Anda di halaman 1dari 20

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/300123854

Physics and astrophysics of neutron stars

Conference Paper · December 2015


DOI: 10.1063/1.4937184

CITATIONS READS

12 589

9 authors, including:

Riccardo Belvedere Federico Cipolletta


Centro Brasileiro de Pesquisas Físicas TIFPA and University of Trento
15 PUBLICATIONS   112 CITATIONS    9 PUBLICATIONS   73 CITATIONS   

SEE PROFILE SEE PROFILE

Sheyse Martins de Carvalho Simonetta Filippi


Universidade Federal Fluminense Università Campus Bio-Medico di Roma
10 PUBLICATIONS   40 CITATIONS    137 PUBLICATIONS   932 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Cells biophysical modeling View project

Spritz - a new fully GRMHD code View project

All content following this page was uploaded by Riccardo Belvedere on 05 May 2016.

The user has requested enhancement of the downloaded file.


Physics and Astrophysics of Neutron Stars
R. Belvedere1 , F. Cipolletta2 , C. Cherubini3,4 , S. M. de Carvalho1,5 , S. Filippi3,4 ,
R. Negreiros5 , Jonas P. Pereira6 , Jorge A. Rueda1,2,7,a) and R. Ruffini1,2,7
1
ICRANet-Rio, Centro Brasileiro de Pesquisas Fı́sicas, Rua Dr. Xavier Sigaud 150, Rio de Janeiro, RJ, 22290-180,
Brazil
2
Dipartimento di Fisica and ICRA, Sapienza Università di Roma, P.le Aldo Moro 5, I–00185 Rome, Italy
3
Nonlinear Physics and Mathematical Modeling Lab, University Campus Bio-Medico of Rome, Via A. del Portillo
21, I–00128 Rome, Italy
4
International Center for Relativistic Astrophysics-ICRA, University Campus Bio-Medico of Rome, Via A. del
Portillo 21, I–00128 Rome, Italy
5
Instituto de Fı́sica, Universidade Federal Fluminense, UFF, Niterói, 24210-346, RJ, Brazil
6
Department of Physics, Astronomy and Geosciences, Towson University, 8000 York Road, Towson, Maryland
21252-0001, USA
7
ICRANet, P.zza della Repubblica 10, I–65122 Pescara, Italy
a)
Corresponding author: jorge.rueda@icra.it

Abstract. Recent progress in the physics and astrophysics of neutron stars is presented.

1. INTRODUCTION
This paper summarizes some of the contributions related to the physics and astrophysics of neutron stars (NSs). In
particular, we here compile the results showed in the talks of R. Belvedere, Sheyse M. de Carvalho, Jonas P. Pereira
and Jorge A. Rueda. The order of the sections has been chosen to preserve the logic and connections within the topics.
We start in section 2 with the results shown by Jorge A. Rueda in his talk on the construction of equilibrium
configurations of uniformly rotating NSs. The solutions are obtained by integrating the Einstein equations in the
axially symmetric case for realistic up-to-date nuclear equations of state (EOS). We outline the main properties of
the configurations: mass, polar and equatorial radii, eccentricity, angular momentum, moment of inertia, and mass
quadrupole moment, for different values of the central density and angular velocity. The maximum rotation frequency,
maximum mass, and limits of stability (mass-shedding and secular instability) are also shown. We find fitting formulas
for relevant quantities in astrophysical applications such as the maximum dimensionless angular momentum and
binding energy of NSs as well as a fitting formula of the secular axisymmetric instability line.
We continue in section 3 on the structure properties of compact stars with the interesting results shown by Jonas
P. Pereira in his talk on how to take into account surface degrees of freedom for the radial stability analyses of stratified
compact stars such as NSs and strange quark stars. He showed that the stability analysis of the radial displacements in
stratified stars with surface degrees of freedom is equivalent to searching for extra boundary conditions for them which
can be obtained in general in the spherically symmetric case with the aid of a generalized distributional approach.
We turn to the astrophysics of NSs in sections 4 and 5. We first show in section 4 the results shown by Sheyse
M. de Carvalho in her talk on the use of the thermal evolution of NSs as a potential test of the structure of the star,
which can reveal the new structure of NS obtained under the assumption of global charge neutrality in contrast to the
assumption of local charge neutrality. Globally neutral NSs were obtained from the solution of the called Einstein-
Maxwell-Thomas-Fermi equations that account for the strong, weak, electromagnetic and gravitational interactions.
The solution of these equations lead us to a new structure of a NS significantly different from the traditional config-
urations obtained through the Tolman-Oppenheimer-Volkoff equations, which impose the condition of local charge
neutrality. We have now a less massive and thiner crust, leading to a new mass-radius relation.
Finally, we discuss the results of the talk by R. Belvedere, who first summarized the properties of the above
NS model based on the solution of the coupled Einstein-Maxwell-Thomas-Fermi equations, taking into account all
fundamental interactions within the framework of general relativity and relativistic mean field theory, both in the
static and in the uniformly rotating cases. The rotation is introduced through the Hartle’s formalism. This model
fulfills global and not local charge neutrality. He then showed that, using the realistic parameters of rotating NSs,
values of the magnetic field and radiation efficiency of pulsars, strongly different from estimates based on fiducial
parameters (mass M = 1.4 M⊙ , radius R = 10 km, moment of inertia I = 1045 g cm2 ), are obtained. The magnetic
field inferred from the traditional Newtonian rotating magnetic dipole model and the one obtained from its general
relativistic analog are compared and contrasted. Applying these considerations to the specific high-magnetic field
pulsars class, we show that all of these sources can be described as canonical pulsars driven by the rotational energy
of the NS, with magnetic fields lower than the quantum critical field for any value of the NS mass.

2. REALISTIC CONFIGURATIONS OF ROTATING NEUTRON STARS IN FULL


GENERAL RELATIVITY

AXIALLY SYMMETRIC EINSTEIN EQUATIONS AND NS EQUATION OF STATE


The interior and exterior metric of a uniformly rotating NS can be written in form of the stationary axisymmetric
spacetime metric
ds2 = −e2ν dt2 + e2ψ (dφ − ωdt)2 + e2λ (dr2 + r2 dθ2 ), (1)
where ν, ψ, ω and λ depend only on variables r and θ. It is useful to introduce the variable eψ = r sin(θ)Be−ν , being
again B = B(r, θ). The energy-momentum tensor of the NS interior is the one of a perfect fluid, given by

T αβ = (ε + P)uα uβ + Pgαβ , (2)

where ε and P denote the energy density and pressure of the fluid, and uα is the fluid 4-velocity. Thus, with the metric
given by equation (1) and the energy-momentum tensor given by equation (2), one can write the field equations as
(setting ζ = λ + ν):
2
" #
1 2 2 3 −4ν 2ζ−2ν (ε + P)(1 + v )
∇ · (B∇ν) = r sin θB e ∇ω · ∇ω + 4πBe + 2P , (3)
2 1 − v2
  (ε + P)v
∇ · r2 sin2 θB3 e−4ν ∇ω = −16πr sin θB2 e2ζ−4ν , (4)
1 − v2
∇ · (r sin(θ)∇B) = 16πr sin θBe2ζ−2ν P, (5)

 #2 −1 
B,r 2
! "
2 B,r  

 2
     1 −1  2
 h  i 
 B r B,rr − 1 − µ2 B,µ − 2µB,µ

ζ,µ = −  1 − µ 1 + r + µ − 1 − µ
B B   2 ,µ

 
(  B,µ ) "  B,µ # "  B,µ #
 B,r 1 B,r 1  3 B,µ 
× −µ + 1 − µ2 +r µ + µr + 1 − µ2 + −µ2 + µ 1 − µ2
B B 2 B 2 B 2 B B
 B,µr !
 B,r       2  
− 1 − µ2 r − µr2 ν,r 2 − 2 1 − µ2 rν,µ ν,r + µ 1 − µ2 ν,µ − 2 1 − µ2 r2 B−1 B,r ν,µ ν,r

1+r
B B
       2    1 1 
+ 1 − µ2 B−1 B,µ r2 ν,r 2 − 1 − µ2 ν,µ + 1 − µ2 B2 e−4ν µr4 ω,r 2 + 1 − µ2 r3 ω,µ ω,r

4 2

1  
2 2
 2 1  
2 4 −1 1  
2 2 −1

2 2  2   2 
− µ 1 − µ r ω,µ + 1 − µ r B B,r ω,µ ω,r − 1 − µ r B B,µ r ω,r − µ ω,µ , (6)
4 2 4

where, in the equation for ζ,µ , we introduced µ ≡ cos(θ).


To integrate the previous set of equations, coupled to the hydrostationary one, we used the RNS public code (see
the web page http://www.gravity.phys.uwm.edu/rns/) by Stergiuolas and Friedman [1]. For a complete review of the
method used we refer the reader to [2, 3, 4, 5].
We can integrate numerically the above Einstein equations once a relation between ε and P is given, namely an
equation of state (EOS). The NS interior is made of a core and a crust. The core of the star has densities higher than
the nuclear value, ρnuc ≈ 3 × 1014 g cm−3 , and it is composed by a degenerate gas of baryons (e.g. neutrons, protons,
hyperons) and leptons (e.g. electrons and muons). The crust, in its outer region (ρ ≤ ρdrip ≈ 4.3 × 1011 g cm−3 ), is
composed of ions and electrons, and in the so-called inner crust (ρdrip < ρ < ρnuc ), there are also free neutrons that drip
out from the nuclei. For the crust, we adopt the Baym-Pethick-Sutherland (BPS) EOS [6]. For the core, we here adopt
modern models based on relativistic mean-field (RMF) theory, which have Lorentz covariance, intrinsic inclusion of
spin, a simple mechanism of saturation for nuclear matter, and they do not violate causality. In particular the set of
parameters for the EOS are the so called NL3 [7], TM1 [8] and GM1 [9, 10]. We use an extension of the formulation of
[11] with a massive scalar meson (sigma) and two vector meson (omega and rho) mediators, and possible interactions
between them.
With the knowledge of the EOS we can compute equilibrium configurations integrating the above equilibrium
equations for given initial conditions, one of which being always the central density and the other a choice between
angular momentum, angular velocity, gravitational mass, baryonic mass or polar to equatorial radius ratio. However
not for all conditions given as input, a model of self-gravitating NS in hydrostationary equilibrium does exist, but they
should be contained in a suitable set defined by conditions of stability which we discuss in the next section.

THE NS STRUCTURE
Stability Region
The uniformly rotating NS equilibrium configurations are located within the region in the mass-central density plane
bounded by the following three limits:

1. the static limit which is composed by non-rotating (spherical symmetric) models;


2. the mass-shedding (Keplerian) limit, namely the models rotating at the maximum admitted angular velocity,
obtained as the limit at which the star’s angular velocity equals the one of a test-particle at the star’s equator;
3. the secular axisymmetric stability limit, i.e. the separatrix of stable and unstable models with respect to axisym-
metric perturbations, obtained as the locus of maxima of constant angular momentum sequences (see, [12], for
details).

For the sake of example, we show hereafter our results for the GM1 EOS (with some exceptions where we
give the results for other nuclear EOS). Analogous results are obtained for other nuclear EOS, for instance we refer
the reader to [13] for the NL3 and the TM1 nuclear parameterizations, in addition to the GM1 EOS shown in this
work. Figure 1 shows the aforementioned region of stability: the solid red, green and black curves represent the static,
Keplerian and secular instability sequences, respectively.

Equatorial Radius, Eccentricity, Moment of Inertia and Quadrupole Moment


The first question one may ask about the equilibrium configuration of self-gravitating NS might concern the shape of
the star since we know that the spherical symmetry breaks down for non-zero angular velocity. In the present case of
uniform rotation the axial symmetry is conserved and the figure of equilibrium is an oblate spheroid. The deformation
of the spheroid can be evaluated via the eccentricity
s
Rpol 2
!
ǫ = 1− , (7)
Req

where Rpol and Req are respectively the polar and equatorial radii.
Figure 2 gives us a general idea about the NS deformation plotting the relations between gravitational mass and
equatorial radius (left panel) and eccentricity and gravitational mass (right panel), for the same ω-constant sequences
in the right panel of figure 1. It is clear from these figures that, for a given rotation frequency, the larger the mass of
the NS the higher its compactness and the lesser its deformation.
An important quantity which is worth to examine is the moment of inertia, which is defined as
J
I= , (8)

.
28 Static sequence .
30
Keplerian sequence

2.6
Secular Instability
.
25

2.4 .
20
M [M ⊙]

M [M ⊙]
2.2
.
15
2.0
Static sequence
Keplerian sequence
.
10
Secular Instability
f
f
=50 Hz
1.8 f
=200 Hz
=300 Hz
. f=500 Hz
05 f
1.6
=716 Hz

14.7 .
14 8 .
14 9 15 0 . .
15 1 .
15 2 .
15 3 .
14 4 .
14 6 14 8. .
15 0 .
15 2
log( εc /c
2
[g cm−3 ]) log( εc /c2
[g cm−3 ])

FIGURE 1. The NS gravitational mass as a function of the central energy density for constant angular momentum sequences (left
panel, colored dashed curves) and constant angular velocity sequences (right panel) obtained with the EOS GM1. In this plot and
hereafter, the red, green and black curves represent respectively the static, the Keplerian, and the secular instability sequences.

where J is the star’s angular momentum.


Figure 3 shows the moment of inertia as a function of the gravitational mass (left panel) and as a function of
the compactness (right panel), the latter defined as GM∗ /(c2 R∗ ), where M∗ and R∗ are the mass and radius of the
spherical configuration with the same central density as the rotating model considered. This figure confirms the above
result namely that deviations from the spherical symmetry increase, along a constant rotation frequency sequence,
for decreasing values of the NS mass (lower compactness). Similarly, for a given mass, the deformation of the figure
of equilibrium becomes more pronounced for increasing values of the rotation frequency, where full rotation effects
beyond the slow rotation approximation are needed for an accurate description.
Following this reasoning, it is important to see how a constant frequency sequence imposes constraints on the
structure of a pulsar. Particularly interesting becomes the case of the f = 716 Hz sequence (blue curve), which
corresponds to the fastest observed pulsar, PSR J1748 − 2446ad. The constant frequency sequence intersects the
stability region in two points: at the maximally rotating Keplerian sequence, defining a minimum mass for the pulsar,
and at the secular axisymmetric instability limit, in the upper part, defining the maximum possible mass for the given
frequency (see the right panel of figure 1). Clearly these minimum and maximum mass values depend upon the EOS.
For the GM1 EOS employed here, we founded that the mass of PSR J1748 − 2446ad has to be & 1.20 M⊙ .
Another important quantity related to the deformation of the star is the mass quadrupole moment. The Geroch-
Hansen mass quadrupole moment [14, 15, 16] extracted via the Ryan’s expansion method [17] is given by
!
corr 4 1
M2 = M2 − + b0 M 3 , (9)
3 4
where a superscript corr stand for corrected, in the sense that this formula corrects the value M2 as computed from
the RNS code (see, [18], for details)
Z 1 ′2 ′ Z 1
1 s ds
M2 = req 3 ′ 4
P2 (µ′ )S̃ ρ (s′ , µ′ )dµ′ , (10)
2 0 (1 − s ) 0

where req is the coordinate equatorial radius, ρ ≡ 2ν − ln(B), s = r/(r + req ) ∈ [0, 1] is a compacted radial coordinate,
µ = cos(θ), P2 (µ) is the Legendre polynomial of second order, and S̃ ρ = r2 S ρ , being S ρ a source function defined as

1 + u2
" #
γ
2λ 2 −2ρ 2 1 2 2 1 1
S ρ (r, µ) = ω,r + 2 (1 − µ )ω,µ + γ,r − 2 µγ,µ

e 2 8πe (ε + P)


2
+r e
1−u r r r
( ! " #)
ρ 1 1 1 1 
+ 16πe2λ − γ,r γ,r + 2
γ,µ γ,µ (1 − µ2 ) − µ , (11)
2 2 r r 2
30 . Static sequence .
1 0
f
Keplerian sequence f
f
= 300 Hz

f f f
Keplerian sequence = 50 Hz = 500 Hz

. f
=50 Hz = 200 Hz = 716 Hz

f .
=200 Hz
25
f
=300 Hz 0 8

f
=500 Hz

.
=716 Hz
20
.
]
M M⊙

0 6

ǫ
[

15
.
0 4

10 .
.
0 2

05 .
12 14 16 18 20 22 24 26
.
0 0
. . . . . .
R M M⊙
0 5 1 0 1 5 2 0 2 5 3 0

eq [km] [ ]

FIGURE 2. NS gravitational mass as a function of the equatorial radius (left panel) and eccentricity as a function of gravitational
mass (right panel) for the same ω-constant sequences of the right panel of figure 1.

6 Keplerian sequence 6 Keplerian sequence


f=50 Hz f=50 Hz
f=200 Hz f=200 Hz
5 f=300 Hz 5 f=300 Hz
f=500 Hz f=500 Hz
f=716 Hz f=716 Hz
g cm2 ]

g cm2 ]

4 4

3 3
45

45
I [10

I [10

2 2

1 1

0
.
05 .
10 .
15 2.0 .
25 .
30
0
0 05 . .
0 10 0.15 0.20 .
0 25 .
0 30
M M ⊙]
[ GM /(c R )

2

FIGURE 3. On the left panel moment of inertia is plotted against gravitational mass, while on the right panel it is plotted as a
function of compactness. The colours are the same as previous figures.

with γ = ln(B), while b0 is the correction term computed in [18, 19] as


√ Z 1 ′3 ′
16 2πreq 4
Z 1
s ds
2
q 1
b0 = − × dµ′
1 − µ′ 2 P(s′ , µ′ )eγ+2λ T 2 (µ′ ), (12)
0
M2 0 (1 − s )
′ 5
0
1 √
where T 02 is the Gegenbauer polynomial of order 0 with normalization T 01/2 = 2/πC0 , with C0 the traditional 0th-
order Gegenbauer polynomial.
It is interesting to compare and contrast the value of the mass quadrupole moment of a NS with the one predicted
by the exterior Kerr solution. For a given mass M and angular momentum J, the mass quadrupole moment of the Kerr
solution is:

J2
. M2Kerr = (13)
M
Figure 4 shows the relation between the ratio between the NS quadrupole moment obtained numerically from
equation (9) and the mass quadrupole moment of the Kerr solution, for selected sequences of constant rotation fre-
quency. In this plot we show both that this ratio never reach unity and that the ratio M2corr /M2Kerr is a decreasing function
.
4 0

f
Keplerian sequence

. f
= 300 Hz

f
= 500 Hz
3 5 = 716 Hz

.
3 0

Kerr
.
2
M /M
2 5
corr

.
2

2 0

.
1 5

.
. . . . . .
1 0

M M⊙
1 8 2 0 2 2 2 4 2 6 2 8

[ ]

FIGURE 4. Ratio of the mass quadrupole moment computed numerically from equation (9) to the value of the Kerr exterior
solution (13) as a function of mass for selected sequences of constant frequency. We show here only the region of large masses
where M2corr starts to approach the Kerr value M2Kerr .

of the NS mass, hence reaching its lowest value at the maximum mass configuration. Indeed, as we can see from figure
4, the largest the maximum mass attained by a NS model, the closest the NS quadrupole moment approaches the Kerr
solution value given by equation (13).

Some Useful Fitting Formulas for Astrophysical Applications


We would like to finish this section by presenting some important relations between NS physical quantities which
are useful for astrophysical applications. First, it is worth to discuss the dimensionless angular momentum (“Kerr
parameter”), a/M ≡ cJ/(GM 2 ), which we show in figure 5 as a function of the total mass for the maximally rotating
configurations, namely the Keplerian sequence, for a variety of EOS. It can be seen how a maximum value is attained
by the NS, (a/M)max ≈ 0.7, and which is approximately the same for all the selected EOS. The existence of such
a particular maximum, EOS-independent, value of a/M possibly implies the existence of universal limiting values
of the NS compactness and the rotational to gravitational energy ratio. This is a conjecture which deserves further
exploration.
Another often useful physical quantity to be computed is the binding energy of the configurations, or the relation
between the baryonic mass and the gravitational mass. For non-rotating NSs, we found that for the analyzed EOS, the
following relation hold:
!2
Mb M 13 M
≈ + , (14)
M⊙ M⊙ 200 M⊙
where Mb is the baryonic mass. We have checked that for other EOS the same relation approximately holds pointing
to a universal property. The maximum relative errors of the above fitting formula for the GM1, TM1 and NL3 EOS
are respectively 1.4%, 1.3% and 0.99%. For rotating configurations one could introduce a dependance on the angular
momentum of the configuration, namely Mb = Mb (M, J), in fact we find that for the set of EOS we udes there is a
common relation given by
!2 !
Mb M 13 M 1 1.7
= + 1− j , (15)
M⊙ M⊙ 200 M⊙ 130

where j ≡ cJ/(GM⊙2 ). This formula is accurate within an error of 2% and duly generalizes equation (14) in the limit
where the angular momentum vanishes ( j → 0).
0. 8
0. 7
0. 6

a/M = cJ/(GM2 )
0. 5 NL3
0. 4 TM1
GM1
0. 3 AU
FPS
C
0. 2 L
O
0. 1 UU
WS
N
0. 0 0. 5 1. 0 1. 5 2.0 2.5 3.0 3. 5
M [M ⊙]
FIGURE 5. Dimensionless angular momentum (“Kerr parameter”), a/M ≡ cJ/(GM 2 ), as a function of the total NS mass along the
Keplerian sequence for the EOS selected in this work (colored curves). For comparison we show the results for additional EOS,
taken from a set supplied by the RNS code (EOS.INDEX file). We refer the reader to the RNS web page and references therein for
further details of these EOS.

TABLE 1. Critical mass and corresponding radius for selected parameterizations of nuclear EOS
J=0 J=0 J,0 J,0
EOS Mcrit (M⊙ ) Rcrit (km) Mmax (M⊙ ) Rmax (km) p k fK (kHz)
NL3 2.81 13.49 3.38 17.35 1.68 0.006 1.34
GM1 2.39 12.56 2.84 16.12 1.69 0.011 1.49
TM1 2.20 12.07 2.62 15.98 1.61 0.017 1.40

Finally, the critical mass of the NS given by the secular axisymmetric instability limit (configurations along the
black curve of figure 1) is fitted approximately by:

crit J=0 p
MNS = MNS (1 + k jNS ), (16)

where the parameters k and p depends of the nuclear EOS (see table 1). These formulas fit the numerical results with
a maximum error of 0.45%.

3. PHYSICAL INSIGHTS INTO THE RADIAL STABILITY OF STRATIFIED STARS

The main aspect of this section will be the elucidation of the physical ideas present in Ref. [20], apropos of the
stability analyses of compact layered stars upon radial perturbations when surface degrees of freedom are present and
play a nontrivial role there. The motivations for this analysis are threefold. Firstly, it is believed that compact stars,
such as NSs, be stratified [21]. This can be qualitatively understood due to the huge range of densities they encompass,
leading the matter content there to likely be in more than one form. Such phase transitions would generically lead to the
discontinuity of some physical quantities, related to the appearance of surface degrees of freedom. An example of that
would be a NS constituted of a hadronic and a color-superconducting quark matter phase, where conformal degrees of
freedom should raise at their interface [22, 23]. Secondly, it seems the relevance and nontriviality of surface degrees
of freedom is being overlooked in the literature. Thirdly, it could play a role in important astrophysical phenomena,
such as glitches and QPOs [21].
POSING THE PROBLEM
We shall cast insightful light on the following problem: how to describe the stability of a spherically symmetric
stratified star endowed with surface degrees of freedom when its phases are perturbed? By surface degrees of freedom
we mean surface energy densities and surface tensions, present in the phase-splitting surfaces. It is simple to perceive
that for describing mathematically a multilayered system, it suffices to investigate only two of its phases and the
surface splitting them. Therefore, we shall work only with a star made up of two phases, split by a surface at r = R.
We shall also assume that the physics in any phase of the system is the same. This signifies that our task is
to connect them, by means of additional boundary conditions to the problem. A convenient mathematical tool for
this purpose is the generalized distributional analysis [24, 25], which deals with compositions of Heaviside functions
θ(r − R) and Dirac delta functions δ(r − R). Physically, the aforesaid quantities will be connected with the phases of
the systems and surface quantities present between any two of them, respectively. In the scope of generalized distribu-
tional analysis, boundary conditions shall raise automatically as constraints to the physically involved distributional
equations. Therefore, the ultimate goal is to properly promote the physical equations valid for continuous systems to
generalized distributions.

CLASSICAL STRATIFIED STARS


Assume a spherically symmetric star constituted of two phases (the physical quantities of the inner phase will be
characterized by the presence of the tag “−”, while the outer ones by “+”), separated by a surface at r = R. Assume
that it hosts a surface mass density σ and a surface tension P. The first equation to be generalized in the classical
case is the one of hydrostatic equilibrium [26]. Since now surface quantities are present, they must also be taken into
account for the total force balance. It is simple to show that this is done by dint of [20]

dP 2P
+ ρg(r) − δ(r − R) = 0, (17)
dr R
where g(r) is the norm of the distributional gravitational field, solution to (the distributional) Poisson’s equation
∇ · ~g = −4πGρ, ~g = −g(r)r̂, that can always be written as
Z r
GM(r)
g(r) = , M(r)  4π ρ(r̄)r̄2 dr̄. (18)
r2 0

The distributional mass density, on the other hand, in the presence of surface densities is expressed as

ρ(r) = ρ− (r)θ(R − r) + ρ+ (r)θ(r − R) + σδ(r − R). (19)

Finally, given that we are dealing with a thin shell, the distributional pressure must be written as

P(r) = P− (r)θ(R − r) + P+ (r)θ(r − R). (20)

At equilibrium, the surface quantities cannot be any. Defining [A(R)]+−  A+ (R) − A− (R) as the discontinuity of A(r)
through the surface r = R, it is simple to see that Eqs. (17)-(19) imply that

R G [M(R)]+−
P= [P(R)]+− + 3
[M 2 (R)]+− , σ = . (21)
2 16πR 4πR2
The first term of the above equation is nothing more than the generalization to the Young-Laplace law for a bubble-like
system [27] for the case it has a non-negligible surface mass. Its second term is the induced surface mass due to a
discontinuity of the gravitational field.
When a distributional volume element is perturbed from its equilibrium position, the generalized Newton’s sec-
ond law describing its dynamics reads [20]
( )
dvr ∂P 2P σ + −
ρ + + ρg(r) − + (v̇r + v̇r ) δ(r − R) = 0, (22)
dt ∂r R 2

where vr is its radial velocity distribution.


Let us investigate the dynamics of perturbations of volume elements (from their equilibrium points) of the form
+ −
ξ̃(r, t) = ξ+ (r)eiω t θ(r − R) + ξ− (r)eiω t θ(R − r), (23)

with ω± given constants. In this way, radial perturbation analyses are summarized by the reality (stable solutions)
or not of ω± . For obtaining the perturbation equations, one should apply the Laplacian operator ∆ [∆A  A(t, r +
ξ̃) − A0 (t, r), A0 and A being a physical quantity in the equilibrium and perturbed cases, respectively] on each side
of Eq. (22). This is simply the case because thermodynamical aspects will also be necessary to fully characterize the
system. Nevertheless, caution should be taken here since were are dealing with distributions. One can show [20] that
for A(r) = A+ (r)θ(r − R) + A− (r) θ (R − r),
∂A
∆A = δA + ξ̃ − [A]+− ξ̃δ(r − R), (24)
∂r
∂ξ̃ ∂A 1 ∂ξ̃+ ∂ξ̃−
! " #
∂A ∂
∆ = (∆A) − + + [A(R)]+− δ(r − R). (25)
∂r ∂r ∂r ∂r 2 ∂r ∂r

Applying ∆ as defined by Eqs. (24) and (25) in Eq. (22), assuming adiabatic properties for the involved fluids and the
constancy of the mass for each phase on average1 , the equations governing the dynamics of Eq. (23) are

4 dP± ±
" #
d ± ±1 d 2 ±
Γ P 2 (r ξ ) − ξ + ω2 ρ± ξ± = 0 (26)
dr r dr r dr

and " #+
d
ΓP (r2 ξ) − 4Rξ(R)[P(R)]+− + 2ξ(R)[3P − 2η2 σ] = 0, (27)
dr −

where Γ  ρ/P(∂P/∂ρ) (adiabatic index [26]) and η2  ∂P/∂σ (squared speed of the sound on the surface of
discontinuity [28]). Notice that the above equations are only meaningful when ω+ = ω−  ω and [ξ(R)]+− = 0.
The second condition simply means that the surface of discontinuity separating the phases is always localizable, while
the first one states that the system can only have a global set of eigenfrequencies. These constraints are exactly what
one expects from the analogy of our description with a system of coupled springs. Equations (26) are the same as one
would obtain for continuous and adiabatic systems [26], while Eq. (27) is our desired boundary condition connecting
any two phases. One sees that such an equation is highly nontrivial, depending even upon the microphysics of fluid at
the surface of separation of the phases, by means of η2 .

STABILITY OF GENERAL RELATIVISTIC COMPACT STARS


This case is much more involved, mathematically speaking, than the previous one analyzed. Nevertheless, the idea here
is simply to promote the associated equations to generalized distributions. The equations describing the perturbations
in the phases of a stratified star are simply the ones obtained for their continuous counterparts. It misses therefore only
the splitting-phases extra boundary condition that would generalize Eq. (26). Long calculations show that it can be
cast as [20]
2η2 ∆σ [cosh β]+−
" # " #
2P β ′ + β0 +
−∆ = − [Pe α ] − ξ̃ + [∆Pe ]− , (28)
R 4πR2 R2
where geometric units were used and for such a derivation it was assumed that the distributional metric to the space-
time is
ds2 = e2α dt2 − e2β dr2 − r2 (dθ2 + sin2 θdϕ2 ). (29)
The precise expression for each Laplacian operator in Eq. (28), as well as the classical limit of the aforesaid equation
can be found in Ref. [20].
We would like to stress that the physically interesting case to talk about the stability of stratified stars when their
phases are perturbed is the one where their surfaces of discontinuity are stable. This is only per se a nontrivial issue
and further details can be found in Ref. [28].
1 This automatically takes place when the phase-splitting surface is stable. The complementary case simply means that the system cannot linger

on in time when radial perturbations take place.


The cases where radial electric fields are present can also be worked out following the same lines as the ones
above. One can show [20] that the same functional boundary condition as in Eq. (28) raises there. The physical reason
for this is the consideration that a shell that splits two phases is thin, which means that electric fields only correct the
components of the associated surface perfect fluid and cannot render it anisotropic.

SUMMARY
We have mainly focused here on conveying the physical insights and ideas featuring in Ref. [20]. We showed that
problem of analyzing the stability upon radial displacements in stratified stars with surface degrees of freedom is
equivalent to searching for extra boundary conditions for them. Finally, such conditions can be generically obtained
in the spherically symmetric case with the help of a generalized distributional approach.

4. THERMAL EVOLUTION OF GLOBALLY NEUTRAL NEUTRON STARS


Recently a new approach has been introduced [29, 30, 31, 32], which developed a new model of NS fulfilling global
and not local charge neutrality. The equilibrium equations of this new treatment supersede the traditional ones based
on the Tolman-Oppenheimer-Volkoff (TOV) system of equations, which obeys local charge neutrality. The new equi-
librium equations, what we called Einstein-Maxwell-Thomas-Fermi (EMTF) equations, introduces self-consistently
the presence of all interactions within the framework of general relativity. The weak interactions are introduced by the
β-stability, and the strong interactions are modeled via the σ-ω-ρ nuclear model, where σ, ω and ρ are the mediator
massive vector mesons. We use the NL3 parameterization of this nuclear model. The supranuclear core is composed
by a degenerate gas of neutrons, protons, and electrons in β-equilibrium. The outer crust region ρ ≤ ρdrip ≈ 4.3 × 1011
g cm−3 is composed ions and electrons and in the inner crust, at ρdrip < ρ < ρnuc , where ρnuc ≈ 2.7 × 1014 g cm−3 is
the nuclear saturation density, there is an additional component of free neutrons dripped out from nuclei.
We obtain a new structure for a NS from the solution of the EMTF equations of equilibrium that are very
different from the traditional configurations obtained throught TOV equations. We have, a positively charged core as
a consequence of the balance between gravitational and Coulomb forces that results in the appearance of a Coulomb
potential energy eV ∼ mπ c2 deep. The transition between core-crust, that starts atρ = ρnuc , is marked by the existence
of a thin, ∆r ∼few hundreds fm, electron layer fully screening the core charge. Where the electric field becomes
overcritical, E ∼ m2π c3 /(e~), and the particle densities decrease until the base of the crust, which is reached when
global charge neutrality is achieved. Consequently, the core is matched to the crust at a density ρcrust ≤ ρnuc .
Configurations with ρcrust > ρdrip possess both inner and outer crust while in the cases with ρcrust ≤ ρdrip the
NS have only outer crust. In the limit ρcrust → ρnuc , both ∆r and E of the transition layer vanish, and the solution
approaches the one given by local charge neutrality (see Figs. 3 and 5 in [31]).
The purpose of this work is to study the consequences of the different values of the density of the crust, ρcrust ,
in the computation of the thermal evolution of the globally neutral NS. Here we cover the configurations with and
without inner crust, all the way to the limit ρcrust ≈ ρnuc , which corresponds to TOV-like solutions satisfying local
charge neutrality.

THERMAL EVOLUTION EQUATIONS


Here we briefly describe the thermal evolution equations that govern the cooling of NSs. The energy balance and
energy transport equations for general relativistic, spherically symmetric star read

∂(Leν ) 4πr2 ∂(T eν/2 )


" #
=−√ ǫν eν + cv , (30)
∂r 1 − 2m/r ∂t

Leν p ∂(T eν/2 )


2
= 1 − 2m/r , (31)
4πr κ ∂r
where ǫν is the neutrino emissivity, cv is the heat capacity per unit volume, κ is the thermal conductivity, T (r, t) is the
interior temperature, L(r, t) is the radiation luminosity, m is the mass enclose within a radius r, and ν = ln g00 , with
g00 the 0–0 component of the metric.
We have two boundary conditions required by the equations 30 and 31, one for the center L(r = 0) = 0 and
another one for the surface T (r = R) = T s [33].
We included all the relevant process of neutrino emission: in the core we consider the direct and modified Urca
processes, neutron-neutron (nn), proton-proton (pp) and neutron-proton (np) Bremsstrahlung. In the crust we have
plasmon decay, e− e+ pair annihilation, electron-nucleus and electron-nuclei Bremsstrahlung. The heat capacity is due
to by electrons, protons and neutrons in the core and by electrons and ions in the crust. Heat capacity and thermal
conductivity follow their traditional formulation as described in [33] and references therein.

COOLING CURVES AND RELAXATION TIME


Now we investigate the thermal evolution of stars with different structures. These NSs has the same gravitational mass
(M ≈ 1.4 M⊙ ), whith different values for ρcrust , which in turn is reflected in the thickness of the crust. Their crust
are increasingly thinner for smaller values of the density at its base, ρcrust . This is illustrated in Fig. 6. We note that
the configuration whose ρcrust = 2.4 × 1014 g cm−3 has approximately the structure of a locally neutral NS, so this
configuration should exhibit traditional cooling properties known in the literature.
The cooling curves are computed by integrating numerically the energy balance and transport equations (30–31)
for the configurations shown in Fig. 6.

100 ρcrust (g cm−3 )


5.0 ×1012
1.0 ×1013
10−1 5.0 ×1013
1.0 ×1014
2.4 ×1014
10−2
ρ/ρnuc

CORE CRUST
10−3

10−4

10−5

10−6
11.5 12.0 12.5 13.0
r (km)

FIGURE 6. Density profiles of globally neutral NS with mass M ≈ 1.4 M⊙ for selected values of the density at the base of the
crust, ρcrust . Notice that at the transition, the density of the core is that of the nuclear saturation density, ρnuc .

In Fig. 7 we show the surface temperature as observed at infinity, as a function of time t in yr for the NS
configurations shown in Fig. 6.
We can see in Figure 7 that different configurations have qualitatively the same behavior but the time at which
each star becomes isothermal, thermal relaxation time tw , [34] is different. The thermal relaxation time is signaled by
the drop in the surface temperature which depends on whether or not the fast neutrino processes are taking place in
the core. For stars younger than t = tw , the crust of the star is hotter than the core and so heat flows from the crust to
the core, or equivalently as a cooling front propagating from the core to the crust. The time tw is therefore the time it
takes for cooling front to reach the star surface.
For local charge neutrality NSs, [35] it is found by numerical simulations that tw is longer for thicker crusts,
following approximately tw ∝ ∆R1.4−1.8
crust , where ∆Rcrust is the crust thickness. In the Fig. 8 we show a enlargement
around the temperature drop at the end of the thermal relaxation phase of Fig. 7. And we can see that for the NSs
we are studying, the aforementioned behavior is only valid for stars whose ρcrust ≥ 5 × 1013 g cm−3 . For stars with
ρcrust ≤ 5 × 1013 g cm−3 , however, we observe the opposite behavior, that is, an increase in tw for thinner crusts. The
thickness of the crust contributes to increase tw , since for a thicker crust the cooling front needs to “travel” a larger
distance and for crusts with larger ρcrust we have more intense emission of neutrinos from the crustal region.
ρcrust (g cm−3 )
5.0 ×1012
1.0 ×1013
5.0 ×1013
106 1.0 ×1014
2.4 ×1014

∞ (K)
Ts

105

10−1 100 101 102 103 104 105 106


t (yr)

FIGURE 7. Surface temperature at infinity T s∞ as a function of time t in yr for the NS configurations shown in Fig. 6.

In the case of traditional NSs the range of densities covered by the crust is roughly the same, thus if the crust
thickness is reduced one naturally obtains a smaller tw (since the crustal neutrino emission stays the same while the
crust is thinner). In our study however, we consider different crust thickness and ρcrust , thus while initially (when the
variation of ρcrust is small) the crustal neutrino emission is roughly the same (while the thickness is reduced), we have
the reduction of tw (as in traditional NSs). However, when ρcrust is significantly reduced, the neutrino emitting region
is also significantly reduced. Therefore, we have two competing processes: the increase of tw due to the reduction of
neutrino emission and the decrease of tw due to the geometrical reduction of the crust thickness. We have found that
for ρcrust . 5 × 1013 g cm−3 the former wins and we have an overall increase of tw for such stars.

2. 0

tw3 tw4 tw5 tw2 tw1


∞ (106 K)

1. 0

tw1 >tw2 >tw3, tw3 <tw4 <tw5

(g cm−3 )
Ts

ρcrust
0. 5 5.0 ×1012
1.0 ×1013
5.0 ×1013
1.0 ×1014
2.4 ×1014
10 50 100 200
t (yr)

FIGURE 8. Enlargement of the evolution of the surface temperature around its drop at the end of the thermal relaxation phase, for
the NSs shown in Fig. 6.
DISCUSSION
We have presented a investigation of the thermal properties of NSs satisfying global charge neutrality. In this config-
uration, the transition from the core to the crust is not contiguos but there is a gap filled by distribution of electrons
and an ultra-high electric field. This description allow us to obtain NSs with very different crusts: thin for the models
whose density at the crust base is low; or thick in the case of larger values of the crust base density; all the way to the
limit of local charge neutrality (where there is a continuous transition from core to crust). Since the observations of
the radii are still far from coming to reality and the fact that the crust plays a small role in the mass of the NS, we use
the thermal evolution of NS to obtain hints on how to constrain the crust thickness.
We have shown that the traditional proportionality between the relaxation time and the crust thickness is violated
for stars whose densities at the crust base is lower than ≈ 5 × 1013 g cm−3 , in which case the opposite behavior occurs,
namely one obtains a longer relaxation time for thinner crust. The reason for this is the reduction of the crustal neutrino
emission (due to the overall reduction of the crust density), which contributes to keeping the crust warm, compensating
the speed up even though the reduction of the crust size tend to speed up cooling. It is important to notice that for
densities higher than this value the traditional behavior is restored, in which the relaxation time increases with the
crust thickness. The fact that in the case in which the density at the base of the crust is equal to the neutron drip
density (ρdrip ≈ 4.3 × 1011 g cm−3 ), we have a situation similar to the quark stars, objects composed of stable quark
matter. Due to charge neutrality at the surface of the quark core of the stranfe star, an ultra-high electric field arises
filling the region between the core and a thin crust made up of ordinary matter, much like in the globally neutral NS.
The studies of the similarities between these two models and a investigation and comparison of the cooling behavior
of compact objects described by each model are currently under way.

5. HIGH-B PULSAR CLASS WITH NOT SO HIGH MANGETIC FIELDS


The surface magnetic field of pulsars is traditionally estimated [36, 37, 38] by equating the rotation energy loss of the
NS to the radiating power of a rotating magnetic point dipole in vacuum, given respectively by


Ėrot = −4π2 I , (32)
P3
and
2 µ2⊥ Ω4
Pdip = − , (33)
3 c3
where Ω is the rotation angular velocity of the star, µ⊥ = µ sin χ is the component of the magnetic dipole µ = BR3
perpendicular to the rotation axis, which is B, the magnetic field at the equator, and the inclination angle χ of the
magnetic dipole with respect to the rotation axis. Under these assumptions, the Newtonian magnetic field is
!1/2
3c3 I
B sin χ = PṖ , (34)
8π2 R6

where P and Ṗ are the rotational period and the spin-down rate of the pulsar, which are observational properties, while
the moment of inertia I and the radius R of the star are model-dependent properties.
In the literature are often extracted information on the nature of the pulsars applying the previous formulas with
fiducial values for the pulsar’s parameters, namely by using a mass M = 1.4 M⊙ , a radius R = 10 km, and a moment
of inertia I = 1045 g cm2 . In this case, for the rotation energy loss and the magnetic field respectively, are obtained the
relations
f Ṗ
Ėrot = −3.95 × 1046 3 erg s−1 , (35)
P
and  1/2
B f sin χ = 3.2 × 1019 PṖ G. (36)
Focusing on the particular class of the so-called High-B pulsars [39, 40], it is straightforward to see that, applying the
previous fiducial values for the mass, radius and moment of inertia, this class appears to have surface magnetic field
values higher than the critical field for quantum electrodynamical effects, Bc = (m2e c2 )/(e~) = 4.41 × 1013 G, as well
TABLE 2. Properties of the high-magnetic field pulsars obtained assuming fiducial NS parameters, R = 10 km and I =
1045 g cm2 , and using Equation (36) with inclination angle χ = π/2 and Equation (35). See [39, 40] for additional details of
these pulsars.
f
Pulsar B f /Bc LX (1033 erg s−1 ) LX /|Ėrot | P (s) Ṗ (10−12 )
J1846–0258 1.11 25 − 28∗ , 120 − 170† 0.0031 − 0.0035a , 0.015 − 0.021b 0.326 7.083

J1819–1458 1.13 1.8 − 2.4 6.21 − 8.28 4.263 0.575
J1734–3333 1.18 0.1 − 3.4 0.0018 − 0.0607 1.169 2.279
J1814–1744 1.24 < 43 < 91.5 1.169 2.279
J1718–3718 1.67 0.14 − 2.6 0.0875 − 1.625 3.378 1.598
J1847–0130 2.13 < 34 < 200 6.707 1.275

In 2000, prior to the 2006 outburst.

During the outburst in 2006.

Classified as a rotating radio transient (RRAT).

f
as, in some cases, to have luminosities higher than the rotational power of the NS, namely, LX > |Ėrot |, (see Table 1
for details). For this reason, has been proposed that this class of pulsars can be the missing link, i.e. transition objects,
between rotation-powered pulsars and the so-called magnetars: NSs powered by the decay of overcritical magnetic
fields. If this would be true, would lead to a large unseen population of magnetars in a quiescence state, which could
be disguised as radio pulsars, [40].
Anyway, we suggest that these conclusions might be premature since the surface magnetic fields inferred by fidu-
cial NS parameters, namely by Equation (36), are in general overestimated. Indeed, the knowledge of a more complex
EOS, structure, and stability conditions of both static and rotating NSs, which were acquired in the intervening years
from the seminal work of Oppenheimer[41], asks for using a more realistic NS configurations. As a result, much lower
values of the magnetic field are obtained when realistic structure parameters are applied and when general relativistic
corrections are introduced to the traditional Newtonian equation. In our analysis, we will compare and contrast both
the configurations subjected to the constraint of global charge neutrality, in which the Coulomb interactions are intro-
duced, and the configurations obtained under the traditional constraint of local charge neutrality. Our conclusion are
that, independently on the theoretical model, different structure parameters as functions of the central density and/or
rotation frequency of the star give rise to quite different quantitative estimates of the astrophysical observables with
respect to the use of fiducial parameters.
We use cgs units throughout the article unless otherwise specified.

NEUTRON STAR EQUILIBRIUM CONFIGURATIONS


Recently, we shown [29, 30, 31] the overriding of the Tolman-Oppenheimer-Volkoff (TOV) system of equations
[41, 42], by the Einstein-Maxwell system of equations coupled to the general relativistic Thomas-Fermi equations
of equilibrium (EMTF for short), in the case of both static and rotating NSs. These new equations account for the
weak, strong, gravitational, and electromagnetic interactions within the framework of general relativity and relativistic
nuclear mean field theory.
Conversely by the TOV-like approach, where the condition of local charge neutrality is applied to each point of
the configuration, in the EMTF equations is demanded the condition of global charge neutrality. Indeed, as shown in
[29, 30], local charge neutrality leads to an inconsistency with the equations of motion of the particles of the system.
Therefore, as shown first in [43] for the case of a self-gravitating one-component system of uncharged particles, the
general relativistic thermodynamic equilibrium of the star is not satisfied when local charge neutrality is applied to a
multi-component system with charged constituents. To ensure the thermodynamical equilibrium, the constancy, along
the whole configuration, of the generalized electro-chemical particle potentials (that we called “Klein potentials”) for
all of the species, has to be provided.
The weak interactions are introduced via the condition of β-equilibrium. The strong interactions are introduced
via the σ-ω-ρ virtual mesons nuclear model within the relativistic mean field theory á la Boguta [44]. To solve the
system of equation, its parameters (coupling constants and masses of the three mesons) have to be experimentally
determined. We here adopt the NL3 parameter set [45]: mσ =508.194 MeV, mω =782.501 MeV, mρ =763.000 MeV,
gσ =10.2170, gω =12.8680, gρ =4.4740, plus two constants that give the strength of the self-scalar interactions, g2 =
−10.4310 fm−1 and g3 = −28.8850.
The solution of the EMTF equations, under the assumption of the constancy of the Klein potentials, leads to
a structure of the NS quite different from the traditional configurations obtained through the TOV equations (see
Figure 4 in [31]): in the center, it is found a positively charged core, composed by a degenerate gas of neutrons,
protons, and electrons in β-equilibrium, whit density spanning from the supra-nuclear central value up to the nuclear
density ρnuc ≈ 2.7 × 1014 g cm−3 at its edge. The charge of the core is fully screened by an electron envelope, with
thickness of few hundreds Fermi. In such electron layer, a supra-critical electric field appears, with value of the order of
E ∼ (mπ /me )2 Ec , where Ec = m2e c3 /(e~) ≈ 1.3 × 1016 Volt cm−1 is the critical field for vacuum polarization. However,
this cores are stable against e+ e− pair creation, thanks to the Pauli blocking due to the electronic layer [46]. In this
layer, the particle densities decrease until the point where global charge neutrality is reached and the crust is found.
Consequently, the core is matched to the crust via this interface at a density ρcrust ≤ ρnuc . In the limit ρcrust → ρnuc ,
the thickness of the transition layer, as well as the electric field inside it, vanishes and the solution approaches the one
given by local charge neutrality (see Figure 3 and 5 in [31]). Configurations with ρcrust = ρdrip divide NSs with and
without inner crust. For stars with density in the core-crust interface ρcrust ≤ ρdrip , namely having only the so-called
outer crust, we used the BPS EOS [47], while for stars having inner crust too, we used, for such region, the BBP EOS
[48].
We introduced the uniform rotation in [32] within the Hartle formalism [49]. From the integration of the equations
of equilibrium, we computed in [32], for different central densities ρc and circular angular velocities Ω, the mass M,
polar R p and equatorial Req radii, angular momentum J, eccentricity ǫ, moment of inertia I, as well as quadrupole
moment Q of the configurations.
The angular momentum J of the star is given by

1 c2 4 dω̄
!
J= R , (37)
6G dr r=R
which is related to the angular velocity Ω by

2G2 J
Ω = ω̄(R) + , (38)
c5 R3
where R is the total radius of the non-rotating star and ω̄(r) = Ω − ω(r) is the angular velocity of the fluid relative to
the local inertial frame, with ω as the fluid angular velocity in the local inertial frame.
The total mass of the configuration is

G2 J 2
M = M0 + δM , δM = m0 (R) + , (39)
c7 R3
where M0 is the mass of the non-rotating star and δM is the contribution to the mass due to the rotation, while m0 is a
second order contribution to the mass related to the pressure perturbation.
The moment of inertia can be computed from the relation
J
I= , (40)

which does not account for deviations from spherical symmetry since within the Hartle formalism J is a first order
function of Ω. It is possible to follow such approximation, because, thanks to the high density of NSs, most of the
observed pulsars are accurately described by a perturbed spherical geometry. Indeed, considering for example the
sequence of configurations with period P = 10 s, shown in Figure 9, it is clear as it practically overlaps the non-
rotating mass-radius relation. The accuracy of the approximation increases for stiffer EOS as the ones given by σ-ω-ρ
relativistic nuclear mean field models, (see [50] for details).
In Figure 9, we show the mass-radius relation that results from the integration of the EMTF equations for the
equilibrium configurations of static and rotating NSs, (see [32] for further details).

Pulsar Properties For Realistic EOS


In this section, we focus on the high-magnetic field pulsar class [39], even if our general qualitative results apply to
all pulsars, to analyze the consequences of using realistic general relativistic structure parameters on the inference of
the magnetic field and efficiency of a pulsar in converting rotational energy into electromagnetic radiation.
30. GCN : Static

P
GCN : Keplerian

25. GCN : =10 s


GCN : Secular Inst
LCN : Static
.
P
LCN : Keplerian

. .
LCN : =10 s
LCN : Secular Inst
20 M JJ ≠
=0
(GCN)
M
min
0
min
(GCN)

M/M ⊙ 15.
10.
05.
00. 9 11 13 15 17
R eq (km)

FIGURE 9. Total mass versus total equatorial radius for the global (red) and local (blue) charge neutrality cases. The dashed curves
represent the static configurations, while the solid lines are the corresponding uniformly maximally rotating (Keplerian) NSs. The
light red and light blue colored lines define the secular instability boundary for the globally and locally neutral cases, respectively.
The horizontal thin red lines define the minimum mass in the globally neutral case. The dots refer to the sequence of constant
period P = 10 s.

To improve the description of the pulsar’s properties, we will introduce a correction to the the magnetic field, by
introducing the finiteness of the mass and size of the star, as measured by the fastness parameter, ΩR/c. The exact
solution of the exterior electromagnetic fields of a (slowly) rotating magnetic dipole aligned with the rotation axis in
general relativity was first found in [51, 52], (see also [53]). There, were solved the Einstein-Maxwell equations in the
Schwarzschild background. The generalization to a general electromagnetic multipolar structure in a Schwarzschild
metric was found in [53]. The solution for the general relativistic case in the slow rotation regime, and for a general
misaligned dipole, was obtained in analytic form in the near zone (r ≪ c/Ω = 1/k = λ/2π, where k is the wave
number and λ the wavelength) in [54, 55] and, for the wave zone (r ≫ c/Ω = 1/k = λ/2π) in [56]. In the latter work,
the radiation power of the dipole was computed as
!2
2 µ2 Ω4 f
G.R.
Pdip = − ⊥3 , (41)
3 c N2

where f and N are the general relativistic corrections


!3 " #
3 R 2M0  M0 
f = − ln(N 2 ) + 1+ , (42)
8 M0 R R
r
2M0
N = 1− . (43)
R
Equating the rotational energy loss to the above electromagnetic radiation power it is possible to obtain the formula
of the surface magnetic field analogous to Equation (34) but with general relativistic corrections:
!1/2
N 2 3c3 I
B sin χ = P Ṗ . (44)
f 8π2 R6

In Figure 10 we have plotted the ratio of the magnetic field obtained via the Newtonian formula (34) and the
general relativistic formula (44) to the fiducial value obtained with (36), for the realistic mass-radius relations of
globally and locally neutral NSs used in this work.
12.
GCN LCN
10.
08.
.
B/Bf
06

04.
02.
. .
. .. ..
Newt Newt
GR GR
00
.
00 .
05 .
10 . 15 .
20 .
25 .
30
M/M ⊙
FIGURE 10. Ratio of the magnetic field given by the Newtonian formula (34) and the general relativistic one (44) to the fiducial
value given by Equation (36). We have used here the realistic mass-radius relations of globally and locally neutral static NSs of
this work and an inclination angle χ = π/2.

We can see from this figure that, in the Newtonian case, the inferred magnetic field increases with increasing NS
mass. This means that the configurations of maximum and minimum mass give us respectively, upper and lower limits
to the magnetic field. It is worth noticing how the general relativistic formula gives us a magnetic field lower than the
Newtonian counterpart, for M/M⊙ & 1.1 and M/M⊙ & 1.2 for the globally and locally neutral configurations respec-
tively. Moreover, we can notice how the magnetic field is extremely close, at very low masses, with the Newtonian
value, as expected; then, it deviates and reaches a maximum value for some value of the mass, and then decreases
for increasing masses. The magnetic field inferred from globally and locally neutral configurations coincides for large
masses close to the critical mass value, as it should be expected since in those massive configurations the structure
parameters are dominated by the NS core, with very little role of the crust. We are here using the parameters of the
static configurations as previously explained.
In Figure 11 we plotted our theoretical prediction for magnetic fields of the pulsars of Table 2 as a function of
the NS mass, using the general relativistic formula (44).
We find that, both in global and local neutrality case, the assumed high-B pulsars have inferred magnetic fields
lower than the critical value for the entire range of NS masses.
On regard the efficiency of pulsars in converting rotational energy into electromagnetic radiation, we show in
Fig. 12 the X-ray luminosity to rotation energy loss ratio, LX /Ėrot , as a function of the NS mass, for both global and
f f
local charge neutrality. We also present in Table 2 the ratio LX /Ėrot , where Ėrot is the rotational energy loss as obtained
from fiducial NS parameters given by Equation (35).
We find that for both globally and locally neutral NSs we have LX < Ėrot in the entire range of stable masses for
any souce.
The only exception to the above result are PSR J1847–0130 and PSR J1819–1458, for which no range of masses
with LX < Ėrot was obtained. However, for PSR J1847–0130 we have only an upper limit for LX , so there is still
room for solutions with LX < Ėrot if future observations lead to an observed value lower than the present upper limit.
In this line, the only object with LX > Ėrot for any mass is PSR J1819–1458. For this particular object there is still
the possibility of being a rotation powered NS since the currently used value of the distance to the source, 3.6 kpc,
inferred from its dispersion measure, is poorly accurate with a considerable uncertainty of at least 25%, see [57] for
details. Indeed, a distance to the source 25% shorter than the above value would imply LX < Ėrot for this object in the
mass range M0 & 0.6 M⊙ .
f
We notice that the efficiency obtained via fiducial parameters, LX /Ėrot , is larger than the actual value obtained
with the realistic NS structure in the entire range of stable masses; see Table 2 and Fig. 12.
J1847−0130 J1734−3333 J1819−1458 J1847−0130 J1734−3333 J1819−1458
J1718−3718 J1814−1744 J1846−0258 J1718−3718 J1814−1744 J1846−0258
.
0 30 .
0 30
GCN LCN
0.25 0.25

0.20 0.20
c

c
B . ./B

B . ./B
0.15 0.15
GR

GR
0.10 0.10

0.05 0.05

0.00 0.00
0. 0 05. .
10 1. 5 .
20 .
25 0. 0 05. .
10 1.5 .
20 .
25
M /M ⊙
0 M /M ⊙
0

FIGURE 11. Magnetic field BG.R. obtained from the general relativistic magneto-dipole formula (44), in units of critical magnetic
field Bc , as function of the mass (in solar masses) for static NSs in the global (left panel) and local (right panel) charge neutrality
cases.

J1847 −0130 up
J1814 −1744up
−0258 . .
J1846
min
J1847 −0130 up
J1814 −1744up
−0258 . .
J1846
min

J1819 −1458 max


J1718 −3718min
−0258
J1846
AO
max
J1819 −1458 max
J1718 −3718min
−0258
J1846
AO
max

J1819 −1458 min


J1734 −3333max
−0258
J1846
min
J1819 −1458 min
J1734 −3333max
−0258
J1846
min

J1718 −3718 max


J1846 −0258 . .
max
AO −3333
J1734
min J1718 −3718 max
J1846 −0258 . .
max
AO −3333
J1734
min
4 4
10 10

10
3 GCN 10
3 LCN
2 2
10 10
1 1
10 10
rot

rot
LX /Ė

LX /Ė

0 0
10 10

10
−1 10
−1

10
−2 10
−2

10
−3 10
−3
.
05 .
10 . 15 .
20 25. .
05 .
10 . 15 .
20 25.
M /M ⊙
0 M /M ⊙
0

FIGURE 12. Ratio between the observed X-ray luminosity LX and the loss of rotational energy Ėrot versus total mass of the rotating
NS, in units of M⊙ . Are drawn the high-B pulsar from the work by Ng and Kaspi [39] for which a magnetic field higher than the
critical field Bc is inferred, once the fiducial value for the moment of inertia I = 1045 g cm2 is taken into account (see Table 2).
Pulsars with luminosity LX defined by an upper limit are labeled with “up”, for pulsars with luminosity LX not well established we
have assumed the existent lower limits (label “min”) and upper limits (label “max”) on it. The values for the pulsar PSR J1846-0258
are dived in prior the 2006 outburst and after the 2006 outburst (label “A.O.”). Left plot: global charge neutrality. Right plot: local
charge neutrality. The magnetic fields shown are referred to the high-magnetic field pulsars of Table 2.
It is also worth to mention that the rotation energy loss (32) depends on the NS structure only through the
moment of inertia, whose quantitative value can be different for different nuclear EOS and/or owing to an improved
value accounting for deviations from the spherical geometry, for instance considering a third-order series expansion
in Ω. However, the latter effect is negligible for this specific case (P ≈ 4.3 s), see for instance Figure 5 in [50], where
no deviations of I from its spherical value appear for such long rotation periods.

ACKNOWLEDGMENTS
C. C. and S. F. would like to acknowledge GNFM-INdAM and ICRANet for partial support. It is a pleasure to
thank D. P. Menezes and R.C.R. de Lima for discussions on the equation of state of NSs and for supplying the
EOS tables. R.B., S.M.C. and J.A.R. acknowledge the support by the International Cooperation Program CAPES-
ICRANet financed by CAPES – Brazilian Federal Agency for Support and Evaluation of Graduate Education within
the Ministry of Education of Brazil. R.N. acknowledges financial support by CAPES and CNPq. J.P.P. acknowledges
CNPq–Conselho Nacional de Desenvolvimento Cientı́fico e Tecnológico within the postdoctoral program “Science
without Borders” of the Brazilian government for the financial support. He is also grateful for the invitation to lecture
at the second César Lattes meeting in Rio de Janeiro and Niterói, Brazil, about the issue of this work.

REFERENCES
[1] N. Stergioulas, and J. L. Friedman, ApJ pp. 306–311 (1995).
[2] N. Stergioulas, Living Reviews in Relativity p. 3 (2003).
[3] H. Komatsu, Y. Eriguchi, and I. Hachisu, MNRAS pp. 355–379 (1989).
[4] G. B. Cook, S. L. Shapiro, and S. A. Teukolsky, ApJ pp. 203–223 (1992).
[5] G. B. Cook, S. L. Shapiro, and S. A. Teukolsky, ApJ pp. 823–845 (1994).
[6] G. Baym, C. Pethick, and P. Sutherland, ApJ p. 299 (1971).
[7] G. A. Lalazissis, J. König, and P. Ring, Phys. Rev. C pp. 540–543 (1997).
[8] Y. Sugahara, and H. Toki, Nuclear Physics A pp. 557–572 (1994).
[9] N. K. Glendenning, and S. A. Moszkowski, Physical Review Letters pp. 2414–1417 (1991).
[10] S. Pal, D. Bandyopadhyay, and W. Greiner, Nuclear Physics A pp. 553–577 (2000).
[11] J. Boguta, and A. R. Bodmer, Nuclear Physics A pp. 413–428 (1977).
[12] J. L. Friedman, J. R. Ipser, and R. D. Sorkin, ApJ pp. 722–724 (1988).
[13] F. Cipolletta, C. Cherubini, S. Filippi, J. A. Rueda, and R. Ruffini, Phys. Rev. D. in press (2015), 1506.05926.
[14] R. Geroch, Journal of Mathematical Physics pp. 1955–1961 (1970).
[15] R. Geroch, Journal of Mathematical Physics pp. 2580–2588 (1970).
[16] R. O. Hansen, Journal of Mathematical Physics pp. 46–52 (1974).
[17] F. D. Ryan, Phys. Rev. D pp. 5707–5718 (1995).
[18] G. Pappas, and T. A. Apostolatos, Physical Review Letters p. 231104 (2012).
[19] K. Yagi, K. Kyutoku, G. Pappas, N. Yunes, and T. A. Apostolatos, Phys. Rev. D p. 124013 (2014).
[20] J. P. Pereira, and J. A. Rueda, ApJ 801, 19 (2015), 1501.02621.
[21] N. Glendenning, Compact Stars. Nuclear Physics, Particle Physics and General Relativity., 1996.
[22] A. Iwazaki, O. Morimatsu, T. Nishikawa, and M. Ohtani, Phys. Rev. D 71, 034014 (2005), hep-ph/0404201.
[23] M. G. Alford, A. Schmitt, K. Rajagopal, and T. Schäfer, Reviews of Modern Physics 80, 1455–1515 (2008),
0709.4635.
[24] C. K. Raju, Journal of Physics A Mathematical General 15, 381–396 (1982).
[25] C. K. Raju, Journal of Physics A Mathematical General 15, 3915 (1982).
[26] S. L. Shapiro, and S. A. Teukolsky, Black Holes, White Dwarfs and Neutron Stars: The Physics of Compact
Objects, 1986.
[27] M. A. Rodrı́guez-Valverde, M. A. Cabrerizo-Vı́lchez, and R. Hidalgo-Álvarez, European Journal of Physics
24, 159–168 (2003).
[28] J. P. Pereira, J. G. Coelho, and J. A. Rueda, Phys. Rev. D 90, 123011 (2014), 1412.1848.
[29] M. Rotondo, J. A. Rueda, R. Ruffini, and S.-S. Xue, Physics Letters B 701, 667–671 (2011).
[30] J. A. Rueda, R. Ruffini, and S.-S. Xue, Nuclear Physics A 872, 286–295 (2011).
[31] R. Belvedere, D. Pugliese, J. A. Rueda, R. Ruffini, and S.-S. Xue, Nuclear Physics A 883, 1–24 (2012).
[32] R. Belvedere, K. Boshkayev, J. A. Rueda, and R. Ruffini, Nuclear Physics A 921, 33–59 (2014).

View publication stats

Anda mungkin juga menyukai