Testing Strong-Field Gravity With Tidal Love Numbers
Testing Strong-Field Gravity With Tidal Love Numbers
Testing Strong-Field Gravity With Tidal Love Numbers
Vitor Cardoso,1,2 Edgardo Franzin,3 Andrea Maselli,4 Paolo Pani,5,1 Guilherme Raposo1
1
CENTRA, Departamento de Fsica, Instituto Superior Tecnico IST,
Universidade de Lisboa UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal
2
Perimeter Institute for Theoretical Physics, 31 Caroline Street North Waterloo, Ontario N2L 2Y5, Canada
3
Dipartimento di Fisica, Universit`
a di Cagliari & Sezione INFN Cagliari,
Cittadella Universitaria, 09042 Monserrato, Italy
4
Theoretical Astrophysics, Eberhard Karls University of Tuebingen, Tuebingen 72076, Germany and
Dipartimento di Fisica, Sapienza Universit`
a di Roma & Sezione INFN Roma1, Piazzale Aldo Moro 5, 00185, Roma, Italy
The tidal Love numbers (TLNs) encode the deformability of a self-gravitating object immersed
in a tidal environment and depend significantly both on the objects internal structure and on the
dynamics of the gravitational field. An intriguing result in classical general relativity is the vanishing
of the TLNs of black holes. We extend this result in three ways, aiming at testing the nature of
compact objects: (i) we compute the TLNs of exotic compact objects, including different families
of boson stars, gravastars, wormholes, and other toy models for quantum corrections at the horizon
scale. In the black-hole limit, we find a universal logarithmic dependence of the TLNs on the location
of the surface; (ii) we compute the TLNs of black holes beyond vacuum general relativity, including
Einstein-Maxwell, Brans-Dicke and Chern-Simons gravity; (iii) We assess the ability of present and
future gravitational-wave detectors to measure the TLNs of these objects, including the first analysis
of TLNs with LISA. Both LIGO, ET and LISA can impose interesting constraints on boson stars,
while LISA is able to probe even extremely compact objects. We argue that the TLNs provide a
smoking gun of new physics at the horizon scale, and that future gravitational-wave measurements
of the TLNs in a binary inspiral provide a novel way to test black holes and general relativity in the
strong-field regime.
CONTENTS
I. Introduction
A. The naturalness problem
B. Quantifying the existence of horizons
C. Executive summary
1
2
2
2
5
5
6
9
9
9
10
V. Detectability
A. Model-independent tests with GWs
B. Detectability of ECOs
C. Detectability of BSs
D. Testing GR
12
13
14
16
16
16
Acknowledgments
17
17
B. Determination of TLNs
17
19
19
20
21
22
23
References
23
I.
INTRODUCTION
Tidal interactions play a fundamental role in astrophysics across a broad range of scales, from stellar objects like ordinary stars and neutron stars (NSs) to large
celestial systems such as galaxies. Several astrophysical
structures (e.g., binaries and tidal tails [1, 2]) are consequences of tidal interactions. Tidal effects can be particularly strong and important in the regime that characterizes compact objects, giving rise to extreme phenomena
such as tidal disruptions.
The deformability of a self-gravitating object immersed
in an external tidal field is measured in terms of its tidal
Love numbers (TLNs) [3, 4]. These leave a detectable imprint in the gravitational-wave (GW) signal emitted by a
neutron-star binary in the late stages of its orbital evolution [57]. So far, a relativistic extension [6, 8, 9] of the
Newtonian theory of tidal deformability has been mostly
motivated by the prospect of measuring the TLNs of NSs
through GW detections and, in turn, understanding the
behavior of matter at supranuclear densities [1016]. The
scope of this paper is to show that tidal effects can also
2
be used to explore more fundamental questions related to
the nature of compact objects and the behavior of gravity
in the strong-field regime.1
An intriguing result in classical general relativity (GR)
is the fact that the TLNs of a black hole (BH) are precisely zero. This property has been originally demonstrated for small tidal deformations of a Schwarzschild
BH [8, 9, 18] and has been recently extended to arbitrarily strong tidal fields [19] and to the spinning case [2022],
at least in the axisymmetric case to quadratic order in the
spin [21] and generically to linear order in the spin [22].
A.
A related, independent work dealing with tidal effects for boson stars, conducted simultaneously to ours, is due to appear
soon [17].
B.
C.
Executive summary
3
sectors can in turn be expanded into a set of multipoles
labeled by an integer l.
Our main results are summarized in Table I and in
Fig. 1, and are discussed in detail in the rest of the paper.
Table I lists the lowest quadrupolar (l = 2) and octupolar (l = 3) polar and axial TLNs for various models of
ECOs in GR, and for some static BHs in other gravity
theories. Table I also compares the TLNs of these objects with the corresponding ones for a typical NS (cf.
also Table III in Appendix A).
One of our main results is that the TLNs of several
ECOs display a logarithmic dependence in the BH limit,
i.e. when the compactness of the object approaches that
of a BH,
C := M/r0 1/2 ,
(1)
where M and r0 are the mass and the radius of the object.
As shown in Table I, this property holds for wormholes,
thin-shell gravastars, and for a simple toy model of a
static object with a perfectly reflecting surface [35, 37].
It is natural to conjecture that this logarithmic behavior
is model independent and will hold for any ECO whose
exterior spacetime is arbitrarily close to that of a BH
in the r0 2M limit. This mild dependence implies
that even the TLNs of an object with r0 2M `P are
not extremely small, contrarily to what one could expect.
Indeed, we estimate that the dimensionless TLNs defined
in Eq. (3) below are
k2E,B O(103 ) ,
k3E,B O(104 ) ,
(2)
It is slightly more common to use the distinction electric/magnetic components rather than polar/axial. Since we
shall discuss also electromagnetic fields, we prefer to use the former distinction.
We adopt the Geroch-Hansen definition of multipole moments [62, 63], equivalent [64] to the one by Thorne [65] in asymptotically mass-centered Cartesian coordinates.
This symmetry is broken if the compact object is spinning due
to spin-tidal couplings. In such case, there exists a series of
selection rules that allow to define a wider class of rotational
TLNs [21, 22, 66, 67]. In this paper, we neglect spin effects to
leading order.
We consider only non-spinning objects, hence the spacetime is
|/|[%]
1000
500
100
50
10
20
30
10
10
10
40
1
0.5
0.1
50
10
M [M]
20
30
40
50
|/|[%]
|/|[%]
0.10
0.01
5 10
50 100
M [10 M]
M [M]
FIG. 1. Relative percentage errors on the average tidal deformability for BS-BS binaries observed by AdLIGO (left panel),
ET (middle panel), and LISA (right panel), as a function of the BS mass and for different BS models considered in this work
(for each model, we considered the most compact configuration in the stable branch; see main text for details). For terrestrial
interferometers we assume a prototype binary at d = 100 Mpc, while for LISA the source is located at d = 500 Mpc. The
horizontal dashed line identifies the upper bound / = 1. Roughly speaking, a measurement of the TLNs for systems which
lie below the threshold line would be incompatible with zero and, therefore, the corresponding BSs can be distinguished from
BHs. Here is given by Eq. (72), the two inspiralling objects have the same mass, and / kE /k2E .
2
TABLE I. Tidal Love numbers (TLNs) of some exotic compact objects (ECOs) and BHs in Einstein-Maxwell theory and modified
theories of gravity; details are given in the main text. As a comparison, we provide the order of magnitude of the TLNs for static NSs
with compactness C 0.2 (the precise number depends on the neutron-star equation of state; see Table III for more precise fits). For BSs,
the table provides the lowest value of the corresponding TLNs among different models (cf. Sec. III A) and values of the compactness. In
the polar case, the lowest TLNs correspond to solitonic BSs with compactness C 0.18 or C 0.20 (when the radius is defined as that
containing 99% or 90% of the total mass, respectively). In the axial case, the lowest TLNs correspond to a massive BS with C 0.16 or
C 0.2 (again for the two definitions of the radius, respectively) and in the limit of large quartic coupling. For other ECOs, we provide
expressions for very compact configurations where the surface r0 sits at r0 2M and is parametrized by := r0 /(2M ) 1; the full results
are available online [57]. In the Chern-Simons case, the axial l = 3 TLN is affected by some ambiguity and is denoted by a question mark
[see Sec. IV C for more details].
k3B
1300
11
70
41.4
402.8
13.6
211.8
Gravastar
4
5(8+3 log )
8
5(7+3 log )
16
5(236 log 2+9 log )
8
105(7+2 log )
8
35(10+3 log )
16
35(316 log 2+9 log )
16
5(31+12 log )
32
5(25+12 log )
32
5(4312 log 2+18 log )
16
7(209+60 log )
32
7(197+60 log )
32
7(30760 log 2+90 log )
Einstein-Maxwell
Scalar-tensor
0
0
0
0
0
0
0
0
Chern-Simons
1.1 MCS
4
NSs
ECOs
BHs
Boson star
Wormhole
Perfect mirror
k2E
k3E
210
factor M 2l+1 was introduced to make the above quantities dimensionless. It is customary to normalize the
TLNs by powers of the objects radius R rather than
by powers of its mass M . Here we adopted the latter
non-standard choice, since the radius of some ECOs (e.g.
11.1 MCS
4 ?
R
M
2l+1
klE,B
HBP .
(4)
Modified theories of gravity and ECOs typically require the presence of extra fields which are
(non)minimally coupled to the metric tensor. Here we
shall consider some representative example of both scalar
5
and vector fields. A full treatment of this problem would
require allowance for an extra degree of freedom, the external scalar and electromagnetic (EM) applied fields. It
is generically expected that, in astrophysical situations,
the ratio of a putative external (scalar or vector) field
to the ordinary gravitational tidal field should be small.
We will therefore focus only on situations where the only
surviving field at large distances is gravitational.
We expand the metric, the scalar field, and the
Maxwell field in spherical harmonics as presented in Appendix B. Since the background is spherically symmetric, perturbations with different parity and different harmonic index l decouple. In the following we discuss the
polar and axial sector separately; due to the spherical
symmetry of the background, the azimuthal number m
is degenerate and we drop it.
Finally, in order to extract the tidal field and the induced multipole moments from the solution, we have
adopted two (related) techniques. The first one relies on an expansion of the metric at large distances
[cf. Eqs. (B9) and (B10)] in terms of the multipole moments. The second technique relies on the evaluation of
the Riemann tensor in Schwarzschild coordinates, whose
tidal correction is related to the total tidal field in the
local asymptotic rest frame [18]. These two procedures
agree with each other and at least in the case of ECOs
the computation of the TLNs is equivalent to the case of
NSs [6, 8]. On the other hand, computing the TLNs of
BHs in extensions of GR presents some subtleties which
are discussed in Sec. IV.
III.
A.
Potential
V (||2 )
Model
2 ||2
Minimal
Massive
Solitonic
Boson stars
4
||2 +
4 ||
2
2||2
2 ||2 1
2
0
Maximum mass
Mmax /M
11
8 10 m eV
S
2
5 ~ 0.1mGeV
S
h 12 i2
10
0
500 GeV
mS
R
S = d4 x g
g ab a b V ||2 . (5)
16
BSs have been extensively studied in the past and have
been proposed as BH mimickers and dark matter candidates, see e.g. Refs. [2527, 68, 69].
BSs are typically classified according to the scalar potential in the above action; here we investigate three of
the most common models: minimal BSs [70, 71], massive BSs [72] and solitonic BSs [73]. The corresponding
scalar potential for these models and the maximum mass
for non-spinning solutions are listed in Table II. A more
comprehensive list of BS models can be found in Ref. [26].
Depending on the model, compact BSs with masses
comparable to those of ordinary stars or BHs require a
certain range of the scalar mass mS := ~. For minimal BSs, the maximum mass in Table II is comparable
to the Chandrasekhar limit for NSs only for an ultralight
field with mS . 1011 eV. For massive BSs, the maximum mass is of the same order of the Chandrasekhar
limit if mS 0.1 GeV and the quartic coupling is large,
~ 1 [72]. Finally, solitonic BSs may reach massive
(M & M ) or supermassive (M & 106 M ) configurations even for heavy bosons with mS 500 GeV if the
coupling parameter in their potential is 0 . 1012 or
0 . 1015 , respectively [73]. For massive and solitonic
BSs, the scaling of the maximum mass in Table II is approximate and valid only when 2 and when 0 1,
respectively. In our numerical analysis, we have considered = 104 2 and 0 = 0.05, whereas the mass term
can be rescaled away (cf., e.g., discussion in Ref. [39]).
Even though BSs have a wide range of compactness,
which depends basically on their total mass (cf. Fig. 9
in Appendix C), interactions between BSs typically leads
to a net weight gain, clustering old BSs close to the mass
peak [27], which also coincides with the peak of compactness.
The details of the numerical procedure to compute the
TLNs of a BS are presented in Appendix C. Figure 2
shows the TLNs of the BS models presented above as a
function of the total mass M , the latter being normalized by the total mass Mmax of the corresponding model.
We only show static configurations in the stable branch,
i.e. with a mass smaller than Mmax (cf. discussion in Appendix C). For minimal BSs and for l = 2 polar case, our
results agree with those recently obtained in Ref. [74].
In addition, we also present the results for l = 2 and
l = 3, for both axial and polar TLNs, and for the three
BS models previously discussed.
The behavior of the TLNs of BSs is in qualitative
agreement with that of NSs. For a given BS model with
a given mass, the magnitude of the polar TLN is larger
than that of an axial TLN with the same l. Furthermore, in the Newtonian regime (M 0) the TLNs scale
as klE C (2l+1) and klB C 2l . This scaling is in
FIG. 2. Polar (top panels) and axial (bottom panels) TLNs for minimal, massive and solitonic BSs. Left and right panels refers
to l = 2 and l = 3, respectively. For massive and solitonic BSs we have considered = 104 2 and 0 = 0.05, respectively. With
these values, the maximum mass scales approximately as shown in Table II. Numerical data are available online [57]. These
plots include only stars in the stable branch.
B.
Several phenomenological models of quantum BHs introduce a Planck-scale modification near the horizon.
In this section, we consider three toy models for microscopic corrections at the horizon scale, namely a wormhole [31], a Schwarzschild geometry with a perfectly reflective surface near the horizon [35, 37], and a thinshell gravastar [29].6 These models have some common
features: (i) the exterior spacetime is described by the
Schwarzschild metric; (ii) the interior is either vacuum
or de Sitter and the tidal perturbation equations can be
solved for in closed form; (iii) simple junction or boundary conditions at the radius r0 of the object can be imposed to connect the perturbations in the interior with
those in the exterior. As a result of these properties, the
TLNs of these models can be computed in closed analytical form. As we show, the qualitative features are the
same and especially in the BH limit do not depend
strongly on the details of the models. Below, we present
explicit formulas for the BH limit, expressions for generic
compactness are provided online [57]. The details of the
computation are given in Appendix D.
1.
Wormholes
Many of these objects are unstable or require exotic matter distributions. We will not be concerned with these issues here.
(6)
7
distinct regions connected by a wormhole with a throat
at r = r0 . Since the wormhole spacetime is composed by
two Schwarzschild metrics, the stress-energy tensor vanishes everywhere except on the throat of the wormhole.
The patching at the throat requires a thin-shell of matter
with surface density and surface pressure
120
0.050
100
0.010
0.005
80
0.001
5. 10-4
1
=
2r0
2M
1
,
r0
1
1 M/r0
p
p=
, (7)
4r0 1 2M/r0
which imply that the weak and the dominant energy conditions are violated, whereas the null and the strong energy conditions are satisfied when r0 < 3M [55]. To
cover the two patches of the spacetime, we use the radial
tortoise coordinate r , which is defined by
dr
2M
= 1
,
(8)
dr
r
where the upper and lower sign refer to the two sides
of the wormhole. Without loss of generality, we can
assume that the tortoise coordinate at the throat is
zero, r (r0 ) = 0, so that one side corresponds to r > 0
whereas the other side corresponds to r < 0.
In Fig. 3, we show the polar and axial TLNs with
l = 2, 3 as functions of := r0 /(2M ) 1. Interestingly,
in this case the TLNs have the opposite sign to those of a
NS. Furthermore, they vanish in the BH limit, i.e. when
r0 2M or 0. The behavior of the TLNs in the BH
limit reads
4
,
5(8 + 3 log )
8
,
k3E
105(7 + 2 log )
16
k2B
,
5(31 + 12 log )
16
k3B
,
7(209 + 60 log )
k2E
(9)
(10)
(11)
(12)
(log )2
get klE,B
k2B 6 103 ,
k3E 4 104 ,
k3E 9 104 ,
(13)
10-35
60
10-25
10-15
10-5
40
20
0
0.0
0.2
0.4
0.6
0.8
1.0
2.
Perfectly-reflective mirror
Thermodynamical arguments suggest that any horizonless microscopic model of BH should act as a mirror,
at least for long wavelength perturbations [35, 37]. Motivated by this scenario, we consider a Schwarzschild geometry with a perfect mirror at r = r0 > 2M and impose
Dirichlet boundary conditions on the Regge-Wheeler and
Zerilli functions, for the axial and polar sector, respectively. Thus, our strategy is to consider the stationary limit of generically dynamical perturbations (in the
Fourier space, where is the frequency of the perturbation) of a Schwarzschild geometry.
The final result, in the 0 limit, reads (cf. Appendix D for details)
8
6 103 ,
5(7 + 3 log )
8
k3E
9 104 ,
35(10 + 3 log )
32
k2B
6 103 ,
5(25 + 12 log )
32
k3B
9 104 ,
7(197 + 60 log )
k2E
(14)
(15)
(16)
(17)
8
Appendix D yields7
0.050
16
4 103 ,
(19)
5(23 6 log 2 + 9 log )
16
k3E
6 104 , (20)
35(31 6 log 2 + 9 log )
32
4 103 , (21)
k2B
5(43 12 log 2 + 18 log )
32
6 104 . (22)
k3B
7(307 60 log 2 + 90 log )
200
k2E
0.010
0.005
150
0.001
10-35
10-25
10-15
10-5
100
50
0
0.0
0.2
0.4
0.6
0.8
1.0
3.
Thin-shell gravastars
40
0.010
0.005
30
0.001
5. 10-4
10-35
10-25
10-15
10-5
20
10
0
0.0
0.2
0.4
0.6
0.8
1.0
e =e
1 2M
r
2
1 2C rr2
0
r > r0
.
r < r0
(18)
FIG. 5. TLNs for a thin-shell gravastar with zero energy density as a function of the compactness. More generic gravastar
models are presented in Ref. [59]. The TLNs are all negative and vanish in the BH limit, r0 2M . Similar to the
perfectly-reflective mirror case, the polar- and axial-type Love
numbers for the same multipolar order coincide in the BH
limit, as shown in the inset.
9
4.
It is remarkable that the models described above display a very similar behavior in the BH limit, when the
radius r0 2M , cf. Table I. Indeed, although all TLNs
vanish in this limit, they have a mild logarithmic dependence. On the light of our results, it is natural to
conjecture that this logarithmic dependence is a generic
feature of ultracompact exotic objects, and will hold true
for any ECO whose exterior spacetime is arbitrarily close
to that of a BH in the r0 2M limit.
Due to this mild dependence, the TLNs are not extremely small, as one would have naively expected if the
scaling with were polynomial. Indeed, in the Planckian
case (r0 2M `P ) the order of magnitude of the TLNs
is the same for all models and it is given by Eq. (2). In
particular, the TLNs of Planckian ECOs are only five orders of magnitude smaller than those a typical NS. The
detectability of these deviations from the zero-Love
rule of BHs in GR is discussed in Sec. V.
IV.
A.
Scalar-tensor theories
G = 2 g + ,
(24)
= 0 .
(25)
As mentioned above, the background solution is
Schwarzschild with a vanishing scalar field.
Following the procedure described in Sec. II, we consider metric perturbations given by Eqs. (B3) and (B4)
for the polar and axial sector, respectively, and a scalar
2(r M )0 l(l + 1)
= 0,
r(r 2M )
(26)
Einstein-Maxwell
(30)
(31)
10
where F = A, A, is the Maxwell tensor and
1
1
g F F F F g ,
(32)
T =
4
4
is the stress-energy tensor of the EM field. The background spacetime is the well-known Reissner-Nordstrom
metric, whose line element reads as in Eq. (B2) with
e = eg = 1
2M
Q2
+ 2 f (r),
r
r
(33)
A = (Q/r, 0, 0, 0) .
Polar TLNs
D1 H0 +
(35)
(36)
d2
2(M r) d
1 2
+
Q r(4M ( 2)r)
dr2
r2 f
dr r6 f 2
r2 4M 2 2M r + r2 2Q4 ,
Q2 r2
d
=
+
,
dr r (r(r 2M ) + Q2 )
d2
4Q2 r2 d
= 2+
,
dr
r4 f
dr
2 M r Q2
d
,
=
+
dr
r3 f
D1 =
(1)
D1
(2)
D2
(1)
D2
2.
(34)
Because the background is electrically charged, gravitational and EM perturbations are coupled to each other.
To compute the tidal deformations, we expand the metric as in Eqs. (B3) and (B4) and the Maxwell field as in
Eqs. (B5) and (B8). As before, we consider the polar and
the axial sectors separately.
1.
(37)
(38)
Axial TLNs
The calculations for gravitational axial TLNs and magnetic TLNs are simpler. The r-component of Einsteins
equations leads to h1 = 0 which automatically satisfies
the -component. The final axial system reads
4Qu04
r2 f r2 h000
2Q2 + r (r 4M ) h0 = 0 ,
(39)
2Q
(Qu04 h0 ) = 0 .
r2 f u004 Qh00 + 2M u04 u4
r
(40)
Also in the axial sector, the coupled system admits an
analytic, closed-form solution. In the absence of EM tidal
fields, the solutions which are regular at the horizon read
r3
f B2 ,
3
r2
= QB2 (1 Q2 /r2 ) .
2
hl=2
=
0
(41)
ul=2
4
(42)
C.
Chern-Simons gravity
1
CS
SCS = d4 x g R g ab a b +
RR ,
2
4
(43)
where CS is the coupling constant of the theory and
RR =
1
Rabcd baef Rcd ef .
2
(44)
11
1
To O(CS
), the only correction is in the scalar-field equation, which reads
Rab =
(45)
(2)
DS (1) =
(46)
12B2 M
,
r2 (r 2M )
(50)
where
cde(a
b)
e R d
dabc
where Cab = c
+ d c R
. We will
focus on spherically symmetric background solutions to
Eqs. (45) and (46). In these conditions, the Pontryagin
scalar vanishes and the background is described by the
Schwarzschild metric with a vanishing scalar field [7678].
1.
Polar TLNs
2.
Axial TLNs
On the other hand, the field transforms as a pseudoscalar and is therefore part of the axial sector. We can
thus expect that non-trivial axial-type TLNs in ChernSimons gravity may exist. In the stationary limit, we
find h1 = 0, and the field equations for the axial sector
reduce to a system of two coupled second-order differential equations for h0 and . This system can be solved
numerically for a generic coupling CS or perturbatively
when CS := CS /M 2 1. The latter case is consistent
with the action (43) being an effective field theory [54].
We have adopted both procedures, described below.
In the perturbative limit, we expand the metric and
scalar perturbations in powers of the coupling CS 1,
2
(2)
h = h(0)
+ CS h + ...
(1)
= CS
+ ...
(47)
(48)
(l)
DS :=
2(M r) d
d2
l(l + 1)
2
.
dr2
r 2M r dr r2 2M r
(51)
(r) =
12
not contribute to the current quadrupole moment S2 .
Interestingly, these integrals can be computed in closed
form although their final expression is cumbersome. We
(2)
report here only the large-distance behavior of h0 (r),
namely
3
M
B2 M 5
9
(2)
,
(57)
h0 (r) [9 8(3)] 2 + O
5
r
r3
where (n) is the Riemann Zeta function. By comparing
the above result with Eq. (B10) and using Eq. (3), it is
straightforward to obtain
k2B =
9
2
2
[8(3) 9]CS
1.10962 CS
.
5
(58)
0.25
0.20
0.15
0.10
0.05
0.00
0.0
0.1
0.2
0.3
0.4
0.5
B3 M 2
=
r (r 2M )(3r 4M ) ,
27
(59)
DS (1) =
16B3 M 3 (3r 5M )
,
3(r 2M )r2
(60)
75600(3) 84331
h0
3r2
3969r3
4
r
M
7560 log
+O
.
(61)
2M
r4
We therefore obtain two terms (the 1/r2 and the log r/r3
term) which decay more slowly than the octupole term,
1/r3 . We believe that these terms arise as subleading
corrections to the external tidal field, which is not captured by the asymptotic expansion in Eq. (B10), because
the latter does not include the effects of a constant scalar
field at infinity. In other words, the multipolar structure
of a tidally deformed BH in Chern-Simons gravity is more
involved than in GR. We anticipate that this issue also
appears in other modified theories and we postpone a
more detailed analysis to the future [80]. In order to get
an estimate for the TLN, we simply consider the ordinary
octupolar correction. By proceeding in the usual way, we
obtain
1
2
2
k3B =
(75600(3) 84331) CS
11.13 CS
, (62)
588
which is the value reported in Table I although, for the
reasons mentioned above, should only be considered as
an estimate.
V.
DETECTABILITY
(63)
(64)
13
~ prior on the parameters [82]. The bracket
with p(0) ()
(|) represents the inner product
Z
)
? (f )
h(f
g ? (f ) + h
g (f )
df
,
(65)
(g|h) = 2
S
(f
)
h
where Sh (f ) is the detectors noise spectral density. According to the principle of the maximum-likelihood estimator, the values of the source parameters can be estimated as those which maximize Eq. (64). In the limit
~
of large signal-to-noise ratio, such that p(|h)
is tightly
peaked around the true values of the source parameters,
~ around
a Taylor expansion of p(|s)
~ leads to
~ p(0) ()e
~ 12 ab a b ,
p(|s)
(66)
where
ab =
h h
a b =~
~
(67)
a = aa ,
(69)
and the correlation coefficients between a and b are
given by
cab =
h a b i
ab
=
.
aa bb
aa bb
(70)
The TLNs enter the GW signal as a fifth-order postNewtonian (PN) correction which adds linearly to the
phase of the waveform,
) = A(f )ei(PP +T ) ,
h(f
2 5 E
M k2 .
3
where q := m1 /m2 > 1 is the mass ratio. For nonspinning objects, the waveform depends on 6 parameters ~ = (ln A, c , tc , ln M, ln , ), i.e. the amplitude, the phase and time at the coalescence, the chirp
mass M = 3/5 (m1 + m2 ), the symmetric mass ratio
= m1 m2 /(m1 + m2 )2 and the average tidal deformability defined in Eq. (73). Nonetheless, ln A is completely
uncorrelated with the other variables, and therefore we
will restrict our analysis by performing derivatives only
with respect to the remaining parameters, leading to a
5 5 Fisher matrix.
The detector properties are encoded in the noise spectral density Sh (f ). We perform the analysis both for
terrestrial and space interferometers.
For the Earth-based detectors, we consider
(i) AdLIGO with its anticipated design sensitivity
curve ZERO DET high P [86] and (ii) the ET design
configuration, with noise described by the analytic fit
provided in Ref. [87]. As space-based detector, we
consider the most optimistic LISA configuration, namely
the N2A5 model defined in Ref. [88], with a 5 106 km
arm-length and an observing time of Tobs = 5 yr.
To compute the errors on the tidal deformability, we numerically integrate Eq. (67) within the freAdLIGO
quency range [fmin , fmax ], where fmin
= 20 Hz,
ET
LISA
fmin = 1 Hz, and fmin
= Max[105 , 4.149
3/8
105 (106 M)5/8 Tobs ] Hz [89]. For the upper freAdLIGO
quency we choose fmax
= (63/2 m)1 , while for LISA
LISA
3/2
fmax = Min[1 Hz, (6 m)1 ], being m = m1 + m2 the
total mass of the system.
A.
(72)
The contribution of higher multipoles and of the axialtype TLNs is subleading and will be neglected in the
following.
In our analysis, we use the so-called TaylorF2 approximant of the GW template in the frequency domain [83],
which is 3.5PN accurate in the point-particle phase and
2PN accurate in the tidal term [84, 85].8 For binary systems for which i=1,2 6= 0, finite-size effects are described
(73)
(71)
14
kE2 = 10
10
00
kE 2 = 10
1000
10
100
10
= 100
1
0.1
107
105
10
15
20
25
30
106
105
104
1000
100
10
1
0.1
100
10
15
20
10
1
0.100
0.010
0.001
1011
M [M]
|/|[%]
|/|[%]
25
30
|/|[%]
100
k E2 = 1
|/|[%]
|/|[%]
1000
|/|[%]
109
107
105
1000
10
0.1
kE
kE
=1
=1
0
kE
2 =
100
2
kE
=1
000
10
k
10 2
=
2
10
= 24
10
= 26
10
=
28
10
=
20
50
100
M [104 M]
M [M]
FIG. 7. Relative percentage errors on the average tidal deformability for equal-mass binaries at 100 Mpc (for AdLIGO and
ET, left and middle panel, respectively) and at 500 Mpc (for LISA, right panel) as functions of the mass of the single object
and for different values of the TLN k2E (top panels) and of (bottom panels) of the two objects. The horizontal dashed line
identifies the upper bound / = 1.
represents the mass range in which terrestrial interferometers will provide new information on matter at supranuclear densities from neutron-star binaries. On the other
hand, our results shown in Fig. 7 do not assume any specific model and extend the analysis of the detectability
of the TLNs to a regime unexplored so far, where more
massive ECOs can contribute to the GW signal through
finite-size effects. Likewise, to the best of our knowledge,
this work presents the first analysis on the detectability
of tidal effects with LISA.
From the bottom panels of Fig. 7 we note that, for a
fixed , the detectability is favored for low-mass systems,
as the tidal phase scales with the inverse of the total mass
T m10/3 (1 + q)3 /q. Moreover, for 2M . M .
5M , AdLIGO will constrain the TLNs for small compactness only (i.e., for large ). This picture improves
for ET, which leads to an upper bound / = 1 up
to M ' 15M . Therefore, as far as terrestrial interferometers are considered, the high-compactness regime for
ECOs seems to be available only for the third generation
of detectors. This result is also evident from the top panels of the left and middle plots in Fig. 7, which show that
AdLIGO will not be able to set any significant constraint
below k2E ' 10, regardless the ECO mass.
On the other hand, space interferometers open a completely new window onto finite-size effects. The top-right
panel of Fig. 7 shows that LISA is capable to bound the
Love numbers with a relative accuracy / . 10% in
almost the entire mass range M [104 , 106 ] M . In other
words, binary systems made of intermediate-mass compact objects will provide interesting constraints on the
TLNs, with k2E ' 1 and below, and therefore also on
the nature of these objects. The exquisite precision of
LISA can be traced back on the dependence of the tidal
deformability (which is the physical parameter entering
the waveform) on the ECOs mass, i.e. M 5 , which
amplifies the tidal effect on the GW signal. This is confirmed by the right-bottom panel of Fig. 7, where the
B.
Detectability of ECOs
Let us now turn our attention to some specific models and, in particular, to the models of ECOs investigated in the previous sections. Based on the previous discussion, as a general setup we consider equalmass binaries at distances d = 100 Mpc with M
[2, 30]M for AdLIGO/ET, and at d = 500 Mpc with
M [104 , 106 ]M for LISA. We note that the GW signal is proportional to 1/d, and therefore the covariance
matrix (68) (i.e., the error on ) scales linearly with the
distance.
In Fig. 8 we show the percentage relative errors /
for models of wormholes, perfect mirrors, and gravastars,
as a function of the mass of the object and for different
values of its compactness. Some qualitative results are independent of the nature of the ECO: the left panels confirm that AdLIGO would be able to constrain the tidal
deformability only for small values of the compactness,
namely C . 0.2. As the errors scale with the distance,
an upper bound / 1 for C = 0.3, would require a
source located at a distance 10 Mpc.
Furthermore, the relative errors decrease for larger
masses, reach a minimum, and then increase again. This
behavior can be explained by looking at the functional
form of Eq. (72). For a fixed compactness (i.e., for
fixed k2E ), the average tidal deformability grows with the
ECO mass, thus making the tidal part of the gravitational waveform more easy to be detected. However,
since the template is truncated at the last stable orbit, fmax m1 , increasing the mass also reduces the
number of effective cycles spent into the detectors bandwidth. This is particular penalizing for tidal effects,
which enter the GW signal as high-PN/high-frequency
15
50
10
5
1000
500
100
50
100
10
10
Gravastar
5 10 15 20 25 30 35 40
100
50
10
5
Perfect Mirror
1
0.5
5 10 15 20 25 30 35 40
M [M]
M [M]
Wormhole
=0
.49
=
0.4
=0
.3
=0
.2
=0
.1
Perfect Mirror
1000
100
10
1
0.10
0.01
Gravastar
1000
100
10
1
0.10
0.01
Perfect Mirror
1
0.1
= 0.1
Wormhole
1000
500
100
= 0.2
1
0.1
= 0.3
|/|[%]
10
5
10
|/|[%]
Wormhole
|/|[%]
100
50
0.1
100
|/|[%]
0.2
=
|/|[%]
= 0.3
|/|[%]
1000
500
|/|[%]
|/|[%]
|/|[%]
Gravastar
1000
100
10
1
0.10
0.01
1
5 10
[104
50 100
M ]
FIG. 8. Relative percentage errors on the tidal deformability for binaries observed by AdLIGO (left panels), ET (middle
panels), and LISA (right panels), as functions of the ECO mass and for different values of the compactness. For terrestrial
interferometers we consider prototype binaries at d = 100 Mpc, while for LISA we set the source at d = 500 Mpc. Top, middle
and bottom panels refer to wormholes, perfect-mirror models, and gravastars, respectively.
corrections.
It is worth noticing that a network of advanced interferometers would improve these results, even though it
will not drastically change the upper bound on the compactness of these objects. Indeed, if we consider that
the experiments are all independent, the Fisher matrices computed for each detector simply sum up, and the
overall error on is given by the inverse of the total
ab . Assuming five detectors with the same sensitivity
of AdLIGO,
the relative error would decrease roughly
by a factor 5 which, from our results in Fig. 8, is still
not enough to constrain objects much more compact than
C 0.2.
This scenario improves drastically for space-based detectors such as LISA (right panels of Fig. 8). Within the
considered mass range, tidal effects may be measured for
ECOs with C . 0.3 up to 1% of accuracy. Moreover,
LISA will be able to put strong constraints even for more
compact objects: for M & 105 M it would be possible
to set an upper bound / = 1 in the entire parameter space. As discussed in the previous section, these
results rely on the magnitude of the ECOs mass, which
strengthens the effect of tidal interactions in the waveform. The right panels of Fig. 8 show indeed that for all
the considered ECO models, LISA leads the analysis to
nearly explore the BH limit C 1/2.
16
this analysis.
C.
Detectability of BSs
Testing GR
Our results suggest that, in generic theories of gravity, the GW signal from a BH coalescence contains a
5PN term which depends on the TLNs. Although the
inclusion of tidal corrections in BH binaries is important
for a correct modeling of the waveform, in most theories this term is subleading relative to other corrections
coming both from dissipative effects (for example due
to scalar- or vector-wave emission) and from corrections
to the Hamiltonian of the binary [41, 42]. Nonetheless,
there are cases in which the tidal deformability found
in this work is the dominant correction relative to the
GR waveform. This is the case of Chern-Simons gravity,
where corrections to the GW phase enter at 2PN order
if the components of the binary are spinning, but only
at 7PN order in the absence of spin [91, 92]. Therefore,
for non-spinning binaries the tidal correction will be the
dominant one and it is interesting to estimate to which
level the Chern-Simons parameter can be constrained by
GW observations of BH binaries.
Unfortunately, the polar TLNs of Schwarzschild BHs
in Chern-Simons gravity are zero as in GR and the axial
The theory of the tidal deformability of compact objects has attracted considerable attention over the last
few years. So far, applications of this theory have been
mostly limited to astrophysics and to the possibility of
constraining the equation of state of NSs with GW observations. In this paper, we argued that tidal effects can
also be used to explore fundamental questions related to
the nature of event horizons, the existence of ECOs, and
the behavior of gravity in the strong-field regime.
Our main results can be summarized as follows:
In the framework of GR, the TLNs of ECOs are
generically non-zero. In the limit that the ECO
compactness C 1/2, all TLNs vanish but only
logarithmically. This result holds for all models of
ECO we have considered. It is therefore natural
to conjecture that this logarithmic dependence is a
generic feature of ultracompact exotic objects.9
The TLNs of a charged BH in Einstein-Maxwell
theory and of an uncharged static BH in BransDicke theory vanish, as in GR. These are both compelling extensions of GR, but our results indicate
that the TLNs of BHs are non-zero in other interesting extensions. In particular, we have explicitly shown that the axial TLNs of a Schwarzschild
17
order. A natural extension of our work is to include
tidal corrections also when they are subleading relative to other, point-particle, beyond-GR terms.
A related point is the fact that, when the TLNs
are small, their effect might be smaller than other
point-particle terms entering the waveform at 5PN
order. In this case, the full 5PN waveform might
be needed in order to extract the TLNs properly.10
ACKNOWLEDGMENTS
We thank Leonardo Gualtieri for interesting discussions. V.C. acknowledges financial support provided
under the European Unions H2020 ERC Consolidator
Grant Matter and strong-field gravity: New frontiers
in Einsteins theory grant agreement no. MaGRaTh
646597. Research at Perimeter Institute is supported
by the Government of Canada through Industry Canada
and by the Province of Ontario through the Ministry
of Economic Development & Innovation. This project
has received funding from the European Unions Horizon 2020 research and innovation programme under the
Marie Sklodowska-Curie grant agreement No 690904, the
NewCompstar COST action MP1304, and from FCTPortugal through the projects IF/00293/2013. The authors thankfully acknowledge the computer resources,
technical expertise and assistance provided by CENTRA/IST. Computations were performed at the clusters
Baltasar-Sete-Sois and Marenostrum, and supported
by the MaGRaTh646597 ERC Consolidator Grant.
10
18
P
ai (1/2 C)i and klB C 2l 4i=0 ai (1/2 C)i for the TLNs of a static neutron
star with a stiff (MS1 [93]) and relatively softer (SLy4 [94]) equation of state. Data are taken from Ref. [21] and agree with the results in
Ref. [8] after using the conversion in Eq. (4).
EOS
P4
i=0
a0
a1
a2
a3
a4
MS1
k2E
k3E
k2M
k3M
-0.581
-0.207
-0.096
-0.034
8.721
3.230
1.369
0.514
-48.69
-18.66
-7.099
-2.910
123.4
47.82
17.53
7.511
-112.8
-43.87
-16.11
-7.009
SLy4
k2E
k3E
k2M
k3M
-0.414
-0.150
-0.063
-0.023
6.227
2.326
0.876
0.346
-35.35
-13.57
-4.483
-1.978
92.70
35.56
11.57
5.288
-87.70
-33.38
-11.24
-5.115
g = g + h ,
(B1)
(0)
(0)
g = diag e , eg , r2 , r2 sin2 .
We decompose h in spherical harmonics and separate the perturbation in even and odd parts, h =
odd
heven
+ h , according to parity. In the Regge-Wheeler
gauge [95], h can be decomposed as
1
2
=
2
lm
lm
0
0
r K (t, r)Y
0
2
2
lm
0
0
0
r sin K (t, r)Y lm
heven
lm
lm
hlm
0
0
hlm
0 (t, r)S
0 (t, r)S
lm
lm
0
0
hlm
hlm
1 (t, r)S
1 (t, r)S
= lm
lm
lm
lm
h0 (t, r)S h1 (t, r)S
0
0
lm
lm
lm
lm
0
0
h0 (t, r)S h1 (t, r)S
hodd
lm
with Slm , Slm Y,
/ sin , sin Y,lm .
In the presence of scalars or vectors, spacetime fluctuations are accompanied by the corresponding fluctuations
in these fields,
(0)
A = A + A ,
(0)
=
(0)
+ ,
(B5)
(B6)
(0)
(B8)
(B2)
(B3)
(B4)
lm
and = l(l + 1). Hereafter we
with Yblm Y,lm , Y,
shall drop the (lm) superscripts on all quantities with the
exception of multipole moments.
Expressions for the metric functions in (B3) and (B4),
the electromagnetic functions in (B8), and the scalar
fields can be obtained by solving the linearized field equations for a given model. The remaining task consists in
extracting the multipole moments and tidal fields from
the spacetime metric. Thorne developed a method to
define the multipole coefficients of any spacetime metric given in coordinates which are asymptotically Cartesian and mass centered (ACMC) [65]. Another definition of the multipole moments of an axisymmetric and
asymptotically flat spacetime was given by Geroch and
Hansen [62, 63]. These two distinct definitions of moments were shown to be equivalent [64].
19
The multipole moments can be extracted from the
asymptotic behavior of the spacetime metric and fields,
#
!
2
4
Ml Y l0 + (l0 < l pole)
rl El Y l0 + (l0 < l pole) ,
rl+1
2l + 1
l(l 1)
l2
"r
!
#
l+1
X
2J
2
4
S
2r
l
gt =
sin2 +
S l0 + (l0 < l pole) +
Bl Sl0 + (l0 < l pole) ,
r
rl
2l + 1 l
3l (l 1)
l2
"r
#
!
X
2
4
2
Q
l0
0
l
l0
0
At = +
Ql Y + (l < l pole)
r El Y + (l < l pole) ,
r
rl+1
2l + 1
l(l 1)
l1
"r
#
!
X 2
4 Jl l0
2rl+1
0
l0
0
A =
S + (l < l pole) +
Bl S + (l < l pole) ,
rl
2l + 1 l
3l (l 1)
l1
X 1
S
l0
0
l
0
= 0 +
l Y + (l < l pole) r El + (l < l pole) .
rl+1
2M X
+
gtt = 1 +
r
"r
(B9)
(B10)
(B11)
(B12)
(B13)
l1
Background solutions
We consider spherically symmetric BSs, with background metric given by Eq. (B2), and the following
ansatz for the background scalar field,
(0)
(C1)
00 ,
d|2 |
2
r
(C4)
0g =
value c of the scalar at the center of the star, the problem is then reduced to an eigenvalue problem for the frequency , which we solve through a standard shooting
method. The value of c is arbitrary and can be tuned
in order to have (r ) = 0.
The total mass of the solution is M = m(r ),
where m(r) is defined by
eg (r) 1
2m(r)
.
r
(C5)
20
FIG. 9. Left panel: ADM mass as a function of the effective radius R for different models of BSs, including some unstable
configurations (to the left of maximum mass). Right panel: Compactness of the models in the stable branch as a function of
the mass. For massive and solitonic BSs we have considered = 104 2 and 0 = 0.05, respectively.
left panel in Fig. 9), an unstable branch that starts after the first maximum roughly at R 50, and then a
second stable branch which starts at R 10 up to the
maximum on the top-right part of the plot.
2.
Polar perturbations
We consider perturbations of the equilibrium configuration, sourced by an external static tidal field. The
metric perturbation is given by (B3) with H1 = 0. We
write the scalar perturbation as
X
=
eit 1 (r)Y lm (, ).
(C6)
m
0
g + 0
2
2
2 2 g
0
02
0
00
0
2
g
H +
8r 0 e
+ 8r0 H + 32 0
0 0 e
1
r
2
r
20
eg (l2 + l + 2) 2
48eg 2 20 + 1602
+
H=0
+ 02
0
r
r2
0
g + 0
2
2
0
00
0
2
g
001 +
8r 2 20 eg + 0 8r02
e
H
0
0
1
0
0
r
2
r
00 0
0g
2 00
l(l + 1)eg
dV g
0 +
+
+ 3202
+
1
e 0 = 0
0
0
2
r 0
r2
d|2 |
00
We now solve the perturbation system supplied by regular boundary conditions at the origin,
(l)
(l)
H0 H0 rl + O rl+2 , 1 1 rl + O rl+2 . (C9)
(l)
(C7)
(C8)
21
radius, Eq. (C7) reduces to
H 00 +
2(r M ) 0 4M 2 2M r + r2
H = 0 , (C10)
H
r(r 2M )
r2 (r 2M )2
8
2
2
(1 2C) [2C(y 1) y + 2] 3(1 2C) [2C(y 1) y + 2] log (1 2C)
5
1
+ 2C 4C 4 (y + 1) + 2C 3 (3y 3) + 2C 2 (13 11y) + 3C(5y 8) 3y + 6
,
(C11)
6
2
2
k3E =
15(1 2C) 2C 2 (y 1) 3C(y 2) + y 3 log(1 2C)
(1 2C) 2C 2 (y 1) 3C(y 2) + y 3
7
1
+ 2C 4C 5 (y + 1) + 2C 4 (9y 2) 20C 3 (7y 9) + 5C 2 (37y 72) 45C(2y 5) + 15(y 3)
.
(C12)
k2E =
h0 h0
+ O rl+3 .
(C14)
(l+1)
b.
Axial perturbations
k2B =
k3B =
4M l(l + 1)r
h0 = 0,
r2 (r 2M )
8
2C(y 2) y + 3
,
5 2C [2C 3 (y + 1) + 2C 2 y + 3C(y 1) 3y + 9] + 3[2C(y 2) y + 3] log(1 2C)
8 2
8C (y 2) 10C(y 3) + 3(y 4) 15 8C 2 (y 2) 10C(y 3) + 3(y 4) log(1 2C)
7
1
+ 2C 4C 4 (y + 1) + 10C 3 y + 30C 2 (y 1) 15C(7y 18) + 45(y 4)
,
where again C = M/Rext but now y = rh00 /h0 evaluated at Rext . Even in this case, the values of klB are
independent of the extraction radius Rext if the latter is
sufficiently large.
(C15)
(C16)
(C17)
22
1.
Exterior spacetime
Let us first consider the exterior spacetime. Einsteins equations for static polar-type perturbations of
the Schwarzschild metric lead to (the notation follows
Appendix B)
2f dH0
l(l + 1) 4M 2
d 2 H0
+
+ 4
H0 = 0 ,(D1)
f
dr2
r dr
r2
r
dK
f 1
=
H0 + H00 ,
(D2)
dr
rf
[1 + f (l2 + l 2 f )]H0 rf (f 1)H00
. (D3)
K=
f (l2 + l 2)
Here f = 1 2M/r and primes stand for derivatives with
respect to r. The above equations can be solved for [6, 8]
H0ext = C1 Pl2 (r/M 1) + C2 Q2l (r/M 1) ,
(D4)
for any value of l, and where C1 and C2 are two integration constants. The term proportional to C1 diverges at large distances and is identified with the external tidal field, whereas the term proportional to C2
is the bodys response. The metric function K follows
straightforwardly from Eq. (D3).
b.
Interior spacetime
(D5)
c.
[[K]] = 0 ,
[[dK/dr ]] = 0 .
(D7)
These two conditions completely specify the matching between the interior and the exterior solution in the wormhole and gravastar cases. In the latter case, the junction
conditions (D7) agree with those derived in Ref. [59].
In the perfect-mirror case, we shall impose a Z2 symmetry on the surface and, therefore, the wavefunction
vanishes at r = r0 . In this case, one can solve for the
Zerilli function Z in the static limit and then reconstruct the metric function H0 through [103]
H0 =
1
l r2 6M 2 3l M r 0Z (r)
r(6M + l r)
Z (r)
36M 3 /r + 18l M 2 + 3l 2 M r
+
6M + l r
+l 2 (l /2 + 1)r2 ,
(D8)
23
2.
Exterior spacetime
DA h0 h000 +
4M l(l + 1)r
h0 = 0 ,
r2 (r 2M )
(D9)
c.
(D10)
where G2,0
2,2 is the Meijer function and 2 F1 is one of the
hypergeometric functions. The terms proportional to c1
and c2 are identified with the external tidal field and
with the body response, respectively. The above solution
reduces to simple expressions for integer values of l, which
can be written in terms of polynomial and logarithmic
functions.
Interior spacetime
hint
0
c2 G2,0
2,2
r
2M
!
1 l, l + 2
.
1, 2
(D11)
11
Note that our definition differs from the standard one by a factor
, which has been included so that h1 0 in the static limit,
whereas h0 remains finite, as expected.
The junction conditions for axial perturbations are easier because they do not couple to the matter of a putative thin shell. In the dynamical case, they simply read
[[h0 ]] = 0 = [[h1 ]] [38]. In the static case, h1 vanishes
identically and one is left with a single second-order differential equation for h0 . Therefore, regularity of the
axial perturbations across the shell imposes that h0 and
its derivative with respect to r be smooth. Thus, for
the wormhole and gravastar cases in the axial sector we
impose
[[h0 ]] = 0 ,
[[dh0 /dr ]] = 0 .
(D13)
For the perfect-mirror model, we follow the same procedure outlined above, namely we impose a Dirichlet condition on the Regge-Wheeler function RW evaluated at
r = r0 . This function is defined as11
h0 =
b.
(D12)
d(rRW )
,
dr
h1 =
ir
RW , (D14)
1 2M/r
and satisfies the Regge-Wheeler equation [95]. The latter can be solved analytically in the static limit, = 0.
Again, the ratio of the two integration constants in
Eq. (D10) is fixed by imposing RW = 0.
After the perturbations are fully specified through the
junction/boundary conditions, the axial TLNs can be
computed by comparing the large-distance behavior of
h0 with Eq. (B10), extracting the multipole moments,
and finally using the definition (3). As in the polar case,
we find a closed-form, cumbersome expression for the axial TLN [57], whose high-compactness limit is provided
in the main text for the various models.
[4] E. Poisson and C. Will, Gravity: Newtonian, PostNewtonian, Relativistic (Cambridge University Press,
Cambridge, UK, 2014).
[5] E.E. Flanagan and T. Hinderer, Phys. Rev. D77,
021502 (2008), arXiv:0709.1915.
[6] T. Hinderer, Astrophys. J. 677, 1216 (2008), Erratum:
ibid. 697, 964 (2009), arXiv:0711.2420.
[7] T. Hinderer et al., Phys. Rev. Lett. 116, 181101 (2016),
arXiv:1602.00599.
[8] T. Binnington and E. Poisson, Phys. Rev. D80, 084018
(2009), arXiv:0906.1366.
24
[9] T. Damour and A. Nagar, Phys. Rev. D80, 084035
(2009), arXiv:0906.0096.
[10] J. Lattimer and M. Prakash, Science 304, 536 (2004),
arXiv:astro-ph/0405262.
[11] T. Hinderer, B.D. Lackey, R.N. Lang, and J.S. Read,
Phys. Rev. D81, 123016 (2010), arXiv:0911.3535.
[12] S. Postnikov, M. Prakash, and J.M. Lattimer, Phys.
Rev. D82, 024016 (2010), arXiv:1004.5098.
[13] J. Vines, E.E. Flanagan, and T. Hinderer, Phys. Rev.
D83, 084051 (2011), arXiv:1101.1673.
[14] T. Damour, A. Nagar, and L. Villain, Phys. Rev. D85,
123007 (2012), arXiv:1203.4352.
[15] W. Del Pozzo, T.G.F. Li, M. Agathos, C. Van
Den Broeck, and S. Vitale, Phys. Rev. Lett. 111, 071101
(2013), arXiv:1307.8338.
[16] A. Maselli, L. Gualtieri, and V. Ferrari, Phys. Rev. D88,
104040 (2013), arXiv:1310.5381.
[17] N. Sennett et al., (in preparation, 2017).
[18] H. Fang and G. Lovelace, Phys. Rev. D72, 124016
(2005), arXiv:gr-qc/0505156.
[19] N. G
urlebeck, Phys. Rev. Lett. 114, 151102 (2015),
arXiv:1503.03240.
[20] E. Poisson, Phys. Rev. D91, 044004 (2015),
arXiv:1411.4711.
[21] P. Pani, L. Gualtieri, A. Maselli, and V. Ferrari, Phys.
Rev. D92, 024010 (2015), arXiv:1503.07365.
[22] P. Landry and E. Poisson, Phys. Rev. D91, 104018
(2015), arXiv:1503.07366.
[23] R.A. Porto, Fortsch. Phys. 64, 723 (2016),
arXiv:1606.08895.
[24] S.D. Mathur, Proceedings, CERN Winter School on
Strings, Supergravity and Gauge Theories. Geneva,
Switzerland, February 913 2009, Class. Quantum
Grav. 26, 224001 (2009), arXiv:0909.1038.
[25] F.E. Schunck and E.W. Mielke, Class. Quantum Grav.
20, R301 (2003), arXiv:0801.0307.
[26] S.L. Liebling and C. Palenzuela, Living Rev. Relat. 15,
6 (2012), arXiv:1202.5809.
[27] R. Brito, V. Cardoso, C.F.B. Macedo, H. Okawa,
and C. Palenzuela, Phys. Rev. D93, 044045 (2016),
arXiv:1512.00466.
[28] P. Grandclement, (2016), arXiv:1612.07507.
[29] P.O. Mazur and E. Mottola,
(2001), arXiv:grqc/0109035.
[30] M. Visser and D.L. Wiltshire, Class. Quantum Grav.
21, 1135 (2004), arXiv:gr-qc/0310107.
[31] M. Visser, Lorentzian wormholes: from Einstein to
Hawking (AIP, Woodbury, NY, USA, 1996).
[32] E.G. Gimon and P. Horava, Phys. Lett. B672, 299
(2009), arXiv:0706.2873.
[33] K. Skenderis and M. Taylor, Phys. Rept. 467, 117
(2008), arXiv:0804.0552.
[34] B. Holdom and J. Ren, (2016), arXiv:1612.04889.
[35] M. Saravani, N. Afshordi, and R.B. Mann, Int. J. Mod.
Phys. D23, 1443007 (2015), arXiv:1212.4176.
[36] S.B. Giddings, Phys. Rev. D90, 124033 (2014),
arXiv:1406.7001.
[37] J. Abedi, H. Dykaar, and N. Afshordi,
(2016),
arXiv:1612.00266.
[38] P. Pani, E. Berti, V. Cardoso, Y. Chen, and R. Norte,
Phys. Rev. D80, 124047 (2009), arXiv:0909.0287.
[39] C.F.B. Macedo, P. Pani, V. Cardoso, and
L.C.B. Crispino, Phys. Rev. D88, 064046 (2013),
arXiv:1307.4812.
25
[72] M. Colpi, S.L. Shapiro, and I. Wasserman, Phys. Rev.
Lett. 57, 2485 (1986).
[73] R. Friedberg, T.D. Lee, and Y. Pang, Phys. Rev. D35,
3658 (1987).
[74] R.F.P. Mendes and H. Yang, (2016), arXiv:1606.03035.
[75] N. Uchikata and S. Yoshida, Class. Quantum Grav. 33,
025005 (2016), arXiv:1506.06485.
[76] R. Jackiw and S.Y. Pi, Phys. Rev. D68, 104012 (2003),
arXiv:gr-qc/0308071.
[77] N. Yunes and C.F. Sopuerta, Phys. Rev. D77, 064007
(2008), arXiv:0712.1028.
[78] C. Molina, P. Pani, V. Cardoso, and L. Gualtieri, Phys.
Rev. D81, 124021 (2010), arXiv:1004.4007.
[79] V. Cardoso and L. Gualtieri, Phys. Rev. D80,
064008 (2009), Erratum: ibid. D81, 089903 (2010),
arXiv:0907.5008.
[80] V. Cardoso, E. Franzin, A. Maselli, P. Pani, and G. Raposo, In preparation (2017).
[81] M. Vallisneri, Phys. Rev. D77, 042001 (2008), arXiv:grqc/0703086.
[82] C. Cutler and E.E. Flanagan, Phys. Rev. D49, 2658
(1994).
[83] T. Damour, B.R. Iyer, and B.S. Sathyaprakash, Phys.
Rev. D62, 084036 (2000).
[84] J. Vines, E.E. Flanagan, and T. Hinderer, Phys. Rev.
D83, 084051 (2011).
[85] D. Bini, T. Damour, and G. Faye, Phys. Rev. D85,
124034 (2012).
[86] D. Shoemaker (LIGO), Advanced LIGO anticipated sensitivity curves, Tech. Rep. T0900288-v3 (2010).
[87] B.S. Sathyaprakash and B.F. Schutz, Living Rev. Relat.
12, 2 (2009), arXiv:0903.0338.
[88] A. Klein, E. Barausse, A. Sesana, A. Petiteau, E. Berti,
[89]
[90]
[91]
[92]
[93]
[94]
[95]
[96]
[97]
[98]
[99]
[100]
[101]
[102]
[103]