Two models for the orbital modulation of γ𝛾\gammaitalic_γ-rays in Cyg X-3

Anton Dmytriiev Centre for Space Research, North-West University, Potchefstroom, 2520, South Africa; anton.dmytriiev@nwu.ac.za Andrzej A. Zdziarski Nicolaus Copernicus Astronomical Center, Polish Academy of Sciences, Bartycka 18, PL-00-716 Warszawa, Poland; aaz@camk.edu.pl Denys Malyshev Institut für Astronomie und Astrophysik Tübingen, Universität Tübingen, Sand 1, D-72076 Tübingen, Germany Valentí Bosch-Ramon Departament de Física Quàntica i Astrofísica, Institut de Ciències del Cosmos, Universitat de Barcelona, Martí i Franquès 1, 08028 Barcelona, Spain Instituto Argentino de Radioastronomía, CONICET, Berazategui Partido, Buenos Aires Province, Argentina Maria Chernyakova School of Physical Sciences and Centre for Astrophysics & Relativity, Dublin City University, Glasnevin, D09 W6Y4, Ireland Dublin Institute for Advanced Studies, 31 Fitzwilliam Place, Dublin 2, Ireland
Abstract

We model the currently available γ𝛾\gammaitalic_γ-ray data from the Fermi Large Area Telescope on Cyg X-3. Thanks to its very strong γ𝛾\gammaitalic_γ-ray activity during 2018–2021, the data quality has significantly improved. We study the strong orbital modulation of the γ𝛾\gammaitalic_γ-rays observed during at high γ𝛾\gammaitalic_γ-ray fluxes. The modulation, as found earlier, is well modeled by anisotropic Compton scattering of the donor blackbody emission by relativistic electrons in a jet strongly misaligned with respect to the orbital axis. We confirm that this model fits well both the average γ𝛾\gammaitalic_γ-ray modulation light curve and the spectrum. However, we find that if the jet is aligned with the spin axis of a rotating black hole, it would undergo geodetic precession with the period of similar-to\sim50 years. However, its presence is ruled out by both the γ𝛾\gammaitalic_γ-ray and radio data. Therefore, we consider an alternative model in which the average jet direction jet is aligned, but it is bent to outside the orbit owing to the thrust of the donor stellar wind, and thus precesses at the orbital period. The γ𝛾\gammaitalic_γ-ray modulation appears then owing to the variable Doppler boosting of synchrotron self-Compton jet emission. The model also fits well the data. However, the fitted bending angle is much larger than the theoretical one based on the binary and wind parameters as currently known. Thus, both models disagree with important aspects of our current theoretical understanding of the system. We discuss possible ways to find the correct model.

\declare@file@substitution

revtex4-1.clsrevtex4-2.cls

1 Introduction

The high-mass binary Cyg X-3 is one of the first discovered (Giacconi et al., 1967) X-ray binaries. Still, after many years of intense research, it remains to be a puzzling (and unique) system. The nature of its compact object remains uncertain. It can be either a neutron star or a black hole (BH), see Koljonen & Maccarone (2017) and Antokhin et al. (2022), hereafter A22, for discussions. Still, those authors favored the presence of a BH, which is also supported by various aspects of the X-ray and radio emission (Hjalmarsdotter et al., 2008, 2009; Szostek & Zdziarski, 2008; Szostek et al., 2008; Koljonen et al., 2010). Its donor is a Wolf-Rayet (WR) star (van Kerkwijk et al., 1992, 1996; Koljonen & Maccarone, 2017), which makes it the only known binary system in the Milky Way consisting of a WR star and a compact object, and a candidate for a future merger (Belczyński et al., 2013). The compact object accretes a fraction of the stellar wind from the donor. A22 estimated the total mass of the system as 18.8M18.8subscript𝑀18.8{M}_{\sun}18.8 italic_M start_POSTSUBSCRIPT ☉ end_POSTSUBSCRIPT, and the compact object mass as Mc7.2Msubscript𝑀c7.2subscript𝑀M_{\rm c}\approx 7.2{M}_{\sun}italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ≈ 7.2 italic_M start_POSTSUBSCRIPT ☉ end_POSTSUBSCRIPT, implying it is a BH. However, they also considered lower masses possible if the wind is significantly clumped. Its distance has recently been determined based on a radio parallax as D9.7±0.5𝐷plus-or-minus9.70.5D\approx 9.7\pm 0.5italic_D ≈ 9.7 ± 0.5 kpc and based on Galactic proper motions and line-of-sight radial velocity measurements as 9±1plus-or-minus919\pm 19 ± 1 kpc (Reid & Miller-Jones, 2023). We hereafter assume D=9𝐷9D=9italic_D = 9 kpc, M=11.6Msubscript𝑀11.6subscript𝑀M_{*}=11.6{M}_{\sun}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 11.6 italic_M start_POSTSUBSCRIPT ☉ end_POSTSUBSCRIPT, Mc=7.2Msubscript𝑀c7.2subscript𝑀M_{\rm c}=7.2{M}_{\sun}italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT = 7.2 italic_M start_POSTSUBSCRIPT ☉ end_POSTSUBSCRIPT.

Cyg X-3 is also the brightest and most variable radio source among X-ray binaries, showing resolved relativistic jets (Mioduszewski et al., 2001; Miller-Jones et al., 2004). Its jets also emit high-energy γ𝛾\gammaitalic_γ-rays, as first discovered by Mori et al. (1997), and later confirmed by observations with the Fermi Large Area Telescope (LAT; Fermi LAT Collaboration et al. 2009), and AGILE (Tavani et al., 2009). As found by Fermi LAT Collaboration et al. (2009), the γ𝛾\gammaitalic_γ-rays emitted during flaring epochs are strongly orbitally modulated. The modulated emission was explained by Dubus et al. (2010) as anisotropic Compton scattering of blackbody photons from the donor by relativistic electrons with a power-law distribution accelerated in the jet (which we hereafter refer to as either the blackbody Compton model or Model 1). The peak of the emission of a jet aligned with the orbital axis would be at the superior conjunction of the compact object. However, Dubus et al. (2010) found that the peak was clearly before the superior conjunction, indicating that the jet is inclined with respect to the orbital axis. With the number of high-energy γ𝛾\gammaitalic_γ-rays increasing with time, the statistical accuracy of the orbital modulation increased, and this model was fitted to the updated data by Zdziarski et al. (2018), hereafter Z18.

Refer to caption

Figure 1: The LAT high-energy γ𝛾\gammaitalic_γ-ray light curve of Cyg X-3 from the beginning of the Fermi observations until MJD 60200. The red and blue symbols represent the detections within a day and upper limits, respectively. Our criterion defining the γ𝛾\gammaitalic_γ-ray bright state for 1-d integration is shown by the horizontal dotted line. Z18 analyzed the data until MJD 57982, which is marked by the vertical dotted line. We see that the later observations have multiplied the number of detected photons by a substantial factor.

The observations analyzed by Z18 were until MJD 57982. As we have found out by analyzing the LAT data, Cyg X-3 has shown very strong activity after that date, see Figure 1. The new data allow us to increase the number of detected γ𝛾\gammaitalic_γ-ray photons by a large factor, and, consequently, to strongly reduce the errors on the light curve averaged and folded over the orbital period.

In our modeling, we consider an important consequence of the jet–binary axis misalignment. Jets powered by the rotation of BHs (Blandford & Znajek, 1977) are launched along the BH spin axis (McKinney et al., 2013). Jets launched via poloidal magnetic field of accretion disks (Blandford & Payne, 1982) will initially be directed along the inner disk rotation axis, which could, in turn, be also aligned with the BH spin (Bardeen & Petterson, 1975). In those cases, the observed jet–orbit misalignment would correspond to the BH spin–orbit one. However, spin precession due to the binary angular momentum–compact object spin coupling has to take place in general relativity, and the period of such geodetic precession for Cyg X-3 is as short as similar-to\sim50 years. Such precession is clearly not seen in the Fermi data, which span \approx14 years, i.e., more than a quarter of that period. It is also not seen in the similar-to\sim30 years of the radio imaging data, in which the jet position angle remains constant along the north-to-south direction. As shown in this work, these observations present an argument against the misaligned jet model.

Therefore, we also consider an alternative model of the orbital modulation in which it is due to the Doppler boosting of the jet synchrotron self-Compton (SSC) emission. This could be achieved if the jet is bent to outside by the thrust of the stellar wind (Yoon & Heinz, 2015; Yoon et al., 2016; Bosch-Ramon & Barkov, 2016; Molina & Bosch-Ramon, 2018; Molina et al., 2019; Barkov & Bosch-Ramon, 2022), with the wind being very strong in this binary. In addition, the Coriolis force due to the binary rotation bends the jet opposite to the rotation direction. We hereafter refer to it as either the orbital precession model or Model 2. We have found that the resulting modulated Doppler boosting can fit the data at an accuracy similar to the blackbody Compton model. In this model, the average jet direction is fully aligned with the orbital axis, and thus no geodetic precession is expected. An important prediction of this model is an orbital modulation of the associated synchrotron emission, with the orbital light curve similar to that of the γ𝛾\gammaitalic_γ-rays. However, we find that the fitted bending angle is much larger than that expected theoretically based on the likely wind and jet parameters. This, in turn, represents a strong argument against this model.

We present our detailed results for both models. We then discuss possible solutions of their problems for each of them. According to our present knowledge, we cannot determine which model is more likely.

Refer to caption

Figure 2: The Lomb-Scargle periodogram in the γ𝛾\gammaitalic_γ-ray bright state in the 0.1–100 GeV range, calculated accounting for the secular orbital period increase and taking into account the measurement uncertainties. The only significant peak is equal within the fitted uncertainties to the orbital period of Cyg X-3.

2 The data analysis

We analyze the data from the direction of Cyg X-3 for MJD 57982–60201, see Figure 1. We used the Fermi Science Tools v. 2.2.0. Details of the analysis generally follow those given in Z18. For the spectral analysis, we used the energy range of 0.05–500 GeV and for the timing analysis, that of 0.1–100 GeV. In this paper, we analyze only γ𝛾\gammaitalic_γ-ray bright states, which we define in the same way as the ‘flaring state’ in Z18. The minimum level of this state is shown by the horizontal line in Figure 1. These fluxes correspond to F(0.1F(0.1italic_F ( 0.1–100 GeV)>5×1010)>5\times 10^{-10}) > 5 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT erg cm-2 s-1 and the test statistics (Mattox et al., 1996) of TS>16TS16{\rm TS}>16roman_TS > 16, i.e., the significance >4σabsent4𝜎>4\sigma> 4 italic_σ. We obtain the average spectrum for that state with positive detections in the 0.05–20 GeV range, and upper limits at higher energies.

We used the quadratic ephemeris given by model 4 of Antokhin & Cherepashchuk (2019) and searched for periodicity in the LAT light curve taking into account their rate of the period increase. We used the Lomb-Scargle method. Our results are shown in Figure 2. We have found the period of P0=0.1996847(2)subscript𝑃00.19968472P_{0}=0.1996847(2)italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.1996847 ( 2 ) d, which is in full agreement with that of Antokhin & Cherepashchuk (2019), P0=0.199684622(15)subscript𝑃00.19968462215P_{0}=0.199684622(15)italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.199684622 ( 15 ) d. We also tested for the effect of using their quadratic+sinusoidal ephemeris, but we found its effect to be very minor. Then, we calculated the folded and averaged light curve in 10 phase bins.

3 Theoretical description

Refer to caption

Figure 3: The geometry corresponding to emission in a compact region of the jet (the counterjet is not shown) for the assumed counter-clockwise rotation. The axes x𝑥xitalic_x and y𝑦yitalic_y are in the binary plane, and the +z𝑧+z+ italic_z direction is parallel to the orbital axis. The x𝑥-x- italic_x direction gives the projection of the direction toward the observer onto the binary plane. The observer is at an angle, i𝑖iitalic_i, with respect to the orbital axis, φ𝜑\varphiitalic_φ is the orbital phase (the superior conjunction is at φ=0𝜑0\varphi=0italic_φ = 0, and the shown configuration is close to the inferior conjunction), θjsubscript𝜃j\theta_{\rm j}italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT and φjsubscript𝜑j\varphi_{\rm j}italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT are the polar and azimuthal angles of the jet with respect to the orbital axis, a𝑎aitalic_a is the binary separation, hhitalic_h is the distance of the γ𝛾\gammaitalic_γ-ray source from the center of the compact object, eorbsubscripteorb{\mathbfit e}_{\rm orb}bold_italic_e start_POSTSUBSCRIPT roman_orb end_POSTSUBSCRIPT is a unit vector perpendicular to the orbital plane, eobssubscripteobs{\mathbfit e}_{\rm obs}bold_italic_e start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT, ecsubscriptec{\mathbfit e}_{\rm c}bold_italic_e start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT and esubscripte{\mathbfit e}_{*}bold_italic_e start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT point from the donor toward the observer, the center of the compact object, and the γ𝛾\gammaitalic_γ-ray source, respectively, and ejsubscriptej{\mathbfit e}_{\rm j}bold_italic_e start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT points from the center of the compact object toward the γ𝛾\gammaitalic_γ-ray source.

3.1 The geometry

Our assumed geometry is shown in Figure 3. The binary rotates with the period P𝑃Pitalic_P, and the azimuth of the rotation with respect to the superior conjunction is φ𝜑\varphiitalic_φ. We assume no eccentricity and the rotation to be counter-clockwise, as found by Veledina et al. (2024a). The γ𝛾\gammaitalic_γ-ray emitting region in the jet is at the distance hhitalic_h from the compact object, which itself is at the binary separation, a𝑎aitalic_a, from the center of the donor. The jet is inclined with respect to the orbital axis as given by the polar and azimuthal angles, θjsubscript𝜃j\theta_{\rm j}italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT and φjsubscript𝜑j\varphi_{\rm j}italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT, respectively, and the distance of the emitting region from the center of the donor is r𝑟ritalic_r. The jet polar angle, φjsubscript𝜑j\varphi_{\rm j}italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT, is either fixed in the case of the blackbody Compton model or dependent on the orbital phase in the orbital precession model.

The unit vectors pointing towards the observer, from the stellar center to the compact object, along the jet, and along z𝑧zitalic_z (parallel to the orbital axis), are

eobs=(sini,0,cosi),ec=(cosφ,sinφ,0),formulae-sequencesubscripteobs𝑖0𝑖subscriptec𝜑𝜑0\displaystyle{\mathbfit e}_{\rm obs}=(-\sin i,0,\cos i),\quad{\mathbfit e}_{% \rm c}=(\cos\varphi,\sin\varphi,0),bold_italic_e start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT = ( - roman_sin italic_i , 0 , roman_cos italic_i ) , bold_italic_e start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT = ( roman_cos italic_φ , roman_sin italic_φ , 0 ) , (1)
ej=(sinθjcosφj,sinθjsinφj,cosθj),eorb=(0,0,1),formulae-sequencesubscriptejsubscript𝜃jsubscript𝜑jsubscript𝜃jsubscript𝜑jsubscript𝜃jsubscripteorb001\displaystyle{\mathbfit e}_{\rm j}=(\sin\theta_{\rm j}\cos\varphi_{\rm j},\sin% \theta_{\rm j}\sin\varphi_{\rm j},\cos\theta_{\rm j}),\quad{\mathbfit e}_{\rm orb% }=(0,0,1),bold_italic_e start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT = ( roman_sin italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT roman_cos italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT , roman_sin italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT roman_sin italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT , roman_cos italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ) , bold_italic_e start_POSTSUBSCRIPT roman_orb end_POSTSUBSCRIPT = ( 0 , 0 , 1 ) , (2)

respectively. We make a simplifying assumption that the modulated emission at a given frequency originates at a single distance, hhitalic_h, from the BH. The vector connecting the donor center with the emission point is aec+hej𝑎subscriptecsubscripteja{\mathbfit e}_{\rm c}+h{\mathbfit e}_{\rm j}italic_a bold_italic_e start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT + italic_h bold_italic_e start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT. The length of this vector gives the distance of that point from the donor, and its square is given by

r2=h2+a2+2hasinθjcos(φjφ).superscript𝑟2superscript2superscript𝑎22𝑎subscript𝜃jsubscript𝜑j𝜑r^{2}=h^{2}+a^{2}+2ha\sin\theta_{\rm j}\cos(\varphi_{\rm j}-\varphi).italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_h italic_a roman_sin italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT roman_cos ( italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT - italic_φ ) . (3)

An important issue concerns the determination of the phase of the superior conjunction. All of the determinations of the ephemeris so far have been based on fitting average profiles of the X-ray emission in various bands (see Bhargava et al. 2017; Antokhin & Cherepashchuk 2019 for the most recent work). The X-ray modulation appears due to the bound-free absorption and scattering by the stellar wind emitted by the donor. Since those profiles are not symmetric, a template designed by van der Klis & Bonnet-Bidaud (1989) was used for the fits. However, the phases corresponding to the observed minimum fluxes using that template are not null (see, e.g., the profiles observed in various X-ray bands and the hard and soft X-ray states of the source in Zdziarski et al. 2012b). This issue was studied in detail by A22. They used the X-ray data from the All Sky Monitor (Levine et al., 1996) onboard Rossi X-ray Timing Explorer, and the infrared data acquired by them. They estimated the superior conjunction based on their three-component wind model fitted to those data to be ϕ0/2π=0.066±0.006subscriptitalic-ϕ02𝜋plus-or-minus0.0660.006\phi_{0}/2\pi=-0.066\pm 0.006italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2 italic_π = - 0.066 ± 0.006, where ϕitalic-ϕ\phiitalic_ϕ is the orbital phase defined by the ephemeris. Since the highest γ𝛾\gammaitalic_γ-ray flux in our folded/averaged light curve is in the phase bin with ϕ/2π=0.15±0.05italic-ϕ2𝜋plus-or-minus0.150.05\phi/2\pi=-0.15\pm 0.05italic_ϕ / 2 italic_π = - 0.15 ± 0.05, the correction to the phase of the superior conjunction is significant. Thus, the true orbital phase of the binary is given by

φ/2π=ϕ/2π+0.066±0.006.𝜑2𝜋plus-or-minusitalic-ϕ2𝜋0.0660.006\varphi/2\pi=\phi/2\pi+0.066\pm 0.006.italic_φ / 2 italic_π = italic_ϕ / 2 italic_π + 0.066 ± 0.006 . (4)

When the jet is inclined with respect to the orbital axis, its viewing angle is different from that of the orbital axis itself. It equals to

ij=arccos(cosicosθjcosφjsinisinθj).subscript𝑖j𝑖subscript𝜃jsubscript𝜑j𝑖subscript𝜃ji_{\rm j}=\arccos(\cos i\cos\theta_{\rm j}-\cos\varphi_{\rm j}\sin i\sin\theta% _{\rm j}).italic_i start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT = roman_arccos ( roman_cos italic_i roman_cos italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT - roman_cos italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT roman_sin italic_i roman_sin italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ) . (5)

Furthermore, the position angle on the sky of the binary differs from that of the jet. The difference between the jet and the binary position angles, ΔλΔ𝜆\Delta\lambdaroman_Δ italic_λ, is given by

cosΔλ=cosicosφjsinθj+sinicosθj1(cosicosθjsinicosφjsinθj)2.Δ𝜆𝑖subscript𝜑jsubscript𝜃j𝑖subscript𝜃j1superscript𝑖subscript𝜃j𝑖subscript𝜑jsubscript𝜃j2\cos\Delta\lambda=\frac{\cos i\cos\varphi_{\rm j}\sin\theta_{\rm j}+\sin i\cos% \theta_{\rm j}}{\sqrt{1-\left(\cos i\cos\theta_{\rm j}-\sin i\cos\varphi_{\rm j% }\sin\theta_{\rm j}\right)^{2}}}.roman_cos roman_Δ italic_λ = divide start_ARG roman_cos italic_i roman_cos italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT roman_sin italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT + roman_sin italic_i roman_cos italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG 1 - ( roman_cos italic_i roman_cos italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT - roman_sin italic_i roman_cos italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT roman_sin italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG . (6)

The value of the binary separation follows via the Kepler law from the assumed total mass of 18.8M18.8subscript𝑀18.8{M}_{\sun}18.8 italic_M start_POSTSUBSCRIPT ☉ end_POSTSUBSCRIPT, yielding a2.66×1011𝑎2.66superscript1011a\approx 2.66\times 10^{11}italic_a ≈ 2.66 × 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT cm. We either assume the binary inclination of i=30°𝑖30°i=30\arcdegitalic_i = 30 ° (A22) or treat it as a free parameter.

3.2 The electrons

The electrons in the emitting regions are assumed to be isotropic in the comoving frame, and move with bulk the velocity βjcsubscript𝛽j𝑐\beta_{\rm j}citalic_β start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT italic_c. Also, they are assumed to have a power law distribution, N(γ)=Kγp𝑁𝛾𝐾superscript𝛾𝑝N(\gamma)=K\gamma^{-p}italic_N ( italic_γ ) = italic_K italic_γ start_POSTSUPERSCRIPT - italic_p end_POSTSUPERSCRIPT from γminsubscript𝛾min\gamma_{\rm min}italic_γ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT to γmaxsubscript𝛾max\gamma_{\rm max}italic_γ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. Hereafter, N(γ)𝑁𝛾N(\gamma)italic_N ( italic_γ ) refers to all of the power-law electrons in the source. However, we normalize the electron distribution to the kinetic energy content (i.e. excluding the rest energy) between γ=1𝛾1\gamma=1italic_γ = 1 to γmaxsubscript𝛾max\gamma_{\rm max}italic_γ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT, Eesubscript𝐸eE_{\rm e}italic_E start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT, instead of K𝐾Kitalic_K.

We also take into account the electrons cooled by the blackbody Compton, synchrotron, SSC and adiabatic losses to below γminsubscript𝛾min\gamma_{\rm min}italic_γ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT. This is important since these particles contribute significantly to the emission at low energies. The electron distribution below γminsubscript𝛾min\gamma_{\rm min}italic_γ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT is determined by a kinetic equation, where the radiative and adiabatic losses compete. Solving the kinetic equation with no injection of fresh electrons below γminsubscript𝛾min\gamma_{\rm min}italic_γ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT yields approximately a broken power-law distribution, with Neγ2proportional-tosubscript𝑁esuperscript𝛾2N_{\rm e}\propto\gamma^{-2}italic_N start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT ∝ italic_γ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT in the range γbrγ<γminsubscript𝛾br𝛾subscript𝛾min\gamma_{\rm br}\leq\gamma<\gamma_{\rm min}italic_γ start_POSTSUBSCRIPT roman_br end_POSTSUBSCRIPT ≤ italic_γ < italic_γ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT (where the radiative losses dominate), and Neγ1proportional-tosubscript𝑁esuperscript𝛾1N_{\rm e}\propto\gamma^{-1}italic_N start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT ∝ italic_γ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT in the range 1γ<γbrless-than-or-similar-to1𝛾subscript𝛾br1\lesssim\gamma<\gamma_{\rm br}1 ≲ italic_γ < italic_γ start_POSTSUBSCRIPT roman_br end_POSTSUBSCRIPT (where the adiabatic losses dominate). The former follows from both the Compton losses in the Thomson regime and the synchrotron ones having γ˙γ2proportional-to˙𝛾superscript𝛾2\dot{\gamma}\propto-\gamma^{2}over˙ start_ARG italic_γ end_ARG ∝ - italic_γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The position of the cooling break γbrsubscript𝛾br\gamma_{\rm br}italic_γ start_POSTSUBSCRIPT roman_br end_POSTSUBSCRIPT is determined self-consistently in each simulation by equating the timescales of total radiative losses and adiabatic losses. We verify that the Compton scattering around γbrsubscript𝛾br\gamma_{\rm br}italic_γ start_POSTSUBSCRIPT roman_br end_POSTSUBSCRIPT is in the Thomson regime.

3.3 Radiation

We assume a one-zone γ𝛾\gammaitalic_γ-ray emitting region in our model. To validate this assumption, we examined the importance of effects of propagation of electrons and the evolution of their distribution along the considered source (Khangulyan et al., 2018), which, if significant, might lead to a departure from the one-zone model. We find that these propagation effects have only a minor impact in our specific case. This is due to the fact that the jet velocity βjsubscript𝛽j\beta_{\rm j}italic_β start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT in all our fits remains below the value identified in Khangulyan et al. (2018), above which such effects become non-negligible.

The treatment of the anisotropic Compton scattering off stellar blackbody photons follows that of Zdziarski et al. (2012c), which employs the full Klein-Nishina (KN) cross section integrated over a broken power law electron distribution. The blackbody photons are calculated for a donor with radius Rsubscript𝑅R_{*}italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT and temperature Tsubscript𝑇T_{*}italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT, but we approximate that emission as coming from a point source. For that process, the total number of electrons determines the observed flux. The emission region has to have a low vertical extent since the orbital profile of the blackbody Compton emission depends sensitively on the height.

For an assumed value of the magnetic field strength, B𝐵Bitalic_B, we compute synchrotron and SSC emission assuming that the jet emitting region has a cylindrical geometry (which is a local approximation to a section of a cone) and is filled with a tangled and uniform magnetic field. The electrons can also be present outside that region, but we neglect their presence in our treatment. The emission is therefore isotropic in the comoving frame. The radius of the cylinder is defined by the opening angle of the jet αjsubscript𝛼j\alpha_{\rm j}italic_α start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT as R=αjh𝑅subscript𝛼jR=\alpha_{\rm j}hitalic_R = italic_α start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT italic_h, and for the height of the cylinder we arbitrarily choose Δh=(1/3)hΔ13\Delta h=(1/3)hroman_Δ italic_h = ( 1 / 3 ) italic_h. We note that the observed synchrotron spectrum for given B𝐵Bitalic_B and K𝐾Kitalic_K does not depend on neither ΔhΔ\Delta hroman_Δ italic_h nor hhitalic_h, with its luminosity being LSB2Kproportional-tosubscript𝐿Ssuperscript𝐵2𝐾L_{\rm S}\propto B^{2}Kitalic_L start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT ∝ italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_K, apart from the synchrotron self-absorbed domain at low frequencies. Then, the shape of the SSC spectrum is also independent of them. However, the ratio of the two luminosities (the Compton dominance), is

LSSCLSLSαjhΔhB2KαjhΔh.proportional-tosubscript𝐿SSCsubscript𝐿Ssubscript𝐿Ssubscript𝛼jΔsuperscript𝐵2proportional-to𝐾subscript𝛼jΔ\frac{L_{\rm SSC}}{L_{\rm S}}\propto\frac{L_{\rm S}}{\alpha_{\rm j}h\Delta hB^% {2}}\propto\frac{K}{\alpha_{\rm j}h\Delta h}.divide start_ARG italic_L start_POSTSUBSCRIPT roman_SSC end_POSTSUBSCRIPT end_ARG start_ARG italic_L start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT end_ARG ∝ divide start_ARG italic_L start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT end_ARG start_ARG italic_α start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT italic_h roman_Δ italic_h italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∝ divide start_ARG italic_K end_ARG start_ARG italic_α start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT italic_h roman_Δ italic_h end_ARG . (7)

In our Model 2, with precessing jet, LSSCsubscript𝐿SSCL_{\rm SSC}italic_L start_POSTSUBSCRIPT roman_SSC end_POSTSUBSCRIPT is fixed by the observed γ𝛾\gammaitalic_γ-ray flux, and hence the predicted LSsubscript𝐿SL_{\rm S}italic_L start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT increases when increasing αjhΔhsubscript𝛼jΔ\alpha_{\rm j}h\Delta hitalic_α start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT italic_h roman_Δ italic_h.

For the single electron synchrotron emissivity, we employ an approximation from eq. (3.40) in Boettcher et al. (2012). We include the synchrotron self-absorption effect by using eq. (3.42) in Boettcher et al. (2012). To obtain the synchrotron radiation energy density, we use the solution of the radiation transfer equation for cylindrical geometry (e.g., eq. 14 in Chiaberge & Ghisellini, 1999), while to calculate the observed synchrotron flux in the absorbed regime, we follow eqs. (18–19) of Zdziarski et al. (2012a). The SSC computation takes into account the KN cross section and follows eqs. (8–12) in Katarzyński et al. (2001) (though excluding the scaling factor of 3/4 for spherical geometry present in their eq. 11). The emission is transformed to the observer’s frame, including contributions from both the jet and counterjet.

We also check the importance of internal γ𝛾\gammaitalic_γγ𝛾\gammaitalic_γabsorption of high-energy γ𝛾\gammaitalic_γ-rays on the soft synchrotron photon field. We find that this effect is negligible within our studied energy range, and becomes pronounced only above 10–20 GeV. At those energies, however, the the source is no longer detected.

3.4 Additional observational constraints

Apart from the LAT γ𝛾\gammaitalic_γ-ray data, we impose some observational constraints in other spectral regions. First, the X-ray EFE𝐸subscript𝐹𝐸EF_{E}italic_E italic_F start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT flux at 100 keV measured in the soft states of Cyg X-3 is \approx0.2 keV cm-2 s-1 (Szostek et al., 2008). The observed similar-to\sim100 keV flux exhibits strong orbital modulation (Zdziarski et al., 2012b) similar to that at lower energies, with the minimum around the superior conjunction, which could be due to wind absorption. This implies that most of the X-ray emission originates in the accretion flow rather than in the jet, whose emission shows a very different modulation pattern. On the other hand, the duration of the flaring state is by a factor of similar-to\sim7 shorter than the one of the soft state. Thus, we require that jet contribution at 100 keV to be <<<0.1 keV cm-2 s11.6×1010{}^{-1}\approx 1.6\times 10^{-10}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT ≈ 1.6 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT erg cm-2 s-1.

Furthermore, the model synchrotron flux in the flaring state cannot exceed available measurements. We use the observational constraint from the Submillimeter Array (SMA) observations during flaring periods in 2018, with the average flux of similar-to\sim300 mJy at 225 GHz (McCollough, 2023). That emission undergoes free-free absorption on electrons of the stellar wind. This process can influence both the observed level of 225 GHz emission and the profile of its orbital modulation. We estimate the distance away from the star at which the optical depth of the free-free absorption of 225 GHz emission drops to unity. Using typical values for the mass loss rate of the donor, velocity and temperature of the stellar wind in Cyg X-3 (e.g., A22), and assuming a fully ionized helium plasma, we find this critical distance to be 20asimilar-toabsent20𝑎\sim 20a∼ 20 italic_a. This provides an approximate constraint on the location of the region responsible for producing the 225 GHz emission.

3.5 The models

We consider two models.

Model 1. The jet is assumed to have a fixed orientation. The dominant radiative process is the anisotropic Compton scattering of stellar blackbody photons in a compact region along the jet. The orbital modulation is entirely due to this process since the Doppler boosting does not depend on the orbital phase. As found before (Dubus et al. 2010; Z18), this model fits well the γ𝛾\gammaitalic_γ-ray folded light curve, while requiring the jet to be significantly misaligned with respect to the orbital axis. Since the jet also contains magnetic field, the SSC contributes to the folded light curve at a constant level, which effect was not calculated before (except for a qualitative consideration in Zdziarski et al. 2012c). Here, we calculate the maximum allowed field strength allowed by the data. The main parameters of this model are the jet angles and the height of the emission region along the jet. The emission of both the jet and the counterjet are taken into account, and we have

φj=const,ecj=ej,φcj=π+φj,θcj=πθj.formulae-sequencesubscript𝜑jconstformulae-sequencesubscriptecjsubscriptejformulae-sequencesubscript𝜑cj𝜋subscript𝜑jsubscript𝜃cj𝜋subscript𝜃j\varphi_{\rm j}={\rm const},\quad{\mathbfit e}_{\rm cj}=-{\mathbfit e}_{\rm j}% ,\quad\varphi_{\rm cj}=\pi+\varphi_{\rm j},\quad\theta_{\rm cj}=\pi-\theta_{% \rm j}.italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT = roman_const , bold_italic_e start_POSTSUBSCRIPT roman_cj end_POSTSUBSCRIPT = - bold_italic_e start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT roman_cj end_POSTSUBSCRIPT = italic_π + italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT roman_cj end_POSTSUBSCRIPT = italic_π - italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT . (8)

Model 2. The jet average direction is aligned with the orbital axis but the thrust of the stellar wind bends the jet outward by an angle θjsubscript𝜃j\theta_{\rm j}italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT along the line connecting the jet with the center of the donor (Yoon & Heinz, 2015; Yoon et al., 2016; Bosch-Ramon & Barkov, 2016; Barkov & Bosch-Ramon, 2022). This effect alone would result in a jet precessing at the orbital period with the azimuth given by φj=φsubscript𝜑j𝜑\varphi_{\rm j}=\varphiitalic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT = italic_φ. However, the Coriolis force arising from the orbital motion will induce a lateral bending of the jet opposite to the direction of the rotation, leading to formation of a helix. Here, we are considering emission of a compact region of the jet, which can be then offset by an arbitrary angle, φosubscript𝜑o\varphi_{\rm o}italic_φ start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT. We note that the Coriolis force acts in the same direction on the jet and counterjet, which is evident upon considering the relevant vector product. This implies that both jet and counterjet experience lateral bending opposite to the direction of rotation. Then, we have

φj=φφo,subscript𝜑j𝜑subscript𝜑o\displaystyle\varphi_{\rm j}=\varphi-\varphi_{\rm o},italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT = italic_φ - italic_φ start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT , (9)
ecj=(sinθjcosφj,sinθjsinφj,cosθj),subscriptecjsubscript𝜃jsubscript𝜑jsubscript𝜃jsubscript𝜑jsubscript𝜃j\displaystyle{\mathbfit e}_{\rm cj}=(\sin\theta_{\rm j}\cos\varphi_{\rm j},% \sin\theta_{\rm j}\sin\varphi_{\rm j},-\cos\theta_{\rm j}),bold_italic_e start_POSTSUBSCRIPT roman_cj end_POSTSUBSCRIPT = ( roman_sin italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT roman_cos italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT , roman_sin italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT roman_sin italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT , - roman_cos italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ) , (10)
φcj=φj,θcj=πθj.formulae-sequencesubscript𝜑cjsubscript𝜑jsubscript𝜃cj𝜋subscript𝜃j\displaystyle\varphi_{\rm cj}=\varphi_{\rm j},\quad\theta_{\rm cj}=\pi-\theta_% {\rm j}.italic_φ start_POSTSUBSCRIPT roman_cj end_POSTSUBSCRIPT = italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT roman_cj end_POSTSUBSCRIPT = italic_π - italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT . (11)

The variable φj,cjsubscript𝜑jcj\varphi_{\rm j,cj}italic_φ start_POSTSUBSCRIPT roman_j , roman_cj end_POSTSUBSCRIPT results then in a variability of the Doppler boosting, which is then responsible for the orbital modulation. We note that our treatment of the geometry is approximate, since we assume a straight jet connecting the origin with the emission region, while the jet actually has a helical shape. The exact details of this helical shape are influenced by the jet acceleration profile, i.e., how the jet acceleration rate (and hence the jet velocity) evolves along the length of the jet. Since we do not have precise information on this acceleration profile, we have to neglect this complication.

A recent study by Yang et al. (2023) suggests the presence of a non-negligible transversal component in the jet of Cyg X-3, albeit observed in the hard state. Nevertheless, it is possible that this transversal component may arise from the wind-induced jet precession, and hence the apparently broadened jet in those observations may be attributed to its helical nature, supporting our assumptions about the jet morphology in the Model 2.

The jet will still be irradiated by the stellar blackbody, which adds a perturbation to the modulation due to the variable Doppler boosting of the precessing jet. We calculate the maximum flux, R2T4proportional-toabsentsuperscriptsubscript𝑅2superscriptsubscript𝑇4\propto R_{*}^{2}T_{*}^{4}∝ italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT, of the blackbody allowed by the model for a given hhitalic_h, constraining in turn the radius and temperature of the donor. The maximum depends on the height of the emission region, hhitalic_h.

We assess the significance of the attenuation of γ𝛾\gammaitalic_γ-rays en route to the observer as they traverse the photon field of the donor star and suffer γ𝛾\gammaitalic_γγ𝛾\gammaitalic_γabsorption. Our analysis shows that this effect is entirely negligible for the jet, with substantial opacity emerging only beyond similar-to\sim100 GeV. For the counterjet, the onset of significant attenuation is, however, found to arise already at 5–6 GeV for Model 1, and 10–15 GeV for Model 2. Still, considering its subdominant contribution to the total emission and the increasingly large uncertainties in the data beyond these energies, we conclude that, in our specific case, this effect can be neglected also for the counterjet.

With either model, we conduct fitting of the phase-averaged spectrum and the light curve. We first obtain the spectrum of the emission as a function of the orbital phase for assumed preliminary values of θjsubscript𝜃j\theta_{\rm j}italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT, either φjsubscript𝜑j\varphi_{\rm j}italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT or φosubscript𝜑o\varphi_{\rm o}italic_φ start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT, βjsubscript𝛽j\beta_{\rm j}italic_β start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT, i𝑖iitalic_i, and hhitalic_h. With that, we fit the observed phase-averaged photon spectrum, obtaining Eesubscript𝐸eE_{\rm e}italic_E start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT, p𝑝pitalic_p, γminsubscript𝛾min\gamma_{\rm min}italic_γ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT and γmaxsubscript𝛾max\gamma_{\rm max}italic_γ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. With those, we fit in turn the observed folded light curve, which yields updated values of the parameters. Using those more accurate estimates, we fit again the phase-averaged spectrum, iterating in this way until convergence. The fitting is performed using the python LMFIT module (employing the default Levenberg-Marquardt minimization method).

4 Results for Model 1: Blackbody Compton

4.1 Fit results

Table 1: Fit results for the phase-averaged spectrum and the light curve for Model 1 without and with magnetic field
αj(°)subscript𝛼j°\alpha_{\rm j}(\arcdeg)italic_α start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ( ° ) Bmaxsubscript𝐵maxB_{\rm max}italic_B start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT(G) h/a𝑎h/aitalic_h / italic_a i(°)𝑖°i(\arcdeg)italic_i ( ° ) θj(°)subscript𝜃j°\theta_{\rm j}(\arcdeg)italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ( ° ) φj(°)subscript𝜑j°\varphi_{\rm j}(\arcdeg)italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ( ° ) βjsubscript𝛽j\beta_{\rm j}italic_β start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT χν2superscriptsubscript𝜒𝜈2\chi_{\nu}^{2}italic_χ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT p𝑝pitalic_p γmin(103)subscript𝛾minsuperscript103\gamma_{\rm min}\,(10^{3})italic_γ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) Ee(1037E_{\rm e}\,(10^{37}italic_E start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT ( 10 start_POSTSUPERSCRIPT 37 end_POSTSUPERSCRIPTerg) χν2superscriptsubscript𝜒𝜈2\chi_{\nu}^{2}italic_χ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT γbrsubscript𝛾br\gamma_{\rm br}italic_γ start_POSTSUBSCRIPT roman_br end_POSTSUBSCRIPT ij(°)subscript𝑖j°i_{\rm j}(\arcdeg)italic_i start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ( ° ) Δλ(°)Δ𝜆°\Delta\lambda(\arcdeg)roman_Δ italic_λ ( ° )
0 1.96±0.09plus-or-minus1.960.091.96\pm 0.091.96 ± 0.09 28±6plus-or-minus28628\pm 628 ± 6 32±7plus-or-minus32732\pm 732 ± 7 188±2plus-or-minus1882188\pm 2188 ± 2 0.59±0.05plus-or-minus0.590.050.59\pm 0.050.59 ± 0.05 7.2/5 4.16±0.05plus-or-minus4.160.054.16\pm 0.054.16 ± 0.05 2.9±0.2plus-or-minus2.90.22.9\pm 0.22.9 ± 0.2 18±1plus-or-minus18118\pm 118 ± 1 5.1/7 426 53+12superscriptsubscript53125_{-3}^{+12}5 start_POSTSUBSCRIPT - 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 12 end_POSTSUPERSCRIPT 22835+118superscriptsubscript22835118228_{-35}^{+118}228 start_POSTSUBSCRIPT - 35 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 118 end_POSTSUPERSCRIPT
5 34 2.3±0.6plus-or-minus2.30.62.3\pm 0.62.3 ± 0.6 33±7plus-or-minus33733\pm 733 ± 7 35±8plus-or-minus35835\pm 835 ± 8 188±3plus-or-minus1883188\pm 3188 ± 3 0.7±0.2plus-or-minus0.70.20.7\pm 0.20.7 ± 0.2 7.3/5 4.8±0.1plus-or-minus4.80.14.8\pm 0.14.8 ± 0.1 3.4±0.4plus-or-minus3.40.43.4\pm 0.43.4 ± 0.4 14±2plus-or-minus14214\pm 214 ± 2 7.5/7 591 52+12superscriptsubscript52125_{-2}^{+12}5 start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 12 end_POSTSUPERSCRIPT 26167+88superscriptsubscript2616788261_{-67}^{+88}261 start_POSTSUBSCRIPT - 67 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 88 end_POSTSUPERSCRIPT

\begin{overpic}[height=138.76157pt]{fit_SED_CygX3_Stable_Jet_zeroB.jpg} \put(0.08,0.285){\large(a)} \end{overpic}
\begin{overpic}[height=138.76157pt]{fit_LC_CygX3_Stable_Jet_zeroB.jpg} \put(0.41,0.28){\large(b)} \end{overpic}
\begin{overpic}[height=138.76157pt]{fit_SED_CygX3_Stable_Jet_maxB_34G_5deg.jpg% } \put(0.08,0.285){\large(c)} \end{overpic}
\begin{overpic}[height=138.76157pt]{fit_LC_CygX3_Stable_Jet_maxB_34G_5deg.jpg} \put(0.41,0.28){\large(d)} \end{overpic}
Figure 4: Fit results for Model 1 without and with magnetic field. Black error bars show the LAT data, the purple SED data point (upper limit) shows the constraint on the X-ray flux at 100 keV of 1.6×1010absent1.6superscript1010\approx 1.6\times 10^{-10}≈ 1.6 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT erg cm-2 s-1, the dark blue solid curves show the best-fit models, and the dashed histograms show the model predictions for the binned fluxes. Cyan dashed and the grey dash-dotted lines display the contributions from the jet and counterjet, respectively. (a) The phase-averaged spectrum for B=0𝐵0B=0italic_B = 0. (b) The 0.1–100 GeV light curve folded and averaged over the orbital phase with respect to the superior conjunction, φ𝜑\varphiitalic_φ. (c) Similar to (a), but for the upper limit value of the magnetic field Bmax=34subscript𝐵max34B_{\rm max}=34italic_B start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 34 G. (d) The corresponding light curve. In (c) and (d), the light-green and red dotted curves show the contributions from the Compton scattering of blackbody photons and the SSC, respectively.

We assume R=1011subscript𝑅superscript1011R_{*}=10^{11}italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT cm and T=105subscript𝑇superscript105T_{*}=10^{5}italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT K (Koljonen & Maccarone, 2017). Notably, that study also reveals that the photosphere extends significantly beyond the WR star, implying that the γ𝛾\gammaitalic_γ-ray emitter might reside within an optically thick wind. This would then influence the blackbody Compton emission, as well as the γ𝛾\gammaitalic_γγ𝛾\gammaitalic_γ absorption, adding complexity to our modeling approach. To address this, we make a simplifying assumption that the jet partially clears its surrounding environment, pushing the optically thick wind to further distances. Thus, we neglect this complication, which was also done in the previous studies (Dubus et al. 2010, Z18). We first include only Compton scattering of blackbody photons. Our results are shown in Figures 4(a,b) and Table 1. Confirming the previous results, e.g., Z18, we find the steady state electron distribution to be steep, with p4.2±0.1𝑝plus-or-minus4.20.1p\approx 4.2\pm 0.1italic_p ≈ 4.2 ± 0.1, with a large low-energy cutoff, γmin(2.9±0.2)×103subscript𝛾minplus-or-minus2.90.2superscript103\gamma_{\rm min}\approx(2.9\pm 0.2)\times 10^{3}italic_γ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ≈ ( 2.9 ± 0.2 ) × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, and the emission height of h(2±0.1)aplus-or-minus20.1𝑎h\approx(2\pm 0.1)aitalic_h ≈ ( 2 ± 0.1 ) italic_a. We find that the high-energy cutoff is not required, and set it to a high value of γmax=106subscript𝛾maxsuperscript106\gamma_{\rm max}=10^{6}italic_γ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT. This implies that the decline due to KN effects alone is sufficient to explain the steepening at high energies in the observed spectrum. We show the folded light curve fitted by our model vs. φ𝜑\varphiitalic_φ, i.e., the phase with respect to the actual superior conjunction, assuming the best fit offset of A22. We see that the peak of the modulation is still significantly before the superior conjunction, implying that the jet is inclined, see Table 1. The value of the φj>180°subscript𝜑j180°\varphi_{\rm j}>180\arcdegitalic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT > 180 ° means that the jet is slightly bent in the direction of the rotation around the superior conjunction, which causes the peak of the scattered flux to occur before it. The jet viewing angle is low, ij5°3+12subscript𝑖j5subscriptsuperscript°123i_{\rm j}\approx 5\arcdeg^{+12}_{-3}italic_i start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ≈ 5 ° start_POSTSUPERSCRIPT + 12 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 3 end_POSTSUBSCRIPT, which is in agreement with radio constraints (Mioduszewski et al., 2001; Miller-Jones et al., 2004). The projection of the orbital axis on the sky shows a large offset with respect to that of the jet, Δλ228°35+118Δ𝜆228subscriptsuperscript°11835\Delta\lambda\approx 228\arcdeg^{+118}_{-35}roman_Δ italic_λ ≈ 228 ° start_POSTSUPERSCRIPT + 118 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 35 end_POSTSUBSCRIPT.

An earlier study by Zdziarski et al. (2012c) found the magnetic field strength limited to B100less-than-or-similar-to𝐵100B\lesssim 100italic_B ≲ 100 G in the γ𝛾\gammaitalic_γ-ray emitting region of the jet, obtained relying only on the spectral measurements. Here, we impose stringent constraints on the magnetic field strength employing the current γ𝛾\gammaitalic_γ-ray data with the spectrum and the light curve, an upper limit for the X-ray flux at 100 keV, and the millimeter measurements.

We have four independent constraints from the data that allow us to infer the maximum allowed magnetic field strength. The first follows from the γ𝛾\gammaitalic_γ-ray modulation light curve, for which the SSC flux cannot exceed its minimum value. Then, the SSC contribution cannot overproduce the emission at 0.1less-than-or-similar-toabsent0.1\lesssim 0.1≲ 0.1 GeV. Depending on the magnetic field, this low-energy emission can become quite strong given the presence of cooled low-energy electrons below γminsubscript𝛾min\gamma_{\rm min}italic_γ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT (see Section 3.2). The remaining two constraints are described in Section 3.4, and are given by the observed 100 keV and 225 GHz fluxes, which cannot be exceeded by the model spectrum.

Following this approach, we derive an upper limit for B𝐵Bitalic_B using the jet opening angle αj=5°subscript𝛼j5°\alpha_{\rm j}=5\arcdegitalic_α start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT = 5 °, a value found on the scale of 1015similar-toabsentsuperscript1015\sim\!10^{15}∼ 10 start_POSTSUPERSCRIPT 15 end_POSTSUPERSCRIPT cm (5.0±0.5°plus-or-minus5.00.5°5.0\pm 0.5\arcdeg5.0 ± 0.5 °; Miller-Jones et al. 2006). Specifically, we obtain Bmax34subscript𝐵max34B_{\rm max}\approx 34italic_B start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ≈ 34 G. This magnetic field strength yields an SSC contribution saturating the minimum of the light curve, while we find that the X-ray and synchrotron emissions remain significantly below the observational constraints. Figures 4(c,d) illustrate the fits to the γ𝛾\gammaitalic_γ-ray data. The resulting fit parameters are given in Table 1, and are similar to those obtained assuming zero magnetic field.

The minimum length of the emitting region can be set by hcooltcool(γmin)βjcsimilar-tosubscriptcoolsubscript𝑡coolsubscript𝛾minsubscript𝛽j𝑐h_{\rm cool}\sim t_{\rm cool}(\gamma_{\rm min})\beta_{\rm j}citalic_h start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT ∼ italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT ( italic_γ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ) italic_β start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT italic_c, where the cooling time, tcool(γ)γ/|γ˙cool|subscript𝑡cool𝛾𝛾subscript˙𝛾coolt_{\rm cool}(\gamma)\equiv\gamma/|\dot{\gamma}_{\rm cool}|italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT ( italic_γ ) ≡ italic_γ / | over˙ start_ARG italic_γ end_ARG start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT |, depends on the total cooling rate, γ˙coolsubscript˙𝛾cool\dot{\gamma}_{\rm cool}over˙ start_ARG italic_γ end_ARG start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT. We calculate γ˙coolsubscript˙𝛾cool\dot{\gamma}_{\rm cool}over˙ start_ARG italic_γ end_ARG start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT due to the stellar Compton process, γ˙cool,SCsubscript˙𝛾coolSC\dot{\gamma}_{\rm cool,SC}over˙ start_ARG italic_γ end_ARG start_POSTSUBSCRIPT roman_cool , roman_SC end_POSTSUBSCRIPT, using eq. 16 in Zdziarski et al. (2014), while the cooling rate due to the SSC (for which the KN effects are negligible) and the synchrotron processes are evaluated as |γ˙cool,SSC/syn|=(4σTγ2Usyn/B)/(3mec)subscript˙𝛾coolSSCsyn4subscript𝜎Tsuperscript𝛾2subscript𝑈syn𝐵3subscript𝑚e𝑐|\dot{\gamma}_{\rm cool,SSC/syn}|=(4\sigma_{\rm T}\gamma^{2}U_{\mathrm{syn}/B}% )/(3m_{\rm e}c)| over˙ start_ARG italic_γ end_ARG start_POSTSUBSCRIPT roman_cool , roman_SSC / roman_syn end_POSTSUBSCRIPT | = ( 4 italic_σ start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT italic_γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_U start_POSTSUBSCRIPT roman_syn / italic_B end_POSTSUBSCRIPT ) / ( 3 italic_m start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT italic_c ), where Usyn/Bsubscript𝑈syn𝐵U_{\mathrm{syn}/B}italic_U start_POSTSUBSCRIPT roman_syn / italic_B end_POSTSUBSCRIPT is the energy density of the target synchrotron emission and of the magnetic field, respectively, and mesubscript𝑚em_{\rm e}italic_m start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT is the electron mass. For the model with the maximum magnetic field we obtain tcool(γmin)7.7subscript𝑡coolsubscript𝛾min7.7t_{\rm cool}(\gamma_{\rm min})\approx 7.7italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT ( italic_γ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ) ≈ 7.7 s. This implies hcool1.6×1011cm0.25h<Δh=h/3subscriptcool1.6superscript1011cm0.25Δ3h_{\rm cool}\approx 1.6\times 10^{11}\,{\rm cm}\approx 0.25h<\Delta h=h/3italic_h start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT ≈ 1.6 × 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT roman_cm ≈ 0.25 italic_h < roman_Δ italic_h = italic_h / 3, as required (since the dissipation region is likely to be extended). Next, we evaluate the angle-integrated jet bolometric luminosity for the same case, obtaining Lj2.4×1037subscript𝐿j2.4superscript1037L_{\rm j}\approx 2.4\times 10^{37}italic_L start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ≈ 2.4 × 10 start_POSTSUPERSCRIPT 37 end_POSTSUPERSCRIPT erg s-1. The ratio Ee/Ljsubscript𝐸esubscript𝐿jE_{\rm e}/L_{\rm j}italic_E start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT gives the average energy loss time scale of all electrons. We find that this time scale is similar to the cooling one at γminsubscript𝛾min\gamma_{\rm min}italic_γ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT.

4.2 Caveats for Model 1

Refer to caption
Figure 5: Comparison of the orbital modulation profiles during different time intervals. We see the profiles remain approximately constant during similar-to\sim5000 days of the LAT monitoring. Here, the horizontal axis gives the phase, ϕitalic-ϕ\phiitalic_ϕ, according to the ephemeris of Antokhin & Cherepashchuk (2019).
Refer to caption
Figure 6: The predicted angular difference between the projections on the sky of the jet and the orbital axis, ΔλΔ𝜆\Delta\lambdaroman_Δ italic_λ, see Equations (6) and (12), for the case of a jet aligned with the spin axis of the compact object but misaligned with respect to the orbital axis. The jet azimuthal angle, φjsubscript𝜑j\varphi_{\rm j}italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT, is predicted to change by 360°360°360\arcdeg360 ° during the precession period of 50 yr, which is clearly not seen. We assumed the best-fit values of θjsubscript𝜃j\theta_{\rm j}italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT, ϕjsubscriptitalic-ϕj\phi_{\rm j}italic_ϕ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT and i𝑖iitalic_i for the case with the maximum magnetic fieldBmax=34subscript𝐵max34B_{\rm max}=34italic_B start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 34 G. Currently, that model gives φj188°subscript𝜑j188°\varphi_{\rm j}\approx 188\arcdegitalic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ≈ 188 °, Δλ261°Δ𝜆261°\Delta\lambda\approx 261\arcdegroman_Δ italic_λ ≈ 261 °, see Table 1.

If the jet axis is linked to the rotation axis of the compact object, prograde geodetic precession is predicted. The De Sitter precession period for either a BH or a neutron star is given by (Barker & O’Connell, 1975; Apostolatos et al., 1994),

Pprec=c2(M+Mc)4/3P5/3(2πG)2/3(2+3M/2Mc)MMc,subscript𝑃precsuperscript𝑐2superscriptsubscript𝑀subscript𝑀c43superscript𝑃53superscript2𝜋𝐺2323subscript𝑀2subscript𝑀csubscript𝑀subscript𝑀cP_{\rm prec}=\frac{c^{2}(M_{*}+M_{\rm c})^{4/3}P^{5/3}}{(2\pi G)^{2/3}(2+3M_{*% }/2M_{\rm c})M_{*}M_{\rm c}},italic_P start_POSTSUBSCRIPT roman_prec end_POSTSUBSCRIPT = divide start_ARG italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT + italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 4 / 3 end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT 5 / 3 end_POSTSUPERSCRIPT end_ARG start_ARG ( 2 italic_π italic_G ) start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT ( 2 + 3 italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT / 2 italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ) italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT end_ARG , (12)

where the Kepler law was used, Mcsubscript𝑀cM_{\rm c}italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT includes the contribution from the rotational energy, and G𝐺Gitalic_G is the gravitational constant. For the best-fit masses of A22, Pprec50yr18000subscript𝑃prec50yr18000P_{\rm prec}\approx 50\,{\rm yr}\approx 18000italic_P start_POSTSUBSCRIPT roman_prec end_POSTSUBSCRIPT ≈ 50 roman_yr ≈ 18000 d, relatively weakly depending on them. During Pprecsubscript𝑃precP_{\rm prec}italic_P start_POSTSUBSCRIPT roman_prec end_POSTSUBSCRIPT, the jet azimuth, φjsubscript𝜑j\varphi_{\rm j}italic_φ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT, changes by 360°360°360\arcdeg360 °. The observations of γ𝛾\gammaitalic_γ-rays by the LAT covered already 5000absent5000\approx 5000≈ 5000 d (see Figure 1), so we would expect substantial changes in the modulation profile. However, we see no evidence for such changes, as shown in Figure 5.

Another argument against the presence of the precession is from the beat between the orbital and precession periods. The two angular frequencies add, leading to a shift of the frequency of the observed orbital modulation of the jet γ𝛾\gammaitalic_γ-ray emission. This shift is \approx0.2 s, while the γ𝛾\gammaitalic_γ-ray period measured by us equals the orbital one with the accuracy of <0.02absent0.02<0.02< 0.02 s, see Section 2.

A consequence of the geodetic precession is also a full rotation of the jet position angle with Pprecsubscript𝑃precP_{\rm prec}italic_P start_POSTSUBSCRIPT roman_prec end_POSTSUBSCRIPT. The predicted values of ΔλΔ𝜆\Delta\lambdaroman_Δ italic_λ are shown in Figure 6. However, we see no evidence for such changes in radio observations from 1985 to 2016 (Molnar et al., 1988; Schalinski et al., 1995; Mioduszewski et al., 2001; Miller-Jones et al., 2004; Egron et al., 2017). The position angle of the approaching jet remains relatively constant at close to 180°180°180\arcdeg180 ° with respect to the north.

Furthermore, this model predicts no Doppler orbital modulation of the synchrotron emission, since it is coming from a jet with a fixed orientation. This may disagree with observations if the preliminary results showing strong orbital modulation at 225 GHz similar to those of the γ𝛾\gammaitalic_γ-rays (McCollough, 2023) are confirmed.

The jet could possibly be misaligned due to some asymmetry in the binary, e.g., the presence of a bow shock (A22), instead of being in the direction of the (misaligned) BH spin. However, strong X-ray linear polarization has been discovered from this system (Veledina et al., 2024a, b). The high polarization degree and the polarization angle perpendicular to the position angle of the radio jet demonstrate we do not see the primary X-ray source directly but instead we see only the X-rays scattered in the inner an accretion funnel (expected for super-critical accretion). This also shows that the funnel, located close to the BH, is aligned with the jet. Thus, a constraint on this origin of the misalignment is that it has to already start on that scale.

5 Results for Model 2: orbital precession

5.1 Fit results

Table 2: Fit results for the phase-averaged spectrum and the light curve for a bent and precessing jet (Model 2, pure SSC) for the maximum value of the magnetic field, B=100𝐵100B=100italic_B = 100 G.
B𝐵Bitalic_B(G) αj(°)subscript𝛼j°\alpha_{\rm j}(\arcdeg)italic_α start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ( ° ) h/a𝑎h/aitalic_h / italic_a i(°)𝑖°i(\arcdeg)italic_i ( ° ) θj(°)subscript𝜃j°\theta_{\rm j}(\arcdeg)italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ( ° ) φo(°)subscript𝜑o°\varphi_{\rm o}(\arcdeg)italic_φ start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT ( ° ) βjsubscript𝛽j\beta_{\rm j}italic_β start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT χν2superscriptsubscript𝜒𝜈2\chi_{\nu}^{2}italic_χ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT p𝑝pitalic_p γminsubscript𝛾min\gamma_{\rm min}italic_γ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT (×103absentsuperscript103\times 10^{3}× 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT) γmaxsubscript𝛾max\gamma_{\rm max}italic_γ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT Ee(1038E_{\rm e}\,(10^{38}italic_E start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT ( 10 start_POSTSUPERSCRIPT 38 end_POSTSUPERSCRIPTerg) χν2superscriptsubscript𝜒𝜈2\chi_{\nu}^{2}italic_χ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT γbrsubscript𝛾br\gamma_{\rm br}italic_γ start_POSTSUBSCRIPT roman_br end_POSTSUBSCRIPT FXsubscript𝐹XF_{\rm X}italic_F start_POSTSUBSCRIPT roman_X end_POSTSUBSCRIPT βeqsubscript𝛽eq\beta_{\rm eq}italic_β start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT
100 5(f) 11(f) 30(f) 41±2plus-or-minus41241\pm 241 ± 2 142±4plus-or-minus1424142\pm 4142 ± 4 0.46±0.02plus-or-minus0.460.020.46\pm 0.020.46 ± 0.02 10.3/7 3.97±0.06plus-or-minus3.970.063.97\pm 0.063.97 ± 0.06 1.1±0.2plus-or-minus1.10.21.1\pm 0.21.1 ± 0.2 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT(f) 24±3plus-or-minus24324\pm 324 ± 3 8.1/7 95 2.8 30.7

\begin{overpic}[height=134.42113pt]{fit_SED_CygX3_JetBend_pure_SSC_1Ledd.jpg} \put(0.08,0.275){\large(a)} \end{overpic}
\begin{overpic}[height=134.42113pt]{fit_LC_CygX3_JetBend_pure_SSC_1Ledd.jpg} \put(0.2,0.27){\large(b)} \end{overpic}
\begin{overpic}[height=130.08731pt]{Synchr_SED_CygX3_JetBend_pure_SSC_1Ledd.% jpg} \put(0.08,0.264){\large(c)} \end{overpic}
\begin{overpic}[height=130.08731pt]{fit_SED_CygX3_JetBend_TWO_HALVES_pure_SSC_% 1Ledd.jpg} \put(0.08,0.264){\large(d)} \end{overpic}
Figure 7: Best-fit solutions for the case of the jet precessing at the orbital period, for the derived upper limit for the magnetic field, Bmax=100subscript𝐵max100B_{\rm max}=100italic_B start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 100 G. (a) The phase-averaged spectrum and (b) the 0.1–100 GeV light curve. The blue solid curves represent the best-fit model, the dashed histograms show the model predictions for the binned fluxes, and the jet and counterjet contributions are depicted with the dashed cyan and dash-dotted grey curves, respectively. (c) The predicted contribution from the synchrotron emission along with the average flux from the SMA, shown by the black triangle. The maximum magnetic field corresponds to the model 225 GHz flux equal to the observed one. The jet and counterjet contributions are shown with dashed turquoise and dash-dotted green curves, respectively. (d) Comparison of the Fermi-LAT data and the phase-averaged models for the data split into two phase parts corresponding to the flux higher and lower than 6×10106superscript10106\times 10^{-10}6 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT erg cm-2 s-1 level (shown in blue and green, respectively).

In this model, the synchrotron and SSC emissions are orbitally modulated due to the varying Doppler factor associated with the variable viewing angle of the jet. The emission peaks at the lowest jet viewing angle, which occurs at φ=π+φo𝜑𝜋subscript𝜑o\varphi=\pi+\varphi_{\rm o}italic_φ = italic_π + italic_φ start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT (at the inferior conjunction for φo=0subscript𝜑o0\varphi_{\rm o}=0italic_φ start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT = 0). Since the Doppler boosting is independent of the emission height, the fitted SSC spectrum and modulation profile do not depend on hhitalic_h. However, the synchrotron flux does depend on it, see Equation (7). Since both the SSC and synchrotron are similarly modulated due to this process, this model could explain the preliminary report of a 225 GHz orbital modulation (McCollough, 2023).

When permitting all parameters in the light curve fit to vary independently, we observe significant uncertainties in their inferred values, indicating the presence of degeneracies among various parameters. Specifically, the same jet viewing angle ijsubscript𝑖ji_{\rm j}italic_i start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT and hence the same Doppler factor modulation profile can be achieved with different combinations of θjsubscript𝜃j\theta_{\rm j}italic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT and i𝑖iitalic_i, implying a correlation between these parameters. To address this, we opt to fix i=30°𝑖30°i=30\arcdegitalic_i = 30 ° (A22).

As previously, we use αj=5°subscript𝛼j5°\alpha_{\rm j}=5\arcdegitalic_α start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT = 5 ° (Miller-Jones et al., 2006). We search for solutions that also reproduce the average flux measured at 225 GHz (McCollough, 2023). We find that configurations assuming a predicted position of the jet–wind recollimation shock of hasimilar-to𝑎h\sim aitalic_h ∼ italic_a (e.g., Yoon et al. 2016) significantly underpredict the 225 GHz flux, regardless of the magnetic field strength. Consequently, we consider h/a1much-greater-than𝑎1h/a\gg 1italic_h / italic_a ≫ 1, which increases the synchrotron flux, see Equation (7). We find h=11a11𝑎h=11aitalic_h = 11 italic_a and B=100𝐵100B=100italic_B = 100 G as the (approximate) minimum distance of the emitting region and the maximum magnetic field strength, respectively, yielding the observed flux at 225 GHz. The fit parameters for this B𝐵Bitalic_B are given in Table 2 and the spectrum and the light curve are shown in Figures 7(a,b). The corresponding model prediction for the phase-averaged synchrotron emission is shown in Figure 7(c).

We then compare our solution to the dependence of the total magnetic field strength in the comoving frame on the distance, B(h)𝐵B(h)italic_B ( italic_h ), of the analytical jet model of Zdziarski et al. (2022). For it, we assume the jet power of Pj=9×1038subscript𝑃j9superscript1038P_{\rm j}=9\times 10^{38}italic_P start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT = 9 × 10 start_POSTSUPERSCRIPT 38 end_POSTSUPERSCRIPT erg s-1, the terminal magnetization parameter of σmin=0.008subscript𝜎min0.008\sigma_{\rm min}=0.008italic_σ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = 0.008, where σ(B2/4π)/(ρc2)𝜎superscript𝐵24𝜋𝜌superscript𝑐2\sigma\equiv(B^{2}/4\pi)/(\rho c^{2})italic_σ ≡ ( italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 4 italic_π ) / ( italic_ρ italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ), ρ𝜌\rhoitalic_ρ is the comoving mass density, and the spin parameter of a=1subscript𝑎1a_{*}=1italic_a start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 1 and other parameters defined in that paper of =0.50.5\ell=0.5roman_ℓ = 0.5, ar=1subscript𝑎𝑟1a_{r}=1italic_a start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = 1, aΓ=0.2subscript𝑎Γ0.2a_{\Gamma}=0.2italic_a start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT = 0.2, q2=1.9subscript𝑞21.9q_{2}=1.9italic_q start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1.9 where assumed. They yield the jet half-opening angle of αj5°subscript𝛼j5°\alpha_{\rm j}\approx 5\arcdegitalic_α start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ≈ 5 ° and B100𝐵100B\approx 100italic_B ≈ 100 G (dominated by the toroidal component) at h/a=11𝑎11h/a=11italic_h / italic_a = 11.

In addition, we generate model predictions for the average spectra corresponding to the Fermi-LAT high and low-flux states, delineated by the 0.1–100 GeV light curve flux being above and below the threshold of 6×10106superscript10106\times 10^{-10}6 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT erg cm-2 s-1, respectively. This threshold approximately corresponds to the average between the maximum and minimum fluxes observed in the light curve. This partitioning yields two distinct phase intervals: (1) 0φ/(2π)<0.080.7φ/(2π)<10𝜑2𝜋0.080.7𝜑2𝜋10\leq\varphi/(2\pi)<0.08\ \cup 0.7\leq\varphi/(2\pi)<10 ≤ italic_φ / ( 2 italic_π ) < 0.08 ∪ 0.7 ≤ italic_φ / ( 2 italic_π ) < 1 where the 0.1 - 100 GeV flux is above the separation threshold, and (2) 0.08φ/(2π)<0.70.08𝜑2𝜋0.70.08\leq\varphi/(2\pi)<0.70.08 ≤ italic_φ / ( 2 italic_π ) < 0.7 where the flux is below the threshold. The comparison of the two spectra with the model predictions is shown in Figure 7(d). We see no noticeable spectral changes between the two substates, which are also well described by our model.

Furthermore, we estimate upper limits on Rsubscript𝑅R_{*}italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT and Tsubscript𝑇T_{*}italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT by including the blackbody Compton component into the model. Based on observations and stellar modeling, Koljonen & Maccarone (2017) found T0.8subscript𝑇0.8T_{*}\approx 0.8italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ≈ 0.81.0×1051.0superscript1051.0\times 10^{5}1.0 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT K at the range of R1.5subscript𝑅1.5R_{*}\approx 1.5italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ≈ 1.51.0×10111.0superscript10111.0\times 10^{11}1.0 × 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT cm, corresponding to RT21.0×1021subscript𝑅superscriptsubscript𝑇21.0superscript1021R_{*}T_{*}^{2}\approx 1.0\times 10^{21}italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≈ 1.0 × 10 start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT cm K2. Interestingly, we find a non-monotonic behavior in the χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT of the light curve fit as a function of RT2subscript𝑅superscriptsubscript𝑇2R_{*}T_{*}^{2}italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT; it is initially decreasing, then increasing again. The best fit is achieved for a value of RT2=1.6×1021subscript𝑅superscriptsubscript𝑇21.6superscript1021R_{*}T_{*}^{2}=1.6\times 10^{21}italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1.6 × 10 start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT cm K2 with χν2=7.8/7superscriptsubscript𝜒𝜈27.87\chi_{\nu}^{2}=7.8/7italic_χ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 7.8 / 7 (Δχ2=2.5Δsuperscript𝜒22.5\Delta\chi^{2}=2.5roman_Δ italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 2.5 with respect to the pure SSC case). However, the χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT of the spectral fit increases monotonically with increasing RT2subscript𝑅superscriptsubscript𝑇2R_{*}T_{*}^{2}italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, with an increase by Δχ2=2.7Δsuperscript𝜒22.7\Delta\chi^{2}=2.7roman_Δ italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 2.7 as compared to the pure SSC fit being achieved coincidentally at the same value of RT2subscript𝑅superscriptsubscript𝑇2R_{*}T_{*}^{2}italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (χν2=10.8/7superscriptsubscript𝜒𝜈210.87\chi_{\nu}^{2}=10.8/7italic_χ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 10.8 / 7). Given the best fit of the light curve and a still satisfactory SED fit, this value of RT2subscript𝑅superscriptsubscript𝑇2R_{*}T_{*}^{2}italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT can be therefore considered as the most optimal one. The best-fit range for RT2subscript𝑅superscriptsubscript𝑇2R_{*}T_{*}^{2}italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for this most optimal solution found by imposing Δχ22.7Δsuperscript𝜒22.7\Delta\chi^{2}\leq 2.7roman_Δ italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≤ 2.7 for the spectral fit (as it degrades much quicker), is RT2=(0R_{*}T_{*}^{2}=(0italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ( 01.7)×10211.7)\times 10^{21}1.7 ) × 10 start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT cm K2. The higher hhitalic_h, the higher the allowed range of RT2subscript𝑅superscriptsubscript𝑇2R_{*}T_{*}^{2}italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. In addition, within this best-fit range, the model-predicted X-ray flux at 100 keV shows a slow decrease with an increasing RT2subscript𝑅superscriptsubscript𝑇2R_{*}T_{*}^{2}italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, thus having consistently lower values compared to the pure SSC case (due to adjusted Eesubscript𝐸eE_{\rm e}italic_E start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT in the fit). We find the X-ray flux at 100 keV for RT2=1.6×1021subscript𝑅superscriptsubscript𝑇21.6superscript1021R_{*}T_{*}^{2}=1.6\times 10^{21}italic_R start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1.6 × 10 start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT cm K2 of 1.4×1010absent1.4superscript1010\approx 1.4\times 10^{-10}≈ 1.4 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT erg cm-2 s-1, which is below the observational constraint.

5.2 Jet parameters

The model of Zdziarski et al. (2022) gives the magnetic flux threading the black hole as a function of Pjsubscript𝑃jP_{\rm j}italic_P start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT in their equation (17), which is based on numerical results of Tchekhovskoy et al. (2009). It yields ΦB5×1021subscriptΦ𝐵5superscript1021\Phi_{B}\approx 5\times 10^{21}roman_Φ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ≈ 5 × 10 start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT G cm2. On the other hand, since Cyg X-3 appears to be a super-Eddington accretor (Veledina et al., 2024a), M˙accr2×1019greater-than-or-equivalent-tosubscript˙𝑀accr2superscript1019\dot{M}_{\rm accr}\gtrsim 2\times 10^{19}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT roman_accr end_POSTSUBSCRIPT ≳ 2 × 10 start_POSTSUPERSCRIPT 19 end_POSTSUPERSCRIPT g s-1 (using the Eddington luminosity for He and the accretion efficiency of 0.1). The magnetic flux at the limit of the magnetically arrested disk (MAD; Bisnovatyi-Kogan & Ruzmaikin 1974; Narayan et al. 2003) is ΦMAD50(M˙accrc)1/2rgsubscriptΦMAD50superscriptsubscript˙𝑀accr𝑐12subscript𝑟g\Phi_{\rm MAD}\approx 50(\dot{M}_{\rm accr}c)^{1/2}r_{\rm g}roman_Φ start_POSTSUBSCRIPT roman_MAD end_POSTSUBSCRIPT ≈ 50 ( over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT roman_accr end_POSTSUBSCRIPT italic_c ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT (Tchekhovskoy et al., 2011), implying ΦMAD4×1022greater-than-or-equivalent-tosubscriptΦMAD4superscript1022\Phi_{\rm MAD}\gtrsim 4\times 10^{22}roman_Φ start_POSTSUBSCRIPT roman_MAD end_POSTSUBSCRIPT ≳ 4 × 10 start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT G cm2. Thus, the jet has the power at least an order of magnitude below the MAD limit.

We then find the electron cooling time at γminsubscript𝛾min\gamma_{\rm min}italic_γ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT of tcool(γmin)23subscript𝑡cool𝛾min23t_{\rm cool}(\gamma{\rm min})\approx 23italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT ( italic_γ roman_min ) ≈ 23 s, translating to the cooling length of 3.2×1011absent3.2superscript1011\approx 3.2\times 10^{11}≈ 3.2 × 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT cm 0.1h<Δh=h/3absent0.1Δ3\approx 0.1h<\Delta h=h/3≈ 0.1 italic_h < roman_Δ italic_h = italic_h / 3, consistent with our assumption of the value of ΔhΔ\Delta hroman_Δ italic_h. We find the total radiative power of Lj1.2×1038subscript𝐿j1.2superscript1038L_{\rm j}\approx 1.2\times 10^{38}italic_L start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ≈ 1.2 × 10 start_POSTSUPERSCRIPT 38 end_POSTSUPERSCRIPT erg s-1. The average energy loss time of all electrons, Ee/Ljsubscript𝐸esubscript𝐿jE_{\rm e}/L_{\rm j}italic_E start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT (Table 2) is then in good agreement with the cooling time scale.

Also, we calculate the Thomson optical depth, τ=neσTR𝜏subscript𝑛esubscript𝜎T𝑅\tau=n_{\rm e}\sigma_{\rm T}Ritalic_τ = italic_n start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT italic_R, where nesubscript𝑛en_{\rm e}italic_n start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT is the number density of electrons from γ=1𝛾1\gamma=1italic_γ = 1 to γmaxsubscript𝛾max\gamma_{\rm max}italic_γ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. We find a low value of τ3.7×105𝜏3.7superscript105\tau\approx 3.7\times 10^{-5}italic_τ ≈ 3.7 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT.

We then calculate the power in both the jet and counterjet. We use eqs. (19–21) in Zdziarski (2014). We assume a homogeneous emitting region (without clumping), pure electron-ion plasma (fully ionized He) without positrons, and a toroidal magnetic field. We first compute the jet power in electrons (excluding their rest energy). We obtain Pe1.2×1038subscript𝑃e1.2superscript1038P_{\rm e}\approx 1.2\times 10^{38}italic_P start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT ≈ 1.2 × 10 start_POSTSUPERSCRIPT 38 end_POSTSUPERSCRIPT erg s-1. Next, we evaluate the total jet power in (cold) iond, as Pi5.1×1038subscript𝑃i5.1superscript1038P_{\rm i}\approx 5.1\times 10^{38}italic_P start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT ≈ 5.1 × 10 start_POSTSUPERSCRIPT 38 end_POSTSUPERSCRIPT erg s-1. The magnetic power is much lower, PB3.8×1036subscript𝑃𝐵3.8superscript1036P_{B}\approx 3.8\times 10^{36}italic_P start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ≈ 3.8 × 10 start_POSTSUPERSCRIPT 36 end_POSTSUPERSCRIPT erg s-1. The total jet power is then 6.3×1038absent6.3superscript1038\approx 6.3\times 10^{38}≈ 6.3 × 10 start_POSTSUPERSCRIPT 38 end_POSTSUPERSCRIPT erg s-1, only slightly lower than that assumed earlier using the model of Zdziarski et al. (2022).

Next, we compute the equipartition parameter, characterizing the degree of equipartition between particles and magnetic field and defined in terms of the energy densities (excluding the rest energy of particles), βeq=ue/(B2/8π)subscript𝛽eqsubscript𝑢esuperscript𝐵28𝜋\beta_{\rm eq}=u_{\rm e}/(B^{2}/8\pi)italic_β start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT = italic_u start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT / ( italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 8 italic_π ), where ue=Ee/(πR2Δh)subscript𝑢esubscript𝐸e𝜋superscript𝑅2Δu_{\rm e}=E_{\rm e}/(\pi R^{2}\Delta h)italic_u start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT / ( italic_π italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ italic_h ) is the electron energy density. We find βeq31subscript𝛽eq31\beta_{\rm eq}\approx 31italic_β start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT ≈ 31. Similarly, the magnetization parameter (defined with respect to the rest mass density, see above) is very low, σ0.001similar-to𝜎0.001\sigma\sim 0.001italic_σ ∼ 0.001. This indicates the jet is strongly dominated by the kinetic energy of particles. However, this value of σ𝜎\sigmaitalic_σ is lower than that assumed in the model of Zdziarski et al. (2022), σmin=0.008subscript𝜎min0.008\sigma_{\rm min}=0.008italic_σ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = 0.008, which is due to the magnetic power in our fit lower than that following from that model.

In addition, we also compute the mass flow rate in ions in the jet (equation 17 in Zdziarski 2014), obtaining M˙j4.6×1018subscript˙𝑀j4.6superscript1018\dot{M}_{\rm j}\approx 4.6\times 10^{18}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ≈ 4.6 × 10 start_POSTSUPERSCRIPT 18 end_POSTSUPERSCRIPT g s-1. We compare this rate to the super-Eddington accretion rate mentioned earlier, M˙accr2×1019greater-than-or-equivalent-tosubscript˙𝑀accr2superscript1019\dot{M}_{\rm accr}\gtrsim 2\times 10^{19}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT roman_accr end_POSTSUBSCRIPT ≳ 2 × 10 start_POSTSUPERSCRIPT 19 end_POSTSUPERSCRIPT g s-1, and the derived jet powers to M˙accrc21.8×1040subscript˙𝑀accrsuperscript𝑐21.8superscript1040\dot{M}_{\rm accr}c^{2}\approx 1.8\times 10^{40}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT roman_accr end_POSTSUBSCRIPT italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≈ 1.8 × 10 start_POSTSUPERSCRIPT 40 end_POSTSUPERSCRIPT erg s-1. We find that the jet is powered by a rather sizeable fraction (23absent23\approx 23≈ 23%) of accretted matter, while only a very small fraction (3absent3\approx 3≈ 3%) of the energy released from accretion (M˙accc2subscript˙𝑀accsuperscript𝑐2\dot{M}_{\rm acc}c^{2}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT roman_acc end_POSTSUBSCRIPT italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT) is supplied to the particle energy, and an almost negligible fraction is converted into the jet magnetic energy.

Next, we calculate the jet bending angle, ΦΦ\Phiroman_Φ, owing to the wind impact. In the limit of Φ1much-less-thanΦ1\Phi\ll 1roman_Φ ≪ 1 rad, we have

ΦαjM˙wvw(Γj1)c4πβjΓjPj,Φsubscript𝛼jsubscript˙𝑀wsubscript𝑣wsubscriptΓj1𝑐4𝜋subscript𝛽jsubscriptΓjsubscript𝑃j\Phi\approx\frac{\alpha_{\rm j}\dot{M}_{\rm w}v_{\rm w}(\Gamma_{\rm j}-1)c}{4% \pi\beta_{\rm j}\Gamma_{\rm j}P_{\rm j}},roman_Φ ≈ divide start_ARG italic_α start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT ( roman_Γ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT - 1 ) italic_c end_ARG start_ARG 4 italic_π italic_β start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT end_ARG , (13)

as given by equation (7) of Bosch-Ramon & Barkov (2016) (see their equation 8 for an expression valid at any ΦΦ\Phiroman_Φ). Here M˙wsubscript˙𝑀w\dot{M}_{\rm w}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT and vvsubscript𝑣vv_{\rm v}italic_v start_POSTSUBSCRIPT roman_v end_POSTSUBSCRIPT are the wind mass loss rate and velocity, respectively, and ΓjsubscriptΓj\Gamma_{\rm j}roman_Γ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT is the jet Lorentz factor. We assume M˙w105Msubscript˙𝑀wsuperscript105subscript𝑀\dot{M}_{\rm w}\approx 10^{-5}{M}_{\sun}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ☉ end_POSTSUBSCRIPT yr-1 (A22) and vw1.5×108subscript𝑣w1.5superscript108v_{\rm w}\approx 1.5\times 10^{8}italic_v start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT ≈ 1.5 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT cm s-1 (van Kerkwijk et al., 1996). However, since the jet likely undergoes a decelaration over the γ𝛾\gammaitalic_γ-ray emission region and possibly upstream of it, we use Γj=2subscriptΓj2\Gamma_{\rm j}=2roman_Γ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT = 2. This yields a low angle of Φ1°Φ1°\Phi\approx 1\arcdegroman_Φ ≈ 1 °.

5.3 Caveats for Model 2

The main problem for this model is the value of the jet wind-induced bending angle, Φ1°Φ1°\Phi\approx 1\arcdegroman_Φ ≈ 1 °, being much lower than the fitted value of θj41°subscript𝜃j41°\theta_{\rm j}\approx 41\arcdegitalic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ≈ 41 °. If the (unknown) value of ΓjsubscriptΓj\Gamma_{\rm j}roman_Γ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT is larger, we can increase ΦΦ\Phiroman_Φ, but only to 1.8°absent1.8°\approx\!1.8\arcdeg≈ 1.8 ° for Γj1much-greater-thansubscriptΓj1\Gamma_{\rm j}\gg 1roman_Γ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ≫ 1. Then, the jet half-opening angle in the bending region can be larger, as seen in the hard state (Yang et al., 2023). Then, the jet power could be lower depending on the poorly known jet parameters. In particular, if the jet composition consists mostly of e± pairs, Pisubscript𝑃iP_{\rm i}italic_P start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT can be negligible compared to Pesubscript𝑃eP_{\rm e}italic_P start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT, which would then increase ΦΦ\Phiroman_Φ by a factor of \approx4. Furthermore, values of M˙wsubscript˙𝑀w\dot{M}_{\rm w}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT significantly higher than 105Msuperscript105subscript𝑀10^{-5}{M}_{\sun}10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ☉ end_POSTSUBSCRIPT yr-1 were obtained by van Kerkwijk (1993) and Ogley et al. (2001). In principle, Φ40°Φ40°\Phi\approx 40\arcdegroman_Φ ≈ 40 ° could be reached if αjsubscript𝛼j\alpha_{\rm j}italic_α start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT, M˙wsubscript˙𝑀w\dot{M}_{\rm w}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT roman_w end_POSTSUBSCRIPT and ΓjsubscriptΓj\Gamma_{\rm j}roman_Γ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT are higher than the adopted values, and Pjsubscript𝑃jP_{\rm j}italic_P start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT is lower.

In addition, if Φ<αj=5°Φsubscript𝛼j5°\Phi<\alpha_{\rm j}=5\arcdegroman_Φ < italic_α start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT = 5 °, there would be only a minor modification of the jet propagation due to the thrust of the stellar wind, which would only induce a weak recollimation shock or a sound wave in the jet (Bosch-Ramon & Barkov, 2016). However, we argued above that ΦΦ\Phiroman_Φ could be larger, and it appears it could be larger than αjsubscript𝛼j\alpha_{\rm j}italic_α start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT, if, in particular, the jet consists of a significant number of e± pairs.

A major consideration for this model is the required large phase offset, φo142°±4°subscript𝜑oplus-or-minus142°4°\varphi_{\rm o}\approx 142\arcdeg\pm 4\arcdegitalic_φ start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT ≈ 142 ° ± 4 °. For our assumed masses, the BH velocity around the center of mass is vBH6×107subscript𝑣BH6superscript107v_{\rm BH}\approx 6\times 10^{7}italic_v start_POSTSUBSCRIPT roman_BH end_POSTSUBSCRIPT ≈ 6 × 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT cm s-1. The average vertical velocity of the jet between the origin and the height of the emission required (via the binary rotation) to reach φosubscript𝜑o\varphi_{\rm o}italic_φ start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT is then h/(Pφo/2π)𝑃subscript𝜑o2𝜋h/(P\varphi_{\rm o}/2\pi)italic_h / ( italic_P italic_φ start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT / 2 italic_π ). Obviously, it is rather low, e.g., 4.3×108absent4.3superscript108\approx 4.3\times 10^{8}≈ 4.3 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT cm s10.01c{}^{-1}\sim 0.01cstart_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT ∼ 0.01 italic_c for h=11a11𝑎h=11aitalic_h = 11 italic_a. This may still possible be if the jet is composed of a slow sheath, formed by a disk outflow, and a fast spine powered by the spin of the rotating BH.

Another issue is the large fitted polar angle of the orbital precession, θj=41°±2°subscript𝜃jplus-or-minus41°2°\theta_{\rm j}=41\arcdeg\pm 2\arcdegitalic_θ start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT = 41 ° ± 2 ° in the γ𝛾\gammaitalic_γ-ray emission region. The jet needs to recollimate above that region in order to satisfy the observational constraint on the semi-opening angle, 5°absent5°\approx\!5\arcdeg≈ 5 ° (Miller-Jones et al., 2004, 2006). This could be achieved by reconfinement, for instance one induced by the stellar wind (e.g., Yoon & Heinz 2015; Yoon et al. 2016), and the symmetry and pressure gradients of the environment at the scales of interest. In fact, the large scale jet half-opening angle can become moderate once its interaction with the wind is not significant any more (see, e.g., equation 19 in Bosch-Ramon & Barkov 2016). However, due to many scales involved, estimating the jet large-scale properties would require numerical simulations, which are out of scope of this work.

An additional complexity can appear if the magnetic field symmetry axis is inclined with respect to the BH spin axis, e.g., due to interaction with the stellar wind, see, e.g., James et al. (2024). Studying such effects is outside the scope of this paper.

6 Discussion

After exhaustive research, we are still unable to choose between the two studied models of the observed γ𝛾\gammaitalic_γ-ray modulation on the orbital period. The blackbody Compton model gives better fits, but it predicts general-relativistic jet precession on the period of 50 years, which is ruled out observationally, see Section 4.2. The model with with a jet precessing on the orbital period due to bending by the thrust of the stellar wind fits the data satisfactorily, but the wind is too weak and/or the jet too strong to account for the fitted large bending angle, see Section 5.3. This phenomenon has been theoretically predicted (e.g., Yoon & Heinz 2015; Yoon et al. 2016; Bosch-Ramon & Barkov 2016; Barkov & Bosch-Ramon 2022), but not yet confirmed observationally in any other source.

An important test possibly allowing us to distinguish between the two models would be either the confirmation or disproval of the presence of orbital modulation with the folded light curve similar to that seen in γ𝛾\gammaitalic_γ-rays at the mm-wavelength modulation. Such modulation was claimed in a preliminary report of (McCollough, 2023). If present, it clearly cannot be explained by the Compton anisotropy effect present in our Model 1, but it can be due to the Doppler boosting present in our Model 2. However, we note that modulation at mm wavelengths could also be due to free-free absorption in the stellar wind, and the orbital phase of the maximum absorption can be shifted due to the jet misalignment with respect to the orbital axis.

We consider then the origin of the structures observed by Mioduszewski et al. (2001) and Miller-Jones et al. (2004), which, if caused by precession, would have much longer periods than the binary one, greater-than-or-equivalent-to\gtrsim60 d and 5.5±0.5plus-or-minus5.50.55.5\pm 0.55.5 ± 0.5 d, respectively. We speculate that these long periods are the result of perturbations in the jet on large scales triggered by instability growth as the jet is being perturbed at its base by the effect of the wind and the orbital motion. This would give rise to a distinct pattern not synchronized with the orbital cycle but rather representing a subharmonic of it. Specifically, if the jet bending is excited with a wavelength N𝑁Nitalic_N times the orbital period multiplied by the jet velocity, the resulting spatial scale would be 103Nasimilar-toabsentsuperscript103𝑁𝑎\sim 10^{3}Na∼ 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_N italic_a.

A consequence of Model 2 is that the jet (on average) and the binary should have the same inclination. The binary inclination determined by A22 is i=29.5±1.2𝑖plus-or-minus29degree51degree2i=29\fdg 5\pm 1\fdg 2italic_i = 29 start_ID start_POSTFIX SUPERSCRIPTOP . ∘ end_POSTFIX end_ID 5 ± 1 start_ID start_POSTFIX SUPERSCRIPTOP . ∘ end_POSTFIX end_ID 2. The jet inclination determined by Miller-Jones et al. (2004) based on the proper motion of the approaching and receding components assuming they are intrinsically symmetric (Mirabel & Rodríguez, 1994) and D=9𝐷9D=9italic_D = 9 kpc is ij26°±6°subscript𝑖jplus-or-minus26°6°i_{\rm j}\approx 26\arcdeg\pm 6\arcdegitalic_i start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ≈ 26 ° ± 6 °, i.e., consistent with the above value of i𝑖iitalic_i. On the other hand, the jet inclination based on fitting the precessing jet model of Hjellming & Johnston (1981) to the curvature observed in the VLBA images of Cyg X-3 is i10.4±3.8𝑖plus-or-minus10degree43degree8i\approx 10\fdg 4\pm 3\fdg 8italic_i ≈ 10 start_ID start_POSTFIX SUPERSCRIPTOP . ∘ end_POSTFIX end_ID 4 ± 3 start_ID start_POSTFIX SUPERSCRIPTOP . ∘ end_POSTFIX end_ID 8 (Miller-Jones et al., 2004) and i14°less-than-or-similar-to𝑖14°i\lesssim 14\arcdegitalic_i ≲ 14 ° (Mioduszewski et al., 2001). Those two estimates are incompatible with the inclination of A22, which would then favor Model 1.

In Model 2, we considered the γ𝛾\gammaitalic_γ-ray emission region at a relatively large distance from the origin, h10a10𝑎h\approx 10aitalic_h ≈ 10 italic_a, which allowed us to obtain the synchrotron emission strong enough to account for the observed average 225 GHz flux. This distance is much larger than that expected for a recollimation shock from the jet-wind interactions, which is at h(1h\sim(1italic_h ∼ ( 12)a2)a2 ) italic_a (Yoon et al., 2016; Bosch-Ramon & Barkov, 2016; Barkov & Bosch-Ramon, 2022). Then, the dissipation at h10asimilar-to10𝑎h\sim 10aitalic_h ∼ 10 italic_a would need to be explained by another process. Alternatively, we can relax the requirement of reproducing the 225 GHz flux by assuming that emission originates in a different part of the jet. Then, the SSC emission alone in that model is independent of hhitalic_h, and we could have h(1h\sim(1italic_h ∼ ( 12)a2)a2 ) italic_a.

7 Conclusions

Our main conclusions are as follows.

We have used a γ𝛾\gammaitalic_γ-ray bright data set from the LAT observations of Cyg X-3 greatly enlarged with respect to the previous study of Z18. We have found that the γ𝛾\gammaitalic_γ-rays are modulated at the period that is compatible with the orbital period based on the X-ray data at the accuracy of <0.02absent0.02<0.02< 0.02 s (accounting for the secular period increase).

We have studied possible models of the orbital modulation. We have significantly improved the model based on anisotropy of Compton scattering, by including KN effects and the presence of a magnetic field. The model fits well both the modulation profile and the average spectrum, and requires that the jet is misaligned with respect to the orbital axis by 35°similar-toabsent35°\sim\!35\arcdeg∼ 35 °. If the jet direction follows the BH spin axis, both will undergo the geodetic precession with a period of similar-to\sim50 yrs. Its presence is ruled out by both the γ𝛾\gammaitalic_γ-ray and radio data. If this is the correct model, the jet misalignment is not related to the BH spin axis.

Therefore, we have proposed an alternative model using the jet bending due to the thrust of the stellar wind from the donor. The jet is aligned with the orbital axis on average, but it is bent to outside of the orbit, causing its precession at the orbital period. The Doppler boosting of both the synchrotron and SSC emission is then orbitally modulated. The modulated SSC emission fits well the γ𝛾\gammaitalic_γ-ray data. The modulation of the SSC in this model does not depend on the distance of the emission region from the center. However, the synchrotron flux (at the SSC flux determined by the γ𝛾\gammaitalic_γ-ray data) depends on it. If we postulate that the observed average mm flux is reproduced by this model, this distance is about 10 orbital separations. The mm emission can still be from a different jet region, and then the γ𝛾\gammaitalic_γ-ray emission distance can be lower. However, a major problem for this model is that the theoretically predicted bending angle at plausible wind and jet parameters is much lower than that fitted to the data.

In both models, the jet velocity is low, βj0.5similar-tosubscript𝛽j0.5\beta_{\rm j}\sim 0.5italic_β start_POSTSUBSCRIPT roman_j end_POSTSUBSCRIPT ∼ 0.5, and the jet is weakly magnetized. The fitted field strengths are well below that predicted by models of the magnetically arrested disk and the magnetically launched jet.

Acknowledgements

We thank I. Antokhin, E. Egron, M. Liska and M. McCollough for discussions. We acknowledge support from the Polish National Science Center grants 2019/35/B/ST9/03944, 2023/48/Q/ST9/00138, and from the Copernicus Academy grant CBMK/01/24. The work of AD is supported by the South African Department of Science and Innovation and the National Research Foundation through the South African Gamma-Ray Astronomy Programme (SA-GAMMA).

References

  • Antokhin & Cherepashchuk (2019) Antokhin, I. I., & Cherepashchuk, A. M. 2019, ApJ, 871, 244, doi: 10.3847/1538-4357/aafb38
  • Antokhin et al. (2022) Antokhin, I. I., Cherepashchuk, A. M., Antokhina, E. A., & Tatarnikov, A. M. 2022, ApJ, 926, 123, doi: 10.3847/1538-4357/ac4047
  • Apostolatos et al. (1994) Apostolatos, T. A., Cutler, C., Sussman, G. J., & Thorne, K. S. 1994, Phys. Rev. D, 49, 6274, doi: 10.1103/PhysRevD.49.6274
  • Bardeen & Petterson (1975) Bardeen, J. M., & Petterson, J. A. 1975, ApJ, 195, L65, doi: 10.1086/181711
  • Barker & O’Connell (1975) Barker, B. M., & O’Connell, R. F. 1975, Phys. Rev. D, 12, 329, doi: 10.1103/PhysRevD.12.329
  • Barkov & Bosch-Ramon (2022) Barkov, M. V., & Bosch-Ramon, V. 2022, MNRAS, 510, 3479, doi: 10.1093/mnras/stab3609
  • Belczyński et al. (2013) Belczyński, K., Bulik, T., Mandel, I., et al. 2013, ApJ, 764, 96, doi: 10.1088/0004-637X/764/1/96
  • Bhargava et al. (2017) Bhargava, Y., Rao, A. R., Singh, K. P., et al. 2017, ApJ, 849, 141, doi: 10.3847/1538-4357/aa8ea4
  • Bisnovatyi-Kogan & Ruzmaikin (1974) Bisnovatyi-Kogan, G. S., & Ruzmaikin, A. A. 1974, Ap&SS, 28, 45, doi: 10.1007/BF00642237
  • Blandford & Payne (1982) Blandford, R. D., & Payne, D. G. 1982, MNRAS, 199, 883, doi: 10.1093/mnras/199.4.883
  • Blandford & Znajek (1977) Blandford, R. D., & Znajek, R. L. 1977, MNRAS, 179, 433, doi: 10.1093/mnras/179.3.433
  • Boettcher et al. (2012) Boettcher, M., Harris, D. E., & Krawczynski, H. 2012, Relativistic Jets from Active Galactic Nuclei (Wiley), doi: 10.1002/9783527641741
  • Bosch-Ramon & Barkov (2016) Bosch-Ramon, V., & Barkov, M. V. 2016, A&A, 590, A119, doi: 10.1051/0004-6361/201628564
  • Chiaberge & Ghisellini (1999) Chiaberge, M., & Ghisellini, G. 1999, MNRAS, 306, 551, doi: 10.1046/j.1365-8711.1999.02538.x
  • Dubus et al. (2010) Dubus, G., Cerutti, B., & Henri, G. 2010, MNRAS, 404, L55, doi: 10.1111/j.1745-3933.2010.00834.x
  • Egron et al. (2017) Egron, E., Pellizzoni, A., Giroletti, M., et al. 2017, MNRAS, 471, 2703, doi: 10.1093/mnras/stx1730
  • Fermi LAT Collaboration et al. (2009) Fermi LAT Collaboration, Abdo, A. A., Ackermann, M., et al. 2009, Science, 326, 1512, doi: 10.1126/science.1182174
  • Giacconi et al. (1967) Giacconi, R., Gorenstein, P., Gursky, H., & Waters, J. R. 1967, ApJ, 148, L119, doi: 10.1086/180028
  • Hjalmarsdotter et al. (2008) Hjalmarsdotter, L., Zdziarski, A. A., Larsson, S., et al. 2008, MNRAS, 384, 278, doi: 10.1111/j.1365-2966.2007.12688.x
  • Hjalmarsdotter et al. (2009) Hjalmarsdotter, L., Zdziarski, A. A., Szostek, A., & Hannikainen, D. C. 2009, MNRAS, 392, 251, doi: 10.1111/j.1365-2966.2008.14036.x
  • Hjellming & Johnston (1981) Hjellming, R. M., & Johnston, K. J. 1981, ApJ, 246, L141, doi: 10.1086/183571
  • James et al. (2024) James, B., Janiuk, A., & Karas, V. 2024, arXiv e-prints, arXiv:2403.14882, doi: 10.48550/arXiv.2403.14882
  • Katarzyński et al. (2001) Katarzyński, K., Sol, H., & Kus, A. 2001, A&A, 367, 809, doi: 10.1051/0004-6361:20000538
  • Khangulyan et al. (2018) Khangulyan, D., Bosch-Ramon, V., & Uchiyama, Y. 2018, MNRAS, 481, 1455, doi: 10.1093/mnras/sty2356
  • Koljonen et al. (2010) Koljonen, K. I. I., Hannikainen, D. C., McCollough, M. L., Pooley, G. G., & Trushkin, S. A. 2010, MNRAS, 406, 307, doi: 10.1111/j.1365-2966.2010.16722.x
  • Koljonen & Maccarone (2017) Koljonen, K. I. I., & Maccarone, T. J. 2017, MNRAS, 472, 2181, doi: 10.1093/mnras/stx2106
  • Levine et al. (1996) Levine, A. M., Bradt, H., Cui, W., et al. 1996, ApJ, 469, L33, doi: 10.1086/310260
  • Mattox et al. (1996) Mattox, J. R., Bertsch, D. L., Chiang, J., et al. 1996, ApJ, 461, 396, doi: 10.1086/177068
  • McCollough (2023) McCollough, M. 2023, Conference presentation, https://drive.google.com/file/d/1l45DBFbzkbRmTcOY6XhIDoXGJmh5I1hW/view
  • McKinney et al. (2013) McKinney, J. C., Tchekhovskoy, A., & Blandford, R. D. 2013, Science, 339, 49, doi: 10.1126/science.1230811
  • Miller-Jones et al. (2004) Miller-Jones, J. C. A., Blundell, K. M., Rupen, M. P., et al. 2004, ApJ, 600, 368, doi: 10.1086/379706
  • Miller-Jones et al. (2006) Miller-Jones, J. C. A., Fender, R. P., & Nakar, E. 2006, MNRAS, 367, 1432, doi: 10.1111/j.1365-2966.2006.10092.x
  • Mioduszewski et al. (2001) Mioduszewski, A. J., Rupen, M. P., Hjellming, R. M., Pooley, G. G., & Waltman, E. B. 2001, ApJ, 553, 766, doi: 10.1086/320965
  • Mirabel & Rodríguez (1994) Mirabel, I. F., & Rodríguez, L. F. 1994, Nature, 371, 46, doi: 10.1038/371046a0
  • Molina & Bosch-Ramon (2018) Molina, E., & Bosch-Ramon, V. 2018, A&A, 618, A146, doi: 10.1051/0004-6361/201833681
  • Molina et al. (2019) Molina, E., del Palacio, S., & Bosch-Ramon, V. 2019, A&A, 629, A129, doi: 10.1051/0004-6361/201935960
  • Molnar et al. (1988) Molnar, L. A., Reid, M. J., & Grindlay, J. E. 1988, ApJ, 331, 494, doi: 10.1086/166575
  • Mori et al. (1997) Mori, M., Bertsch, D. L., Dingus, B. L., et al. 1997, ApJ, 476, 842, doi: 10.1086/303667
  • Narayan et al. (2003) Narayan, R., Igumenshchev, I. V., & Abramowicz, M. A. 2003, PASJ, 55, L69, doi: 10.1093/pasj/55.6.L69
  • Ogley et al. (2001) Ogley, R. N., Bell Burnell, S. J., & Fender, R. P. 2001, MNRAS, 322, 177, doi: 10.1046/j.1365-8711.2001.04129.x
  • Reid & Miller-Jones (2023) Reid, M. J., & Miller-Jones, J. C. A. 2023, ApJ, 959, 85, doi: 10.3847/1538-4357/acfe0c
  • Schalinski et al. (1995) Schalinski, C. J., Johnston, K. J., Witzel, A., et al. 1995, ApJ, 447, 752, doi: 10.1086/175914
  • Szostek & Zdziarski (2008) Szostek, A., & Zdziarski, A. A. 2008, MNRAS, 386, 593, doi: 10.1111/j.1365-2966.2008.13073.x
  • Szostek et al. (2008) Szostek, A., Zdziarski, A. A., & McCollough, M. L. 2008, MNRAS, 388, 1001, doi: 10.1111/j.1365-2966.2008.13479.x
  • Tavani et al. (2009) Tavani, M., Bulgarelli, A., Piano, G., et al. 2009, Nature, 462, 620, doi: 10.1038/nature08578
  • Tchekhovskoy et al. (2009) Tchekhovskoy, A., McKinney, J. C., & Narayan, R. 2009, ApJ, 699, 1789, doi: 10.1088/0004-637X/699/2/1789
  • Tchekhovskoy et al. (2011) Tchekhovskoy, A., Narayan, R., & McKinney, J. C. 2011, MNRAS, 418, L79, doi: 10.1111/j.1745-3933.2011.01147.x
  • van der Klis & Bonnet-Bidaud (1989) van der Klis, M., & Bonnet-Bidaud, J. M. 1989, A&A, 214, 203
  • van Kerkwijk (1993) van Kerkwijk, M. H. 1993, A&A, 276, L9
  • van Kerkwijk et al. (1996) van Kerkwijk, M. H., Geballe, T. R., King, D. L., van der Klis, M., & van Paradijs, J. 1996, A&A, 314, 521, doi: 10.48550/arXiv.astro-ph/9604100
  • van Kerkwijk et al. (1992) van Kerkwijk, M. H., Charles, P. A., Geballe, T. R., et al. 1992, Nature, 355, 703, doi: 10.1038/355703a0
  • Veledina et al. (2024a) Veledina, A., Muleri, F., Poutanen, J., et al. 2024a, Nature Astronomy, doi: 10.1038/s41550-024-02294-9
  • Veledina et al. (2024b) Veledina, A., Poutanen, J., Bocharova, A., et al. 2024b, arXiv e-prints, arXiv:2407.02655, doi: 10.48550/arXiv.2407.02655
  • Yang et al. (2023) Yang, J., García, F., del Palacio, S., et al. 2023, MNRAS, 526, L1, doi: 10.1093/mnrasl/slad111
  • Yoon & Heinz (2015) Yoon, D., & Heinz, S. 2015, ApJ, 801, 55, doi: 10.1088/0004-637X/801/1/55
  • Yoon et al. (2016) Yoon, D., Zdziarski, A. A., & Heinz, S. 2016, MNRAS, 456, 3638, doi: 10.1093/mnras/stv2954
  • Zdziarski (2014) Zdziarski, A. A. 2014, MNRAS, 445, 1321, doi: 10.1093/mnras/stu1835
  • Zdziarski et al. (2012a) Zdziarski, A. A., Lubiński, P., & Sikora, M. 2012a, MNRAS, 423, 663, doi: 10.1111/j.1365-2966.2012.20903.x
  • Zdziarski et al. (2012b) Zdziarski, A. A., Maitra, C., Frankowski, A., Skinner, G. K., & Misra, R. 2012b, MNRAS, 426, 1031, doi: 10.1111/j.1365-2966.2012.21635.x
  • Zdziarski et al. (2012c) Zdziarski, A. A., Sikora, M., Dubus, G., et al. 2012c, MNRAS, 421, 2956, doi: 10.1111/j.1365-2966.2012.20519.x
  • Zdziarski et al. (2014) Zdziarski, A. A., Stawarz, Ł., Pjanka, P., & Sikora, M. 2014, MNRAS, 440, 2238, doi: 10.1093/mnras/stu420
  • Zdziarski et al. (2022) Zdziarski, A. A., Stawarz, Ł., Sikora, M., & Nalewajko, K. 2022, MNRAS, 515, L17, doi: 10.1093/mnrasl/slac060
  • Zdziarski et al. (2018) Zdziarski, A. A., Malyshev, D., Dubus, G., et al. 2018, MNRAS, 479, 4399, doi: 10.1093/mnras/sty1618