ACCRETION OF SUPERSONIC WINDS ONTO BLACK HOLES IN 3D: STABILITY OF THE SHOCK CONE

and

Published 2015 October 6 © 2015. The American Astronomical Society. All rights reserved.
, , Citation M. Gracia-Linares and F. S. Guzmán 2015 ApJ 812 23 DOI 10.1088/0004-637X/812/1/23

0004-637X/812/1/23

ABSTRACT

Using numerical simulations we present the accretion of supersonic winds onto a rotating black hole in three dimensions. We study five representative directions of the wind with respect to the axis of rotation of the black hole and focus on the evolution and stability of the high-density shock cone that is formed during the process. We explore both the regime in which the shock cone is expected to be stable in order to confirm previous results obtained with two-dimensional simulations, and the regime in which the shock cone is expected to show a flip–flop (FF) type of instability. The methods used to attempt a triggering of the instability were (i) the accumulation of numerical errors and (ii) the explicit application of a perturbation on the velocity field after the shock cone was formed. The result is negative, that is, we did not find the FF instability within the parameter space we explored, including cases that are expected to be unstable.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The Bondi–Hoyle–Littleton or wind accretion process occurs when a compact object moves with respect to a uniformly distributed perfect fluid considered to be free of self-gravity (Bondi & Hoyle 1944; Bondi 1952). This problem has been studied analytically and numerically in Newtonian (Fryxell et al. 1987; Matsuda et al. 1987, 1991, 1992; Shima et al. 1988; Sawada et al. 1989) and relativistic regimes (Petrich et al. 1988; Font & Ibañez 1998; Dönmez et al. 2011; Cruz-Osorio et al. 2012; Lora-Clavijo & Guzmán 2013). In the relativistic regime this process has been associated with high-energy sources (for a review see Edgar 2004). The process has been analyzed for various types of compact objects acting as accretors and in this paper we concentrate on the case of a black hole. In its most advanced fashion, the BHL accretion onto rotating black holes includes the radiation of the fluid restricted to 2D (Zannoti et al. 2011; Park & Ricotti 2013), the accretion in 2.5D magnetized fluids in (Penner 2011; Takahashi & Ohsuga 2014), and the case when the fluid has density gradients (Loxra-Clavijo et al. 2015).

One of the important aspects of this accretion process is that when the relative speed between the accretor and the fluid is supersonic, a high-density region with a conic shape called a shock cone is formed on the rear part of the accretor. When the compact object is rotating, the fluid of the shock cone gets dragged by rotation, eventually becoming distorted, and its motion may be unstable (see, e.g., Foglizzo et al. 2005). The stability is important because it would provide a model for an abrupt acceleration of gas in high-density regions near a compact object, which in turn could help with the modeling of short- and high-energy emissions (Dönmez et al. 2011; Lora-Clavijo & Guzmán 2013).

The stability of the shock cone is described in terms of whether or not the flow is steady near the accretor. Particularly interesting is the instability of the flip–flop (FF) type, which happens when the angular momentum of the shock cone starts changing sign and eventually the cone blows up. On one hand this instability involves the nature of the accretor. For instance, a black hole with an event horizon and a neutron star may show different behavior due to the different nature of their surfaces. On the other hand the stability depends on the dynamical conditions of the wind, such as its velocity and the speed of sound associated with its equation of state.

Empirical criteria have been proposed to establish a regime of stability for the shock cone that is independent of the accretor and more focused on the relation of the accretor size and the Bondi accretion radius (e.g., Foglizzo et al. 2005). Given that this analysis requires the evolution of the wind by numerical means, for example, on top of the curved background spacetime of a black hole, there is still controversy about whether the instability is physical or due to the numerical methods used. For instance, in Shima et al. (1988); Sawada et al. (1989; Matsuda et al. (1991, 1992); and Foglizzo et al. (2005), some dynamical arguments are found that show that the instability is physical, whereas in Fryxell et al. 1987; Shima et al. 1998; and Cruz-Osorio et al. 2012 it is shown that it might be due to an inadequate numerical implementation.

In this paper we focus on the FF instability of the shock cone for the particular case in which the accretor is a spinning black hole and the wind is a relativistic fluid obeying an ideal gas equation of state in 3D. This problem has been analyzed in 2D in Font & Ibañez (1998); Dönmez et al. (2011); and Cruz-Osorio et al. (2012), where a relativistic study of nonspherical supersonic Bondi–Hoyle accretion onto a spinning black hole with relative accretor size ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.1,0.26$ was made on the equatorial plane assuming s-lab symmetry, where rhor is an approximate black hole radius and racc is the Bondi accretion radius. We present these cases in full 3D because as far as we can tell such results have never been presented before.

According to Foglizzo et al. (2005), the instability is expected to happen when the properties of the system involve a supersonic wind velocity with Mach numbers ${{\mathcal{M}}}_{\infty }^{R}\geqslant 2.4$ and the size of the accretor with respect to the Bondi accretion radius is such that ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}\gtrsim 0.05.$ Therefore, in order to confirm or discard the stability in this regime, we explore the parameter space with this high speed and small black holes that lie within this regime, where ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}$ ranges from 0.011 to 0.025. We explore the accretion onto black holes with angular momentum parameters as high as a = 0.95 and five representative orientations of the wind flow with respect to the axis of rotation of the black hole. In all of the cases analyzed we find that there is no FF instability.

The paper is organized as follows. In Section 2 we describe the relativistic hydrodynamics (RHD) equations modeling the fluid and the numerical methods used. In Section 3 we present the set of configurations that we experiment with, and finally in Section 4 we describe the conclusions from our analysis.

2. NUMERICAL SET UP

2.1. The Equations and Numerical Methods

The equations of general relativistic hydrodynamics are written as a flux conservative system, which assumes a standard 3 + 1 decomposition of the spacetime, known as the Valencia formulation (Banyuls et al. 1997):

Equation (1)

where ${\boldsymbol{u}}=\{D,{S}_{i},\tau \}$ is the set of conserved variables, D is the generalized rest-mass density of the fluid, Si are the generalized momenta in each direction, τ is the internal energy, ${{\boldsymbol{F}}}^{(i)}({\boldsymbol{u}})$ is the fluxes, and ${\boldsymbol{S}}({\boldsymbol{u}})$ is the source. In terms of the primitive variables of the fluid elements, ρ the rest-mass density, p the pressure, vi the fluid 3-velocity, and epsilon the specific internal energy, the conserved variables are defined by

Equation (2)

where γ is the determinant of the spatial 3-metric ${\gamma }_{{ij}}$ of the 3D spatial slices that the spacetime is foliated with, $h\equiv 1+\epsilon +p/\rho $ is the specific enthalpy and $W={(1-{\gamma }_{{ij}}{v}^{i}{v}^{j})}^{-1/2}$ is the Lorentz factor. As usual, the system of equations is closed with an equation of state. In terms of primitive and conservative variables, the fluxes and sources are:

where ${g}_{\mu \nu }$ are the components of the 4-metric and ${{\rm{\Gamma }}}^{\delta }{}_{\mu \nu }$ are the Christoffel symbols of the spacetime, α is the lapse function, and ${\beta }^{i}$ are the components of the shift vector associated to the 3 + 1 decomposition of the spacetime.

We solve the system of Equations (1) using the Cactus Einstein Toolkit (ETK; Einstein Toolkit 2005; Löffler et al. 2012), which provides the necessary computational tools to evolve this relativistic fluid in full 3D. In particular we use the GRHydro Thorn (Baiotti et al. 2005), which contains high-resolution shock capturing methods to solve these equations. For these methods we specifically use the HLLE and Marquina numerical flux formulae and minmod and MC linear reconstructors. For the integration in time we use a fourth order Runge–Kutta method. As part of the diagnostics tools, we use the Outflow Thorn that allows us to measure the flux of rest-mass across a given spherical surface and monitor the behavior of the accretion rate.

2.2. Initial Conditions

We describe the spacetime of the rotating black hole using Kerr–Schild (KS) horizon-penetrating coordinates, with the axis of rotation parallel to $\hat{z}$ as follows:

Equation (3)

where M is the mass of the black hole and we have assumed geometric units $G=c=1$ to hold.

The fluid is set initially as a spatially constant rest-mass density ideal gas, moving toward the black hole and with a given asymptotic velocity ${v}_{\infty }^{2}={v}^{i}{v}_{i}.$ We assume the fluid obeys a gamma-law equation of state $p=({\rm{\Gamma }}-1)\rho \epsilon $ and thus we introduce the asymptotic speed of sound ${c}_{s\infty }.$ Once we define the value of cs and assume the density to initially be a constant $\rho ={\rho }_{\mathrm{ini}},$ the pressure can be written as ${p}_{\mathrm{ini}}={c}_{{\rm{s}}\infty }^{2}{\rho }_{\mathrm{ini}}/({\rm{\Gamma }}-{c}_{{\rm{s}}\infty }^{2}{{\rm{\Gamma }}}_{1}),$ where ${{\rm{\Gamma }}}_{1}={\rm{\Gamma }}/({\rm{\Gamma }}-1).$ In order to avoid negative and zero values of the pressure, the condition ${c}_{{\rm{s}}\infty }\lt \sqrt{{\rm{\Gamma }}-1}$ has to be satisfied. Finally, with this value for ${p}_{\mathrm{ini}},$ the initial internal specific energy is reconstructed using the equation of state.

In order to parameterize the initial data, we define the relativistic Mach number at infinity ${{\mathcal{M}}}_{\infty }^{R}={{Wv}}_{\infty }/{W}_{s}{c}_{{\rm{s}}\infty },$ where W is the Lorentz factor of the gas and Ws is the Lorentz factor calculated with the speed of sound. When ${{\mathcal{M}}}_{\infty }^{R}$ is bigger than one, the wind is said to be supersonic and otherwise subsonic.

According to Foglizzo et al. (2005), the two parameters deciding on the stability of the process are the Mach number of the gas ${{\mathcal{M}}}_{\infty }^{R}$ and the relative size of the accretor with respect to the Bondi accretion radius racc. In our case the compact object is a black hole and we define an approximate black hole radius to be ${r}_{\mathrm{hor}}=2\;M,$ even though our hole is rotating. Thus the relative size of the accretor is given by ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}={v}_{\infty }^{2}+{c}_{{\rm{s}}\infty }^{2}.$

This formula indicates that a relatively small black hole is equivalent to having a case with the wind moving with a small velocity v. This explains why the cases where ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}$ is small are difficult to track, and even more difficult in 3D. The reason is that the numerical domain is set such that it contains at least a sphere of radius racc in order to see the accretion process. Slower winds imply a bigger racc, therefore demanding a bigger numerical domain and consequently more computational resources.

In order to observe the behavior of the process in a general scenario, we choose five representative different directions of the wind, which we characterize in terms of its orientation with respect to $\hat{z}$, which is the axis of rotation of the black hole denoted by $\Uparrow .$ These five different orientations of the wind are represented by $\uparrow \;\downarrow \;\to \;\swarrow \;\nwarrow ,$ and correspond to directions parallel to $\hat{z},\;-\hat{z},\;\hat{x},\;-\hat{x}-\hat{z}$, and $-\hat{x}+\hat{z}$ respectively.

2.3. Boundary Conditions

The ETK allows the solution of the equations in Cartesian coordinates, and thus the domain is a cubic box of a given size. In the case of the evolution of a wind, two boundary conditions are required (see, e.g., Dönmez et al. 2011; Cruz-Osorio et al. 2012), namely an inflow boundary condition in the upstream region of the domain, where the gas is being pumped into the box, and outflow boundary conditions in the piece of the boundary where the wind is expected to exit the numerical domain.

The ETK does not provide the tools to automatically incorporate different boundary conditions on different faces of the cubic domain and we programmed a module that allows one to do it.

There is another very important boundary to consider. Since we are using horizon-penetrating coordinates we are allowed to perform an excision inside the black hole horizon as performed in Cruz-Osorio et al. (2012) and Lora-Clavijo & Guzmán (2013). The excision method consists of the removal of a chunk of the numerical domain inside the black hole horizon. The removed chunk is usually assumed to be a lego sphere that defines a lego-shaped boundary inside the horizon. Considering the fact that the light cones in KS coordinates remain open and point toward the singularity inside the black hole horizon, no boundary conditions are needed, and extrapolation of the fluid variables suffices to guarantee that the fluid is not reflected back toward the exterior of the black hole. This method is already implemented in the ETK, as described in Hawke et al. (2005).

3. NUMERICAL EXPERIMENTS AND RESULTS

Showing the instability of a given fluid configuration requires the perturbation theory analysis, which applies when equilibrium configurations are given a priori. Another possibility uses the numerical evolution of equilibrium configurations and then numerical errors; for instance, truncation or discretization errors act as the perturbation and eventually (or never) trigger an instability. When the configuration is stable, the perturbation is simply expelled and the system remains in a state pretty much like the unperturbed one, whereas in unstable cases the variables of the system change notably.

In the present case we deal with the stability of a shock cone, which is a configuration that actually forms as a consequence of the evolution of the fluid, and there is no equilibrium configuration given a priori. Thus one needs to first evolve the system until the shock cone forms. The shape and properties of the cone are not constructed based on a set of assumptions about the density profile or velocity field, but are merely given by numerical evolution. Therefore, in order to know about its stability, the method consists of tracking the further evolution of the system; we simply need to evolve the system during a sufficiently long time and wait for the numerical errors to accumulate and then possibly make the cone oscillate in a FF fashion or show any other type of instability.

The different configurations of a wind accretion process are characterized by the following parameters: (1) the asymptotic velocity of the wind, (2) the initial rest-mass density of the wind, (3) the adiabatic index Γ, (4) the black hole angular momentum a, (5) the orientation of the wind with respect to the axis of rotation of the black hole, and (6) the relative accretor size ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}.$

According to Foglizzo et al. (2005) the size of the accretor and the velocity of the wind are the parameters that determine whether the shock cone is stable or unstable, and it has been estimated that in the range of relative accretor size ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}\gtrsim 0.05$ with Mach number ${{\mathcal{M}}}_{\infty }^{R}\geqslant 2.4$ the shock cone is unstable.

In this paper we explore two sizes of the black hole. We first consider a mildly small black hole with ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.1,0.26.$ With these sizes the shock cone should be stable, however, since they have not been shown before in 3D we present them here. On the other hand, a second set of experiments considers a very small accretor ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.011,0.012,0.025,$ which is in the regime of instability.

3.1. Mildly Small Black Hole

In the first set of experiments we use the following parameters: ${\rho }_{0}={10}^{-6},$ ${\rm{\Gamma }}=4/3,$ cs = 0.1. The velocities are supersonic with ${{\mathcal{M}}}_{\infty }=3,\;5$ and we use two values of the wind velocity ${v}_{\infty }=0.3,\;0.5,$ which correspond to the sizes of the accretor ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.1$ and 0.26, respectively. These parameters correspond to mildly small black holes with sizes bigger than those expected for objects showing FF instability according to Foglizzo et al. (2005). We also carried out these simulations for two values of the black hole angular momentum $a=0.5,\;0.95.$

The evolution was performed using resolution ${\rm{\Delta }}{xyz}=0.25\;M,$ on a domain ${[-40\;M,40\;M]}^{3},$ where M is the mass of the black hole. This is a modest resolution compared to that used in the analysis using s-lab symmetry (Font & Ibañez 1998; Dönmez et al. 2011; Cruz-Osorio et al. 2012); however, this should trigger an instability faster, if it triggers any at all. In all of the cases, that is, for the two velocities, the two values of the black hole angular momentum and the five orientations of the wind, we did not find the FF instability, as expected for these parameters. In Table 1 we list the set of configurations for which we track the evolution of the wind.

Table 1.  Configurations Analyzed for ${\rho }_{0}={10}^{-6},$ ${\mathcal{M}}=3,\;5,$ ${\rm{\Gamma }}=4/3,$ cs = 0.1, and ${v}_{\infty }=0.3$

Config. ${\mathcal{M}}=3$     ${\mathcal{M}}=5$  
  ${\dot{M}}_{3}({10}^{-4})$ ${\dot{M}}_{3\mathrm{BHL}}({10}^{-4})$   ${\dot{M}}_{5}({10}^{-4})$ ${\dot{M}}_{5\mathrm{BHL}}({10}^{-4})$
a = 0.5
$\uparrow \Uparrow $ 8.29   3.81
$\downarrow \Uparrow $ 8.24   3.80
$\to \Uparrow $ 8.21 4.00   3.63 1.00
$\swarrow \Uparrow $ 8.27   3.60
$\nwarrow \Uparrow $ 8.32   3.61
a = 0.95
$\uparrow \Uparrow $ 8.23   3.64
$\downarrow \Uparrow $ 8.22   3.64
$\to \Uparrow $ 8.82 4.00   3.63 1.00
$\swarrow \Uparrow $ 8.57   3.64
$\nwarrow \Uparrow $ 8.50   3.58

Note. The value of $\dot{M}$ corresponds to the value measured on a spherical surface located at $r=3\;M$ after the process of shock cone formation has been settled down. No instability was found in all of the presented cases. In order to compare how well the BHL formula works, we also show the accretion rate for each case.

Download table as:  ASCIITypeset image

In these simulations there are already some previously unknown results, for example, the case $\to \Uparrow ,$ which is the generalization of the s-lab studies shown in, e.g., Font & Ibañez (1998); Dönmez et al. (2011); and Cruz-Osorio et al. (2012). In Dönmez et al. (2011) it was shown that there was a FF instability, whereas in Cruz-Osorio et al. (2012) the use of horizon-penetrating coordinates to describe the black hole indicated that there is no such instability. However, both analyses were performed using s-lab symmetry and the result in 3D was still uncertain. We show, for the first time, in Figure 1, a snapshot of the cone in 3D for the case with ${v}_{\infty }=0.5,$ ${c}_{s\infty }=0.1$ and black hole angular momentum a = 0.95. No signs of instability were found up to t = 10,000 M, which is a comparable time with that in Font & Ibañez (1998; Dönmez et al. (2011); and Cruz-Osorio et al. (2012).

Figure 1.

Figure 1. Isocurves of the density at $t=3000\;M$ for the mildly accretor size case $\to \Uparrow ,$ a = 0.95, and ${v}_{\infty }=0.5,$ ${c}_{s\infty }=0.1,$ ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.26.$ The thick line defining the conic shape is actually a set of nearby isocurves that show how steep the density gradient is in that region. At the top we show the cone morphology as seen at the plane z = 0, where the dragging of the density due to the black hole rotation can be seen. In the bottom panel we show the shock cone structure on the plane y = 0, which is the angle that is being ignored in simulations assuming s-lab symmetry.

Standard image High-resolution image

Aside from the morphology of the shock cone that looks very stable, we monitor the stability of the accretion process by measuring the accretion rate $\dot{M}$ at a spherical surface located at $r=3\;M.$ In all of our experiments we find this rate to stabilize after an initial transient corresponding to the shock cone formation lapse. We measure this rate for the different cases and check that the slower the wind, the bigger the accretion rate, as expected from the Bondi formula. We also show the accretion rate in Table 1 for the two cases ${{\mathcal{M}}}_{\infty }=3$ and ${{\mathcal{M}}}_{\infty }=5$, and for comparison we additionally show the values obtained using the BHL accretion rate formula (Edgar 2004)

Equation (4)

we label the results with ${\dot{M}}_{3\mathrm{NHL}}$ and ${\dot{M}}_{5\mathrm{NHL}}$ for the two wind velocities considered. We point out that the accretion rate, as measured for a black hole spacetime in our simulations for this size of black hole, is considerably bigger than the accretion rate calculated with the BHL formula.

3.1.1. Particular Cases

Some special configurations may be interesting. For instance one may wonder whether there is a whirlpool behind the black hole when the wind moves parallel or antiparallel to the axis of rotation. In Figure 2 we show slices of the velocity field at three planes of z = constant, and a lateral view of the process at $t=8000\;M.$ In this case the black hole size is ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.26$ and no dragging effects are noticeable in the velocity field due to the rotation of the black hole. What can actually be seen is the convergence of the stream lines at the shock cone boundary.

Figure 2.

Figure 2. Velocity field at t = 8000 M for the case $\uparrow \Uparrow ,$ ${v}_{\infty }=0.5,$ spin a = 0.95, and ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.26.$ The slices are as follows: (top left) is taken at z = 3, (top right) at z = 6, (bottom-left) at z = 9,  and (bottom-right) at y = 0.

Standard image High-resolution image

Another case that has never been shown in 3D is the accretion of a diagonal wind. We show a snapshot of the shock cone density at $t=5000\;M$ in Figure 3, for the wind orientation $\swarrow \Uparrow .$

Figure 3.

Figure 3. Snapshot of the density isocurves for the diagonal case $\swarrow \Uparrow ,$ a = 0.5, and ${v}_{\infty }=0.5,$ ${c}_{s\infty }=0.1,$ ${\rm{\Gamma }}=4/3$, and ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.26.$

Standard image High-resolution image

3.2. Very Small Black Hole

A second set of tests consists of the evolution of the wind for a smaller black hole, particularly three cases: ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.011,0.012,0.025.$ Because these numbers are smaller than the threshold of 0.05, the FF instability is expected to happen as suggested in Foglizzo et al. (2005). According to the analysis in Foglizzo et al. (2005), even axisymmetric simulations should become unstable when considering ${\rm{\Gamma }}=5/3,4/3$ and ${{\mathcal{M}}}_{\infty }=2.4,\;3.0.$ We evolved these particular cases in 3D for all the wind orientations described before. These cases are presented in Table 2. In this case we only perform the simulations for the bigger black hole angular momentum a = 0.95. We use the same resolution ${\rm{\Delta }}{xyz}=0.25\;M,$ however, because a domain containing at least a sphere of radius racc is required, we used the domain ${[-100\;M,100\;M]}^{3}.$ Once again we did not find the FF instability after 10,000 M. We point out that unlike the mildly small accretor size, the accretion rate for this size of accretor is considerably smaller than the accretion calculated with the BHL formula.

Table 2.  Configurations Analyzed for ${\rho }_{0}={10}^{-6}$ and a = 0.95

${\rm{\Gamma }}=5/3$ ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.011$ ${\mathcal{M}}=2.4$
  ${\dot{M}}_{2.4}(\times {10}^{-2})$ ${\dot{M}}_{\mathrm{BHL}}({10}^{-2})$
$\uparrow \Uparrow $
$\downarrow \Uparrow $
$\to \Uparrow $ 0.14 1.00
$\swarrow \Uparrow $
$\nwarrow \Uparrow $
${\rm{\Gamma }}=4/3$ ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.012$ ${\mathcal{M}}=2.5$
  ${\dot{M}}_{2.5}(\times {10}^{-2})$ ${\dot{M}}_{\mathrm{BHL}}({10}^{-2})$
$\uparrow \Uparrow $
$\downarrow \Uparrow $
$\to \Uparrow $ 0.48 1.00
$\swarrow \Uparrow $
$\nwarrow \Uparrow $
${\rm{\Gamma }}=4/3$ ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.025$ ${\mathcal{M}}=3$
  ${\dot{M}}_{3.0}(\times {10}^{-2})$ ${\dot{M}}_{\mathrm{BHL}}({10}^{-3})$
$\uparrow \Uparrow $
$\downarrow \Uparrow $
$\to \Uparrow $ 0.34 3.72
$\swarrow \Uparrow $
$\nwarrow \Uparrow $

Note. The value of $\dot{M}$ corresponds to the value measured on a spherical surface located at $r=3\;M$ after the process of shock cone formation has been settled down. None of the cases show instability. In order to compare how well the BHL formula works, we also show the accretion rate for each case. Unlike in the case of mildly small black holes, in this case the numerical accretion rate is about the same, independent of the orientation of the wind.

Download table as:  ASCIITypeset image

In order to illustrate how the dragging effects due to the rotation of a small black hole on the fluid are more significant compared to the bigger black hole case, in Figure 4 we show a snapshot of the density for the case with orientation $\to \Uparrow ,$ ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.011,$ ${\rm{\Gamma }}=5/3$, and ${\mathcal{M}}=2.4$ from Table 2. In Figure 5 we also show a snapshot on a vertical plane of the diagonal case, which shows a more dragged fluid than in the case of Figure 3.

Figure 4.

Figure 4. Isocurves of the density at time $t=8000\;M,$ for the case $\leftarrow \Uparrow ,$ a = 0.95, and ${v}_{\infty }=0.1,$ ${c}_{s\infty }=0.04,$ ${\rm{\Gamma }}=5/3,$ corresponding to a small black hole with ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.011.$ We show two views, at the z = 0 and y = 0 planes. This case is expected to be unstable because the shock is detached from the black hole, and the accretor size is small. Nevertheless, we did not encounter signs of FF instability until t = 10,000 M.

Standard image High-resolution image
Figure 5.

Figure 5. Density isocurves for the diagonal case $\swarrow \Uparrow ,$ $a=0.95,$ ${\rm{\Gamma }}=4/3$,  and ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.012.$ The snapshot is taken at $t=3800\;M.$

Standard image High-resolution image

3.3. Perturbed Bondi–Hoyle Accretion

As we did not find any unstable case, in addition to the analysis of simply tracking the evolution of the gas and expecting the numerical error to trigger the FF instability, we designed a numerical experiment in which we evolve the gas until the shock cone forms and then apply a perturbation to the velocity field of the form

Equation (5)

where $k\geqslant 0$ is a random number between 0 and A. We applied this random perturbation perpendicular to the direction of the shock cone and expected for it to trigger the instability. We performed this experiment using the two extreme relative accretor sizes in this paper, ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.011$ and ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.26,$ and all the orientations we have considered here. In all of the cases no instability was triggered for all the directions of the wind.

In Figure 6 we show the accretion rate for the perturbed case, where a feature can be seen at the moment at which we apply the perturbation. We used $A=0.2,$ which means that the velocity perpendicular to the shock cone has been added randomly, even one-fifth of the speed of light, and nevertheless the flux restores and becomes stable. In Figures 7 and 8 we show snapshots of the density before, during, and after the perturbation is applied, and show how the morphology of the cone restores.

Figure 6.

Figure 6. Accretion rate for the perturbed case $\to \Uparrow ,$ a = 0.95, and ${v}_{\infty }=0.5,$ ${c}_{s\infty }=0.1,$ ${\rm{\Gamma }}=4/3$ and a perturbation amplitude A = 0.2. In the inset we show the instant when the perturbation is applied by time $t=939\;M.$

Standard image High-resolution image
Figure 7.

Figure 7. Snapshots of the density for a perturbed case with orientation $\to \Uparrow ,$ a = 0.95, and ${v}_{\infty }=0.5,$ ${c}_{s\infty }=0.1,$ ${\rm{\Gamma }}=4/3$ with a perturbation amplitude A = 0.2. In the top left panel we show the density before applying the perturbation. In the top right panel we show the morphology at $t=939\;M$ when the perturbation was applied. In the bottom panels we show how the shock cone restores at times $t=945\;M$ and $t=955\;M.$

Standard image High-resolution image
Figure 8.

Figure 8. Snapshots of the density for a perturbed case with orientation $\to \Uparrow ,$ $a=0.95,$ ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.011,$ ${v}_{\infty }=0.1,$ ${c}_{s\infty }=0.04,$ ${\rm{\Gamma }}=4/3$ with a perturbation amplitude A = 0.2. In the top left panel we show the density before applying the perturbation. In the top right panel we show the morphology at $t=1500\;M$ when the perturbation was applied. In the bottom panels we show how the shock cone restores at times $t=1530\;M$ and $t=1600\;M.$

Standard image High-resolution image

3.4. Influence of Numerical Methods

We wanted to keep the physical system as free as possible of numerical sophistication in order to avoid undesirable effects. Thus, in the production runs we used a single refinement level because we wanted to avoid the numerical dissipation required to damp instabilities at refinement level interfaces, which could hide the potential unstable modes of the shock cone. This has been discussed as a potential damping of the instability in FF-type phenomena (Foglizzo et al. 2005).

In order to show the effects due to the use of fixed mesh refinement (FMR) in Figure 9 we show the accretion rate when using FMR with two refinement levels and when using a single refinement level. The deviation is of the order of 3% in the mass accretion rate. Thus, even though FMR allows  more efficient execution of long runs, the accuracy is not a good as with unigrid, which is important in long runs.

Figure 9.

Figure 9. (Top): Accretion rate comparison for the case $\to \Uparrow ,$ $a=0.95,$ ${\mathcal{M}}=5,$ between two runs, one using two refinement levels, and the other one using unigrid mode. In the unigrid mode, the whole domain ${[-20\;M,20\;M]}^{3}$ is covered using a resolution $0.25\;M$ in all directions. In the case of two refinement levels, the full domain is covered with a resolution $0.5\;M$ and a small box with a resolution $0.25\;M$ is set in the domain ${[-10\;M,10\;M]}^{3}.$ The accretion rate is measured at a sphere of radius $3\;M$ and the differences are of $3\%,$ which eventually might be important. (Bottom): Accretion rate comparison for the case $\to \Uparrow ,$ $a=0.95,$ ${\mathcal{M}}=5,$ between four runs, using Marquina and HLLE as Riemann solvers, both with MC and minmod as reconstruction methods, using the domain ${[-20\;M,20\;M]}^{3}$ in unigrid mode covered with a resolution $0.25\;M.$ The difference in the accretion rate among the four combinations of numerical methods lies within 1%.

Standard image High-resolution image

We also verified our results on the stability of the shock cone in all our runs using various combinations of approximate Riemann solvers and reconstruction methods. We evolved all the configurations in the tables using four combinations, namely: (1) Marquina with MC and minmod and (2) HLLE with MC and minmod. In all of the cases we found the shock cone to be stable in all of the described orientations and velocities of the wind. In Figure 9 we compare the accretion rate estimated using the four combinations of fluxes and reconstructors for a particular case. The difference among the accretion rate lies within 1%, which is smaller than the error introduced when using FMR.

4. DISCUSSION AND CONCLUSIONS

We presented the supersonic accretion of a wind onto black holes in 3D and studied whether or not there is an FF instability using a range of parameters that are expected to show this instability. We track the formation of the shock cone considering a high black hole spin and a general set of orientations of the wind with respect to the orientation of the axis of rotation of the black hole.

We focus on two regimes, one in which the relative accretor size is ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.1,\;0.26$, where the FF instability is not expected; however, we present the accretion in three dimensions. A second regime in which the relative accretor size is ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.011,\;0.012,\;0.025,$ which should show the FF instability, according to Foglizzo et al. (2005). In both cases we did not see the instability during the evolution time where it was expected to happen.

A first set of experiments assumes the numerical errors could trigger the instability of the shock cone. Since we did not find the instability we ran a second set of experiments in which we applied a transverse perturbation to the shock cone and also failed to see the instability.

Our production runs were performed in unigrid mode in order to avoid the numerical artifacts induced by mesh refinement, in particular the artificial dissipation. We illustrate, with a particular case, the estimated deviation from unigrid mode induced by the use of FMR.

Another particular ingredient of our simulations is that the description of the black hole spacetime uses KS horizon-penetrating coordinates, and we use excision inside the event horizon. This inhibits the propagation of numerical errors generated at the inner boundary. As described in Cruz-Osorio et al. (2012) this simple fact makes the difference between stability and instability of the shock cone when the accretor is a black hole. Unfortunately this technique can only be applied to black holes and not to other compact objects.

One more result is that the BHL formula is not accurate for both of our regimes. For the cases where ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.1,\;0.26,$ the accreted material on a surface located at $3\;M$ is bigger than twice the accretion predicted by the BHL formula, whereas for the cases of small black holes with ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}$ ranging from 0.011 to 0.025, the accretion rate ranges from nearly ∼0.15 to ∼0.5 times those calculated with the BHL formula, depending on the value of Γ and how supersonic the wind is.

Given the rich combination of wind orientations and values of the black hole angular momentum explored, we can safely conclude that even for our most extreme case, with a wind moving at ${{\mathcal{M}}}_{\infty }=2.4$ and black holes as small as ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.011,$ the shock cone is stable. And for faster winds with ${{\mathcal{M}}}_{\infty }=3$ and ${r}_{\mathrm{hor}}/{r}_{\mathrm{acc}}=0.025$, within the regime of expected instability there are no signs of instability either.

This research is partly supported by grant CIC-UMSNH-4.9.

APPENDIX

The implementation of the relativistic hydrodynamics, both in special and general relativity of the ETK has been tested severely (Einstein Toolkit 2005; Löffler et al. 2012). However, we have added a module that deals with specific boundary conditions that need to be tested. We present a test relevant to the results presented here, the exact solution of radial accretion of an ideal gas onto a Schwarzschild black hole known as a Michel solution (Papadopoulos & Font 1998). The first step toward constructing the exact solution is to integrate the evolution equations in spherical symmetry ${\partial }_{\mu }\left(\sqrt{-g}\rho {u}^{\mu }\right)=0$ and ${{\rm{\nabla }}}_{\mu }{T}_{\ 0}^{\mu }=0,$

Equation (6)

The result from dividing both equations and differentiating with respect to ρ, ur, and r is

Equation (7)

and after we substitute $d\rho /\rho $ in Equation (7) and reorder terms we obtain $\frac{{{du}}^{r}}{{u}^{r}}\left[{V}^{2}-{\left(\frac{{u}^{r}}{{u}_{0}}\right)}^{2}\right]+\frac{{dr}}{r}\left[2\;{V}^{2}-\frac{M}{{{ru}}_{0}^{2}}\right]=0,$ where ${V}^{2}\equiv \frac{d\mathrm{ln}(\rho h)}{d\mathrm{ln}\rho }-1.$ In order to find a solution for ur and r, the condition must at the same time satisfy the relations

Equation (8)

, and the results of solving (8) are the critical points ${u}_{c}^{r}=\frac{M}{2{r}_{c}}$ and ${V}_{c}^{2}=\frac{{u}_{c}^{r}}{{({u}_{c}^{r})}^{2}-{({g}_{00})}_{c}}.$

On the other hand, if we consider that the gas obeys a polytropic equation of state $P=\kappa {\rho }^{{\rm{\Gamma }}},$ and if a value of Γ and ${\rho }_{c}$ is fixed, we can calculate the polytropic constant κ.

Once we know the critical values ${u}_{c}^{r},$ Vc, and κ, the constants c1 and c2 can be calculated and Equations (6) can be solved. Although we have constructed an exact solution, the system cannot be solved analytically and for this reason we programmed a Newton–Raphson root finder to perform this task.

In order to set the Michel solution as an initial condition into the ETK GRHydro thorn, we first interpolate the hydro variables into the ETK 3D Cartesian grid, and then perform the tensor coordinate transformation of the velocity field. In order to set a non-rotating black hole spacetime we only set the black hole rotation parameter to a = 0 on the KS metric. Given that Michel accretion is steady, all of the hydrodynamical variables have to remain time-independent. In Figure 10, we show the superposed snapshots of the evolution of density, pressure, and velocity profile projected along the x-axis so as the accretion rate measured at a surface located at r = 3 M. The critical values used in this case are ${\rho }_{c}=1\times {10}^{-2},$ ${r}_{c}=400\;M,$ whereas the polytropic index is ${\rm{\Gamma }}=4/3$ Papadopoulos & Font (1998). It can be observed that the density, pressure, velocity profile, and the accretion rate remain time-independent.

Figure 10.

Figure 10. Superposed snapshots of the rest-mass energy density, pressure, and vx velocity every unit of time from t = 0 to t = 400 M along the x-axis. We compare the solution with the Michel exact solution we started with. We also show the accretion mass rate $\dot{M},$ which shows an initial transient and stabilizes afterward.

Standard image High-resolution image

This simple test shows that the inflow boundary conditions we implemented are correct, in all six faces of the boundary. We set the numerical domain to ${[15\;M,15\;M]}^{3},$ covered it using a resolution of $0.25\;M$ in all directions, and set an excision radius to ${r}_{\mathrm{exc}}=1.5\;M.$ We ran the test using HLLE as a Riemann solver and the minmod limiter.

Please wait… references are loading.
10.1088/0004-637X/812/1/23