This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\volnopage

Vol.0 (20xx) No.0, 000–000

11institutetext: South-Western Institute for Astronomy Research, Yunnan University, University Town, Chenggong, Kunming 650500, People’s Republic of China; [email protected].
22institutetext: Hakodate Campus, Hokkaido University of Education, Hachiman-Cho 1-2, Hakodate, Hokkaido 040-8567, Japan
33institutetext: Department of Astronomy, Xiamen University, Xiamen, Fujian 361005, People’s Republic of China
\vs\no

Received  20xx month day; accepted  20xx  month day

Effects of resistivity on standing shocks in low angular momentum flows around black holes

Chandra B. Singh 11    Toru Okuda 22    Ramiz Aktar 33
Abstract

We study two dimensional low angular momentum flow around the black hole using the resistive magnetohydrodynamic module of PLUTO code. Simulations have been performed for the flows with parameters of specific angular momentum, specific energy, and magnetic field which may be expected for the flow around Sgr A*. For flows with lower resistivity η=106\eta=10^{-6} and 0.010.01, the luminosity and the shock location on the equator vary quasi-periodically. The power density spectra of luminosity variation show the peak frequencies which correspond to the periods of 5 ×105\times 10^{5}, 1.4 ×105\times 10^{5}, and 5×104\times 10^{4} seconds, respectively. These quasi-periodic oscillations (QPOs) occur due to the interaction between the outer oscillating standing shock and the inner weak shocks occurring at the innermost hot blob. While for cases with higher resistivity η=0.1\eta=0.1 and 1.0, the high resistivity considerably suppresses the magnetic activity such as the MHD turbulence and the flows tend to be steady and symmetric to the equator. The steady standing shock is formed more outward compared with the hydrodynamical flow. The low angular momentum flow model with the above flow parameters and with low resistivity has a possibility for the explanation of the long-term flares with \sim one per day and \sim 5 – 10 days of Sgr A* in the latest observations by Chandra, Swift, and XMM-Newton monitoring of Sgr A*.

keywords:
accretion, accretion disks — magnetohydrodynamics (MHD) — methods:numerical — shock waves — Galaxy: center

1 Introduction

Black hole accretion is the most efficient process which can address the issue of power generated in the neighbourhood of a black hole. Historically, the study of black hole accretion has been based on two extreme cases of accretion process: radiatively inefficient flow called Bondi flow (Bondi (1952); Michel (1972)) and radiatively efficient one called Keplerian disk (Shakura & Sunyaev (1973); Novikov & Thorne (1973)). Both suffer from certain limitations. Spherical Bondi flow with zero angular momentum is quite fast and cannot explain the high luminosities associated with observational signatures around the black hole. However, in reality, accretion flow is supposed to have some amount of angular momentum associated with it. On the other hand, cold, thin, Keplerian disk cannot explain the issue of change of spectral states and associated temporal variabilities and it is not applicable for the close region around the black hole as pressure gradient and advective radial velocity terms are ignored.

Accretion flow onto the black hole is supposed to be supersonic at the event horizon and subsonic at a large distance as the accretion flow approaches the speed of light at the horizon with sound speed being of lesser value. So the flow with angular momentum must pass through at least one sonic point before plunging onto the black hole and should be advective (Liang & Thompson (1980)). In case of accretion onto a star, even a small angular momentum will stop the matter fall onto its surface because of the infinite potential barrier associated with the Newtonian potential. Whereas in the case of a black hole, gravity always wins over centrifugal force because of higher-order terms (Chakrabarti (1993)). Not only that, for given values of specific energy and specific angular momentum of accretion flow around the black hole, multiple sonic points may also exist with the possibility of standing shocks (Fukue (1987); Chakrabarti (1989)). There have been several interesting works which explored the solutions with standing shocks in pseudo-Newtonian potential (Paczyńsky & Wiita (1980)) taking into account various prescriptions for alpha parameter (Chakrabarti (1996); Becker, Das & Le (2008); Kumar & Chattopadhyay (2013)). General relativistic solutions for inviscid (Das et al. (2015)) and viscous disks (Chattopadhyay & Kumar (2016)) with standing shocks have also been reported. Even in the presence of magnetic field, formation of standing shocks in accretion flows have been explored (Takahashi et al. (2006); Fukumura et al. (2007)).

In recent time the model which has wide recognition is the advection dominated accretion flow (ADAF) (Narayan & Yi (1994); Narayan, Kato & Honma (1997)) solution which takes care of inner boundary condition around the black hole however, has only one sonic point close to the black hole. It should be noted that advective flow with multiple sonic points may not necessarily be ADAF-type especially when standing shock exists in the accretion flow (Chakrabarti (1996)). Overall, ADAF solutions occupy a small region of parameter space for given specific energy and specific angular momentum (Lu, Gu, Yuan (1999); Kumar & Chattopadhyay (2013, 2014)). The need of a sub-Keplerian component was presented in addition to Keplerian one, the sub-Keplerian component can undergo shock transition and form a hot, puffed up region like corona (Chakrabarti & Titarchuk (1995)). The properties of post-shock region formed from the natural course of flow dynamics can address issues like state transitions (Mandal & Chakrabarti (2010)), origin of hard power-tail and low frequency QPOs (Chakrabarti, Mondal & Debnath (2015)) and also the origin of outflows (Das et al. (2001); Singh & Chakrabarti (2011); Aktar, Das & Nandi (2015)).

In the last 25 years, there have been a significant amount of simulation works dedicated to explore the formation of standing shock in low angular momentum sub-Keplerian advective flows around black holes. Using smoothed particle hydrodynamics (SPH) simulations, stable standing shocks were shown to form in one-dimensional (Chakrabarti & Molteni (1993)) and two-dimensional setups (Molteni, Lanzafame & Chakrabarti (1994)) as predicted by semi-analytical solutions of inviscid flows (Chakrabarti (1989)). For the first time, the dependence of standing shock stability on values of viscosity parameters was shown by SPH simulations as well (Chakrabarti & Molteni (1995)). The origin of outflows from the post-shock region in accretion disks were shown in simulations using SPH (Molteni, Ryu & Chakrabarti (1996)), Eulerian total variation diminishing (TVD) (Ryu, Chakrabarti & Molteni (1997); Okuda, Teresi & Molteni (2007)) and Lagrangian TVD (Lee et al. (2016)). In the presence of cooling, the post-shock region may oscillate as the cooling time scale becomes comparable to free-fall time scale and can be responsible for quasi-periodic oscillations(QPOs) in case of stellar mass as well as supermassive black holes (Molteni, Sponholz & Chakrabarti (1996); Okuda, Teresi & Molteni (2007)). Besides the case of inviscid flow, viscosity can also induce shock oscillations and give rise to QPOs (Lanzafame, Molteni & Chakrabarti (1998); Chakrabarti, Acharrya & Molteni (2004); Lanzafame et al. (2008); Lee, Ryu & Chattopadhyay (2011); Das et al. (2014); Lee et al. (2016)). There have been some works regarding stability or instability of the shock and shocks seem to be stable against axisymmetric (Nakayama (1992, 1994); Nobuta & Hanawa (1994); Le et al. (2016)) as well as non-axisymmetric perturbations (Molteni, Toth & Kunetsov (1999); Gu & Foglizzo (2003); Gu & Lu (2006)). Recently it has established through numerical simulations that advective flow can be segregated into two components, Keplerian as well as sub-Keplerian, in the presence of viscous heating and cooling processes (Giri & Chakrabarti (2013); Giri, Garain & Chakrabarti (2015); Roy & Chakrabarti (2017)). All the above-mentioned simulation works addressed the accretion flow behaviour around a non-rotating black hole using pseudo-Newtonian potential. Recently general relativistic high-resolution shock-capturing simulation code was used to study the scenario in the Schwarzschild (Kim et al. (2017)) and Kerr (Kim et al. (2019)) space-time which further established the formation of standing shock in hydrodynamic flow around non-rotating as well as rotating black hole in full general relativistic treatment. However, till now there has been only one work taking into account different magnitudes of magnetic field strength in such flows in the presence of standing shocks (Okuda et al. (2019)). The long term evolution properties were investigated and long term flares in connection with Sgr A* could be explained.

Simulation works dealing with advective flows usually take into account two types of setup : torus equilibrium solution (e.g., (Stone & Pringle (2001); McKinney, Gammie (2002)) and Bondi flow along with arbitrary choice of specific angular momentum (Proga & Begelman (2003)). Our study involves a third and different kind of set up where we take initial conditions based on exact solutions of hydrodynamic equations (Chakrabarti (1989)). We are dealing with a big gap in parameter space which lie between the regime of high angular momentum torus and zero angular momentum Bondi flow. To make our study simpler, we are dealing with inviscid flows having constant and small specific angular momentum value which are lower than that of Keplerian value of the specific angular momentum for an inner-most stable orbit. Such low angular momentum flows are likely to be present in binary systems accreting winds from companion beside Roche lobe overflow as well as active galactic nuclei where winds from stellar clusters collide and lose angular momentum before getting accreted onto central black holes (Chakrabarti & Titarchuk (1995); Smith et al. (2001, 2002); Moscibrodzka et al. (2006)). The objective of our work is to study the effects of resistivity with the varying magnitude on the formation and stability of standing shocks in low angular momentum accretion flows around the black holes which have not been explored before.

Section 2 shows details of 1.5 dimensional (1.5D) theoretical solutions which have been used for the simulation set up. In section 3, basic equations solved by simulation code are presented. Besides, the details of computational domain, initial, and boundary conditions are described in section 3. Section 4 contains details of numerical results followed by section 5 where we present a summary and discussion of our work.

2 Theoretical solution

We consider a semi-analytical approach of solving the standard conservation equations under hydrodynamic (HD) framework. The calculations are done in cylindrical co-ordinates with co-ordinates RR and zz. Axi-symmetry is assumed for the angular ϕ\phi co-ordinate. For simplicity, we further assume that the flow velocity along the vertical direction is zero and therefore only integrate along the radial co-ordinate assuming vertically averaged dynamical quantities.

We define the scale radius as Schwarzschild radius rgr_{g}= 2GM/c22GM/c^{2}, with MM being the mass of the central compact object, G as the gravitational constant and cc is the speed of light. The matter that is accreted onto the central compact object has radial velocity given by uRu_{\rm R}, specific angular momentum LL, and total specific energy \cal E. As the semi-analytical calculations are carried out using non-dimensional quantities, we define the following -

r=Rrg;h=zrg;vR=uRc;λ=Lrgc;ϵ=c2r=\frac{R}{r_{g}};\,\,\,\,\,h=\frac{z}{r_{g}};\,\,\,\,\,v_{\rm R}=\frac{u_{\rm R}}{c};\,\,\,\,\,\lambda=\frac{L}{r_{g}c};\,\,\,\,\,\,\epsilon=\frac{\cal E}{c^{2}} (1)

For studying ideal, inviscid flow onto a compact object, we deal with mass conservation equation,

M˙=4πρvRrh,\dot{M}=4\pi\rho v_{\rm R}rh, (2)

where ρ\rho is the density and hh being the half-thickness of flow. Energy conservation gives us a relation of specific energy of the flow or Bernoulli constant,

ϵ=vR22+cs2Γ1+λ22r2+Φ.\epsilon=\frac{v_{\rm R}^{2}}{2}+\frac{c_{s}^{2}}{\Gamma-1}+\frac{\lambda^{2}}{2r^{2}}+\Phi. (3)

Here csc_{s} is adiabatic sound speed in units of cc and Γ\Gamma is the adiabatic index. cs=Γpρc2c_{s}=\sqrt{\frac{\Gamma p}{\rho c^{2}}}. pp is the thermal pressure and Φ\Phi is the non-dimensional gravitational potential given by 1/2(x1)-1/2(x-1) for non-rotating black hole (Paczyńsky & Wiita (1980)) where x=rsp/rgx=r_{sp}/r_{g} and rspr_{sp} being the spherical radius.

Using vertical equilibrium condition, we evaluate the radial dependence for hh

h=csx(x1)h=c_{s}\sqrt{x}(x-1) (4)

For given values of ϵ\epsilon and λ\lambda, we solve equations (2) and (3) and look for transonic conditions. Differentiating the equations (2) and (3), we obtain

dvRdr[vR2cs2(Γ+1)vR]=2cs2Γ+1dlnfdrdGdr,\frac{dv_{\rm R}}{dr}[v_{\rm R}-\frac{2c_{s}^{2}}{(\Gamma+1)v_{R}}]=\frac{2c_{s}^{2}}{\Gamma+1}\frac{dlnf}{dr}-\frac{dG}{dr}, (5)

where G=λ2/2r21/2(r1)G=\lambda^{2}/2r^{2}-1/2(r-1) and f=2r3/2(r1)f=2r^{3/2}(r-1)
(Chakrabarti (1989)). At critical points, the vanishing of left-hand side gives radial velocity,

(vR)crit=2Γ+1(cs)crit,(v_{\rm R})_{crit}=\sqrt{\frac{2}{\Gamma+1}}(c_{s})_{crit}, (6)

and the vanishing of right hand side gives sound speed,

(cs)crit=(Γ+1)(rcrit2)rcrit2(λK2λ2)(5rcrit2).(c_{s})_{crit}=\frac{(\Gamma+1)(r_{crit}-2)}{r_{crit}^{2}}\frac{(\lambda_{K}^{2}-\lambda^{2})}{(5r_{crit}-2)}. (7)
Refer to caption
Figure 1: Mach number versus radial distance from semi-analytical estimates for flow parameters, specific energy, ϵ\epsilon = 1.98 ×\times 10610^{-6} and specific angular momentum, λ=1.35\lambda=1.35, with Γ\Gamma = 1.6 (adopted from Okuda et al. (2019)).

The subscript critcrit and KK represent quantities at the critical points and Keplerian orbits respectively. In the case we obtain multiple critical points, we also check whether shock conditions are satisfied or not along the accretion flow. If shock conditions are satisfied, then there is a possibility of standing axisymmetric shock otherwise not. The shock location can be determined using an invariant quantity, CC, across the shock which is given by

C=[+(3Γ1)+(2/+)]22+(Γ1)+2=[(3Γ1)+(2/)]22+(Γ1)2.C=\frac{[\mathcal{M_{+}}(3\Gamma-1)+(2/\mathcal{M_{+}})]^{2}}{2+(\Gamma-1)\mathcal{M}_{+}^{2}}=\frac{[\mathcal{M_{-}}(3\Gamma-1)+(2/\mathcal{M_{-}})]^{2}}{2+(\Gamma-1)\mathcal{M}_{-}^{2}}. (8)

Here, =vR/cs\mathcal{M}=v_{\rm R}/c_{s} is the Mach number of the accretion flow. The subscripts - and ++ represent quantities in the pre-shock and post-shock region respectively. Further details of the semi-analytical approach can be found in (Chakrabarti (1989)).

Fig. 1 shows the variation of Mach number, \mathcal{M}, of flow with radial distance from black hole obtained from exact theoretical solution solving conservation equations. The transonic flow passes through the outer critical point “a” and continues its journey towards the black hole. The flow chooses to undergo shock transition along “bc”, becomes subsonic then again accelerates towards the black hole and passes through inner critical point “d” to become supersonic before entering the black hole horizon. The shocked flow is preferrable in nature as the entropy generation is relatively higher compared to no shock flow.

3 Numerical Setup

3.1 Basic Equations

The numerical setup for the present work uses grid-based, finite volume computational fluid dynamics code, PLUTO (Mignone et al. (2007, 2012)). Numerical simulations are carried out by solving the equations of classical resistive magnetohydrodynamics (MHD) in the conservative form:

ρt+(ρv)\displaystyle\frac{\partial\rho}{\partial t}+\nabla\cdot\left(\rho\textbf{v}\right) =\displaystyle= 0,\displaystyle 0, (9)
(ρv)t+[ρvvBB]+pt\displaystyle\frac{\partial(\rho\textbf{v})}{\partial t}+\nabla\cdot[\rho\textbf{v}\textbf{v}-\textbf{B}\textbf{B}]+\nabla p_{t} =\displaystyle= ρΦ,\displaystyle-\rho\nabla\Phi, (10)
Et+[(E+pt)v(vB)B+η(×B)×B]\displaystyle\frac{\partial E}{\partial t}+\nabla\cdot[(E+p_{t})\textbf{v}-(\textbf{v}\cdot\textbf{B})\textbf{B}+\eta(\nabla\times\textbf{B})\times\textbf{B}] =\displaystyle= ρvΦ,\displaystyle-\rho\textbf{v}\cdot\nabla\Phi, (11)
Bt×(v×Bη×B)\displaystyle\ \frac{\partial\textbf{B}}{\partial t}-\nabla\times(\textbf{v}\times\textbf{B}-\eta\nabla\times\textbf{B}) =\displaystyle= 0.\displaystyle 0. (12)

Here, ptp_{t} being the total pressure with contribution from thermal pressure, pp, and magnetic pressure, B2/2B^{2}/2. EE is the total energy density given by

E=pΓ1+12(ρv2+B2).\displaystyle E=\frac{p}{\Gamma-1}+\frac{1}{2}({\rho\textbf{v}^{2}}+\textbf{B}^{2}). (13)

η\eta is the resistivity for which range of values have been chosen, 10610^{-6}, 0.010.01, 0.10.1 and 11. Vector potential A is prescribed to generate the magnetic field B as B=×A\textbf{B}=\nabla\times\textbf{A}. Following Proga & Begelman (2003), the components of A are as follows AR=0A_{R}=0, Aϕ=A0zrspRA_{\phi}=\frac{A_{0}z}{r_{sp}R} and Az=0A_{z}=0. Here, A0=sign(z)(8πpoutβout)1/2Rout2A_{0}=sign(z)(\frac{8\pi p_{out}}{\beta_{out}})^{1/2}R_{out}^{2} and βout=8πpout/Bout2\beta_{out}=8\pi p_{out}/B_{out}^{2}, subscript ”out” denotes parameters at the outer boundary of the computational domain, RoutR_{out}. Following Okuda et al. (2019), we take a typical value of 5000 for βout\beta_{out}.

Refer to caption
Figure 2: Flow parameters on the equator, namely density (ρ\rho in gcm3cm^{-3}), radial velocity (vRv_{R}) and temperature (T), for final state of HD simulation run. The standing shock is at 64.8Rg64.8R_{g} (adapted from Okuda et al. (2019)).

3.2 Initial and Boundary Conditions

First, we have studied advective flows onto black holes in axisymmetric 2D cylindrical geometry (RR, zz) in HD framework. The theoretical solutions given in section 2 provide initial conditions of primitive variables, radial, and azimuthal velocity components, density, and pressure for 2D HD set up. Once the HD flow with the standing shock achieves a steady state, we use the solutions as the initial conditions for the magnetized flow with resistivity and let the simulation evolve further. The computational domain is 0R200Rg0\leq R\leq 200R_{g} and 200Rgz200Rg-200R_{g}\leq z\leq 200R_{g} with the resolution of 1640×8201640\times 820 cells (for details, see Okuda et al. (2019)). Though we performed some simulation runs at half resolution, 820×410820\times 410, and double resolution, 3280×16403280\times 1640, the results remain unchanged.
In both HD and MHD runs, the same boundary conditions are imposed. At the outer radial boundary, Rout=200RgR_{out}=200R_{g}, there are two domains: the disk region where the matter is injected and the atmosphere above the disk region. The flow parameters given by 1.5D theoretical solutions are provided in the region houtzhout-h_{out}\leq z\leq h_{out} where houth_{out} is the vertical equilibrium height at RoutR_{out}. For the atmosphere region, the matter is allowed to leave the domain but not enter. The axisymmetric boundary condition is implemented at the inner radial boundary. At R=2RgR=2R_{g}, the absorbing condition is imposed in the computational domain. In the vertical direction, z=±200Rgz=\pm 200R_{g}, standard outflow boundary conditions are imposed. In the case of MHD run, the constant magnetic field is imposed on the outer radial boundary. Fig. 2 shows profiles of density (ρ\rho in gcm3cm^{-3}), radial velocity (vRv_{R}) and temperature(T) of the HD flow. The standing shock location from the simulation run is at 65Rg\sim 65R_{g} which is significantly different from the predicted location from the theoretical solution, i.e. 20Rg\sim 20R_{g}. This is due to the assumption of the vertical hydrostatic equilibrium used in the 1.5D transonic solutions, which is valid as far as the disk thickness hh is sufficiently small compared with the radius rr (that is, h/rh/r\ll 1). However, in many cases of the low angular momentum flow with the standing shock, h/rh/r\sim 0.1 – 0.5 because such flows are intrinsically advective and geometrically thick. Therefore if the outer radial boundary is chosen to be very far from the predicted theoretical shock location, the difference between the numerical and theoretical shock locations becomes significant (Okuda et al. (2019)).

4 Results and discussion

To get the characteristic features of the flow, we examine the time evolution of shock location RsR_{s} on the equator and total luminosity LL of the flow. The luminosity LL is calculated as follows, assuming that the gas is optically thin

L=qff𝑑V,\displaystyle L=\int q_{ff}dV, (14)

where qffq_{ff} represents the free-free emission rate per unit volume and integration is performed over all the whole computational domain. In order to have some estimate of how much matter is lost as an outflow from the system, mass outflow rate in the z-direction is given by,

M˙out=0Rout2πρ(R,zout)vz(R,zout)R𝑑R0Rout2πρ(R,zout)vz(R,zout)R𝑑R,\displaystyle\dot{M}_{out}=\int_{0}^{R_{out}}2\pi\rho(R,z_{out})v_{z}(R,z_{out})RdR-\int_{0}^{R_{out}}2\pi\rho(R,-z_{out})v_{z}(R,-z_{out})RdR, (15)

where vzv_{z} is the vertical velocity.

In the present study, the time variability of the flow will correlate with the magnetorotational instability (MRI). We check whether the flow is subject to the MRI and whether we can resolve the fastest growing MRI mode or not. The stringent diagnostics of spatial resolution for the MRI instability has been examined in 3D magnetized flow. Therefore, its application to our 2D magnetized flow may be limited. The critical wavelength of the instability mode is given by λc=2πvA/3Ω\lambda_{\rm c}=2\pi v_{\rm A}/\sqrt{3}\Omega, where vAv_{\rm A} and Ω\Omega are the Alfven velocity and the angular velocity (Hawley & Balbus (1991); Balbus & Hawley (1998)). A criterion value QxQ_{\rm x} of the MRI resolution is defined by

Qx=λcΔx,Q_{\rm x}=\frac{\lambda_{\rm c}}{\Delta x}, (16)

where Δx\Delta x is the mesh sizes ΔR\Delta R and Δz\Delta z in the radial and vertical directions, respectively. When QxQ_{\rm x}\gg 1, the flow is unstable against the MRI instability, otherwise, the flow is stable. Fig. 3 shows 2D contours of radial MRI-criterion QrQ_{r} for cases of η=106\eta=10^{-6} and 1.0 at times tt= 7×1067\times 10^{6} and 8.7×1068.7\times 10^{6} seconds, respectively. The analyses show QrQ_{r}\gg 1 in most regions for both cases. So, both flows are unstable to the MRI. The contours are asymmetric to the equator in the former but symmetric in the latter.

Refer to caption Refer to caption
Figure 3: 2D contours of MRI-criterion QrQ_{r} for cases of η=106\eta=10^{-6} and 1.0 at times tt = 7×1067\times 10^{6} and 8.7×1068.7\times 10^{6} seconds, respectively. For both cases, QrQ_{r}\gg 1 in most regions. The contours are asymmetric to the equator in the former but symmetric in the latter.

4.1 Luminosity, shock location and mass in/outflow rates

Fig. 4 shows how luminosity LL and shock location RsR_{s} vary with time for different levels of resistivity in the flow. For lower values of resistivity, η=106\eta=10^{-6} and 0.010.01, there are features of irregular oscillation in the luminosity LL and the standing shock location RsR_{s}. As the shock moves towards the black hole, there is an increase in luminosity while the luminosity decreases when the shock recedes away. The shock and the luminosity oscillate irregularly with time scales of 105106\sim 10^{5}-10^{6} s, and the luminosity varies maximumly by a factor of ten around the average LL is 3.0×1034\sim 3.0\times 10^{34} erg s-1. On the other hand, for relatively high resistivity, 0.10.1 and 1.01.0, the oscillatory nature disappears and LL and RsR_{s} show small modulations or almost steady value at later times. The highly resistive flows behave qualitatively similar to that of HD flow.

Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 4: Variation of Luminosity (L) and shock location (RsR_{s}) with time for resistive MHD flow with different values of resistivity, η\eta = 10610^{-6}, 0.01, 0.1 and 1 (in clockwise direction).
Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 5: Mass inflow (M˙edge\dot{M}_{edge}) as well as outflow (M˙out\dot{M}_{out}) rate evolving with time for resistive MHD flow with different values of resistivity, η\eta = 10610^{-6}, 0.01, 0.1 and 1 (in clockwise direction).

The mass inflow M˙edge\dot{M}_{edge} and outflow M˙out\dot{M}_{out} rates are presented in Fig. 5 correspondings to Fig. 4. Similar to our previous work Okuda et al. (2019), it has been established that there is a correlation between LL and M˙edge\dot{M}_{edge} and between RsR_{s} and M˙out\dot{M}_{out}. While the variation of LL seems to be opposite in behaviour compared to RsR_{s}. That means when the post-shock region shrinks, the emission increases, and vice versa. Since the low angular momentum flows are very advective, most of the input gas M˙input\dot{M}_{input} (3×1020\sim 3\times 10^{20} g s-1) falls into the event horizon and M˙edge\dot{M}_{edge} is comparable to M˙input\dot{M}_{input} in all cases. However, the mass outflow rate M˙out\dot{M}_{out} in low resistivity case is considerably high as a few tens of percent of the input accretion rate but in the high resistive case with η\eta = 0.1 and 1.0, M˙out\dot{M}_{out} amounts to \sim 10 percent. Such mass outflow rates are very high compared with the mass outflow rate found in the usual accretion flow. It should be noted that the very high mass outflow rate in the low resistivity case may be correlated to the MRI turbulence, that is, the MRI turbulence plays important roles not only in the outward transfer of the angular momentum but also in outward mass transfer.

Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 6: Radial profiles of normalized Reynolds stress αgas\alpha_{gas} and normalized Maxwell stress αmag\alpha_{mag} for resistive MHD flow with different resistivity, η\eta = 10610^{-6}, 0.01, 0.1 and 1 (in clockwise direction). These values are space-averaged (between 2Rg-2R_{g} and 2Rg2R_{g} in z-direction) and time averaged (between 1.1×1071.1\times 10^{7} and 1.2×1071.2\times 10^{7} seconds).
Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 7: Radial profiles of gas pressure and magnetic pressure which are space-averaged (between 2Rg-2R_{g} and 2Rg2R_{g} in z-direction) and time averaged (between 1.1×1071.1\times 10^{7} and 1.2×1071.2\times 10^{7} seconds) for resistive MHD flow with different resistivity, η\eta = 10610^{-6}, 0.01, 0.1 and 1 (in clockwise direction).

4.2 Effects of the resistivity on magnetized flow

The resistivity has dissipative and diffusive characters in the magnetic field through the current density, similar to the viscosity in hydrodynamical flow and we expect the higher resistivity to suppress the magnetic activity like magnetic turbulence. We examine the effects of resistivity through the time evolution of the magnetized flows with η=106\eta=10^{-6} to 1.0. In the case with the lowest resistivity 10610^{-6}, after a transient initial time evolution, the magnetic field is amplified rapidly by the MRI and the MHD turbulence develops near the equatorial plane. Fig. 6 shows radial profiles of normalized Reynolds stress αgas\alpha_{gas} and normalized Maxwell stress αmag\alpha_{mag} for resistive MHD flow with different resistivity, η\eta = 10610^{-6}, 0.01, 0.1 and 1 (in clockwise direction). These values are space-averaged (between 2Rg-2R_{g} and 2Rg2R_{g} in the z-direction) and the time averaged over the last duration time. Here we see that the maxwell stress Σmag\Sigma_{mag} is larger by a factor of a few to ten than the Reynolds stress Σgas\Sigma_{gas} in cases of lower η=106\eta=10^{-6} and 0.01 over the most region, while in the higher η\eta the Reynolds stress mostly dominates over the Maxwell stress. From this, we confirm that the higher resistivity suppresses the Maxwell stress and then MHD turbulence, that is, hydrodynamical mode dominates over magnetohydrodynamical mode. As the result, in the case with the highest resistivity, the flow is dominated by the hydrodynamical quantities at the outer radial boundary which are symmetric to the equator and the flow achieves a steady and symmetric state. Fig. 7 denotes radial profiles of the gas pressure and the magnetic pressure for MHD flow with resistivity, η\eta = 10610^{-6}, 0.01, 0.1 and 1 (in clockwise direction). In all cases with different resistivity, the gas pressure dominates the magnetic pressure and the pressure distributions are not so different each other.

Fig. 8 shows 2D density contours and velocity vectors at the later evolution of the flow with η=106\eta=10^{-6} and 1.0 at times tt = 7 ×\times 10610^{6} and 8.7 ×\times 10610^{6} seconds respectively. Here, the location of standing shock is distinguished as the thick black contour lines, and the velocity vectors are taken to be an arbitrary unit. In the low resistivity case, the density contours are asymmetric to the equator and turbulent motions are observed in the shocked region. While in the high resistivity case variables become symmetric to the equator and no turbulent motion is observed. The flow features seem to return to the initial hydrodynamical steady-state but with a bit larger shock location Rs70RgR_{s}\sim 70R_{g} than Rs65RgR_{s}\sim 65R_{g} in Fig. 2, because the magnetic pressure contributes to the pressure balance to some extent in the shock location.

Refer to caption Refer to caption
Figure 8: 2D density contours and velocity vectors of flows with η=106\eta=10^{-6} and 1.0 at times tt = 7 ×\times 10610^{6} and 8.7 ×\times 10610^{6} seconds, respectively. In the former, variables of density and temperature are asymmetric to the equator and turbulent motions are observed within the post-shock region but in the latter case the flow is almost symmetric to the equator, and no turbulent motion appears. The shock locations are denoted by thick black contour lines.

4.3 Astrophysical significance

The present results for cases with low resistivity of η=106\eta=10^{-6} and 0.01 are very similar to those for the previous magnetized flow without resistivity (Okuda et al. (2019)). Adopting the same parameters of the flow and magnetic field as the present study, they found that the centrifugally supported shock moves back and forth between 60 RgR170RgR_{\rm g}\leq R\leq 170R_{\rm g} and that another inner weak shock appears irregularly with rapid variations due to the interaction of the expanding high magnetic blob with the accreting matter below the outer shock. The process repeats irregularly with an approximate time-scale of (4 – 5) ×105\times 10^{5} s (\sim 5 days) with an accompanying smaller amplitude modulation with a period of 0.9×105\sim 0.9\times 10^{5} s (25 hrs). In this respect, we also analyzed the time variability of the resistive magnetized flows. Fig. 9 show the power density spectra of luminosity for different values of resistivity. For η\eta = 10610^{-6} and 0.010.01, the peak (fundamental) frequency is estimated roughly to be at 2 ×\times 10610^{-6} along with two weak signatures (harmonics) at 7 ×\times 10610^{-6} and 2 ×\times 10510^{-5} Hz. These correspond to the periods of 5 ×105\times 10^{5}s (5.8 days), 1.4 ×105\times 10^{5}s (1.6 days), and 5×104\times 10^{4}s (0.6 day) , respectively and are comparable to two QPOs periods 5\sim 5 days and 1\sim 1 day found in the non-resistive magnetized flow. Therefore the QPO peak frequencies can be associated with periods of \sim 5 – 10 days and \sim 1 day X-ray flares observed in the latest observations by Chandra, Swift, and XMM-Newton monitoring of Sgr A* (Degenaar et al. (2013); Neilsen et al. (2013, 2015); Ponti et al. (2015)). On the other hand, for η\eta = 0.10.1 and 1.01.0, there is no clear peak frequency. The average mass outflow rate 106\sim 10^{-6} Myr1M_{\odot}yr^{-1} and mass inflow rate 3×106Myr1\sim 3\times 10^{-6}M_{\odot}yr^{-1} obtained in small η\eta cases show a roughly good correspondence with the Chandra observations (Wang et al. (2013)) which suggest the presence of a high outflow rate that nearly balances the inflow rate.

Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 9: Power density spectra for resistive MHD flow with different resistivity values, η\eta = 10610^{-6}, 0.01, 0.1 and 1 (in clockwise direction).

5 Summary

We studied the effect of resistivity on standing shock in the magnetized flow around a black hole. The flow parameters of specific energy, ϵ\epsilon = 1.98 ×\times 10610^{-6} and specific angular momentum, λ=1.35\lambda=1.35, with Γ\Gamma = 1.6 have been considered to address the flow behaviour around Sgr A*. For flows with lower resistivity η=106\eta=10^{-6} and 0.010.01, the luminosity and the shock location on the equator vary quasi-periodically. These quasi-periodic oscillations are attributed to the interactive result between the outer oscillating standing shock and the inner weak shocks occurring at the innermost hot blob. The luminosity varies maximumly by a factor of ten around the average LL 3.0×1034\sim 3.0\times 10^{34} erg s-1. The mass outflow rate is very large as a few tens of percent of the input accretion rate. The MHD turbulence seems to play important roles in the outward transport of not only angular momentum but also accreting gas. The power density spectra of luminosity variation show the peak frequencies which correspond to the periods of 5 ×105\times 10^{5}s (5.8 days), 1.4 ×105\times 10^{5}s (1.6 days) and 5×104\times 10^{4}s (0.6 day) , respectively. While for cases with higher resistivity η=0.1\eta=0.1 and 1.0 the flow becomes steady and symmetric to the equator. Variable features of the luminosity disappear here and the steady standing shock is formed more outward compared with the hydrodynamical flow. The mass outflow rate is also high as \sim a few tens %\% of the input gas. The high resistivity considerably suppresses the magnetic activity such as the MHD turbulence and tends to form the magnetized flow to be stable and symmetric to the equator. The low angular momentum magnetized flow model with low resistivity has possibility for the explanations of the high mass outflow rate \sim as 10%\% of the Bondi accretion rate \sim 1 ×\times 10510^{-5} Myr1M_{\odot}yr^{-1} as suggested by Chandra observations (Wang et al. (2013)) and of the long-term flares with \sim one per day and \sim 5 – 10 days of Sgr A* in the latest observations by Chandra, Swift, and XMM-Newton monitoring of Sgr A*.

Acknowledgements.
CBS is supported by the National Natural Science Foundation of China under grant no. 12073021. RA acknowledges support from National Science Foundation of China under grant No. 11373002, and Natural Science Foundation of Fujian Province of China under grant No. 2018J01007.

References

  • Aktar, Das & Nandi (2015) Aktar R., Das S., Nandi A., 2015, MNRAS, 453, 3414
  • Balbus & Hawley (1998) Balbus S. A., Hawley J. F., 1998, Rev. Mod. Phys, 70, 1
  • Becker, Das & Le (2008) Becker P.A., Das S., Le T., 2008, ApJ, 677, L93
  • Bondi (1952) Bondi H., 1952, MNRAS, 1952, 112, 195
  • Chakrabarti (1989) Chakrabarti S. K., 1989, ApJ, 347, 365
  • Chakrabarti (1993) Chakrabarti S. K., 1993, MNRAS, 261, 625
  • Chakrabarti (1996) Chakrabarti S. K., 1996, ApJ, 464, 664
  • Chakrabarti, Acharrya & Molteni (2004) Chakrabarti S. K., Acharyya K., Molteni D., 2004, A&A, 421, 1
  • Chakrabarti & Molteni (1993) Chakrabarti S. K., Molteni D., 1993, ApJ, 417, 671
  • Chakrabarti & Molteni (1995) Chakrabarti S. K., Molteni D., 1995, ApJ, 272, 80
  • Chakrabarti, Mondal & Debnath (2015) Chakrabarti S. K., Mondal S., Debnath D., 2015, MNRAS, 452, 3451
  • Chakrabarti & Titarchuk (1995) Chakrabarti S. K., Titarchuk L., 1995, ApJ, 455, 623
  • Chattopadhyay & Kumar (2016) Chattopadhyay I., Kumar R., 2016, MNRAS, 459, 3792
  • Das et al. (2001) Das S., Chattopadhyay I., Nandi A., Chakrabarti S. K., 2001, A&A, 379, 683
  • Das et al. (2014) Das S., Chattopadhyay I., Nandi A., Molteni D., 2014, MNRAS, 442, 251
  • Das et al. (2015) Das T. K., Nag S., Hegde S., Bhattacharya S., Maity I., Czerny B., Barai P., Wiita P.J., Karas V., Naskar T., 2015, New Astronomy, 37, 81
  • Degenaar et al. (2013) Degenaar N., Miller J. M., Kennea J., Gehrels N., Reynolds M. T., Wijnands, R. 2013, ApJ, 769, 155
  • Fukue (1987) Fukue J., 1987, PASJ, 39, 309
  • Fukumura et al. (2007) Fukumura K., Takahashi M., Tsuruta S., 2007, ApJ, 657, 415
  • Giri & Chakrabarti (2013) Giri K., Chakrabarti S.K., 2013, MNRAS, 430, 2836
  • Giri, Garain & Chakrabarti (2015) Giri K., Garain S.K., Chakrabarti S.K., 2013, MNRAS, 448, 3221
  • Gu & Foglizzo (2003) Gu W.M., Foglizzo T., 2003, A&A, 409, 1
  • Gu & Lu (2006) Gu W.M., Lu J.F., 2006, MNRAS, 365, 647
  • Hawley & Balbus (1991) Hawley, J. F., Balbus, S. A., 1991, ApJ, 376, 223
  • Kim et al. (2017) Kim J., Garain S.K., Balsara D.W., Chakrabarti S.K., 2017, MNRAS, 472, 542
  • Kim et al. (2019) Kim, J., Garain, S. K., Chakrabarti, S. K., Balsara, D. S., 2019, MNRAS, 482, 3636
  • Kumar & Chattopadhyay (2013) Kumar R., Chattopadhyay I., 2013, MNRAS, 430, 386
  • Kumar & Chattopadhyay (2014) Kumar R., Chattopadhyay I., 2014, MNRAS, 443, 3444
  • Lanzafame et al. (2008) Lanzafame G., Cassaro P., Schilliró F., Costa V., Belvedere G., Zappalá R. A., 2008, A&A, 482, 473
  • Lanzafame, Molteni & Chakrabarti (1998) Lanzafame G., Molteni D., Chakrabarti S., 1998, MNRAS, 299, 799
  • Le et al. (2016) Le T., Wood K.S., Wolff M.T., Becker P.A., Putney J., 2016, ApJ, 819, 112
  • Lee, Ryu & Chattopadhyay (2011) Lee S.-J., Ryu D., Chattopadhyay I., 2011, ApJ, 728, 142
  • Lee et al. (2016) Lee S.-J., Chattopadhyay I., Kumar R., Hyung S., Ryu D., 2016, ApJ, 831, 33
  • Liang & Thompson (1980) Liang E.P.T., Thompson K.A., 1980, ApJ, 240, 271L
  • Lu, Gu, Yuan (1999) Lu J.F., W.-M. Gu, Yuan F., 1999, ApJ, 523, 340
  • Mandal & Chakrabarti (2010) Mandal S., Chakrabarti S. K., 2010, ApJ, 710, L147
  • McKinney, Gammie (2002) McKinney J.C., Gammie C.F., 2002, MNRAS, 573, 728
  • Michel (1972) Michel F.C., 1972, AP&SS, 112, 195
  • Mignone et al. (2007) Mignone A., Bodo G., Massaglia S., Matsakos T., Tesileanu O., Zanni C., Ferrari A., 2007, ApJS, 170, 228
  • Mignone et al. (2012) Mignone A., Zanni C., Tzeferacos P., van Straalen B., Colella P., Bodo G., 2012, ApJS, 198, 7
  • Molteni, Lanzafame & Chakrabarti (1994) Molteni D., Lanzafame G., Chakrabarti S. K., 1994, ApJ, 425, 161
  • Molteni, Ryu & Chakrabarti (1996) Molteni D., Lanzafame G., Chakrabarti S. K., 1996, ApJ, 470, 460
  • Molteni, Sponholz & Chakrabarti (1996) Molteni D., Sponholz H., Chakrabarti S. K., 1996, ApJ, 457, 805
  • Molteni, Toth & Kunetsov (1999) Molteni D., Toth G., Kuznetsov, 1999, ApJ, 516, 411
  • Moscibrodzka et al. (2006) Moscibrodzka M, Das T.K., Czerny B., 2006, MNRAS, 370, 219
  • Nakayama (1992) Nakayama K., 1992, MNRAS, 259, 259
  • Nakayama (1994) Nakayama K., 1992, MNRAS, 270, 871
  • Narayan, Kato & Honma (1997) Narayan R., Kato S., Honma F., 1997, ApJ, 476, 49
  • Narayan & Yi (1994) Narayan R., Yi I., 1994, ApJ, 428, L13
  • Neilsen et al. (2013) Neilsen J., et al. 2013, ApJ, 774, 42
  • Neilsen et al. (2015) Neilsen J., et al. 2015, ApJ, 799, 199
  • Nobuta & Hanawa (1994) Nobuta K., Hanawa T., 1994, PASJ, 46, 257
  • Novikov & Thorne (1973) Novikov I.D., Throne K.S., 1973, in Black Holes, ed. B.S. Dewitt & C. Dewitt (New York: Gordon and Breach) 343
  • Okuda et al. (2019) Okuda T., Singh C. B., Das S., Aktar R., Nandi A., de Gouveia Dal Pino E.M., 2019, PASJ, 71, 49
  • Okuda, Teresi & Molteni (2007) Okuda T., Teresi V., Molteni D., 2007, MNRAS, 377, 1431
  • Paczyńsky & Wiita (1980) Paczyńsky B., Wiita P. J., 1980, A&A, 88, 23
  • Ponti et al. (2015) Ponti, G., et al. 2015, MNRAS, 454, 1525
  • Proga & Begelman (2003) Proga D., Begelman M.C., 2003, ApJ, 592, 767
  • Roy & Chakrabarti (2017) Roy A., Chakrabarti S. K., 2017, MNRAS, 472, 4689
  • Ryu, Chakrabarti & Molteni (1997) Ryu D., Chakrabarti S. K., Molteni D., 1997, ApJ, 474, 378
  • Shakura & Sunyaev (1973) Shakura N.L., Sunyaev R.A., 1973, A&A, 24, 337
  • Singh & Chakrabarti (2011) Singh C. B., Chakrabarti S.K., 2011, MNRAS, 410, 2414
  • Smith et al. (2001) Smith D.M., Heindl W.A., Markwardt C.B., Swank J.H., 2001, ApJ, 554, 41
  • Smith et al. (2002) Smith D.M., Heindl W.A., Swank J.H., 2002, ApJ, 569, 362
  • Stone & Pringle (2001) Stone J.M., Pringle J.E., 2001, MNRAS, 322, 461
  • Takahashi et al. (2006) Takahashi M., Goto J., Fukumura K., Rillet D., Tsuruta S., 2006, ApJ, 645, 1408
  • Wang et al. (2013) Wang Q.D., Nowak M.A., Markoff S.B., et al. 2013. Science, 341, 981