You are on page 1of 21

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

net/publication/364953477

Bosonic Dark Matter in Light of the NICER Precise Mass-Radius Measurements

Preprint · October 2022


DOI: 10.48550/arXiv.2210.17308

CITATIONS READS
0 13

2 authors, including:

Davood Rafiei Karkevandi


Isfahan University of Technology
4 PUBLICATIONS   19 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Davood Rafiei Karkevandi on 02 November 2022.

The user has requested enhancement of the downloaded file.


Draft version November 1, 2022
Typeset using LATEX default style in AASTeX63

Bosonic Dark Matter in Light of the NICER Precise Mass-Radius Measurements

1, 2 3
Soroush Shakeri and Davood Rafiei Karkevandi
1 Department of Physics, Isfahan University of Technology, Isfahan 84156-83111, Iran
2 ICRANet,Piazza della Repubblica 10, I-65122 Pescara, Italy
3 ICRANet-Isfahan, Isfahan University of Technology, Isfahan 84156-83111, Iran
arXiv:2210.17308v1 [astro-ph.HE] 31 Oct 2022

ABSTRACT
We explore the presence of self-interacting bosonic dark matter (DM) whithin neutron stars (NSs)
in light of the latest mass-radius measurements of the Neutron Star Interior Composition Explorer
(NICER). The bosonic DM is distributed in NSs as a core for DM particles with high mass, low
fraction and low self-coupling constant or as a halo for particles with low mass, high fraction and high
self-coupling constant leading to formation of DM admixed NSs. We focus on the evolution of the
visible and dark radius of the mixed object due to DM model parameters and fractions. It is shown
that DM core formation reduces the visible radius and total mass pushing them below observational
limits while halo formation is in favor of the latest mass-radius constraints. Applying joint constraints
for radius of 1.4M NSs and the maximum mass from NICER and LIGO/Virgo observations, we scan
over the parameter space of the bosonic DM model to obtain an allowed region. It turns out that the
maximum mass limit provides a more stringent constraint compare to the radius one. Our investigation
allows for the exclusion of a large portion of DM fractions for sub-GeV bosons and limited the amount
of accumulated DM to be less than ∼ 4% for the entire mass range at the strong coupling regime λ = π.
In this paper, we introduce main features of the pulse profile corresponding to the DM admixed NS
and the effect of DM halo on the light bending is considered extensively as an independent probe for
the DM model. We find that the depth of minimum fluxes in the pulse profiles crucially depend on the
amount of DM distributed around NS and its compactness. The current/future astrophysics missions
via precise measurements of compact object properties may test the possibility of the existence of DM
within NSs and break the degeneracies between different scenarios to interpret exotic observations.

1. INTRODUCTION
The advent of multi-messenger observations of compact astrophysical objects have provided a unique opportunity to
explore the dense matter equation of state (EOS) and to answer the question, whether neutron stars (NSs) are composed
of normal nuclear matter, or they contain more exotic type of matter. In this regard, the joint precise measurements
of mass, radius and tidal deformability of NSs by electromagnetic (EM) and gravitational waves (GWs), have led to
remarkable improvements in our understanding of their interior structures Most et al. (2018); Fattoyev et al. (2018);
Annala et al. (2018); Rezzolla et al. (2018). Among different possibilities, dark matter (DM) either as a bosonic
or fermionic particle could be mixed with ordinary matter leading to a new type of compact objects so called DM
admixed NS Tolos et al. (2015); Ellis et al. (2018b); Nelson et al. (2019); Ivanytskyi et al. (2020); Dengler et al. (2021);
Karkevandi et al. (2022). Depending on the DM model parameters, i.e. mass, self-coupling constant and its fraction
within the mixed star, DM could reside as a dense core inside NS or form an extended halo around it. The distribution
of DM in the DM admixed NS causes some noticeable impacts on the mass-radius profile and the deformation of this
configuration in a binary system which encodes in the tidal deformability parameter Karkevandi et al. (2022, 2021).
Among different DM candidates the scalar/pseudoscalar bosons such as axions are of great interests from various
aspects in astrophysics and cosmology Marsh (2016); Shakeri & Hajkarim (2022). In this regard, self-gravitating
objects made of bosonic DM could be formed whether as non self-interacting scalar fields or with repulsive/attractive

s.shakeri@iut.ac.ir, d.rafiei@alumni.iut.ac.ir
2 Shakeri & Karkevandi

self-interactions Agnihotri et al. (2009); Eby et al. (2016); Wystub et al. (2021); Visinelli (2021). In the absence of
interaction between bosons, the Heisenberg uncertainty principle provides the required quantum pressure to avoid from
a gravitational collapse. The stable self-gravitating configurations of complex scalar fields called boson stars (BSs) were
first introduced by Kaup (1968) and Ruffini & Bonazzola (1969). Afterwards, Colpi et al. (1986) considered a repulsive
self-interaction among bosons which alters the mass and radius of BSs drastically and become comparable to the ones
of NSs. Following these efforts, the attention has been directed to the possible role of BSs in astrophysics whether as a
DM candidate Jetzer (1992); Arbey et al. (2003); Amaro-Seoane et al. (2010); Liebling & Palenzuela (2017); Chavanis
(2019) or as black hole mimickers Mielke & Schunck (2000); Guzman & Rueda-Becerril (2009); Bustillo et al. (2021);
Chavanis (2020).
Asymmetric Dark Matter (ADM) which usually do not annihilate to standard model (SM) particles can be substan-
tially accumulated in NSs through different processes Kouvaris & Tinyakov (2011a,b); Li et al. (2012); Nelson et al.
(2019); Ivanytskyi et al. (2020); Del Popolo et al. (2020b,c); Ciancarella et al. (2021); Berezhiani et al. (2021); Bell et al.
(2021); Kalashnikov & Chechetkin (2022). Accretion of bosonic ADM by stars has been discussed by nonlinear numer-
ical relativity methods, where it was showed that bosonic cores may develop inside the star and these configurations
are stable throughout most of the parameter space Brito et al. (2015, 2016). We may also have gravitationally stable
configurations composed entirely of fermionic or bosonic ADM named “dark stars” Narain et al. (2006); Kouvaris &
Nielsen (2015); Eby et al. (2016); Maselli et al. (2017); Visinelli et al. (2018); Chavanis (2018). ADM stars could also
serve as DM cores and accretion centers for baryonic matter (BM) which may lead to the formation of a DM admixed
NS with high fractions of DM Goldman et al. (2013); Xiang et al. (2014); Ellis et al. (2018a,b). Furthermore, a large
amount of DM maybe provided in a binary system of a dark compact object and a NS, by capturing of DM particles
or the coalescence of these objects Sandin & Ciarcelluti (2009); Ciarcelluti & Sandin (2011); Giudice et al. (2016);
Gresham & Zurek (2019).
On the other hand, the existence of self-annihilating DM in NS lead to some evidences such as the variation of
luminosity, effective temperature and cooling of the mixed object Kouvaris (2008); de Lavallaz & Fairbairn (2010);
Ángeles Pérez-Garcı́a et al. (2022). In this regard, the decay of DM into SM particles can behave as a power source
in a new type of the first stars, instead of the nuclear fusion in ordinary stars, which make them very bright. These
astrophysical objects which are composed mostly of ordinary matter called also dark stars in which DM heating
prevents the star from collapsing Wu et al. (2022). In both above cases the observable effects produced by DM could
be detectable by James Webb Space Telescope (JWST) at very high redshift Spolyar et al. (2008); Freese et al. (2010);
Baryakhtar et al. (2017); Raj et al. (2018); Chatterjee et al. (2022); Baryakhtar et al. (2022)
Astrophysical objects which are a mixture of fermions and bosons, interacting only gravitationaly, have been discussed
based on two different methods: i) fermion-boson star Henriques et al. (1989, 1990b,a); Valdez-Alvarado et al. (2020);
Di Giovanni et al. (2020); Nyhan & Kain (2022) and ii) DM admixed NS Sandin & Ciarcelluti (2009); Ciarcelluti
& Sandin (2011); Leung et al. (2011); Ellis et al. (2018b); Nelson et al. (2019). In both above cases, the energy
momentum tensor of the Einstein equation takes separate contributions of fermionic and bosonic components. For the
former, fermionic component is described by a perfect-fluid while the scalar bosonic field is governed by Klein–Gordon
equation. In order to compute equilibrium configurations of fermion-boson stars, one needs to solve the coupled
Einstein-Klein–Gordon equations Liebling & Palenzuela (2017). In the DM admixed NSs, fermions and bosons are
considered as perfect fluids where the two-fluid formalism of Tolman-Oppenheimer-Volkof (TOV) equations are utilized
to describe the mixed compact star. Recently, the equilibrium configurations of DM admixed NSs have been extensively
considered in Karkevandi et al. (2022, 2021), where the DM component was modeled by self-repulsive bosons. In a
related study for the fermion-boson stars Di Giovanni et al. (2021), which was done following the similar scheme in
Karkevandi et al. (2022) but with a vanishing bosonic self-interaction, the dependence of DM core/halo formation to
the particle mass is in agreement with the results presented in Karkevandi et al. (2022).
During the last decade the measurement techniques of the mass and radius of NSs have been much improved,
the Shapiro delay method or orbital measurements of binary systems determine the gravitational mass, such as the
measured mass of the heaviest NSs, PSR J0348+0432 (2.01 ± 0.04M ) Antoniadis et al. (2013) and PSR J0740+6620
(2.14+0.10
−0.09 M ) Cromartie et al. (2019). In this direction, the Keck-telescope optical spectrophotometry and imaging
of the companions of PSR J1810+1744 Romani et al. (2021) and PSR J0952-0607 Romani et al. (2022) provide novel
mass results 2.13 ± 0.04M and 2.35 ± 0.17M , respectively. Furthermore, the GW data of LIGO-Virgo collaboration
with the tidal deformability measurement of the event GW170817 Abbott et al. (2017, 2018) enables us to constrain
the radius of 1.4M NS, hereafter called R1.4 Fattoyev et al. (2018); Most et al. (2018); Annala et al. (2018); Tews
3

et al. (2018); De et al. (2018); Radice & Dai (2019); Coughlin et al. (2019); Capano et al. (2020); Essick et al. (2020);
Breschi et al. (2021); Gamba et al. (2021); Pang et al. (2022). Afterwards, the Neutron Star Interior Composition
Explorer (NICER) announced the joint precise measurements of mass and radius for PSR J0030+0451 and PSR
J0740+6620 Miller et al. (2019); Riley et al. (2019); Miller et al. (2021); Riley et al. (2021), which opens a new window
towards better understanding of NSs structure. Ever since, many attempts have been made to impose limits on dense
matter EoSs and NS properties specially R1.4 via combined analyses of GW, X-ray observations and nuclear theory or
experiments Raaijmakers et al. (2019, 2020, 2021); Bogdanov et al. (2019); Dietrich et al. (2020); Reed et al. (2021);
Pang et al. (2021); Al-Mamun et al. (2021). Considering the aforementioned results from multi-messenger observation
+0.86
of NSs, gives the value of R1.4 to be 11.75−0.81 km at 90% confidence level at 1σ uncertainty Dietrich et al. (2020) or
more recent work indicating that R1.4 = 12.01+0.78 −0.77 km at 95% confidence level Huth et al. (2022). Moreover, Pang
et al. (2021) estimated the radius of a typical 1.4M NS to be 11.94+0.76 −0.87 at 90% confidence and in Breschi et al.
(2021), an average over all current estimations has been done, giving R1.4 = 12.0+1.2 −1.2 with the 90% credible region.
In this paper, we mainly focus on the precise radius measurements in addition to the maximum mass parameter
to probe DM admixed NSs. We take MTmax & 2M as the total maximum mass constraint and R1.4 & 11 km as a
conservative limit for the visible radius of NSs composed of a fraction of bosonic self-interacting DM (SIDM). In the
following, we investigate the effect of DM core/halo formation on the visible and dark radius of the mixed object in the
light of the latest multi-messenger constraints. We examine different DM model parameters such as the particle mass,
self-coupling constant and also the amount of DM within the mixed NS in order to check their consistency with the
allowed parameter space given by NICER. We show that the existence of bosonic DM as a core of a composed object
leads to a significant reduction in the radius and mass of the mixed star which is disfavored by recent measurements.
Moreover, an upper limit of DM fraction in NSs has been placed for the whole sub-GeV mass range of bosonic SIDM
to be in agreement with the latest joint radius and mass measurements of compact objects.
In fact, the precise mass-radius measurements by X-ray telescopes (e.g. NICER/XMM-Newton) rely on the surface
emission of the star. Depending on the compactness of a NS, the path of the photons is distorted by the gravitational
warping of spacetime in the vicinity of the object, leading to some modifications in the visibility of hot spots and the
corresponding pulse profiles (Beloborodov 2002; Turolla & Nobili 2013; Ozel et al. 2016). As a new observable, we
extensively consider the variation of the EM pulse profile in a NS with a DM halo around it for different DM model
parameters. Recently, this effect has been investigated in Miao et al. (2022), where they observed a peak flux deviation
from pure NS which depends on the ratio of the DM halo mass to its radius. In the current study, we present a more
detailed analysis of Pulse Profile Modeling (PPM) taking into account different amount of DM and boson masses.
Moreover, the compactness of DM admixed NSs, which is a crucial parameter for PPM, has been investigated for a
wide range of boson masses considering various DM factions and self-coupling constants. Our results demonstrate the
necessity of inclusion of the DM admixed NS among other possibilities in order to obtain posterior distributions of
samples during analysis of NICER/XMM-Newton.
This paper is organized as follows. In Sec. 2 we present a framework to model the DM admixed NS based on two-fluid
TOV formalism which includes two EoSs describing bosonic DM and BM components. The equilibrium configurations
of the mixed compact object through the mass-radius profiles are compared with the latest measurements of NICER
(Sec. 3). We consider the evolution of the visible and dark radius of DM admixed NSs under the influence of the boson
mass and self-interaction strength as a function of DM fractions in Sec. 4. The pulse profile as a new observable will
be introduced in Sec. 5 where the effect of DM halo will be taken into account in light bending in order to track the
surface emission of a star with a baryonic core surrounded by a DM halo. In Sec. 6 a scan over bosonic DM mass
and fraction is presented regarding two key observable features MTmax ≥ 2M and R1.4 ≥ 11 km. Finally, in the last
section we summarize our results. In this paper we use units in which ~ = c = G = 1.
2. MODELING OF THE DM ADMIXED NS
In this work to model DM, we use bosonic SIDM which is described through the following Lagrangian
1 m2χ ∗ λ
L= ∂ µ φ∗ ∂ µ φ − φ φ − (φ∗ φ)2 , (1)
2 2 4
we assume mean-field approximation and expand the interaction term λ(φ∗ φ)2 /4 in terms of φ∗ φ − hφ∗ φi which leads
to a linearized mean-field Lagrangian as
1 m∗2
χ λ
LM F = ∂µ φ∗ ∂ µ φ − φ∗ φ + hφ∗ φi2 , (2)
2 2 4
4 Shakeri & Karkevandi

as a result the effective mass m∗χ depends on the expectation value of φ∗ φ as m∗2 2 ∗
χ = mχ + λhφ φi.
The pressure arise from the DM perfect fluid described by this Lagrangian is

λ ∗ 2
P = ζ 2 (µ2χ − m∗2
χ )+ hφ φi , (3)
4
where ζ represents the amplitude of zero mode and µχ is bosonic chemical potential. After some straightforward
computations presented previously in Appendix of Karkevandi et al. (2022), it turns out that the total pressure of
SIDM particles is given by
s !2
m4χ 3λ
P = 1+ 4ρ−1 . (4)
9λ mχ
This EoS has been originally obtained by Colpi et al. (1986) for boson stars (BSs) in an alternative way by solving
Klein-Gordon-Einstein equations for a spherically symmetric configuration of the scalar fields. In the strong-coupling
limit λ  4π(mχ /MP l )2 applied in Colpi et al. (1986), the spatial derivatives of the scalar fields have been neglected
leading to a local solution of field equations. In this regime the system can be approximated as a prefect fluid and the
anisotropy of pressure is ignored. The maximum mass of a BS with the quartic self-interaction was found to be
 2
BS 1/2 100 MeV
Mmax ≈ 10 M λ , (5)

it was shown Amaro-Seoane et al. (2010) that the maximum compactness Cmax = (M/R)max grows with the coupling
constant λ and reaches to a saturated value 0.16 at strong coupling limit. Working in this regime, we can find the
minimum radius Rmin corresponding to Mmax as

 2
BS 1/2 100 MeV
Rmin ≈ 92.3 km λ , (6)

For the BM component, we utilize a realistic EoS including induced surface tension (IST) where both the short-range
repulsion and long-range attraction interactions between baryons have been taken into account Sagun et al. (2019b,a).
This EoS is in agreement with the latest NS observations providing the maximum mass Mmax = 2.08M and the
radius 11.37 km fora NS with mass 1.4M Sagun et al. (2020). In our work the crust part of the NS is described
via the polytropic EoS with γ = 4/3 Ivanytskyi et al. (2020). In order to model DM admixed NSs, we assume an
µν µν
energy-momentum tensor composed of BM and DM fluids T µν = TDM + TBM with only gravitational interaction
between them, then the Einstein equation turns out to be

1 µν µν
Gµν = Rµν − Rgµν = 8π(TDM + TBM ). (7)
2
Each energy-momentum tensor is conserved separately giving rise to a set of the equations of motion called two-fluid
formalism of Tolman-Oppenheimer-Volkof (TOV) Tolman (1939); Oppenheimer & Volkoff (1939); Sandin & Ciarcelluti
(2009); Ciarcelluti & Sandin (2011).

dpB m + 4πr3 (pB (r) + pD (r))


= − (pB + B ) , (8)
dr r(r − 2m)
dpD m + 4πr3 (pB (r) + pD (r))
= − (pD + D ) , (9)
dr r(r − 2m)
dm
= 4πr2 (B + D ) , (10)
dr
where m = mD (r) + mB (r) and B and D indices stand for BM and DM components. The solution of above equations
describe the hydrostatic configuration of DM admixed NSs which express the DM distribution within NSs lead to three
5

possibilities (i) DM resides as a core inside NS, (ii) DM froms a halo around NS (iii) DM dispersed through the whole
NS. The total mass of the object in all the cases is defined by

MT = mB (RB ) + mD (RD ) , (11)

where mB (RB ) and mD (RD ) are the enclosed masses of BM and DM, respectively. The gravitational radius of the
DM admixed NS is defined by the radius of BM fluid (RB ) for the DM core formation, and is determined based on
the DM halo radius (RD ) for the DM halo formation, while for the third case RB ≈ RD . The fraction of DM inside
the mixed object is defined as Fχ = mDM(R T
D)
which is realized as a model parameter in our computation. In principle,
there are various accumulation scenarios for DM in NSs leading to different values for Fχ (see Sec. VII of Karkevandi
et al. (2022)).
Since the radius of BM component RB is detectable through surface EM radiations, in this work the total mass of
the DM admixed NS is shown as a function of the visible radius RB . This approach has been utilized extensively in
previous studies Tolos et al. (2015); Ellis et al. (2018b); Deliyergiyev et al. (2019); Ivanytskyi et al. (2020); Di Giovanni
et al. (2021); Dengler et al. (2021); Lee et al. (2021). However, in a recent novel study presented in Karkevandi et al.
(2022), the mass-radius (M-R) relation were plotted in terms of the outermost radius of DM admixed NSs instead of
visible radius RB . Knowing that the outermost radius can be whether RB or RD along with stable M-R sequence
depending on DM model parameters and the amount of DM in the star. Both the above mentioned approaches are
identical in the case of DM core formation where both outermost radius of the object and the radius of BM component
are the same.
It has been shown that SIDM particles with self-coupling constant in order of π and for the mass range mχ . 200
MeV distribute as a halo around NS while for heavier sub-GeV masses with relatively low DM fractions, a core inside
the NS will be formed Karkevandi et al. (2022). Note that core and halo formation crucially depend on the DM
fraction, where the DM core causes a decrease in the total mass, radius and tidal deformability of the compact object
while the DM halo formation increases all of these quantities. It was seen that for a given mass and self coupling
constant, by growing Fχ a transition like behaviour from DM core to DM halo can occur in DM admixed NSs.

3. THE MASS-RADIUS PROFILE OF THE DM ADMIXED NS AND NICER CONSTRAINTS


In this section, we consider the bosonic DM model parameters, i.e. mass of bosons (mχ ), self-coupling constant (λ)
and the DM fraction (Fχ ) via the M-R profile of DM admixed NSs. In the following, we compare our results with the
latest mass-radius measurements of NICER Miller et al. (2019); Riley et al. (2019); Miller et al. (2021); Riley et al.
(2021). In figures 1 and 2, M-R credible regions ( dark: 68% (1σ), light: 95% (2σ) ) for PSR J0030+0451 and PSR
J0740+6620 measured by NICER team are colored in blue and red, respectively. As it is shown, according to the
NICER results larger radii (i.e. R & 11km) are more favorable. The gray dashed lines illustrate the maximum mass
limit for NSs, M = 2M and the black solid lines show the M-R relation for the pure BM component (IST EoS).
In Fig.1, the coupling constant is fixed at λ = π and various boson masses and DM fractions are considered. Left
panel shows the effect of the boson mass varied from 100-500 MeV for Fχ = 20%. It is seen that by increasing mχ
which corresponds to DM core formation, the DM admixed NS model disfavors NICER results, more specifically, for
mχ = 400, 500 MeV, the M-R profiles are completely outside the red and blue regions. Note that light bosons, for
which a DM halo is formed, e.g. mχ = 100 MeV, are in agreement with NICER data and even cause the M-R stable
sequence to be more compatible with the red part in comparison with black solid curve (only BM). However, notice
that for DM halo formation specially for large RD , gravitational effects such as the tidal deformability should be taken
into account which has been discussed extensively in Karkevandi et al. (2022, 2021). In the right panel, we show that
how the DM fraction influences the M-R sequence of DM admixed NSs for DM core/halo configurations. Two boson
masses are chosen for which mχ = 100 MeV corresponds to DM halo formation (solid line) and mχ = 400 MeV leads
to a DM core inside NS (dashed line). We can see that increasing Fχ from 5% to 20% for 400 MeV bosons shifts
the M-R plots out of the allowed parameter space which is due to the DM core formation and decreasing of mass
and radius of the mixed object. It is depicted that for mχ = 100 MeV, the M-R diagrams are more compatible with
NICER data for higher Fχ , but it should be mentioned that these fractions cause the halo to be very large and could
affect other observable features of the mixed compact object.
Here we also examine the impact of various self-coupling constants between bosonic particles in DM admixed NSs.
Thus in Fig. 2, we vary λ from relatively low values 0.001π to higher ones up to π for mχ = 100 MeV and Fχ = 20%. It
can be seen that by decreasing coupling constant, the M-R curves move towards left and for the lowest value λ = 0.001π
6 Shakeri & Karkevandi

Figure 1. M-R profile for DM admixed NSs is considered for various boson masses and DM fractions at λ = π. In the left panel
total mass is presented as a function of the visible radius (RB ) for different mχ as labeled and Fχ = 20%. Right panel shows
the effect of the DM fraction on the M-R profile changing from 5% to 20% for mχ = 400 MeV (dashed lines) and mχ = 100
MeV (solid lines). Red and blue regions correspond to the latest NICER results and the gray dashed line shows the maximum
mass limit for NSs.

goes entirely out of the NICER permitted parameter space. In fact for low self-coupling constant, even light bosons
such as mχ = 100 MeV would reside as a DM core inside NSs which decreases the mass and radius of the compact
object in comparison with only BM.
In summary, we show that sub-GeV DM bosons with smaller masses and higher self-coupling constants cause the
DM admixed NS to be more compatible with the latest NICER mass-radius data. In addition, if a DM core is formed
inside NS, only low DM fractions are in agreement with PSRJ0030+0451 and PSR J0740+662 regions, while for DM
halo configuration those regions are respected in whole considered range of fractions. Thus, by considering the above-
mentioned conditions, the NS could contain bosonic DM as a core or halo and be consistent with NICER results. These
results are in excellent agreement with those presented in Rutherford et al. (2022) where a different self-interacting
asymmetric bosonic model Agnihotri et al. (2009); Nelson et al. (2019) has been adopted to describe DM component
of a mixed object, a more detailed comparison with our study will be presented in the following section.

Figure 2. M-R profile of DM admixed NSs for mχ = 100 MeV and Fχ = 20% considering various coupling constants as labeled.
Red and blue regions correspond to the latest NICER results and the gray dashed line shows the maximum mass limit for NSs.

4. THE RADIUS EVOLUTION OF THE DM ADMIXED NS


Concerning the distribution of DM in NSs as a core or halo, in this section we extensively examine the behaviour
of the visible and dark radius (RB and RD ) for various DM model parameters (i.e. mχ and λ) along with a range of
DM fraction. Note that the visible radius of the mixed compact object is always RB , so the corresponding radius of a
DM admixed NS with MT = 1.4M is compared with the lower observational limit of radius ' 11 km Dietrich et al.
(2020); Huth et al. (2022). The analyses presented in this section as well as the radius constraint R1.4 & 11 and the
maximum mass limit MT & 2M will be used in Sec. 6 in order to produce a scan plot over DM parameter space.
7

In Fig. 3, we show variation of the outermost radius of DM admixed NSs with MT = 1.4M as a function of Fχ
for different boson masses and self-coupling constants. It is seen that by increasing the DM fraction, the outermost
radius decreases up to a turning point and then it goes sharply upwards. In these plots, the radius of the mixed object
will change from the visible radius RB (solid lines) for which a DM core is formed, to RD where a DM halo is made
(dashed lines). In fact, we see that depending on mχ and λ, a DM core will be formed generally in low Fχ while by
increasing the amount of DM, the dark radius RD will grow and becoming larger than RB which indicates that a DM
halo is formed around NS. Therefore, a transition occurs at the turning point of the diagrams from DM core formation
(RB > RD ) to DM halo (RD > RB ) by increasing the fraction of DM. This critical point (RB ≈ RD ) is equivalent
to the formation of a DM admixed NS in which DM dispersed throughout the whole object and belongs to the third
family of the mixed object introduced in Sec 2. In the left panel of Fig. 3, mχ is varied from 100 MeV to 300 MeV for
fixed λ = π, we see that for lighter bosons, the DM halo is started to form in very low fractions, however, one can show
that for the massive boson mχ = 300 MeV the DM halo will not form even up to Fχ ' 30%. Moreover, the impact of
the coupling constant on the outermost radius is considered in the right panel where it shows that for higher coupling
constants the transition point and the DM halo appear at smaller values of Fχ .
In the following Figs. 4 and 5, the variation of RB and RD for the DM admixed NS with total gravitational mass
1.4M are shown separately as a function of Fχ for both DM core (solid lines) and DM halo (dashed lines) formations.
In these plots black dashed horizontal line indicates R1.4 = 11 km as a lower bound for the visible radius of the mixed
object.

Figure 3. Variation of the outermost radius (R) of DM admixed NSs with total mass 1.4M as a function of DM fractions.
Solid curves show the variation of RB (visible radius) for which a DM core is formed, however, dashed lines indicate how RD
varies when a DM halo is formed around NS. Different boson masses are considered as labeled for λ = π in the right panel while
left panel is related to mχ = 200 MeV and various coupling constants.

In Fig. 4, the impact of different boson masses mχ ∈ [100, 300] MeV are examined at a fixed coupling constant
λ = π for a given range of DM fractions Fχ ∈ (0, 20]%. In the left panel, we see that for both DM core and DM halo
formations RB is a decreasing function of Fχ where for higher mχ the more reduction rate is observed. The variation
rate for light bosons (mχ . 150 MeV) becomes slow down reaching to a constant value for larger Fχ , even a slight
growth could be seen in the visible radius at relatively high fractions for light masses such as mχ = 100 MeV. This
shows that while for massive bosons one can define an upper limit for the DM fraction to satisfy a given radial limit
(e.g. 11 km), for light enough bosons RB lies above the radius limit for the whole considered range of DM fractions.
It turns out that by increasing Fχ for mχ & 150 MeV, RB becomes less than 11 km which is not in favour of the
recent lower observational limit of R1.4 . As it is depicted, the DM core formation inside NSs (solid line) observes up to
relatively high fraction for heavy bosons while at tiny fractions we may have core formations even for the light bosons
mχ . 150 MeV. By doing more precise calculations one can find mχ ' 153.5 MeV as the heaviest boson for which
RB & 11 km condition is met for the whole range of DM fraction till 20% . Regarding the mass range mχ ∈ [100, 300]
MeV we find that for DM fraction Fχ . 5%, visible radii of the mixed objects are higher than 11 km which is in
agreement with the latest radius constraints for NSs. From the right panel of Fig. 4, we see that RD is an increasing
function of Fχ for all the considered boson mass range, meanwhile, the DM radius of lighter particles increases with
higher slopes in low fractions compare to more massive ones.
8 Shakeri & Karkevandi

Figure 4. Variation of RB (left) and RD (right) are indicated separately with respect to Fχ for MT = 1.4M . Different
boson masses are considered as labeled for λ = π. For each mχ , the solid and dashed curves depict the DM core and DM halo
formations, respectively. The black dashed line in the left panel shows the lower observational constraint for R1.4 .

Figure 5. Variation of RB (left) and RD (right) are indicated separately with respect to Fχ for MT = 1.4M . Different
coupling constants are considered as labeled for mχ = 200 MeV. For each λ, the solid and dashed curves depict the DM core
and DM halo formations, respectively. The black dashed line in the left panel shows the lower observational constraint for R1.4 .

We also investigate the effect of self-coupling constant on the visible and dark radius of DM admixed NSs with
different amount of DM (Fig. 5) which the mass of DM particles is fixed at mχ = 200 MeV while λ is changed between
0.5π and 4π. It is seen that for both DM core/halo formations, RB decreases more towards lower λ going faster below
11 km limit which is disfavoured by the latest observations. In Fig. 5 (left panel) one might denote an upper bound
for Fχ in each cases, where below this value the radius constraint is satisfied, note that for some coupling constants
(e.g. λ = 4π) the curve is always above 11 km. We find that for mχ = 200 MeV as a moderate boson mass, λ ' 2.9π is
the lowest value of self-coupling constant for which the condition RB & 11 km is fulfilled for DM fractions up to 20%.
The DM radius as it is illustrated in the right panel of Fig. 5 is increasing with Fχ and higher values of λ corresponds
to sharper rise in RD . From both left and right panels, it turns out that higher coupling constants lead to the DM
halo formation at lower Fχ , however, for smaller λ, DM resides as a core for a greater range of DM fractions.
In this section, we show that for boson masses and DM fractions in range mχ ∈ [100, 300] MeV and Fχ ∈ (0, 20]%,
respectively, the outermost radius of the mixed object changes from visible radius (RB ) to dark radius (RD ) by
increasing Fχ for which the corresponding fraction point, where RB ≈ RD , will be increased for heavier bosons
and/or lower coupling constants. Moreover, it is seen that generally for both DM core and DM halo formations
and Fχ . 20%, the visible radius (RB ) of the DM admixed NS will be decreased by increasing Fχ while RD is a
rising function. Therefore, the lower limit of R1.4 (11 km), which is inferred from various observational measurements
Dietrich et al. (2020); Huth et al. (2022), is disfavoured for heavier bosons and/or smaller coupling constants. In fact,
we conclude that smaller mχ and larger λ are in favor of radius constraints for NSs with M = 1.4M . It is worth
saying that the reduction of the visible radius for DM particles with mχ = 300 MeV which constitute 20% of a DM
9

admixed NS is about 14.4% of the corresponding radius of a pure NS. This decline for the same amount of DM but
with mχ = 100 MeV is only 0.05% of RB in a pure case which shows contrary to low mass particles, the presence of
massive bosons in NS significantly reduces the radius of the composed object. The drop of RB for a mixed compact
object composed of 20% DM component with mχ = 200 MeV and λ = 0.5π is 11.18% of the radius of a normal
NS, this value for the same condition but for λ = 4π is only 2.22%. Note that the current NICER uncertainty in
determination of the radius of a NS is ∼ 10% Rutherford et al. (2022). The next generation of X-ray telescopes such
as STROBE-X are capable to reach 5% uncertainty level (and even incredible 2% uncertainty for longer observations)
in measurements of mass and radius of the compact objects Ray et al. (2019) and hence they will be served as unique
opportunities to probe the presence of DM within the NSs. The above mentioned examinations further demonstrate
the effect of the mass and self-coupling constant of the DM particles on the variation of the visible radius at a constant
fraction.
It is instructive to compare our results with those obtained in a recent study Rutherford et al. (2022) where the
authors tried to explore a different bosonic asymmetric DM model and they attempted to constrain the amount of
DM within the compact objects. The DM model parameters space which considered in that study were Fχ ∈ [0, 20]%,
mχ ∈ [10−2 , 108 ] MeV with the repulsive self-interaction gχ /(mφ /MeV) ∈ [10−2 , 103 ], where mφ is the mass of the
mediator vector bosons between DM particles. The role of DM particles only as a core component was considered in
Rutherford et al. (2022) showing that regardless of the BM EoS, low DM factions in the range of Fχ ∈ [0, 4]% (as the
most constrained parameter) were favored by the latest NICER mass-radius measurements. It was concluded that while
the mass and the self-coupling constant of DM particles as separated quantities are unconstrained, one can constrain
the ratio of these parameters. Similar to us it was shown that the heavier bosons with low self-interaction strength
have been disfavoured in the light of the latest observational data. It has been also argued that the presence of the DM
cores in the mixed object relaxes the constraints on BM EoSs for interpreting the latest mass-radius measurements to
be consistent with them.

5. PULSE PROFILE AND LIGHT BENDING DUE TO THE DM HALO


X-ray telescopes measure the radius of a NS by tracking the X-ray emission from hot spots on the surface as the star
rotates. The trajectory of photons are affected by the gravitational field of compact objects leads to the gravitational
light bending which is taken into account in the analysis of NICER via Pulse Profile Modeling (PPM) (Beloborodov
2002; Poutanen & Beloborodov 2006; Turolla & Nobili 2013; Ozel et al. 2016; Gendreau et al. 2012). In this section, we
consider the effect of bosonic DM halo on the self-light bending from DM admixed NSs. The thermal surface photons
usually originate from the baryonic radius at RB , while it is possible to have some non-thermal emission produces at
higher latitude r > RB Lorimer & Kramer (2004). Due to the lack of interaction between DM particles and photons,
the DM halo up to radius RD remains invisible in the light of EM interactions but it affects the light propagation
through the gravitational interaction. We assume a spherically symmetric non-rotating space-time (Schwarzschild
metric) outside RB as

ds2 = −g(r)dt2 + f (r)dr2 + r2 dθ2 + r2 sin2 θdφ2 , (12)

where g(r) = f (r)−1 = 1 − 2M (r)/r, the total mass M (r) similar to Eq. (11) has two contributions coming from both
BM and DM fluids at r ≥ RB . In a NS without DM halo M (r ≥ RB ) is a constant value, the similar situation happens
for a DM admixed NS with DM core formation. However, the formation of the DM halo around NS can potentially
change the geometry outside the surface of NS and also the light propagation characteristics in this region. In this
section in all considered cases we fix the total mass of DM admixed NSs at MT (RD ) = 1.4M and we focus only on
those region of DM model parameter space in which we have the DM halo formation.
As it was discussed in preceding sections, the bosonic DM is confined inside a BM shell for heavy DM particles
with low DM fractions while the light DM particles usually form halo and distributed up to large radius out of the
NS. The presence of DM inside or around the NS would change the compactness and the gravitational potential of
the object, consequently, it contributes to the deflection of the light originating from the visible surface of the star.
The compactness of DM admixed NSs is displayed in Fig. 6 as a function of boson masses, self-coupling constants
and for different DM fractions as labeled. Here we consider the total compactness of the mixed star up to the visible
and dark radius as C(RB ) = MT (RB )/RB and C(RD ) = MT (RD )/RD , respectively. We see in Fig. 6 that the total
compactness at RD and RB increases with the boson masses, while the increasing rate of the BM core compactness
is larger for higher DM fractions towards more massive bosons. Along with each curves for a constant Fχ towards
10 Shakeri & Karkevandi

massive bosons, the radius of DM halo and BM core decrease and MT (RB ) increases which lead to forming a more
compact object. We also examine the impact of different values of self-coupling constants (i.e. λ = 0.5π, π and 2π)
on the compactness of the DM admixed NS in Fig. 6. Assuming a constant DM fraction for a given boson mass, it is
seen that for larger values of λ the total compactness decreases, this effect is more evident for larger mχ . Note that
the maximum compactness for various self-coupling constant is equal for each fraction which happens in RB ≈ RD at
the end point of each curves where C(RB ) ≈ C(RD ). In both left and right panels of Fig. 6 for light bosons mχ . 200
MeV, increasing the DM fraction and self-coupling constant lead to a decrease in the compactness.
In order to compute the pulse profile of a DM admixed NS, we need to determine the metric function g(r) outside
the surface of NS taking into account the DM halo contribution which is shown in Fig. 7. The variation of g(r) in
terms of different DM fraction and boson masses is illustrated as a function of radial distance starting from RB . We
see that all curves started from distinct points as a result of different values of MT (RB )/RB , but for larger radius all
the way to RD , the lines merge to the pure BM case with M = 1.4M . Note that for an observer outside the DM
halo radius, the mixed object is similar to a star with total mass 1.4M .

Figure 6. Compactness up to the visible radius C(RB ) = MT (RB )/RB (left), and till the dark radius C(RD ) = MT (RD )/RD
(right), as a function of boson masses for various DM fractions and self-coupling constants. In all of the cases the total mass of
the objects are MT (RD ) = 1.4M while Fχ is varied from 10% to 30% as labeled and coupling constants λ = 0.5π, π and 2π
are shown by solid, dashed and dotted lines, respectively. All of the diagrams are plotted in the presence of DM halo until the
mχ for which RD ≈ RB indicating a marginal radius to form DM core.

Figure 7. Variation of metric function g(r) = 1 − 2M (r)/r with respect to radial distance for various DM fractions and
mχ = 100 MeV (left) and different boson masses as labeled and Fχ = 30% (right). In both of the plots coupling constant is
fixed at λ = π and the pure BM is shown by dashed black lines. For all the considered DM admixed NSs the variation of g(r)
is plotted from RB which is the visible radius of the object till RD where the halo is ended and MT (RD ) = 1.4M for all of
them. Note that the metric function for pure NS is calculated from RB ≈ 11.37 km up to 112 km which is the radius of the
biggest DM halo in our cases, for a NS with mass equal to 1.4M .
11

We consider the trajectory of photons emitting from a small spot (dS 0 ) on the star surface (BM core) where the
emission angle α is evaluated with respect to the normal vector of the emission surface. The star is spinning and the
visible spot changes its position periodically and light rays would be bent by the gravity and reach to an observer at
far distance D appearing as observed pulse profiles.
The bolometric observed flux of a visible spot is given by Beloborodov (2002); Poutanen & Beloborodov (2006):

d cos α dS 0
F = g(RB )δ 5 I 0 (α0 ) cos α , (13)
d cos Ψ D2
here δ is the Doppler factor and is appearing due to the rotation of the star, in the case of a slowly rotating star
δ = 1. The local intensity of radiation I 0 (α0 ) depends on the applied emission model, in the following without loss of
generality, we normalize pulse profile to value F0 = I 0 (α0 )dS 0 /D2 . The bending angle Ψ which is defined as an angle
between the line of sight (LOS) and the local radial direction of a visible spot on RB is as follows

cos Ψ = cos i cos θ + sin i sin θ cos φ , (14)

where i measures the angle between the rotational axis of the star and LOS, θ is the colatitude of the visible spot and φ
is the rotational phase of the pulsar. The bending angle can also be extracted through general relativity computations
using the null geodesic equation taking into account modified metric function g(r) in the presence of DM halo outside
RB as Misner et al. (1973)
Z ∞ h
dr 1 g(r) i−1/2
Ψ= 2 b2
− , (15)
RB r r2

where b is the impact parameter which is known as the distance between the LOS and the point that the photon hits
the observer sky at infinity and is connected to α as

p
b = RB sin α/ g(RB ) , (16)

The maximum value of Ψ is obtained at α = π/2 and the visibility condition of a spot is defined by Ψ < Ψmax (α =
π/2). In order to compute the normalized pulse profile F/F0 , one need to determine Ψ and α and the relation between
them through Eqs. (15) and (16) for a given RB . Note that in Eq. (16) the metric function is evaluated at baryonic
surface where the radiation originated while in Eq. (15) we use g(r) in which the contribution of DM halo is taken
into account (see Fig. 7 ). In order to express the pulse profile in terms of the pulsar phase φ, we have to construct a
correspondence between the pulsar phase φ and the bending angle Ψ via Eq. (14) on one hand and a combination of
Eqs. (15) and (16) on the other hand following the approach presented in (Poutanen & Beloborodov 2006).
The pulse profile of DM admixed NSs for different fractions of DM (left panel) and also different boson masses (right
panel) are displayed in Fig. 8. In these figures the total mass of the mixed object at RD is assumed to be 1.4M , the
angular position of the visible spot is defined by i = θ = π/4 and the rotation frequency of the star is 400Hz. The fall
and grow of the flux (even in pure BM shown by dashed line) as a function of the rotational phase is due to changing
the position of the spot compare to a distant observer, the minimum flux corresponds to the far-side position. The
depth of the minimum crucially depends on the compactness of the object which affect the gravitational light-bending
and alters the invisible surface of the star, the less compact object gives more deeper minimum. In the left panel of Fig.
8, we consider bosonic SIDM with mass mχ = 100 MeV and self-coupling constant λ = π which distributes as a halo
around the star, by increasing the amount of DM the minimum fluxes of the profiles decrease. As it was mentioned
earlier, the compactness of the object encodes in the minimum depth, it is seen that the higher DM fraction leads to
formation of less compact objects (see Fig. 6). The deviation of the minimum amplitude of the highest considered
BM BM 30% BM
fraction Fχ = 30% from pure NS (i.e. ∆F/Fmin = (Fmin − Fmin )/Fmin ) is about 25.7% which is notable effect of
the presence of DM halo around NS. It is worth mentioning that the difference between the total compactness at
BM
RB for DM admixed NS and pure NS (∆C(RB )) over C(RB ) following approximately similar ratio which is about
BM
∆C(RB )/C(RB ) ≈ 28.7% for DM fraction 30% (see Table 1). These values for mχ = 100 MeV and Fχ = 20% and
BM BM
10% are (∆F/Fmin , ∆C(RB )/C(RB )) ≈ (15.3%, 18.2%) and (7.6%, 7.7%), respectively. It is remarkable to see that
there is an approximate linear correlation between the deviation in the peak fluxes and the compactness at RB for
different Fχ . Such a relation may not be seen for more massive bosons (e.g. mχ = 250, 290 MeV) due to relatively
12 Shakeri & Karkevandi

Figure 8. The bolometric pulse profile of a visible spot on the BM surface of a DM admixed NS as a function of the observed
rotational phase. The gravitational light bending has taken into account for a rotating DM admixed NS (ν = 400Hz) with
total mass MT (RD ) = 1.4M . We take the inclination angle of the spin axis with respect to the line of sight i = π/4 and
the colatitude of the visible spot θ = π/4. (Left) the DM fraction of bosonic SIDM particle with mχ = 100 MeV and λ = π
inside the mixed object is changing from a pure BM, 10%, 20% and 30% as labeled. (Right) the mass of bosonic SIDM particle
is changing from 100 MeV to 290 MeV at fixed DM fraction Fχ = 30% and λ = π . As it shows the pulse profile is crucially
influenced by the size of the DM halo which depends on DM model parameters (e.g. mχ and λ) and DM fraction Fχ . The pulse
profile of the pure NS is shown by black dashed lines.

small RB as a result of experiencing a significant DM core formation along with higher DM fractions before forming
the halo (right panel of Fig. 8).
The dependency of the pulse profile to boson masses is examined in the right panel of Fig. 8, where the DM fraction
and the self-coupling constant are fixed at Fχ = 30% and λ = π, respectively. We expect to have more compact objects
at baryonic radius for larger mχ which makes the spot more visible during star’s rotation and rises the minimum flux
for massive bosons. Since the compactness C(RB ) associated with boson masses mχ = 250 and 290 MeV are higher
than the compactness of a pure NS, the minimum luminosities are observed at upper values compare to pure NS,
BM
∆F/Fmin = 2.1% and 19.6%, respectively (Fig. 8 right panel). On the other hand, the BM core of mixed objects
composed of bosonic SIDM particles with masses 100, 150 and 200 MeV have lower compactness and hence lower
minimum fluxes of the pulse profiles are seen. To clarify this point, it is instructive to compare the compactness of
the BM core from Fig. 6 and Table 1. It is worth mentioning that for mχ = 250 MeV with 30% fraction of DM, we
have about the same pulse profile to pure NS owing to the fact that DM is distributed in a radius relatively close to
the one of normal NS.

mχ (MeV) Fχ MT (RB )(M ) RB (km) RD (km) C(RB ) C(RD )


100 10% 1.282 11.331 57.201 0.167 0.036
100 20% 1.139 11.365 87.515 0.148 0.023
100 30% 0.995 11.363 112.133 0.129 0.018
150 30% 1.056 11.061 45.121 0.140 0.045
200 30% 1.194 10.387 22.177 0.169 0.093
250 30% 1.347 9.522 12.582 0.208 0.164
290 30% 1.399 8.962 9.012 0.23 0.229

Table 1. Properties of DM admixed NSs with MT (RD ) = 1.4M for SIDM bosons at λ = π and for different masses and DM
fractions as listed. For the pure BM with total mass 1.4M and radius 11.37 km, the compactness is C = M/R = 0.181.

Emitted photons depending on their initial positions on NS surface will pass trajectories with different lengths. One
can define the difference of photon arrival times to an observer sky at infinity with respect to a specific reference point,
which is selected to be aligned with LOS (Poutanen & Beloborodov 2006). In fact, the time delay of photons is scaled
with the travel time of photons emitted from the closest region to the observer as ∆t = t(b) − t(0), since the integral
of photon geodesic equation is diverged for vanishing impact parameter. The scaled time delay is given by
13


b2 g(r) −1/2
Z
−1 dr h i
∆t(b) = c 1− − 1 , (17)
RB g(r) r2

the above time delay should also be taken into account in the photon arrival phase to the observer φobs = φ+2πν∆t[b(φ)]
(Poutanen & Beloborodov 2006), note that for a slowly rotating star φobs ≈ φ and the phase shift is negligible. The
time delay (17) scaled with RB /c is illustrates in Fig. (9) for mχ = 100 MeV, λ = π and different Fχ as labeled.
Due to the increasing of the compactness C(RB ) for lower fractions, the higher gravitational light bending enhances
the time delay. In fact the baryonic core with larger compactness produces higher values of gravitational potential
around the visible surface and causes more deflection of light and therefore the time dilation will be increased. As it
is shown the time delay in our case has almost negligible effect which is appeared as a very small phase shift in the
corresponding pulse profile of the DM admixed NS compare to pure NS.

Figure 9. Time delay of surface photons as a function of bending angle for different fraction of DM. Time delay is scaled
by RB /c which is the time needed for photons to propagate radially from center of the star to the baryonic surface. The
corresponding time delay for the pure NS is shown by black dashed lines.

In this section, we showed that the presence of DM halo will change the geometry around NS and therefore it
modifies the light trajectory and light bending effect in the vicinity of DM admixed NSs. Our results show that the
mixed objects should be considered as a possibility in the numerical simulation codes based on ray tracing methods in
order to interpret observations of X-ray telescopes from compact objects. Therefore by taking into account DM halo
around compact objects, the observational constraints may change and this will also affect the interpretation of those
compact objects and the determination of their mass and radius. It is seen that the compactness of the mixed object,
which depends on DM model parameters and fractions, plays an important role to impact the observed pulse profile
and the minimum flux during star rotation. Recently, the effect of DM halo on the pulse profile of emitted photons
from DM admixed NSs has been investigated in Miao et al. (2022) where the total mass up to the baryonic radius RB
is assumed to be MT (RB ) = 1.4M and obviously the total gravitational mass of the object (i.e. up to RD ) is more
than 1.4M . In that study, the halo mass (Mhalo ) was defined as the mass which occupies in range RB ≤ r ≤ RD and
deviation of the peak flux compare to pure NS depends on Mhalo /RD with an approximate linear relation. In Miao
et al. (2022) contrary to our study, C(RB ) is the same for all considered cases and instead the compactness of DM halo
is changed, it has been reported that for compact halo with Mhalo /RD ≈ 0.01 the peak flux is less than pure BM by
around 10%. In our case we use a different definition for DM halo mass including all DM content and starting from
r = 0 all the way to RD , besides both C(RB ) and C(RD ) change with mχ , λ and Fχ while the total gravitational mass
of the DM admixed NS (M (RD )) is fixed at 1.4M for all the considered case which is more reasonable assumption
for comparison. In fact in order to consider the validity of our scenario, there could be other observational methods
to measure the total gravitational mass of the object rather than the surface emission measured by X-ray telescopes.
In general our results are in good agreement with those presented in Miao et al. (2022), while our applied methods
are different. Meanwhile, it is not clear for us why the effect of DM halo in their study is more significant at the peak
14 Shakeri & Karkevandi

of the profile while we would expect that the modification in the observed luminosity due to the light bending is more
evident at minimums (rather than maximums) where the spot is at the far-side position. Moreover their modeling of
pulse profile from a two point-like spot is ambitious, according to (Beloborodov 2002) by assuming i = θ = π/4 the
primary spot is visible all the time while the antipodal one is never seen and therefore one spot model should give the
same results.

6. JOINT CONSTRAINTS FROM MASS AND RADIUS OBSERVATIONS


In this section, regarding two key observable properties of NSs, MTmax ≥ 2M and R1.4 ≥ 11 km, we attempt
to investigate the DM admixed NS parameter space by the visible (BM) radius and the total maximum mass of the
object. A precise scan has been done over bosonic DM mass and fraction in given ranges mχ ∈ (0, 1000] MeV and
Fχ ∈ (0, 10]% for a fixed self-interaction strength λ = π. In fact the distribution of DM in or around a NS would affect
its astrophysical properties, the DM core formation can significantly decrease both the mass and visible radius which
has the potential to violate obtained constraints for NSs. However, formation of a DM halo around the NS will increase
its total gravitational mass and slightly decline the visible radius, this may favour mass-radius constraints specially
when an applied BM EoS confirms aforementioned observational limits. The discussion presented in this section is
complementary to the one in Sec. VI of Karkevandi et al. (2022) where the tidal deformability and the total maximum
mass constraints were considered to probe the DM model parameter space and its fraction, which consequently led to
an allowed parameter space limited to low DM fractions for sub-GeV bosons at λ = π.

Figure 10. The fraction of DM as a function of its particle mass for λ = π. The blue dashed curve represents the maximum
total gravitational mass to be equal 2M , the top right region of the dashed line is in disagreement with 2M constraint. The
black curve indicates the visible radius 11 km for a mixed compact object with total mass to be 1.4M and determines the
disallowed parameter space in which R1.4 < 11 km. The area below and on the left of the dashed line is consistent with both
the heaviest known NSs and the latest NICER and LIGO/Virgo constraints for R1.4 .

In Fig. 10, the variation of the total maximum mass (MTmax ) is indicated as a density plot where the blue dashed
line defines the mass boundary MTmax = 2M which splits Fχ − mχ parameter space into two regions with maximum
mass less than 2M (top right & blur blue) as the non-allowed part and higher than 2M (left bottom & blur yellow)
as the allowed region. Note that the color density represents the total maximum gravitational mass of DM admixed
NSs as labeled. In addition, the black solid line corresponds to the mixed compact objects with MT = 1.4M for
which their visible radius RB is equal to 11 km. In a similar fashion to the mass constraint, the black solid line divides
the parameter space into allowed and non-allowed regions in terms of the radius constraint, the top right region of
15

the plot belongs to objects with R1.4 < 11 km which is disfavoured in the light of the latest observations of X-ray
telescopes and GW detectors Huth et al. (2022); Dietrich et al. (2020). As it is seen both mass and radius constraints
are following the same trend which simultaneously are measured by NICER and support underlying reasons for our
consideration. Obviously, we see that the maximum mass of the DM admixed NS provides a more stringent constraint
compare to the R1.4 limit.
Moreover, there is a particle mass limit below which both constraints are satisfied for any amount of DM in NSs.
In this regard, the MTmax ≥ 2M limit is met for mχ ≤ 105 MeV, while for the visible radius constraint R1.4 ≥ 11
km, the validated mass range of DM particles is mχ ≤ 155. Consequently, the joint constraints are respected for
mχ ≤ 105 MeV considering all possible DM fractions. On the other hand, there is an upper amount of DM where both
MTmax = 2M and R1.4 = 11 km lines merge together and indicates 4% as a fraction limit for the whole sub-GeV
bosons. This fact demonstrates that every observed NS could in principle include low fractions of DM limited to 4%.
Furthermore, SIDM bosons with masses mχ & 200 MeV can form core inside NSs, which reduces the radius and mass
reaching to fraction boundary limits towards more massive particles. Meanwhile, for mχ ∈ [500, 1000] MeV the impact
of DM particles on the properties of DM admixed NS is saturated, this is originated from the behaviour of bosonic DM
EoS for massive particles. In this mass range, the mean reduction of the radius and mass with respect to pure NS at
Fχ = 10% is 8.72% and 13.84%, respectively. We see that light particles are in favour of mass and radius constraints,
however, as it was shown previously the tidal deformability significantly restricted available fractions in low boson
mass regime Karkevandi et al. (2022).
The results discussed above are related to fixed λ = π, for smaller values of coupling constants, the limiting lines of
both constraints (MTmax = 2M and R1.4 = 11 km) will be shifted to the left and moved downwards, which leads to
increasing the exclusion region. A scan over a range of coupling constants in addition to masses and fractions, will be
subject of a following paper in which combined multi-messenger data of compact objects will be used to tighten the
DM parameter space. In this direction, a recent study has been done, where a different bosonic DM model applied to
scan over mass-coupling parameter space for some fixed DM fractions Giangrandi et al. (2022).

7. CONCLUSION AND REMARKS


In this article, we have considered bosonic particles with self-repulsive interaction to model DM in NSs for a new
family of compact objects called DM admixed NSs. The distribution of DM as a core or halo in mixed compact objects
crucially depends on DM model parameters such as boson masses, self-coupling strength and also the amount of DM
in NSs. The equilibrium configurations of DM admixed NSs are considered to probe DM model parameter space in
the light of the latest NICER observations, where the total mass of the object is illustrated with respect to the visible
radius. Owing to the fact that DM core formation reduces both the total mass and visible radius, we have shown that
lighter bosons, larger coupling constants and lower DM fractions are in favor of the NICER measurements for PSR
J0030+0451 and PSR J0740+662.
Regarding the lower limit of radius ∼ 11 km for NSs with total mass 1.4M obtained from joint analysis of NICER
and LIGO/Virgo, we focused on the variation of the visible and dark radius and their dependency upon Fχ , mχ and
λ. It is turned out that by increasing DM fractions, RB is mainly a decreasing function for DM core/halo formation,
however, RD always rises. We found that the reduction of RB going below 11 km by the presence of bosonic DM with
larger mχ and/or smaller λ, which can be used to further constrain DM model. It was demonstrated that the variation
rate of the visible radius of a DM admixed NS is within the sensitivity range of the upcoming X-ray telescopes such as
STROBE-X Ray et al. (2019), ATHENA Barcons et al. (2012) and eXTP Watts et al. (2019) (see Sec. 4). Moreover,
a shift from DM core to DM halo could occur for all DM particles by increasing Fχ where the outermost radius of
the mixed compact object changes from RB to RD . The core-halo transition point crucially depends on the mass and
self-coupling constant of bosonic DM particles.
We introduced the pulse profile as a new observable in Sec. 5, to look for the evidences of the bosonic DM halo
around NSs. The presence of DM around NSs via changing the geometry of space-time outside the BM radius and also
the compactness of the object will impact the trajectory of surface photons. The effect of DM model parameters and
fraction on the pulse profile has been investigated comprehensively which can be served as an independent probe for
bosonic DM. It was shown that the deviation of the minimum fluxes due to mχ and Fχ in DM admixed NSs compare
to pure NSs is a remarkable signature of the DM halo. Our results might be also included into numerical methods for
PPM and ray tracing from NS surface in X-ray telescopes giving rise DM admixed NS as a new promising possibility
to interpret observations.
16 Shakeri & Karkevandi

Finally, we perform a precise scan over bosonic mass and fraction at ranges mχ ∈ (0, 1000] MeV and Fχ ∈ (0, 10]%
for a fixed self-interaction strength λ = π. In this regard, combined astrophysical constraints of NSs, MTmax ≥ 2M
and R1.4 ≥ 11 km, are taken into account, it was seen that the maximum mass limit gives wider restrictions on the DM
model parameter space compare to the radius constraint. We found that for sub-GeV bosonic DM particles, each NS
could contain DM up to 4% of the total mass without violating the astrophysical limits for mass and radius. Besides,
the presence of light bosons (i.e. mχ ≤ 105 MeV) in NS are consistent with both permitted values of MTmax and R1.4
at any DM fractions. Note that the allowed parameter space for low mass bosons would be more tighten to very low
fractions by considering the tidal deformability inferred from GW detections Karkevandi et al. (2022). In addition
to LIGO/Virgo/KAGRA results Abbott et al. (2021), this bound could be further constrain thanks to the upcoming
GW missions such as LISA Baker et al. (2019) and Einstein telescopes Maggiore et al. (2020). Note that the presence
of DM core/halo should be precisely taken into account in numerical-relativity simulations to compute GW spectrum
and waveforms during merger and postmerger phases of a binary system containing at least a DM admixed NS Ellis
et al. (2018a); Giudice et al. (2016); Bezares et al. (2019); Emma et al. (2022).
The study presented in this article can be extended to include numerous BM/DM EoSs and in order to find the
best case scenario, Bayesian analysis can also be done regarding the latest multi-messenger data. The effect of DM
particles may be viewed as an effective softening/stiffening of the EoS corresponding to the whole structure of DM
admixed NSs. This effect in principle represents similar behaviours to BM EoSs at high density such as those happen
in the hybrid stars or twin stars as a deconfinement phase transition at the core of NSs Christian & Schaffner-Bielich
(2021); Sen et al. (2022) or hyperon puzzle Del Popolo et al. (2020a). It is worth noting that DM admixed NSs can
provide alternative explanations for exotic compact objects which have been detected so far such as the secondary
component in the GW190814 event Abbott et al. (2020) with the total mass ∼ 2.6M Das et al. (2021); Lee et al.
(2021); Di Giovanni et al. (2021), this object can be described within our DM model by the presence of ∼ 20% bosonic
SIDM with mχ = 50 MeV and λ = π subject to a visible radius of about 10 km Karkevandi et al. (2021). The NICER
collaboration reported the radius of PSR J0740+6620 with a gravitational mass of 2.08 ± 0.07M to be 12.35 ± 0.75km
Miller et al. (2021) which is approximately similar to the corresponding radius for PSR J0030+0451 as 12.45 ± 65km
Miller et al. (2021) while it is around 1.5 times less massive (M ∼ 1.4 M ). This observation is debatable in terms of
NSs theoretical approaches and the corresponding mass-radius profiles Legred et al. (2021); Li et al. (2021); Christian
& Schaffner-Bielich (2021); Drischler et al. (2022) and could also be explained by the existence of DM admixed NS.
In addition to the pulse profile which is mentioned in this paper, other observational features can be applied to
explore the DM features in or around stars. For instance, gravitational Microlensing due to the dark halo surrounding
NS could cause measurable changes in the brightness of the lensed source which would give hints to discriminate them
from other dense objects, the presence of large halo may also affect Shapiro time delay. Our research can potentially
have a significant impact on the discovery of the DM admixed NSs with self-repulsive bosonic DM component. The
upcoming astrophysical instruments thanks to precise measurements of the compact object properties may shed light
on the nature of DM and the possibility of the existence of DM within NS.

ACKNOWLEDGMENTS
S.S would like to thank Remo Ruffini for supporting his visit at ICRANet Pescara as an adjunct professor where
the last part of this work was done. D.R.K. is really grateful for fruitful discussions during Dark Matter in Compact
Objects, Stars, and in Low Energy Experiments workshop at University of Washington (INT Program INT-22-2b) in
which some parts of this research were presented. S.S and D.R.K appreciate for valuable discussions and comments
during the 5th (virtual) workshop on Transient events and Multi-messenger Astrophysics organized by Iranian National
Observatory (INO) in 28-29 July 2022.
”The wound is the place where the Light enters you.” Rumi

REFERENCES
Abbott, B., Abbott, R., Abbott, T., et al. 2017, Physical Abbott, B. P., et al. 2018, Phys. Rev. Lett., 121, 161101,

Review Letters, 119, doi: 10.1103/physrevlett.119.161101 doi: 10.1103/PhysRevLett.121.161101


17

Abbott, R., et al. 2020, Astrophys. J. Lett., 896, L44, Brito, R., Cardoso, V., & Okawa, H. 2015, Phys. Rev.
doi: 10.3847/2041-8213/ab960f Lett., 115, 111301, doi: 10.1103/PhysRevLett.115.111301
—. 2021, Astrophys. J. Lett., 915, L5, Bustillo, J. C., Sanchis-Gual, N., Torres-Forné, A., et al.
doi: 10.3847/2041-8213/ac082e 2021, Phys. Rev. Lett., 126, 081101,
Agnihotri, P., Schaffner-Bielich, J., & Mishustin, I. N. 2009, doi: 10.1103/PhysRevLett.126.081101
Phys. Rev. D, 79, 084033, Capano, C. D., Tews, I., Brown, S. M., et al. 2020, Nature
doi: 10.1103/PhysRevD.79.084033 Astron., 4, 625, doi: 10.1038/s41550-020-1014-6
Al-Mamun, M., Steiner, A. W., Nättilä, J., et al. 2021, Chatterjee, S., Garani, R., Jain, R. K., et al. 2022.
Phys. Rev. Lett., 126, 061101, https://arxiv.org/abs/2205.05048
doi: 10.1103/PhysRevLett.126.061101 Chavanis, P.-H. 2018, Phys. Rev. D, 98, 023009,
Amaro-Seoane, P., Barranco, J., Bernal, A., & Rezzolla, L. doi: 10.1103/PhysRevD.98.023009
2010, JCAP, 11, 002, —. 2019, Eur. Phys. J. Plus, 134, 352,
doi: 10.1088/1475-7516/2010/11/002 doi: 10.1140/epjp/i2019-12734-7
Ángeles Pérez-Garcı́a, M., Grigorian, H., Albertus, C., —. 2020, Phys. Rev. D, 101, 063532,
Barba, D., & Silk, J. 2022, Phys. Lett. B, 827, 136937, doi: 10.1103/PhysRevD.101.063532
doi: 10.1016/j.physletb.2022.136937 Christian, J.-E., & Schaffner-Bielich, J. 2021.
Annala, E., Gorda, T., Kurkela, A., & Vuorinen, A. 2018, https://arxiv.org/abs/2109.04191
Phys. Rev. Lett., 120, 172703, Ciancarella, R., Pannarale, F., Addazi, A., & Marciano, A.
2021, Phys. Dark Univ., 32, 100796,
doi: 10.1103/PhysRevLett.120.172703
doi: 10.1016/j.dark.2021.100796
Antoniadis, J., et al. 2013, Science, 340, 6131,
Ciarcelluti, P., & Sandin, F. 2011, Phys. Lett. B, 695, 19,
doi: 10.1126/science.1233232
doi: 10.1016/j.physletb.2010.11.021
Arbey, A., Lesgourgues, J., & Salati, P. 2003, Phys. Rev. D,
Colpi, M., Shapiro, S., & Wasserman, I. 1986, Phys. Rev.
68, 023511, doi: 10.1103/PhysRevD.68.023511
Lett., 57, 2485, doi: 10.1103/PhysRevLett.57.2485
Baker, J., et al. 2019. https://arxiv.org/abs/1907.06482
Coughlin, M. W., Dietrich, T., Margalit, B., & Metzger,
Barcons, X., Barret, D., Decourchelle, A., et al. 2012, arXiv
B. D. 2019, Mon. Not. Roy. Astron. Soc., 489, L91,
e-prints, arXiv:1207.2745.
doi: 10.1093/mnrasl/slz133
https://arxiv.org/abs/1207.2745
Cromartie, H. T., et al. 2019, Nature Astron., 4, 72,
Baryakhtar, M., Bramante, J., Li, S. W., Linden, T., &
doi: 10.1038/s41550-019-0880-2
Raj, N. 2017, Phys. Rev. Lett., 119, 131801,
Das, H. C., Kumar, A., & Patra, S. K. 2021, Phys. Rev. D,
doi: 10.1103/PhysRevLett.119.131801
104, 063028, doi: 10.1103/PhysRevD.104.063028
Baryakhtar, M., et al. 2022, in 2022 Snowmass Summer
De, S., Finstad, D., Lattimer, J. M., et al. 2018, Phys. Rev.
Study. https://arxiv.org/abs/2203.07984
Lett., 121, 091102, doi: 10.1103/PhysRevLett.121.091102
Bell, N. F., Busoni, G., Motta, T. F., et al. 2021, Phys. Rev. de Lavallaz, A., & Fairbairn, M. 2010, Phys. Rev. D, 81,
Lett., 127, 111803, doi: 10.1103/PhysRevLett.127.111803 123521, doi: 10.1103/PhysRevD.81.123521
Beloborodov, A. M. 2002, Astrophys. J. Lett., 566, L85, Del Popolo, A., Deliyergiyev, M., & Le Delliou, M. 2020a,
doi: 10.1086/339511 Phys. Dark Univ., 30, 100622,
Berezhiani, Z., Biondi, R., Mannarelli, M., & Tonelli, F. doi: 10.1016/j.dark.2020.100622
2021, Eur. Phys. J. C, 81, 1036, Del Popolo, A., Deliyergiyev, M., Le Delliou, M., Tolos, L.,
doi: 10.1140/epjc/s10052-021-09806-1 & Burgio, F. 2020b, Phys. Dark Univ., 28, 100484,
Bezares, M., Viganò, D., & Palenzuela, C. 2019, Phys. Rev. doi: 10.1016/j.dark.2020.100484
D, 100, 044049, doi: 10.1103/PhysRevD.100.044049 Del Popolo, A., Le Delliou, M., & Deliyergiyev, M. 2020c,
Bogdanov, S., et al. 2019, Astrophys. J. Lett., 887, L25, Universe, 6, doi: 10.3390/universe6120222
doi: 10.3847/2041-8213/ab53eb Deliyergiyev, M., Del Popolo, A., Tolos, L., et al. 2019,
Breschi, M., Perego, A., Bernuzzi, S., et al. 2021, Mon. Not. Physical Review D, 99, doi: 10.1103/physrevd.99.063015
Roy. Astron. Soc., 505, 1661, Dengler, Y., Schaffner-Bielich, J., & Tolos, L. 2021.
doi: 10.1093/mnras/stab1287 https://arxiv.org/abs/2111.06197
Brito, R., Cardoso, V., Macedo, C. F. B., Okawa, H., & Di Giovanni, F., Fakhry, S., Sanchis-Gual, N., Degollado,
Palenzuela, C. 2016, Phys. Rev. D, 93, 044045, J. C., & Font, J. A. 2020, Phys. Rev. D, 102, 084063,
doi: 10.1103/PhysRevD.93.044045 doi: 10.1103/PhysRevD.102.084063
18 Shakeri & Karkevandi

Di Giovanni, F., Sanchis-Gual, N., Cerdá-Durán, P., & Henriques, A. B., Liddle, A. R., & Moorhouse, R. G. 1990b,
Font, J. A. 2021. https://arxiv.org/abs/2110.11997 Nucl. Phys. B, 337, 737,
Dietrich, T., Coughlin, M. W., Pang, P. T. H., et al. 2020, doi: 10.1016/0550-3213(90)90514-E
Science, 370, 1450, doi: 10.1126/science.abb4317 Huth, S., et al. 2022, Nature, 606, 276,
Drischler, C., Han, S., & Reddy, S. 2022, Phys. Rev. C, doi: 10.1038/s41586-022-04750-w
105, 035808, doi: 10.1103/PhysRevC.105.035808 Ivanytskyi, O., Sagun, V., & Lopes, I. 2020, Phys. Rev. D,
Eby, J., Kouvaris, C., Nielsen, N. G., & Wijewardhana, L. 102, 063028, doi: 10.1103/PhysRevD.102.063028
2016, JHEP, 02, 028, doi: 10.1007/JHEP02(2016)028 Jetzer, P. 1992, PhR, 220, 163,
doi: 10.1016/0370-1573(92)90123-H
Ellis, J., Hektor, A., Hütsi, G., et al. 2018a, Phys. Lett. B,
Kalashnikov, I. Y., & Chechetkin, V. M. 2022, Mon. Not.
781, 607, doi: 10.1016/j.physletb.2018.04.048
Roy. Astron. Soc., 514, 1351,
Ellis, J., Hütsi, G., Kannike, K., et al. 2018b, Phys. Rev. D,
doi: 10.1093/mnras/stac1319
97, 123007, doi: 10.1103/PhysRevD.97.123007
Karkevandi, D. R., Shakeri, S., Sagun, V., & Ivanytskyi, O.
Emma, M., Schianchi, F., Pannarale, F., Sagun, V., &
2021, in 16th Marcel Grossmann Meeting on Recent
Dietrich, T. 2022, Particles, 5, 273,
Developments in Theoretical and Experimental General
doi: 10.3390/particles5030024
Relativity, Astrophysics and Relativistic Field Theories.
Essick, R., Landry, P., & Holz, D. E. 2020, Phys. Rev. D, https://arxiv.org/abs/2112.14231
101, 063007, doi: 10.1103/PhysRevD.101.063007 Karkevandi, D. R., Shakeri, S., Sagun, V., & Ivanytskyi, O.
Fattoyev, F. J., Piekarewicz, J., & Horowitz, C. J. 2018, 2022, Phys. Rev. D, 105, 023001,
Phys. Rev. Lett., 120, 172702, doi: 10.1103/PhysRevD.105.023001
doi: 10.1103/PhysRevLett.120.172702 Kaup, D. J. 1968, Phys. Rev., 172, 1331,
Freese, K., Ilie, C., Spolyar, D., Valluri, M., & doi: 10.1103/PhysRev.172.1331
Bodenheimer, P. 2010, ApJ, 716, 1397, Kouvaris, C. 2008, Phys. Rev. D, 77, 023006,
doi: 10.1088/0004-637X/716/2/1397 doi: 10.1103/PhysRevD.77.023006
Gamba, R., Breschi, M., Bernuzzi, S., Agathos, M., & Kouvaris, C., & Nielsen, N. G. 2015, Phys. Rev. D, 92,
Nagar, A. 2021, Phys. Rev. D, 103, 124015, 063526, doi: 10.1103/PhysRevD.92.063526
doi: 10.1103/PhysRevD.103.124015 Kouvaris, C., & Tinyakov, P. 2011a, Phys. Rev. D, 83,
Gendreau, K. C., Arzoumanian, Z., & Okajima, T. 2012, in 083512, doi: 10.1103/PhysRevD.83.083512
Society of Photo-Optical Instrumentation Engineers —. 2011b, Phys. Rev. Lett., 107, 091301,
(SPIE) Conference Series, Vol. 8443, Space Telescopes doi: 10.1103/PhysRevLett.107.091301
and Instrumentation 2012: Ultraviolet to Gamma Ray, Lee, B. K. K., Chu, M.-c., & Lin, L.-M. 2021, Astrophys.
ed. T. Takahashi, S. S. Murray, & J.-W. A. den Herder, J., 922, 242, doi: 10.3847/1538-4357/ac2735
844313, doi: 10.1117/12.926396 Legred, I., Chatziioannou, K., Essick, R., Han, S., &
Giangrandi, E., Sagun, V., Ivanytskyi, O., Providência, C., Landry, P. 2021, Phys. Rev. D, 104, 063003,
doi: 10.1103/PhysRevD.104.063003
& Dietrich, T. 2022. https://arxiv.org/abs/2209.10905
Leung, S., Chu, M., & Lin, L. 2011, Phys. Rev. D, 84,
Giudice, G. F., McCullough, M., & Urbano, A. 2016,
107301, doi: 10.1103/PhysRevD.84.107301
JCAP, 10, 001, doi: 10.1088/1475-7516/2016/10/001
Li, J. J., Sedrakian, A., & Alford, M. 2021, Phys. Rev. D,
Goldman, I., Mohapatra, R., Nussinov, S., Rosenbaum, D.,
104, L121302, doi: 10.1103/PhysRevD.104.L121302
& Teplitz, V. 2013, Phys. Lett. B, 725, 200,
Li, X., Wang, F., & Cheng, K. 2012, Journal of Cosmology
doi: 10.1016/j.physletb.2013.07.017
and Astroparticle Physics, 2012, 031–031,
Gresham, M. I., & Zurek, K. M. 2019, Phys. Rev. D, 99,
doi: 10.1088/1475-7516/2012/10/031
083008, doi: 10.1103/PhysRevD.99.083008 Liebling, S. L., & Palenzuela, C. 2017, Living Rev. Rel., 20,
Guzman, F. S., & Rueda-Becerril, J. M. 2009, Phys. Rev. 5, doi: 10.12942/lrr-2012-6
D, 80, 084023, doi: 10.1103/PhysRevD.80.084023 Lorimer, D. R., & Kramer, M. 2004, Handbook of Pulsar
Henriques, A., Liddle, A. R., & Moorhouse, R. 1989, Astronomy, Vol. 4
Physics Letters B, 233, 99, Maggiore, M., et al. 2020, JCAP, 03, 050,
doi: https://doi.org/10.1016/0370-2693(89)90623-0 doi: 10.1088/1475-7516/2020/03/050
—. 1990a, Physics Letters B, 251, 511, Marsh, D. J. E. 2016, Phys. Rept., 643, 1,
doi: https://doi.org/10.1016/0370-2693(90)90789-9 doi: 10.1016/j.physrep.2016.06.005
19

Maselli, A., Pnigouras, P., Nielsen, N. G., Kouvaris, C., & Rezzolla, L., Most, E. R., & Weih, L. R. 2018, Astrophys.
Kokkotas, K. D. 2017, Phys. Rev. D, 96, 023005, J. Lett., 852, L25, doi: 10.3847/2041-8213/aaa401
doi: 10.1103/PhysRevD.96.023005 Riley, T. E., et al. 2019, Astrophys. J. Lett., 887, L21,
Miao, Z., Zhu, Y., Li, A., & Huang, F. 2022. doi: 10.3847/2041-8213/ab481c
https://arxiv.org/abs/2204.05560 —. 2021, Astrophys. J. Lett., 918, L27,
Mielke, E. W., & Schunck, F. E. 2000, Nucl. Phys. B, 564, doi: 10.3847/2041-8213/ac0a81
185, doi: 10.1016/S0550-3213(99)00492-7 Romani, R. W., Kandel, D., Filippenko, A. V., Brink,
Miller, M. C., et al. 2019, Astrophys. J. Lett., 887, L24, T. G., & Zheng, W. 2021, Astrophys. J. Lett., 908, L46,
doi: 10.3847/2041-8213/ab50c5 doi: 10.3847/2041-8213/abe2b4
—. 2021, Astrophys. J. Lett., 918, L28, —. 2022, Astrophys. J. Lett., 934, L18,
doi: 10.3847/2041-8213/ac089b doi: 10.3847/2041-8213/ac8007
Misner, C. W., Thorne, K. S., & Wheeler, J. A. 1973, Ruffini, R., & Bonazzola, S. 1969, Phys. Rev., 187, 1767,
Gravitation (San Francisco: W. H. Freeman) doi: 10.1103/PhysRev.187.1767
Most, E. R., Weih, L. R., Rezzolla, L., & Schaffner-Bielich, Rutherford, N., Raaijmakers, G., Prescod-Weinstein, C., &
J. 2018, Phys. Rev. Lett., 120, 261103, Watts, A. 2022. https://arxiv.org/abs/2208.03282
doi: 10.1103/PhysRevLett.120.261103 Sagun, V., Lopes, I., & Ivanytskyi, A. 2019a, Nucl. Phys.
Narain, G., Schaffner-Bielich, J., & Mishustin, I. N. 2006, A, 982, 883, doi: 10.1016/j.nuclphysa.2018.10.024
Phys. Rev. D, 74, 063003, Sagun, V., Panotopoulos, G., & Lopes, I. 2020, Phys. Rev.
doi: 10.1103/PhysRevD.74.063003 D, 101, 063025, doi: 10.1103/PhysRevD.101.063025
Nelson, A., Reddy, S., & Zhou, D. 2019, JCAP, 07, 012, Sagun, V. V., Lopes, I., & Ivanytskyi, A. I. 2019b,
doi: 10.1088/1475-7516/2019/07/012 Astrophys. J., 871, 157, doi: 10.3847/1538-4357/aaf805
Nyhan, J. E., & Kain, B. 2022, Phys. Rev. D, 105, 123016, Sandin, F., & Ciarcelluti, P. 2009, Astroparticle Physics,
doi: 10.1103/PhysRevD.105.123016 32, 278–284, doi: 10.1016/j.astropartphys.2009.09.005
Oppenheimer, J. R., & Volkoff, G. M. 1939, Phys. Rev., 55, Sen, D., Alam, N., & Chaudhuri, G. 2022, Phys. Rev. D,
374, doi: 10.1103/PhysRev.55.374 106, 083008, doi: 10.1103/PhysRevD.106.083008
Ozel, F., Psaltis, D., Arzoumanian, Z., Morsink, S., & Shakeri, S., & Hajkarim, F. 2022.
Baubock, M. 2016, Astrophys. J., 832, 92, https://arxiv.org/abs/2209.13572
doi: 10.3847/0004-637X/832/1/92 Spolyar, D., Freese, K., & Gondolo, P. 2008, Phys. Rev.
Pang, P. T. H., Tews, I., Coughlin, M. W., et al. 2021, Lett., 100, 051101, doi: 10.1103/PhysRevLett.100.051101
Astrophys. J., 922, 14, doi: 10.3847/1538-4357/ac19ab Tews, I., Margueron, J., & Reddy, S. 2018, Phys. Rev. C,
Pang, P. T. H., et al. 2022. 98, 045804, doi: 10.1103/PhysRevC.98.045804
https://arxiv.org/abs/2205.08513 Tolman, R. C. 1939, Phys. Rev., 55, 364,
Poutanen, J., & Beloborodov, A. M. 2006, Mon. Not. Roy. doi: 10.1103/PhysRev.55.364
Astron. Soc., 373, 836, Tolos, L., Schaffner-Bielich, J., & Dengler, Y. 2015, Phys.
doi: 10.1111/j.1365-2966.2006.11088.x Rev. D, 92, 123002, doi: 10.1103/PhysRevD.92.123002
Raaijmakers, G., et al. 2019, Astrophys. J. Lett., 887, L22, Turolla, R., & Nobili, L. 2013, Astrophys. J., 768, 147,
doi: 10.3847/2041-8213/ab451a doi: 10.1088/0004-637X/768/2/147
—. 2020, Astrophys. J. Lett., 893, L21, Valdez-Alvarado, S., Becerril, R., & Ureña López, L. A.
doi: 10.3847/2041-8213/ab822f 2020, Phys. Rev. D, 102, 064038,
Raaijmakers, G., Greif, S. K., Hebeler, K., et al. 2021, doi: 10.1103/PhysRevD.102.064038
Astrophys. J. Lett., 918, L29, Visinelli, L. 2021. https://arxiv.org/abs/2109.05481
doi: 10.3847/2041-8213/ac089a Visinelli, L., Baum, S., Redondo, J., Freese, K., & Wilczek,
Radice, D., & Dai, L. 2019, Eur. Phys. J. A, 55, 50, F. 2018, Phys. Lett., B777, 64,
doi: 10.1140/epja/i2019-12716-4 doi: 10.1016/j.physletb.2017.12.010
Raj, N., Tanedo, P., & Yu, H.-B. 2018, Phys. Rev. D, 97, Watts, A. L., Yu, W., Poutanen, J., et al. 2019, Science
043006, doi: 10.1103/PhysRevD.97.043006 China Physics, Mechanics, and Astronomy, 62, 29503,
Ray, P. S., et al. 2019. https://arxiv.org/abs/1903.03035 doi: 10.1007/s11433-017-9188-4
Reed, B. T., Fattoyev, F. J., Horowitz, C. J., & Wu, Y., Baum, S., Freese, K., Visinelli, L., & Yu, H.-B.
Piekarewicz, J. 2021, Phys. Rev. Lett., 126, 172503, 2022, Phys. Rev. D, 106, 043028,
doi: 10.1103/PhysRevLett.126.172503 doi: 10.1103/PhysRevD.106.043028
20 Shakeri & Karkevandi

Wystub, S., Dengler, Y., Christian, J.-E., & Xiang, Q.-F., Jiang, W.-Z., Zhang, D.-R., & Yang, R.-Y.
Schaffner-Bielich, J. 2021. 2014, Phys. Rev. C, 89, 025803,
https://arxiv.org/abs/2110.12972 doi: 10.1103/PhysRevC.89.025803

View publication stats

You might also like