You are on page 1of 12

Home Search Collections Journals About Contact us My IOPscience

A uniform treatment of the orbital effects due to a violation of the strong equivalence principle

in the gravitational Stark-like limit

This content has been downloaded from IOPscience. Please scroll down to see the full text.

2013 Class. Quantum Grav. 30 025006

(http://iopscience.iop.org/0264-9381/30/2/025006)

View the table of contents for this issue, or go to the journal homepage for more

Download details:

IP Address: 129.173.72.87
This content was downloaded on 18/05/2015 at 20:32

Please note that terms and conditions apply.


IOP PUBLISHING CLASSICAL AND QUANTUM GRAVITY
Class. Quantum Grav. 30 (2013) 025006 (11pp) doi:10.1088/0264-9381/30/2/025006

A uniform treatment of the orbital effects due to a


violation of the strong equivalence principle in the
gravitational Stark-like limit

L Iorio 1,2
Ministero dell’Istruzione, dell’Università e della Ricerca (MIUR)-Istruzione, Italy

E-mail: lorenzo.iorio@libero.it

Received 30 July 2012, in final form 2 November 2012


Published 14 December 2012
Online at stacks.iop.org/CQG/30/025006

Abstract
We consider a binary system made of self-gravitating bodies embedded in a
constant and uniform external field g. We analytically work out several orbital
effects induced by a putative violation of the Strong Equivalence Principle
(SEP) due to g. In our calculation, we do not assume e ∼ 0, where e is
the binary’s orbit eccentricity. Moreover, we do not a priori choose any
specific preferred spatial orientation for the fixed direction of g. Our results
do not depend on any particular SEP-violating theoretical scheme. They can be
applied to general astronomical and astrophysical binary systems immersed in
an external constant and uniform polarizing field.

PACS numbers: 04.80.−y, 04.80.Cc

1. Introduction

If bodies with non-negligible gravitational binding self-energies, like astronomical and


astrophysical objects (planets and their natural satellites, main-sequence stars, white dwarfs,
neutron stars), moved with different accelerations in a given external field, the so-called strong
equivalence principle (SEP) would be violated. While the general theory of relativity assumes
the validity of SEP, it is generally violated in alternative theories of gravity; see [1, 2] and
references therein.
Potential SEP violations in the Earth–Moon system freely falling in the external
gravitational field of the Sun, theoretically predicted by Nordtvedt long ago [3, 4], are
currently searched for with the lunar laser ranging (LLR) technique [5–7] in the framework
1 Fellow of the Royal Astronomical Society (FRAS); International Institute for Theoretical Physics and Advanced

Mathematics Einstein–Galilei; International Astronomical Union.


2 Permanent address: Viale Unit’a di Italia 68, I-70125 Bari (BA), Italy.

0264-9381/13/025006+11$33.00 © 2013 IOP Publishing Ltd Printed in the UK & the USA 1
Class. Quantum Grav. 30 (2013) 025006 L Iorio

of the parametrized post–Newtonian (PPN) formalism. There are projects to test SEP with the
Martian moon Phobos and the planetary laser ranging technique [8] as well.
As shown by Damour and Schäfer [9], the acceleration due to SEP violation occurring
for a two-body system in an external gravitational field g like, e.g., the Galactic field at the
location of a binary pulsar, is, to leading order,
Ag = (p − c )g, (1)
where p and c denote the pulsar and its less compact companion; for a generic body with
non-negligible gravitational self-energy3 Egrav ,  accounts for the SEP-violating difference
between the inertial mass mi and the gravitational mass mg: at first post–Newtonian level, it is
. mg . Egrav
= − 1 = η1 εgrav , εgrav = , (2)
mi mi c2
where c is the speed of light in vacuum. The external field g in equation (1) is assumed here to
be constant and uniform, so that it induces a gravitational analogue of the Stark effect [9]. To
maximize the size of potential SEP violations, one should look at compact objects like neutron
stars: indeed, it is
 (       
ε  ∼ 1.9 × 10−11 , ε⊕  ∼ 4.2 × 10−10 , ε  ∼ 1.3 × 10−6 , εNS  ∼ 10−1 . (3)
grav grav grav grav

The SEP-violating parameter η1 can be expressed in various ways depending on the theoretical
framework adopted; see [1, 2] for recent overviews. As far as the Earth–Moon system and
the PPN framework is concerned, latest published bounds ([6] and [7], respectively) on the
Nordtvedt parameter are
η1 ≡ ηN = (4.0 ± 4.3) × 10−4 , (4)

η1 ≡ ηN = (−0.6 ± 5.2) × 10−4 . (5)


Williams et al in a recent analysis [13] released
η1 ≡ ηN = (−1.8 ± 2.9) × 10−4 . (6)
In this paper, we deal with the orbital effects due to equation (1) in a uniform way by obtaining
explicit and transparent analytical expressions valid for the main observable quantities
routinely used in empirical studies. In section 2, we work out the SEP-violating rates of change
of the osculating Keplerian orbital elements, which are commonly used in solar system and
binaries studies. The SEP-violating shift of the projection of the binary’s orbit onto the line of
sight, which is the basic observable in pulsar timing, is treated in section 3. In section 4, we
calculate the SEP-violating perturbation of the radial velocity, which is one of the standard
observable in spectroscopic studies of binaries: in systems suitable for testing SEP, the pulsar’s
companion is a white dwarf or a main sequence star for which radial velocity curves may be
obtained as well. The primary-to-companion range and range-rate SEP-violating perturbations,
potentially occurring in systems like the Earth–Moon one, are calculated in section 5. Section 6
is devoted to summarizing and discussing our results. We will not resort to a priori simplifying
assumptions about the binary’s orbital eccentricity. Moreover, we will not restrict ourselves to
any specific spatial orientation for g. Our results are, thus, quite general. They can be used for
better understanding and interpreting present and future SEP experiments, hopefully helping
in designing new tests as well. Finally, we remark that our calculations are model-independent
in the sense that they do not depend on any specific theoretical mechanism yielding SEP
3 For a non-relativistic spherical body of uniform density, it can be cast [10, 11] E
grav = 3M G/5R, where G is the
2

Newtonian constant of gravitation, M is the body’s mass and R is its radius. For highly relativistic neutron stars, see,
e.g., [12].

2
Class. Quantum Grav. 30 (2013) 025006 L Iorio

violations. Thus, they may be useful when different concurring SEP-violating scenarios are
considered like, e.g., MOND and its external field effect [14, 15], Galileon-based theories with
the Vainshtein mechanism [16, 17] and variations of fundamental coupling constants as well
[18–20]. Moreover, there are also other effects, like Lorentz symmetry violations [21], which
affect the same SEP-violating peculiar observables like the eccentricity [22]. For a seminal
paper on Lorentz-violating orbital effects in binaries, see [23].

2. The long-term rates of change of the osculating Keplerian orbital elements

The SEP-violating acceleration of equation (1) is much smaller that the standard two-body
Newtonian and post–Newtonian ones; thus, its orbital effects can be treated perturbatively. It
can be thought as due to the following perturbing potential:
Ug =  g·r, (7)
where
.
 = c − p : (8)
r is the relative position vector of the binary.
The Lagrange planetary equations [24] allow to work out all the orbital effects caused by
equation (7) in an effective way which encompasses just one integration.
Damour and Schäfer [9] used a different formalism and an approach requiring six
independent4 integrations. The work by Damour and Schäfer [9] is important also for the
relevance for astrophysical tests found in the literature. Freire et al [2] recently used the same
formalism as Damour and Schäfer [9].
As a first step, the perturbing potential of equation (7) must be evaluated onto a reference
trajectory assumed as unperturbed with respect to the effect we are interested in. To this aim,
we will adopt the standard Keplerian ellipse, although it would be possible, in principle, to
use a post–Newtonian orbit as well [25, 26]. We do not use such a relativistic unperturbed
path as a reference trajectory. Indeed, in addition to the usual 1PN pericenter precession and
the SEP-violating/0PN effects which will turn out to be per se small, it would only yield
additional SEP-violating/1PN mixed terms. It must be stressed that, strictly speaking, the
previous considerations—and the consequent calculations that we will show below—hold
only in the case of a SEP violation due to an external polarizing field g. In fact, a SEP-violating
gravity theory is expected to modify the 1PN pericenter precession rate even if such an external
field is absent; it is accounted for by the multiplicative factor F put by Damour and Schäfer
[9] in front of the standard Einsteinian 1PN pericenter precession, where F depends on the
binary’s masses in alternative theories. Moving to our approach, an average of equation (7)
over one full orbital period of the binary must be taken; by using the eccentric anomaly E as
fast variable of integration, the result is
3aeg
Ug = − {cos ω(ĝx cos  + ĝy sin ) + sin ω[ĝz sin I + cos I(ĝy cos  − ĝx sin )]},
2
(9)
where g = |g| , ĝ = g/g, and a, e, I, ω,  are the semimajor axis, the eccentricity, the
inclination, the argument of pericenter and the longitude of the ascending node of the binary
orbit referred to a generic coordinate system. As far as binary pulsar systems are concerned,
the reference {X, Y } plane is the plane of the sky tangent to the celestial sphere at the location
4 Averages of the time derivatives of the orbital energy, the orbital angular momentum and the Laplace–Runge–Lenz

vector were taken in [9]. The orbital angular momentum and the Laplace–Runge–Lenz vector are perpendicular.

3
Class. Quantum Grav. 30 (2013) 025006 L Iorio

of the system considered. The inclination I of the binary’s orbital plane refers just to the plane
of the sky, so that the reference Z axis is directed from the solar system’s barycenter to the
binary along the line-of-sight direction; the reference Y axis is usually directed toward the
North Celestial Pole, and the reference X axis, from which the node  is counted, is in the east
direction. The unit vectors I0 , J0 , K 0 of such a coordinate system can be expressed in terms of
the celestial right ascension α and declination δ of the binary as [27–29]

⎨− sin α,
I0 = cos α, (10)

0,

⎨− sin δ cos α,
J0 = − sin δ sin α, (11)

cos δ,

⎨cos δ cos α,
K 0 = cos δ sin α, (12)

sin δ.
Instead, in studies pertaining our solar system the reference {X, Y } plane is customarily chosen
as coincident with the mean equator at J2000.0, so that the reference X axis is directed toward
the Vernal Equinox  at J2000.0 and the Z axis points toward the North Celestial Pole. In
obtaining equation (9), we have reasonably assumed that g does not vary over a full orbital
revolution; moreover, we did not make any a priori simplifying assumption about the spatial
orientation of g which, in general, will not be directed along any particular direction of the
coordinate system adopted. Nonetheless, the SEP-violating field g can reasonably be assumed
to be approximately directed toward the Galactic Center (GC), whose Celestial coordinates
are [30]
αGC = 17h 45m 37s .224 = 266.4051 deg, (13)

δGC = −28◦ 56 10 .23 = −28.936175 deg. (14)
Thus, for a Sun–planet pair in the Solar System it is

⎨cos δGC cos αGC ,
ĝ ≡ĝSS = cos δGC sin αGC (15)

sin δGC .
The Lagrange planetary equations [24] and equation (9) straightforwardly yield the
following long-term rates of changes of the six standard osculating Keplerian orbital elements5,
and of the longitude of the pericenter6
as well
 
da
= 0, (16)
dt
  √
de 3g 1 − e2
=− [ĝz sin I cos ω + cos I cos ω(ĝy cos  − ĝx sin ) −
dt 2nb a
− sin ω(ĝx cos  + ĝy sin )], (17)
. . 
5 M = nb (t − t0 ) is the mean anomaly, where t0 is the time of passage at pericenter and nb = GM/a3 is
the Keplerian mean motion. In general, SEP-violating theories induce modifications of the (effective) gravitational
coupling parameter G in the binary interactions (cf [9]); here we will neglect them.
.
6 It is a ‘dogleg’ angle defined as
=  + ω. It is used in some specific cases such as, e.g., analyses of planetary
motions in our solar system.

4
Class. Quantum Grav. 30 (2013) 025006 L Iorio

 
dI 3ge cos ω[ĝz cos I + sin I(ĝx sin  − ĝy cos )]
= √ , (18)
dt 2nb a 1 − e2
 
d 3ge csc I sin ω[ĝz cos I + sin I(ĝx sin  − ĝy cos )]
= √ , (19)
dt 2nb a 1 − e2
 
dω 3g
= √ {(−1 + e2 ) cos ω(ĝx cos  + ĝy sin )
dt 2enb a 1 − e 2

+ sin ω[−ĝy cos I cos  + ĝz (e2 csc I − sin I) + ĝx cos I sin ]}, (20)

 
d
3g
= √ (1 − e2 ) cos ω(ĝx cos  + ĝy sin )
dt 2enb a 1 − e2


I
+ sin ω (cos I − e2 )(ĝy cos  − ĝx sin ) + ĝz (1 − e2 + cos I) tan ,
2
(21)
 
dM 3g(1 + e2 )
− nb = − {cos ω(ĝx cos  + ĝy sin )
dt 2enb a
+ sin ω[ĝz sin I + cos I(ĝy cos  − ĝx sin )]}. (22)
We recall that, in obtaining the rate of the mean anomaly perturbatively, the mean motion
nb has to be considered constant; thus, equation (22) is proportional to the rate of the time
of pericenter passage. About the pericenter precession, it must recalled that, in addition to
the effect of equation (21) due to the external polarizing field g, a further SEP-violating
rate of change is, in principle, present. It is formally equal to the usual 1PN relativistic rate
re-scaled by a multiplicative factor F which may depend on the masses of the binary in
SEP-violating alternative theories of gravity [9]. In obtaining equations (16)–(22), we did not
make any a priori simplifying assumption on the orbital configuration, i.e. we did not assume
e → 0, I → 0. In the limit e → 0, the orbital plane remains unchanged, as found by Damour
and Schäfer [9], since the long-term changes of both the inclination I and the node  vanish.
Instead, the rate of change of the eccentricity e remains substantially unaffected since it is not
of order O(e). In the limit of small eccentricities and small inclinations, the mean longitude
.
l =
+ M is customarily used. From equations (21)–(22), it turns out that its rate is of order
O(e):
 

dl 3eg
− nb = − 3 cos ω(ĝx cos  + ĝy sin ) + sin ω ĝy (2 + cos I) cos 
dt 4nb a

I
− ĝx (2 + cos I) sin  + ĝz sin I + 2ĝz tan . (23)
2
Thus, in the small eccentricity limit, all the osculating Keplerian orbital elements remain
unaffected by a SEP violation, apart from the eccentricity. Note that equations (16)–(22) and
(23) are all proportional to n−1 b a
−1
= a1/2 , so that they are larger for detached binaries. For
large eccentricities the node and the inclination variations can, in principle, be used as well as
further observables. Interestingly, equations (18)–(19) show that their ratio is independent of
both g and ĝ; moreover, it depends neither on a nor on e, nor on  amounting to
˙

= csc I tan ω. (24)
˙
I
5
Class. Quantum Grav. 30 (2013) 025006 L Iorio

For systems exhibiting large inclinations, equation (24) may have relevant observational
implications, provided that both the inclination and the node are accessible to observations.
Also the ratio of equation (18) and equation (17) could be considered [2]; however, if on the
one hand the ratio of the rates of the inclination and the eccentricity does not depend on g and
a, on the other hand it depends on ĝ and on e, I, , ω as well, contrary to equation (24) which
is simpler.
Damour and Schäfer [9] define the following vector directed along the external field7
3A
f= . (25)
2nb a
Then, they use the unit vector â directed along the semimajor axis toward the pericenter, the
unit vector k̂ directed along the orbital angular momentum and b̂ = k̂ × â. Their expression
for the averaged rate of change of the eccentricity is [9]
  
de
= 1 − e2 (b̂ · f ). (26)
dt
A direct comparison of equation (26) with our equation (17) can actually be performed by
expressing the unit vectors k̂ and â in terms of the Keplerian orbital elements as

⎨sin I sin ,
k̂ = − sin I cos , (27)

cos I,
and ⎧
⎨cos  cos ω − cos I sin  sin ω,
â = sin  cos ω + cos I cos  sin ω, (28)

sin I sin ω.
By using equations (27)–(28), it turns out that equation (26) coincides with our equation (17).
As far as the inclination I is concerned, Freire et al [2], by adopting the same formalism
of Damour and Schäfer [9], obtain
d cos I e
=√ (K̂ 0 · b̂)(k̂ · fˆ ), (29)
dt 1 − e2
where K̂ 0 is the unit vector of the line of sight. Also in this case, with the aid of equations
(27)–(28) it can be shown that equation (29) agrees with our equation (18).
Neither Damour and Schäfer [9] nor Freire et al [2] explicitly considered the SEP-violating
rates of change of the node (equation (19)), the pericenter (equations (20)–(21)) and the time
of pericenter passage (equation (22)).
In binaries hosting a radiopulsar, the rate of change of the inclination can be expressed in
.
terms of the rate of change of the pulsar’s projected semimajor axis xp = ap sin I/c, where c
is the speed of light in vacuum, as I˙ = (ẋp /xp ) tan I. About the measurability of the node 
in wide binaries hosting a radiopulsar, it could be determined only if they are close enough to
the Earth. Indeed, in this case the orbital motion of the Earth changes the apparent inclination
angle I of the pulsar orbit on the sky, an effect known as the annual-orbital parallax [27]. It
causes a periodic change of the projected semi-major axis. There is also a second contribution
due to the transverse motion in the plane of the sky [28], yielding a secular variation of the
projected semi-major axis. By including both these effects in the model of the pulse arrival
times,  can be determined, as in the case of PSR J0437-4715 [29], located at only 140 pc
from us. Interestingly, such an approach may also be used in optical spectral observations
of binary stars possessing a sufficiently well-determined radial velocity curve [28]. As far as
7 Note that the quantity f has the physical dimensions of reciprocal time, i.e. [ f ] = T−1 .

6
Class. Quantum Grav. 30 (2013) 025006 L Iorio

exoplanets are concerned, spectroscopic variations of the hosting star during the transits can,
in principle, be used to measure  [31]. According to Fluri and Berdyugina [32], the analysis
of the polarization of the light scattered by the planetary atmospheres may allow to determine,
among other things,  as well.
The SEP-violating orbital rates of equations (16)–(23) can be defined as long-term effects
since they were obtained by taking averages over one full orbital period Pb of the binary
system. It must be noted that, strictly speaking, they cannot be considered as secular rates.
Indeed, in calculating them, it was assumed that the external field g was constant over Pb . The
extent to which such an approximation can be considered valid depends on the specific system
considered. It is certainly true for binary pulsar systems, for which timing data span some
decades at most, when g is due to the Galaxy: in this case, g can reasonably be considered as
constant over any foreseeable data analysis based on existing (and future) pulsar timing records
because of the extremely low variation of the Galactic field at the pulsar’s location during such

time spans. If, instead, g is due to a remote third body with a small orbital frequency nb with
respect to nb , a relatively low modulation may be introduced by a slowly varying g(t ). A quite
different situation occurs for systems like the Earth–Moon one whose external field g is due

to a relatively close body like the Sun having an orbital frequency nb comparable to nb . In this
case, g cannot be considered as constant, and a second integration has to be performed over

the orbital period Pb of the third body.
In equations (16)–(23) the Keplerian orbital elements I, , ω determining the orientation
of the binary orbit in space are present. They are fixed only in the two-body Keplerian problem
for two pointlike masses. In any realistic situation, they are slowly varying with frequencies
typically far smaller than nb . It is true even if no other bodies are part of the system because
of general relativity itself causing the well-known 1PN gravitoelectric and gravitomagnetic
precessions.
Thus, even if g can be considered constant, in general it is not so for I, , ω. Again, the
peculiarities of the specific binaries at hand may greatly reduce the variability of them. For
example, the more the system’s orbit is large, the weaker are the effects of general relativity.
Our results of equations (16)–(23) are quite general, and can be applied to any
gravitationally bound two-body system immersed in a (weak) external field: they are valid for
any orbital geometry of the binary system and for a quite generic spatial orientation of g in a
given coordinate system.

3. The line-of-sight perturbation ρ

In timing of binary systems typically hosting radiopulsars, the basic observable quantity is the
projection ρ of the pulsar’s barycentric orbit along the line of sight [9] since its variations ρ
are straightforwardly related to the timing via τ = ρ/c. In view of the fact that ρ̂ = Ẑ,
one has to consider the Z component of the pulsar’s barycentric position vector r
Z = r sin I sin (ω + ν) , (30)
where ν is the true anomaly. The SEP-violating perturbation of ρ due to equation (1) can be
straightforwardly obtained in the following way:
 Pb  2π  
dZ ∂Z dE dM  ∂Z dκ dt
ρ = dt = + dE, κ = a, e, I, ω,
0 dt 0 ∂E dM dt κ
∂κ dt dE
(31)
where
r = a (1 − e cos E ) , (32)
7
Class. Quantum Grav. 30 (2013) 025006 L Iorio

cos E − e
cos ν = , (33)
1 − e cos E

1 − e2 sin E
sin ν = , (34)
1 − e cos E

dE 1
= , (35)
dM 1 − e cos E

dt 1 − e cos E
= , (36)
dE nb
and dκ/dt are the instantaneous rates of change of the Keplerian orbital elements computed
onto the unperturbed Keplerian orbital ellipse according to, e.g., the right-hand sides of the
Gauss perturbative equations8 [24]. We have

π g 1 − e2 sin I
ρ = − {(4 − 6e2 + 3e4 )ĝz sin I sin 2ω
2e4 n2b
+(4 − 6e2 + 3e4 ) cos I sin 2ω(ĝy cos  − ĝx sin )
+[−e4 + (4 − 6e2 + 3e4 ) cos 2ω](ĝx cos  + ĝy sin )}. (37)
No approximations in ĝ, e and I were used in deriving equation (37). A purely formal singularity
appears in equation (37) for e → 0: actually, it is unphysical since it is cured by using properly
chosen non-singular orbital elements (see, e.g., [33]).

4. The radial velocity perturbation ρ̇

A standard observable in spectroscopic studies of binaries is the radial velocity. Up to the


velocity of the binary’s center of mass V0 , it is, by definition, dρ/dt, i.e. the temporal rate
of change of the line-of-sight projection r of the barycentric orbit of the binary’s component
whose light curve is available.
The unperturbed, Keplerian expression for ρ̇ can straightforwardly be obtained from
equations (30) and (40) as
∂Z dν nb a sin I
ρ̇ = nb = √ [e cos ω + cos (ω + ν)] . (38)
∂ν dM 1 − e2
The SEP-violating perturbation ρ̇SEP  on ρ̇ can be worked out by replacing Z with equations
(38) in (31). For computational purposes, it turns out more convenient to use the true anomaly
ν as fast variable of integration instead of the eccentric anomaly E. By using
a(1 − e2 )
r= , (39)
1 + e cos ν
dν  a 2 
= 1 − e2 , (40)
dM r
dt (1 − e2 )3/2
= , (41)
dν nb (1 + e cos ν)2
 
8 In principle, dM/dt includes also nb , which yields the purely Keplerian part ρKep of the shift of the line-of-sight
component of the orbit. It vanishes.

8
Class. Quantum Grav. 30 (2013) 025006 L Iorio

one finally obtains



9π g 1 − e2 sin I
ρ̇ = − {ĝz sin I cos 2ω
e2 nb
+ cos I cos 2ω(ĝy cos  − ĝx sin ) − sin 2ω(ĝx cos  + ĝy sin )}. (42)
Also in this case, the use of suitably defined non-singular orbital elements cures the formal
singularity in equation (42) for e → 0.

5. The range and range-rate perturbations r and ṙ

By proceeding as in the previous sections, it is possible to work out the SEP-violating range
and range-rate perturbations r and ṙ, respectively. More precisely, we used equation
√ (31)
in which we replaced Z with equation (39) for r, and with dr/dt = nb ae sin ν/ 1 − e2
for ṙ. Generally speaking, such observable quantities are routinely measured in, e.g., the
Earth–Moon LLR experiment9 and in several spacecraft-based interplanetary missions. The
following calculations refer to the vector r connecting the centers of mass of the two bodies
constituting the binary system; thus, they are valid when the distances of the ranging devices
(laser retroreflectors, transponders on probes, etc) from the centers of mass of the bodies are
negligible with respect to r.
We obtain

3π  1 − e2
r = [ĝz cos ω sin I + cos I cos ω(ĝy cos  − ĝx sin )
n2b
− sin ω(ĝx cos  + ĝy sin )], (43)

3π [1 + e(2 − e)]


ṙ = {cos ω(ĝx cos  + ĝy sin ) +
(1 − e2 )nb
+ sin ω[ĝz sin I + cos I(ĝy cos  − ĝx sin )]}. (44)
No approximations in e were used in deriving equations (43) and (44).

6. Summary and conclusions

We worked out the first-time derivatives of some orbital effects induced by a Stark-like SEP
violation in a binary system made of two self-gravitating bodies immersed in an external
polarizing field g, assumed constant and uniform with respect to the characteristic temporal
and spatial scales of the binary.
We provided the reader with full analytical expressions for the SEP-violating rates of
change of all the six osculating Keplerian orbital elements (section 2), for the projection of
the binary’s orbit along the line of sight (section 3) and its time derivative (section 4) and
for the range and range-rate (section 5). We did not make any a priori assumption about the
orientation of g in space. We did not make simplifying assumptions about the orbital geometry
of the binary.
It is important to have at disposal explicit expressions of the different relevant SEP-
violating effects since each one has a peculiar temporal pattern which may be helpful in
separating it from other possible competing signatures. For example, PSR B1620-26 [34, 35]
is a pulsar-white dwarf binary in a relatively wide orbit (Pb = 16540653(6) s ∼ 2 × 102 d
9 If the polarizing external field is due to the Sun, our calculation are not strictly applicable to the Earth–Moon system

since the temporal variations of g may not be neglected over the characteristic timescales of such a binary.

9
Class. Quantum Grav. 30 (2013) 025006 L Iorio

[36]) orbited by quite distant circumbinary planet-like companion10 (Pb ∼ 3 × 104 d). For a
discussion of the rate of change of the eccentricity in binaries hosting compact objects, see,
e.g., [2].
Our results are important also to accurately assess the overall uncertainty which can
be obtained when constraints on  are inferred by comparing them with the corresponding
empirically determined quantities. Indeed, taking into account only the accuracy with which
an observable like, say, de/dt can be determined is, in principle, not enough; a propagation
of the errors affecting all the parameters such as g [40–42], and its orientation ĝ, entering
the corresponding theoretical prediction must be done as well to correctly infer the total,
systematic uncertainty in .
Our expressions are also useful in designing suitable SEP tests and in better interpreting
present and future experiments, not necessarily limited to binary pulsars, especially when
different competing theoretical mechanisms yielding SEP violations are considered. Indeed,
we found that the ratio of the node and inclination SEP-violating precessions depend only on
the inclination itself and on the pericenter.

Acknowledgments

I thank P Freire for useful correspondence and references. I am also grateful to an anonymous
referee for her/his helpful comments.

References

[1] Damour T 2012 Theoretical aspects of the equivalence principle Class. Quantum Grav. 29 184001
[2] Freire P C C, Kramer M and Wex N 2012 Tests of the universality of free fall for strongly self-gravitating bodies
with radio pulsars Class. Quantum Grav. 29 184007
[3] Nordtvedt K 1968 Equivalence principle for massive bodies: I. Phenomenology Phys. Rev. 169 1014–6
[4] Nordtvedt K 1968 Equivalence principle for massive bodies: II. Theory Phys. Rev. 169 1017–25
[5] Nordtvedt K 1968 Testing relativity with laser ranging to the Moon Phys. Rev. 170 1186–7
[6] Turyshev S and Williams J 2007 Space-based tests of gravity with laser ranging Int. J. Mod. Phys. D 16 2165–79
[7] Hofmann F, Müller J and Biskupek L 2010 Lunar laser ranging test of the Nordtvedt parameter and a possible
variation in the gravitational constant Astron. Astrophys. 522 L5
[8] Turyshev S, Farr W, Folkner W, Girerd A, Hemmati H, Murphy T, Williams J and Degnan J 2010 Advancing
tests of relativistic gravity via laser ranging to Phobos Exp. Astron. 28 209–49
[9] Damour T and Schäfer G 1991 New tests of the strong equivalence principle using binary-pulsar data Phys. Rev.
Lett. 66 2549–52
[10] Chandrasekhar S 1939 An Introduction to the Study of Stellar Structure (Chicago, IL: University of Chicago
Press)
[11] Lang K 1980 Astrophysical Formulae (Berlin: Springer)
[12] Lattimer J and Prakash M 2001 Neutron star structure and the equation of state Astrophys. J. 550 426–42
[13] Williams J G, Turyshev S G and Boggs D H 2012 Lunar laser ranging tests of the equivalence principle Class.
Quantum Grav. 29 184004
[14] Milgrom M 2009 MOND effects in the inner Solar system Mon. Not. R. Astron. Soc. 399 474–86
[15] Blanchet L and Novak J 2011 External field effect of modified Newtonian dynamics in the Solar system Mon.
Not. R. Astron. Soc. 412 2530–42
[16] Hui L and Nicolis A 2012 An observational test of the Vainshtein mechanism Phys. Rev. Lett. 109 051304
[17] Iorio L 2012 Constraints on Galileon-induced precessions from solar system orbital motions J. Cosmol.
Astropart. Phys. JCAP12(2012)001
[18] Damour T and Donoghue J F 2011 Spatial variation of fundamental couplings and Lunar Laser Ranging Class.
Quantum Grav. 28 162001
[19] Iorio L 2011 Orbital effects of spatial variations of fundamental coupling constants Mon. Not. R. Astron.
Soc. 417 2392–400
10 For exoplanets, see [37–39].

10
Class. Quantum Grav. 30 (2013) 025006 L Iorio

[20] Iorio L 2011 Observational constraints on spatial anisotropy of G from orbital motions Class. Quantum
Grav. 28 225027
[21] Bailey Q 2010 Lorentz-violating gravitoelectromagnetism Phys. Rev. D 82 065012
[22] Iorio L 2012 Orbital effects of Lorentz-violating Standard Model Extension gravitomagnetism around a static
body: a sensitivity analysis Class. Quantum Grav. 29 175007
[23] Damour T and Esposito-Farèse G 1992 Testing local Lorentz invariance of gravity with binary-pulsar data Phys.
Rev. D 46 4128–32
[24] Bertotti B, Farinella P and Vokrouhlický D 2003 Physics of the Solar System (Dordrecht: Kluwer)
[25] Calura M, Fortini P and Montanari E 1997 Post-Newtonian Lagrangian planetary equations Phys. Rev.
D 56 4782–8
[26] Calura M, Montanari E and Fortini P 1998 Lagrangian planetary equations in Schwarzschild spacetime Class.
Quantum Grav. 15 3121–9
[27] Kopeikin S 1995 On possible implications of orbital parallaxes of wide orbit binary pulsars and their
measurability Astrophys. J. 439 L5–8
[28] Kopeikin S 1996 Proper motion of binary pulsars as a source of secular variations of orbital parameters Astrophys.
J. 467 L93–5
[29] van Straten W, Bailes M, Britton M, Kulkarni S R, Anderson S B, Manchester R N and Sarkissian J 2001 A test
of general relativity from the three-dimensional orbital geometry of a binary pulsar Nature 412 158–60
[30] Reid M J and Brunthaler A 2004 The proper motion of Sagittarius A∗ . II. The mass of Sagittarius A∗ Astrophys.
J. 616 872–84
[31] Queloz D, Eggenberger A, Mayor M, Perrier C, Beuzit J, Naef D, Sivan J and Udry S 2000 Detection of a
spectroscopic transit by the planet orbiting the star HD209458 Astron. Astrophys. 359 L13–7
[32] Fluri D and Berdyugina S 2010 Orbital parameters of extrasolar planets derived from polarimetry Astron.
Astrophys. 512 A59
[33] Broucke R and Cefola P 1972 On the equinoctial orbit elements Celest. Mech. Dyn. Astron. 5 303–10
[34] Sigurdsson S, Richer H, Hansen B, Stairs I and Thorsett S 2003 A young white dwarf companion to pulsar
B1620-26: evidence for early planet formation Science 301 193–6
[35] Richer H B, Ibata R, Fahlman G G and Huber M 2003 The pulsar/white dwarf/planet system in messier 4:
improved astrometry Astrophys. J. 597 L45–7
[36] Arzoumanian Z, Joshi K, Rasio F A and Thorsett S E 1996 Orbital parameters of the PSR B1620-26 triple
system IAU Colloq 160: Pulsars: Problems and Progress (Astronomical Society of the Pacific Conference
Series vol 105) ed S Johnston, M A Walker and M Bailes (San Francisco, CA: Astronomical Society of the
Pacific) pp 525–530
[37] Wright J T et al 2011 The exoplanet orbit database Publ. Astron. Soc. Pac. 123 412–22
[38] Schneider J, Dedieu C, Le Sidaner P, Savalle R and Zolotukhin I 2011 Defining and cataloging exoplanets: the
exoplanet.eu database Astron. Astrophys. 532 A79
[39] Perryman M 2011 The Exoplanet Handbook (Cambridge: Cambridge University Press)
[40] Kuijken K and Gilmore G 1989 The mass distribution in the galactic disc-part III-the local volume mass density
Mon. Not. R. Astron. Soc. 239 651–64
[41] Paczyński B 1990 A test of the galactic origin of gamma-ray bursts Astrophys. J. 348 485–94
[42] Xue X X et al 2008 The milky way’s circular velocity curve to 60 kpc and an estimate of the dark matter halo
mass from the kinematics of 2400 SDSS blue horizontal-branch stars Astrophys. J. 684 1143–58

11

You might also like