Next Article in Journal
W 50 Morphology and the Dynamics of SS 433 Formation—The Origin of TeV Gammas from the Microquasar
Previous Article in Journal
Analytical Inverse QCD Coupling Constant Approach and Its Result for αs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Galactic Stellar Black Hole Binaries: Spin Effects on Jet Emissions of High-Energy Gamma-Rays

by
Dimitrios Rarras
1,†,
Theocharis Kosmas
1,*,†,
Theodora Papavasileiou
1,2,† and
Odysseas Kosmas
1,†,‡
1
Department of Physics, University of Ioannina, GR-45110 Ioannina, Greece
2
Department of Informatics, University of Western Macedonia, GR-52100 Kastoria, Greece
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Current address: Conigital Ltd., 51 Parkside, Coventry CV1 2HG, UK.
Particles 2024, 7(3), 792-804; https://doi.org/10.3390/particles7030046
Submission received: 10 July 2024 / Revised: 26 August 2024 / Accepted: 30 August 2024 / Published: 3 September 2024

Abstract

:
In the last few decades, galactic stellar black hole X-ray binary systems (BHXRBs) have aroused intense observational and theoretical research efforts specifically focusing on their multi-messenger emissions (radio waves, X-rays, γ -rays, neutrinos, etc.). In this work, we investigate jet emissions of high-energy neutrinos and gamma-rays created through several hadronic and leptonic processes taking place within the jets. We pay special attention to the effect of the black hole’s spin (Kerr black holes) on the differential fluxes of photons originating from synchrotron emission and inverse Compton scattering and specifically on their absorption due to the accretion disk’s black-body radiation. The black hole’s spin (dimensionless spin parameter a * ) enters into the calculations through the radius of the innermost circular orbit around the black hole, the R I S C O parameter, assumed to be the inner radius of the accretion disk, which determines its optical depth τ d i s k . In our results, the differential photon fluxes after the absorption effect are depicted as a function of the photon energy in the range 1 GeV E 10 3 GeV. It is worth noting that when the black holes’ spin ( α * ) increases, the differential photon flux becomes significantly lower.

1. Introduction

In the last few decades, one of the powerful techniques to test the Kerr nature of astrophysical black holes, based on the study of the electromagnetic reflection spectrum of accretion disks, is considered X-ray reflection spectroscopy (see, for example, Refs. [1,2,3,4,5,6,7,8,9]). As it is well known, the accretion disk of black holes emits thermal photons that can exhibit inverse Compton scattering off free electrons in the compact object’s corona, which is a hot and optically thin medium close to the stellar black hole object.
Stellar-mass black holes are formed either from the gravitational collapse of massive stars or from the merger of two neutron stars. For our purposes, stellar-mass black holes have masses ranging from M B H 3   M up to M B H 50   M , specifically similar to the mass of the donor star of the BHXRB system.
In this work, we deal with uncharged ( Q = 0 ) black holes which are classified in two categories, i.e., Schwarzschild black holes (non-rotating, J = 0 ) that were the first exact solution of Einstein’s field equations found by Karl Schwarzschild in 1916, and Kerr black holes (rotating, J 0 ) that were found by Roy Kerr (1963) (see Ref. [10]). We mention that charged ( Q 0 ) black holes are either Reissner–Nordström (non-rotating, J = 0 ) or Kerr–Newman (rotating, J 0 ) types (see Ref. [11] pp. 33–58). The Kerr–Newman metric of a charged spinning black hole is the most general BH solution, which was found by Ezra “Ted” Newman in 1965 (see Ref. [12]). It is worth mentioning that all astrophysical black holes (BHs) can be described by the Kerr metric with only two parameters: (i) the black hole mass ( M b h ) and (ii) the black hole angular momentum (J).
In the present work, we study jet emissions of radiation and neutrinos created through several hadronic and leptonic processes occurring inside the jets. We focus on the effect of the black hole’s spin (Kerr black holes) on the differential photon fluxes and specifically on the absorption effect which is mainly due to the accretion disk’s black-body radiation. We mention that the black hole’s spin (dimensionless spin parameter a * ) enters this type of calculation through the radius of the innermost circular orbit around the black hole, i.e., the R I S C O geometrical parameter, assumed to be the inner radius of the accretion disc. This parameter determines the optical depth, τ d i s k , of the accretion disc.
Recently, progress has been made in measuring the spins of stellar-mass BHs in the Milky Way Galaxy (see Ref. [13] and references therein). Independent determinations of the BHs’ spins are of fundamental importance for testing general relativity (GR) and for a deep understanding of space and time.
The rest of this paper is organized as follows. In Section 2, we make a brief description of the state of the art of Kerr black holes with the assumed metric. In Section 3, analytic calculations of the emitted photon differential flux carried out within the above model are presented and discussed. In addition, the absorption effects due to photon annihilation and e + , e pair creation are considered. Finally, in Section 4, the extracted conclusions from the aforementioned investigation are summarized and future work that is planned is briefly shortly discussed.

2. Background—Description of the Model

In this section, we discuss the radius of the innermost stable circular orbit in the Kerr geometry—the Kerr R I S C O —the main interactions that take place inside the jets, the transfer equation which describes the particles’ energy distribution inside the jets, the source function of gamma-rays, the intensity and integral flux of gamma-rays, and the absorption of photons.

2.1. Kerr R I S C O

In previous works (see Refs. [14,15,16]), the black hole of the binary system was supposed to be a Schwarzschild black hole, with its only property being that of the mass. However, this case is not realistic, as most of the black holes, besides possessing mass M, also have angular momentum J. For this reason, in this work, we consider the black hole to be a Kerr black hole, one that has mass and is rotating.
The line element of the Kerr geometry (see Ref. [11] pp. 49–52 and also Ref. [17]), describing the space–time around a Kerr black hole, is
d s 2 = c 2 1 2 μ r ρ 2 d t 2 + 4 μ α c r sin 2 θ ρ 2 d t d ϕ ρ 2 Δ d r 2 ρ 2 d θ 2 r 2 + α 2 + 2 μ r α 2 sin 2 θ ρ sin 2 θ d ϕ 2
where μ = r g = r S 2 = G M c 2 and α = J M c are constants. Furthermore,
ρ 2 = r 2 + α 2 cos 2 θ , Δ = r 2 2 μ r + α 2 .
The above expression for d s 2 is known as the Boyer–Lindquist form and that for ( t , r , θ , ϕ ) as Boyer–Lindquist coordinates. It is useful to define the dimensionless spin parameter  α * α / μ . The larger this parameter, the more aspherical the line element. The case of α * = 1 corresponds to an extreme Kerr black hole. Schwarzschild black holes are the special case for α * = 0 .
We are now going to derive the innermost stable circular orbit, R I S C O , around a Kerr black hole. To do so, we work on the equatorial plane; setting θ = π / 2 in the Kerr line element of Equation (1), we obtain
d s 2 = c 2 1 2 μ r d t 2 + 3 μ α c r d t d ϕ r 2 Δ d r 2 r 2 + α 2 + 2 μ α 2 r d ϕ 2
from which we can read the covariant metric components g μ ν and calculate the contravariant metric components g μ ν . We take a particle with unit rest mass such that p μ = d x μ d τ . Since the metric does not depend explicitly on t and ϕ , we obtain p t and p ϕ that are conserved:
p t = k c 2 , p ϕ = h
where k and h are constants that represent the energy and the angular momentum of the particle. We use the fact that g μ ν p μ p ν = c 2 . Since p θ = 0 and substituting Equation (3), after some elaboration, we obtain the energy equation:
1 2 d r d τ 2 + V e f f ( r ; h , k ) = 1 2 c 2 ( k 2 1 )
where we have identified the effective potential per unit mass as
V e f f ( r ; h , k ) = μ c 2 r + h 2 α 2 c 2 ( k 2 1 ) 2 r 2 μ ( h α c k ) 2 r 3
Circular motion occurs when d r d τ 2 = 0 , and to remain in a circular motion, the radial acceleration d 2 r d τ 2 must also vanish. In terms of the effective potential, we require
V e f f ( r ; h , k ) = 1 2 c 2 ( k 2 1 ) a n d d V e f f d r = 0
For a circular orbit to also be stable, we require that
d 2 V e f f d r 2 = 0
From this, we obtain an implicit equation for the coordinate radius of the innermost stable circular orbit, see Ref. [11] pp. 49–52 and also Ref. [17]:
r 2 6 μ r 3 α 2 ± 8 α μ r = 0
where the upper sign corresponds to an orbit of a co-rotating particle, with respect to the black hole’s spin, and the lower sign to an orbit of a counter-rotating particle. The solution to Equation (8) (see Ref. [18]), the R I S C O , is found to be
R I S C O = μ { 3 + Z 2 ( 3 Z 1 ) ( 3 + Z 1 + 2 Z 2 ) 1 / 2 }
where
Z 1 1 + 1 α * 2 1 / 3 1 + α * 1 / 3 + 1 α * 1 / 3 Z 2 3 α * 2 + Z 1 2 1 / 2
In the limit α * = 0 , we recover R I S C O = 6 r g , the Schwarzschild case. In the extreme Kerr limit α * = 1 , we find R I S C O = 9 r g for the counter-rotating orbit and R I S C O = r g for the co-rotating case. We conclude that R I S C O strongly depends on the spin (angular momentum) of the black hole and, respectively, whether the particle is counter- or co-rotating. This radius plays a crucial role when we calculate the absorption from the accretion disc as, in our model, it is the inner radius of the accretion disk, R i n .

2.2. Main Reactions Inside the Jets

For the production of gamma-rays and neutrinos, two chains of reactions take place inside the jet. Specifically, the relativistic protons of the jet interact mainly with the cold hadronic matter of the jet, as well as with photons created from internal and external emission sources (see Refs. [19,20,21]).
The first chain of reactions begins with the proton–proton, p p , collisions which are
p p p p + α π 0 + β π + π p p p n + π + + α π 0 + β π + π p p n n + 2 π + + α π 0 + β π + π
where α and β are multiplicities depending on the proton energy (see Ref. [22]).
The second chain of reactions begins with proton–photon, p γ scatterings, which are
p γ n π + p γ p π 0 p γ p π 0 π + π
Subsequently, the pion decay processes take place as follows. The neutral pion decays to two gamma-ray photons:
π 0 γ + γ
While the charged pions decay to muons and muon-neutrinos, the muons (antimuons) decay to electrons (positrons) and electron-neutrinos (electron-antineutrinos):
π + μ + ν μ e + ν ¯ μ ν e + ν μ π μ ν ¯ μ e ν μ ν ¯ e + ν ¯ μ

2.3. The Transfer Equation

The governing equation that describes the matter inside the jet, in its simplest form, is the steady-state transfer equation (see Refs. [23,24,25]):
N ( E , z ) b ( E , z ) E + t 1 N ( E , z ) = Q ( E , z )
where its solution N ( E , z ) is the energy distributionof the particles inside the jet. Q ( E , z ) is the particle source function giving the respective production rate. b ( E ) is the energy loss rate, which has the form b ( E ) = d E / d t = E · t l o s s 1 , with t l o s s 1 being the rate that the particles lose energy. The rate t 1 is given as the sum t 1 = t e s c 1 + t d e c 1 , with t e s c 1 = c / ( z m a x z 0 ) being the rate that the particles escape the acceleration zone and t d e c 1 their decay rate.
The solution of the transfer equation is (see Refs. [26,27,28]):
N ( E , z ) = 1 | b ( E ) | E E m a x Q ( E , z ) e τ ( E , E ) d E
where
τ ( E , E ) = E E t 1 | b ( E ) | d E
The transfer equation is to be solved numerically. For each particle in the chains of reactions, we have a different transfer equation. The particles that are produced from others have a dependence on them through their source function. So, the set of transfer equations becomes a set of coupled equations.
In recent research, a more general form of the transfer equation is being used [29,30], that is,
N ( E , t , z ) t + Γ ( t , z ) v ( t , z ) N ( E , t , z ) z + N ( E , t , z ) b ( E , t , z ) E + t 1 N ( E , t , z ) = Q ( E , t , z )
The first term in this equation represents the temporal evolution of the energy distribution, and the second term represents the propagation of particles along the jet, where Γ ( t , z ) is the Lorentz factor along the jet and v ( t , z ) its bulk velocity. Calculations including contributions from such terms of the transfer equation are going to be published in future works.

2.4. Gamma-Ray Source Function

In this subsection, we see the form of the source function of gamma-rays. For the source function of the other particles in the chains of reactions, one can see [31,32]. We have assumed that gamma-rays are mainly produced by π 0 decay. Besides this decay, gamma-rays can also be produced by the η meson decay as
η γ + γ
The gamma-ray spectra are calculated for photons of energy E γ = x E p through the expression (see Ref. [33])
F γ ( x , E p ) = B γ l n x x 1 x β γ 1 + k γ x β γ ( 1 x β γ ) 4 1 l n x 4 β γ x β γ 1 x β γ 4 k γ β γ x β γ ( 1 2 x β γ ) 1 + k γ x β γ ( 1 x β γ )
where B γ = 1.3 + 0.14 L + 0.011 L 2 , β γ = 1 / ( 0.008 L 2 + 0.11 L + 1.79 ) , and k γ = 1 / ( 0.014 L 2 + 0.049 L + 0.801 ) , where L = ln ( E p / 1 TeV). These results are consistent with proton energies in the range 0.1 TeV < E p < 10 5 TeV. Equation (19) includes the contribution of η decay.
In the energy range E γ 100 GeV, the gamma-ray source function is
Q γ ( E γ , z ) = c n ( z ) E γ / E p m a x 1 d x x N p E γ x , z F γ x , E γ x σ p p ( i n e l ) E γ x
The inelastic cross-section for the p p scattering is given by (see Refs. [33,34,35])
σ p p ( i n e l ) ( E p ) = ( 0.25 l 2 + 1.88 L + 34.3 ) 1 E t h E p 4 2 × 10 27 cm 2
where L = ln ( E p / 1000 ) , with E p in GeV and E t h = 1.2 GeV being the threshold energy for the production of a single neutral pion. For E γ < 100 GeV, the delta-function approximation is used.

2.5. Intensities and Integral Fluxes

If we integrate the source function of Equation (20) over the acceleration zone, we obtain the respective intensity (see Refs. [36,37]) as
I γ , o ( E γ ) = V Q γ ( E γ , z ) d 3 r = π tan 2 ξ z 0 z m a x Q γ ( E γ , z ) z 2 d z
where the second equality holds because of the conical geometry of the jet.
However, part of the photons emitted from the jet can be absorbed due to e + e production through the interaction:
γ + γ e + + e
This can take place if a photon emitted from the jet interacts with a lower-energy photon emitted from the accretion disk, the donor (companion) star, or the corona. This leads to the corresponding optical depths τ d i s k , τ d o n o r , and τ c o r , which depend on various parameters of the binary system.
So, their initial intensity I γ , 0 will be reduced to
I γ = I γ , o e τ a l l
where τ a l l is the sum of all the optical depths.
The corresponding photon integral flux emitted is calculated through
Φ γ = 100 E m a x I γ ( E γ ) d E γ 4 π d 2

2.6. Absorption from the Accretion Disk

In this subsection, we discuss the photon absorption due to the interaction described by Equation (23). The accretion disk consists of overheated matter and gas rotating around the system’s black hole. The mass accretion rate M ˙ a c c r is defined as the rate at which mass from the donor star accumulates in the equatorial region of the black hole in a binary system. For the accretion disk of the system, we use the Shakura–Sunyaev “standard” disk model (see Ref. [38]), the best known and studied theoretical model. In this model, the accretion disk’s inner radius is located at R i n = R I S C O .
Due to the presence of the magnetic field and high temperature, the disk emits soft X-ray photons capable of absorbing the gamma-ray photons when they collide with an angle θ 0 . The cross-section of this interaction is (see Ref. [39])
σ γ γ ( E γ , ϵ , θ 0 ) = π e 4 2 m e 2 c 4 ( 1 β 2 ) ( 3 β 4 ) ln 1 + β 1 β 2 β ( 2 β 2 )
where ϵ is the energy of the less energetic photon originating from the accretion disk. Also,
β = 1 1 s , s = ϵ E γ ( 1 cos θ 0 ) 2 m e 2 c 4
For pair creation to occur, s > 1 .
The disk optical depth τ d i s k is given by integrating over the X-ray photon energy, the distance that the high-energy photons travels from the emission source to the observer l, and the disk’s radius R and angle ϕ :
τ d i s k = 0 0 2 π R i n R o u t ϵ m i n d n d ϵ d Ω ( 1 cos θ 0 ) σ γ γ ρ cos ω D 3 R d ϵ d R d ϕ d l
where ϵ m i n = 2 m e 2 c 4 / E γ ( 1 cos θ 0 ) , ρ is the collision point distance to the central object, D is the distance between the disk’s surface element and the collision point, and ω is the angle between ρ and the z axis.
Analytically, the above quantities are
ρ = z 2 + l 2 + 2 l z cos i 1 / 2 , D = R 2 + ρ 2 2 R ρ cos ψ 1 / 2 ψ = cos 1 l sin i cos ϕ ρ , ω = cos 1 z + l cos i ρ
where ψ is the angle between ρ and R, and i is the inclination angle of the jet’s axis to the observer’s line of sight. If we assume thermal equilibrium between the infinitesimal surface elements of the disk, the emitted photon density follows the black-body radiation spectrum:
d n d ϵ d Ω = 2 h 2 c 3 ϵ 2 e T ( R ) 1
where T ( R ) = ϵ / k B T ( R ) with k B being the Boltzmann constant and h the Planck constant. As we have already mentioned, the disc begins at R i n = R I S C O , with R I S C O given by Equation (9).
For the temperature of the disk T ( R ) , we use the temperature profile which is obtained by the Paczyński–Wiita potential (see Ref. [40]):
Φ P W = G M b h r r S
The latter equation gives the temperature profile (see Ref. [41]):
T ( r ) = T o r ¯ 2 / 3 r ¯ ( r ¯ 2 ) 3 1 3 3 / 2 ( r ¯ 2 ) 2 1 / 2 r ¯ 3 / 2 1 / 4
where r ¯ = r / r g . Furthermore, the dependence on the black hole’s mass and the accretion rate is sustained through the following constant:
T o = 3 G M B H M ˙ 8 π σ S B r g 3 1 / 4 .
For a more realistic description of the accretion disk around a rotating black hole, we could incorporate the potential proposed from Mukhopadhyay [42]. From this potential, we can extract a temperature profile for the accretion disk [43] as
T ( r ) = T o 4 R ( r ¯ ) 2 3 r ¯ 4 L ( r ¯ ) 1 / 4 1 R ( r ¯ i n ) R ( r ¯ ) 1 / 4 ,
where we have defined the following functions:
R ( r ¯ ) = r ¯ 2 2 α * r ¯ + α * 2 r ¯ ( r ¯ 2 ) + α * L ( r ¯ ) = 2 R ( r ¯ ) 3 r ¯ 3 r ¯ ( r ¯ 2 ) + α * 2 r ¯ α * r ¯ 3 r ¯ 2 2 r ¯ R ( r ¯ ) .
In future works, we will use the above temperature profile in our calculations [44].

3. Results and Discussion

The results obtained in this work refer to an ideal BHXRB system with parameters commonly found in typical BHXRBs. We set different values for the spin parameter of the black hole of the system in order to determine what effect it has on the results while changing it. The main parameters of the system are found in Table 1.
For the above parameters, we calculate the luminosity of the disk ( L d i s k ) as a function of the photon energy of the disk’s black-body spectrum, which is depicted in Figure 1 for various values of the dimensionless spin parameter α * . As can be seen, the luminosity becomes significant for larger energy values as α * increases. This occurs because as α * increases, the R I S C O becomes smaller, so the disk’s inner radius ( R i n ) is located at a smaller radius where its temperature is higher. As a result of this, the black-body spectrum of the accretion disk increases.
We also consider that the jet’s hadron-to-lepton ratio is α k = L p / L e = 1 , where L p ( L e ) stands for the proton (electron) kinetic luminosity. At this point we assume that a small portion of the particles, q r = 10 4 , is accelerated in the acceleration zone of the jet.
In general, in BHXRB systems the acceleration zone of the jet is taken to be from z 0 = 10 8 cm (where z 0 represents the distance from the center of the black hole to the start of the acceleration zone) up to z m a x = 5 z 0 (where z m a x denotes the distance from the center of the black hole to the end of the acceleration zone). The acceleration efficiency is taken to be η = 0.1 .
For these parameters, we now calculate the exponential of minus all the optical depths that are found in Equation (24), which is shown in Figure 2. What we can see is that the larger the spin parameter, the larger the absorption, and when we go to negative values of α * , the absorption is even smaller.
Considering a leptonic model, the radiation emission is produced by synchrotron emission from accelerated jet electrons and protons, and inverse Compton scattering [45] between disk photons and jet electrons. We have calculated the differential flux after absorption. It is depicted in Figure 3. We observe that the larger the spin, the lower the differential flux, while for negative values of α * , the differential flux is larger.
We can conclude that for larger values of the spin parameter, we have more absorption and thus a lower differential flux. This can be explained by the value of the R I S C O . For more negative values of the spin parameter, the R I S C O is larger, so the accretion disk is smaller, and in this way, it absorbs less photons from the jet. On the other hand, going to larger positive values of the spin parameter, the R I S C O becomes smaller, so the accretion disk is larger and has higher temperatures, and in this way, it contributes more to the absorption.
In order to determine a more concrete conclusion from our results, our model should be applied to real observed BHXRB systems. Such systems could be the MAXI J1820+070 (see Refs. [46,47,48,49,50,51,52,53,54]), the XTE J1550-564 (see Refs. [55,56,57,58,59]), and the XTE J1859+226 (see Refs. [60,61,62,63]). These three systems have different values of α * , while their other parameters are close, so by performing calculations for them, we can further determine the effect of the black hole’s spin on the emitted spectra.

4. Conclusions

In this work we first addressed the R I S C O in the Kerr geometry which depends on the dimensionless spin parameter a * . We presented the main interactions that take place inside the jets leading to the production of gamma-rays, and the transfer equation that describes the particles’ energy distribution. Also, we discussed the gamma-ray source function, from which the gamma-ray intensity as well as integral flux emanate. Further, we examined how photons emitted from the jet can be absorbed due to e + e pair creation.
Moreover, we obtained results for an ideal BHXRB system. Specifically we studied the accretion disk’s luminosity, the absorption, and the differential flux remaining after the absorption effect. Having performed calculations for various values of the spin parameter α * , we concluded the following. When the spin has more negative values, the absorption effect becomes weaker and, thus, the differential flux is larger. When the spin parameter increases (becomes more positive), a larger luminosity of the accretion disk occurs; therefore, the absorption effect is more intense and, thus, the differential flux becomes smaller.
Our next goal is to apply our model to real observed BHXRB systems, such as the MAXI J1820+070, the XTE J1550-564, and the XTE J1859+226, in order to obtain results for their neutrino and gamma-ray intensities and integral fluxes expected to be observed at Earth (see Ref. [44]).

Author Contributions

Conceptualization, T.K. and O.K.; methodology, O.K.; software, O.K.; validation, T.K. and O.K.; formal analysis, D.R. and T.P.; investigation, D.R. and T.P.; resources, T.K.; data curation, T.K. and O.K.; writing—original draft preparation, D.R.; writing—review and editing, D.R. and T.K.; visualization, T.P.; supervision, T.K. and O.K.; project administration, T.K. and O.K.; funding acquisition, T.K. All authors have read and agreed to the published version of the manuscript.

Funding

The authors Th.P., O.T.K., and T.S.K. wish to acknowledge financial support from the Association for Advancement of Research on Open Problems (OPRA Association, Tel Aviv, Israel) in Nuclear Physics & Particle Physics, project No 83241/ELKE-UoI.

Data Availability Statement

The datasets generated during and/or analyzed in the present study are available upon request from the corresponding author.

Conflicts of Interest

Author Odysseas Kosmas was employed by the company Conigital Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
BHXRBBlack Hole X-ray Binary
R I S C O Radius of the Innermost Stable Circular Orbit

References

  1. Schee, J.; Stuchlík, Z. Profiles of emission lines generated by rings orbiting braneworld Kerr black holes. Gen. Relativ. Gravit. 2009, 41, 1795–1818. [Google Scholar] [CrossRef]
  2. Bambi, C. Testing the space-time geometry around black hole candidates with the analysis of the broad K α iron line. Phys. Rev. 2013, 87, 023007. [Google Scholar]
  3. Bambi, C. Broad K α iron line from accretion disks around traversable wormholes. Phys. Rev. 2013, 87, 084039. [Google Scholar]
  4. Johannsen, T.; Psaltis, D. Testing the no-hair theorem with observations in the electromagnetic spectrum. IV. Relativistically broadened iron lines. Astrophys. J. 2013, 773, 57. [Google Scholar] [CrossRef]
  5. Jiang, J.; Bambi, C.; Steiner, J.F. Testing the Kerr nature of black hole candidates using iron line spectra in the CPR framework. Astrophys. J. 2015, 811, 130. [Google Scholar] [CrossRef]
  6. Ni, Y.; Zhou, M.; Cardenas-Avendano, A.; Bambi, C.; Herdeiro, C.A.; Radu, E. Iron Kα line of Kerr black holes with scalar hair. J. Cosmol. Astropart. Phys. 2016, 2016, 049. [Google Scholar] [CrossRef]
  7. Zhou, M.; Cardenas-Avendano, A.; Bambi, C.; Kleihaus, B.; Kunz, J. Search for astrophysical rotating Ellis wormholes with X-ray reflection spectroscopy. Phys. Rev. D 2016, 94, 024036. [Google Scholar] [CrossRef]
  8. Bambi, C.; Abdikamalov, A.B.; Ayzenberg, D.; Cao, Z.; Liu, H.; Nampalliwar, S.; Tripathi, A.; Wang-Ji, J.; Xu, Y. relxill_nk: A Relativistic Reflection Model for Testing Einstein’s Gravity. Universe 2018, 4, 79. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Abdikamalov, A.B.; Ayzenberg, D.; Bambi, C.; Dauser, T.; García, J.A.; Nampalliwar, S. About the Kerr Nature of the Stellar-mass Black Hole in GRS 1915+105. Astrophys. J. 2019, 875, 41. [Google Scholar]
  10. Kerr, R.P. Gravitational field of a spinning mass as an example of algebraically special metrics. Phys. Rev. Lett. 1963, 11, 237. [Google Scholar] [CrossRef]
  11. Romero, G.E.; Vila, G.S. Introduction to Black Hole Astrophysics; Springer: Berlin/Heidelberg, Germany, 2013; Volume 876. [Google Scholar]
  12. Newman, E.T.; Couch, E.; Chinnapared, K.; Exton, A.; Prakash, A.; Torrence, R. Metric of a rotating, charged mass. J. Math. Phys. 1965, 6, 918–919. [Google Scholar] [CrossRef]
  13. McClintock, J.E.; Narayan, R.; Davis, S.W.; Gou, L.; Kulkarni, A.; Orosz, J.A.; Penna, R.F.; Remillard, R.A.; Steiner, J.F. Measuring the spins of accreting black holes. Class. Quantum Gravity 2011, 28, 114009. [Google Scholar] [CrossRef]
  14. Papavasileiou, T.V.; Kosmas, O.T.; Sinatkas, I. Prediction of gamma-ray emission from Cygnus X-1, SS 433, and GRS 1915+105 after absorption. Astron. Astrophys. 2023, 673, A162. [Google Scholar] [CrossRef]
  15. Papavasileiou, T.V.; Kosmas, O.T.; Sinatkas, I. Studying the Spectral Energy Distributions Emanating from Regular Galactic XRBs. Universe 2023, 9, 312. [Google Scholar] [CrossRef]
  16. Kosmas, O.T.; Papavasileiou, T.V.; Kosmas, T.S. Integral Fluxes of Neutrinos and Gamma-Rays Emitted from Neighboring X-ray Binaries. Universe 2023, 9, 517. [Google Scholar] [CrossRef]
  17. Hobson, M.P.; Efstathiou, G.P.; Lasenby, A.N. General Relativity: An Introduction for Physicists; Cambridge University Press: New York, NY, USA, 2006; pp. 310–354. [Google Scholar]
  18. Bardeen, J.M.; Press, W.H.; Teukolsky, S.A. Rotating black holes: Locally nonrotating frames, energy extraction, and scalar synchrotron radiation. Astrophys. J. 1972, 178, 347–370. [Google Scholar] [CrossRef]
  19. Böttcher, M.; Dermer, C.D. Photon-Photon Absorption of Very High Energy Gamma Rays from Microquasars: Application to LS 5039. Astrophys. J. 2005, 634, L81–L84. [Google Scholar] [CrossRef]
  20. Cerutti, B.; Dubus, G.; Malzac, J.; Szostek, A.; Belmont, R.; Zdziarski, A.A.; Henri, G. Absorption of high-energy gamma rays in Cygnus X-3. Astron. Astrophys. 2011, 529, A120. [Google Scholar] [CrossRef]
  21. Romero, G.E.; Vila, G.S. The proton low-mass microquasar: High-energy emission. Astron. Astrophys. 2008, 485, 623–631. [Google Scholar] [CrossRef]
  22. Mannheim, K.; Schlickeiser, R. Interactions of cosmic ray nuclei. Astron. Astrophys. 1994, 286, 983–996. [Google Scholar]
  23. Smponias, T.; Kosmas, O.T. Neutrino Emission from Magnetized Microquasar Jets. Adv. High Energy Phys. 2017, 2017, 4962741. [Google Scholar] [CrossRef]
  24. Romero, G.E.; Torres, D.F.; Kaufman Bernadó, M.M.; Mirabel, I.F. Hadronic gamma-ray emission from windy microquasars. Astron. Astrophys. 2003, 410, L1–L4. [Google Scholar] [CrossRef]
  25. Zhang, J.F.; Li, Z.R.; Xiang, F.Y.; Lu, J.F. Electron transport with re-acceleration and radiation in the jets of X-ray binaries. Mon. Not. R. Astron. Soc. 2017, 473, 3211–3222. [Google Scholar] [CrossRef]
  26. Kosmas, O.T.; Leyendecker, S. Phase lag analysis of variational integrators using interpolation techniques. PAMM Proc. Appl. Math. Mech. 2012, 12, 677–678. [Google Scholar]
  27. Kosmas, O.T.; Leyendecker, S. Family of high order exponential variational integrators for split potential systems. J. Phys. Conf. Ser. 2015, 574, 012002. [Google Scholar] [CrossRef]
  28. Kosmas, O.T.; Vlachos, D.S. A space-time geodesic approach for phase fitted variational integrators. Phys. Conf. Ser. 2016, 738, 012133. [Google Scholar] [CrossRef]
  29. Kantzas, D.; Markoff, S.; Beuchert, T.; Lucchini, M.; Chhotray, A.; Ceccobello, C.; Tetarenko, A.J.; Miller-Jones, J.C.A.; Bremer, M.; Garcia, J.A.; et al. A new lepto-hadronic model applied to the first simultaneous multiwavelength data set for Cygnus X–1. Mon. Not. R. Astron. Soc. 2021, 500, 2112–2126. [Google Scholar] [CrossRef]
  30. Carulli, A.M.; Reynoso, M.M.; Romero, G.E. Neutrino production in Population III microquasars. Astropart. Phys. 2021, 128, 102557. [Google Scholar] [CrossRef]
  31. Papavasileiou, T.; Kosmas, O.; Sinatkas, I. Simulations of Neutrino and Gamma-Ray Production from Relativistic Black-Hole Microquasar Jets. Galaxies 2021, 9, 67. [Google Scholar] [CrossRef]
  32. Papavasileiou, T.; Kosmas, O.; Sinatkas, I. Relativistic Magnetized Astrophysical Plasma Outflows in Black-Hole Microquasars. Symmetry 2022, 14, 485. [Google Scholar] [CrossRef]
  33. Kelner, S.R.; Aharonian, F.A.; Bugayov, V.V. Energy spectra of gamma rays, electrons, and neutrinos produced at proton-proton interactions in the very high energy regime. Phys. Rev. D 2006, 74, 034018. [Google Scholar] [CrossRef]
  34. Smponias, T.; Kosmas, O.T. High Energy Neutrino Emission from Astrophysical Jets in the Galaxy. Adv. High Energy Phys. 2015, 2015, 921757. [Google Scholar] [CrossRef]
  35. Kosmas, O.; Smponias, T. Simulations of Gamma-Ray Emission from Magnetized Microquasar Jets. J. Adv. High Energy Phys. 2018, 2018, 9602960. [Google Scholar] [CrossRef]
  36. Reynoso, M.M.; Romero, G.E.; Christiansen, H.R. Production of gamma rays and neutrinos in the dark jets of the microquasar SS433. Mon. Not. R. Astron. Soc. 2008, 387, 1745–1754. [Google Scholar] [CrossRef]
  37. Reynoso, M.M.; Romero, G.E. Magnetic field effects on neutrino production in microquasars. Astron. Astrophys. 2009, 493, 1–111. [Google Scholar] [CrossRef]
  38. Shakura, N.I.; Sunyaev, R.S. Black holes in binary systems. Observational appearance. Astron. Astrophys. 1973, 24, 337–355. [Google Scholar]
  39. Gould, R.J.; Schréder, G.P. Pair production in photon-photon collisions. Phys. Rev. 1967, 155, 1404. [Google Scholar] [CrossRef]
  40. Paczynsky, B.; Wiita, P.J. Thick accretion disks and supercritical luminosities. Astron. Astrophys. 1980, 88, 23–31. [Google Scholar]
  41. Gierliński, M.; Zdziarski, A.A.; Poutanen, J.; Coppi, P.S.; Ebisawa, K.; Johnson, W.N. Radiation mechanisms and geometry of Cygnus X-1 in the soft state. Mon. Not. R. Astron. Soc. 1999, 309, 492–512. [Google Scholar] [CrossRef]
  42. Mukhopadhyay, B. Description of pseudo-Newtonian potential for the relativistic accretion disks around Kerr black holes. Astrophys. J. 2002, 581, 427. [Google Scholar] [CrossRef]
  43. Papavasileiou, T.V.; Kosmas, O.; Kosmas, T.S. A direct method for reproducing fully relativistic spectra from standard accretion disks by modifying their inner boundary. arXiv 2024, arXiv:2408.02415. [Google Scholar]
  44. Rarras, D.; Kosmas, O.; Papavasileiou, T.; Kosmas, T. Black Hole’s Spin-Dependence of γ-ray and Neutrino Emissions from MAXI J1820+070, XTE J1550-564, and XTE J1859+226. Particles 2024. to be published. [Google Scholar]
  45. Blumenthal, G.R.; Gould, R.J. Bremsstrahlung, synchrotron radiation, and compton scattering of high-energy electrons traversing dilute gases. Rev. Mod. Phys. 1970, 42, 237. [Google Scholar] [CrossRef]
  46. Torres, M.A.P.; Casares, J.; Jiménez-Ibarra, F.; Álvarez-Hernández, A.; Muñoz-Darias, T.; Padilla, M.A.; Jonker, P.G.; Heida, M. The binary mass ratio in the black hole transient MAXI J1820+070. Astrophys. J. Lett. 2020, 893, L37. [Google Scholar] [CrossRef]
  47. Mikołajewska, J.; Zdziarski, A.A.; Ziółkowski, J.; Torres, M.A.; Casares, J. The Donor of the Black Hole X-ray Binary MAXI J1820+070. Astrophys. J. 2022, 930, 9. [Google Scholar] [CrossRef]
  48. Poutanen, J.; Veledina, A.; Berdyugin, A.V.; Berdyugina, S.V.; Jermak, H.; Jonker, P.G.; Kajava, J.J.E.; Kosenkov, I.A.; Kravtsov, V.; Piirola, V.; et al. Black hole spin–orbit misalignment in the X-ray binary MAXI J1820+070. Science 2022, 375, 874–876. [Google Scholar] [CrossRef]
  49. Kalogera, V. Spin-orbit misalignment in close binaries with two compact objects. Astrophys. J. 2000, 541, 319. [Google Scholar] [CrossRef]
  50. Atri, P.; Miller-Jones, J.C.A.; Bahramian, A.; Plotkin, R.M.; Deller, A.T.; Jonker, P.G.; Maccarone, T.J.; Sivakoff, G.R.; Soria, R.; Altamirano, D.; et al. A radio parallax to the black hole X-ray binary MAXI J1820+070. Mon. Not. R. Astron. Soc. Lett. 2020, 493, L81–L86. [Google Scholar] [CrossRef]
  51. Zdziarski, A.A.; Tetarenko, A.J.; Sikora, M. Jet Parameters in the Black Hole X-ray Binary MAXI J1820+070. Astrophys. J. 2022, 925, 189. [Google Scholar] [CrossRef]
  52. Zhao, X.; Gou, L.; Dong, Y.; Tuo, Y.; Liao, Z.; Li, Y.; Jia, N.; Feng, Y.; Steiner, J.F. Estimating the black hole spin for the X-ray binary MAXI J1820+070. Astrophys. J. 2021, 916, 108. [Google Scholar] [CrossRef]
  53. Draghis, P.A.; Miller, J.M.; Zoghbi, A.; Reynolds, M.; Costantini, E.; Gallo, L.C.; Tomsick, J.A. A systematic view of ten new black hole spins. Astrophys. J. 2023, 945, 19. [Google Scholar] [CrossRef]
  54. Bhargava, Y.; Belloni, T.; Bhattacharya, D.; Motta, S.; Ponti, G. A timing-based estimate of the spin of the black hole in MAXI J1820+070. Mon. Not. R. Astron. Soc. 2021, 508, 3104–3110. [Google Scholar] [CrossRef]
  55. Orosz, J.A.; Steiner, J.F.; McClintock, J.E.; Torres, M.A.; Remillard, R.A.; Bailyn, C.D.; Miller, J.M. An improved dynamical model for the microquasar XTE J1550-564. Astrophys. J. 2011, 730, 75. [Google Scholar] [CrossRef]
  56. Miller-Jones, J.C.A.; Fender, R.P.; Nakar, E. Opening angles, Lorentz factors and confinement of X-ray binary jets. Mon. Not. R. Astron. Soc. 2006, 367, 1432–1440. [Google Scholar] [CrossRef]
  57. Motta, S.E.; Munoz-Darias, T.; Sanna, A.; Fender, R.; Belloni, T.; Stella, L. Black hole spin measurements through the relativistic precession model: XTE J1550-564. Mon. Not. R. Astron. Soc. Lett. 2014, 439, L56–L69. [Google Scholar] [CrossRef]
  58. Steiner, J.F.; Reis, R.C.; McClintock, J.E.; Narayan, R.; Remillard, R.A.; Orosz, J.A.; Gou, L.; Fabian, A.C.; Torres, M.A.P. The spin of the black hole microquasar XTE J1550-564 via the continuum-fitting and Fe-line methods. Mon. Not. R. Astron. Soc. 2011, 416, 941–958. [Google Scholar] [CrossRef]
  59. Kaaret, P.; Corbel, S.; Tomsick, J.A.; Fender, R.; Miller, J.M.; Orosz, J.A. Tzioumis, A.K.; Wijnands, R. X-ray Emission from the Jets of XTE J1550-564. Astrophys. J. 2003, 582, 945. [Google Scholar] [CrossRef]
  60. Nandi, A.; Mandal, S.; Sreehari, H.; Radhika, D.; Das, S.; Chattopadhyay, I.; Iyer, N.; Agrawal, V.K.; Aktar, R. Accretion flow dynamics during 1999 outburst of XTE J1859+ 226-modeling of broadband spectra and constraining the source mass. Astrophys. Space Sci. 2018, 363, 1–12. [Google Scholar] [CrossRef]
  61. Kimura, M.; Done, C. Evolution of X-ray irradiation during the 1999–2000 outburst of the black hole binary XTE J1859+ 226. Mon. Not. R. Astron. Soc. 2019, 482, 626–638. [Google Scholar] [CrossRef]
  62. Yanes-Rizo, I.V.; Torres, M.A.P.; Casares, J.; Motta, S.E.; Muñoz-Darias, T.; Rodríguez-Gil, P.; Armas Padilla, M.; Jiménez-Ibarra, F.; Jonker, P.G.; Corral-Santana, J.M.; et al. A refined dynamical mass for the black hole in the X-ray transient XTE J1859+ 226. Mon. Not. R. Astron. Soc. 2022, 517, 1476–1482. [Google Scholar] [CrossRef]
  63. Motta, S.E.; Belloni, T.; Stella, L.; Pappas, G.; Casares, J.; Muñoz-Darias, A.T.; Torres, M.A.P.; Yanes-Rizo, I.V. Black hole mass and spin measurements through the relativistic precession model: XTE J1859+ 226. Mon. Not. R. Astron. Soc. 2022, 517, 1469–1475. [Google Scholar] [CrossRef]
Figure 1. The luminosity ( L d i s k ) of the disk with respect to photon energy (E) for various values of α * .
Figure 1. The luminosity ( L d i s k ) of the disk with respect to photon energy (E) for various values of α * .
Particles 07 00046 g001
Figure 2. The exponential of minus the sum of the optical depths ( e τ a l l ) with respect to photon energy (E) for various values of α * .
Figure 2. The exponential of minus the sum of the optical depths ( e τ a l l ) with respect to photon energy (E) for various values of α * .
Particles 07 00046 g002
Figure 3. The differential flux ( d Φ γ / d E ) of synchrotron emission from accelerated jet electrons and protons and inverse Compton scattering between disk photons and jet electrons with respect to log E for various values of α * .
Figure 3. The differential flux ( d Φ γ / d E ) of synchrotron emission from accelerated jet electrons and protons and inverse Compton scattering between disk photons and jet electrons with respect to log E for various values of α * .
Particles 07 00046 g003
Table 1. Main parameters of the ideal BHXRB system.
Table 1. Main parameters of the ideal BHXRB system.
BHXRB ParameterSymbol (Units)Value
Black hole mass M b h   ( M ) 20
Donor star mass M d   ( M ) 15
Distance to Earth d   ( k p c ) 2
Donor star luminosity L s t a r   ( L ) 10 4
Donor star temperature T d   ( K ) 10,000
Jet’s inclination i   ( ) 30
Jet’s Lorentz factor Γ b   ( c ) 2.29
Jet’s half-opening angle ξ ( ) 2.0
Black hole’s spin parameter α * ± 0 , 0.6 , 0.9
Mass accretion rate M ˙   ( M ˙ E d d ) 0.07
Jet’s emitting region height z 0   ( c m ) 10 8
Seperation distance of the system s   ( c m ) 3.16 × 10 12
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Rarras, D.; Kosmas, T.; Papavasileiou, T.; Kosmas, O. Galactic Stellar Black Hole Binaries: Spin Effects on Jet Emissions of High-Energy Gamma-Rays. Particles 2024, 7, 792-804. https://doi.org/10.3390/particles7030046

AMA Style

Rarras D, Kosmas T, Papavasileiou T, Kosmas O. Galactic Stellar Black Hole Binaries: Spin Effects on Jet Emissions of High-Energy Gamma-Rays. Particles. 2024; 7(3):792-804. https://doi.org/10.3390/particles7030046

Chicago/Turabian Style

Rarras, Dimitrios, Theocharis Kosmas, Theodora Papavasileiou, and Odysseas Kosmas. 2024. "Galactic Stellar Black Hole Binaries: Spin Effects on Jet Emissions of High-Energy Gamma-Rays" Particles 7, no. 3: 792-804. https://doi.org/10.3390/particles7030046

Article Metrics

Back to TopTop