A publishing partnership

STOCHASTIC PARTICLE ACCELERATION IN TURBULENCE GENERATED BY MAGNETOROTATIONAL INSTABILITY

, , , and

Published 2016 May 10 © 2016. The American Astronomical Society. All rights reserved.
, , Citation Shigeo S. Kimura et al 2016 ApJ 822 88 DOI 10.3847/0004-637X/822/2/88

0004-637X/822/2/88

ABSTRACT

We investigate stochastic particle acceleration in accretion flows. It is believed that magnetorotational instability (MRI) generates turbulence inside accretion flows and that cosmic rays (CRs) are accelerated by the turbulence. We calculate equations of motion for CRs in the turbulent fields generated by MRI with the shearing box approximation and without back reaction to the field. Our results show that the CRs randomly gain or lose their energy through interaction with the turbulent fields. The CRs diffuse in the configuration space anisotropically: the diffusion coefficient in the direction of the unperturbed flow is about 20 times higher than the Bohm coefficient, while those in the other directions are only a few times higher than the Bohm. The momentum distribution is isotropic and its evolution can be described by the diffusion equation in momentum space where the diffusion coefficient is a power-law function of the CR momentum. We show that the shear acceleration works efficiently for energetic particles. We also cautiously note that in the shearing box approximation, particles that cross the simulation box many times along the radial direction undergo unphysical runaway acceleration by the Lorentz transformation, which needs to be taken into account with special care.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

When the mass accretion rate onto a black hole is sufficiently lower than the Eddington rate, a radiation inefficient accretion flow (RIAF) is formed (Narayan & Yi 1994; Abramowicz et al. 1995; Ohsuga & Mineshige 2011; Yuan et al. 2012; Ressler et al. 2015). Inside RIAFs, the plasmas are so hot and tenuous that non-thermal particles (cosmic rays; CRs) can naturally be generated (e.g., Quataert & Gruzinov 1999; Toma & Takahara 2012; Niedźwiecki et al. 2013, 2015; Kimura et al. 2014). Generally, in accretion flows, magnetorotational instability (MRI) is believed to play a crucial role in the transport of angular momentum. MRI generates strong and turbulent magnetic fields inside RIAFs in which the CRs are likely to be accelerated through the stochastic acceleration process (Dermer et al. 1996; Liu et al. 2004). This process has been analytically formulated using the quasi-linear theory, and the evolution of the momentum distribution function of the CRs has usually been described by the diffusion equation in momentum space (e.g., Schlickeiser & Miller 1998; Stawarz & Petrosian 2008; Ohira 2013). Kimura et al. (2015) investigate stochastic acceleration inside RIAFs of low-luminosity active galactic nuclei by using this formulation, and they find that CRs can be accelerated up to ∼10 PeV if the turbulence is on the Kolmogorov spectrum and that high-energy neutrinos generated through the interaction between CRs and background matter are compatible with the IceCube events (Aartsen et al. 2013, 2015). Protons escaping from RIAFs interact with circumnuclear matter and emit gamma rays (Fujita et al. 2015), which can be consistent with the observed TeV flux from Sgr A* (Aharonian et al. 2009b) and Cen A (Aharonian et al. 2009a). These examples indicate that understanding the acceleration process in accretion flows is imperative for studying the sources of such astrophysical neutrinos and gamma rays.

There are many other candidate sites of the stochastic acceleration process, such as clusters of galaxies (e.g., Blasi 2000; Petrosian 2001; Brunetti & Lazarian 2007), gamma-ray bursts (e.g., Asano & Terasawa 2009; Murase et al. 2012), and blazars (e.g., Katarzyński et al. 2006; Asano et al. 2014; Kakuwa et al. 2015). Also, these studies use the quasi-linear theory, which is constructed based on the assumption of weak and isotropic turbulence, while some of those sites may include strong and/or anisotropic turbulence. Some numerical works support the validity of the quasi-linear theory, even for the case with strong turbulence (e.g., Casse et al. 2002; Dmitruk et al. 2003; O'Sullivan et al. 2009), but recent particle simulations show substantial differences between the simulation results and the quasi-linear theory (e.g., Fatuzzo & Melia 2014; Lynn et al. 2014; Teaca et al. 2014). In these simulations, turbulence is provided with some Fourier spectra or driven as fluctuations based on some algorithms in magnetohydrodynamic fluids. Although these simulations are useful for investigating the physics of stochastic acceleration due to the controllability of the turbulence, it is also important to study particle acceleration in realistic astrophysical turbulence, which is  generated by MHD instabilities, such as the Kelvin–Helmholtz (KH) instability and MRI. The anisotropy of the flow is essential for the growth of these instabilities, and MHD simulations show that the nonlinear growth of these instability generates turbulent magnetic fields that are stronger than the mean magnetic fields (e.g., Hawley et al. 1995; Suzuki & Inutsuka 2009; Zhang et al. 2009). We study particle acceleration under the turbulence induced by such physical instabilities.

Recently, particle in cell (PIC) simulations have been performed to study plasma kinetics in RIAFs (Riquelme et al. 2012, 2015; Hoshino 2013, 2015). These simulations show that turbulent fields are generated by MRI and the magnetic reconnections generate CRs. These PIC simulations need to resolve the gyro-motion of thermal particles, and so practically they have to adopt unrealtistic conditions such as fairly strong shears or weak magnetic fields.

In this paper, we assume that thermal particles are described by MHD and calculate the orbital motions of test particles in the turbulent fields generated by MRI. Currently, only MHD simulations can treat large-scale turbulent fields, which are essential to determine the maximum energy of accelerated particles in accretion flows because the particles with higher energy have larger gyro-radii. Our approach makes it possible to investigate stochastic acceleration with more realistic turbulences and shear flows. However, our treatment cannot take into account the injection process for thermal particles into non-thermal acceleration. In this sense, this study is complementary to the PIC simulations of MRI. Our method is similar to that of Roh et al. (2016), who solve the particle orbits under turbulences generated by Richtmyer–Meshkov instability in supernova remnants (Inoue et al. 2013).

This paper is organized as follows. Section 2 describes the numerical scheme and characteristics of turbulence generated in MHD simulations. The method and results of particle simulations are presented in Section 3. We discuss some features of our simulations in Section 4, and Section 5 is devoted to a summary.

2. MAGNETO-HYDRODYNAMICAL SIMULATION

2.1. Numerical Scheme

First, we perform ideal MHD simulations to obtain turbulent fields generated by MRI. We ignore back reaction by CRs for simplicity. We use the isothermal equation of state and do not solve the energy equation. We use the shearing box approximation without vertical gravity for simplicity (Hawley et al. 1995). The shearing box is a local approximation of differentially rotating plasma. We use the co-rotational frame and ignore the curvature effect. Then, the set of MHD equations is written as

Equation (1)

Equation (2)

Equation (3)

where ${{\boldsymbol{e}}}_{i}$ is the unit vector of the xi direction, ${{\boldsymbol{V}}}_{{\rm{fl}}}$ is the fluid velocity, ${{\boldsymbol{\Omega }}}_{{\rm{K}}}={{\rm{\Omega }}}_{{\rm{K}}}{{\boldsymbol{e}}}_{z}$ is the angular velocity at the box center (${{\rm{\Omega }}}_{{\rm{K}}}$ is the Keplerian angular velocity), T is the time for the MHD simulation, and the other letters have their usual meanings. In the shearing box, the directions $x,\;y,$ and z correspond to $r,\;\phi $, and z in the cylindrical coordinate, respectively. In this paper, we ignore the effect of resistivity, which may affect particle acceleration through the generation of electric fields (see Section 4.3).

The boundary conditions of the shearing box for the magnetic fields are described as

Equation (4)

Equation (5)

Equation (6)

The same conditions are used for the velocity fields and density. The background shear velocity in a shearing box is ${{\boldsymbol{V}}}_{{\rm{shear}}}=-1.5{{\rm{\Omega }}}_{{\rm{K}}}x{{\boldsymbol{e}}}_{y}$, where the factor of 1.5 comes from the index of the radial dependence of angular velocity, ${{\rm{\Omega }}}_{{\rm{K}}}\propto {r}^{-1.5}$. This generates relative velocities in the y direction between the adjacent boxes in the x direction, so that the y coordinate in Equation (4) depends on the MHD time T. As the initial condition for MHD simulations, we set the plasma to be of uniform density ${\rho }_{0}$. The initial plasma has a vertical magnetic field, ${B}_{0}=(0,\;0,\;{B}_{z})$ with ${\beta }_{{\rm{pl}}}={10}^{4}$, where ${\beta }_{{\rm{pl}}}=8\pi {\rho }_{0}{c}_{{\rm{s}}}^{2}/{B}_{0}^{2}$ (${c}_{{\rm{s}}}$ is the sound speed). We perform MHD simulations with the above setup using a modified version of an MHD code used for investigating MRI turbulence in protoplanetary disks (Suzuki & Inutsuka 2009; Suzuki et al. 2010) which adopts a second-order Godunov-CMoCCT scheme.

The most unstable wavelength of MRI, ${\lambda }_{{\rm{cr}}}$, is represented as (e.g., Balbus & Hawley 1998; Sano et al. 2004)

Equation (7)

where ${v}_{{\rm{A,0}}}={B}_{0}/\sqrt{4\pi {\rho }_{0}}$ is the Alfvén velocity and $H=\sqrt{2}{c}_{{\rm{s}}}/{{\rm{\Omega }}}_{{\rm{K}}}$ is the scale height. For the particle simulation, the box size and mesh number are fixed as $({L}_{x}/{N}_{x},\;{L}_{y}/{N}_{y},\;{L}_{z}/{N}_{z})$ = (2H/128, 4H/256, 2H/128), for which ${\lambda }_{{\rm{cr}}}$ is resolved by several grids. We briefly discuss the effect of the box size in the next subsection.

2.2. Characteristics of MRI Turbulence

The nonlinear growth of MRI generates strongly turbulent fields. After several orbital periods, a quasi-steady state is realized. The strength of the turbulent fields is indicated by the α parameter introduced by Shakura & Sunyaev (1973), which is expressed as

Equation (8)

where wxy is the xy component of the stress tensor, ${\boldsymbol{\delta }}{\boldsymbol{V}}$ is the turbulent velocity, ${P}_{0}={\rho }_{0}{c}_{{\rm{s}}}^{2}$ is the initial pressure, and $\langle {\rangle }_{{\rm{v}}}$ denotes the volume average quantities. The turbulent velocity is defined as ${\boldsymbol{\delta }}{\boldsymbol{V}}={{\boldsymbol{V}}}_{{\rm{fl}}}-{{\boldsymbol{V}}}_{{\rm{shear}}}$. Figure 1 shows the evolution of α for two models of different box sizes, $({L}_{x},\;{L}_{y},\;{L}_{z})$ = (2H, 4H, 2H) and (0.8H, 1.6H, 0.8H). We can see that for the model with the larger box, α is almost constant ($\sim 0.02-0.03$) for $T\gtrsim 7\;{T}_{{\rm{rot}}}$, where ${T}_{{\rm{rot}}}=2\pi /{{\rm{\Omega }}}_{{\rm{K}}}$ is the rotation time. On the other hand, α varies between 0.01 and 0.06 for the smaller box. This indicates that the smaller the box size is, the stronger are the fluctuations in the quasi-steady state (Bodo et al. 2008; Jiang et al. 2013). Hereafter, we will use the data set of the larger box model in order to avoid the effect of the temporal fluctuation of turbulent fields. The initial value of ${\beta }_{{\rm{pl}}}$ also affects the saturation level of MRI, but we only treat the data set for ${\beta }_{{\rm{pl}}}={10}^{4}$ as the first step of our study.

Figure 1.

Figure 1. Evolution of α for the turbulent fields generated by MRI. The solid and dotted lines represent models with $({L}_{x},\;{L}_{y},\;{L}_{z})$ = (2H, 4H, 2H) and (0.8H, 1.6H, 0.8H), respectively. The quasi-steady state in the model with the smaller box fluctuates more strongly than that with the larger box.

Standard image High-resolution image

We tabulate the components of the averaged magnetic field $\langle {B}_{i}^{2}{\rangle }_{{\rm{v}}}$, turbulent velocity $\langle \delta {V}_{i}^{2}{\rangle }_{{\rm{v}}}$, and electric field $\langle {E}_{i}^{2}{\rangle }_{{\rm{v}}}$ in Table 1, for which we use MHD data of $T=20\;{T}_{{\rm{rot}}}$. The electric field is computed from the ideal MHD condition:

Equation (9)

These fields satisfy $\langle {B}_{y}^{2}{\rangle }_{{\rm{v}}}$ > $\langle {B}_{x}^{2}{\rangle }_{{\rm{v}}}$ ≳ $\langle {B}_{z}^{2}{\rangle }_{{\rm{v}}}$, $\langle \delta {V}_{y}{\rangle }_{{\rm{v}}}\gg \langle \delta {V}_{x}{\rangle }_{{\rm{v}}}$ ≳ $\langle \delta {V}_{z}{\rangle }_{{\rm{v}}}$, and $\langle {E}_{z}^{2}{\rangle }_{{\rm{v}}}\gtrsim \langle {E}_{x}^{2}{\rangle }_{{\rm{v}}}\gg \langle {E}_{y}^{2}{\rangle }_{{\rm{v}}}$. The magnetic field is stretched by the shear motion, so that By is stronger than the other directions. The shear velocity is faster than turbulent velocity, $\langle {V}_{{\rm{shear}}}^{2}{\rangle }_{{\rm{v}}}\gt \langle \delta {V}_{i}^{2}{\rangle }_{{\rm{v}}}$. The electric field is then mainly generated by the shear motion, so that $\langle {E}_{y}^{2}{\rangle }_{{\rm{v}}}$ is weaker than $\langle {E}_{x}^{2}{\rangle }_{{\rm{v}}}$ and $\langle {E}_{z}^{2}{\rangle }_{{\rm{v}}}$.

Table 1.  Mean Values of Background Fields

$\langle {B}_{x}^{2}{\rangle }_{{\rm{v}}}/(8\pi {P}_{0})$ $6.58\times {10}^{-3}$
$\langle {B}_{y}^{2}{\rangle }_{{\rm{v}}}/(8\pi {P}_{0})$ $3.62\times {10}^{-2}$
$\langle {B}_{z}^{2}{\rangle }_{{\rm{v}}}/(8\pi {P}_{0})$ $3.23\times {10}^{-3}$
$\langle \delta {V}_{x}^{2}{\rangle }_{{\rm{v}}}{/c}_{{\rm{s}}}^{2}$ $1.79\times {10}^{-2}$
$\langle \delta {V}_{y}^{2}{\rangle }_{{\rm{v}}}{/c}_{{\rm{s}}}^{2}$ $2.66\times {10}^{-2}$
$\langle \delta {V}_{z}^{2}{\rangle }_{{\rm{v}}}{/c}_{{\rm{s}}}^{2}$ $1.10\times {10}^{-2}$
$\langle {E}_{x}^{2}{\rangle }_{{\rm{v}}}/(8\pi {P}_{0})$ $2.63\times {10}^{-5}$
$\langle {E}_{y}^{2}{\rangle }_{{\rm{v}}}/(8\pi {P}_{0})$ $5.07\times {10}^{-7}$
$\langle {E}_{z}^{2}{\rangle }_{{\rm{v}}}/(8\pi {P}_{0})$ $5.45\times {10}^{-5}$

Download table as:  ASCIITypeset image

We calculate the power spectra of turbulent magnetic, velocity, and electric fields. First, we take the Fourier transformation of the variables

Equation (10)

Then, we define the one-dimensional power spectrum as

Equation (11)

where we write the definition of the power spectrum of magnetic field for the x direction. The power spectra of the other variables and the other directions are also defined in the same manner. The power spectra of ${\boldsymbol{B}}$, ${\boldsymbol{\delta }}{\boldsymbol{V}}$, and ${\boldsymbol{E}}$ for the x and y directions of the data of $T=20\;{T}_{{\rm{rot}}}$ are shown in Figure 2. The Nyquist wavenumber is ${k}_{{\rm{N}}}\simeq 201/H$ for all of the directions, and then the small-scale turbulence ${k}_{i}\gtrsim {k}_{{\rm{N}}}/4\simeq 50/H$, shown by the vertical dotted line, suffers from numerical dissipation. On the other hand, the large-scale turbulence for ${k}_{i}\lesssim 10/H$ is probably affected by the injection process of turbulence. This scale becomes larger as MRI grows in the nonlinear regime, although it is given by Equation (7) in the linear regime, whose scale is close to the vertical dashed line in Figure 2 (e.g., Sano & Inutsuka 2001). Here, we consider $10\lesssim {k}_{i}H\lesssim 50$ as the inertial range. The shapes of the power spectra in the inertial range are similar for all of the variables. The power spectra have anisotropy. The spectra for the x and z directions are almost the same, which is almost consistent with the Kolmogorov turbulence, ${B}_{{k}_{x}}^{2}\sim {B}_{{k}_{z}}^{2}\propto {k}_{x}^{-5/3}$. On the other hand, ${B}_{{k}_{y}}^{2}\propto {k}_{y}^{-2}$, which is steeper than those for the other directions. Since the magnetic field is almost parallel to the y direction, these properties are consistent with the theory of incompressible MHD turbulence (Goldreich & Sridhar 1995), which is also confirmed by the previous simulations (e.g., Cho & Lazarian 2003). These properties of the turbulent fields are unchanged even for snapshot data of different MHD times. Note that in order to see the waves moving with the background shear, one should change the coordinates and wavenumbers according to Hawley et al. (1995). Although this slightly modifies the inertial range of the power spectra, we do not have to deal with it because the motions of the test particles are calculated in the coordinates of the shearing box (see Section 3.1).

Figure 2.

Figure 2. Power spectra of the data of $T=20\;{T}_{{\rm{rot}}}$. The upper and lower panels show the power spectra for the x and y directions, respectively. The thick solid, thick dashed, and thick dotted lines represent ${B}_{k}^{2}/(8\pi {P}_{0})$, $\delta {V}_{k}^{2}/{c}_{s}^{2}$, and ${c}^{2}{E}_{k}^{2}/(8\pi {P}_{0})$, respectively. The thin solid lines show the spectrum predicted by Goldreich & Sridhar (1995), which is almost consistent with the simulation results for $10\lesssim {kH}\lesssim 50$. The power spectra rapidly decrease for ${kH}\gtrsim 50$ due to numerical dissipation. The vertical dashed and dotted lines show the initial gyro-scales for models A1 and A3, respectively.

Standard image High-resolution image

3. TEST PARTICLE SIMULATION

3.1. Numerical Procedure

We calculate the orbits of CRs as test particles in the turbulence generated in the MHD simulations. The equation of motion for a test particle is

Equation (12)

where ${\boldsymbol{p}}=\gamma m{\boldsymbol{v}}$ and ${\boldsymbol{v}}$ are the momentum and velocity of the test particle, respectively (γ is the Lorentz factor of the particle). The tidal and Coriolis forces are included in ${E}_{{\rm{mod}}}$ and ${B}_{{\rm{mod}}}$, respectively (Hoshino 2013),

Equation (13)

Equation (14)

Note that the tidal and Coriolis forces are much weaker than the electromagnetic force, and they have little effect on the orbits of CRs in our simulations. We assume CRs as protons and neglect radiative and Coulomb losses. We use the Boris method (e.g., Birdsall & Langdon 1991) to integrate this equation, which is often used in PIC simulations.

Since the MHD data are defined only on the grid points, it is necessary to interpolate the data to the particle position. The MHD data consist of ${\boldsymbol{B}}$ and ${{\boldsymbol{V}}}_{{\rm{fl}}}$. First, we compute ${\boldsymbol{B}}$ and ${{\boldsymbol{V}}}_{{\rm{fl}}}$ at the particle position through linear interpolation. Then, we calculate ${\boldsymbol{E}}$ according to Equation (9). This procedure guarantees ${\boldsymbol{E}}\cdot {\boldsymbol{B}}=0$, which is necessary to avoid artificial acceleration due to interpolation (cf., Roh et al. 2016). In reality, ${\boldsymbol{B}}$ and ${\boldsymbol{E}}$ evolve on an MHD timescale. However, as the first step, we use the snapshot MHD data for simplicity.

As the initial conditions, the particle positions and velocities are given such that a uniform and isotropic distribution is realized in the calculation box (hereafter, we call this box the initial box). We set monoenergetic CRs whose gyro radii ${r}_{{\rm{gyro,0}}}$ satisfy

Equation (15)

where ${E}_{p,0}$ is the initial energy of CRs, ${\epsilon }_{{\rm{gyro}}}$ is a parameter of the initial energy of CRs, and ${\rm{\Delta }}x={L}_{x}/{N}_{x}$ is the mesh size. The initial Lorentz factor is then represented as

Equation (16)

The value of ${\gamma }_{0}$ depends on the models of the accretion flows. We tabulate ${\gamma }_{0}$ in Table 2, supposing that the accretion flow has ${c}_{{\rm{s}}}$ ≃ 2.2 × 109 cm s−1, ${{\rm{\Omega }}}_{{\rm{K}}}$ ≃ 4.4 × 10−6 s−1, and ${n}_{0}={\rho }_{0}/m\simeq 9.0\times {10}^{7}\;{\mathrm{cm}}^{-3}$. These values correspond to the self-similar solution for RIAFs (Yuan & Narayan 2014) with $m={10}^{8}$ (black hole mass in unit of M), ${\dot{m}}_{{\rm{BH}}}={10}^{-3}$ (mass accretion rate in unit of the Eddington ratio), r = 30 (disk radius in unit of the Schwarzschild radius), s = 0 (parameter of mass loss rate), and $\alpha =0.03$.4 In reality, the energies of these CRs are so high that they probably escape from the flow in the vertical direction. It is desirable to calculate CRs of much lower energies, such as ${\gamma }_{0}\sim 10-100$. However, the gyro-scale of such low-energy CRs is too small compared to the grid scale of the MHD simulations. These particles do not significantly change their energies in interactions with the large-scale turbulence that is likely to determine the maximum energy of the CRs in accretion flows. To investigate the interaction between CRs and large-scale turbulence, we calculate the orbital motions of the high-energy CRs by prohibiting them from escaping in the vertical direction. Since these CRs are ultrarelativistic, we can write ${E}_{p,0}={p}_{0}c$, where p0 is the initial momentum of CRs.

Table 2.  Model Parameters and Physical Quantities

Model T ${\epsilon }_{{\rm{gyro}}}$ ${C}_{{\rm{esc}}}$ ${\gamma }_{0}$ ${t}_{{\rm{end}}}$ a $\delta t$ a Dxb ${D}_{y}$ b ${D}_{z}$ b q D0c A
A1 $20{T}_{{\rm{rot}}}$ 4 2 $3.4\times {10}^{8}$ 416 50 2.3 17 1.6 2.38 $1.59\times {10}^{-4}$ 0.30
A2 $20{T}_{{\rm{rot}}}$ 1 2 $8.5\times {10}^{7}$ 5628 200 2.4 24 1.6 1.91 $3.17\times {10}^{-5}$ 0.25
A3 $20{T}_{{\rm{rot}}}$ 8 2 $6.8\times {10}^{8}$ 94 40 2.0 16 2.0 2.79 $5.38\times {10}^{-4}$ 0.25
B1 $15{T}_{{\rm{rot}}}$ 4 2 $3.5\times {10}^{8}$ 447 50 2.2 19 1.5 2.38 $1.50\times {10}^{-4}$ 0.31
C1 $25{T}_{{\rm{rot}}}$ 4 2 $3.7\times {10}^{8}$ 438 50 2.2 17 1.6 2.46 $1.45\times {10}^{-4}$ 0.28
X1 $20{T}_{{\rm{rot}}}$ 4 $\infty $ $3.4\times {10}^{8}$ $\infty $ 200 3.1 19 2.0 0.969 $5.83\times {10}^{-5}$ 0.54
X2 $20{T}_{{\rm{rot}}}$ 1 $\infty $ $8.5\times {10}^{7}$ $\infty $ 200 2.5 20 1.5 1.31 $2.88\times {10}^{-5}$ 0.46
X3 $20{T}_{{\rm{rot}}}$ 8 $\infty $ $6.8\times {10}^{8}$ $\infty $ 200 3.5 21 3.0 0.654 $6.44\times {10}^{-5}$ 0.58

Notes.

aIn unit of ${t}_{{\rm{gyro,0}}}$. bIn unit of ${D}_{{\rm{Bohm}}}$. cIn unit of ${\bar{D}}_{p}$.

Download table as:  ASCIITypeset image

Due to Equations (4)–(6), we can continue to calculate the orbits of CRs outside of the initial box. Because of the relative velocities between boxes, we should convert the energies and momenta of CRs crossing the box boundary in the x direction by Lorentz transformation as

Equation (17)

Equation (18)

Equation (19)

where ${\beta }_{{\rm{box}}}=\mp 1.5{{\rm{\Omega }}}_{{\rm{K}}}{L}_{x}/c$ is the relative velocity between boxes (we use the − sign for the CRs crossing $x=+{L}_{x}/2$ boundary, and vice versa) and ${\rm{\Gamma }}={(1-{\beta }_{{\rm{box}}}^{2})}^{-1/2}$. These Γ and ${\beta }_{{\rm{Box}}}$ depend on ${c}_{{\rm{s}}}$ and ${L}_{x}/H$. We get ${\beta }_{{\rm{box}}}\simeq 0.306$ and ${\rm{\Gamma }}-1$ ≃ 5.01 × 10−2 for ${c}_{{\rm{s}}}$ ≃ 2.2 × 109 cm s−1 and ${L}_{x}/H=2$. Since ${\boldsymbol{v}}={c}^{2}{\boldsymbol{p}}/{E}_{p}$, vx and vz are also changed by the transformation. We solve Equation (12) in the rest frame of a box where each CR is located.

This Lorentz transformation causes an artificial acceleration for sufficiently energetic CRs. This acceleration happens when the energetic CRs continuously cross several boxes without changing their directions of motion (see Section 4.1 for details). In order to avoid this, we remove those CRs that have crossed a boundary at $x={x}_{{\rm{esc}}}$ as escape particles. We define ${x}_{{\rm{esc}}}=({C}_{{\rm{esc}}}+0.5){L}_{x}$, where ${C}_{{\rm{esc}}}$ is the escape parameter, and set ${C}_{{\rm{esc}}}=2$ such that the relative Lorentz factor between the escape boundaries at $x=-{x}_{{\rm{esc}}}$ and ${x}_{{\rm{esc}}}$ is less than two. Although the Lorentz contraction may also affect the length scale that CRs travel, we ignore its effect for simplicity.

We note that the boundary condition for CRs described above is different from that for the MHD simulations. The latter use the Galilean transformation, since the fluid velocity is non-relativistic, ${\rm{\Gamma }}-1\ll 1$. We cannot use the Galilean transformation as the boundary condition for CRs because the velocity of CRs after the Galilean transformation could exceed the speed of light, ${v}_{y}^{\prime }={v}_{y}+1.5{{\rm{\Omega }}}_{{\rm{K}}}{L}_{x}\sim c+1.5{{\rm{\Omega }}}_{{\rm{K}}}{L}_{x}\gt c$.

We set the time step as ${\rm{\Delta }}t={\rm{min}}({\rm{\Delta }}{t}_{{\rm{gyro}}},\;{\rm{\Delta }}{t}_{{\rm{cell}}})$, where ${\rm{\Delta }}{t}_{{\rm{gyro}}}={C}_{{\rm{safe}}}{E}_{p,0}/({{ceB}}_{{\rm{max}}})$ and ${\rm{\Delta }}{t}_{{\rm{cell}}}={C}_{{\rm{safe}}}{\rm{\Delta }}x/c$. Note that the velocities of CRs are always almost the speed of light because we focus on the ultrarelativistic particles. We use the maximum value of the magnetic field in the box ${B}_{{\rm{max}}}$ to estimate ${\rm{\Delta }}{t}_{{\rm{gyro}}}$, and set ${C}_{{\rm{safe}}}=0.01$.

3.2. Results

We show the results of the simulations and discuss the evolution of the distribution function. The parameter sets are tabulated in Table 2. The letters A, B, and C represent the different times T of snapshot data, and we use the numbers 1, 2, and 3 to distinguish between the initial energy ${\epsilon }_{{\rm{gyro}}}$. Group X will be discussed in Section 4.2. We use ${N}_{p}={2}^{15}=32768$ CR particles and calculate their orbits until half of CRs escape from the system, the times of which are tabulated in Table 2 in units of the initial gyro-period, ${t}_{{\rm{gyro,0}}}=2\pi {r}_{{\rm{gyro,0}}}/c$. For all the models, the CRs randomly gain or lose their energies through interactions with turbulent fields, and diffuse in both the configuration and momentum spaces.

3.2.1. Lab Frame, Box Frame, and Shear Frame

There are three frames for evaluating the positions and momenta of CRs. One is the rest frame of the initial box where the CRs are initially located (lab frame), another is the rest frame of a box where each CR is located at the evaluation time (box frame), and the other is the rest frame of the MHD fluid element in the mean flow (i.e., the unperturbed flow) in each box (shear frame).

Figure 3 shows the evolution of the energy of a CR in the shear frame (thick solid line) and the box frame (dashed line). When we measure the energy in the box frame, the CR energy jumps due to the Lorentz transformation. In Figure 3, we can see two jumps of energy at $t\sim 240\;{t}_{{\rm{gyro,0}}}$ and $t\sim 243\;{t}_{{\rm{gyro,0}}}$, which coincide with the CR's crossing the box boundary (shown by the dotted lines) in the x direction. The CR position is represented by the thin solid line. On the other hand, the energy measured in the shear frame does not have such jumps but smoothly evolves with time. Since the box boundaries are not special surfaces in nature, we use the shear frame for discussing the evolution of CR energy unless otherwise noted.

Figure 3.

Figure 3. Evolution of the energy of a CR in the shear frame (thick solid line) and the box frame (dashed line) for model A1. The thin solid and dotted lines show the position x of the CR and the box boundary, respectively. The particle energy jumps in the box frame, while it smoothly changes in the shear frame. These jumps coincide with the CR's crossing the box boundaries in the x direction.

Standard image High-resolution image

As mentioned above, we find that CRs randomly gain or lose a small amount of energy through their interaction with turbulent fields. A small fraction of CRs continuously gain (lose) energies, so that they can reach several times higher (lower) energies than their initial energies. Figure 4 shows the long-term evolution of the energies of such CRs in the shear frame. The most energetic CR at $t=400\;{t}_{{\rm{gyro,0}}}$ has about six times higher energy than the average value. This gradual change of particle energy implies that there is no "hot spot," where CRs gain energy efficiently, in the MRI turbulence. Note that the average energy, shown as the dotted line in Figure 4, is gradually increasing. This is consistent with the quasi-linear theory of stochastic acceleration (e.g., Stawarz & Petrosian 2008).

Figure 4.

Figure 4. Long-term evolution of the energy of CRs in the shear frame for model A1. The thick solid line shows the evolution of energy for the most energetic CR at $t=400\;{t}_{{\rm{gyro,0}}}$. The thin solid line shows that for the minimum energy CR. The dotted line shows the average energy of CRs. The most energetic CR has about six times higher energy than the average value.

Standard image High-resolution image

3.2.2. Evolution in Configuration Space

The diffusion coefficient in configuration space is estimated to be

Equation (20)

where $\delta {x}_{i}={x}_{i}(t+\delta t)-{x}_{i}(t)$ is the displacement of a particle (${x}_{i}=x,\;y$, or $\;z$), $\delta t$ is the time span for estimating the diffusion coefficient, and $\langle {\rangle }_{p}$ denotes the average over the CRs. The estimate of the diffusion coefficient is affected by the gyro-motion for too short $\delta t$, while it is affected by excluding the escaped CRs for too long $\delta t$. We calculate ${D}_{{x}_{i}}$ for several values of $\delta t$ and choose suitable $\delta t$ for estimating the diffusion coefficient, which is tabulated in Table 2.

We show the temporal evolution of ${D}_{x},{D}_{y},$ and Dz for model A1 in the upper panel of Figure 5. We normalize ${D}_{{x}_{i}}$ by the Bohm coefficient,

Equation (21)

The diffusion coefficients in configuration space have anisotropy. Since the magnetic field is stretched in the y direction, CRs easily move in the y direction. Thus, ${D}_{y}\gt {D}_{x}\sim {D}_{z}$ is satisfied. The diffusion coefficients are almost constant in time, and so super diffusion is not observed in our simulations (see Xu & Yan 2013; Lazarian & Yan 2014; Roh et al. 2016).

Figure 5.

Figure 5. Diffusion coefficient in configuration space for model A1. The upper panel shows the temporal evolution of Dx (thick solid), Dy (dashed), and Dz (dotted) normalized by the Bohm coefficient. CRs diffuse anisotropically, ${D}_{y}\gt {D}_{x}\gtrsim {D}_{z}$. The lower panel shows the momentum dependence of ${D}_{{x}_{i}}$. The plus, cross, and asterisk points show Dx, Dy, and Dz, respectively. The Bohm coefficients (thin solid lines) are also shown in both panels. The energy dependence of the diffusion coefficients is almost the same as that of the Bohm one.

Standard image High-resolution image

We do not discuss the diffusion coefficient in the direction parallel or perpendicular to the averaged magnetic field. Although we can define the volume-averaged magnetic field $\langle {\boldsymbol{B}}{\rangle }_{{\rm{v}}}$, it does not represent the local (in the scale of gyro radius) direction of the magnetic field due to the strong turbulence.

Dividing CRs into a few momentum bins, we estimate the momentum dependence of ${D}_{{x}_{i}}$. We use 20 momentum bins of equal interval in the logarithmic space for $0.1\leqslant p/{p}_{0}\leqslant 20$. To reduce the statistical fluctuation, we use only those momentum bins which have more than 100 CRs. In the lower panel of Figure 5, the momentum dependence of ${D}_{{x}_{i}}$ is shown at $t=400{t}_{{\rm{gyro,0}}}$. We can see that ${D}_{{x}_{i}}\propto p$, which is the same dependence of the Bohm coefficient. These are common features for all of the models. The diffusion coefficients in configuration space for the other models are tabulated in Table 2. All of the models have similar values of the diffusion coefficients, ${D}_{y}\sim 20{D}_{{\rm{Bohm}}}$ and ${D}_{x}\sim {D}_{z}\sim 2{D}_{{\rm{Bohm}}}$.

3.2.3. Evolution in Momentum Space

We discuss the evolution of the momentum distribution function $f({\boldsymbol{p}})$ of CRs. Figure 6 shows the momentum distribution of each direction ${dN}/{{dp}}_{i}$, normalized as $\int ({dN}/{{dp}}_{i}){{dp}}_{i}=1$ (${p}_{i}={p}_{x},\;{p}_{y},$ or pz). CRs diffuse in all of the directions in momentum space. From the upper panel, it can be seen that the momentum distribution is isotropic in the shear frame. The turbulent fields isotropize the motion of CRs in the shear frame. The dispersion of the momentum distribution monotonically increases with time. The momentum distribution in the box frame is almost the same as that in the shear frame because the relative shear velocity inside of a box is sufficiently smaller than the particle velocities. On the other hand, in the lab frame, the distribution is anisotropic due to the velocity difference between the boxes. The dispersion in the y direction is larger than in the other directions. In this section, we will use the shear frame when discussing the evolution of the distribution function. Note that the isotropy of the momentum distribution does not conflict with the anisotropic diffusion in configuration space. This situation is possible when the timescale of pitch angle scattering is much shorter than the diffusion timescale in configuration space.

Figure 6.

Figure 6. Momentum distribution of each direction for model A1. The thin solid lines show the initial distribution. The solid, dashed, and dotted lines show the final states of ${dN}/{{dp}}_{x},{dN}/{{dp}}_{y}$, and ${dN}/{{dp}}_{z}$, respectively. The upper and lower panels show the distribution in the shear frame and the lab frame, respectively. The momentum distribution is isotropic in the shear frame while it is anisotropic in the lab frame.

Standard image High-resolution image

Since the momentum distribution is isotropic, we can write $f({\boldsymbol{p}})=f(p)$, where $p=| {\boldsymbol{p}}| $. Below, we show that the evolution of f(p) obtained by our simulations is well described by the diffusion equation in momentum space:

Equation (22)

where Dp is the diffusion coefficient in momentum space, ${t}_{{\rm{esc}}}={({C}_{{\rm{esc}}}+0.5)}^{2}{L}_{x}^{2}/(2{D}_{x})$ is the escape time, and ${\dot{f}}_{{\rm{inj}}}={N}_{p}\delta (t)\delta (p-{p}_{0})$ is the injection term. We write $p={E}_{p}/c=\gamma {mc}$ because CRs are ultrarelativistic. Equation (22) is expected to describe the evolution of the distribution function when the fractional energy change per scattering is small.

To calculate the evolution of f(p) using Equation (22), we estimate the diffusion coefficient in momentum space Dp from our simulation results. The diffusion coefficient is likely to be represented as

Equation (23)

where $\delta p=p(t+\delta t)-p(t)$ and A is the numerical factor. We estimate Dp using the momentum bins defined in the previous subsection. Although $A=1/6$ when we consider the isotropic random walk in three-dimensional space, the values of A are not obvious due to the dependence of Dp on p. In fact, the values of A are different among the models (see Table 2). We plot $\langle \delta {p}^{2}{\rangle }_{p}/\delta t$ for model A1 in Figure 7. We can fit $\langle \delta {p}^{2}{\rangle }_{p}/\delta t$ by a power-law function, $\langle \delta {p}^{2}{\rangle }_{p}/\delta t\propto {(p/{p}_{0})}^{q}$, for all of the models. The resultant q is tabulated in Table 2.

Figure 7.

Figure 7. Momentum dependence of $\langle \delta {p}^{2}{\rangle }_{p}/\delta t$ for model A1. The points show the simulation results, which can be fit by a power-law function, as shown by the solid line.

Standard image High-resolution image

We compute A as follows. If the temporal evolution of f(p) is described by Equation (22), then the temporal evolution of the averaged momentum $\langle \dot{p}{\rangle }_{p}$ is written as (see Becker et al. 2006)

Equation (24)

where we use Equation (22) and ignore the terms ${\dot{f}}_{{\rm{inj}}}$ and $f(p)/{t}_{{\rm{esc}}}$. Integrating this equation by part twice, we obtain

Equation (25)

Writing ${D}_{p}={D}_{0}{(p/{p}_{0})}^{q}$, ${dN}/{dp}=4\pi {p}^{2}f$, and $p/{p}_{0}=\xi $, we obtain

Equation (26)

We compute D0 taking the time average of this equation over the whole calculation time. We tabulate the values of D0 and A in Table 2, where D0 is normalized by $\bar{{D}_{p}}={p}_{0}^{2}/{t}_{{\rm{gyro,0}}}$. Since $\bar{{D}_{p}}/{D}_{0}={t}_{{\rm{accel}}}/{t}_{{\rm{gyro,0}}}$, where we define ${t}_{{\rm{accel}}}={p}_{0}^{2}/{D}_{0}$, we find that the acceleration time is about ${10}^{3}\mbox{--}{10}^{4}$ times longer than the gyro-period in our simulations.

We solve Equation (22) with Dp obtained using the above procedure, and compare the distribution functions calculated by Equation (22) to those obtained by the particle simulations. In Figure 8, we plot f(p) of $t=100\;{t}_{{\rm{gyro,0}}}$ and $t=400\;{t}_{{\rm{gyro,0}}}$ calculated by both the simulation and the diffusion equation for model A1. We find that the distribution functions of the diffusion equation are always in agreement with those of the particle simulation. We confirm that the diffusion equation reproduces the evolution of the distribution function in all of the models. We can conclude that the stochastic acceleration inside the accretion flows can be described by Equation (22).

Figure 8.

Figure 8. Evolution of the momentum distribution functions obtained using the particle simulation (points) and the diffusion equation (solid lines) for model A1. The results of the particle simulation are well described by the diffusion equation in momentum space.

Standard image High-resolution image

We also calculate the orbits of CRs with snapshots in different times, $T=15\;{T}_{{\rm{rot}}}$ (model B1) and $T=25\;{T}_{{\rm{rot}}}$ (model C1). The differences of the values in Table 2 among models A1, B1, and C1 are less than 15%, which indicates that our results are almost unchanged by the temporal fluctuation of the turbulent fields in our current setup of the particle simulations. However, we should note that ${t}_{{\rm{gyro,0}}}\simeq 6.36\times {10}^{-3}\;{T}_{{\rm{rot}}}$ in model A1, and then the calculation time of CR orbits is longer than the time step of the MHD simulation. Thus, in reality, CRs should move in dynamically evolving turbulence. Although the current simulations with the snapshots of different times provide almost the same results, it is unclear how much the dynamically evolving fields affect the orbits of CRs. This effect should be investigated as the next step in a separate paper.

Finally, we briefly mention the scales of turbulence interacting with CRs. We show the initial gyro-scales of CRs as the vertical dashed and dotted lines for models A1 and A3, respectively, in Figure 2. Except for model A3, the initial gyro-scales are in the dissipation scale of the MHD simulation. Although the turbulence in the dissipation scale with steep spectra is artificially realized, the stochastic acceleration of CRs through interaction with such turbulence is physical. Accelerated CRs can interact with the turbulence in the inertial range due to their larger gyro-scales. However, highly accelerated CRs suffer from artificial acceleration due to the Lorentz transformation (see Section 4.1). Thus, it is difficult to keep a large fraction of CRs in the inertial range. We need MHD data with much higher resolution to observe the CRs that continue to interact with the realistic turbulent fields.

4. DISCUSSION

4.1. Runaway Particles

As briefly discussed in Section 3.1, sufficiently energetic CRs in the shearing box simulation become runaway particles. When CRs of ${p}_{y}\gt 0$ (${p}_{y}\lt 0$) cross the $x=+{L}_{x}/2$ ($x=-{L}_{x}/2$) boundary, their energies in the box frame increase. If their gyro radii are large enough to cross several boxes without changing the direction of motion, the CRs continuously gain energies.

To observe the behavior of the runaway particles, we perform a simulation without an escape boundary. Figure 9 shows the temporal evolution of the energy for a runaway particle in the box frame. It is seen that the energy of the runaway particle increases stepwise every time it crosses the box boundaries in the x direction. We find that the runaway particle has an almost constant box crossing time, ${\rm{\Delta }}t={L}_{x}/{v}_{x}\simeq {L}_{x}/({\beta }_{x}c)$. The energy gain per crossing boundary is ${\rm{\Delta }}{E}_{p}={\rm{\Gamma }}({E}_{p}+{\beta }_{{\rm{Box}}}{{cp}}_{y})-{E}_{p}=\{{\rm{\Gamma }}(1+{\beta }_{{\rm{box}}}{\beta }_{y})-1\}{E}_{p}$. Thus,

Equation (27)

Solving this equation, we obtain ${E}_{p}\propto \mathrm{exp}(t/{t}_{{\rm{grow}}})$, where ${t}_{{\rm{grow}}}\simeq \{{\rm{\Gamma }}(1+{\beta }_{{\rm{box}}}{\beta }_{y})-1\}{\beta }_{x}{L}_{x}/c\simeq 0.30{T}_{{\rm{rot}}}$. Here, we use ${\beta }_{{\rm{box}}}\simeq 0.306$, ${\beta }_{x}\simeq 0.31c$, and ${\beta }_{y}\simeq 0.92c$, which are obtained using the simulation result. This solution is consistent with the temporal evolution of the energy of the runaway particle, as shown in Figure 9. One should care about such runaway particles when using the shearing box approximation without the escape boundaries. A more realistic global simulation without a shearing box approximation does not suffer from this kind of unphysical effect. Particle simulations with global turbulent fields are desirable to obtain a solid conclusion, although it is computationally too expensive to perform with the required resolution.

Figure 9.

Figure 9. Temporal evolution of the energy of a runaway particle in the box frame (solid line), which is compared to the analytic estimate (dotted line). The energy of the runaway particle grows exponentially, and the energy gain rate is consistent with the analytic estimate.

Standard image High-resolution image

4.2. Physical Interpretation of the Diffusion Coefficient

In this subsection, we discuss the physical interpretation of Dp obtained in our simulations. As seen below, we find that the shear acceleration works efficiently, although the origin of Dp is not fully understood.

The shear acceleration is expected to be efficient in accretion flows (e.g., Katz 1991; Subramanian et al. 1999). The shear motion seems to be important in our simulation because the electric field is mainly produced by shear motion. To understand its relevance clearly, we perform particle simulations assuming ${V}_{{\rm{shear}}}=0$ and ${\beta }_{{\rm{box}}}=0$. In this treatment, we do not deal with the Lorentz transformation for the orbital evolutions of CRs (Equations (17) and (18)), and the electric field induced by shear motion vanishes. Since the runaway particle does not appear due to ${\beta }_{{\rm{box}}}=0$, we set ${C}_{{\rm{esc}}}=\infty $. As a result, we find that the evolution of the distribution function is described by Equation (22), and Dp and ${D}_{{x}_{i}}$ are obtained in the same manner as in Section 3.2. The resultant values are tabulated in Table 2 as model group X. We can see that both q and D0 for group X are lower than those for group A. The differences of q and D0 between groups X and A are larger since the initial gyro radius is larger. This implies that shear acceleration works efficiently for particles with larger gyro radii and that it is a dominant process of particle acceleration in model A3.

For the shearing box, Dp by the shear acceleration is analytically estimated as (see Earl et al. 1988; Rieger & Duffy 2004; Ohira 2013)

Equation (28)

where τ is the collision time. Since we found that the diffusion coefficient in configuration space is proportional to the Bohm coefficient, we may assume

Equation (, 29)

where $\eta ={D}_{x}/{D}_{{\rm{Bohm}}}$ and ${v}_{x}^{2}={c}^{2}/3$. Then, we can write

Equation (30)

We find ${D}_{p,\mathrm{shear}}\propto {p}^{3}$ and ${D}_{p,\mathrm{shear}}/\bar{{D}_{p}}\simeq 6.1\times {10}^{-4}$ for $p={p}_{0}$, using the parameter set of model A3 ($\eta \simeq 2.0$ and ${t}_{{\rm{gyro,0}}}/{T}_{{\rm{rot}}}\simeq 1.3\times {10}^{-2}$). The momentum dependence and value of this estimate are almost consistent with Dp of the simulation result, which also shows that the shear acceleration is the dominant process for model A3.

Then, a remaining problem is the physical origin of Dp in model X, in which CRs are accelerated by the electric field induced only by the turbulent fluid motion with the velocity ${\boldsymbol{\delta }}{\boldsymbol{V}}$. According to the quasi-linear theory of gyro-resonance by isotropic Alfvén modes, the power index of Dp is the same as the index of the power spectrum of the turbulent magnetic field (e.g., Stawarz & Petrosian 2008), in which $q=5/3$ is expected for the Kolmogorov turbulence. For our turbulence, the power spectra are steeper for smaller scales (see Figure 2), and so this theory would predict higher values of q for lower p0 models. Model group A shows the opposite trend, while model group X is qualitatively consistent with this feature. However, quantitatively, $q\sim 4$ for model X1 (large k due to low ${E}_{p,0}$) and $q\sim 5/3$ for model X3 (small k due to high ${E}_{p,0}$) are expected in this theory, which are far from the particle simulation results.

Our turbulence has anisotropic power spectra (see Figure 2). Yan & Lazarian (2002) investigate the scattering frequency of such anisotropic MHD turbulence and show that the scattering frequency by the Alfvén modes is sufficiently suppressed compared to the case of isotropic turbulences. Instead, gyro-resonance by fast modes plays a crucial role in scattering CRs under MHD turbulence. The fast modes are almost isotropic and the index of the power spectrum is 1.5 (Cho & Lazarian 2003), which leads to q = 1.5 (see Cho & Lazarian 2006). Model X2 has a value close to q = 1.5, but it is unlikely that the fast modes have substantial power in the MRI turbulence because velocity fluctuations are sub-sonic in our turbulent field.

Lynn et al. (2014) claim that for sub-sonic MHD turbulence, the fast modes have only a small fraction of the turbulent power and slow modes dominate over the other modes in the acceleration of relativistic CRs. They consider the mirror force acting on CRs, and analytically derive ${D}_{p}\propto {p}^{2}$ for stochastic acceleration in the weakly compressible turbulence. In this case, q should always be equal to 2, which is often called the hard-sphere model, but this theory does not seem to be applicable to our model since the resultant q of our simulations is not close to 2 for both groups X and A.

4.3. Plasma Properties in RIAFs

As the first step of our study, we use the shearing box approximation for RIAFs, which was also used in the previous studies (e.g., Sharma et al. 2006; Hoshino 2013). However, the shearing box approximation has a few points which are at odds with RIAFs. First, the local approximation is not good for RIAFs because the aspect ratio, H/r, is close to unity due to the the high temperature. The centrifugal balance is also violated due to the strong pressure gradient force in RIAFs. The angular velocity is usually sub-Keplerian, which seems to weaken the shear acceleration. In addition, the radial advection of heat is important for the global structure of RIAFs. The high temperature makes viscous stress strong under the α prescription, which makes the radial velocity fast compared to the standard disk (e.g., Narayan & Yi 1994; Kimura et al. 2014). The global simulations can consistently take into account these effects.

Magnetic reconnection is supposed to play an important role in the saturation of MRI turbulence (e.g., Sano & Inutsuka 2001). Recent PIC simulations also suggest that dissipation by the reconnection of stretched magnetic fields is important for both the saturation of MRI turbulence and the generation of CRs (Hoshino 2015). However, since we solve a set of ideal MHD equations in which magnetic reconnection occurs due to numerical resistivity, we ignore the electric field generated by finite resistivity in the simulation of CRs. In reality, the electric field generated at the reconnection site is expected to enhance particle acceleration in MRI turbulence. To include this effect, the MHD simulation that includes the resistivity model based on collisionless plasma physics is required, which is beyond the scope of the present paper.

The effect of back reaction by CRs on the nonlinear evolution for MRI is not well understood. Kuwabara & Ko (2015) analyze the linear growth of MRI including the effect of CR diffusion. They show that CR diffusion barely affects the linear growth of MRI in shearing box calculations, even if the CR pressure is comparable to the thermal pressure. However, the effect of CRs on the saturation level of MRI turbulence has not yet been studied. In the case close to ideal MHD, the saturation level is supposed to be determined by the parasitic instability of the KH modes that induce the dissipation by magnetic reconnection (Goodman & Xu 1994; Sano & Inutsuka 2001; Sano et al. 2004; Pessah 2010). Since the growth rate of KH instability is modified by CRs if they have comparable energy density to the thermal particles (Suzuki et al. 2014), CRs probably affect the saturation level of MRI turbulence. To investigate this, three-dimensional MHD simulations including back reaction by CRs are necessary (see Bai et al. 2015).

Although we assume that the thermal component is described by the MHD formalism, it is unclear whether or not collisionless plasma is described by MHD. As an alternative method of describing the collisionless plasma in accretion flows, Quataert et al. (2002) used the anisotropic MHD formalism where the plasma has anisotropic pressure. In this formalism, as the magnetic field becomes stronger, the pressure anisotropy increases, which suppresses the growth of MRI (Sharma et al. 2006). In reality, the thermal particles are expected to be isotropized by some plasma instabilities, such as the mirror mode, so that the pressure anisotropy decreases. This allows MRI to grow in the nonlinear regime, which has been confirmed by recent PIC simulations with the shearing box approximation (Riquelme et al. 2012; Hoshino 2013, 2015). However, as noted in Section 1, there is the scale gap between the gyro-scales of electrons and the scale heights of accretion flows. The hybrid simulations in which we treat protons as particles and electrons as a fluid may be useful for investigating larger-scale turbulences with kinetic effects (see Shirakawa & Hoshino 2014), but it is currently difficult to bridge the scale gap completely.

5. SUMMARY

We have investigated the stochastic particle acceleration process in accretion flows. The CRs are likely to exist inside accretion flows when the mass accretion rate is sufficiently lower than the Eddington rate (e.g., Kimura et al. 2014). The existence of CRs is expected from the theoretical estimate (e.g., Kimura et al. 2015), numerical simulations (e.g., Hoshino 2015), and modeling of the observed photon spectra (e.g., Nemmen et al. 2014), although gamma rays from accretion flows have not yet been detected (Wojaczynski et al. 2015).

We have calculated the orbits of CRs as test particles in turbulent fields generated by MRI with the shearing box approximation. Within this approximation, when a CR particle crosses the box boundary in the x direction, it changes energy as measured in the box frame. Sufficiently energetic CRs that can cross the box boundary in the x direction many times without changing the direction of motion tend to increase their energies in a runaway manner. We have avoided this artificial acceleration by introducing escape boundaries: the CRs crossing these boundaries are allowed to escape from the system. From our simulations, we have found that the energies of CRs randomly change. Some lucky CRs continuously increase their energies, and reach energies which are several times higher than their initial energies after $t\sim 400{t}_{{\rm{gyro,0}}}$. The CRs anisotropically diffuse in the configuration space. Since the magnetic field is stretched mainly in the y direction, the diffusion coefficients satisfy ${D}_{y}\gt {D}_{x}\sim {D}_{z}$. The energy dependence of ${D}_{{x}_{i}}$ is consistent with the Bohm coefficient, and we can approximately write ${D}_{y}\sim 20{D}_{{\rm{Bohm}}}$ and ${D}_{x}\sim {D}_{z}\sim 2{D}_{{\rm{Bohm}}}$.

We have also shown that the evolution of the distribution function is described by the diffusion equation in momentum space (Equation (22)). The diffusion coefficient in momentum space can be fit by a power-law form, ${D}_{p}={D}_{0}{(p/{p}_{0})}^{q}$. Both D0 and q depend on the initial momentum p0. Although the shear acceleration is efficient for sufficiently energetic particles, quantitative understanding of the acceleration mechanism in our simulation is not complete. To apply our simulation results to astrophysical objects, e.g., calculating the spectrum of non-thermal protons and predicting the neutrino and gamma-ray fluxes more precisely than those in Kimura et al. (2015), we need to understand the acceleration process and the dependence of Dp on physical quantities.

We thank Masahiro Hoshino, Jun Kakuwa, Yutaka Ohira, Ryo Yamazaki, Daisuke Nakauchi, and Sanemichi Takahashi for fruitful discussions. This work is partly supported by JST grant "Building of Consortia for the Development of Human Resources in Science and Technology" (SSK and KT) and JSPS Grants-in-Aid for Scientific Research 15H05437 (KT).

Footnotes

  • The definition of the scale height in our paper, $H=\sqrt{2}{c}_{{\rm{s}}}/{{\rm{\Omega }}}_{{\rm{K}}}$, is different from that in Yuan & Narayan (2014), ${H}_{{\rm{YN}}}={c}_{{\rm{s}}}/{{\rm{\Omega }}}_{{\rm{K}}}$, so that the value of the density is slightly modified. In addition, we assume ${\rm{\Omega }}={{\rm{\Omega }}}_{{\rm{K}}}$ in our paper, although this is not a good approximation for RIAFs (see Section 4.3).

Please wait… references are loading.
10.3847/0004-637X/822/2/88