Brought to you by:

Articles

SAGITTARIUS A* ACCRETION FLOW AND BLACK HOLE PARAMETERS FROM GENERAL RELATIVISTIC DYNAMICAL AND POLARIZED RADIATIVE MODELING

, , and

Published 2012 August 2 © 2012. The American Astronomical Society. All rights reserved.
, , Citation Roman V. Shcherbakov et al 2012 ApJ 755 133 DOI 10.1088/0004-637X/755/2/133

0004-637X/755/2/133

ABSTRACT

We obtain estimates of Sgr A* accretion flow and black hole parameters by fitting polarized submillimeter observations with spectra computed using three-dimensional general relativistic (GR) magnetohydrodynamical (MHD) (GRMHD) simulations. Observations are compiled from averages over many epochs from reports in 29 papers for estimating the mean fluxes Fν, linear polarization (LP) fractions, circular polarization (CP) fractions, and electric vector position angles. GRMHD simulations are computed with dimensionless spins a* = 0, 0.5, 0.7, 0.9, 0.98 over a 20, 000M time interval. We perform fully self-consistent GR polarized radiative transfer using our new code to explore the effects of spin a*, inclination angle θ, position angle (P.A.), accretion rate $\dot{M}$, and electron temperature Te (Te is reported for radius 6M). By fitting the mean submillimeter fluxes and LP/CP fractions, we obtain estimates for these model parameters and determine the physical effects that could produce polarization signatures. Our best-bet model has a* = 0.5, θ = 75°, P.A. = 115°, $\dot{M}=4.6\times 10^{-8}\ M_\odot \ {\rm {{\rm yr}}}^{-1}$, and Te = 3.1 × 1010 K at 6M. The submillimeter CP is mainly produced by Faraday conversion as modified by Faraday rotation, and the emission region size at 230 GHz is consistent with the very long baseline interferometry size of 37 μas. Across all spins, model parameters are in the ranges θ = 42°–75°, $\dot{M}=(1.4\hbox{--}7.0)\times 10^{-8}\ M_\odot \ {\rm {{\rm yr}}}^{-1}$, and Te = (3–4) × 1010 K. Polarization is found both to help differentiate models and to introduce new observational constraints on the effects of the magnetic field that might not be fit by accretion models so far considered.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The mass of the Galactic center black hole (BH) is M ≈ 4.5 × 106M (Ghez et al. 2008; Reid et al. 2008; Gillessen et al. 2009), and the spin is uncertain (Huang et al. 2009b; Broderick et al. 2009, 2011; Moscibrodzka et al. 2009; Dexter et al. 2010). It resides at a distance d ≈ 8.4 kpc. Because of its proximity, it has been observed in many wavelengths: γ-rays, X-rays, IR, (sub)millimeter, and radio. X-ray bremsstrahlung emission originates from hot gas at large radii where the BH's gravity becomes important (Narayan et al. 1995, 1998; Shcherbakov & Baganoff 2010), and Compton-scattered emission originates from close to the horizon (Moscibrodzka et al. 2009). X-rays at large radii are spatially resolved and have been used to constrain dynamical models for this region (Shcherbakov & Baganoff 2010). The submillimeter emission is cyclo-synchrotron emission originating from close to the BH. Cyclo-synchrotron emission is polarized, and both linear polarization (LP) and circular polarization (CP) have been observed from Sgr A* at several submillimeter wavelengths. The accretion flow was recently resolved at 230 GHz (Doeleman et al. 2008; Fish et al. 2011). General relativistic (GR) effects were deemed necessary to explain the small size with FWHM of 37 μas. Radio emission is also produced by cyclo-synchrotron at larger distances from the BH. Relativistic frame dragging is important near the BH, so submillimeter polarized observations and the Compton-scattered X-rays might help to constrain the BH spin. The goal of the present paper is to model the submillimeter in the range of 88–857 GHz in order to estimate the accretion flow and BH parameters.

Sgr A* is a variable source with a variability amplitude routinely reaching 30% in submillimeter. A popular approach is to fit simultaneous observations (e.g., Yuan et al. 2004; Broderick et al. 2009), in particular, the set from Falcke et al. (1998). However, in such an approach, one would use a single simultaneous set of observations. However, simultaneous observations of fluxes and LP and CP fractions at several frequencies are not available. So we consider non-simultaneous statistics of all observations at all frequencies and find the mean values and standard errors of quantities at each frequency.

Numerous accretion flow models have been applied to the Galactic center: advection-dominated accretion flow (ADAF; Narayan & Yi 1995), advection-dominated inflow–outflow solution (ADIOS; Blandford & Begelman 1999), jet-ADAF (Yuan et al. 2002), jet (Maitra et al. 2009), and viscous and magnetohydrodynamical (MHD) numerical simulations. The quasi-analytical models are useful because there is little expense in changing parameters. However, they have a large number of free parameters and also incorporate many assumptions that are not justifiable from first principles (Huang et al. 2008, 2009a), which leads to systematic uncertainties in all fits. Numerical simulations require fewer inputs and are useful for more quantitative modeling of the plasma near a rotating BH. General relativistic (GR) MHD (GRMHD) simulations (especially three-dimensional (3D) simulations), which are run over a sufficiently long duration, are still computationally expensive and involve state-of-the-art codes that are still being developed (McKinney & Blandford 2009; Fragile et al. 2009; Noble & Krolik 2009; Moscibrodzka et al. 2009; Penna et al. 2010). Yet, these expensive 3D simulations are required to model the turbulent disk flow, because two-dimensional (2D) axisymmetric simulations cannot sustain turbulence as shown by generalizations of Cowling's anti-dynamo theorem (Hide & Palmer 1982). Given their expense, such 3D GRMHD simulations are limited to a region relatively close to the BH (Dexter et al. 2009; Moscibrodzka et al. 2009), whereas some emission and some Faraday rotation might happen far from the BH. Thus, we analytically extend the modeled region out to 20, 000M, perform radiative transfer, and find the best fit to the data. The extension to large radius allows us to define the electron temperature more consistently (Sharma et al. 2007). We find a posteriori (see the Appendix) that the simulated polarized spectra are not overly sensitive to the details of the analytic extensions of density and temperature but may depend on the extension of the magnetic field.

The radiation close to the BH has been modeled in Newtonian (Yuan et al. 2004) and quasi-Newtonian approximations (Goldston et al. 2005; Chan et al. 2009). It has been modeled in GR assuming unpolarized (Fuerst & Wu 2004; Dexter et al. 2009; Dolence et al. 2009) and polarized (Broderick et al. 2009; Shcherbakov & Huang 2011) light. Fitting the total flux spectrum might not be sufficient to estimate the spin, and naturally one expects polarization to provide extra observational constraints. Spin values from a* = 0 (Broderick et al. 2009) to a* = 0.9 (Moscibrodzka et al. 2009) have been estimated. We neglect Comptonization (Moscibrodzka et al. 2009) and radiation from non-thermal electrons (Mahadevan 1998; Özel et al. 2000; Yuan et al. 2004). Emissivities are calculated in the synchrotron approximation (Legg & Westfold 1968; Sazonov 1969; Pacholczyk 1970; Melrose 1971) with an exact thermal electron distribution. Discrepancies with the exact cyclo-synchrotron emissivities (Leung et al. 2011; Shcherbakov & Huang 2011) are negligible as estimated in Section 5. Exact Faraday rotation and conversion expressions are used (Shcherbakov 2008).

We compare simulated spectra to observed ones at many frequencies simultaneously, extending an approach pioneered by Broderick et al. (2009) and Dexter et al. (2009). We compute the average observed spectra, find the deviations of the means, and then compare them to the average simulated spectra. In the search for the best-fit models, we are guided by the value of χ2/dof, which is the normalized sum of squares of normalized deviations. Yet, we leave the exploration of the statistical meaning of χ2/dof to future work. We search the space of all parameters, spin a*, inclination θ, ratio of proton to electron temperatures Tp/Te (Tp/Te is reported for radius 6M), and accretion rate $\dot{M}$, to find the minimum χ2 models.

We summarize the radio/submillimeter observations of Sgr A* in Section 2. Our 3D GRMHD simulations are described in Section 3 together with the physically motivated extension to large radii and the electron heating prescription. We run simulations for dimensionless spins a* = a/M = 0, 0.5, 0.7, 0.9, 0.98. The GR polarized radiative transfer technique is described in Section 5.

The set of observations we consider consists of the spectral energy distribution (SED) within the 88–857 GHz frequency range, LP fractions at 88, 230, and 349 GHz, and CP fractions at 230 and 349 GHz. In Section 6 we discuss our results: the best-fit models to the observations, the importance of various physical effects in producing the observed CP and LP and electric vector position angle (EVPA), and image size estimates. We produce the simulated images of total and polarized intensities. Discussion in Section 7 compares the results to previous estimates, emphasizes the significance of polarization, notes the sources of error, and outlines prospects for future work. In the Appendix we describe a number of convergence tests of our GR polarized radiative transfer code and the radial extension of the dynamical model. Throughout the paper we measure distance and time in the units of BH mass M by setting the speed of light c and gravitational constant G to unity.

2. OBSERVATIONS

Sgr A* is known to be a highly variable source, yet quiescent models of Sgr A* emission are popular and useful. Unlike the drastic variations of X-ray and NIR fluxes (Baganoff et al. 2001; Genzel et al. 2003), submillimeter fluxes do not vary by more than a factor of 2–3 (Zhao et al. 2003). We compile the set of observed polarized fluxes at each frequency and then find the mean spectrum and the errors of the mean fluxes.

Previously, the observed flux spectra were compiled by Yuan et al. (2004) and Broderick et al. (2009). However, both papers summarize a limited set of observations and concentrate on simultaneously observed fluxes. Submillimeter flux data reported in Yuan et al. (2004) consist of a short set of observations by Falcke et al. (1998) and one set of the Submillimeter Array (SMA) observations by Zhao et al. (2003). Broderick et al. (2009) add to these the rest of the SMA total flux data (Marrone et al. 2006a, 2006b, 2007, 2008). Hence 6 out of at least 29 papers on submillimeter observations of Sgr A* were taken into account. We compute an averaged spectrum based on 29 papers reporting submillimeter observations of Sgr A*.

The reported observations vary in covered period from several hours (An et al. 2005) to several years (Zhao et al. 2003; Krichbaum et al. 2006). We know that variations of a factor of two may happen within several hours (Yusef-Zadeh et al. 2009), whereas variations by more than a factor of several are never observed in the submillimeter. So, fluxes observed more than a day apart are weakly correlated. The issue of autocorrelation in timescales will be addressed in future work. We consider the following averaging technique to sample the distributions of fluxes. First, we define groups of close frequencies, the frequencies in each group being different by no more than several percent from the mean. There are 11 such groups (see Table 1). We exclude papers reporting single frequencies far from the mean of each group. In particular, the 94 and 95 GHz observations of Li et al. (2008) and Falcke et al. (1998) and the 112 GHz observations of Bower et al. (2001) are excluded. A mean frequency is ascribed to represent each group. Then, we take all reported observations of each polarization type (total flux, LP and CP fraction, EVPA) for each group and draw the largest sample of fluxes/polarization fractions, taking observations separated by at least 24 hr. When several fluxes are reported over a period of several hours (Yusef-Zadeh et al. 2009), we draw one data point from the very beginning of the observation, unless a flare is reported to occur at that time. Some of the published observations have large error bars. Often such data are produced by observing in submillimeter with large beam size, but light from Sgr A* is blended with dust and other sources. In particular, the Submillimeter Telescope (SMT) data (Yusef-Zadeh et al. 2009), early Caltech Submillimeter Observatory (CSO) measurements (Serabyn et al. 1997), and early James Clerk Maxwell telescope (JCMT) measurements (Aitken et al. 2000) may have such issues, so we exclude these data from the sample. The interferometric observations, especially with very long baseline interferometry (VLBI), help to reduce the error from otherwise unreliable observations, e.g., with the Berkeley–Illinois–Maryland Association (BIMA) array (Bower et al. 2001). However, some inconsistencies still exist for simultaneous observations at the same frequency with different instruments (Yusef-Zadeh et al. 2009).

Table 1. Summary of Sgr A* Radio/Submillimeter Observations: Means and 1σ Uncertainties

ν Telescopes Fν LP CP EVPA
(GHz)   (Jy) (%) (%) (°)
8.45 VLA $\bf 0.683\pm 0.032$ (Serabyn et al. 1997; Falcke et al. 1998; Bower et al. 1999a; An et al. 2005) ... $\bf -0.26\pm 0.06$a (Bower et al. 1999a)  
14.90 VLBA, VLA $\bf 0.871\pm 0.012$ (Serabyn et al. 1997; Falcke et al. 1998; Bower et al. 2002; Herrnstein et al. 2004; An et al. 2005; Yusef-Zadeh et al. 2009) ... $\bf -0.62\pm 0.26$a (Bower et al. 2002) ...
22.50 VLBA, VLA $\bf 0.979\pm 0.016$ (Serabyn et al. 1997; Falcke et al. 1998; Bower et al. 1999b; Herrnstein et al. 2004; An et al. 2005; Lu et al. 2008; Yusef-Zadeh et al. 2007, 2009) $\bf 0.20\pm 0.01$a (Bower et al. 1999b; Yusef-Zadeh et al. 2007) ... ...
43 GMVA, VLBA, VLA $\bf 1.135\pm 0.026$ (Falcke et al. 1998; Lo et al. 1998; Bower et al. 1999b; Herrnstein et al. 2004; An et al. 2005; Shen et al. 2005; Krichbaum et al. 2006; Yusef-Zadeh et al. 2007; Lu et al. 2008; Yusef-Zadeh et al. 2009) $\bf 0.55\pm 0.22$a (Bower et al. 1999b; Yusef-Zadeh et al. 2007) ... ...
88 BIMA array, MPIfR, VLBA, VLA, Nobeyama, NMA, CARMA $\bf 1.841\pm 0.080$ (Falcke et al. 1998; Krichbaum et al. 1998; Bower et al. 1999b; Doeleman et al. 2001; Miyazaki et al. 2004; Shen et al. 2005; Krichbaum et al. 2006; Macquart et al. 2006; Lu et al. 2008; Yusef-Zadeh et al. 2009) $\bf 1.42\pm 0.5$a,b (Bower et al. 1999b; Macquart et al. 2006) ... −4c (Bower et al. 1999b; Shen et al. 2005; Macquart et al. 2006)
102 OVRO, CSO-JCMT, Nobeyama, NMA, IRAM $\bf 1.91\pm 0.15$ (Serabyn et al. 1997; Falcke et al. 1998; Miyazaki et al. 2004; Mauerhan et al. 2005; Yusef-Zadeh et al. 2009) ... ... ...
145 Nobeyama, NMA, IRAM, JCMT $\bf 2.28\pm 0.26$ (Falcke et al. 1998; Aitken et al. 2000; Miyazaki et al. 2004; Yusef-Zadeh et al. 2009) ... ... ...
230 IRAM, JCMT, BIMA array, SMA, OVRO $\bf 2.64\pm 0.14$ (Serabyn et al. 1997; Falcke et al. 1998; Aitken et al. 2000; Bower et al. 2003, 2005; Zhao et al. 2003; Krichbaum et al. 2006; Marrone et al. 2006a, 2007, 2008; Doeleman et al. 2008; Yusef-Zadeh et al. 2009) $\bf 7.40\pm 0.66$ (Bower et al. 2003, 2005; Marrone et al. 2007, 2008) $\bf -1.2\pm 0.3$a (Munoz et al. (2009, 2012)) $\bf 111.5\pm 5.3$ (Bower et al. 2003, 2005; Marrone et al. 2007, 2008)
349 SMA, CSO, JCMT $\bf 3.18\pm 0.12$ (Aitken et al. 2000; An et al. 2005; Marrone et al. 2006b, 2007, 2008; Yusef-Zadeh et al. 2009) $\bf 6.50\pm 0.61$ (Marrone et al. 2006b, 2007) $\bf -1.5\pm 0.3$a (Munoz et al. (2012)) $\bf 146.9\pm 2.2$ (Marrone et al. 2006b, 2007)
674 CSO, SMA $\bf 3.29\pm 0.35$ (Marrone et al. 2006a, 2008; Yusef-Zadeh et al. 2009) ... ... ...
857 CSO $\bf 2.87\pm 0.24$ (Serabyn et al. 1997; Marrone et al. 2008; Yusef-Zadeh et al. 2009) ... ... ...

Notes. aThe uncertainty of the mean of these quantities is given by instrumental errors. bThe mean LP at 3.5 mm is computed based on lower and upper sidebands in Macquart et al. (2006). The error is based on 0.5% systematic error reported therein. cThe mean EVPA at 88 GHz is uncertain due to ±180° degeneracy; e.g., the reported EVPA = 80° could as well be interpreted as −100°.

Download table as:  ASCIITypeset image

After the sample of fluxes, polarization fractions, and EVPAs are found for each frequency group, we compute the mean and the standard error. The summary of results is presented in Table 1. CP fractions of −1.2% at 230 GHz and −1.5% at 349 GHz are based on SMA observations by Munoz et al. (2012) with the reported ±0.3% instrumental error. Note that standard errors in our total flux samples are smaller than the error bars of prior observations (Falcke et al. 1998; Yuan et al. 2004; Broderick et al. 2009), but still larger compared to contemporary single-observation instrumental errors (Marrone et al. 2007). That is, we do not incorporate instrumental error in our estimates of standard error of the mean fluxes or LP and EVPA at 230 and 349 GHz (even though the instrumental error of LP at 88 GHz is large). We do not incorporate the source size measurements (Doeleman et al. 2008) in calculating χ2/dof, but we check that the best bet model is not inconsistent with those observations. Figure 1 shows a compilation of the mean quantities and their Gaussian standard errors. The data are represented by both error bars and the interpolated shaded area. A red dashed curve on the Fν plot represents the analytic approximation Fν = 0.248ν0.45exp (− (ν/1100)2), where flux is in Jy and frequency is in GHz.

Figure 1.

Figure 1. Mean observed SEDs of specific flux Fν, linear polarization (LP) fraction, electric vector position angle (EVPA), and circular polarization (CP) fraction. The error bars show 1σ standard error of the mean. The dashed line on the Fν plot represents the analytic approximation Fν(Jy) = 0.248ν0.45exp (− (ν/1100)2) for frequency ν in GHz (not the simulated SED). As noted in Table 1, the error is instrumental for CP at high frequencies and LP at 88 GHz, whereas it is computed from a sample of observed quantities for flux, EVPA at all frequencies, and LP at high frequencies.

Standard image High-resolution image

3. THREE-DIMENSIONAL GRMHD SIMULATIONS

Our radiative transfer calculations take the results of simulations of accretion flows onto BHs as input. These simulations are similar to those in Penna et al. (2010). Below, we review the methodology.

3.1. Governing Equations

We simulate radiatively inefficient accretion flows (RIAFs) onto rotating BHs using a 3D fully GR code (see Section 3.3). The BH is described by the Kerr metric. We work with Heaviside–Lorentz units. Our five simulations correspond to different choices of the dimensionless BH spin parameter: a* = 0, 0.5, 0.7, 0.9, and 0.98. The self-gravity of the RIAF is ignored.

The RIAF is a magnetized fluid, so we solve the GRMHD equations of motion (Gammie et al. 2003). Mass conservation gives

Equation (1)

where ρ is the fluid-frame rest-mass density, uμ is the contravariant four-velocity, and ∇μ is the covariant derivative. Energy–momentum conservation gives

Equation (2)

where the stress energy tensor Tμν includes both matter and electromagnetic terms,

Equation (3)

where ugas is the internal energy density and pgas = (Γ − 1)ugas is the ideal gas pressure with Γ = 4/3. Models with Γ = 5/3 show minor differences compared to models with Γ = 4/3 (McKinney & Gammie 2004; Mignone & McKinney 2007). The contravariant fluid-frame magnetic four-field is given by bμ and is related to the lab-frame three-field Bmu via bμ = Bνhμν/ut, where hμν = uμuν + δμν is a projection tensor and δμν is the Kronecker delta function (Gammie et al. 2003). We often employ $\bf b$ below, which is the orthonormal magnetic field vector in a comoving locally flat reference frame (Penna et al. 2010). The magnetic energy density (ub) and magnetic pressure (pmag) are then given by umag = pmag = bμbμ/2 = b2/2 = b2/2. Note that the angular velocity of the gas is Ω = uϕ/ut.

Magnetic flux conservation is given by the induction equation

Equation (4)

where vi = ui/ut and g = Det(gμν) is the determinant of the metric. No explicit resistivity or viscosity is included, but we use a shock-capturing Godunov method that fully conserves energy. So, all dissipation from shocks and numerical diffusivity (e.g., in shear flows or current sheets) is fully captured, as required to study RIAFs.

In Penna et al. (2010), we studied both RIAFs and geometrically thin radiatively efficient disks. For the latter case, a cooling term was added to the energy–momentum Equation (2) to describe radiative losses and keep the disk thin. The current set of models are all RIAFs, so no cooling term is used. Entropy generated by viscous or resistive dissipation is advected along with the inflow or transported out via convection or in a wind.

3.2. Physical Models

The initial mass distribution is an isentropic equilibrium torus (Chakrabarti 1985a, 1985b; De Villiers et al. 2003) with pressure p = K0ρ4/3 for K0 = 0.009. The torus inner edge is at rin = 20M, and the maximum density and pressure are at Rmax = 65M. We initialize the solution so that ρ = 1 at the pressure maximum. As in Chakrabarti (1985a), the angular velocity distribution of the initial torus is a power law, where for the Chakrabarti (1985a) q-parameter we choose q = 1.65 (at large radii Ω ∼ (r/M)q). The thickness of the torus at the pressure maximum is then |h/r| ∼ 0.3, where

Equation (5)

where $dA_{\theta \phi }\equiv \sqrt{-g} d\theta d\phi$ is an area element in the θ − ϕ plane and the integral over dt is a time average over the period when the disk is in a steady state (see Section 3.6). A tenuous atmosphere fills the space outside the torus. It has the same polytropic equation of state as the torus, $p=K_0 \rho ^\Gamma$, with Γ = 4/3, and an initial rest-mass density of ρ = 10−6(r/M)−3/2, corresponding to a Bondi-like atmosphere. The torus is threaded with three loops of weak, poloidal magnetic field: the initial gas-to-magnetic pressure ratio is β = pgas, max/pmag, max = 100, where pgas, max and pmag, max are the maximum values of the gas and magnetic pressure in the torus. This approach to normalizing the initial field is used in many other studies (Gammie et al. 2003; McKinney & Gammie 2004; McKinney 2006a; McKinney & Narayan 2007b; Komissarov & McKinney 2007; Penna et al. 2010).

Recent GRMHD simulations of thick disks indicate that the results for the disk (but not the wind jet, which for us is less important) are roughly independent of the initial field geometry (McKinney & Narayan 2007a, 2007b; Beckwith et al. 2008, but see also McKinney et al. 2012). The magnetic vector potential we use is given by

Equation (6)

with all other Aμ initially zero. This is the same Aμ as used in Penna et al. (2010). We use Q = (ugas/ugas, max − 0.2)(r/M)3/4 and set Q = 0 if either r < S or Q < 0. Here ug, max is the maximum value of the internal energy density in the torus. We choose S = 22M and λfield/(2πr) = 0.28, which give initial poloidal loops that are roughly isotropic such that they have roughly 1:1 aspect ratio in the poloidal plane. The form of the potential in Equation (6) ensures that each additional field loop bundle has opposite polarity. Perturbations are introduced to excite the magnetorotational instability (MRI). The second term on the right-hand side of Equation (6) is a random perturbation: ranc is a random real number generator for the domain 0–1. Random perturbations are introduced in the initial internal energy density in the same way, with an amplitude of 10%. In Penna et al. (2010), it was found that similar simulations with perturbations of 2% and 10% became turbulent at about the same time, the magnetic field energy at that time was negligibly different, and there was no evidence for significant differences in any quantities during inflow equilibrium.

3.3. Numerical Methods

We perform simulations using a fully 3D version of HARM that uses a conservative shock-capturing Godunov scheme (Gammie et al. 2003; Shafee et al. 2008; McKinney 2006b; Noble et al. 2006; Mignone & McKinney 2007; Tchekhovskoy et al. 2007; McKinney & Blandford 2009). We use horizon-penetrating Kerr–Schild coordinates for the Kerr metric (Gammie et al. 2003; McKinney & Gammie 2004), which avoids any issues with the coordinate singularity in Boyer–Lindquist coordinates. The code uses uniform internal coordinates (t, x(1), x(2), x(3)) mapped to the physical coordinates (t, r, θ, ϕ). The radial grid mapping is

Equation (7)

which spans from Rin = 0.9rH to Rout = 200M, where rH is the radius of the outer event horizon. This just ensures that the grid never extends inside the inner horizon, in which case the equations of motion would no longer be hyperbolic. The parameter R0 = 0.3M controls the resolution near the horizon. For the outer radial boundary of the box, absorbing (outflow, no inflow allowed) boundary conditions are used.

The θ-grid mapping is

Equation (8)

where x(2) ranges from 0 to 1 (i.e., no cutout at the poles) and Y = 0.65 is chosen to concentrate grid zones toward the equator. Reflecting boundary conditions are used at the polar axes. The ϕ-grid mapping is given by ϕ(x(3)) = 2πx(3), such that x(3) varies from 0 to 1/2 for a box with Δϕ = π. Periodic boundary conditions are used in the ϕ-direction. Penna et al. (2010) considered various Δϕ for thin disks and found little difference in the results. In all of their tests, Δϕ > 7|h/r| and we remain above this limit as well. In what follows, spatial integrals are renormalized to refer to the full 2π range in ϕ, even if our computational box size is limited in the ϕ-direction. For the purpose of radiative transfer, we combine two identical regions of size Δϕ = π, preserving the orientation to obtain the span of full 2π.

3.4. Resolution and Spatial Convergence

The resolution of the simulations is Nr × Nθ × Nϕ = 256 × 64 × 32. This is the fiducial resolution of Penna et al. (2010). Shafee et al. (2008) found this resolution to be sufficient to obtain convergence compared to a similar 512 × 128 × 32 model. In the vertical direction, we have about seven grid cells per density scale height. Turbulence is powered by the MRI, which is seeded by the vertical component of the magnetic field (Balbus & Hawley 1998). The characteristic length scale of the MRI is the wavelength of the fastest-growing mode:

Equation (9)

where vzA is the vertical component of the Alfvén speed. We find that the MRI is well resolved in the midplane of the disk both initially and in the saturated state.

Penna et al. (2010) studied convergence in Nr, Nθ, and Nϕ and found that models with Nr = 256 or Nr = 512, Nθ = 64 or Nθ = 128, and Nϕ = 64 or Nϕ = 32 behaved similarly for disks with similar resolution across the disk. Our resolution of the MRI and prior convergence testing by Penna et al. (2010) for similarly resolved disks justify our choice of grid resolution. It is currently not computationally feasible to perform a similar spin parameter study at much higher resolutions, and future studies will continue to explore whether such simulations are fully converged (Hawley et al. 2011; McKinney et al. 2012).

3.5. Ceiling Constraints

During the simulation, the rest-mass density and internal energy densities can become low beyond the corona, but the code remains accurate and stable for a finite value of b2/ρ, b2/ugas, and ugas/ρ for any given resolution. We enforce b2/ρ ≲ 10, b2/ugas ≲ 100, and ugas/ρ ≲ 10 by injecting a sufficient amount of mass or internal energy into a fixed zero angular momentum observer (ZAMO) frame with four-velocity uμ = { − α, 0, 0, 0}, where $\alpha =1/\sqrt{-g^{tt}}$ is the lapse.

We have checked that the ceilings are rarely activated in the regions of interest of the flow. Figure 2 shows the constrained ratios, b2/ρ, b2/ugas, and ugas/ρ, as a function of θ at six radii (r = 4, 6, 8, 10, 12, and 14M) for the a* = 0 model. The data have been time-averaged over the steady-state period from t = 14, 000M to 20, 000M. The ceiling constraints are shown as dashed red lines. The solution stays well away from the ceilings. Thus, the ceilings are sufficiently high.

Figure 2.

Figure 2. Ratios of b2/ρ, b2/ugas, and ugas/ρ vs. θ. Black curves correspond to different radii in the flow; from top to bottom, r = 4, 6, 8, 10, 12, and 14M. The data are time-averaged over the steady state period of the flow, from t = 14, 000M to 20, 000M. Numerical ceilings constrain the solution to lie below the dashed red lines, but we see that the solution does not approach these limits.

Standard image High-resolution image

3.6. Approach to Steady State

We run the simulations from t = 0M to t = 20, 000M. The accretion rate, the height- and ϕ-averaged plasma β, and other disk parameters fluctuate turbulently about their mean values. The simulation reaches a quasi-steady state when the mean parameter values are time independent. Figure 3 shows the accretion rate and height- and ϕ-averaged 1/β at the event horizon as a function of time for all five models. We take the period from t = 14, 000M to t = 20, 000M to define steady state.

Figure 3.

Figure 3. Accretion rate and height- and ϕ-averaged σ = pmag/pgas = 1/β vs. time at the event horizon for all five models: a* = 0 (solid light cyan), a* = 0.5 (solid dark red), a* = 0.7 (long-dashed green), a* = 0.9 (short-dashed brown), and a* = 0.98 (dot-dashed orange).

Standard image High-resolution image

As shown in Penna et al. (2010), for disk models like the one considered, the disk outside the innermost stable circular orbit (ISCO) behaves like the α-disk model with α ∼ 0.1 across disk thicknesses of h/r ∼ 0.05–0.4. This allows one to accurately infer the timescale for reaching "inflow equilibrium," corresponding to a quasi-steady flow across all quantities, at a given radius. For h/r ∼ 0.3 by t ∼ 15, 000M–20, 000M (the simulation runs until 20, 000M, but the initial 5000M are transients not necessarily associated with achieving inflow equilibrium for a simple viscous disk), we use the results in Appendix B of Penna et al. (2010) and find that inflow equilibrium is achieved within a radius of r ∼ 25M–30M for models with a* ∼ 1 and r ∼ 35M for models with a ∼ 0. Even for a doubling of the viscous timescale, inflow equilibrium is achieved by r ∼ 20M–25M depending on the BH spin. This motivates using an analytical extension of the simulation solution for radii beyond r ∼ 25M as described later in Section 4.1.

3.7. Evolved Disk Structure

Figure 4 shows matter stream lines as vectors and number density ne as a gray-scale map. The large-scale vortices existing on a single time shot (a) almost disappear when averaged over the duration 6000M (b) from times 14, 000M to 20, 000M. The density is highest in the equatorial plane on average, but deviations are present on the instantaneous map. The ISCO does not have any special significance: density and internal energy density increase through ISCO toward the BH horizon.

Figure 4.

Figure 4. Stream lines of velocity (red vectors) and number density ne (gray-scale map) for spin a* = 0.9 at ϕ = 0 in the meridional plane (rc as cylindrical radius): single time snapshot at t = 14, 000M (a) and time average between t = 14, 000M and t = 20, 000M (b). The corresponding calibration bars of ne are shown on the right. Number density is normalized by its maximum value, and the vectors show the poloidal velocity direction.

Standard image High-resolution image

Figure 5 shows magnetic field lines as vectors and comoving electromagnetic energy density ∝b2 as a gray-scale map. The structure of magnetic field at early times remembers the initial multi-loop field geometry (Penna et al. 2010) but switches at late times to a helical magnetic field structure resembling a split monopole in meridional projection. Such switching of magnetic field structure suggests that the final helix with projected split monopole is a natural outcome of any vertical flux being dragged into the BH (although the amount of magnetic flux threading the hole and disk may be chosen by initial conditions as described in McKinney et al. 2012). The magnetic field structure of a single snapshot (a) looks similar to the structure of the linear average between 14, 000M and 20, 000M (b). The polar region of the flow has the strongest magnetic field. The magnetic field lines on Figure 5 illustrate only the direction of the field's poloidal component. The toroidal magnetic field is stronger above and below the midplane of the disk outside of ISCO. The toroidal field strength is comparable to the poloidal field strength inside the ISCO and near the disk midplane.

Figure 5.

Figure 5. Magnetic field lines (red vectors) and comoving electromagnetic energy density ∝b2 (gray-scale map) for spin a* = 0.9 at ϕ = 0 in the meridional plane (rc as cylindrical radius): single time snapshot at t = 14, 000M (a) and time average between t = 14, 000M and t = 20, 000M (b). The corresponding calibration bars of comoving b2 are shown on the right. Magnetic field energy density is normalized by its maximum value. The magnetic field lines illustrate only the direction of the field's poloidal component.

Standard image High-resolution image

4. DYNAMICAL MODEL BASED ON SIMULATIONS

We now discuss extensions of the numerical simulations, which we need to perform radiative transfer computations. We extend the simulations to large radii and define the electron temperature.

4.1. Extension to Large Radius

The flow is evolved in a quasi-steady state for 6000M from 14, 000M until 20, 000M, which corresponds to eight orbits at r = 25M. The flow is not sufficiently settled at larger radii. However, outside 25M, some Faraday rotation and emission might occur. So, we extend the dynamical model to larger radii (i.e., r > 25M) in a continuous way and check (see the Appendix) how variations of our large radius prescriptions change the results of radiative transfer. The outer radial boundary of radiative transfer is situated at r = 20, 000M. The profiles of number density ne, internal energy density ugas, magnetic field $\bf b$, and velocity $\bf v$ are extended as power laws until radius r = 20, 000M. The power-law index for number density β is obtained by matching the known value ne = 130 cm−3 at about 1farcs5 ≈ 3 × 105M (Baganoff et al. 2003) and the average ne, cut value at r = 25M in the equatorial plane for each model. The value of β may be different for different models. The radial flow velocity vr is then obtained from the continuity relation in the equatorial plane nevrr2 = const. The power law of internal energy density ugas is obtained in a similar way by matching the values Te = Tp = 1.5 × 107 K and ne = 130 cm−3 at distance 3 × 105M (Baganoff et al. 2003; Shcherbakov & Baganoff 2010). The meridional physical velocity is extended as $v_{\hat{\theta }}\propto (r/M)^{-3/2}$ and toroidal velocity as $v_{\hat{\phi }}\propto (r/M)^{-1/2}$ to approximately match the power law between 15M and 25M, where the relationship $v_{\hat{i}}\approx u^i\sqrt{g_{ii}}$ is used to connect the four-velocity components with physical velocity components. All components of comoving magnetic field are extended as br, bθ, bϕ∝(r/M)−1, which appears valid across a diverse set of GRMHD models (McKinney et al. 2012). This power-law slope corresponds roughly to equipartition of magnetic field energy density, since constant fraction magnetic field is $b\propto \sqrt{n T_p}\propto (r/M)^{-1}$ for n∝(r/M)−1. Exploration of various extensions of the magnetic field will be the topic of future studies.

After defining the extension power laws for quantities in the equatorial plane, we extend the quantities radially at arbitrary θ and ϕ in a continuous way. For example, for density at arbitrary θ and ϕ and at r > 25M we have

Equation (10)

where ne(25M, θ, ϕ) is taken from the simulations. We similarly extend other quantities. As shown in the Appendix, small variations in power-law indices of number density and temperature have little influence on radiation intensities and LP/CP fluxes, but variations of magnetic field slope can make a substantial difference.

4.2. Electron Temperature

Neither the proton Tp nor the electron Te temperature is given directly by the simulations. However, it is crucial to know the electron temperature Te to determine the emission. Our solution is to split the total internal energy density ugas, given by the simulations and their power-law extension, between the proton energy and the electron energy. The energy balance states

Equation (11)

where cp = 3/2 and ce ⩾ 3/2 are the respective heat capacities, ρ is the rest-mass density, and kB is Boltzmann's constant. The difference of temperatures TpTe is influenced by three effects: equilibration by Coulomb collisions at large radii, the difference in heating rates fp and fe of protons and electrons operating at intermediate radii, and the difference in heat capacities operating close to the BH. Radiative cooling is ignored since, according to Sharma et al. (2007), the radiative efficiency of the flow is negligible for realistic $\dot{M}\lesssim 10^{-7}\ M_\odot \ {\rm {{\rm yr}}}^{-1}$. The relevant effects can be summarized by the equation

Equation (12)

where

Equation (13)

is the non-relativistic temperature equilibration rate by collisions (Shkarofsky et al. 1966), all quantities being measured in cgs units. We consider protons to always have non-relativistic heat capacity and collisions to always obey the non-relativistic formula. The magnitudes of errors introduced by these simplifications are negligible. The exact expressions for total electron heat capacity and differential heat capacity are approximated as

Equation (14)

Equation (15)

correspondingly, with the error <1.3%, where

Equation (16)

is the dimensionless electron temperature. It was recently shown (Sharma et al. 2007) that the ratio of heating rates in the non-relativistic regime in a disk can be approximated as

Equation (17)

with coefficient C. This formula is adopted in the relativistic regime as well, since no better prescription is available. Sharma et al. (2007) found the value C = 0.33 in simulations, whereas we find C = 0.36–0.42 for the best-fit models (see Table 2 and Section 6).

Table 2. Properties of the Best-fit Models with Different Spins

Model Inclination Spin Position Angle, Heating Ratio Tp/Te Electron Te Accretion Rate
  Angle, θ (deg) P.A. (deg) Constant, C at 6M at 6M (K) $\dot{M}$ (M yr−1)
Spin a* = 0 42.0 171.0 0.42107 15.98 3.343 × 1010 7.005 × 10−8
Spin a* = 0.5 74.5 115.3 0.37012 20.14 3.087 × 1010 4.594 × 10−8
Spin a* = 0.7 64.5 84.7 0.37239 20.16 3.415 × 1010 2.694 × 10−8
Spin a* = 0.9 53.5 123.4 0.39849 18.16 4.055 × 1010 1.402 × 10−8
Spin a* = 0.98 57.2 120.3 0.41343 17.00 4.190 × 1010 1.553 × 10−8
Spin a* = 0.5 short period 1 70.0 79.3 0.38934 18.50 3.334 × 1010 3.513 × 10−8
Spin a* = 0.5 short period 2 72.8 113.1 0.40507 17.31 3.541 × 1010 3.452 × 10−8
Spin a* = 0.5 short period 3 73.4 57.4 0.37302 19.87 3.125 × 1010 3.897 × 10−8
Spin a* = 0.5 short period 4 74.4 115.4 0.36147 20.95 2.978 × 1010 4.508 × 10−8
Spin a* = 0.5 short period 5 71.9 95.7 0.37420 19.79 3.137 × 1010 5.334 × 10−8
Spin a* = 0.5 short period 6 76.4 116.7 0.38853 18.59 3.320 × 1010 6.080 × 10−8
Spin a* = 0 fast light 41.4 187.5 0.41929 16.09 3.322 × 1010 7.044 × 10−8
Spin a* = 0.5 fast light 72.7 105.9 0.39804 17.83 3.447 × 1010 3.957 × 10−8
Spin a* = 0.7 fast light 59.4 131.8 0.35708 21.62 3.204 × 1010 2.966 × 10−8
Spin a* = 0.9 fast light 53.3 123.3 0.40215 17.86 4.116 × 1010 1.340 × 10−8
Spin a* = 0.98 fast light 57.7 119.6 0.41720 16.73 4.246 × 1010 1.515 × 10−8

Notes. Mean values are shown for ratio Tp/Te, electron temperature Te, and the accretion rate $\dot{M}$. These are the simple means over all simulation snapshots, which were employed for radiative transfer in a particular model. The values of χ2/dof range from 2 to 5 across all models.

Download table as:  ASCIITypeset image

The proton and electron temperatures are determined at each point in the following way. We first take a single snapshot of a simulation with spin a* and extend the flow quantities to r = 20, 000M (see Section 4.1). Then we compute azimuthal averages of radial velocity vr, number density ne, and ugas/ρ at the equatorial plane, extend them as power laws to rout = 3 × 105M, and solve the Equations (11) and (12) from rout down to the inner grid cell point. Temperatures are set to Te = Tp = 1.5 × 107 K at rout (Baganoff et al. 2003; Shcherbakov & Baganoff 2010). In the next step we compare the values of ugas/ρ to the calculated Te and Tp and determine the functional dependence Te = Te(ugas/ρ) and Tp = Tp(ugas/ρ). At each point of the simulation (including off the equator), we draw temperatures from this correspondence. That is, GRMHD simulation directly provides ugas and ρ at the equatorial plane, so the function Te = Te(ugas/ρ) gives Te at each point in space. Typical profiles of proton and electron temperatures are shown in Figure 6. Temperatures stay equal until r ∼ 104M due to collisions, despite different heating prescriptions. Within r = 3 × 103M the timescale of collisional equilibration becomes relatively long and electrons become relativistic; thus, Te deviates down from Tp. The electron and proton temperature profiles in the region r < 20, 000M are used to conduct the radiative transfer. For a given accretion rate, we find that there exists a unique dependence of the ratio of temperatures Tp/Te (measured at r = 6M at the equator) on the heating coefficient C, so that we can use Tp/Te and C interchangeably.

Figure 6.

Figure 6. Temperatures of protons Tp (upper red line) and electrons Te (lower blue line) for the dynamical model with spin a* = 0.5 giving the best fit to polarization observations (see Table 2 and Section 6).

Standard image High-resolution image

5. GENERAL RELATIVISTIC POLARIZED RADIATIVE TRANSFER

5.1. Description of Radiative Transfer

Now we convert the dynamical model of the accretion flow into a set of observable quantities using polarized radiative transfer (Broderick et al. 2009; Shcherbakov & Huang 2011). We closely follow Shcherbakov & Huang (2011) for the transfer technique. Similar to Huang et al. (2009a), we define the polarized basis in the picture plane, where one vector points north, another vector points east, and the wavevector points toward the observer. We parallel transport this basis in the direction of the BH and do the radiative transfer along the ray in the opposite direction toward the observer. At each point along the ray we go into the locally flat comoving frame, calculate the angles between the magnetic field and basis vectors, and compute the Faraday conversion, Faraday rotation, emissivities, and absorptivities.

Radiative transfer involves shooting a uniform grid of PN × PN geodesics from the picture plane down to the BH. The total polarized fluxes are computed by integration of intensities along each ray backward to the picture plane. We found that PN = 111 is good enough to compute the spectrum (Dexter et al. 2009 used PN = 150). For radiative transfer we employ all 3D data in each numerical simulation snapshot and, following Moscibrodzka et al. (2009), perform multilinear interpolation in three dimensions for the quantities in between the grid points. We make no approximations in the use of spatial 3D data. We self-consistently take into account the evolution of the numerical simulation as the light geodesics travel around the BH. Since it is too time-consuming to look up simulation data over a long period of time, we only evolve the simulation between t − Δt and t + Δt to get a spectrum at time t + 20, 000M. The offset 20, 000M appears, since the picture plane is located 20, 000M away from the BH center. The extension to the large radius outside 25M, however, is not evolved with time. It is taken to be that of a single snapshot at time t. The snapshots at times t − Δt and t + Δt are taken to represent the numerical simulations at earlier and later times, respectively. We find that Δt = 60M is large enough to achieve accurate simulated spectra. The total fluxes are found at regular time intervals within a period of quasi-steady accretion from 14, 000M until 20, 000M, e.g., for t = 14, 000M, 14, 300M, ..., 19, 700M, 20, 000M. We compute Nperiods = 21 spectra over the quasi-steady accretion phase and average them to find the mean simulated spectra. To compute the polarized fluxes, we take the integration domain in the picture plane to be a square with a side

Equation (18)

in the units of rgM, where frequency ν is in GHz. This square is centered at the BH. The size based on Equation (18) is larger than the photon orbit visible diameter dph ≈ 10.4M and follows the intrinsic size dependence on frequency (Shen et al. 2005; Doeleman et al. 2008) at low frequencies. An important radiative transfer parameter is the distance from the BH, where intensity integration starts. The dependence of synchrotron emissivity on temperature and magnetic field strength is so strong that it overwhelms the sole effect of gravitational redshift close to the BH. We obtain accurate results in the submillimeter for computation out from rmin = 1.01rH, where $r_H=M(1+\sqrt{1-a_*^2})$ is the horizon radius. To quantify the needed accuracy of computations, we define a quantity χ2H/dof in the Appendix. We conduct multiple tests of radiative transfer convergence for best-fit models at each spin. In the Appendix, we justify the chosen values of radiative transfer parameters PN, Δt, Nperiods, rmin, etc.

Our calculation of plasma response is different from that of Shcherbakov & Huang (2011). They offered a way to find exact emissivities, absorptivities, Faraday rotation, and conversion coefficients for thermal and other isotropic particle distributions. Here, for simplicity, we employ fitting formulae for Faraday rotation and Faraday conversion and synchrotron approximation for emissivities for a thermal plasma. We define

Equation (19)

where θB is the kb angle, γ is the electron gamma factor, and νB = eb/(2πmec) is the cyclotron frequency. Then following Legg & Westfold (1968) and Melrose (1971), we write down emissivities in the I, Q, and V modes as

Equation (20)

Here Kz(x) is the Bessel function of the second kind of order z. We employed IEEE/IAU definitions of Stokes Q, U, and V (Hamaker & Bregman 1996), and we define counterclockwise rotation of the electric field as seen by the observer as corresponding to positive V > 0—as also chosen in Shcherbakov & Huang (2011). So, the sign of the V emissivity (Equation (20)) is opposite to the sign in Rybicki & Lightman (1979). A variation of emissivity formulae (19) and (20) exists: Sazonov (1969) and Pacholczyk (1970) define X = 2ν/(3νB(γ − 1)2sin θB), integrating over particle energy instead of γ. This approximation appears to give significantly larger errors at low particle energies.

Next, one needs to identify the accurate thermal particle distribution N(γ). Various N(γ) correspond to various synchrotron approximations. The ultrarelativistic thermal approximation (Pacholczyk 1970; Huang et al. 2009a) has the simplest distribution N(γ) = exp (− (γ − 1)/θe)(γ − 1)2/2/θ3e. However, the exact thermal distribution of particles

Equation (21)

allows for more precise computation of radiation. Synchrotron emissivities based on Equations (19) and (20) with the exact thermal distribution (21) agree with the exact cyclo-synchrotron emissivities εI, εQ, and εV (Leung et al. 2011; Shcherbakov & Huang 2011) to within 2% for typical dynamical models and frequencies >100 GHz. Emissivities integrated over the ultrarelativistic thermal distribution typically have ∼10% error.

Thermal absorptivities are found from emissivities (Equation (20)) via Kirchhoff's law

Equation (22)

where Bν = 2kBTeν2/c2 is the source function for low photon energies (hν ≪ kBTe). Faraday rotation ρV and Faraday conversion ρQ coefficients are taken from Shcherbakov (2008):

Equation (23)

Here

Equation (24)

and

Equation (25)

are the fitting formulae for deviations of ρV and ρQ from analytic results for finite ratios of νB/ν. The deviation of f(Z) from 1 is significant for the set of observed frequencies ν, temperatures θe, and magnetic fields found in the typical models of Sgr A*. These formulae constitute a good fit to the exact result for the typical parameters of the dynamical model (Shcherbakov 2008).

Polarized radiative transfer can take much longer to perform compared to non-polarized radiative transfer when using an explicit integration scheme to evolve the Stokes occupation numbers NQ, NU, and NV. Large Faraday rotation measure (RM) and Faraday conversion measure lead to oscillations between occupation numbers. One of the solutions is to use an implicit integration scheme, while another solution is to perform a substitution of variables. In the simple case of Faraday rotation leading to interchange of NQ and NU, our choice of variables is the amplitude of oscillations and the phase. Thus, the cylindrical polarized coordinates arise as follows:

Equation (26)

Then, the amplitude NQU slowly changes along the ray and the angle ϕ changes linearly, and this translates into a speed improvement. In the presence of substantial Faraday conversion, the polarization vector precesses along some axis on a Poincaré sphere, adding an interchange of circularly and linearly polarized light. So, polar polarized coordinates are more suitable in this case:

Equation (27)

where Npol is the total polarized intensity, ϕ angle changes are mainly due to Faraday rotation, and ψ angle changes are mainly due to Faraday conversion. The application of this technique speeds up the code enormously at low frequencies of ν < 100 GHz.

5.2. Search for the Best Fits

We define χ2/dof quantities to discriminate between models. We define χ2F for fitting total fluxes as

Equation (28)

for the set of seven frequencies ν = 88, 102, 145, 230, 349, 680, and 857 GHz, where σ(F) are the errors of the means. We incorporate LP fractions at 88, 230, and 349 GHz and CP fractions at 230 and 349 GHz to obtain

Equation (29)

Then we define dof (as degrees of freedom) to be dofF = 7 − 3 = 4 for flux fitting and dof = 12 − 3 = 9 for fitting all polarized data. The quantity χ2/dof would be drawn from χ2 statistics if σ values were the true observational errors and if the observed fluxes were drawn from a Gaussian distribution. However, for the purpose of the present work, we only employ χ2/dof as a measure of fitting the data. That is, lower χ2/dof indicates better agreement with the data. We do not attempt to ascribe any statistical meaning to the quantity χ2/dof.

We explore models with four parameters: spin a*, inclination angle θ, accretion rate $\dot{M}$, and the ratio of proton to electron temperature Tp/Te (Tp/Te is reported for radius r = 6M). For the radiative transfer calculations, the density from the simulations is scaled to give the desired accretion rate.

6. RESULTS

In previous sections, we described our compiled observations, GRMHD numerical simulations of the flow structure, our method for obtaining the electron temperature, and our method for polarized radiative transfer. In this section, we discuss our results for accretion flow and BH parameters, as guided by a minimization of χ2/dof for our model applied to the observations.

Figure 7 shows best fits to observations by models with five different spins. Inclination angle θ, accretion rate $\dot{M}$, and heating coefficient C were adjusted to reach the lowest χ2/dof. Fits to fluxes Fν (upper left) are not substantially different, although models with higher spins fit better at high frequencies. Larger deviations can be seen on LP (lower left) and CP (lower right) plots. Models with high spins require a lower accretion rate (i.e., density) to fit the flux spectrum. As a consequence, they are not subject to Faraday depolarization, which leads to a decrease of LP at low ν, and the models end up having larger LP fractions at 88 GHz. Not all models reproduce the observed decrease of mean LP fraction between 230 and 349 GHz groups. The discrepancies in fitting the CP fraction are also large: all the lowest χ2 models give |CP| < 1.5% at 349 GHz. The best-bet model with spin a* = 0 reproduces LP and CP fractions well but fails in fitting the total flux. Most solutions predict the wrong sign of the EVPA(349 GHz)–EVPA(230 GHz) difference, which could be fixed with stronger magnetic field (e.g., as seen in models by McKinney et al. 2012) to yield stronger Faraday rotation. In sum, crude agreement of simulated polarized spectra with the observed ones was achieved, but the improved dynamical models may be needed for better fits.

Figure 7.

Figure 7. Fits to the observed fluxes and LP and CP fractions by best models for each spin. The inclination angle θ, accretion rate $\dot{M}$, a ratio of temperatures Tp/Te were adjusted for each spin to minimize χ2/dof. Fits to total flux Fν are in the upper left panel, LP fraction in the lower left, and CP fraction in the lower right. Shown are the best-fit models with spin a* = 0 (short-dashed brown), spin a* = 0.5 (solid dark red), spin a* = 0.7 (long-dashed green), spin a* = 0.9 (solid light cyan), and spin a* = 0.98 (dot-dashed orange). The upper right panel shows the dependence of EVPA on frequency for the best models. Note that EVPAs are not included in our fitting procedure. The thick blue curve represents observations. Simulated EVPA curves are arbitrarily shifted to approximate EVPA at 349 GHz. The addition of an external (to the emitting region) Faraday rotation screen helps to fit EVPA(349 GHz)–EVPA(230 GHz).

Standard image High-resolution image

We now isolate the physical effects responsible for the observed polarized quantities for our best-bet model with spin a* = 0.5 that has the lowest χ2/dof (see Section 6.1).

There are several radiative transfer effects that contribute similarly to the polarized fluxes. Let us consider the production of CP in the flow. Figure 8 shows the consequence of switching off each physical effect for our best-bet model with spin a* = 0.5. The solid red curve is the result with all physics on. The dot-dashed orange line below is for zero circular emissivity having εV = 0. The brown dashed line corresponds to zero Faraday conversion (ρQ = 0). Switching off εV emissivity leads to a minor correction, whereas setting Faraday conversion to zero results in CP of the opposite sign with several times smaller absolute value. Most of the CP in this model is produced by Faraday conversion. It would be incorrect, however, to think that the simple linear-to-circular conversion explains the observed CP. The dashed green line in Figure 8 shows the CP fraction, when Faraday rotation is switched off (ρV = 0). The effect of Faraday rotation is insignificant at ν > 350 GHz, but the rotation of the plane of LP simultaneous with conversion between LP and CP produces a unique effect at lower ν. This is the so-called rotation-induced conversion (Homan et al. 2009). Sign oscillations of V with frequency do not happen when the Faraday rotation is on, but they do happen when ρV = 0. For the best-fit model it is the rotation-induced Faraday rotation, which is responsible for most of the circularly polarized light.

Figure 8.

Figure 8. Contributions of different effects to the CP fraction as a function of frequency for our best-bet model with BH spin a* = 0.5. Shown are observations (blue error bars), the best-bet model (solid red line), the same dynamical model computed with zero V emissivity (εV = 0) in radiative transfer so that CP is produced by Faraday conversion (dot-dashed orange), the same model with zero Faraday conversion (ρQ = 0) (short-dashed brown), and the same model with zero Faraday rotation (ρV = 0) (long-dashed green). Emissivity in circular V mode contributes little to the observed CP, which is mainly due to Faraday conversion.

Standard image High-resolution image

In Figure 9 we illustrate the influence of Faraday rotation on LP fraction (left panel) and EVPA (right panel). The solid curves are produced with all physics on for our best-bet model with spin a* = 0.5. The green dashed lines are computed when switching off Faraday rotation (ρV = 0). The Faraday rotation is small at high frequencies, and LP curves look similar at ν > 200 GHz. As the rotation of polarization plane is much stronger at low ν, a significant phase shift accumulates between different rays at the low end of the spectrum and cancellations of LP become strong at ν < 150 GHz. This illustrates the effect of Faraday depolarization (Bower et al. 1999a). In the absence of Faraday rotation, the dependence of EVPA on frequency is not constant: the variations of intrinsic emitted EVPA are significant. Thus, the change of EVPA with ν should not always be ascribed to the effect of Faraday rotation. The positive observed slope of EVPA with ν at high ν, acquired due to negative Faraday rotation measure (RM < 0), is comparable to the slope of intrinsic emitted EVPA.

Figure 9.

Figure 9. Contributions of different effects to the LP fraction (on the left) and EVPA (on the right) as functions of frequency for the best-bet model with spin a* = 0.5. Shown are observations (blue error bars and thick blue line), the best-bet model (solid red line), and the same dynamical model computed with zero Faraday rotation (ρV = 0) in radiative transfer (long-dashed green). Beam depolarization is weak: if Faraday rotation is absent, then LP stays high at low frequencies. Even when the Faraday rotation is set to zero, the EVPA depends on frequency due to varying intrinsic emission EVPA. Faraday rotation in the best-bet model is too weak to reproduce EVPA observations, so stronger magnetic fields or more magnetic flux near the black hole than in the simulations may be required.

Standard image High-resolution image

There is an alternate way to test dynamical models against observations. The intrinsic image size was recently measured (Doeleman et al. 2008) with the VLBI technique. The measured correlated flux at 230 GHz was Fcorr ≈ 0.35 Jy at the 3.5Gλ SMT-JCMT baseline. Similar values of correlated flux were observed later by the same group (Fish et al. 2011). We plot this correlated flux with 3σ error bar in Figure 10 and compare it to simulated correlated fluxes. To simulate the correlated flux, we follow Fish et al. (2009) and employ a Gaussian interstellar scattering ellipse with half-widths at half-maximum 7.0  ×  3.8Gλ with position angle (P.A.) 170° east of north. The correlated fluxes for the best-fit models with spin a* = 0.5 (darker red lines) and a* = 0.98 (lighter orange lines) are shown. We vary the P.A. of the BH spin axis and plot correlated flux curves with the largest (upper solid lines) and the smallest (lower dashed lines) correlated flux at 3.5Gλ. Since we do not fit EVPA directly, models with different P.A. have the same χ2/dof. The size in our best-bet model with spin a* = 0.5 is consistent with observations, whereas the best-fit model with spin a* = 0.98 has larger correlated flux, so that the size of the shadow is slightly underpredicted. The simulated source size is in crude agreement to the observed one.

Figure 10.

Figure 10. Correlated flux as a function of baseline at 230 GHz normalized to the averaged observed flux at 2.82 Jy for the best-fit models with spin a* = 0.5 (darker red lines) and a* = 0.98 (lighter orange lines). The upper solid lines show the smallest size (largest correlated flux) over all position angles of BH spin axis, and the lower dashed lines show the largest size (smallest correlated flux) over all position angles. An observational result presented in Doeleman et al. (2008) with 3σ error bars at baseline 3.5Gλ is depicted as a vertical black bar for comparison. The size in our best bet model with spin a* = 0.5 is consistent with observations, whereas the best-fit model with spin a* = 0.98 has larger correlated flux, so that the size of the shadow is slightly underpredicted.

Standard image High-resolution image

Table 2 summarizes the properties of several best-fit models. Rows 1–5 show the model parameters for best fits with spins from a* = 0 to a* = 0.98. The simulated spectra are computed every 300M from t = 14, 000M until t = 20, 000M for Δt = 60M. Then Nperiod = 21 spectra are averaged to compare to observations. Rows 6–11 show the model parameters for models with spin a* = 0.5 for spectra averaged over shorter periods. That is, Nperiod = 21 spectra are computed from t = 14, 000M until t = 15, 000M for the first short period, while the second short period covers the time interval from t = 15, 000M until t = 16, 000M, etc. When comparing the best-fit models with spin a* = 0.5 computed over different simulation periods, we find variations in inclination angle Δθ = 3° from the mean, the electron temperature ΔTe/Te = 10%, and the accretion rate $\Delta \dot{M}/\dot{M}=30\%$. The spin position angle varies by as much as ΔP.A. = 30°.

The last five rows in Table 2 show the model parameters for best fits within the "fast light" approximation. In this approximation, simulated spectra are computed over single frozen snapshots, e.g., for Δt = 0. When the fast light approximation is used instead of the correct simultaneous evolution of photon field and MHD, the models with spins a* = 0, 0.9, 0.98 produce almost identical best fits with variations Δθ < 0.6°, ΔTe/Te < 1.5%, and $\Delta \dot{M}/\dot{M}<5\%$. However, the models with a* = 0.5, 0.7 settle to different χ2/dof minima with larger changes in quantities: Δθ = 5°, ΔTe/Te = 10%, $\Delta \dot{M}/\dot{M}=10\%$. These variations are still smaller than variations between models with different spins. Switching to the fast light approximation results in significant changes Δχ2/dof ∼ 1 between the best-fit models for the same spins, which emphasizes the importance of precise radiation transfer calculations.

6.1. Model Parameters

We now discuss the estimated parameters obtained for the best-fit models. The best-bet model with spin a* = 0.5 has inclination angle θ = 74fdg5, mean accretion rate $\dot{M}=4.6\times 10^{-8}\ M_\odot \ {\rm {{\rm yr}}}^{-1}$, and ratio of temperatures Tp/Te = 20.1 at r = 6M, which gives Te = 3.1 × 1010 K at r = 6M in the equatorial plane. The best-fit models with other spins give the inclination angles: θ = 42°, 64fdg5, 53fdg5, 57fdg2 at a* = 0, 0.7, 0.9, 0.98, respectively. Thus, the inclination angle for the five models lies within θ = 42°–75°. Our modeling favors neither edge-on nor face-on orientations. The electron temperature Te at r = 6M is surprisingly uniform over a set of best-fit models. All five best-fit models with spins from a* = 0 to a* = 0.98 presented in Table 2 have electron temperature within the tight range

Equation (30)

The accretion rate depends strongly on spin. The model with spin a* = 0 has an accretion rate $\dot{M}=7.0\times 10^8\ {M_\odot \ \rm {{\rm yr}}}^{-1}$, which is five times larger than the accretion rate $\dot{M}=1.4\times 10^8\ {M_\odot \ \rm {{\rm yr}}}^{-1}$ for the model with spin a* = 0.9. Higher spin values give lower accretion rates. A natural outcome of fitting a polarized spectrum is the P.A. of the BH spin axis. Similar to Huang et al. (2009a), we rely on the observed intrinsic EVPA ≈111fdg5 at 230 GHz and EVPA ≈146fdg9 at 349 GHz (see Section 2). For the model to fit the difference in EVPA, we add a Faraday rotation screen far from the BH with constant RM. Then we compute the required RM and the intrinsic P.A. to fit the simulated EVPAs at 230 and 349 GHz. The best-bet model with a* = 0.5 gives P.A. = 115fdg3 east of north, whereas the next best-fit model with spin a* = 0.98 requires P.A. = 120fdg3. However, P.A. is different by 90° between the models with spin a* = 0 and a* = 0.7, which indicates that P.A. can lie within a wide range. In sum, some parameters, such as Te, are estimated to be in narrow ranges, while only order-of-magnitude estimates are available for other parameters, such as $\dot{M}$.

With the estimated orientation of the BH spin axis, we can plot an image of average radiation intensity from near the event horizon. Figure 11 shows images of total intensity Iν for the best-bet model with spin a* = 0.5 (upper left panel), the best fit for spin a* = 0.98 (lower left panel), and LP intensity and CP intensity plots for the best-bet model with a* = 0.5 (upper right and lower right panels, correspondingly). The LP average intensity plot was made by averaging U and Q intensities separately and then finding the total LP fraction and EVPA. Blue (predominant) color on the CP plot depicts the regions with negative CP intensity, and red (scarce) color depicts the regions with positive CP intensity. The total V flux from this solution is negative (V < 0). The streamlines on the LP plot are aligned with EVPA direction at each point. The spin axis is rotated by P.A. = 115fdg3 east of north for the best-bet model with spin a* = 0.5 and by P.A. = 120fdg3 for the best-fit model with spin a* = 0.98. The spin axis is inclined at θ to the line of sight, so that either the right (west) or left (east) portions of the flow are closer to the observer. The color schemes on all plots are nonlinear with corresponding calibration bars plotted on the sides. The numbers at the top of calibration bars denote normalizations.

Figure 11.

Figure 11. Images of polarized intensities for the best-fit models: total intensity for spin a* = 0.98 model (lower left); intensities for a* = 0.5 model: total intensity (upper left), linear polarized intensity and streamlines along EVPA (upper right), and circular polarized intensity (lower right). Distances are in the units of BH mass M. Images are rotated in the picture plane to correspond to the best spin P.A.: P.A. = 115fdg3 for the a* = 0.5 model and P.A. = 120fdg3 for the a* = 0.98 model. Individual calibration bars are on the sides of corresponding plots. The ill-defined polar region does not contribute significantly to the emission.

Standard image High-resolution image

7. DISCUSSION AND CONCLUSIONS

Let us compare our results with estimates of Sgr A* accretion flow and BH parameters made by other researchers.

Two separate searches for spin based on GRMHD numerical simulations have been reported so far: Moscibrodzka et al. (2009) and Dexter et al. (2010). Moscibrodzka et al. (2009) considered the set of spins from a* = 0.5 to a* = 0.98 for 2D GRMHD simulations and then fitted the X-ray flux, the 230 GHz flux, and the flux slope at 230 GHz. They found that at least one model for each spin is crudely consistent with the observations (see their Table 3), and their best-bet model has a* = 0.9. Dexter et al. (2010) focused on a set of 3D GRMHD and then fitted the 230 GHz flux and size estimates, and they provided a table of spin probabilities with a* = 0.9 having the highest P(a). When we fitted the spectrum and LP/CP fractions, the model with a* = 0.5 has the lowest χ2/dof. As for these two groups, we are also unable to provide a statistically significant constraint on a*. Other spin estimates have been based on analytic models. Broderick et al. (2009, 2011) favor a* = 0 solutions, while Huang et al. (2009b) favor a* < 0.9 (although they do not explore their full model parameter space).

Another poorly constrained quantity is the mass accretion rate. Our estimate $\dot{M}_{\rm est}=(1.4\hbox{--}7.0)\times 10^{-8}\ M_\odot \ {\rm {{\rm yr}}}^{-1}$ is broad. Acceptable models in Moscibrodzka et al. (2009) give $\dot{M}$ from 0.9 × 10−8 to $12\times 10^{-8}\ M_\odot \ {\rm {{\rm yr}}}^{-1}$, which agrees with our range. Dexter et al. (2010) reported a 90% confidence interval of $\dot{M}$ for spin a* = 0.9 solutions, while incorporating flow size in χ2 analysis. Our estimates have somewhat higher accretion rates than the range $\dot{M}=5^{+15}_{-2}\times 10^{-9}\ M_\odot \ {\rm {{\rm yr}}}^{-1}$ (90%) in Dexter et al. (2010) because models with lower spin naturally need higher $\dot{M}$ to fit the data. Note that Dexter et al. (2009) found an even lower accretion rate $\dot{M}(a_*=0.9)=(1.0\hbox{--}2.3)\times 10^{-9}\ M_\odot \ {\rm {{\rm yr}}}^{-1}$ when they assumed equality of proton and electron temperatures (Tp = Te).

In addition to spin and accretion rate, we can try to estimate the inclination angle θ and electron temperature Te (Te is reported at r = 6M in the equatorial plane). Our range is θest = 42°–75°, which agrees with estimates by other groups. Broderick et al. (2009) and Dexter et al. (2010) reported θ ∼ 50°. Huang et al. (2009a, 2009b) favor slightly lower θ ∼ 40° and 45°, respectively, but they have large error bars. To estimate Te, Moscibrodzka et al. (2009) and Dexter et al. (2010) used a constant Tp/Te, whereas Huang et al. (2009a) and the present work calculated the profile of Te. In all models, Te is a shallow function of radius, which made Dexter et al. (2010) estimate a "common" Te = (5.4 ± 3.0) × 1010 K (calculated at some distance from the BH center). We measure Te at r = 6M, and we obtain a narrower range (likely owing to fitting of polarized observations) of Te = (3.0–4.2) × 1010 K.

One can use two types of observations to estimate the BH spin axis P.A.: the 230 GHz correlated flux and the EVPA. Using the correlated flux gave Broderick et al. (2009) and Dexter et al. (2010) a result of P.A. = (− 70°)–(− 20°) = (110°)–(160°). Using the EVPA data has given slightly different results: Meyer et al. (2007) predict the range P.A. = 60°–108°, whereas Huang gets either P.A. ≈ 115° (Huang et al. 2009b) or P.A. ≈ 140° (Huang et al. 2009a). Our values of P.A. are within the range 85°–171°, which is consistent with predictions in Meyer et al. (2007) and with estimates based on the observed correlated flux. The size of the flow may depend substantially on the luminosity state (Broderick et al. 2009) or the presence of non-thermal structures, spiral waves, and other features. In some astrophysical sources, P.A. is directly known from spatially resolved jets, and Sgr A* may be one of such sources. A tentative jet feature was revealed in X-rays by Muno et al. (2008) in their Figure 8 showing P.A.jet = 120°. This value is close to P.A. = 115fdg3 or P.A. = 120fdg3 for the best-fit models with spins a* = 0.5 and a* = 0.98, respectively.

Besides the estimates of accretion rate and flow properties based on the inner flow, there exist estimates based on the outer flow. Shcherbakov & Baganoff (2010) constructed an inflow–outflow model with conduction and stellar winds with radiation matching the X-ray surface brightness profile observed by Chandra. Their model7 had an accretion rate $\dot{M}=6\times 10^{-8}\ M_\odot \ {\rm {{\rm yr}}}^{-1}$ and electron temperature Te = 3.6 × 1010 K at r = 6M, which is consistent with present results. Shcherbakov & Baganoff (2010) constrained the density in the outer radial flow from X-ray observations, while the present work constrains the density in the inner radial flow from submillimeter observations. The density profile is then found to be

Equation (31)

The density power-law index β lies between β = 1.5 for ADAF (Narayan & Yi 1995) and β = 0.5 for the convection-dominated accretion flow (Narayan et al. 2000; Quataert & Gruzinov 2000). However, the modification of the power-law index from the steep ADAF profile is likely due to conduction for Sgr A*, not convection (Shcherbakov & Baganoff 2010). Newer GRMHD simulations of radially extended disks show a comparable power-law index for density (McKinney et al. 2012).

Our dynamical model has limitations and relies on several approximations. More convergence testing, like done in McKinney et al. (2012), is required to ensure that the 3D GRMHD simulation results are reliable. The amount of initial magnetic flux and the field geometry might have a pronounced effect on simulation results. For example, magnetically choked accretion flows (MCAFs; Igumenshchev 2008; McKinney et al. 2012) may have more desirable properties (such as larger Faraday rotation as discussed related to Figure 9) for Sgr A* compared to MRI-dominated disks described in the present work. The dependence of the estimated accretion flow and the BH parameters on the simulation type and the initial setup should be carefully explored in future works. The polarization is expected to be able to best highlight changes in the magnetic field geometry and strength, and so our work is an important stepping stone to distinguish whether Sgr A* is a classical MRI-dominated disk or an MCAF.

The limited dynamical range of our simulations leads to another caveat. We fix electron density ne and temperatures Tp and Te in the outer flow and extend them down to the event horizon. The slopes of these quantities break at 25M radius, where the power-law radial extrapolation starts. Thus, the density and temperature slopes in the inner flow may need to be determined more self-consistently. Future simulations will need to cover a larger range of radii and plasma physics effects, such as conduction (Johnson & Quataert 2007; Sharma et al. 2008; Shcherbakov & Baganoff 2010). Simulations with larger outer radial boundaries that are run for longer will also help to fit the Faraday rotation, which happens for the present models partially outside of the simulated domain. A proper simulation of the polar region of the flow may be important as well. At present, we artificially limit the lowest density and highest temperature there. If we do not, then numerical artifacts associated with excessive numerical dissipation and heating appear (similar to those in Moscibrodzka et al. 2009).

We found the lowest χ2/dof for the model with spin a* = 0.5, whereas other groups found a* = 0 and a* = 0.9 to provide the best fits in their modeling. So, there still appears to be no reliable estimate of BH spin for Sgr A*. One common shortcoming of recent papers is the use of thermal electron distribution. If non-thermal electrons provide most of the energy for the submillimeter peak, then this would invalidate all prior spin estimates (Shcherbakov & Huang 2011).

The radiative transfer we performed has its shortcomings. The emissivities in our special synchrotron approximation provide, e.g., 2% agreement with exact emissivities (Leung et al. 2011; Shcherbakov & Huang 2011) for b = 20 G, θB = 1 rad, Te = 6.9 × 109 K, and observed frequency ν = 100 GHz. Agreement is better for larger Te. Non-polarized radiative transfer methods (Moscibrodzka et al. 2009; Dexter et al. 2010) have an intrinsic error that is comparable with our polarized radiative transfer method for the same total emissivity εI, but the error is still 1%–5%.

There are other unaccounted sources of error. The mass of the BH in the Galactic center is known to within 10% (Ghez et al. 2008; Gillessen et al. 2009), and the distance is known to 5%. We do not expect these uncertainties to lead to significant changes in our predictions. A shift to slightly lower spin may be able to mimic the effect of smaller BH or a BH at larger distance.

An improvement in observations can lead to further insights on the flow and BH parameters. For example, the detailed comparison of flux, LP, and CP curves in Figure 7 shows that the models with different spins have discrepancies at frequencies not yet probed by observations. In particular, the CP fractions at 88 and 690 GHz are different. The EVPA data need improvement as well. EVPA observations are available at 230 and 349 GHz, but these frequencies are affected by Faraday rotation. The observations at higher frequency, where the Faraday rotation effect is weaker, should provide a better estimate of BH spin axis P.A. Another important quantity, LP at 88 GHz, has a largely unknown value. Its observations are reported in two papers. Variations in simulated LP (88 GHz) are large between the best-fit models (see Figure 7). Refinement of the observed mean LP (88 GHz) could potentially help discriminate between different spins. A measurement of the emitting region size or the correlated flux is also promising. Despite the correlated flux at 230 GHz being measured at the SMT-JCMT 3.5Gλ baseline, the statistics of this measurement need to be improved toward being comparable with the statistics of total flux. The correlated flux observations are currently being accumulated (Fish et al. 2011). The correlated flux at this baseline is exponentially sensitive to the physical flow size. As a caveat, the conclusion on image sizes may depend on the behavior of matter in the ill-defined polar regions. Our models do not exhibit significant emission from high latitudes at 230 GHz (see Figure 11) or anywhere above 88 GHz.

Future work should incorporate rigorous statistical analysis, and such analysis should include temporal information from the observations. The time variability properties can be found from the simulations and compared to the observed ones. In particular, "jet lags" (Yusef-Zadeh et al. 2008; Maitra et al. 2009) and quasi-periodic oscillations (Genzel et al. 2003; Eckart et al. 2006; Miyoshi et al. 2011) should be investigated using the simulations (Dolence et al. 2012). Also, future 3D GRMHD simulations will model more radially extended flows, account for ADAF/ADIOS-type scale heights of |h/r| ∼ 1, capture outflows, and account for the effects of accumulated magnetic flux near the BH (McKinney et al. 2012). Lastly, for the radiative transfer, adding Comptonization would be one way to test the quiescent X-ray luminosity L ≈ 4 × 1032 erg s−1 within 2–10 keV (Shcherbakov & Baganoff 2010).

The authors are grateful to Lei Huang for checking various emissivity prescriptions; to Ramesh Narayan for extensive discussions and comments; to Avi Loeb, Avery Broderick, James Moran, Alexander Tchekhovskoy, Cole Miller, Julian Krolik, Steven Cranmer for insightful comments; and to Jim Stone for encouragement with self-consistent radiative transfer. We thank the anonymous referees for their extensive feedback, which helped to improve the manuscript. The numerical simulations and the radiative transfer calculations in this paper were partially run on the Odyssey cluster supported by the FAS Sciences Division Research Computing Group at Harvard, Deepthought cluster at the University of Maryland, and were partially supported by NSF through TeraGrid resources provided by NCSA (Abe), LONI (QueenBee), and NICS (Kraken) under grant numbers TG-AST080025N and TG-AST080026N. The paper is partially supported by NASA grants NNX08AX04H (R.V.S. and Ramesh Narayan), NNX08AH32G and NNX11AE16G (Ramesh Narayan), NASA Hubble Fellowship grant HST-HF-51298.01 (R.V.S.), NSF Graduate Research Fellowship (R.F.P.), and NASA Chandra Fellowship PF7-80048 (J.C.M.).

APPENDIX: RADIATIVE TRANSFER CONVERGENCE

We have devised a novel code for GR polarized radiative transfer. As with any new code, we need to conduct a set of convergence tests to ensure that it works accurately. First, we need to come up with metrics for assessing accuracy. In the present paper we model fluxes at seven frequencies between 88 and 857 GHz, LP fractions at three frequencies, and CP fractions at two frequencies and define χ2 as to characterize the goodness of fit. We employ a similar quantity χ2H/dof to characterize the accuracy of radiative transfer. We define

Equation (A1)

where Qi, 1 are simulated polarized fluxes for one set of radiative transfer parameters and Qi, 2 are the fluxes for another set. The errors σ(Q) are the observed errors of the mean, and the index i runs through all fitted fluxes and LP and CP fractions. When one of the models fits the data exactly, χ2H/dof coincides with χ2/dof. We vary the following radiative transfer and dynamical model parameters:

  • 1.  
    Number of points PN along north–south axis and along east–west axis in the picture plane.
  • 2.  
    Distance from the center Pss measured in horizon radii rH, where radiative transfer starts.
  • 3.  
    Dimensionless scale Pfact of the integration region in the picture plane.
  • 4.  
    Number of simulated spectra Nperiods for a single model to compute the mean spectrum.
  • 5.  
    Time interval Δt of simultaneous propagation of rays and evolution of numerical simulations.
  • 6.  
    Extension power-law slope of density profile Prhopo.
  • 7.  
    Extension slope of temperature profile PUpo.
  • 8.  
    Extension slope of magnetic field profile PBpo.

Since fluctuations and differences in χ2/dof between different models reach 1, values χ2H/dof ≲ 0.1 are acceptable, but, in general, we strive for χ2H/dof < 0.02. We set constant Pfact, Pss, Psnxy for all radiative transfer computations, but we cannot check the code accuracy for all models. We check the convergence a posteriori for the best-fit model at each spin value.

We find values of parameters by trial and error. The resulting set has Pfact = 1, Pss = 1.01rH, Psnxy = 111, Nperiods = 21, Δt = 120M. The values of Prhopo and PUpo are fixed by extensions to large radii of temperature and density in the inner flow.

The tests and the values of χ2H/dof are summarized in Table 3. The second column describes the test. In particular, Pfact: 1 → 0.8 means that we test convergence of the integration region relative size. We change one parameter at a time. Since the power-law slopes Prhopo and PUpo can vary from model to model, we change them in such a way that Prhopo is increased by 0.2 and PUpo is decreased by 0.1. We also estimate the influence of magnetic field extension power-law slope PBpo by making it shallower from (r/M)−1.0 to (r/M)−0.8. We chose to test relatively small variations ΔPrhopo = 0.2 and ΔPUpo = 0.1, because density and temperature at rout = 3 × 105M are known to within a factor of several (Baganoff et al. 2003; Shcherbakov & Baganoff 2010), while these variations correspond to changes by factors of 7 and 2.5 in density and temperature, respectively, at rout.

Table 3. Values of χ2H/dof for Radiative Transfer Convergence Tests and Sensitivity to Model Parameter Tests for Best-fit Models

Number Description Spin a* = 0 Spin a* = 0.5 Spin a* = 0.7 Spin a* = 0.9 Spin a* = 0.98
1 PN: 75 → 111 0.00081 0.00138 0.00101 0.00047 0.01175
2 PN: 111 → 161 0.00017 0.00072 0.00018 0.00007 0.00084
3 Pss: 1.003rH → 1.01rH 0.00036 0.00059 0.00110 0.00095 0.00051
4 Pss: 1.01rH → 1.03rH 0.00778 0.00982 0.01616 0.01468 0.00920
5 Pfact: 0.8 → 1.0 0.01358 0.07278 0.06905 0.02893 0.02686
6 Pfact: 1.0 → 1.2 0.00087 0.05532 0.07173 0.02681 0.03534
7 Nperiods : 11 → 21 0.15665 0.40611 0.12233 0.21397 0.13588
8 Nperiods : 21 → 41 0.02474 0.04505 0.13244 0.02684 0.04834
9 Interval Δt: 120M → 180M 0.06987 0.06851 0.13549 0.02948 0.10979
10 Interval Δt: 80M → 120M 0.09095 0.04103 0.03094 0.03881 0.06246
11 Interval Δt: 0M → 120M 0.09051 0.35296 0.53045 0.02731 0.07881
12 Prhopo: QQ + =0.2 0.04493 0.04057 0.03134 0.01587 0.05241
13 PUpo: QQ − =0.1 0.01200 0.02726 0.00977 0.01088 0.04174
14 PBpo: −1.0 → −0.8 0.02401 1.05214 0.15156 0.05486 0.04941

Download table as:  ASCIITypeset image

The results of the tests are as follows. The first 11 tests represent variations of radiative transfer parameters, and the last 3 tests explore the variations of power-law extension slopes. Tests 1–4 produce small χ2H/dof, so that PN can be lowered and Pss can be increased. The changes in the integration region scale (Pfact) result in high χ2H/dof ≈ 0.07, as indicated by tests 5 and 6. Low Pfact leads to systematic underproduction of total flux, whereas high Pfact mainly leads to different LP fractions. Test 7 results in high χ2H/dof ≈ 0.4, so that a small number of simulated spectra (e.g., Nperiods = 11) cannot be justified. Lower values χ2H/dof ≈ 0.13 attained in test 8 indicate that Nperiods = 21 periods might be acceptable. With tests 9–11, we tested variations in the time interval Δt of simultaneous propagation of rays and evolution of numerical simulations. It is expected that longer intervals lead to convergence. However, switching from Δt = 120M to Δt = 180M and switching Δt = 80M to Δt = 120M both lead to χ2H/dof ≲ 0.1. Since these values of χ2H/dof are acceptable, we implement Δt = 120M for radiative transfer runs. As elucidated by test 11, freezing simulations in time lead to χ2H/dof ≈ 0.5, which is too high. Thus, conducting radiative transfer over frozen simulation snapshots is not acceptable. Changes in extension slopes of density and temperature (tests 12 and 13) result in small χ2H/dof ≲ 0.05. Variations of magnetic field slope (test 14) lead to large χ2H/dof ≈ 1, which means that the modifications of b extensions will change the best fits. Extensions as shallow as |b|∝(r/M)−0.5 may provide better fits to Faraday RM and should be carefully explored. Various extensions of the fluid velocity lead to practically the same polarized intensities and are not included in tests.

Footnotes

  • Note that gravitational radius is defined as rg = 2M in Shcherbakov & Baganoff (2010).

Please wait… references are loading.
10.1088/0004-637X/755/2/133