Electron Preacceleration in Weak Quasi-perpendicular Shocks in High-beta Intracluster Medium

, , and

Published 2019 May 7 © 2019. The American Astronomical Society. All rights reserved.
, , Citation Hyesung Kang et al 2019 ApJ 876 79 DOI 10.3847/1538-4357/ab16d1

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/876/1/79

Abstract

Giant radio relics in the outskirts of galaxy clusters are known to be lit up by the relativistic electrons produced via diffusive shock acceleration (DSA) in shocks with low sonic Mach numbers, Ms ≲ 3. The particle acceleration at these collisionless shocks critically depends on the kinetic plasma processes that govern the injection to DSA. Here, we study the preacceleration of suprathermal electrons in weak, quasi-perpendicular (Q) shocks in the hot, high-β (β = Pgas/PB) intracluster medium (ICM) through two-dimensional particle-in-cell simulations. Guo et al. showed that, in high-β Q-shocks, some of the incoming electrons could be reflected upstream and gain energy via shock drift acceleration (SDA). The temperature anisotropy due to the SDA-energized electrons then induces the electron firehose instability (EFI), and oblique waves are generated, leading to a Fermi-like process and multiple cycles of SDA in the preshock region. We find that such electron preacceleration is effective only in shocks above a critical Mach number ${M}_{\mathrm{ef}}^{* }\approx 2.3$. This means that, in ICM plasmas, Q-shocks with Ms ≲ 2.3 may not efficiently accelerate electrons. We also find that, even in Q-shocks with Ms ≳ 2.3, electrons may not reach high enough energies to be injected to the full Fermi-I process of DSA, because long-wavelength waves are not developed via the EFI alone. Our results indicate that additional electron preaccelerations are required for DSA in ICM shocks, and the presence of fossil relativistic electrons in the shock upstream region may be necessary to explain observed radio relics.

Export citation and abstract BibTeX RIS

1. Introduction

Weak shocks with low sonic Mach numbers, Ms ≲ 3, form in the hot intracluster medium (ICM) during major merges of galaxy clusters (e.g., Gabici & Blasi 2003; Ryu et al. 2003; Ha et al. 2018). Radiative signatures of those merger shocks have been detected in X-ray and radio observations (e.g., Markevitch & Vikhlinin 2007; van Weeren et al. 2010; Bruggen et al. 2012; Brunetti & Jones 2014). In the case of the so-called radio relics, the radio emission has been interpreted as synchrotron radiation from the relativistic electrons accelerated via diffusive shock acceleration (DSA) in the shocks. Hence, the sonic Mach numbers of relic shocks, Mradio (radio Mach number), have been inferred from the radio spectral index (e.g., van Weeren et al. 2010, 2016), based on the DSA test-particle power-law energy spectrum (e.g., Bell 1978; Blandford & Ostriker 1978; Drury 1983). In X-ray observations, the sonic Mach numbers, MX (X-ray Mach number), have been estimated for merger-driven shocks, using the discontinuities in temperature or surface brightness (e.g., Markevitch et al. 2002; Markevitch & Vikhlinin 2007).

While Mradio and MX are expected to match, Mradio has been estimated to be larger than MX in some radio relics (e.g., Akamatsu & Kawahara 2013). In the case of the Toothbrush radio relic in merging cluster 1RXS J060303.3, for instance, van Weeren et al. (2016) estimated that Mradio ≈ 2.8 while MX ≈ 1.2–1.5. In the so-called reacceleration model, weak shocks with ∼MX are presumed to sweep through fossil electrons with power-law energy spectrum, Nfossil ∝ γp (γ is the Lorentz factor), and then the radio spectra with observed spectral indices, αsh = (p − 1)/2, are supposed to be generated (e.g., Kang 2016a, 2016b). This model may explain the discrepancy between Mradio and MX in some cases. However, it may not be realistic to assume the presence of fossil electrons with flat power-law spectra up to γ ∼ 104 over length scales of 400–500 kpc, because such high-energy electrons cool with timescales of ∼100 Myr (Kang et al. 2017). On the other hand, with mock X-ray and radio observations of radio relics using simulated clusters, Hong et al. (2015) argued that the surfaces of merger shocks are highly inhomogeneous in terms of Ms (see also Ha et al. 2018), and X-ray observations preferentially pick up the parts with lower Ms (higher shock energy flux) while radio emissions manifest the parts with higher Ms (higher electron acceleration). As a result, MX could be be smaller than Mradio. However, the true origin of this discrepancy has yet to be understood.

Hence, for the full description of radio relics, it is necessary to first understand shocks in the ICM. They are collisionless shocks, as in other astrophysical environments (e.g., Brunetti & Jones 2014). The physics of collisionless shocks involves complex kinetic plasma processes well beyond the MHD Rankine–Hugoniot jump condition. DSA, for instance, depends on various shock parameters, including the sonic Mach number, Ms, the plasma beta, β = Pgas/PB (the ratio of thermal to magnetic pressures), and the obliquity angle between the upstream background magnetic field direction and the shock normal, θBn (see Balogh & Truemann 2013).

In general, collisionless shocks can be classified by the obliquity angle as quasi-parallel (Q, hereafter) shocks with θBn ≲ 45° and quasi-perpendicular (Q, hereafter) shocks with θBn ≳ 45°. In situ observations of Earth's bow shock indicate that protons are effectively accelerated at the Q-portion, while electrons are energized preferentially in the Q-configuration (e.g., Gosling et al. 1989). In such shocks, one of the key processes for DSA is particle injection, which involves the reflection of particles at the shock ramp, the excitation of electromagnetic waves/turbulences by the reflected particles, and the energization of particles through ensuing wave–particle interactions (e.g., Treumann & Jaroschek 2008; Treumann 2009). Because the thickness of the shock transition zone is on the order of the gyroradius of postshock thermal ions, both ions and electrons need to be preaccelerated to suprathermal momenta greater than a few times the momentum of thermal ions, ${p}_{\mathrm{th},i}$, in order to diffuse across the shock transition layer and fully participate in the first-order Fermi (Fermi-I, hereafter) process of DSA (e.g., Kang et al. 2002; Caprioli et al. 2015). Here, ${p}_{\mathrm{th},i}=\sqrt{2{m}_{i}{k}_{{\rm{B}}}{T}_{i2}}$, where Ti2 is the postshock ion temperature and kB is the Boltzmann constant. Hereafter, the subscripts 1 and 2 denote the preshock and postshock quantities, respectively.

Kinetic processes in collisionless shocks can be studied through, for instance, particle-in-cell (PIC) and hybrid plasma simulations (e.g., Caprioli & Spitkovsky 2014a; Guo et al. 2014a, 2014b; Park et al. 2015). Previous studies have mostly focused on shocks in β ≲ 1 plasmas, where the Aflvén Mach number MA is about the same as Ms (${M}_{{\rm{A}}}\approx \sqrt{\beta }{M}_{{\rm{s}}}$), investigating shocks in solar wind and the interstellar medium (ISM) (see also Treumann (2009) and references therein). If plasmas have very low-β (sometimes referred to as cold plasmas), even the thermal motions of particles can be neglected. In hot ICM plasmas on the other hand, β ∼ 100 (e.g., Ryu et al. 2008; Porter et al. 2015), and shocks have low sonic Mach numbers of Ms ≲ 3 but relatively high Alfvén Mach numbers up to MA ≈ 30. In such shocks, kinetic processes are expected to operate differently from low-β shocks.

Recently, we investigated proton acceleration in weak (Ms ≈ 2–4) "Q-shocks" in high-β (β = 30–100) ICM plasmas through one-dimensional (1D) and two-dimensional (2D) PIC simulations (Ha et al. 2018; hereafter Paper I). The main findings can be recapitulated as follows. (1) Q-shocks with Ms ≳ 2.3 develop overshoot/undershoot oscillations in their structures and undergo quasi-cyclic reformation, leading to a significant amount of incoming protons being reflected at the shock. The backstreaming ions excite resonant and nonresonant waves in the foreshock region, leading to the generation of suprathermal protons that can be injected into the Fermi-I process. (2) Q-shocks with Ms ≲ 2.3, on the other hand, have relatively smooth and steady structures. The development of suprathermal population is negligible in these shocks. (3) In Q-shocks, a substantial fraction of incoming ions are reflected and gain energy via shock drift acceleration (SDA), but the energized ions advect downstream along with the background magnetic field after about one gyromotion without being injected to the Fermi-I acceleration. (4) For the description of shock dynamics and particle acceleration in high-β plasmas, the sonic Mach number is a more relevant parameter than the Alfvén Mach number, because the reflection of particles is mostly controlled by Ms.

As a sequel, in this work, we explore the electron preacceleration in low Mach number, Q-shocks in high-β ICM plasmas. Such shocks are thought be the agents of radio relics in merging clusters. Previously, the pre-energization of thermal electrons at collisionless shocks (i.e., the injection problem), which involves kinetic processes such as the excitation of waves via microinstabilities and wave–particle interactions, was studied through PIC simulations (e.g., Amano & Hoshino 2009; Matsukiyo et al. 2011; Riquelme & Spitkovsky 2011; Guo et al. 2014a, 2014b; Park et al. 2015, and references therein). For instance, Amano & Hoshino (2009) showed that, in high-MA Q-shocks with β ∼ 1, strong electrostatic waves are excited by Buneman instability and confine electrons in the shock foot region, where electrons gain energy by drifting along the motional electric field—a phenomenon known as shock surfing acceleration (SSA). On the other hand, Riquelme & Spitkovsky (2011) found that, in Q-shocks with ${M}_{{\rm{A}}}\,\lesssim \,{({m}_{i}/{m}_{e})}^{1/2}$ (where mi/me is the ion-to-electron mass ratio) and β ∼ 1, the growth of oblique whistler waves in the shock foot by modified two-stream instabilities (MTSIs) may play important roles in confining and pre-energizing electrons.

Through 1D PIC simulations, Matsukiyo et al. (2011) showed that, in weak shocks with fast Mach numbers, Mf ≈ 2–3 and β ≈ 3, a fraction of incoming electrons are accelerated and reflected through SDA, and these form a suprathermal population. This was confirmed by Guo et al. (2014a) (hereafter GSN14a) who carried out 2D PIC simulations of Ms = 3, Q-shocks in plasmas with β = 6–200. In particular, GSN14a presented the "relativistic SDA theory" for oblique shocks, which can be briefly summarized as follows. The incoming electrons that satisfy certain criteria (i.e., those with pitch angles larger than the loss-cone angle) are reflected and gain energy through SDA at the shock ramp. The energized electrons backstream along the background magnetic field lines with small pitch angles, generating the temperature anisotropy of ${T}_{e\parallel }\gt {T}_{e\perp }$. Guo et al. (2014b) (hereafter GSN14b) then showed that the "electron firehose instability" (EFI) is induced by the temperature anisotropy and oblique waves are excited (Gary & Nishimura 2003). The electrons are scattered back and forth between magnetic mirrors at the shock ramp and self-generated upstream waves (a Fermi-I type process), being further accelerated mostly through SDA. At this stage, the electrons are still suprathermal and do not have sufficient energies to diffuse downstream of the shock; instead, they stay upstream of the shock ramp. The authors referred to this as a "Fermi-like process," as opposed to the full, bona fide Fermi-I process. GSN14a also pointed out that SSA does not operate in weak ICM shocks, for two reasons: (1) the Buneman instability is suppressed in hot plasmas; and (2) in high-β shocks, the preacceleration via SDA dominates over the energization through interactions with the oblique whistler waves generated via MTSIs in the shock foot.

For electron preacceleration in weak ICM shocks, however, there are still issues to be further addressed. Most of all, there should be a critical Mach number, below which the preacceleration is not efficient. Even though electrons are pre-energized at shocks with Ms ≈ 3, as shown in GSN14a and GSN14b, it is not clear whether they could be further accelerated by the full Fermi-I process of DSA. We will investigate those issues in this paper, using 2D PIC simulations.

The paper is organized as follows. Section 2 includes the descriptions of our simulations, along with the definitions of various parameters involved. In Section 3, we give a brief review of the background physics of Q-shocks, in order to facilitate the understanding of our simulation results in the subsequent section. In Section 4, we present shock structures and electron preacceleration in simulations, and examine the dependence of our findings on various shock parameters. A brief summary is given in Section 5.

2. Numerics

Simulations were performed using TRISTAN-MP, a parallelized electromagnetic PIC code (Buneman 1993; Spitkovsky 2005). The geometry is 2D planar, while all three components of particle velocity and electromagnetic fields are followed. The details of our simulation setups can be found in Paper I, but some basic definitions of parameters and features are described below, in order to make this paper self-contained. In Paper I, we used the variable v to represent flow velocities—for example, v0 and vsh. Here, however, we use the variable u for "flow" velocities, while v is reserved for "particle" velocities.

Plasmas, which are composed of ions and electrons of Maxwellian distributions, move with the bulk velocity ${{\boldsymbol{u}}}_{0}=-{u}_{0}\hat{{\boldsymbol{x}}}$ toward a reflecting wall at the leftmost boundary (x = 0), and a shock forms and propagates toward the $+\hat{{\boldsymbol{x}}}$ direction. Hence, simulations are performed in the rest frame of the shock downstream flow. For the given preshock ion temperature, Ti, the flow Mach number, M0, is related to the upstream bulk velocity as

Equation (1)

where cs1 is the sound speed in the upstream medium and Γ = 5/3 is the adiabatic index. Thermal equilibrium is assumed for incoming plasmas, i.e., Ti1 = Te1, where Te1 is the preshock electron temperature. In typical PIC simulations, because of severe requirements for computational resources, reduced ion-to-electron mass ratios, mi/me < 1836, are assumed. Here, we consider the mass ratio of mi/me = 100 and 400; electrons have the rest mass of me = 511 keV/c2, while "ions" have reduced masses emulating the proton population. In the limit of high β, the upstream flow speed in the shock rest frame can be expressed as ${u}_{\mathrm{sh}}\approx {u}_{0}\cdot r/(r-1)$, where $r=({\rm{\Gamma }}+1)/({\rm{\Gamma }}-1+2/{M}_{s}^{2})$ is the shock compression ratio, and the sonic Mach number, Ms, of the induced shock is given as

Equation (2)

The magnetic field carried by incoming plasmas, B0, lies in the xy plane, and the angle between B0 and the shock normal direction is the obliquity angle θBn, as defined in the introduction. The initial electric field in the flow frame is zero everywhere, but the motional electric field, E0 = $-{{\boldsymbol{u}}}_{0}/c$ × B0, is induced along the $+\hat{{\boldsymbol{z}}}$ direction, where c is the speed of light. The strength of B0 is parameterized by β as

Equation (3)

where MA ≡ ush/uA is the Alfvén Mach number of the shock. Here, ${u}_{{\rm{A}}}={B}_{0}/\sqrt{4\pi {{nm}}_{i}}$ is the Alfvén speed, and n = ni = ne are the number densities of incoming ions and electrons. We consider β = 50 and 100, along with ${k}_{{\rm{B}}}{T}_{1}={k}_{{\rm{B}}}{T}_{i1}={k}_{{\rm{B}}}{T}_{e1}=0.0168{m}_{e}{c}^{2}=8.6\,\mathrm{keV}$ (or Ti1 = Te1 = 108 K), relevant for typical ICM plasmas (Ryu et al. 2008; Porter et al. 2015).

The fast Mach number of MHD shocks is defined as Mf ≡ ush/uf, where the fast wave speed is ${u}_{{\rm{f}}}^{2}=\{{({c}_{{\rm{s}}1}^{2}+{u}_{{\rm{A}}}^{2})+{[({c}_{{\rm{s}}1}^{2}+{u}_{{\rm{A}}}^{2})}^{2}-4{c}_{{\rm{s}}1}^{2}{u}_{{\rm{A}}}^{2}{\cos }^{2}{\theta }_{\mathrm{Bn}}]}^{1/2}\}/2$. In the limit of high β (i.e., cs1 ≫ uA), Mf ≈ Ms.

The model parameters of our simulations are summarized in Table 1. We adopt β = 100, θBn = 63°, and mi/me = 100 as the fiducial values of the parameters. The incident flow velocity, u0, is specified to induce shocks with Ms ≈ 2–3, which are characteristic for cluster merger shocks (e.g., Ha et al. 2018), as noted in the introduction. Models with different Ms are named with the combination of the letter "M" and sonic Mach numbers (for example, the M2.0 model has Ms = 2.0). Models with parameters different from the fiducial values have names that are appended by a character for the specific parameter and its value: the M2.3-θ73 model has θBn = 73°, for example, while the M2.3-m400 model has mi/me = 400.

Table 1.  Model Parameters of Simulations

Model Name Ms MA u0/c θBn β Te1 = Ti1[K(keV)] mi/me ${L}_{x}[c/{w}_{\mathrm{pe}}]$ Ly [c/wpe] Δx[c/wpe] ${t}_{\mathrm{end}}[{w}_{\mathrm{pe}}^{-1}]$ ${t}_{\mathrm{end}}[{{\rm{\Omega }}}_{\mathrm{ci}}^{-1}]$
M2.0 2.0 18.2 0.027 63° 100 108(8.6) 100 × 103 80 0.1 1.13 × 105 30
M2.15 2.15 19.6 0.0297 63° 100 108(8.6) 100 × 103 80 0.1 1.13 × 105 30
M2.3 2.3 21 0.0325 63° 100 108(8.6) 100 × 103 80 0.1 1.13 × 105 30
M2.5 2.5 22.9 0.035 63° 100 108(8.6) 100 × 103 80 0.1 1.13 × 105 30
M2.75 2.75 25.1 0.041 63° 100 108(8.6) 100 × 103 80 0.1 1.13 × 105 30
M3.0 3.0 27.4 0.047 63° 100 108(8.6) 100 1.2 × 104 80 0.1 2.26 × 105 60
M2.15-θ53 2.15 19.6 0.0297 53° 100 108(8.6) 100 × 103 80 0.1 1.13 × 105 30
M2.15-θ73 2.15 19.6 0.0297 73° 100 108(8.6) 100 × 103 80 0.1 1.13 × 105 30
M2.3-θ53 2.3 21 0.0325 53° 100 108(8.6) 100 × 103 80 0.1 1.13 × 105 30
M2.3-θ73 2.3 21 0.0325 73° 100 108(8.6) 100 × 103 80 0.1 1.13 × 105 30
M2.0-β50 2.0 12.9 0.027 63° 50 108(8.6) 100 × 103 80 0.1 8.0 × 104 30
M2.3-β50 2.3 14.8 0.0325 63° 50 108(8.6) 100 × 103 80 0.1 8.0 × 104 30
M3.0-β50 3.0 19.4 0.047 63° 50 108(8.6) 100 × 103 80 0.1 8.0 × 104 30
M2.0-m400 2.0 18.2 0.013 63° 100 108(8.6) 400 × 103 80 0.1 1.5 × 105 10
M2.3-m400 2.3 21 0.016 63° 100 108(8.6) 400 × 103 80 0.1 1.5 × 105 10
M3.0-m400 3.0 27.4 0.023 63° 100 108(8.6) 400 × 103 80 0.1 1.5 × 105 10
M2.3-r2 2.3 21 0.0325 63° 100 108(8.6) 100 × 103 80 0.05 3.8 × 104 10
M2.3-r0.5 2.3 21 0.0325 63° 100 108(8.6) 100 × 103 80 0.2 3.8 × 104 10

Download table as:  ASCIITypeset image

Simulations are presented in units of the plasma skin depth, c/wpe, and the electron plasma oscillation period, ${w}_{\mathrm{pe}}^{-1}$, where ${w}_{\mathrm{pe}}=\sqrt{4\pi {e}^{2}n/{m}_{e}}$ is the electron plasma frequency. The Lx and Ly columns of Table 1 denote the x- and y-sizes of the computational domain. Except for the M3.0 model (see below), the longitudinal and transverse lengths are Lx = 7 × 103c/wpe and Ly = 80c/wpe, respectively, which are represented by a grid of cells with size Δx = Δy = 0.1c/wpe. The last two columns show the end times of simulations in units of ${w}_{\mathrm{pe}}^{-1}$ and the ion gyration period, ${{\rm{\Omega }}}_{\mathrm{ci}}^{-1}$, where Ωci = eB0/mic is the ion gyrofrequency. The ratio of the two periods scales as ${w}_{\mathrm{pe}}/{{\rm{\Omega }}}_{\mathrm{ci}}\propto ({m}_{i}/{m}_{e})\sqrt{\beta }$. For most models, simulations run up to tend wpe ≈ 1.13 × 105, which corresponds to tend Ωci ≈ 30 for β = 100 and mi/me = 100. The M3.0 model extends twice as far, up to tend wpe ≈ 2.26 × 105 or tend Ωci ≈ 60, and correspondingly has a longer longitudinal dimension of Lx = 1.2 × 104c/wpe. Comparison models with smaller β, M2.0-β50, M2.30-β50, and M3.0-β50 also go up to tend Ωci ≈ 30 (tend wpe ≈ 8.0 × 104). The models with mi/me = 400, on the other hand, are calculated only up to tend Ωci ≈ 10 (tend wpe ≈ 1.5 × 105). Models with different Δx/(c/wpe), M2.3-r2, and M2.3-r0.5 are also considered, in order to inspect the effects of spatial resolution. In each cell, 32 particles (16 per species) are placed. The time step is ${\rm{\Delta }}t=0.045[{w}_{\mathrm{pe}}^{-1}]$.

Compared to the reference model reported by GSN14a and GSN14b, our fiducial models have higher β (100 versus 20) and lower Ti1 = Te1 (108 K versus 109 K). As a result, our simulations run for a longer time—for instance, ωpetend ≈ 1.13 × 105 to reach tendΩci ≈ 30. Our shocks are also less relativistic. More importantly, this work also includes weaker shock models with Ms < 3.0, while GSN14a and GSN14b considered only shocks with Ms = 3.0.

3. Physics of ${Q}_{\perp }$-shocks

3.1. Critical Mach Numbers

The structures and time variations of collisionless shocks are primarily governed by the dynamics of reflected ions and the waves excited by the relative drift between reflected and incoming ions. In theories of collisionless shocks, therefore, a number of critical shock Mach numbers have been introduced to describe ion reflection and upstream wave generation; for a review, see Balogh & Truemann (2013). Although the main focus of this paper is the electron acceleration at Q-shocks, we here present a brief review of the "shock criticalities" due to reflected ions.

The reflection of ions has been often linked to the "first critical Mach number," ${M}_{{\rm{f}}}^{* }(\beta ,{\theta }_{\mathrm{Bn}});$ it was found for un2 = cs2 by applying the Rankine-Hugoniot jump relation to fast MHD shocks, i.e., the condition that the downstream flow speed normal to the shock surface equals the downstream sound speed (e.g., Edmiston & Kennel 1984). In supercritical shocks with ${M}_{{\rm{f}}}\gt {M}_{{\rm{f}}}^{* }$, the shock kinetic energy cannot be dissipated sufficiently through resistivity and wave dispersion, and hence a substantial fraction of incoming ions should be reflected upstream in order to sustain the shock transition from the upstream to the downstream. In subcritcal shocks below ${M}_{{\rm{f}}}^{* }$, on the other hand, the resistivity alone can provide enough dissipation to support a stable shock structure. In collisionless shocks, however, the reflection of ions occurs at the shock ramp, due to the magnetic deflection and the cross shock potential drop—the physics of which are beyond the fluid description. Hence, it should be investigated with simulations resolving kinetic processes.

In Q-shocks, the first critical Mach number also denotes the minimum Mach number, above which kinetic processes trigger overshoot/undershoot oscillations in the density and magnetic field, and the shock structures may become nonstationary under certain conditions. The reflection of ions is mostly due to the deceleration by the shock potential drop, and resonant and nonresonant waves are excited via streaming instabilities induced by reflected ions (e.g., Caprioli & Spitkovsky 2014a, 2014b). Such processes depend on shock parameters. For instance, in shocks with higher Ms and consequently higher shock kinetic energies, the structures tend to more easily fluctuate and become unsteady. In high-β plasmas, on the other hand, shocks could be stabilized against certain instabilities owing to fast thermal motions, which can subdue the relative drift between reflected and incoming particles; thus, theoretical analyses based on the cold plasma assumption could be modified in high-β plasmas. In Paper I, we found that ${M}_{{\rm{f}}}^{* }\approx 2.3$ for Q-shocks in ICM plasmas with β ≈ 100, which is higher than the fluid prediction by Edmiston & Kennel (1984). In Q-shocks, the kinetic processes involved in determining ${M}_{{\rm{f}}}^{* }$ are also part of the preacceleration of ions, and hence the injection to the Fermi-I process of DSA.

In Q-shocks, both ions and electrons are reflected through the magnetic deflection; the two populations are subject to deceleration by the magnetic mirror force due to converged magnetic field lines at the shock transition. In addition, the shock potential drop decelerates incident ions while it accelerates electrons toward the downstream direction. The reflected particles gain energy through the gradient drift along the motional electric field at the shock surface (SDA). Most of reflected ions, however, are trapped—mostly at the shock foot—before they advect downstream with the background magnetic field after about one gyromotion. As a result, streaming instabilities are not induced in the upstream, and hence the ensuing CR proton acceleration is ineffective, as previously reported with hybrid and PIC simulations (e.g., Caprioli & Spitkovsky 2014a, 2014b; Paper I). However, still the dynamics of reflected ions is primarily responsible for the main features of the transition zone of Q-shocks (e.g., Treumann & Jaroschek 2008; Treumann 2009). For instance, the current due to the drift motion of reflected ions generates the magnetic foot, ramp, and overshoot, and the charge separation due to reflected ions generates the ambipolar electric shock potential drop at the shock ramp.

In Q-shocks, the accumulation of reflected ions at the upstream edge of the foot may lead to the cyclic self-reformation of shock structures over ion gyroperiods and result in the excitation of low-frequency whistler waves in the shock foot region (e.g., Matsukiyo & Scholer 2006; Scholer & Burgess 2007). This leads to the so-called "second or whistler critical Mach number," ${M}_{{\rm{w}}}^{* }\approx (1/2)\sqrt{{m}_{i}/{m}_{e}}\cos {\theta }_{\mathrm{Bn}}$ in the β ≪ 1 limit (Kennel et al. 1985; Krasnoselskikh et al. 2002). In subcritical shocks with ${M}_{{\rm{f}}}\lt {M}_{{\rm{w}}}^{* }$, linear whistler waves can phase-stand in the shock foot upstream of the ramp. Dispersive whistler waves were found far upstream in interplanetary, subcritical shocks (e.g., Oka et al. 2006). Those waves interact with the upstream flow and contribute to the energy dissipation, effectively suppressing the shock reformation. Above ${M}_{{\rm{w}}}^{* }$, stationary linear wave trains cannot stand in the region ahead of the shock ramp.

The "third or nonlinear whistler critical Mach number," ${M}_{\mathrm{nw}}^{* }\approx \sqrt{{m}_{i}/2{m}_{e}}\cos {\theta }_{\mathrm{Bn}}$ in the β ≪ 1 limit, was introduced to describe the nonstationarity of shock structures. Krasnoselskikh et al. (2002) predicted that, in supercritical shocks with ${M}_{{\rm{f}}}\gt {M}_{\mathrm{nw}}^{* }$, nonlinear whistler waves turn over because of the gradient catastrophe, leading to the nonstationarity of the shock front and quasi-periodic shock-reformation (see Scholer & Burgess 2007). However, Hellinger et al. (2007) used 2D hybrid and PIC simulations to show that phase-standing oblique whistlers can be emitted in the foot even in supercritical Q-shocks, so the shock-reformation is suppressed in 2D. In fact, the nonstationarity and self-reformation of shock structures are important, long-standing problems in the study of collisionless shocks, which have yet to be fully understood (e.g., Scholer et al. 2003; Lembeǵe et al. 2004; Matsukiyo & Scholer 2006).

In the β ≪ 1 limit (i.e., in cold plasmas), ${M}_{{\rm{w}}}^{* }=2.3$ and ${M}_{\mathrm{nw}}^{* }=3.2$ for the fiducial parameter values adopted for our PIC simulations (mi/me = 100 and θBn = 63°). Hence, in some of the models considered here, ${M}_{{\rm{w}}}^{* }\lt {M}_{{\rm{s}}}\lt {M}_{\mathrm{nw}}^{* }$, so the whistler waves induced by reflected ions could be be confined within the shock foot without overturning. However, these critical Mach numbers increase to ${M}_{{\rm{w}}}^{* }=9.7$ and ${M}_{\mathrm{nw}}^{* }=13.8$ for the true ratio of mi/me = 1836. Therefore, in the ICM, weak Q-shocks are expected to be subcritical with respect to the two whistler critical Mach numbers, and so they would not be subject to self-reformation. The confirmation of these critical Mach numbers, or improved estimations for β ≳ 1, through numerical simulations is very challenging, as noted above. The excitation of oblique whistler waves and the suppression of shock-reformation via surface rippling require at least 2D simulations (e.g., Lembeǵe & Savoini 2002; Burgess 2006). The additional degree of freedom in higher-dimensional simulations tends to stabilize some instabilities revealed in lower-dimensional simulations (e.g., Matsukiyo & Matsumoto 2015). Moreover, simulation results are often dependent on mi/me, as well as the magnetic field configuration, i.e., whether B0 is in-plane or off-plane (Lembeǵe et al. 2009). Furthermore, adopting the realistic ratio of mi/me in PIC simulations is very computationally expensive, as pointed in the previous section.

As mentioned in the introduction, GSN14a and GSN14b showed that in high-β, Q-shocks, electrons can be preaccelerated via multiple cycles of SDA, due to the scattering by the upstream waves excited via the EFI. Here, we introduce the additional "EFI critical Mach number," ${M}_{\mathrm{ef}}^{* }$, above which the electron preacceleration is effective. We seek it in the next section, along with the relevant kinetic processes involved.

The space physics and ISM communities have been mainly interested in shocks in low-β plasmas (β ≲ 1), and hence the analytic relations simplified for cold plasmas are often quoted (e.g., the dispersion relation for fast magnetosonic waves used by Krasnoselskikh et al. (2002)). In such works, MA is commonly used to characterize shocks. However, in hot ICM plasmas, shocks have Ms ≈ Mf ≪ MA, and magnetic fields play dynamically less important roles. Moreover, the ion reflection at the shock ramp is governed mainly by Ms rather than MA (e.g., Paper I). Thus, in the rest of this paper, we will use the sonic Mach number Ms to characterize shocks.

3.2. Energization of Electrons

As mentioned in the introduction, GSN14a discussed the relativistic SDA theory for electrons in Q-shocks, which involves the electron reflection at the shock and the energy gain due to the drift along the motional electric field. In a subsequent paper, GSN14b showed that these electrons can induce the EFI, which leads to the excitation of oblique waves. The electrons return back to the shock due to the scattering by those self-excited upstream waves, and are further accelerated through multiple cycles of SDA (Fermi-like process). Below, we follow these previous papers to discuss how the physical processes depend on parameters such as Ms, θBn, and T1 for the shocks considered here (Table 1). Inevitably, we cite below some of equations presented in GSN14a and GSN14b.

3.2.1. Shock Drift Acceleration

GSN14a derived the criteria for electron reflection by considering the dynamics of electrons in the so-called de Hoffmann–Teller (HT, hereafter) frame, in which the flow velocity is parallel to the background magnetic field—and hence the motional electric field disappears both upstream and downstream of the shock (de Hoffmann & Teller 1950). In the HT frame, the upstream flow has ${u}_{t}={u}_{\mathrm{sh}}\sec {\theta }_{\mathrm{Bn}}$ along the background magnetic field. Hereafter, v and v represent the velocity components of incoming electrons, parallel and perpendicular to the background magnetic field, respectively, in the upstream rest frame, and ${\gamma }_{t}\,\equiv \,{(1-{u}_{t}^{2}/{c}^{2})}^{-1/2}$ is the Lorentz factor of the upstream flow in the HT frame.

The reflection criteria can be written as

Equation (4)

(Equation (19) of GSN14a), and

Equation (5)

assuming that the normalized cross-shock potential drop is ${\rm{\Delta }}\phi (x)\,\equiv \,e[{\phi }^{\mathrm{HT}}(x)-{\phi }_{0}^{\mathrm{HT}}]/{m}_{e}{c}^{2}\ll 1$. Here, $G\,\equiv \,{(1-{v}_{\parallel }{u}_{t}/{c}^{2})}^{2}$, $F\,\equiv \,{[1-{({v}_{\parallel }-{u}_{t})}^{2}/({{Gc}}^{2}{\cos }^{2}{\alpha }_{0})]}^{1/2}$, and ${\alpha }_{0}\,\equiv \,{\sin }^{-1}(1/\sqrt{b})$ with the magnetic compression ratio $b\,\equiv \,B{(x)}^{\mathrm{HT}}/{B}_{0}^{\mathrm{HT}}$. The superscript HT denotes the quantities in the HT frame. Note that, for Δϕ(x) = 0, Equation (5) becomes the same as Equation (20) of GSN14a.

In Figure 1(a), the red and blue solid lines mark the boundaries of the reflection criteria in Equations (4) and (5) for the M2.0 and M3.0 models, respectively. The red and blue dashed curves to the right of the vertical lines are the post-reflection velocities calculated via Equations (25) and (26) of GSN14a by inserting the boundary values of Equations (4) and (5). For b and Δϕ, the values estimated at the shock surface from simulation data were used. The solid black half-circle shows $v\,\equiv \,{({v}_{\parallel }^{2}+{v}_{\perp }^{2})}^{1/2}=c$, while the dashed black half-circle shows v = vth,e, where ${v}_{\mathrm{th},e}=\sqrt{2{k}_{{\rm{B}}}{T}_{e1}/{m}_{e}}$ is the electron thermal speed of the incoming flow.

Figure 1.

Figure 1. (a) Velocity diagram to analyze the electron reflection in weak ICM shocks; v and v are the electron velocity components—parallel and perpendicular to the background magnetic field, respectively—in the upstream rest frame. The black solid half-circle shows v = c, while the black dashed half-circle shows v = vth,e. The red (for the M2.0 model with Ms = 2 and θBn = 63°) and blue (for the M3.0 model with Ms = 3 and θBn = 63°) vertical lines draw the reflection condition for v in Equation (4), while the red and blue solid curves to the left of the vertical lines draw the reflection condition for v in Equation (5). The red and blue dashed curves to the right of the vertical lines draw the post-reflection velocity given in Equations (25) and (26) of GSN14a with the boundary values for the pre-reflection velocity given in Equations (4) and (5) of this paper. Electrons located in the region bounded by the colored vertical and solid lines are reflected to the region right of the vertical lines bounded by the dashed lines. (b) The fraction of reflected electrons, R, in percentage (black), and the average energy gain via a single SDA, $\langle {\rm{\Delta }}\gamma \rangle $, in units of ${m}_{e}{c}^{2}/{k}_{{\rm{B}}}T$ (red), as a function of Ms. The solid lines are for Q-shocks with θBn = 63°, while the dashed lines are for Q-shocks with θBn = 13°. (c) $R\cdot \langle {\rm{\Delta }}\gamma \rangle $ as a function of θBn for different Ms. (d) The EFI parameter, I, in Equation (7) as a function of Ms for models with θBn = 63° (black circles). The red squares are for models with θBn = 73°, while the blue triangles are for θBn = 53°. The instability condition is I > 0.

Standard image High-resolution image

As in GSN14b, we semi-analytically estimated the amount of the incoming electrons that satisfy the reflection condition, i.e., those bounded by the colored solid curves and the colored vertical lines together with the black circle in Figure 1(a). In Figure 1(b), the fraction of the reflected electrons, R, estimated for Q-shocks with θBn = 63° is shown by the black filled circles connected by a black solid line, while R for Q-shocks with θBn = 13° is shown by the black filled circles connected by a black dashed line. In Q-shocks, the reflection fraction, R, is quite high and increases with Ms, ranging ∼20–25% for 2 ≤ Ms ≤ 3. In Q-shocks, R is also high, ranging ∼17–20%, for 2.15 ≤ Ms ≤ 3, but it drops sharply at Ms = 2.

We should point out that the electron reflection becomes ineffective for superluminal shocks with large obliquity angles (i.e., ${u}_{\mathrm{sh}}/\cos {\theta }_{\mathrm{Bn}}\geqslant c$), because the electrons streaming upstream along the background field cannot outrun the shocks (see GSN14b). The obliquity angle for the superluminal behavior is ${\theta }_{\mathrm{sl}}\,\equiv \,\arccos ({u}_{\mathrm{sh}}/c)=86^\circ $ for the shock in M3.0 with mi/me = 100 and T1 = 108 K, and it is larger for smaller Ms. This angle is larger than θBn of our models in Table 1, and hence all the shocks considered here are subluminal.

For given T1 and Ms, the reflection of electrons is basically determined by b(x) and Δϕ(x), which quantify the magnetic deflection and the acceleration at the shock potential drop. Both b(x) and Δϕ(x) increase with increasing θBn. Larger b enhances the electron reflection (positive effect), while larger Δϕ(x) suppresses it (negative effect). In GSN14b, shocks are semi-relativistic with Δϕ ∼ 0.1–0.5, and hence the negative effect of the potential drop is substantial. However, in our models, shocks are less relativistic because of the lower temperature adopted, and ${\rm{\Delta }}\phi \sim {m}_{i}{u}_{\mathrm{sh}}^{2}/2{m}_{e}{c}^{2}\ll 1$. As a result, the magnetic deflection dominates over the acceleration by the cross-shock potential, leading to higher R at higher θBn. GSN14b showed that SDA becomes inefficient for ut ≳ vth,e ($\cos {\theta }_{\mathrm{Bn}}\lesssim \cos {\theta }_{\mathrm{limit}}={M}_{{\rm{s}}}\sqrt{{m}_{e}/{m}_{i}}$), which is more stringent than the superluminal condition (ut > c). Thus, the electron reflection fraction begins to decrease for θBn ≳ 60° in their models. Although not shown here, in our models, R monotonically increases with the obliquity angle for a given Ms, because the adopted θBn (≤73°) is smaller than the limiting obliquity angle, θlimit, for Ms = 2–3 and mi/me = 100.

The reflected electrons gain energy via SDA. We estimated the energy gain from a single SDA cycle as

Equation (6)

where γi and γr are the Lorentz factors for the pre-reflection and post-reflection electron velocities, respectively (Equation (24) of GSN14a). For given T1 (or given cs1), ut and Δγ depend on Ms and θBn. For the shocks considered here, γi ≈ 1 and ut ≪ c, so ${\rm{\Delta }}\gamma \approx 2[{({u}_{t}/c)}^{2}-{u}_{t}{v}_{\parallel }/{c}^{2}]$. In Figure 1(b), the red filled circles connected with the red solid line show the average energy gain, $\langle {\rm{\Delta }}\gamma \rangle $, in units of ${m}_{e}{c}^{2}/{k}_{{\rm{B}}}{T}_{e}$, estimated for Q-shocks with θBn = 63°. The red filled circles with the red dashed line show the quantity for Q-shocks with θBn = 13°. Here, the average was taken over the incoming electrons of Maxwellian distributions, so the $\langle {\rm{\Delta }}\gamma \rangle $ shown are representative values during the initial development stage of suprathermal particles.

In addition, the product of R and $\langle {\rm{\Delta }}\gamma \rangle $ is plotted as a function of θBn for different Ms in Figure 1(c). For the models in Table 1, R was calculated using b and Δϕ estimated at the shock surface from simulation data, as mentioned above. For the rest, the values of b and Δϕ for the models with θBn = 13° presented in Paper I were adopted for Q-shocks, while the values for the models with θBn = 63° presented in this work were adopted for Q-shocks. Panels (b) and (c) of Figure 1 show that more electrons are reflected and higher energies are achieved at higher Ms and larger θBn.

3.2.2. Electron Firehose Instability

GSN14b performed periodic-box simulations with beams of streaming electrons, in order to isolate and study the EFI due to the reflected and SDA-energized electrons. They found the following: nonpropagating (ωr ≈ 0), oblique waves with wavelengths ∼(10–20)c/wpe are excited dominantly, δBz is stronger than δBx and δBy (the initial magnetic field is in the xy plane), and both the growth rate and the dominant wavelength of the instability are insensitive to the mass ratio mi/me. These results are consistent with the expectations from the previous investigations of oblique EFI (e.g., Gary & Nishimura 2003).

The EFI criterion in weakly magnetized plasmas can be defined as

Equation (7)

where ${\beta }_{e\parallel }\,\equiv \,8\pi {n}_{e}{k}_{{\rm{B}}}{T}_{e\parallel }/{B}_{0}^{2}$ is the electron beta parallel to the initial magnetic field (Equation (10) of GSN14b). Equation (7) indicates that the instability parameter, I, is larger for higher βe for a given value of Te/Te. For higher Ms, R is larger and Te/Te is smaller, leading to larger I.

Figure 1(d) shows the instability parameter of shocks with θBn = 63°, as a function of Ms, estimated using the velocity distributions of the electrons that are located within (0–1) rL,i (rL,i is the ion Larmor radius with the upstream field B0) upstream from the shock position in simulation data. For the Ms = 2.0 model, I ≲ 0 with almost no temperature anisotropy, so the upstream plasma should be stable against the EFI. This finding, which will be further updated with simulation results in the next section, suggests that the preacceleration of electrons due to the EFI may not operate effectively in very weak shocks. For Ms > 2, on the other hand, the EFI criterion is satisfied and I increases with increasing Ms, implying that larger temperature anisotropies (Te > Te) at higher Ms shocks induce stronger EFIs. The figure also indicates that I increases steeply around Ms ≈ 2.2–2.3. Additional data points, marked using the blue triangles connected by the blue dashed line (θBn = 53°) and the red squares connected with the red dashed line (θBn = 73°), show that Q-shocks with higher obliquity angles are more unstable to the EFI.

4. Results

4.1. Shock Structures

As discussed in Section 3.1, the criticality defined by the first critical Mach, ${M}_{{\rm{f}}}^{* }$, primarily governs the structures and time variations of collisionless shocks. In subcritical shocks, most of the shock kinetic energy is dissipated at the shock transition, resulting in relatively smooth and steady structures. In supercritical shocks, on the other hand, reflected ions induce overshoot/undershoot oscillations in the shock transition and ripples along the shock surface (e.g., Lowe & Burgess 2003, Krasnoselskikh et al. 2013, and Trotta & Burgess 2019). Q-shocks with ${M}_{{\rm{f}}}\gt {M}_{{\rm{f}}}^{* }$ may undergo quasi-periodic reformation owing to the accumulation of upstream low-frequency waves. Q-shocks are less prone to reformation, because reflected ions mostly advect downstream after about one gyromotion.

Figure 2 compares the magnetic field structure for Q and Q-shocks with different Ms. In the Q-shocks, the overshoot/undershoot oscillation becomes increasingly more evident for higher Ms, but the shock structure seems to be quasi-stationary without any signs of reformation. This is consistent with the fact that the nonlinear whistler critical Mach number for our fiducial models is ${M}_{\mathrm{nw}}^{* }=3.2$. On the other hand, the Q-shock in the M3.2-2D model exhibits quasi-periodic reformations.

Figure 2.

Figure 2. Stack plots of the total magnetic field strength, averaged over the transverse direction, B, in the M2.0, M2.3, and M3.0 models from tΩci = 20 (bottom) to tΩci = 30 (top). The M3.2-2D model represents the Q-shock with Ms = 3.2 and θBn = 13°, taken from Paper I. Here, B0 is the magnetic field strength far upstream.

Standard image High-resolution image

According to the fluid description of Edmiston & Kennel (1984), ${M}_{{\rm{f}}}^{* }\approx 1$ for Q-shocks in high-β plasmas, so the fraction of reflected ions is expected be relatively high in all the shock models under consideration. As can be seen in the phase-space distribution of protons in Figure 3(a)–(c), the back-streaming ions turn around mostly within about one ion gyroradius in the shock ramp (x − xsh ≲ 60c/wpe). Note that, with mi/me = 100 in the M3.0 model, the shock ramp corresponds to the region of x − xs < 60c/ωpe, while the foot extends to x − xs ≈ rL,i ≈ 200c/ωpe (e.g., Balogh & Truemann 2013). In the M2.0 model, rL,i is smaller and so the characteristic widths of the ramp and foot are accordingly smaller as well. Figure 3(d) shows that the ion reflection fraction, ${\alpha }_{\mathrm{ref},i}={n}_{\mathrm{ref},i}/{n}_{i}$, increases with increasing Ms, and such a trend is almost independent of the mass ratio mi/me. Here, ${n}_{\mathrm{ref},i}$ was calculated as the number density of ions with vx > 0 in the shock rest frame in the ramp region. Because ${\alpha }_{\mathrm{ref},i}$ increases abruptly at Ms ≈ 2.2–2.3, we may regard ${M}_{{\rm{f}}}^{* }\approx 2.3$ as an effective value for the first critical Mach number, above which high-β Q-shocks reflect a sufficient amount of incoming ions and become supercritical. From Figure 2, we can see that the ensuing oscillations in shock structures appear noticeable in earnest only for Mf ≳ 2.3. Our estimation of ${M}_{{\rm{f}}}^{* }$ is higher than the prediction of Edmiston & Kennel (1984). This might be partly because, in high-β plasmas, kinetic processes due to fast thermal motions could suppress some of microinstabilities driven by the relative drift between backstreaming and incoming ions, as mentioned before.

Figure 3.

Figure 3. Ion phase-space distributions in the $x-{p}_{{ix}}$ plane for the M2.0 model (a), the M2.3 model (b), and the M3.0 model (c), at tΩci ≈ 30. The x-coordinate is measured relative to the shock position, xsh, in units of c/wpe. The bar at the top displays the color scale for the log of the ion phase-space density (arbitrary units). In panel (d), the black circles show the fraction of reflected ions in the shock ramp region of 0 ≤ x − xsh ≤ 60c/wpe at tΩci ≈ 30 for the fiducial models with mi/me = 100, while the red circles show the same fraction in 0 ≤ x − xsh ≤ 240c/wpe at tΩci ≈ 10 for the three models with mi/me = 400.

Standard image High-resolution image

4.2. Electron Preacceleration

Reflected electrons are energized via SDA at the shock ramp. The consequences of this can be observed in the phase-space distribution of electrons shown in Figure 4: see panels (a)–(c) for the M2.0 model and panels (e)–(g) for the M3.0 model. Because B0 is in the xy plane, electrons initially gain the z-momentum, pez, through the drift along the motional electric field, E0 = $-{{\boldsymbol{v}}}_{0}/c$ × B0, and then the gain is distributed to pex and pey during gyration motions. In addition, reflected electrons streaming along the background magnetic field with small pitch angles in the upstream region have larger positive py than px. Panels (d) and (h) of Figure 4 also show the distributions of electron density ne (black curve) and By (red curve) around the shock transition.

Figure 4.

Figure 4. Electron phase-space distributions and shock structures for the M2.0 model (left panels) and the M3.0 model (right panels) at wpet ≈ 1.13 × 105 (tΩci ≈ 30). The x-coordinate is measured relative to the shock position, xsh, in units of c/wpe. The distributions in x − pex, x − pey, and x − pez, and the distributions of electron number density ne and transverse magnetic field By in units of upstream values are shown from top to bottom. The bar at the top displays the color scale for the log of the electron phase-space density (arbitrary units).

Standard image High-resolution image

If electrons are accelerated via the full Fermi-I process (i.e., DSA) and in the test-particle regime, the momentum distribution follows the so-called DSA power-law:

Equation (8)

where fN is the normalization factor and $q({M}_{{\rm{s}}})=3r/(r-1)$ is the slope (Drury 1983; Kang & Ryu 2010). Here, pmax is the maximum momentum of accelerated electrons that increases with the shock age before any energy losses set in. The injection momentum, pinj, is the minimum momentum with which electrons can diffuse across the shock and be injected into the full Fermi-I process as described in the introduction. It marks the approximate boundary between the thermal and nonthermal momentum distributions. The momentum spectrum in Equation (8) can be transformed to the energy spectrum in terms of the Lorentz factor as

Equation (9)

where the slope is s(Ms) = q(Ms) − 2. For instance, s = 2.5 for Ms = 3.0, while s = 2.93 for Ms = 2.0.

The injection momentum, which can be estimated as ${p}_{\mathrm{inj}}\sim 3{p}_{\mathrm{th},i}$ (e.g., Kang et al. 2002; Caprioli et al. 2015, Paper I), marks the approximate boundary between the suprathermal and nonthermal momentum distributions and it is well beyond the highest momentum that electrons can achieve in our PIC simulations. In the M3.0 model, for example, pinj corresponds to γinj ≈ 10, while electrons of the highest momenta reach only γ ≲ 2 (see Figure 5). In other words, our simulations could follow only the preacceleration of suprathermal electrons, which are not energetic enough to diffuse across the shock. Thus, the DSA slope, s(Ms), is not necessarily reproduced in the energy spectra of electrons. However, the development of power-law tails with s(Ms) may indicate that the preaccelerated electrons have undergone a Fermi-like process, as proposed by GSN14a and GSN14b. Although not shown here, the trajectories of suprathermal electrons in our simulations demonstrated that, in supercritical shocks, the electrons gain energies via multiple cycles of SDA through scattering between the shock ramp and upstream waves, just like Figure 8 of GSN14a.

Figure 5.

Figure 5. Upstream electron energy spectra at tΩci = 10 (blue lines), tΩci = 30 (red), and tΩci = 60 (green) in various models. The spectra were taken from the region of (0–1)rL,i upstream of the shock. The black dotted–dashed lines indicate the test-particle power laws of Equation (9), while the purple dashed lines show the Maxwellian distributions in the upstream region.

Standard image High-resolution image

The upper panels of Figure 5 compare the electron energy spectra, (γ − 1)dN/, taken from the upstream region of (0–1)rL,i, ahead of the shock, at tΩci = 10 (blue lines) and 30 (red lines), in the models with different Ms. In the case of the M3.0 model, the simulation is performed longer and the spectrum at tΩci = 60 is also shown with the green line (which almost overlaps with the red line). As described in Section 3.1, reflected electrons gain energy initially via SDA, and may continue to be accelerated via a Fermi-like process and multiple cycles of SDA, if oblique waves are excited by the EFI. Two points are evident here. (1) In the M2.0 model, the blue and red lines almost coincide, indicating almost no change of the spectrum from tΩci = 10 to 30. The spectrum is similar to that of the electrons energized by a single cycle of SDA, which was illustrated in Figure 7 of GSN14a. Thus, the Fermi-like process followed by the EFI may not operate efficiently in this model. (2) The M3.0 model, on the other hand, exhibits a further energization from tΩci = 10 to 30, demonstrating the presence of a Fermi-like process. However, there is no difference in the spectra of tΩci = 30 and 60. As a matter of fact, the energy spectrum of suprathermal electrons seems to saturate beyond tΩci ≈ 20 (not shown in the figure). This should be due to the saturation of the EFI and the lack of further developments of longer wavelength waves (see the next subsection for further discussions).

The middle and lower panels of Figure 5 show the electron energy spectra in models with different parameters. The models with mi/me = 400 were followed only up to tend Ωci = 10 (blue lines), because longer computing time is required for larger mi/me. Comparison of the two sets of models with different values of mi/me confirms that the EFI is  almost independent of mi/me for sufficiently large mass ratios, as previously shown by Gary & Nishimura (2003) and GSN14b, and so is the electron acceleration. Figure 5(d) for the M2.3-θ73 model indicates that SDA—and hence the EFI—is more efficient at higher obliquity angles, which is consistent with Figure 1(c). Panels (e) and (f) of Figure 5, displaying the models with β = 50, demonstrate that the EFI is more efficient at higher β. All the models with Ms ≳ 2.3 show marginal power law–like tails beyond the spectra energized by a single cycle of SDA.

With the M2.3-r2 and M2.3-r0.5 models, we examined how the electron energy spectrum depends on the grid resolution, although the comparison plots are not shown. Our simulations with different Δx produced essentially the same spectra, especially for the suprathermal part.

In Paper I, we calculated the injection fraction, ξ(Ms, θBn, β), of nonthermal protons with p ≥ pinj for Q shocks, as a measure of the DSA injection efficiency. Because the simulations in this paper can follow only the preacceleration stage of electrons via an upstream Fermi-like process, we define and estimate the "fraction of suprathermal electrons" as follows:

Equation (10)

where $\langle f(p)\rangle $ is the electron distribution function, averaged over the upstream region of (0–1) rL,i, ahead of the shock. For the "suprathermal momentum," above which the electron spectrum changes from a Maxwellian to a power law–like distribution, we use pspt ≈ 3.3pth,e. Note that ${p}_{\mathrm{spt}}\approx {p}_{\mathrm{inj}}{({m}_{i}/{m}_{e})}^{-1/2}$. For the M3.0 model, for instance, pspt corresponds to γ ≈ 1.25. Different choices of pspt result in different values of ζ, of course, but the dependence on parameters such as Ms and θBn does not change much.

In Figure 6, the circles connected with solid lines show the suprathermal fraction, ζ(Ms), for the fiducial models with θBn = 63° at tΩci = 10–30. This fraction is expected to increase with increasing Ms, because the EFI parameter, I, is larger for higher Ms (see Figure 1(d)). Moreover, it increases with time until tΩci ≈ 20, due to a Fermi-like process, as shown in Figure 5—except for the M2.0 model where the increase in time is insignificant. However, ζ seems to stop growing for tΩci ≳ 20, indicating the saturation of electron preacceleration. This is related to the reduction of temperature anisotropy via electron scattering and the ensuing decay of EFI-induced waves, which will be discussed more in the next section.

Figure 6.

Figure 6. Suprathermal fraction, ζ, defined in Equation (10), as a function of Ms for the fiducial models (θBn = 63°) at tΩci = 10 (blue circles), 15 (cyan circles), 20 (green circles), and tΩci = 30 (red circles). The triangles are for the models with θBn = 53° at tΩci = 10 (blue) and 30 (red), while the squares are for the models with θBn = 73° at tΩci = 10 (blue) and 30 (red).

Standard image High-resolution image

The red solid line in Figure 6 is represented roughly by $\zeta \propto {M}_{{\rm{s}}}^{4}$ in the range of 2.3 ≲ Ms ≤ 3, but it drops rather abruptly below 2.3, deviating from the power-law behavior. We note that the Mach number dependence of ζ is steeper than that of the ion injection fraction for Q-shocks, which is roughly $\xi \propto {M}_{{\rm{s}}}^{1.5}$, as shown in Paper I. This implies that the kinetic processes involved in electron preacceleration might be more sensitive to Ms (see Section 3.2).

Figure 6 also shows ζ for models with θBn = 53° (triangles) and θBn = 73° (squares). For shocks with larger θBn, the reflection of electrons and the average SDA energy gain are larger, resulting in larger I, as shown in Figure 1(c) and (d). Hence, ζ should be larger at a higher obliquity angle. However, for θBn > θlimit ≈ 73–78°, ζ should begin to decrease, as mentioned in Section 3.2.1.

Based on the above results, we propose that the preacceleration of electrons is effective only in Q-shocks with Ms ≳ 2.3 in the hot ICM, i.e., ${M}_{\mathrm{ef}}^{* }\approx 2.3$. We should point out that this is close to the first critical Mach number for ion reflection, ${M}_{{\rm{f}}}^{* }\approx 2.3$, estimated from the Mach number dependence of the fraction of reflected ions, ${\alpha }_{\mathrm{ref},i}$, shown in Figure 3(d). As shown in Figure 2, overshoot/undershoot oscillations develop in the shock transition, owing to a sufficient amount of reflected ions, in shocks with Ms ≳ 2.3; with larger magnetic field compression due to the oscillations, more electrons are reflected and energized via SDA (see Section 3.2.1). Hence, we expect that the electron reflection is directly linked with the ion reflection, so ${M}_{\mathrm{ef}}^{* }$ would be related with ${M}_{{\rm{f}}}^{* }$. Note that the critical Mach number, ${M}_{{\rm{f}}}^{* }\approx 2.3$, is also similar to the first critical Mach number for ion reflection and injection into DSA in Q-shocks in high-β plasmas, ${M}_{{\rm{s}}}^{* }\approx 2.25$ (Paper I).

4.3. Upstream Waves

The nature and origin of upstream waves in collisionless shocks have long been investigated through both analytical and simulation studies, with the help of in situ observations of Earth's bow shock. In Q-shocks with low Mach numbers, magnetosonic waves such as phase-standing whistlers and long-wavelength whistlers are known to be excited by backstreaming ions via an ion/ion-beam instability (e.g., Krauss Varban & Omidi 1991). This is especially so in supercritical Q-shocks, where the foreshock region is highly turbulent with large-amplitude waves and the shock transition can undergo quasi-periodic reformation due to the nonlinear interaction of accumulated waves and the shock front (see Paper I and Figure 2(d)). In Q-shocks, a sufficient amount of incoming protons can be reflected at the shock, which in turn may excite fast magnetosonic waves. As discussed in Section 3.1, two whistler critical Mach numbers, ${M}_{{\rm{w}}}^{* }$ and ${M}_{\mathrm{nw}}^{* }$, are related with the upstream emission of whistler waves and the nonlinear breaking of whistler waves in the shock foot. In some of our models, ${M}_{{\rm{w}}}^{* }\lt {M}_{{\rm{s}}}\lt {M}_{\mathrm{nw}}^{* }$, and hence whistler waves are confined within the shock foot and shock reformation does not occur.

In supercritical shocks with Ms ≳ 2.3, we expect to see the following three kinds of waves: (1) nearly phase-standing whistler waves with kc/ωpi ∼ 1 (kc/ωpe ∼ 0.1) excited by reflected ions (e.g., Hellinger et al. 2007; Scholer & Burgess 2007), where ${w}_{\mathrm{pi}}=\sqrt{4\pi {e}^{2}n/{m}_{i}}$ is the ion plasma frequency (wpi = 0.1wpe for me/mi = 100); (2) phase-standing oblique waves with kc/ωpe ∼ 0.4 and larger θBk (the angle between the wave vector k and B0) excited by the EFI; and (3) propagating waves with kc/ωpe ∼ 0.3 and smaller θBk, also excited by the EFI (e.g., Hellinger et al. 2014).

Here, we focus on the waves excited by the EFI described in Section 3.2.2. Previous studies on the EFI and the EFI-induced waves showed the following characteristics (Gary & Nishimura 2003; Camporeale & Burgess 2008; Hellinger et al. 2014; Lazar et al. 2014, GSN14b). (1) The magnetic field fluctuations in the EFI-induced waves are predominantly along the direction perpendicular to both k and B0, i.e., $| \delta {B}_{z}| $ is larger than $| \delta {B}_{x}| $ and $| \delta {B}_{y}| $ in our geometry. (2) Phase-standing oblique waves with almost zero oscillation frequencies (ωr ≈ 0) have higher growth rates than propagating waves (${\omega }_{r}\,\ne \,0$). (3) Nonpropagating modes decay to propagating modes with longer wavelengths and smaller θBk. (4) The EFI-induced waves scatter electrons, resulting in a reduction of the electrons' temperature anisotropy, which in turn leads to the damping of the waves.

Figure 7 shows the distribution of magnetic field fluctuations, $\delta {\boldsymbol{B}}$, in the upstream region for the M2.0 and M3.0 models. The epoch shown, tΩci ≈ 7 (wpe t ≈ 2.63 × 104) is early—yet in the M3.0 model, waves are well-developed (see also Figure 9), while the energization of electrons is still undergoing (see Figure 5). For the supercritical shock of the M3.0 model, we interpret that there are ion-induced whistlers in the shock ramp region of 0 ≲ x − xs ≲ 60c/ωpe, while EFI-induced oblique waves are present over the whole region shown. As shown in GSN14b, the EFI-excited waves are oblique, with θBk ∼ 60° and $| \delta {B}_{z}| \gt | \delta {B}_{x}| $ and $| \delta {B}_{y}| $. The increase in δBy toward x − xsh = 0 is due to the compression in the shock ramp. In the subcritical shock of the M2.0 model, on the other hand, the fractions of reflected ions and electrons are not sufficient for either the emission of whistler waves or the excitation of EFI-induced waves, so no substantial waves are present in the shock foot. This is consistent with the instability condition shown in Figure 1(d).

Figure 7.

Figure 7. Magnetic field fluctuations δBx in (a) and (d), δBy in (b) and (e), and δBz in (c) and (f), normalized to B0, in the upstream region of 0 < (x − xsh)wpe/c < 100 at wpet ≈ 2.63 × 104 (tΩci ≈ 7) for the M2.0 model (top panels) and the M3.0 model (bottom panels).

Standard image High-resolution image

Figure 8 compares δBz in six different models at tΩci ≈ 10. The wave amplitude increases with increasing Ms, and the EFI seems only marginal in the M2.3 models. This result confirms our proposal for the "EFI critical Mach number" ${M}_{\mathrm{ef}}^{* }\approx 2.3$, presented in Section 4.2.  Moreover, this figure corroborates our findings that the EFI is more efficient at larger θBn and higher β. From δBz of the M2.3 and M3.0 models in Figures 7 and 8, the dominant waves in the shock foot seem to have λ ∼ 15–20c/ωpe, so they are consistent with the EFI-induced waves (GSN14b).

Figure 8.

Figure 8. Magnetic field fluctuations, δBz, normalized to B0, in the upstream region of $0\lt (x-{x}_{\mathrm{sh}}){w}_{\mathrm{pe}}/c\lt 100$ at wpet ≈ 3.76 × 104 (tΩci ≈ 10) for six different models.

Standard image High-resolution image

Figure 9 shows the time evolution of the average of the magnetic field fluctuations, $\left\langle \delta {B}_{z}^{2}/{B}_{0}^{2}\right\rangle $, and the magnetic energy power, ${P}_{{Bz}}(k)\propto | \delta {B}_{z}(k){| }^{2}k$, of upstream waves for the M3.0 model. According to the linear analysis by Camporeale & Burgess (2008), the growth rate of the EFI peaks at kmaxc/ωpe ∼ 0.4 for βe = 10 and Te/Te = 0.7. Thus, we interpret that the powers in the range of kc/ωpe ∼ 0.2–0.3 are due to the oblique waves induced by the EFI, while those of kc/ωpe ≲ 0.15 are contributed by the phase-standing whistler waves induced by reflected ions.

Figure 9.

Figure 9. (a) Time evolution of $\left\langle \delta {B}_{z}^{2}/{B}_{0}^{2}\right\rangle $, the square of the magnetic field fluctuations normalized to the background magnetic field, averaged over the square region of $80\times 80{(c/{\omega }_{\mathrm{pe}})}^{2}$ covering $0\lt (x-{x}_{\mathrm{sh}}){w}_{\mathrm{pe}}/c\lt 80$, for the M3.0 model. (b) Time evolution of ${P}_{{B}_{z}}(k)\propto | \delta {B}_{z}(k){| }^{2}k$, the magnetic energy power of δBz in the square region of $80\times 80{(c/{\omega }_{\mathrm{pe}})}^{2}$ for the M3.0 model. (c) ${P}_{{B}_{z}}(k)$ vs. ${kc}/{w}_{\mathrm{pe}}$ at five different time epochs.

Standard image High-resolution image

Moreover, through periodic box simulations of the EFI, Camporeale & Burgess (2008) and Hellinger et al. (2014) demonstrated that initially nonpropagating oblique modes grow and then saturate, followed by the transfer of wave energy into propagating modes with longer wavelengths and smaller θBk. Figure 8 of Camporeale & Burgess (2008) and Figure 4 of Hellinger et al. (2014) show that a cycle of the EFI-induced wave growth and decay occurs on timescales of roughly several hundreds of tΩce. We suggest that the oscillatory behaviors of the excited waves with the timescales of ${{tw}}_{\mathrm{pe}}\sim 2\times {10}^{4}\mbox{--}4\times {10}^{4}$ (tΩce ∼ 500–1000) shown in Figure 9 would be related to those characteristics of the EFI. Figure 9(c) illustrates such a cycle during the period of t wpe ≈ (1.8–2.1) × 105: excitation with kmaxc/wpe ≈ 0.3 → inverse cascade with kmaxc/wpe ≈ 0.2 → damping of waves.

Our results indicate that the EFI-induced waves do not further develop into longer wavelength modes with λ ≫ λmax, where λmax ≈ 15–20c/ωpe is the wavelength of the maximum linear growth. Note that λmax is close to the gyroradius of electrons with γ ≲ 2. Thus, the acceleration of electrons via resonant scattering by the EFI-induced waves is saturated. As a consequence, the energization of electrons stops at the suprathermal stage (γ < 2) and does not proceed all the way to the DSA injection momentum (γinj ≈ 10). We interpret that this result should be due to the intrinsic properties of the EFI, rather than the limitations or artifacts of our simulations,  as shown by the studies of Camporeale & Burgess (2008) and Hellinger et al. (2014). Hence, we conclude that the preacceleration via the EFI alone may not explain the injection of electrons to DSA in weak ICM shocks. However, the conclusion needs to be further verified through a more detailed study of the EFI and EFI-induced waves for high-β ICM plasmas, including kinetic linear analyses and numerical simulations, which we leave for a future work.

5. Summary

In Q-shocks, a substantial fraction of incoming particles are reflected at the shock ramp. Most of these reflected ions are advected downstream along with the underlying magnetic field after about one gyromotion, but the structures of the shocks are still primarily governed by the dynamics of reflected ions. Especially in supercritical shocks, the accumulation of reflected ions in the shock ramp generates overshoot/undershoot oscillations in the magnetic field, ion/electron densities, and electric shock potential. Reflected electrons, on the other hand, can stream along the background magnetic field with small pitch angles in the upstream region. As presented in GSN14a and GSN14b, the SDA reflected electrons produce the temperature anisotropy, Te > Te, which induces the EFI; the EFI in turn excites oblique waves in the upstream region. Electrons are then scattered between the shock ramp and the upstream waves, and gain energies via a Fermi-like process involving multiple cycles of SDA. All these processes depend most sensitively on Ms among a number of other shock parameters; for instance, the development of the EFI and the energization of electrons are expected to be inefficient in very weak shocks with Ms close to unity.

In this paper, we have used 2D PIC simulations to study the preacceleration of electrons facilitated by the EFI in Q-shocks with Ms ≲ 3 in the high-β ICM. Various shock parameters have been considered, as listed in Table 1. Our findings can be summarized as follows:

  • 1.  
    For ICM Q-shocks, ion reflection and overshoot/undershoot oscillations in the shock structures become increasingly more evident for Ms ≳ 2.3, while the shock structures seem relatively smooth and quasi-stationary for lower Mach number shocks. Hence, we suggest that the effective value of the first critical Mach number would be ${M}_{{\rm{f}}}^{* }\approx 2.3$, which is higher than previously estimated from the MHD Rankine–Hugoniot jump condition by Edmiston & Kennel (1984).
  • 2.  
    Because electron reflection is affected by ion reflection and the ensuing growth of overshoot/undershoot oscillation, the EFI critical Mach number, ${M}_{\mathrm{ef}}^{* }\approx 2.3$, seems to be closely related with ${M}_{{\rm{f}}}^{* }$. Oscillations in the shock structures enhance the magnetic mirror in the shock ramp, providing a favorable condition for the efficient reflection of electrons. Only in shocks with ${M}_{{\rm{s}}}\gt {M}_{\mathrm{ef}}^{* }$ are the reflection and SDA of electrons efficient enough to generate sufficient temperature anisotropies, which can trigger the EFI and the excitation of oblique waves.
  • 3.  
    We have presented the fraction of suprathermal electrons, ζ(Ms, θBn), defined as the number fraction of electrons with p ≥ pspt = 3.3pth,e in the upstream energy spectrum. The suprathermal fraction increases with increasing Ms, roughly as $\zeta \propto {M}_{{\rm{s}}}^{4}$ for the fiducial models. Below ${M}_{\mathrm{ef}}^{* }\approx 2.3$, ζ drops sharply, indicating inefficient electron preacceleration in low Mach number shocks. This fraction also increases with increasing θBn. For shocks with larger θBn, the reflection of electrons and the average SDA energy gain are larger, and hence ζ is larger.
  • 4.  
    In the supercritical M3.0 model, the suprathermal tail of electrons extends to higher γ in time, but it saturates beyond tΩci ≈ 20 with the highest energy of γ ≲ 2. In order for suprathermal electrons to be injected into DSA, their energies should reach at least γinj ≳ 10. We interpret that such saturation is due to the lack of wave powers with long wavelengths. The maximum growth of the EFI in the linear regime is estimated to be at λmax ≈ 15–20c/ωpe. The EFI becomes stabilized owing to the reduction of electron temperature anisotropy, before waves with λ ≫ λmax develop. This implies that the preacceleration of electrons due to a Fermi-like process and multiple cycles of SDA, facilitated by the upstream waves excited via the EFI, may not proceed all the way to DSA in high-β, Q-shocks.

Our results indicate that processes other than those considered in this paper may be crucial to understand the origin of radio relics in galaxy clusters. For instance, in the reacceleration model, pre-existing fossil electrons are assumed (e.g., Kang 2016a, 2016b). In particular, fossil electrons with γ ∼ 10–100 could be scattered by ion-induced waves and/or pre-existing turbulent waves and participate in DSA. Park et al. (2015), for instance, showed through 1D PIC simulations that electrons can be injected to DSA and accelerated via the full Fermi-I process, even in Q with MA ≈ Ms = 20 and θBn = 30°. In addition, if shock surfaces are highly nonuniform with varying Ms and θBn (e.g., Hong et al. 2015; Ha et al. 2018), the features of Q and Q-shocks may be mixed up, facilitating the upstream environment of abundant waves for electron scattering. However, all these processes need to be investigated in detail before their roles are discussed, and we leave such investigations for future works.

The authors thank the anonymous referee for constructive comments. H.K. was supported by the Basic Science Research Program of the National Research Foundation of Korea (NRF) through grant 2017R1D1A1A09000567. D.R. and J.-H. H. were supported by the NRF through grants 2016R1A5A1013277 and 2017R1A2A1A05071429. J.-H. H. was also supported by the Global PhD Fellowship of the NRF through 2017H1A2A1042370.

Please wait… references are loading.
10.3847/1538-4357/ab16d1