This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.

Articles

COMPTONIZED PHOTON SPECTRA OF SUPERCRITICAL BLACK HOLE ACCRETION FLOWS WITH APPLICATION TO ULTRALUMINOUS X-RAY SOURCES

, , , , , and

Published 2012 May 22 © 2012. The American Astronomical Society. All rights reserved.
, , Citation T. Kawashima et al 2012 ApJ 752 18 DOI 10.1088/0004-637X/752/1/18

0004-637X/752/1/18

ABSTRACT

Radiation spectra of supercritical black hole accretion flows are computed using a Monte Carlo method by post-processing the results of axisymmetric radiation hydrodynamic simulations. We take into account thermal/bulk Comptonization, free–free absorption, and photon trapping. We found that a shock-heated region (∼108 K) appears at the funnel wall near the black hole where the supersonic inflow is reflected by the centrifugal barrier of the potential. Both thermal and bulk Comptonization significantly harden photon spectra although most of the photons upscattered above 40 keV are swallowed by the black hole due to the photon trapping. When the accretion rate onto the black hole is ${\dot{M}}$  ≈ 200LE/c2, where LE is the Eddington luminosity, the spectrum has a power-law component which extends up to ∼10 keV by upscattering of photons in the shock-heated region. In higher mass accretion rates, the spectra roll over around 5 keV due to downscattering of the photons by cool electrons in the dense outflow surrounding the jet. Our results are consistent with the spectral features of ultraluminous X-ray sources, which typically show either a hard power-law component extending up to 10 keV or a rollover around 5 keV. We found that the spectrum of NGC 1313 X-2 is quite similar to the spectrum numerically obtained for high accretion rate (${\dot{M}} \,{\approx}\, 1000L_{\rm E}/c^{2}$) source observed with low viewing angle (i = 10°–20°). Our numerical results also demonstrate that the face-on luminosity of supercritically accreting stellar mass black holes (10 M) can significantly exceed 1040  erg s−1.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Ultraluminous X-ray sources (ULXs) are bright X-ray sources found in off-center regions of nearby galaxies. The typical luminosity of ULXs is 1039–1041 erg s−1, which exceeds the Eddington luminosity for neutron stars and stellar mass black holes. Such large luminosities can be explained either by subcritical accretion (i.e., accretion rates lower than the Eddington accretion rate) onto an intermediate-mass black hole (IMBH; Miller et al. 2003, 2004; Cropper et al. 2004; Strohmayer & Mushotzky 2009) or supercritical accretion onto a stellar mass black hole (King et al. 2001; Ebisawa et al. 2003; Okajima et al. 2006; Tsunoda et al. 2006; Vierdayanti et al. 2006, 2008; Poutanen et al. 2007; Berghea et al. 2008). Studying the photon spectra of supercritical accretion flows may give a clue to determine the mass of ULXs.

X-ray observations revealed that spectra of ULXs often accompany hard X-ray emission. Gladstone et al. (2009) showed that they are well fitted by an energetically coupled disk and Comptonized corona model, assuming a stellar mass black hole. However, it is not clear whether the geometry and energy state of the coronae assumed in their spectral fitting models are realized in supercritical accretion flows. Theoretical studies indicate that advection-dominated super-Eddington accretion flows, i.e., slim disks, are formed when the luminosity exceeds the Eddington luminosity (Abramowicz et al. 1988; Watarai et al. 2000, see also Kato et al. 2008, and references therein). Since the spectra of slim disks are dominated by a thermal disk component, we should explain how the power-law components are produced in supercritical accretion flows.

The dynamics of supercritical accretion flows have been studied by radiation hydrodynamic simulations (Eggum et al. 1988; Okuda 2002; Ohsuga et al. 2005; Ohsuga 2006) and radiation magnetohydrodynamic (MHD) simulations (Ohsuga & Mineshige 2011; Ohsuga et al. 2009). Kawashima et al. (2009) performed axisymmetric two-dimensional radiation hydrodynamic simulations of supercritical black hole accretion flows incorporating the Compton cooling/heating. They pointed out that a mildly hot (∼107.5–108 K) outflow can upscatter the soft photons emitted by the disk via a thermal Compton effect. Bulk Compton scattering (Blandford & Payne 1981) can also harden the spectra of black hole accretion flows. For subcritical accretion flows, Laurent & Titarchuk (2011) showed that the bulk Comptonization produces hard X-ray power-law components. However, their models did not take into account the disk dynamics and the inflow and outflow components.

In this paper, we report the results of Monte Carlo calculations of photon spectra of supercritical black hole accretion flows obtained by post-processing the radiation hydrodynamic simulation results. This paper is organized as follows. In Section 2, we describe the numerical method for radiation hydrodynamic simulations and calculations of photon spectra by using the Monte Carlo method. In Section 3, we present the results for the computation of photon spectra. In Section 4, we summarize the structure of supercritical accretion flows and the photon trajectory, and compare the computed spectra with those of ULXs. The details of numerical methods for radiation hydrodynamical simulation are summarized in Appendix A, and the modification of the optical depth due to fluid motions and thermal motions of electrons, which we applied in Monte Carlo calculations, is summarized in Appendix B.

2. NUMERICAL METHOD

2.1. Radiation Hydrodynamic Simulations of Supercritical Accretion Flows Incorporating Compton Cooling/Heating

We solve the axisymmetric two-dimensional radiation hydrodynamic equations in spherical coordinates (r, θ, ϕ) by adopting the flux-limited diffusion approximation (Levermore & Pomraning 1981; Turner & Stone 2001). General relativistic effects are incorporated by a pseudoNewtonian potential (Paczyńsky & Wiita 1980), Ψ = −GM/(rrs), where rs(= 2GM/c2) is the Schwarzschild radius, G is the gravitational constant, M = 10 M is the mass of the central black hole, and c is the speed of light. We adopt the α prescription of viscosity (Shakura & Sunyaev 1973) and set α = 0.1. Basic equations and numerical methods to solve the radiation hydrodynamic equations are the same as in Kawashima et al. (2009), and they are summarized in Appendix A.

We improve on the investigation by Kawashima et al. (2009) by using a larger computational domain 2rsr ⩽ 1000rs and 0 ⩽ θ ⩽ π/2, and a higher resolution (Nr, Nθ) = (192, 192). We assume azimuthal symmetry and mirror symmetry with respect to the equatorial plane. The grid points in radial and polar directions are distributed such that Δln r = constant and Δcos θ = 1/Nθ, respectively. We start the calculations with a hot, rarefied, optically thin atmosphere (i.e., without an initial optically thick disk). In order to simulate the formation and the evolution of the accretion disks, we continuously supply mass through the outer boundary at r = 1000rs near the equatorial plane (0.45π  ⩽  θ  ⩽  0.5π) with a constant rate ${\dot{M}}_{\rm input}$. The injected matter is assumed to have a specific angular momentum corresponding to the Keplerian angular momentum at r = 300rs. Viscous processes allow the angular momentum of the gas to be transported outward, which drives the gas inflow. Eventually, the gas falls onto the black hole in a quasi-steady fashion. At the outer boundary except for the mass injection region (i.e., r = 1000rs and 0  ⩽  θ  <  0.45π), we allow matter to escape freely but not to enter the computational domain. We further adopt an absorbing boundary condition at r = 2rs. We note that the initial hot atmosphere does not affect the results of our calculations because it is swept away from the computational domain by the accretion flow and the outflow.

The left panel of Figure 1 shows the time-averaged distribution of mass density and temperature for a model with accretion rate ${\dot{M}} \,{\approx}\, 200L_{\rm E}/c^{2}$, where LE is the Eddington luminosity and ${\dot{M}}$ is measured at r = 2rs. The supercritical accretion flow produces four regions: (1) a radiation pressure dominant accretion disk, (2) a sub-relativistic, mildly hot (Tgas ∼ 107.5–108 K) funnel jet around the rotation axis, (3) a cool (Tgas ∼ 106.5 K), dense (ρ  ∼ 10−6 g cm−3), and slow (≲ 0.01c) outflow emerging from the disk, and (4) a horn-like hot region around the funnel wall near the black hole (r  ≲  5rs), where the gas temperature increases to Tgas  ∼  108 K due to shock heating by supersonic inflow bounced by the centrifugal barrier of the potential (the right panel of Figure 1). The reflection shocks are identified by entropy jumps and kinks of the streamlines. Hot regions are also created in the funnel owing to the heating by internal shocks.

Figure 1.

Figure 1. (a) Gas temperature and mass density distributions obtained by time-averaging the results of the axisymmetric two-dimensional radiation hydrodynamic simulation in quasi-steady states for models with ${\dot{M}} \,{\approx}\, 200L_{\rm E}/c^2$. Arrows represent the poloidal velocities of the outflows, with their length being proportional to log vp when poloidal speed vp > 10−5c. The black square (−300rsx, y, z ⩽ 300rs) indicates the region where the photon transfer is computed. (b) The snapshot of gas temperature and density distributions near the black hole at t = 52.5 s when the accretion flow is quasi-steady.

Standard image High-resolution image

2.2. Monte Carlo Calculations of Radiative Transfer

The photon spectra are computed by solving the three-dimensional radiative transfer using the Monte Carlo method (e.g., Pozdnyakov et al. 1983), by post-processing the results of radiation hydrodynamic simulations. Since non-gray radiation hydrodynamic simulations fully coupled with the Monte Carlo radiative transfer require huge computational resources, we calculate photon spectra by post-processing the simulation results. Here, we should emphasize that our radiation hydrodynamic calculations incorporate the effects of Comptonization (i.e., Compton cooling/heating and subsequent change in fluid dynamics) by applying the Kompaneets equation and by solving Equations (A1)–(A6) in Appendix A. Although the Compton cooling/heating rate used in our radiation hydrodynamic simulations might not be fully consistent with the results of full radiative transfer, our spectral calculations do not significantly overestimate the effects of Comptonization on spectra compared to the spectral calculations using the simulation results obtained without taking into account the Compton cooling.

We ignored the curvature of photon trajectories and gravitational redshift because most of the photons affected by these general relativistic effects are trapped by the inflowing plasma, and swallowed by the black hole, as we shall show in Section 3.

We adopt Cartesian coordinates to calculate the radiative transfer in the region −300rsx, y, z  ⩽ 300rs (depicted by the square box in Figure 1(a)). The grid points are equally spaced with a number of grid points of (Nx, Ny, Nz) = (160, 160, 160). The physical quantities at each grid point are obtained by interpolating the time-averaged simulation results. We averaged the quantities at each grid points for 5 s after accretion flows become quasi-steady. We employed the model for the mass input rate from the outer boundary ${\dot{M}_{\rm input}} =$ 10,000, 3000, and 1000LE/c2, whose eventual mass accretion rates onto the black hole are ${\dot{M}}$  ≈ 1000, 500, and 200LE/c2, respectively.

We calculate radiation spectra by Monte Carlo simulations, incorporating emission, absorption, and the scattering by the following algorithm.

  • 1.  
    In order to limit the emission point to avoid too many scatterings in the optically thick region, we identify the surface where the effective optical depth ${\tau }_{\rm eff}={\sqrt{3{\tau _{\rm a}}({\tau }_{\rm a} + {\tau }_{\rm e})}}$ equals 10. Here, τa and τe are the optical depth for the absorption and the Thomson scattering, respectively, calculated by integrating the opacity in the −z-direction starting from the upper boundary at z = 300rs.The absorption coefficient is evaluated by free–free absorption coefficient αffν obtained from Kirchhoff's law εffν/4π = ανffBν, where εffν is the emissivity of free–free emission (i.e., bremsstrahlung) and Bν is the intensity of the blackbody radiation at frequency ν. We consider the emissivity of the electron–electron and the electron–proton bremsstrahlung (see Manmoto et al. 1997). For Thomson scattering, we adopted αesν = neσT, where ne is the number density of the plasma and σT is the cross-section of the Thomson scattering. Since the aim of calculating τeff in this calculation step is just restricting the region for emissions of seed photons, for simplicity, we used the cross-section for Thomson scattering and took into account the corrections for αffν and αesν by fluid motions only (see Appendix B). As we explain later, when we calculate the escape probability in order to trace photon packets, we calculate scattering coefficients by using the Klein–Nishina formula taking into account the correction due to bulk motions and thermal motions of electrons. We carried out calculations limiting the emission region of seed photon packets above the surface τeff = 1, 3, 5, 10, 20 and confirmed that the spectral shape of the photons when τeff = 10 is essentially the same as that obtained by employing the emission region deeper than τeff = 10.It should be noted that the surface where τeff = 1 is located inside the disk when the distance from the rotation axis $R= \sqrt{x^{2} + y^{2}}$ satisfies R ≲ (several)× 10rs, while it is found in the dense outflow in the outer disk. In our calculation, therefore, seed photons created in the inner disk and in the outflow covering the outer disk mainly contribute to the observable spectra.
  • 2.  
    We place a seed photon packet above the surface τeff = 10. The point where the seed photon is emitted is determined such that the emission probability at each grid cell is proportional to the emissivity, by using the inverse function method for a function defined as
    Equation (1)
    where n and ntotal are the sequential number and the total number of grid points which satisfy the condition τeff ⩽ 10, respectively. The emissivity of the seed photon εffν is given by the thermal bremsstrahlung emission. The intensity of thermal bremsstrahlung emission approaches that of a blackbody deep inside the disk or in the dense outflow, where τeff significantly exceeds 1. We ignored the cyclotron and synchrotron emissivity. This is partly because our radiation hydrodynamic simulation does not include magnetic fields, and partly because the disk or outflow is dense and hot enough to attain blackbody intensity via only the thermal bremsstrahlung emission/absorption in the soft X-ray band (see Section 4.4 for the evaluation of cyclotron emissivity).The direction of propagation of seed photons is determined by using a uniform random number, assuming isotropic emission in the fluid rest frame. After the direction is determined, the frequency and the direction of each seed photon are converted to those for the observer frame via the Lorentz transformation, by using the fluid velocity at the point where the seed photon is generated.
  • 3.  
    We solve the photon transport problem taking into account free–free absorption, photon trapping (Begelman 1978; Ohsuga et al. 2002, 2003), and thermal/bulk Comptonization.

Each photon packet is given an initial weight w = 1. Along the photon trajectory between scatterings, the weight of the photon packet is reduced by the escape to the observer, by being swallowed by the black hole, and by free–free absorption. We note that the word "escape" is used in referring to "leak" in this paper. Due to the escape from the calculation box, the weight of photon packets is reduced for each scattering by a fraction 1 − Pesc, where Pesc  ≡  exp (− τesc) is the escape probability (i.e., the probability that photons will escape to observer or will be swallowed by the black hole without scattering or absorption). The optical depth τesc is obtained by integrating the opacity of the free–free absorption and the electron scattering from the current position of the photon packet to the outer boundary or to the inner boundary at r = 2rs, along its current direction. When the ray crosses the inner boundary, the fraction Pescw of the photons is regarded as swallowed by the black hole. This mimics the effects of photon trapping because we take into account the beaming of photons due to the fluid motion. In the other cases, the fraction Pescw of the photons is summed up as observable photons. In addition to the reduction due to the escape of photons, the free–free absorption decreases the weight of photon packets by a fraction 1 − Pabs, where Pabs is the absorption probability determined by the optical thickness for free–free absorption between the current and the next scattering positions, that is, Pabs = 1 − exp (− τabs).

In order to obtain Pesc, we calculate τesc determined by

Equation (2)

where k is the sequential number of the grid cells along the ray counted from the current position of the photon packet, N is the total number of grid cells which the ray crosses, and Δτ(k) is the optical thickness of the kth cell computed by integrating the effective extinction coefficient αν, eff = αffν, eff + αKNν, eff along the segment of the ray inside the kth grid cell, where αKNν, eff is the effective scattering coefficient corresponding to the Klein–Nishina cross-section which is modified by the fluid motions and thermal motions of electrons (see Appendix B for detail).

Here, αffν, eff is the effective absorption coefficient modified by fluid motions only, since the modification of absorption cross-section due to thermal motion of ions (or electrons for the electron–electron bremsstrahlung) is already taken into account when we obtain the absorption coefficient from emissivity by using Kirchhoff's law. The extinction coefficients inside the cell are evaluated by linearly interpolating the coefficients at eight grid points at the corner of the cell.

By using the Pesc obtained by calculating τesc, we store the fraction of escaping photons in the packet as wPesc for each energy bin in order to calculate the observable spectra (or spectra of photons which are swallowed by the black hole), and reduce the weight of the photon packet remaining in the computational domain as w(1 − Pesc).

In order to calculate the position for the next scattering event and the reduction of the weight of the photon packet due to absorption, we store the optical depth for scatterings from the current position ${\bm x}(n=1)$ toward the kth cell τKN(k) = ∑kn = 1ΔτKN(n), that for absorption τabs(k) = ∑kn = 1Δτabs(n), and the position ${\bm x}(k)$ where the ray enters the k th cell. Here, ΔτKN(n) and Δτabs(n) are the optical thickness for scattering and the absorption of the nth cell computed by integrating αKNν, eff and αffν, eff.

The next scattering point is determined by computing the optical depth where the next scattering will occur, which is determined by ${\tau }_{\rm scat} = -{\ln } [1-{\xi } (1- e^{-{\tau }_{\rm esc}^{\rm KN}})]$. Here ξ is a uniform random number in the range (0, 1) and τKNesc is the optical depth for the Compton scattering between the current position of the photon and the boundary (i.e., τKNesc = τKN(N)). By finding the kth cell along the ray which satisfies τKN(k)  ⩽  τscat < τKN(k + 1), the next scattering point ${\bm x}_{\rm scat}$ is obtained by

Equation (3)

The optical depth for absorption between the current and the next scattering positions τabs is obtained from τabs(k), which is stored when we calculated τesc described above. We evaluate the absorption probability Pabs = 1 − exp (− τabs) and reduce the weight of the photon packet by a fraction (1 − Pabs). The weight of the photon packet w(1 − Pesc)(1 − Pabs) is available for the next scattering.

After the position of the next scattering is determined and the weight of photon packet remaining in the calculation box is evaluated, we calculate the energy shift of the photon packet due to thermal and bulk Compton scatterings. We obtain the electron temperature and the fluid velocity at that position by linear interpolation from those at the eight grid points at the corners of the cell including the scattering point. The details of the algorithm of the thermal Compton scattering are described in Pozdniakov et al. (1977) and Pozdnyakov et al. (1983). We incorporate the bulk Comptonization by transforming the photon four-momentum from the observer frame to the comoving frame of the fluid before the thermal Compton scattering. We convert the four-momentum into the observer frame after the Compton scattering.

For a frequency range of 1016 Hz < ν < 1022 Hz, we divide the photons into 50 frequency bins with Δlog ν = constant. We generate 4 × 105 seed photons for each frequency bin and apply the above method to calculate their trajectories. These steps are repeated for photons belonging to the same frequency bin. We obtain the spectral energy distribution (SED) by summing up the energy of the escaping photons multiplied by the photon weight.

3. RESULTS

Figure 2 shows the SEDs for models with accretion rates ${\dot{M}}$ ≈ 1000, 500, and 200 LE/c2. The solid curves display the isotropic luminosity obtained by summing up the escaping photons with viewing angle between 0° and 10°. The resulting SED for the lower mass accretion rate model (${\dot{M}} \,{\approx}\, 200 L_{\rm E}/c^2$) exhibits a power-law spectrum extending from ∼1 keV to ∼10 keV. In contrast, the spectra for higher mass accretion rate models (${\dot{M}} \,{\approx}\, 1000, 500 L_{\rm E}/c^2$) feature a spectrum with a rollover at ∼5 keV. The spectral shape below ∼1 keV is similar to that of the disk blackbody emission of a standard accretion disk model, i.e., νLν∝ν4/3. The power-law index for each model reproduces the typical photon index of the spectra of ULXs, νLν ∝ ν0.3. The isotropic luminosity for a model with ${\dot{M}}$ ≈ 500LE/c2 is higher than that for ${\dot{M}}$ ≈ 1000LE/c2 because the funnel becomes Thomson thick when ${\dot{M}}$ ≈ 1000LE/c2. Since the Thomson thick funnel prevents the photons from escaping to the observer without being scattered, the photons in the funnel diffuse out into the cool outflow, where they are downscattered. Moreover, the photons trapped by the inflow increase with ${\dot{M}}$. These are the reasons why the isotropic luminosity for ${\dot{M}}$  ≈ 1000LE/c2 is lower than that for ${\dot{M}}$ ≈ 500LE/c2.

Figure 2.

Figure 2. SEDs for models with mass accretion rates ${\dot{M}} \,{\approx}\, 1000$ (red), 500 (green), and 200LE/c2 (blue). The dotted line shows the disk blackbody spectrum of a standard thin accretion disk model, while the dashed one shows the typical photon index of the spectra of ULXs.

Standard image High-resolution image

Figure 3 displays the dependence of isotropic X-ray luminosity on viewing angle. The isotropic X-ray luminosity significantly exceeds the Eddington luminosity when the supercritical accretion flow is viewed by a nearly face-on observer. As we explained above, isotropic luminosity for ${\dot{M}}$  ≈ 1000LE/c2 is lower than that for ${\dot{M}}$  ≈ 500LE/c2, when the viewing angle is smaller. For the model with ${\dot{M}}$ ≈ 200LE/c2, the reduction of the isotropic luminosity from i = 0°–10° to i = 10°–20° is smaller than that for the higher mass accretion models, since the opening angle of the funnel is wider and the collimation of radiative flux is weaker.

Figure 3.

Figure 3. Dependence of isotropic X-ray luminosity (0.3–10 keV) on viewing angle i. The red, green, and blue lines display the isotropic luminosity for ${\dot{M}}$ ≈ 1000, 500, and 200LE/c2, respectively.

Standard image High-resolution image

Figure 4 shows the dependence of SEDs on viewing angle. When viewing angle is smaller, SEDs become harder, because the high-energy photons upscattered in the shock-heated region near the black hole escape to the observer through the funnel jet. For ${\dot{M}}$ ≈ 200LE/c2, the rollover at ∼5 keV appears for observers with larger viewing angle. These rollovers are formed by downscattering due to the thermal electrons in the cool outflow surrounding the funnel jet. Meanwhile the power law extends up to ∼10 keV for observers with smaller viewing angle.

Figure 4.

Figure 4. Dependence of SEDs on viewing angle i for the models with ${\dot{M}}$ ≈ 1000LE/c2 (upper panel) and ${\dot{M}}$ ≈ 200LE/c2 (lower panel).

Standard image High-resolution image

Figure 5 demonstrates the effects of Comptonization on the spectra for ${\dot{M}}$ ≈ 200LE/c2. The solid curve is the same as that in Figure 2. The dotted curve is that without Comptonization. The dashed curve shows the spectra obtained by calculations including the thermal Compton scattering only and without the bulk Comptonization. The SEDs become harder through the thermal Compton scattering in the shock-heated region at the funnel wall. Let us evaluate the Compton y-parameter in the shock-heated region. The Thomson optical thickness across the shock-heated region is τe ∼ 10, where we obtained τe by integrating the Thomson scattering coefficient along a segment between (R, z) = (3rs, 2rs) and (2rs, 3rs). The Compton y-parameter for thermal Comptonization in the shock-heated region is y = (4kBTe/mec2)max(τe, τ2e) ∼ 10, where kB is the Boltzmann constant, me is the electron mass, and Te (∼108 K) is the electron temperature. This value of y is large enough to form the Wien-like convex spectra in 5 keV  <  hν  <  15 keV of the dashed curve in Figure 5. We confirmed by tracing the trajectories of photons and by separating the contribution from photons generated at different radii that the Comptonized photon spectrum shown by the dashed curve in Figure 5 is mainly formed in the shock-heated region around the funnel wall.

Figure 5.

Figure 5. Effects of Comptonization on the SEDs for a model with ${\dot{M} = 200 L_{\rm E}/c^{2}}$. The dotted curve displays the spectrum for a face-on observer (0°–10°) without Compton scattering. The solid curve shows the spectrum obtained by including the bulk and the thermal Compton scattering. The dashed curve shows the spectrum obtained by calculations with the thermal Compton scattering only. The thin solid curve shows the spectrum of photons swallowed by the black hole (i.e., photons which enter the region r < 2rs) for calculations with the bulk and the thermal Compton scattering.

Standard image High-resolution image

The SED becomes harder when we take into account the bulk Comptonization (dot-dashed curve), although the spectral hardening is less significant than that by the thermal Compton scattering. We note that the bulk Compton scattering due to inflow scatters the photons toward the black hole, and most of the upscattered photons are swallowed. When the upscattered photons diffuse out into the funnel, they can escape.

The thin solid curve in Figure 5 shows the SEDs of the photons trapped by the inflow. For this calculation, both thermal and bulk Compton scatterings are included. Numerical results indicate that most of the Comptonized photons above 40 keV are swallowed by the black hole. Accordingly, the observable photons produce the power-law component whose photon number decreases around 40 keV.

4. DISCUSSION

4.1. Summary of Numerical Results

We have shown that the Comptonized SEDs are harder than those of a slim disk model (e.g., Watarai et al. 2005) by post-processing the results of the radiation hydrodynamic simulations of supercritical accretion onto a black hole. The X-ray spectrum of mildly supercritical accretion flows (${\dot{M}}$  ≈ 200 LE/c2) has a power-law component in 1–10 keV and a rollover appears around 5 keV in highly supercritical accretion flows (${\dot{M}}$ ≳ 500 LE/c2).

Figure 6 schematically shows the structure of Comptonizing supercritical accretion flows and the photon trajectories. Hard X-ray photons are produced near the black hole. Photons generated in the disk are scattered many times and are diffusively transported to the black hole because the beaming due to accreting gas motion helps photons to move toward the black hole. The sub-relativistic accretion near the black hole makes the photon frequency increase by the bulk Compton scattering. The photons are significantly upscattered by the thermal Compton scatterings at the shock-heated region. Some photons which are upscattered near the black hole are injected into the funnel jet. Some of them escape along the rotation axis. The others are scattered by electrons in the funnel jet and injected into the cool outflow surrounding the funnel jet. They are downscattered by the thermal electrons in the cool outflow. Some fraction of the downscattered photons escape to the observer. The soft photons (≲1 keV) are produced and self-Comptonized in the cool outflow. We note that most photons upscattered near the black hole are swallowed by it.

Figure 6.

Figure 6. Schematic picture of a supercritical accretion flow consisting of (1) a radiation pressure dominant accretion disk, (2) a sub-relativistic, mildly hot funnel jet, (3) a cool, dense, and slow outflow, and (4) a shock-heated region. The left panel displays the overall structure of a supercritical accretion flow obtained by our simulations. The right one shows the structure near the black hole (r ≲ 10rs). The shock-heated regions are formed because the accreting gas near the rotation axis is bounced by the centrifugal barrier. Photons are upscattered by electron scatterings in the converging inflow (bulk Compton scattering) and the thermal Comptonization by hot electrons in the shock-heated region.

Standard image High-resolution image

4.2. Spectral Softening by Cool Outflows

The spectral softening due to thermal electrons in the cool outflow becomes more significant when ${\dot{M}}$ is higher, because the narrower and denser funnel jet decreases the number of photons which escape without entering cool outflow surrounding the jet. When ${\dot{M}}$ ≈ 200 LE/c2, the SED is significantly hardened by thermal Comptonization in a shock-heated region as shown in Figure 5. In contrast, when the mass accretion rate is higher (${\dot{M}}$  ≈ 1000 LE/c2), the SED calculated by taking into account the thermal Comptonization only becomes softer than that calculated without Comptonization, because the downscatterings in the cool outflow overwhelm the upscatterings in the shock-heated region. When we take into account the bulk Comptonization by the sub-relativistic outflow, the outflow motion in the funnel assists photons to escape without significant downscattering and the SED is eventually hardened.

When the mass accretion rate is higher, our calculations in this paper show that SEDs become softer, although our previous paper (Kawashima et al. 2009) expected that SEDs become harder because of Compton upscattering by thermal electrons in the jet. Kawashima et al. (2009) evaluated the photon indices by calculating the Compton y-parameters in the region τeff ⩽ 1. The effective optical depth is calculated by integrating the free–free absorption and the scattering coefficients from the outer boundary at r = 500rs to the inner boundary at r = 3rs along the rays with polar angle θ = constant, for various θ. (Here, we note that the computational domain for radiation hydrodynamic simulation in Kawashima et al. 2009 was smaller than that in this work.) Since Kawashima et al. (2009) did not take into account the downscattering by cool outflows, they overestimated the spectral hardening when the accretion rate is high. When the mass accretion rate is higher, a greater number of escaping photons tends to enter into the cool outflow surrounding the jet and they are subsequently downscattered, because the funnel becomes narrower and jet becomes thicker for Thomson scatterings. This is the reason why the SEDs become softer when the mass accretion rate is higher, in this work. Nevertheless, the conclusion in Kawashima et al. (2009) that the spectral state changes from the slim disk one to the Comptonizing outflows is valid in highly supercritical accretion flows because the SED of the photons escaping from such highly supercritical flows is determined by the Compton downscatterings in the outflow.

4.3. Comparison with Observations of ULXs

Recent X-ray observatories (Suzaku, XMM-Newton, and Chandra) provide us with precise spectra of ULXs below 10 keV. In Figure 7, we compare the SEDs obtained in this work and the XMM-Newton pn data of three ULXs: NGC 1313 X-2, IC 342 X-1, and NGC 5204 X-1, which are kindly provided by J. C. Gladstone. The SEDs of our results are normalized by the X-ray luminosity shown in Figure 3, while the SEDs of ULXs are normalized by the X-ray luminosity shown in Gladstone et al. (2009). The SEDs obtained by our numerical simulations are similar to those of ULXs. The observed X-ray spectra of ULXs with photon index Γ ≲ 2 can be fitted well by the models with i ≲ 30°. On the other hand, for ULXs with photon index Γ ≳ 2, spectra with 20° ≲ i ≲ 50° are fitted well.

Figure 7.

Figure 7. Comparison of SEDs obtained from our radiation hydrodynamic simulations with those of ULXs. Top panel: the SED for the simulation model with ${\dot{M}}$ ≈ 1000LE/c2 and viewing angle i = 10°–20° and the SED of NGC 1313 X-2. Middle panel: ${\dot{M}}$ ≈ 200LE/c2 (i = 0°–10°) and IC 342 X-1. Bottom panel: ${\dot{M}}$ ≈ 200LE/c2 (i = 40°–50°) and NGC 5204 X-1. Red or blue solid curves display the SEDs calculated in this work. Black points with error bars show the XMM-Newton EPIC pn data corrected for absorption, which are shown in Gladstone et al. (2009) and are provided by J. C. Gladstone. Each SED is normalized by its isotropic X-ray luminosity.

Standard image High-resolution image

For IC 342 X-1 and NGC 5204 X-1, the soft X-ray component below 1 keV of our results is weaker than that of the ULXs. This is because the contribution from the cool outflow in the region R ≳ 100rs is underestimated because the temperature of the outflow might be too cool to significantly contribute to the SED at ≲1 keV or our computation box might be too small. We might be able to resolve this discrepancy by carrying out MHD simulations taking into account the dissipation of magnetic energy in the disk corona and in the outflow, which can heat the outflow. In addition, the magnetic fields affect the geometrical structure of jets. Takeuchi et al. (2010) performed axisymmetric two-dimensional radiation MHD simulations and found that the jet is collimated by magnetic stress, while it is accelerated by radiative force. Therefore, the structure of the jet and the outflow is affected by magnetohydrodynamical processes. It will be our future work to calculate photon spectra by using the results of radiation MHD simulations.

Let us compare the luminosity obtained from our simulations and ULXs. It should be noted that our simulation assumed a black hole with M = 10 M. The luminosity increases or decreases according to the black hole mass. The X-ray luminosity for the models with ${\dot{M}}$  ≈ 1000LE/c2 and i = 10°–20° is ∼1040 erg s−1 (Figure 7), while that of NGC 1313 X-2 evaluated in Gladstone et al. (2009) is ∼5 × 1039 erg s−1. Since the X-ray spectrum of NGC 1313 X-2 is well fitted by our simulation results with these parameters for the accretion rate and the viewing angle, the luminosity difference suggests that NGC 1313 X-2 may harbor a black hole whose mass is 5 M, since the luminosity of accretion flows is proportional to the mass of black holes when ${\dot{M}}/(L_{\rm E}/c^{2})$ is constant. On the other hand, the X-ray luminosity for the models with ${\dot{M}}$ ≈ 200LE/c2 and i = 40°–50° is ∼3 × 1039 erg s−1, while that of NGC 5204 X-1, whose spectrum in the 1–10 keV band is well fitted by this accretion rate and viewing angle in our simulations, is ∼5 × 1039 erg s−1. This indicates that NGC 5204 X-1 possesses a slightly massive black hole with mass M ∼ 20 M. We leave it as a future work to carry out radiation hydrodynamical (or radiation magnetohydrodynamical) simulations and spectral calculations for supercritical accretion onto a black hole whose mass is smaller or larger than 10 M.

Recently, T. Yoshida et al. (in preparation) confirmed that IC 342 X-1 shows power-law spectra with Γ ∼ 1.6–1.8 extending up to ∼20 keV during the lower luminosity phase in Suzaku observations, while it shows power-law spectra with photon index Γ ∼ 1.8–2.0 accompanying a rollover at ∼5 keV during the higher luminosity phase in some XMM-Newton observations. In our results, for viewing angle i = 10°–20°, the power-law component extends up to 10 keV in a lower ${\dot{M}}$ model, while the rollover at ∼5 keV appears in a higher ${\dot{M}}$ model. Furthermore, each photon index of power-law components obtained from our numerical simulations is consistent with that of X-ray observations. This suggests that supercritical accretion occurs in IC 342 X-1.

We note that the SEDs for the model with low mass accretion rate (${\dot{M}}$  ≈ 200LE/c2) are similar to those of galactic black hole candidates in the very high state, except that obtained SEDs possess a harder power-law component. Our calculations indicate that even when the ULXs show SEDs similar to that of the very high state, whose luminosity is L  ≲  LE, we cannot conclude that the ULXs are subcritically accreting IMBHs because supercritically accreting stellar mass black holes can also produce SEDs similar to the very high state.

4.4. Effects of Cyclotron/Synchrotron Emission

In this study, we ignored the cyclotron and synchrotron emission. As we mentioned in Section 2, this is partly because our radiation hydrodynamic simulation does not treat the effect of magnetic field, and partly because the disk (or the outflow covering the outer disk) is dense and hot enough to produce the radiation field reaching blackbody via only the bremsstrahlung emission processes in the soft X-ray band.

Here, we show that the cyclotron emission might not contribute to the X-ray spectra. If we apply Emag ∼ 10Egas from the radiation MHD simulations (Ohsuga & Mineshige 2011) to the result of our radiation hydrodynamic simulation, we obtain B ∼ 104 G in the funnel jet, where Emag and Egas are the magnetic and the gas energy density, respectively. The physical condition in the funnel jet (ρ ∼ 10−9g cm−3, Tgas ∼ 108 K, and B ∼104 G) is close to one examined by Takaraha & Tsuruta (1982), which shows that the emission of self-absorbed cyclotron higher harmonics becomes optically thin when ν  ≳ ν* = 3 × 1012 Hz in that condition, where ν* is the critical frequency. The self-absorbed cyclotron–synchrotron intensity has a peak in Iν = 2ν2*kBTgas/c2 at ν = ν* (∼3 × 1012 Hz, i.e., far-IR band). This peak intensity due to self-absorbed cyclotron emission from the jet is lower than that of the blackbody intensity in the UV or soft X-ray band arising from the disk (Tgas ∼ 107 K). Since the intensity due to cyclotron–synchrotron emission in the UV or soft X-ray band is much less than that of the peak intensity, it is also less than that of the blackbody emission from the disk. This suggests that the seed photons contributing to the Comptonized X-ray spectra are dominated by the bremsstrahlung emission from the disk rather than the cyclotron emission from the jet.

We note that the cyclotron emission from the shock-heated region is also too weak to affect the soft X-ray spectra. In the shock-heated region, since the estimated magnetic field is B ∼ 106 G and the critical frequency is ν*∼ 1014 Hz, i.e., near-IR band, the corresponding peak intensity is still lower than that of ∼107 K blackbody in the UV or soft X-ray band. It does not affect the X-ray spectra even if they are strongly self-Comptonized.

The cyclotron emissions and their Compton upscatterings are expected to be important in frequencies from the radio to the visible band. For the calculation of spectra from the radio to hard X-ray band including the cyclotron–synchrotron emission, it is important to carry out the radiation MHD simulations in order to obtain the structure of magnetized supercritical accretion flow. This remains as a future work.

4.5. General Relativistic Effects

Our spectral calculations do not incorporate the general relativistic effects. By the gravitational redshift, the energy of photons Comptonized up to the hard X-ray band by heated electrons at r  ≲  5rs will be decreased by 10%–30% in Schwarzschild spacetime. Subsequently, the hard X-ray tail would become softer than that of the calculation for flat spacetime. Niedźwiecki & Zdziarski (2006) calculated the photon spectra Comptonized by spherically free-falling accretion flow, whose seed photons are emitted from the accretion disk outside the spherical accretion flows (but sometimes the disk extended down to the black hole) in flat, Schwarzschild, and Kerr spacetime. They compared the results for flat spacetime and those for Schwarzschild coordinates, and found that the effect of gravitational redshift softens the photon spectra in the hard X-ray band. Their results indicate that the difference between the spectra for the flat and the Schwarzschild spacetime becomes small in significant supercritical accretion model ${\dot{M}}$ = 12 LE/c2 because most of the photons scattered up to the hard X-ray band by the bulk Compton effect are swallowed by the black hole via the photon trapping effect. Although this seems to indicate that the general relativistic treatment does not affect the observable SEDs, it might not be negligible because our results show that the funnel jet assists photons, which are upscattered in the shock-heated region near the black hole, to escape to observers. The spectral calculation of supercritical accretion flows taking into account the general relativistic effect remains as a future work.

4.6. Dependence on the Viewing Angle

We emphasize that the funnel jet affects the dependence of the observable spectra on viewing angle. Niedźwiecki & Zdziarski (2006) showed that the photon spectra become softer when the viewing angle is lower, because the soft photons are emitted from the geometrically thin accretion disk and the isotropic high-energy photons arise from the spherical accretion flows via Compton scatterings. On the other hand, our result shows that the radiation spectra become harder when the viewing angle is smaller (Figure 4), because the hard X-ray photons escape through the funnel (see Figure 6). We conclude that it is important to take into account the disk geometry for spectral calculation of accretion flows.

We thank T. Hanawa and O. Blaes for useful discussion. We are grateful to J. Gladstone for providing us the XMM-Newton data corrected for absorption. The numerical simulations were carried out on the XT4 at the Center for Computational Astrophysics, National Astronomical Observatory of Japan. This work was supported by the Grant-in-Aid for Scientific Research of the Ministry of Education, Culture, Sports, Science, and Technology (20340040, R.M.) and the Grant-in-Aid for JSPS Fellows (22.4587, T.K.).

APPENDIX A: SUMMARY OF NUMERICAL METHOD FOR SOLVING THE RADIATION HYDRODYNAMIC EQUATIONS

In this appendix, we describe the basic equations and numerical methods adopted in radiation hydrodynamical simulations presented in this paper. See Ohsuga et al. (2005), Ohsuga (2006), and Kawashima et al. (2009) for the details of numerical methods.

We solved the set of radiation hydrodynamic equations. The basic equations are the continuity equation,

Equation (A1)

the equations of motion,

Equation (A2)

Equation (A3)

Equation (A4)

the energy equation of the gas,

Equation (A5)

and the energy equation of the radiation,

Equation (A6)

Here, ρ is the mass density, ${\bm v} = (v_{r}, v_{\theta }, v_{\phi })$ is the velocity, p is the gas pressure, e is the internal energy density, trϕ is the viscous stress tensor, Φvis is the viscous dissipative function, E0 is the radiation energy density, ${\bm F}_{0}=(F_{0r},F_{0{\theta }})$ is the radiative flux, P0 is the radiation pressure tensor, B is the blackbody intensity, κ is the absorption opacity, χ = κ + ρσT/mp is the total opacity, where σT is the cross-section of Thomson scattering, and mp is the proton mass. In the energy equation, ΓComp denotes the energy transport rate from the gas to the radiation field via the Comptonization, i.e., Compton heating/cooling rate. We write ΓComp as

Equation (A7)

where Tgas is the gas temperature, Trad[ = (E0/a)1/4] is the radiation temperature, a is the radiation constant, and me is the electron mass. This equation is obtained by integrating the Kompaneets equation over the frequency. We assume a one-temperature plasma in which the electron temperature equals the ion temperature. We employed the equation of state such that p = (γ − 1)e, where γ is the specific heat ratio. The gas temperature Tgas is, hence, calculated from p = ρkBTgas/(μmp), where kB is the Boltzmann constant and μ is the mean molecular weight. The viscous stress tensor trϕ and viscous dissipative function Φvis are written as

Equation (A8)

Equation (A9)

where η is the dynamical viscous coefficient described by α-prescription being proportional to the gas and the radiation pressure as

Equation (A10)

Here, ΩK is the Keplerian angular speed and λ represents the flux limiter of the flux-limited diffusion approximation (see Ohsuga et al. 2005 for the detail).

We solve the radiation hydrodynamic equations by using an explicit–implicit scheme. Our code is the same as Ohsuga (2006) except that this code solves the gas and the radiation energy equations including Compton cooling/heating terms. We employ the operator-splitting method in radiation hydrodynamic equations. The energy equations are solved by updating the following terms: (1) the gas and radiation energy exchange terms except the Compton cooling/heating terms,

Equation (A11)

Equation (A12)

(2) the Compton cooling/heating terms,

Equation (A13)

Equation (A14)

(3) the radiation flux divergence term,

Equation (A15)

(4) the advection term,

Equation (A16)

Steps (1) and (2) are solved by using the Newton–Raphson iteration, with the bisection method when Newton–Raphson failed. Steps (3) and (4) are solved by the explicit method. These modules to solve the radiation transfer were implemented in a hydrodynamic code Virginia Hydrodynamics One (VH-1). VH-1 is based on the Lagrange-remap version of the piecewise parabolic method given by Colella & Woodward (1984), which is the high-order Godunov's scheme using a quadratic interpolation of the fluid variables at each cell boundary and is written as a Lagrangian hydrodynamic program coupled with a remap onto the original Eulerian grid. The Riemann problem is solved at each cell boundary as described in van Leer (1979). This code provides us third-order accuracy in space and second-order accuracy in time. We note that the update of vϕ due to the viscous stress tensor in Equation (A4) is solved implicitly by using the Thomas method.

APPENDIX B: MODIFICATION OF SCATTERING COEFFICIENT DUE TO BULK AND THERMAL MOTION OF ELECTRONS

We summarize the effective scattering coefficient of Compton scattering, which is modified by the bulk and thermal motions of electrons. Since the scattering cross-section is well defined in the rest frame of each electron, we should use the effective cross-section when we calculate the optical depth in the observer frame. The derivation is almost the same as that by Abramowicz et al. (1991) for fluid motion and that by Pozdnyakov et al. (1983) for thermal motion of electrons. We derive the effective scattering coefficient on the basis of Lorentz invariance of the optical depth, and we do not assume the uniform plasma, which is assumed in the derivation by Pozdnyakov et al. (1983).

At first, we consider the modification of the scattering coefficient due to fluid motion. In order to obtain the effective scattering coefficient, we have to derive the relation between dl and dl0, where dl and dl0 are the lengths of the line element along the ray that the photon actually travels in the observer frame and in the comoving frame of the fluid, respectively. Hereafter, the subscript "0" denotes the physical quantity measured in the fluid rest frame. We introduce four-velocity of the fluid ${\bm u}$ = ${\it {\Gamma }}(c, {\bm v})$ and a null vector which describes an infinitesimal tangent vector of the photon trajectory $d{\bm x}$ = $ (cdt, d{\bm l})$, such that dl = $|d {\bm l}|$. In our study, the photon trajectory is a line since the spacetime is assumed to be flat. From uμdxμ = udxμ0, we obtain the following relation:

Equation (B1)

where we used the property of a massless particle 0 = −c2dt20 + dl02 = −c2dt2 + dl2, and defined θ as an angle between the direction of photon propagation and that of fluid motion in an observer frame such that ${\bm v}{\cdot }d{\bm l} = v \cos \theta dl$. We note that Equation (B1) is the same as the transformation of photon frequency and does not coincide with the Lorentz contraction form. In order to calculate the optical depth, we have to use the transformation of the length (Equation (B1)) rather than the form of the Lorentz contraction, because we should compare the distances which a photon travels between a couple of world points (i.e., cdt and cdt0 in the observer frame and the fluid rest frame, respectively).

The optical depth is Lorentz invariant because it represents a number or a probability of interactions of a photon with matter during the propagation along a world line connecting the two world points. Therefore, we obtain the following relation:

Equation (B2)

where we used the relation (B1) and N = ΓN0, which is the result of Lorentz contraction. We can regard [1 − (v/c)cos θ]σ0 as an effective cross-section due to fluid motion. The scattering cross-section σ0 depends on the photon frequency, i.e., σ0 = σ00), where ν0 = νΓ[1 − (v/c)cos θ]. It should be noted that we defined dτ0 = σ0N0dl0. On the other hand, Abramowicz et al. (1991) fixed dl in the observer frame and defined dτ0 = σ0N0dl. They corrected the optical depth as dτ = (lmfp0/lmfp)dτ0, where lmfp is the mean free path of a photon for electron scattering. However, the derivation in this paper and in Abramowicz et al. (1991) is equivalent because dl/lmfp is Lorentz invariant, that is, (lmfp0/lmfp0N0dl = σ0N0dl0.

Next we take into account the correction of σ0 by the thermal motion of electrons. Similar to Equation (B2), if we consider the scattering due to the electron with a particular velocity ${\bm v}_{\rm e}$ which is measured in the fluid rest frame, we obtain the following relation:

Equation (B3)

where each value with the subscript "0e" is measured in the comoving frame of the electron, θe is an angle formed by the ray and an electron with velocity ${\bm v}_{\rm e}$, ${\bm p}_{\rm e}$ is the space component of the four-momentum of the electron, and $f_{0}({\bm p}_{\rm e})$ is the distribution function for the electrons defined in the comoving frame of the fluid. We used the relation $N_{\rm 0e}({\bm v}_{\rm e})$ = $(N_{\rm 0}/{\it {\Gamma }}_{\rm e})f_{0}({\bm p}_{\rm e})$, where Γe is the Lorentz factor for the motion of the electron in the fluid rest frame. We assume that the electrons are thermalized, i.e., $f_{0}({\bm p}_{\rm e})$ is the Maxwell–Boltzmann distribution function:

Equation (B4)

where Θe  ≡  (kBTe)/(mec2). From Equation (B3), we can regard σ0 as the Klein–Nishina cross-section averaged over the momentum space of thermal electrons:

Equation (B5)

Here, the Klein–Nishina cross-section σKN is a function of the dimensionless variable x, which is defined as x = (2hν0/mec2)Γe[1 − (ve/c)cos θe], and is written as

Equation (B6)

where $\hat{\sigma }(x)$ is the dimensionless function introduced for the sake of convenience, and is described as

Equation (B7)

By fixing the value of ve (i.e., fixing pe) during the integration over the angle, dx is written as dx = −(2hν0/mec2)(ve/c)Γed(cos θe), and we obtain

Equation (B8)

where we define ${\it {\Gamma }}^{\pm }_{\rm e}\,{\equiv }\,{\it {\Gamma }}_{\rm e} \pm \sqrt{{\it {\Gamma }}_{\rm e}^2 - 1}$, and Φ(x) is written as the following:

Equation (B9)

Therefore, the averaged cross-section in Equation (B5) is rewritten as

Equation (B10)

where we use the relation $p_{\rm e} = m_{\rm e}c\sqrt{{\it {\Gamma }}_{\rm e}^{2}-1}$. We compute the integral in both the numerator and the denominator in this equation. The numerator and the denominator are described as the form ∫1exp (− Γee)Ψ(Γe)e, where Ψ(Γe) = $\left[{\Phi }(x)\right]^{x=(2h{\nu }_{0}/m_{\rm e}c^{2}){\it {\Gamma }}_{\rm e}^{+}}_{x=(2h{\nu }_{0}/m_{\rm e}c^{2}){\it {\Gamma }}_{\rm e}^{-}}$ (for the numerator) and ${\it {\Gamma }}_{\rm e}\sqrt{{\it {\Gamma }}_{\rm e}^{2}-1}$ (for the denominator). In order to calculate in the finite integral range, a new variable u = exp [(1 − Γe)/Θe] is introduced and used instead of Γe. We use the following formula for computation:

Equation (B11)

where ${\tilde{u}}$ and ${\tilde{u}}_{\rm max}$ are integers. By substituting Equation (B11) into Equation (B5), we obtain the numerical formula for the Klein–Nishina cross-section averaged by momentum of electrons as follows:

Equation (B12)

where ${\tilde{\it \Gamma }}_{\rm e}({\tilde{u}})$ =$1 - {\Theta }_{\rm e}{\ln }\left[({\tilde{u}} - 1/2)/{\tilde{u}}_{\rm max}\right]$ $({\tilde{u}} = 1, 2, \ldots , {\tilde{u}}_{\rm max})$ and ${\tilde{\it \Gamma }}_{\rm e}^{\pm }$ = ${\tilde{\it \Gamma }}_{\rm e} ({\tilde{u}}) {\pm } \sqrt{({\tilde{\it \Gamma }}_{\rm e}({\tilde{u}}))^{2} - 1}$.

Finally, from Equation (B2), we obtain the effective scattering coefficient αKNν, eff, which is modified by both the bulk motion and the thermal motion of electron, as

Equation (B13)

where we set ${\tilde{u}}_{\rm max} = 20$ when we calculate σ0 described in Equation (B12).

Please wait… references are loading.
10.1088/0004-637X/752/1/18