Abstract
We perform general relativistic radiation magnetohydrodynamics simulations of super-Eddington accretion flows around a neutron star with a dipole magnetic field for modeling the Galactic ultraluminous X-ray source exhibiting X-ray pulsations, Swift J0243.6+6124. Our simulations show the accretion columns near the magnetic poles, the accretion disk outside the magnetosphere, and the outflows from the disk. It is revealed that the effectively optically thick outflows, consistent with the observed thermal emission at ∼107 K, are generated if the mass accretion rate is much higher than the Eddington rate and the magnetospheric radius is smaller than the spherization radius. In order to explain the blackbody radius (∼100–500 km) without contradicting the reported spin period (9.8 s) and spin-up rate (), a mass accretion rate of is required. Since the thermal emission was detected in two observations with of −2.22 × 10−8 s s−1 and −1.75 × 10−8 s s−1 but not in another with , the surface magnetic field strength of the neutron star in Swift J0243.6+6124 is estimated to be between 3 × 1011 G and 4 × 1012 G. From this restricted range of magnetic field strength, the accretion rate would be when the thermal emission appears and when it is not detected. Our results support the hypothesis that the super-Eddington phase in the 2017–2018 giant outburst of Swift J0243.6+6124 is powered by highly super-Eddington accretion flows onto a magnetized neutron star.
Export citation and abstract BibTeX RIS
Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.
1. Introduction
Ultraluminous X-ray sources (ULXs) are bright X-ray compact objects, whose X-ray luminosities exceed the Eddington luminosity for stellar-mass black holes ∼1039 erg s−1, and have been discovered in off-nuclear regions of our Galaxy and nearby galaxies (see reviews by Kaaret et al. 2017; Fabrika et al. 2021). Recently, several ULXs have been shown to have coherent pulsations at periods, ∼1–10 s (e.g., Bachetti et al. 2014; Fürst et al. 2016; Israel et al. 2017a, 2017b). These objects are called ULX pulsars (ULXPs). It is widely believed that these pulsations can be observed if the radiation from the vicinity of the neutron star (NS) periodically changes via rotation of the NS. In this case, the ULXP is powered by a super-Eddington accretion onto the magnetized NS since the luminosity of the ULXP exceeds the Eddington luminosity for the NS, LEdd.
The super-Eddington accretion flows around the magnetized NS consist of the accretion disk, the outflows, and the accretion columns. The super-Eddington accretion disk is truncated at a certain radius, called the truncation radius, by the magnetic fields of the NS, provided that the magnetic field strength of the NS is high enough to prevent disk accretion. It is considered that radiatively driven outflows are launched from the super-Eddington accretion disk, which forms on the outside of the truncation radius, since the vertical radiation pressure force is larger than the vertical gravitational force (Shakura & Sunyaev 1973; Poutanen et al. 2007). The gas moves along the magnetic field lines inside the truncation radius. Then, column accretion flows, called the accretion columns, are formed near the magnetic poles of the NS. The radiative luminosity at the base of the accretion columns exceeds the Eddington luminosity (Basko & Sunyaev 1976; Mushtukov et al. 2015; Kawashima et al. 2016). If the magnetic axis is misaligned with the rotation axis of the NS, then the pulsation exceeding the Eddington luminosity can be observed via the precession of the accretion columns (Mushtukov et al. 2018; Inoue et al. 2020). This inflow-outflow structure around the magnetized NS has been shown by general relativistic radiation magnetohydrodynamics (GR-RMHD) simulations of super-Eddington accretion flows onto an NS with a dipole magnetic field (Takahashi & Ohsuga 2017; Abarca et al. 2021).
Recent observations suggested that outflows exist in ULXs (see, e.g., Pinto et al. 2016) and ULXPs (see, e.g., Kosec et al. 2018). In the Galactic ULXP, Swift J0243.6+6124, the outflows, as well as a relativistic jet (van den Eijnden et al. 2018) were reported (see, e.g., Tao et al. 2019; van den Eijnden et al. 2019). Tao et al. (2019) showed that the radiation spectrum of Swift J0243.6+6124, at the peak of the 2017–2018 giant outburst, can be explained by the thermal blackbody at 0.4–0.6 keV ∼ 107 K, with a radius of 100–500 km (however, see also Jaisawal et al. 2019). This thermal emission is considered to originate from the optically thick outflows. It is not known whether the outflows appearing in the numerical simulations mentioned above can explain such thermal emission. Although a semianalytical model was proposed to account for the effects of the outflows (Chashkina et al. 2019; Mushtukov et al. 2019), it remains unclear whether such powerful outflows can occur or not.
In this paper, we perform GR-RMHD simulations of super-Eddington accretion flows onto an NS with a dipole magnetic field for modeling Swift J0243.6+6124. The dependence of the magnetospheric and outflow structure on the surface magnetic field strength at the NS surface BNS and the mass accretion rate , which is not investigated in previous studies (Takahashi & Ohsuga 2017; Abarca et al. 2021), is reported. It is revealed that the super-Eddington accretion onto the magnetized NS reproduces the thermal emission observed in this object. We also restrict BNS and based on the condition of the thermal emission. This paper is organized as follows: we will describe the numerical methods in Section 2 and then present the results in Section 3. Section 4 is devoted to the discussion. Finally, we give our conclusion in the Section 5.
2. Numerical Methods
In this study, we numerically solve the GR-RMHD equations in Schwarzschild polar coordinates (t, r, θ, ϕ), assuming axisymmetry with respect to θ = 0, π. We use the GR-RMHD simulation code, UWABAMI (Takahashi & Ohsuga 2017). The moment formalism is adopted to describe the radiation field (Thorne 1981). M1 closure is used as the closure relation (Levermore 1984; Kanno et al. 2013; Sadowski et al. 2013). Using the closure, the radiation fields can be updated by solving the zeroth and first moment equations of the radiation without solving the radiative transfer equation. Hereafter, Greek suffixes and Latin suffixes represent spacetime components and space components, respectively. The speed of light c and the gravitational constant G are set to 1 unless otherwise noted.
2.1. Basic Equations
The basic equations are as follows (see, e.g., Takahashi et al. 2018): the mass conservation law,
the energy-momentum conservation laws for magnetohydrodynamics (MHD),
the energy-momentum conservation laws for the radiation field,
the induction equations,
where ρ is the proper mass density, uμ is the four-velocity of the gas, Bi is the magnetic field vector in the laboratory frame, bμ is the magnetic four-vector in the fluid frame, and g is the determinant of the metric, . The energy-momentum tensor of ideal MHD Tμ ν consists of that of fluid and electromagnetic field Mμ ν :
where e is the internal energy density, pgas = (Γ − 1)e is the gas pressure (Γ = 5/3), and pmag = b2/2 is the magnetic pressure in the fluid frame. The energy-momentum tensor of the radiation field in the M1 formalism is given by (Sadowski et al. 2013)
Here, , , and are the radiation energy density, the radiation pressure in the radiation rest frame, and the four-velocity of the radiation rest frame, respectively. The radiation four-force Gμ , which describes the interaction between the ideal MHD and radiation field, is defined as
where κabs and κsca are the opacity for free–free absorption and isotropic electron scattering, respectively:
Here, Te is the electron temperature. The blackbody intensity is given by , where a is the radiation constant. In this study, we consider the thermal Comptonization defined as follows (Fragile et al. 2018; Utsumi et al. 2022):
Here, is the radiation energy density in the fluid frame, is the radiation temperature, and me is the electron rest mass. We take Te = Tg for simplicity, where Tg is the gas temperature. The gas temperature can be derived from
where mp is the proton mass, k is the Boltzmann constant, and μw = 0.5 is the mean molecular weight. We consider the mean-field dynamo in our simulations (Sadowski et al. 2015).
2.2. Numerical Models
In this study, we take MNS = 1.4M⊙ and rNS = 10 km, where MNS and rNS are the mass and radius of the NS, and consider the NSs with dipole magnetic fields (Wasserman & Shapiro 1983). The dipole magnetic field strength at the NS surface BNS applied in this study, which is relatively weaker than the typical value observed in the X-ray pulsar, 1011−13 G (see Section 4.5), is described in Table 1. Table 1 also reports the maximum gas density of the initial torus ρ0, the end time of the simulation tend, the time after which the mass outflow rate is almost constant teq, and numerical grid points (Nr , Nθ ). The light-crossing time for the gravitational radius of the NS, rg = MNS = 2.1 km, is denoted by tg. The computational domain consists of r = [rNS, rout], where rout = 2100 km, and θ = [0, π]. The size of the radial grid exponentially increases with r, and the size of the polar grid is uniform (Takahashi & Ohsuga 2017). Multipole components of the NS magnetic field are not considered. GR-RMHD simulations considering the multipole components are left as important future work, although the GR-MHD simulations employing quadrupole fields were recently performed by Das et al. (2022). The rotation of the NSs is ignored since the rotation period of the NSs observed in ULXPs, 1–10 s, is much longer than the rotational timescale of the accretion disk, ∼10−2 s, even within r ∼ 100 km. There remains the possibility that ULX is a millisecond pulsar (Kluzniak & Lasota 2015), which we will discuss later. We assume that the magnetic axis coincides with the rotation axis of the accretion disk. The convergence of the simulation results is confirmed using models B1e10D001, B1e10D001_a, B1e10D001_b, and B1e10D001_c. These models have the same initial parameters except for the resolution.
Table 1. Parameters for Different Models
Parameter | BNS | ρ0 | tend | teq | (Nr , Nθ ) |
---|---|---|---|---|---|
Unit | (G) | (g cm−3) | (tg) | (tg) | |
B3e9D001 | 3.3 × 109 | 0.01 | 50,000 | 30,000 | (592, 412) |
B3e9D01_np | 3.3 × 109 | 0.1 | 50,000 | 40,000 | (592, 412) |
B1e10D001 | 1010 | 0.01 | 50,000 | 30,000 | (592, 512) |
B1e10D002 | 1010 | 0.021 | 50,000 | 30,000 | (592, 512) |
B1e10D004 | 1010 | 0.046 | 50,000 | 35,000 | (592, 512) |
B1e10D01 | 1010 | 0.1 | 50,000 | 35,000 | (592, 412) |
B1e10D1_np | 1010 | 1 | 50,000 | 40,000 | (592, 512) |
B3e10D001 | 3.3 × 1010 | 0.01 | 50,000 | 40,000 | (732, 512) |
B3e10D01 | 3.3 × 1010 | 0.1 | 50,000 | 35,000 | (732, 512) |
B3e10D02 | 3.3 × 1010 | 0.21 | 50,000 | 40,000 | (592, 512) |
B3e10D04 | 3.3 × 1010 | 0.46 | 50,000 | 40,000 | (592, 512) |
B3e10D1 | 3.3 × 1010 | 1 | 50,000 | 40,000 | (592, 412) |
B1e10D001_a | 1010 | 0.01 | 40,000 | 25,000 | (732, 512) |
B1e10D001_b | 1010 | 0.01 | 40,000 | 30,000 | (592, 412) |
B1e10D001_c | 1010 | 0.01 | 40,000 | 30,000 | (592, 326) |
Note. BNS, ρ0, and (Nr , Nθ ) are the surface magnetic field strength of the NS, the maximum gas density of the initial torus, and numerical grid points, respectively. The simulation continues until t = tend. The mass outflow rate is almost constant after t = teq. The model without accretion columns is denoted by \_np at the end of the model name.
Download table as: ASCIITypeset image
We initially set the equilibrium torus given by Fishbone & Moncrief (1976), as can be seen in Figure 1, where . The gas pressure pgas inside the torus is replaced with pgas + prad by assuming a local thermodynamic equilibrium, Tg = Tr. The inner edge of the torus and the maximum pressure radius are set to 210 km and 304.5 km, respectively. The poloidal-loop magnetic fields inside the torus are considered in addition to the dipole magnetic field of the NS. The vector potential of this loop magnetic field is proportional to ρ. The embedded loop magnetic field is antiparallel to the dipole magnetic field at the inner edge of the torus (Romanova et al. 2011; Parfrey & Tchekhovskoy 2017; Takahashi & Ohsuga 2017). We set the maximum (pgas + prad)/pmag to 100 inside the torus. We give a perturbation on pgas + prad by 10% to break an equilibrium state of the torus.
The NS and torus are initially surrounded by a relatively hot and low-density corona,
where ρc is determined by , and Tinit is set to 5 × 1011 K. The floor value of the gas density ρflr and gas pressure pflr are described in Section 2.3. The radiation energy density of the corona is , where . The velocity of the gas ui and radiation rest frame are set to zero in the corona. The present result is not affected by the corona gas, which is set initially. This is because these hot gases are blown away by the radiation force from the accretion disk after some time has elapsed since the simulations began.
2.3. Boundary and Floor Condition
We adopt the outgoing boundary at r = rout, and the reflective boundary at θ = 0, π. We assume that the inner boundary, r = rNS, is the boundary at which the gas can be swallowed by the NS but the energy cannot be swallowed by the NS (Ohsuga 2007). At inner ghost cells, ρ, pgas, and are the free boundary, and ui and are set to zero. For the magnetic field, the radial component is fixed to be the dipole field (Wasserman & Shapiro 1983), while the tangential component is set according to the free boundary condition (Parfrey & Tchekhovskoy 2017; Abarca et al. 2021). The numerical flux in mass conservation law, energy conservation law for MHD, and energy conservation law for the radiation field are zero at the NS surface. The numerical flux of the induction equation is also set to zero at the NS surface to satisfy the ideal MHD condition. We also impose the following condition on the gas density and pressure at the NS surface:
where ρ* and Tg,* are the gas density and gas temperature calculated from the conservative variables, respectively. In this case, the gas density is set to the initial value by keeping the gas temperature constant at the NS surface. The amount of reduced kinetic and thermal energy densities by applying Equations (13) and (14) is added to the radiation energy density Rt t . Here, the kinetic and thermal energy densities are defined as follows:
The mass reduced by the treatment of Equation (13) is considered to be swallowed by the NSs.
We impose a floor condition on ρ and pgas to solve the GR-RMHD equations stably:
where is set to 103. The upper and lower limits of the gas temperature are set to 5 × 1011 K and 5 × 105 K, respectively.
The gas density should be low near θ = 0, π because the strong magnetic pressure and tension due to the open poloidal magnetic field lines inhibit the mass flow toward the rotation axis in the very vicinity of the rotation axis. In the numerical simulation, however, the gas reaches the region of θ = 0, π due to the numerical diffusion. Then, the unphysical mass outflow driven by the radiation force is formed close to the polar axis. We decided to ignore the gas–radiation interaction to avoid this problem in this region. Practically, we set the absorption and scattering opacity to zero at σ > σrad, where σ = b2/ρ is the magnetization. In this study, we set σrad = 10 in all models. This value is set so that the opacity near the rotation axis is zero while the radiation from the dense accretion flows at the NS surface can be solved.
3. Results
In all models, the initial torus deviates from the equilibrium state due to the magnetorotational instability caused by the differential rotation in the torus after the simulation starts. The angular momentum of the gas is transported outward, and then the accretion disk is formed around the equatorial plane (z = 0 plane). The radiatively driven outflows are launched from the accretion disk. The accretion disk in models B3e9D001, B1e10D001, B1e10D002, B1e10D004, B1e10D01, B3e10D001, B3e10D01, B3e10D02, B3e10D04, and B3e10D1 is truncated before it reaches to the NS surface. Then, the accreting gas moves along magnetic field lines, and column accretion flows (accretion columns) are formed near the poles of the NS. On the other hand, the accretion disk in models B3e9D01_np and B1e10D1_np directly connects to the NS surface without being truncated.
3.1. Quasi-steady-state Structure
We use the numerical data from period teq to tend shown in Table 1 for investigating a quasi-steady-state structure in the present work, since the flows become quasi-steady and the mass outflow rate is almost constant after teq. First, we introduce the quasi-steady-state structure of the accretion flows and outflows around the magnetized NS. Figure 2 shows the time-averaged overview of models B1e10D001 (left) and B1e10D01 (right) in the quasi-steady state. Colors represent ρ in panels (a) and (c), and in panels (b) and (d). The white region with r < 10 km is the NS, and its center is located at the origin, (R, z) = (0, 0). The magnetic field lines are shown by cyan solid lines. We find that the accretion disk is formed near the equatorial plane at 20 km ≲ R ≲ 200 km (the green region in panels (a) and (c)). In panels (b) and (d), it is clear that is large in the accretion disk. The accretion column is formed near the NS magnetic poles, r ≲ 20 km, the details of which are described in Figure 3. Streamlines follow the poloidal velocity vectors in panels (a) and (c). It can be seen that the outflows emanate from the disk toward the low-density region. In addition, turbulent motions exist inside the accretion disk. The radiative flux is represented by streamlines in panels (b) and (d). We can see that the outward radiative flux from the accretion disk appears. The accretion disk, the outflows, and the outward radiative flux from the accretion disk are common to all models. Both ρ and in the accretion disk tend to increase as ρ0 becomes large. Actually, both of ρ and in model B1e10D01 (ρ0 = 0.1 g cm−3) are an order of magnitude higher than those in model B1e10D001(ρ0 = 0.01 g cm−3).
Download figure:
Standard image High-resolution imageDownload figure:
Standard image High-resolution imageWe can see that the gas accretes onto the NS around the pole outside the accretion column (r ≲ 100 km and θ ≲ 40°, 140° ≲ θ in panel (a)). This is because the interaction between the gas and radiation is artificially switched off at the high-σ region (σ > σrad) in our simulation to maintain numerical stability. The radiation force does not work around the polar axis, and the gas falls down to the NS. We note that the density in this region is too small to contribute to the mass accretion rate, and the effects on the overall structure are negligibly small.
Next, we explain the dependence of the magnetospheric structure on ρ0. Colors in Figure 3 show ρ, and magnetic field lines are represented by cyan lines. We take models B1e10D001, B1e10D01, and B1e10D1_np. We plot the time-averaged magnetization, , by yellow contours. Here, the angle brackets denote the time-averaged value. The magnetization on the high-density side of the yellow line is < 10. In the left panel (model B1e10D001), the high-density accretion column, around (R, z) = (5 km, 12 km), is formed (one-side accretion). Although not shown in Figure 3, the same is true of models B3e9D001, B1e10D002, B3e10D001, B3e10D01, and B3e10D02. In the middle panel (model B1e10D01), the accretion columns are formed at (R, z) = (6 km, ± 10 km), which is true of models B1e10D004, B3e10D04, and B3e10D1. At the truncation radius, which is the radius of the inner edge of the accretion disk, the gas begins to move away from the equatorial plane toward the NS poles. The truncation radius is R ∼ 25 km in model B1e10D001 and R ∼ 13 km in model B1e10D001, and approximately corresponds to the position where at the equatorial plane. The density inside this radius is <1% of the density of the disk, except inside the accretion column. Also, the truncation radius is roughly comparable to the magnetospheric radius, rM, ∼27 km for model B1e10D001 and ∼13 km for model B1e10D01. In the present work, rM is defined as the maximum radius of the region, where , inside the maximum pressure radius of the initial torus, 304.5 km. Here, is uϕ -weighted plasma beta,
Although prad is the radiation pressure in the radiation rest frame, it is approximately equal to the radiation pressure in the fluid frame since the MHD and radiation field inside the accretion disk are in local thermal equilibrium, and therefore the radiation rest frame coincides with the fluid frame. We note that the gas pressure is negligible in the super-Eddington accretion disk (Takahashi & Ohsuga 2017), so that pmag ∼ prad at . The coincidence between the truncation radius and rM indicates that the truncation radius is determined by the balance between pmag and prad. As can be seen from the left and middle panels in Figure 3, the truncation radius decreases as ρ0 increases.
In model B1e10D1_np, since the radiation pressure of the accretion disk is larger than the magnetic pressure of the dipole field even at the NS surface, the accretion disk reaches the NS surface without being truncated and no accretion column is formed. In this case, the region where reaches the NS surface, and rM cannot be determined. Although not shown in this figure, accretion columns are not formed in model B3e9D01_np for the same reason. Based on the hypothesis that the radiation from the accretion column is the origin of the X-ray pulse, models B3e9D01_np and B1e10D1_np are inappropriate for explaining ULXPs. Incidentally, the reason why the one-side accretion occurs in model B1e10D001 is described in Section 4.4.
Table 2 lists the time-averaged value of the mass accretion rate at r = 11 km, mass outflow rate at r = rout, radiative luminosity Lrad at r = rout, and kinetic luminosity Lkin at r = rout of each model. The Eddington mass accretion rate is denoted by . The mass accretion rate and outflow rate are defined as
The radiative and kinetic luminosity can be calculated using following formula (Sadowski et al. 2016):
As can be seen from this table, , , Lrad, and Lkin tend to be large as ρ0 increases. We also find that Lkin/Lrad tends to be large as increases. Indeed, the mass accretion rate is and Lkin/Lrad ∼ 0.083 in the case of model B1e10D001, mass accretion rate is and Lkin/Lrad ∼ 0.86 in the case of model B1e10D01, and mass accretion rate is and Lkin/Lrad ∼ 1.2 in the case of model B1e10D1_np. That is, the kinetic luminosity is comparable to the radiation luminosity for the model with high mass accretion rate. These trends are also reported in nonrelativistic radiation hydrodynamics simulation (Ohsuga 2007).
Table 2. Time-averaged Value for the Mass Accretion Rate, Mass Outflow Rate, Radiative Luminosity, and Kinetic Luminosity
Tbb | rbb | |||||
---|---|---|---|---|---|---|
Model | [] | [] | [LEdd] | [LEdd] | [107 K] | [km] |
B3e9D001 | 32 | 36 | 3.0 | 0.15 | 0.72 | 57 |
B3e9D01_np | 220 | 2700 | 130 | 47 | 0.63 | 230 |
B1e10D001 | 54 | 74 | 4.1 | 0.34 | 0.71 | 58 |
B1e10D002 | 110 | 110 | 9.5 | 1.0 | 0.73 | 65 |
B1e10D004 | 160 | 730 | 21 | 7.1 | 0.66 | 130 |
B1e10D01 | 500 | 2600 | 42 | 36 | 0.66 | 240 |
B1e10D1_np | 4700 | 23000 | 500 | 410 | 0.50 | 1100 |
B3e10D001 | 98 | 170 | 5.5 | 0.89 | 0.75 | 50 |
B3e10D01 | 720 | 2900 | 140 | 56 | 0.69 | 270 |
B3e10D02 | 650 | 4600 | 150 | 37 | 0.63 | 250 |
B3e10D04 | 1500 | 13000 | 320 | 210 | 0.59 | 660 |
B3e10D1 | 2300 | 28000 | 490 | 460 | 0.55 | 1300 |
Note. The mass accretion rate is measured at r = 11 km, and the others are calculated at r = rout.
Download table as: ASCIITypeset image
Figure 4 shows rM and the time derivative of the NS rotation period (spin-up rate) estimated using our numerical models as a function of the mass accretion rate. Points in Figure 4(a) indicate the resulting rM for BNS = 3.3 × 109 G (red), 1010 G (blue), and 3.3 × 1010 G (black), in which all models have accretion columns. The magnetospheric radius rM decreases with increasing and increases with increasing BNS. Dashed lines indicate the analytical solution for rM, which can be calculated assuming that the magnetic pressure of the dipole field balances with the radiation pressure of the accretion disk (Takahashi & Ohsuga 2017, see also Appendix B for details):
where α is the viscous parameter (Shakura & Sunyaev 1973), and we take α = 0.1. We can find that the resulting rM is consistent with the analytical one. We note that the results of models B1e10D1_np and B3e9D01_np are not plotted since rM is not determined in these models as mentioned above.
Download figure:
Standard image High-resolution imageWe estimate as follows:
(Shapiro & Teukolsky 1986; Takahashi & Ohsuga 2017). Here, and P are the specific angular momentum and rotation period of the NS, respectively. For numerical estimation, we assume P = 9.8 s, which corresponds to that observed in Swift J0243.6+6124. Figure 4(b) shows that is in the range from −10−9 to −10−7 s s−1, which is not inconsistent with the observed value of Swift J0243.6+6124, (see Section 3.3 for details). It can also be clearly seen that the larger and higher BNS lead to large . Dashed lines indicate the analytical solution for , which can be calculated assuming that the Keplerian angular momentum at rM is transported to the NS without dissipation (Takahashi & Ohsuga 2017; see also Appendix B):
The resulting is roughly comparable to the analytical one.
3.2. Powerful Outflows
In Figure 5, we show that powerful outflows are launched from the super-Eddington accretion disk and are driven by radiation and centrifugal forces. The top panel in Figure 5 shows time-averaged (thick solid line), (dashed line), and net flow rate (thin solid line) as a function of r in the case of model B1e10D01. The position of rM in this model is represented by a vertical line. This panel indicates that outflows mainly originate from the accretion disk. This is clearly understood from the fact that increases significantly with increasing r, and for r < rM is much smaller than that for r > rM. The mass accretion rate decreases with decreasing r due to efficient mass ejection from the accretion disk. The mass outflow rate is much smaller than the mass accretion rate at r < rM, and we find inside the magnetosphere.
Download figure:
Standard image High-resolution imageNext, we investigate the radial profile of forces acting on fluid elements. The steady-state axisymmetric equations of motion in the r- and θ-directions in the statistic observer frame are
where superscripts (r) and (θ) denote the r- and θ-components of the vectors in the static observer frame. The forces acting on fluid elements consist of the gravitational force fgrav, thermal force fthermal, radiation force frad, Lorentz force fmag, centrifugal force fcent, inertial force finertial, and advection force fadv. The centrifugal force is divided into two components, fθ,cent and fϕ,cent, which are caused by the polar (θ-) and azimuthal (ϕ-) motions, respectively (see Appendix C for details). We define the total force ftot as the sum of all forces without fgrav, and therefore holds.
The middle panel in Figure 5 shows time- and shell-averaged mass flux weighted radial forces normalized by the gravitational force as a function of r, which can be calculated as follows:
The bottom panel in Figure 5 shows the time-averaged mass flux weighted z-component of the forces normalized by the vertical gravitational force,
at the disk surface. The polar angle of the disk surface is defined as θ = π/2 − ΘH, where
It is clear that the outflows are mainly driven by the radiation force for 20 km < r < 50 km, where the radiation force is stronger than the other forces in both the r- and z-directions. For 50 km < r < 200 km, the gas is launched from the accretion disk by the radiation force and moves outward by centrifugal force. The Lorentz force is dominant at r < rM, but as already mentioned, most of the outflows are of disk origin (r > rM). The thermal force, centrifugal force due to the motion in the θ-direction, advection force, and inertial force are significantly small over the whole region. The radial distribution of forces for r > rM in models B3e9D001, B3e9D01, B1e10D001, B1e10D002, B1e10D004, B3e10D001, B3e10D01, B3e10D02, B3e10D04, and B3e10D1 is similar to that in model B1e10D01. The radial profile of forces outside r = rNS in models B3e9D01_np and B1e10D1_np, where the magnetosphere is not formed, is the same as that in the disk region outside r = rM in model B1e10D01.
In Figure 6, we show the gas temperature distribution and the location of the photosphere. In all models, the gas temperature tends to be high near the z-axis and in the equatorial plane at r < 750 km. Red lines indicate the photospheres, , where is the time-averaged effective optical depth measured as,
where . As can be seen in this figure, the photosphere extends to surround the disk region. The larger the mass accretion rate, the wider the region surrounded by the photosphere. This can be understood from a comparison of models B1e10D001, B1e10D01, and B1e10D1_np. Also, within the range of parameters employed in the present study, the position of the photosphere does not depend much on BNS. It is clear from the fact that the photosphere of models B1e10D01 and B3e10D01 hardly changes (see the red lines in Figure 6). In all models, the gas temperature on the photosphere is ∼107 K over a wide range of polar angles. From this, it is expected that blackbody radiation with a temperature of ∼107 K can be observed.
Download figure:
Standard image High-resolution imageHereafter, we define the thermalization radius rth as the radius where . That is, the radius of the photosphere is represented as rth. In Figure 7, we plot rth (upper panel) and the time-averaged gas temperature at the photospheres Tth (middle panel) as a function of polar angle θ. In model B1e10D1_np, which has the largest in our models, rth ≳ 1000 km at 30° < θ < 150°. Models with smaller tend to have smaller rth. The thermalization radius rth is slightly less than 1000 km at 40° < θ < 140° in models B1e10D01 and B3e10D01 and is ∼500 km at 60° < θ < 120° in model B1e10D001. The thermalization radius cannot be determined close to the rotation axis (θ ≲ 30°, 150° ≲ θ) since the gas is optically thin and even at the NS surface. The range of θ, within which rth can be obtained, narrows as decreases. In all models, Tth is almost constant at the angular range where rth can be determined. The bottom panel in Figure 7 shows the polar angle dependence of the radial gas velocity at the photosphere, v(r)(rth), where v(r) = u(r)/u(t). We can see v(r)(rth) > 0 on the photosphere, which indicates that most of the photosphere is located inside the optically thick outflows. Hereafter, we refer to such an outflow as an effectively optically thick outflow (Urquhart & Soria 2016).
Download figure:
Standard image High-resolution imageHere, we estimate the blackbody temperature Tbb by taking θ-averaged Tth on the photosphere as
This integration is performed on the surface where the thermalization radius can be estimated. The corresponding blackbody radius rbb can be calculated as
where the θ-averaged isotropic luminosity is
Here, θ integration is also conducted only in the range where the thermalization radius can be obtained. The resulting Tbb and rbb are presented in Table 2. The blackbody temperature is not so sensitive to the accretion rate and is ∼107 K, which is consistent with the blackbody temperature observed in Swift J0243.6+6124 at the super-Eddington phase (Tao et al. 2019). The resulting rbb is roughly fitted as
Tao et al. (2019) showed that the thermal photons would come from r = 100–500 km for Swift J0243.6+6124. Using Equation (38), our model can explain this radius when the accretion rate is . In the models treated in the present study, there is no clear result that rbb depends on BNS. We will discuss this point later.
We run three models, B1e10D001_a, B1e10D001_b, and B1e10D001_c, which have the same initial parameters as model B1e10D001 except for the resolution, to assess the effect of the resolution. We confirm that rM and of models B1e10D001_a and model B1e10D001_b are almost the same as those of model B1e10D001, while those of model B1e10D001_c are different from those of model B1e10D001. The resolution of the simulation models adopted in this study is the same as or higher than that of model B1e10D001_b. Therefore, our simulation results are independent of the resolution.
3.3. Magnetic Field Strength of the Neutron Star in Swift J0243.6+6124
Here, we suggest how to constrain BNS using analytical solutions of rM and and the conditions under which effectively optically thick outflows, which can reproduce the blackbody radiation observed in Swift J0243.6+6124, occur.
It is necessary to simultaneously satisfy rM < rco, where rco is the corotation radius, and rM can be defined (that is, rM > rNS) in ULXPs. The condition of rM < rco is for gas to accrete onto the NS (Illarionov & Sunyaev 1975). Otherwise, the gas would all be blown away due to the propeller effect. The accretion columns are formed in the case that the second condition is satisfied (see Figure 3 for details). Substituting Equations (24) and (27) into these conditions, we get
and
If the magnetic field strength at the NS surface satisfies the inequalities of Equations (39) and (40), then the column accretion onto the magnetic poles of the NS appears, and such objects can be the candidates for the ULXPs.
Next, we consider the emergence of the outflows from the super-Eddington disks. The radiatively driven outflows are thought to be launched in the slim disk region, which is inside the spherization radius (or trapping radius), rsph. This is discussed in Shakura & Sunyaev (1973) and revealed by the numerical simulations (Kitaki et al. 2021; Yoshioka et al. 2022). The spherization radius is roughly estimated as . Since the disk is truncated at r = rM, the ULXPs accompanying the outflows appear if the condition of rM < rsph is satisfied in addition to rNS < rM < rco (Mushtukov et al. 2019). Using Equations (24) and (27), rM < rsph becomes
Equations (39)–(41) are the general conditions to explain the ULXPs, while we concentrate on the discussion about Swift J0243.6+6124 in the following. As described in Section 3.2, the range of the mass accretion rate that can reproduce the blackbody radiation observed in this object is from the relation in Equation (38). Using Equation (27), this condition is transformed as
Among ULXPs with outflows, objects that meet the conditions of Equations (41) and (42) would have effectively optically thick outflows, and the blackbody radiation with rbb > 100 km would appear in the radiation spectrum.
Figure 8 summarizes the conditions shown above in the plane. Here, we assume P = 9.8 s, which is the rotation period observed in Swift J0243.6+6124 (Doroshenko et al. 2018; Chen et al. 2021). We take α = 0.1, MNS = 1.4M⊙, and rNS = 10 km. The red shaded region indicates the ULXP with outflow, which satisfies both rNS < rM < rco and rM < rsph. The ULXP without outflow is in the white region, where rNS < rM < rco is satisfied but rM < rsph is not satisfied. The condition of ULXPs such as rM > rco (propeller regime) or rM < rNS (nonpulsating regime) is not satisfied in gray shaded regions. In the blue hatched region where the accretion rate is , effectively optically thick outflows with rbb = 100–500 km and Tbb ∼ 107K appear so that the blackbody radiation, as observed in Swift J0243.6+6124, is thought to be reproduced. Here, it is confirmed that the models that can successfully reproduce such effectively optically thick outflows and accretion columns (B1e10D004, B1e10D01, B3e10D01, and B3e10D02) are located in the blue hatched region (filled circles). In contrast, other models are located outside the blue hatched region (filled squares). In observations of Swift J0243.6+6124, P is almost constant, while is reported to be −2.22 × 10−8 s s−1 (Obs. 1), −1.75 × 10−8 s s−1 (Obs. 2), and −6.8 × 10−9 s s−1 (Obs. 3; Doroshenko et al. 2018; Chen et al. 2021). These observed spin-up rates are denoted by black dashed lines in Figure 8. In Obs. 1 and 2, the observed isotropic luminosity exceeds LEdd, and the thermal emission of which the observed blackbody radius is 100–500 km is observed (obsID: 90302319004, 90302319006 and 90302319008 in Tao et al. 2019). Then, the magnetic field strength is restricted to be 2 × 1010 G < BNS < 5 × 1012 G for Obs. 1 and 1010 G < BNS < 4 × 1012 G for Obs. 2, which are represented by black solid lines. In Obs. 3, on the other hand, the luminosity is lower than LEdd, and the thermal emission is not observed. Such features are realized in the condition that the magnetic field is in the range of 3 × 1011 G < BNS < 1014 G (see black solid line for Obs. 3). Therefore, the magnetic field strength that satisfies the three observations is between 3 × 1011 G and 4 × 1012 G. Using Equation (27), the mass accretion rate corresponding to this range is estimated to be for Obs. 1 and 2, and for Obs. 3. Since the luminosity has been reported to be ∼1039 erg s−1 at the super-Eddington phase in Swift J0243.6+6124 (Obs.1 and 2), the radiative efficiency of this object is 1%–5%, which is smaller than the standard accretion efficiency. This indicates that Swift J0243.6+6124 is observed at a relatively large viewing angle within the range where the pulse emission is detected. The magnetic field strength evaluated in the present study is not inconsistent with previous studies (Tsygankov et al. 2018; Doroshenko et al. 2020). Tsygankov et al. (2018) estimated the upper limit for BNS in Swift J0243.6+6124, <6 × 1012 G, from the fact that the transition to the propeller regime was not detected even at a luminosity of 6 × 1035 erg s−1. Doroshenko et al. (2020) concluded that BNS in Swift J0243.6+6124 is in (3–9) × 1012G and likely at the lower limit of this range. We note that BNS here is assumed not to change during the observation. This assumption would be reasonable since it has been pointed out that the NS magnetic field strength does not change for a sufficiently long period, >1 Myr, based on the observations of some X-ray pulsars and the ULXP, NGC 1313 X-2 (Makishima et al. 1999; Sathyaprakash et al. 2019). Although the NSs might have multipole magnetic fields, we focus on the case of the dipole magnetic fields in this study. We will discuss multipole fields later.
Download figure:
Standard image High-resolution image4. Discussions
4.1. Geometrical Beaming
In all of our models, the radiation is highly beamed by the outflows from the accretion disk. The amplification factor exceeds 100 at maximum, which is higher than that indicated by King & Lasota (2019; typically 1/b ∼ 1–10; see also Takahashi & Ohsuga 2017; Abarca et al. 2021). The cause of this difference is probably the magnetic field strength. The difference will be smaller if we employ a strong BNS by which rM nearly equals rsph, although rM is much smaller than rsph in the present study. As rM approaches rsph, the outflow rate would decrease since the outflows mainly occur in rM < r < rsph. In this case, the outflow can no longer effectively collimate the radiation, and therefore 1/b would decrease (Abarca et al. 2021). Although simulating rM ∼ rsph is difficult and beyond the scope of the present study (see also Section 4.5), we plan to conduct such simulations in the future.
4.2. Comparison with Observations of NGC 5907 ULX1 and NGC 1313 X-2
Here, we show, based on the present simulations, that the ULXPs NGC 5907 ULX1 and NGC 1313 X-2 are thought to be powered by the highly super-Eddington accretion onto NSs with relatively weak magnetic fields. In NGC 5907 ULX1, the mechanical power (), which is evaluated from the observed nebula emission of the ULX bubbles, is reported to 1.3 × 1041 erg s−1 (Belfiore et al. 2020), and is comparable to for the higher-mass accretion rate model (B3e10D1), ∼8 × 1040 erg s−1. In addition, the isotropic luminosity in model B3e10D1 at θ ∼ 30° or 150° is almost the same as the observed X-ray luminosity, (Israel et al. 2017a), where the isotropic luminosity is calculated as . Thus, NGC 5907 ULX1 probably has a mass accretion rate of and is viewed from the polar angle 20°–30° away from the z-axis. The magnetic field strength of the NS is stronger than 1010 G. This is because and of model B1e10D1_np are almost the same as those of model B3e10D1, but no accretion column appears in this model.
Also, for our slightly lower-mass accretion models (B1e10D01, B3e10D01, and B3e10D02) are comparable to in NGC 1313 X-2, ∼1040 erg s−1 (Pakull & Grisé 2008). In these models, the viewing angle, at which LISO is consistent with , is about 30°–40° measured from the z-axis. Therefore, NGC 1313 X-2 would be almost a face-on object with a mass accretion rate of several . Since no accretion column is formed in model B3e9D01_np, of which the kinetic and isotropic luminosity are similar to those of the above three models, BNS would be stronger than 3.3 × 109 G.
According to our numerical models, both NGC 5907 ULX1 and NGC 1313 X-2 would also have effectively optically thick outflows with rbb > 100 km. Indeed, the thermal emission, probably originated from outflows, was reported in NGC 1313 X-2 (Qiu & Feng 2021). However, such emission is not detected in NGC 5907 ULX1. This inconsistency can be resolved by conducting simulations of a large rM. If BNS is so strong that rM is comparable to rsph, then the outflow rate is thought to decrease, and thermal emission is not detected. In this case, 1/b decreases as discussed in Section 4.1, and the observer's viewing angle should be smaller than that estimated above for to be consistent with . However, increasing BNS may change the intrinsic luminosity, and therefore simulations with a high BNS are necessary to perform. Postprocessing radiative transfer simulations are needed to obtain detailed radiation spectra and compare the observation (Kawashima et al. 2012; Kitaki et al. 2017; Narayan et al. 2017).
On the other hand, no nebula has been detected around Swift J0243.6+6124. The reason for this is that the object is a transient source. In order to form a nebula extending over 100 pc, an energy injection, ∼1040 erg s−1, from the central object is needed to last for at least 105 yr (Pakull & Grisé 2008; Belfiore et al. 2020).
4.3. Ultraluminous Supersoft Sources
Our simulations are consistent with the hypothesis proposed by Urquhart & Soria (2016) in which ULXs are observed as ultraluminous supersoft sources (ULSs) by an edge-on observer, as far as the angular range of τeff > 1 and its accretion rate dependence are concerned. According to their model, the optical depth of the outflows τeff is larger than unity at 35° < θ < 145° for , and the polar angle at which τeff > 1 widens as the mass accretion rate increases. These features are also obtained in our simulations. Indeed, simulated outflows are effectively optically thick at polar angles > 40° away from the z-axis for models B1e10D01 and B3e10D01 (see Figures 6 and 7 for details). The mass accretion rates in these models are and , respectively (Table 2). This angular range and the mass accretion rates do not so contradict the suggestion by Urquhart & Soria (2016) mentioned above. Other models, in which the accretion rate is and the accretion columns form, show approximately the same results. In addition, the angular range where τeff > 1 widens with an increase in the mass accretion rate (see Figure 7). However, the blackbody temperature estimated from the simulation, ∼107 K, is higher than that observed in the ULSs, ∼106 K. Such a discrepancy might be resolved by the simulations with the initial torus located far away since the winds launched from the outer part of the accretion disks work to expand the photosphere and lower the temperature there (we will discuss this point later). If this is the case, our simulations are consistent with the hypothesis in which ULSs are edge-on sources of ULXs.
4.4. Why Does the One-side Accretion Occur?
The accretion to only one of the poles (one-side accretion) occurs in models with a relatively large truncation radius (B3e9D001, B1e10D001, B1e10D002, B3e10D001, B3e10D01, and B3e10D02) since the distorted magnetic field lines inhibit the gas accretion onto the opposite side. Initially, the shape of the NS magnetic field is symmetric with respect to the equatorial plane. Then, the magnetic field on the equatorial plane is perpendicular to the equatorial plane. However, when the gas accumulates at the truncation radius and flows toward either pole, the magnetic field lines are distorted (see magnetic field lines in the left panel of Figure 3). The magnetic field on the equatorial plane at the truncation radius will be tilted with respect to the equatorial plane. Since the gas tends to move to the side with the lower gravitational potential, the gas flows in the direction in which the preceding gas accretes. The accretion column on the opposite side is less likely to be formed. This one-side accretion is more pronounced for the model with a large truncation radius (B3e9D001, B1e10D001, B1e10D002, B3e10D001, B3e10D01, and B3e10D02). In contrast, for the case where the truncation radius is very small, the gas on the surface of the thick disk accretes to the opposite pole, forming two accretion columns (models B1e10D01, B1e10D004, B3e10D04, and B3e10D1). The one-side accretion described above was also shown by some MHD simulations (see, e.g., Lii et al. 2014; Takasao et al. 2022). We discuss the importance of three-dimensional simulations to study the detailed structure of accretion flow in the subsequent section.
4.5. Future Issues
Three-dimensional simulations of nonaxisymmetric accretion flows onto a rotating NS have yet to be performed. In this study, we assume an axisymmetric structure, in which the magnetic axis of the NS coincides with the rotation axis of the accretion disk, and perform two-dimensional simulations. However, the gas is thought to accrete onto the NS surface through multiple streams due to the magnetic Rayleigh–Taylor or interchange instability (Kulkarni & Romanova 2008). Furthermore, the pulse emission is produced by the misalignment between the magnetic axis and rotation axis of the NS. Thus, we need to treat a nonaxisymmetric structure with three-dimensional simulations. The misalignment of the two axes might hinder the one-side accretion since the gas in the accretion disk would preferentially accrete to the closest pole (see, e.g., Romanova et al. 2003). In addition, if the NS rotates as fast as millisecond pulsars, the outflow rate might increase due to the effective acceleration by a rapidly rotating magnetosphere of the NS (e.g., Lovelace et al. 1999). Since the size of the photosphere increases as the mass outflow rate increases, the gas temperature decreases. As a result, rbb increases, and BNS is larger than that estimated in Figure 8. Simulations taking account of the rotation of the NS have been performed by Parfrey & Tchekhovskoy (2017) and Das et al. (2022).
In our simulations, the gas–radiation interaction is switched off in the very vicinity of the rotation axis near the NS where σ > σrad (see Section 2.3). This treatment would not affect the resulting size and temperature of the photosphere in Section 3.2. In the present simulations, the radiatively driven outflows are launched from the accretion column base as well as the accretion disk (see also Abolmasov & Lipunova 2023). However, the opacity of free–free absorption of such outflows is quite small due to the low density and high temperature. This leads to the smaller size of the photosphere for θ ≲ π/6, 5π/6 ≲ θ compared with that formed by outflows from the accretion disk (rth ∼ several 100 km; see the red lines in Figure 6). If we do not ignore the gas–radiation interaction near the rotation axis, the outflowing gas launched from the accretion column is more effectively accelerated by the radiation force near the rotation axis. This further reduces the gas density and size of the photosphere. Therefore, neglecting the gas–radiation interaction in the vicinity of the rotation axis near the NS does not affect the size and temperature of the photosphere presented in Section 3.2.
We need to investigate the influence of the boundary condition. In this study, we assumed that the kinetic and thermal energy reaching the inner boundary (NS surface) is immediately converted into radiative energy. Under this boundary condition, the energy carried to the NS core is ignored. In reality, a fraction of the energy reaching the NS surface is converted into the thermal energy of the NS atmosphere. The heat flux from the atmosphere to the core occurs due to thermal conduction (see Wijnands et al. 2017, and references therein). Then, the radiative energy generated on the NS surface might decrease, leading to a decrease in radiative luminosity. Also, the accretion flow geometry might change if we set different boundary conditions (see Figure 9 in Appendix A for details). Investigation of how the accretion flow depends on the boundary conditions is left as future work.
GR-RMHD simulations of large-scale accretion flows are also needed to evaluate rbb and Tbb, more accurately. It has been pointed out that the simulations in which the initial torus is located inside the spherization radius overestimate the mass outflow rate (see Kitaki et al. 2021; Yoshioka et al. 2022). Therefore, the accretion rate to reproduce effectively optically thick outflows would be higher than that estimated in the present study. This point should be clarified in long-term simulations with the initial torus placed far enough away from the spherization radius. In such simulations, the disk wind is expected to blow from the outer region of the accretion disks (r ≲ rsph). The blackbody radius would be larger, and the blackbody temperature would be smaller in the region near the equatorial plane. If this is the case, edge-on observers might identify the objects as the ULSs.
In the present study, we adopt a relatively weak magnetic field of the NS 3.3 × 109−10 G compared to the typical X-ray pulsar 1011–13 G. As BNS increases while the accretion rate remains fixed, the amount of the outflowing gas probably reduces since the area where radiatively driven outflows mainly occur, , decreases. This leads the photosphere to be small and the gas temperature at the photosphere to be high. It is expected that the blackbody radius tends to decrease with an increase in BNS. In this case, the magnetic field strength of Swift J0243.6+6124 estimated in Section 3.3 would decrease. Although rbb may depend on BNS as described above, rbb is a function of mass accretion rate only (see Table 1 and Equation (38)) in our results since simulations of rM ≪ rsph are performed. In addition, it has been reported that when BNS is strong enough that rM > rsph, rM is no longer dependent on BNS (Chashkina et al. 2017). We need to perform simulations for the strong BNS case.
The M1 closure, which is employed in the present study, leads to unphysical solutions when the system is optically thin and the radiation fields are anisotropic (see, e.g., Ohsuga & Takahashi 2016). In our simulations, especially in the models where the one-side accretion occurs, the radiation fields are quite anisotropic. In order to accurately calculate the radiation fields, we have to solve the radiative transfer equation. Such radiation MHD simulations, in which the radiative transfer equation is solved, are performed by Jiang et al. (2014), Ohsuga & Takahashi (2016), Asahina et al. (2020), Zhang et al. (2022), and Asahina & Ohsuga (2022).
The modeling of Swift J0243.6+6124 considering multipolar magnetic field components is to be further investigated. Although we constrain the dipole magnetic field strength at the NS surface, 3 × 1011 G < BNS < 4 × 1012 G, in the present work (see also Tsygankov et al. 2018; Doroshenko et al. 2020), a cyclotron resonance scattering feature (CRSF) corresponding to 1.6 × 1013 G was reported by Kong et al. (2022). This discrepancy is resolved if CRSF originates from multipole magnetic fields, as has already been suggested by some authors (see, e.g., Israel et al. 2017a). In the case that the multipole magnetic field component is dominant over the dipole component, the accretion flow geometry is expected to be more complex (Long et al. 2007; Das et al. 2022), and the position of the magnetospheric radius changes. We plan to perform the GR-RMHD simulations of super-Eddington accretion flows onto the NS with multipole magnetic fields.
Our models might exhibit a high polarization degree, as recently observed in the X-ray binary (Ratheesh et al. 2023; Veledina et al. 2023). The hard X-ray photons produced in the accretion column base are scattered by the inner wall of the funnel, which consists of the effectively optically thick outflows, and they pass through the low-density region near the rotation axis. If such photons are observed, a high polarization degree might be detected. Indeed, Ratheesh et al. (2023) reported that the polarization degree increases from approximately 6% at 2 keV to 10% at 8 keV. Postprocess polarized radiative transfer simulations are needed to compare our models with such observations.
5. Conclusions
We performed two-dimensional axisymmetric GR-RMHD simulations of super-Eddington flows around NSs with a dipole magnetic field for modeling the Galactic ULXP, Swift J0243.6+6124. In our simulations, the accretion columns near the magnetic poles, the super-Eddington disk outside the magnetospheric radius, and outflows launched from the disk appear. If the magnetospheric radius (truncation radius) is smaller than the spherization radius, the outflows are generated since the radiation force mainly drives the outflows inside the spherization radius. When the accretion rate is large enough while satisfying the above condition, effectively optically thick outflows are launched from the disk and would be responsible for the blackbody radiation. The blackbody temperature of the effectively thick outflow, ∼107 K, is roughly consistent with the observations by Tao et al. (2019). The blackbody radius increases with an increase in the mass accretion rate and agrees with observations, 100–500 km, when the accretion rate is about . Since the blackbody radiation was detected in two observations with of −2.22 × 10−8 s s−1 and −1.75 × 10−8 s s−1, but not in another with , the surface magnetic field strength of the NS in Swift J0243.6+6124 is limited to be 3 × 1011 G ≲ BNS ≲ 4 × 1012 G. The accretion rate is evaluated to be when the blackbody radiation is detected and when the blackbody radiation is not observed. Our results support the hypothesis that the super-Eddington phase in the 2017–2018 giant outburst of Swift J0243.6+6124 is powered by super-Eddington accretion flows onto a magnetized NS.
Acknowledgments
We would like to thank the anonymous reviewer for providing fruitful comments. This work was supported by JSPS KAKENHI grant Nos. JP21J21040 (A.I.), JP21H04488, JP18K03710 (K.O.), JP20K11851, JP20H01941, JP20H00156 (H.R.T) and JP18K13591 (Y.A.). A part of this research has been funded by the MEXT as "Program for Promoting Researches on the Supercomputer Fugaku" (Toward a unified view of the universe: from large-scale structures to planets, JPMXP1020200109; K.O., H.R.T., Y.A., and A.I.), and by Joint Institute for Computational Fundamental Science (JICFuS; K.O.). Numerical computations were performed with computational resources provided by the Multidisciplinary Cooperative Research Program in the Center for Computational Sciences, University of Tsukuba, Oakforest-PACS operated by the Joint Center for Advanced High-Performance Computing (JCAHPC), Cray XC 50 at the Center for Computational Astrophysics (CfCA) of the National Astronomical Observatory of Japan (NAOJ), the FUJITSU Supercomputer PRIMEHPC FX1000 and FUJITSU Server PRIMERGY GX2570 (Wisteria/BDEC-01) at the Information Technology Center, The University of Tokyo.
Appendix A: Dependence of the Accretion Flows on the Boundary Conditions
Here, we show the dependence of the accretion flows on the inner boundary condition of the radiative flux. Figure 9 shows the time evolution of the gas density and radiation energy density with different boundary conditions. The top panel is the result of model Be10D01 where the radiative flux at the NS surface is set to zero. On the other hand, the results of the model in which the radiative flux is set according to the free boundary condition are presented in the bottom panel (free boundary model). Cyan lines show the magnetic field lines. We can see that the low-density void region is formed between two accretion columns (10 km < r < 13 km and 50° < θ < 130°). A part of the photons emitted from the accretion columns is transported to the void region. These photons push the gas outward and the void suffers from expansion (see the top-right panel of Figure 9). This void repeatedly contracts and expands. The period of repeated expansion is about ∼500tg ≃ 10−3 s. However, the variation in the mass accretion rate at the NS surface due to the expansion is negligibly small. Such a structure cannot be seen in the free boundary model. This is because the photon can freely penetrate the NS surface, and the outward radiation force becomes considerably small compared with model B1e10D01.
Download figure:
Standard image High-resolution imageWe also find that the thickness of the accretion column in the free boundary model is thinner than that of model B1e10D01. It may originate from the decrease in the radiation energy of the accretion column due to the free boundary condition. It can be seen that the radiation energy at the base of the accretion column in model B1e10D01 is about 100 times larger than that in the free boundary model.
Appendix B: Analytic Formula of Magnetospheric Radius and Spin-up Rate
Takahashi & Ohsuga (2017) derived the analytical formulas of the magnetospheric radius of Equation (24), and spin-up rate of Equation (27), applying the self-similar solutions for the slim disk proposed by Watarai & Fukue (1999). Here we summarize the derivations of these formulas.
In the slim disk model, the energy balances are (Narayan & Yi 1994), where is the advected energy, is the viscous-dissipated energy, is the radiative cooling energy, and f represents the fraction of the advective cooling to the viscous-dissipated energy. Using the height-integrated radiation pressure described in Equation (12) in Watarai & Fukue (1999) and the scale height of the accretion disk H, presented in Equation (11) in Watarai & Fukue (1999), we get
where α is the viscous parameter (Shakura & Sunyaev 1973), c1 and c3 are the constants depending on Γ, f, and α (see Watarai & Fukue 1999, for details). In this study, we take Γ = 1.5, f = 0.1 for the analytical formula. The reason why we do not employ Γ = 5/3 is that the rotational velocity of the disk is zero with Γ = 5/3 (Narayan & Yi 1994). The magnetic pressure at radius r by the dipole magnetic field is . The magnetospheric radius rM is defined as the radius where pmag balances with prad:
Next we give the derivation of the spin-up rate. The spin-up rate can be calculated by assuming that the Keplerian angular momentum at rM is transported to the NS without dissipation. In this case, spin-up rate is represented by (Shapiro & Teukolsky 1986), where is the specific Keplerian angular momentum. Using Equation (B2), the spin-up can be calculated as follows:
Appendix C: Equations of Motion in General Relativity
Here, we derive the equations of motion, which are used in this study, according to Mihalas & Mihalas (1984). Equation (2) can be transformed to
Subtracting the time component of Equation (C1) multiplied by ui /ut from the spatial component of Equation (C1), the following relation can be obtained:
We define the relativistic enthalpy, w = ρ + pgas + e + b2, and Equation (C2) can be written as
where is the orthonormal tetrad that transforms vectors in the laboratory frame (e.g., Boyer–Lindquist coordinate) to the zero angular momentum observer frame (ZAMO frame; Bardeen et al. 1972), is the four-velocity of the gas in the ZAMO frame, is the inverse matrix of , and d/d τ = uμ ∂μ is the derivative of proper time, τ (Mihalas & Mihalas 1984). Now using Equations (C2) and (C3), we get
We calculate the equation of motion in the r-direction. In order to obtain a concrete expression, it is necessary to fix the spacetime. We consider the Boyer–Lindquist metric with a spin parameter of zero. In this case, all nondiagonal components of are zero. Assuming steady (∂t = 0) and axisymmetric (∂ϕ = 0) flow, the following equation of motion in the radial direction is obtained:
where , , , , , , and describe the gravitational force, thermal force, centrifugal force due to the poloidal motion, centrifugal force due to the toroidal motion, radiation force, Lorentz force, and advection force, respectively. These forces can be described as follows:
where γ = u(t) is the Lorentz factor in the ZAMO frame (i.e., statistic observer frame), and v(r) = u(r)/r(t). The equation of motion in the θ-direction can also be obtained by the same procedure.
where is the inertial force caused by the motion in the direction of θ. These forces can be written as follows:
where v(θ)=u(θ)/u(t). The relativistic equations of motion derived by this method reduce to the Newtonian equations of motion for , , and.