-hole inspirals [446].
Bounds obtainable from gravitational radiation effects should be compared with the solid bound
[381] derived from solar system dynamics, which limit the presence of a Yukawa
modification of Newtonian gravity of the form

Mirshekari et al. [287] studied bounds that could be placed on more general graviton dispersion relations that could emerge from alternative theories with Lorentz violation, in which the effective propagation speed is given by
where
