Articles

MAGNETIC FIELD AMPLIFICATION IN NONLINEAR DIFFUSIVE SHOCK ACCELERATION INCLUDING RESONANT AND NON-RESONANT COSMIC-RAY DRIVEN INSTABILITIES

, , , and

Published 2014 June 23 © 2014. The American Astronomical Society. All rights reserved.
, , Citation Andrei M. Bykov et al 2014 ApJ 789 137 DOI 10.1088/0004-637X/789/2/137

0004-637X/789/2/137

ABSTRACT

We present a nonlinear Monte Carlo model of efficient diffusive shock acceleration where the magnetic turbulence responsible for particle diffusion is calculated self-consistently from the resonant cosmic-ray (CR) streaming instability, together with non-resonant short- and long-wavelength CR-current-driven instabilities. We include the backpressure from CRs interacting with the strongly amplified magnetic turbulence which decelerates and heats the super-Alfvénic flow in the extended shock precursor. Uniquely, in our plane-parallel, steady-state, multi-scale model, the full range of particles, from thermal (∼eV) injected at the viscous subshock to the escape of the highest energy CRs (∼PeV) from the shock precursor, are calculated consistently with the shock structure, precursor heating, magnetic field amplification, and scattering center drift relative to the background plasma. In addition, we show how the cascade of turbulence to shorter wavelengths influences the total shock compression, the downstream proton temperature, the magnetic fluctuation spectra, and accelerated particle spectra. A parameter survey is included where we vary shock parameters, the mode of magnetic turbulence generation, and turbulence cascading. From our survey results, we obtain scaling relations for the maximum particle momentum and amplified magnetic field as functions of shock speed, ambient density, and shock size.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The existence of strong, super-Alfvénic collisionless shocks in many astrophysical objects, such as supernova remnants (SNRs), extra-galactic radio jets, clusters of galaxies, and compact accreting sources, has been revealed to us by nonthermal radiation produced by relativistic particles. Diffusive shock acceleration (DSA), the most favored acceleration mechanism for producing these relativistic particles (e.g., Jones & Ellison 1991; Malkov & Drury 2001), is expected to be efficient and to produce highly non-equilibrium particle populations. A high acceleration efficiency implies strong coupling between the accelerated particle population, the shock structure, and the electromagnetic fluctuations, with a wide range of scales, responsible for scattering the particles, making DSA a difficult nonlinear (NL) problem.

Efficient DSA produces a hard energy spectrum and the backpressure of these high-energy cosmic rays (CRs) penetrating into the cold incoming plasma creates a precursor with a length-scale on the order of the diffusion length of the highest energy CRs (e.g., Blandford & Eichler 1987). Imbedded in this precursor is a short-scale, viscous subshock that is largely responsible for heating the plasma and injecting particles into the Fermi mechanism. Basic considerations of momentum and energy conservation require that the production of relativistic particles involving the largest magnetic turbulence scales must impact the injection of thermal particles and the structure of the shock on the smallest ion inertial scales making the shock intrinsically multi-scale.

Because hard spectra can be produced in efficient DSA, CRs with the longest diffusion lengths can escape at an upstream boundary and carry away a sizable fraction of the shock ram pressure (e.g., Ellison et al. 1981; Berezhko & Krymskiĭ 1988; Berezhko & Ellison 1999; Caprioli et al. 2010; Ellison & Bykov 2011; Drury 2011; Malkov et al. 2013).5 This escaping energy flux plays an important role in DSA and must be self-consistently included in determining the shock structure. It will cause the overall shock compression ratio to increase and the escaping CR current is certain to generate turbulence that will serve as seed turbulence for compression and amplification as it is overtaken by the shock (e.g., Blandford & Funk 2007; Blasi et al. 2007; Bell et al. 2013).

Global conservation considerations aside, the detailed formation and structure of collisionless shocks is by no means certain. Early on it was suggested that a Weibel-type instability from counter streaming plasma flows could result in the formation of a gas subshock; a narrow region filled with strong magnetic fluctuations deflecting the incoming particles (e.g., Moiseev & Sagdeev 1963; Tidman & Krall 1971). This basic picture was supported later by direct spacecraft observations of heliospheric shocks (e.g., Kennel et al. 1984; Tsurutani & Stone 1985; Gosling et al. 1989).

In addition to observations, hybrid and particle-in-cell (PIC) plasma simulations6 have investigated collisionless shocks coupled with particle acceleration. To our knowledge, the first plasma simulation to show clear evidence of DSA was the parallel-shock, hybrid simulation of Quest (1988). Since then, a great deal of work with both hybrid and full-particle PIC simulations has been done (see Winske & Quest 1988; Winske et al. 1990; Ellison et al. 1993; Giacalone et al. 1997; Kato & Takabe 2008, 2010; Spitkovsky 2008a, 2008b; Treumann 2009; Gargaté & Spitkovsky 2012; Burgess & Scholer 2013; Caprioli & Spitkovsky 2013a and references therein for a sampling of this large body of work). The direct simulation of the microscopic structure of the shock on scales of a few thousand ion inertial lengths7 has been extremely important for resolving the microscopic structure of the plasma shock transition and for clearly demonstrating the initialization of the Fermi acceleration process. However, PIC/hybrid simulations are computationally demanding and the runtime, box size, number of particles, and particle momentum range that can be simulated are limited. In non-relativistic shock simulations, they are currently limited to two or three decades in the dynamical ranges of the spectra of energetic particles and in the k-space of the magnetic fluctuations.

In contrast, models of strong magnetic field amplification (MFA) applicable to SNRs require dynamical ranges for particle momentum and turbulence extending nine or ten decades. Therefore, to capture the NL physics connecting the highest energy CRs to the self-generated, broadband wave turbulence, and to the viscous subshock structure, coarse grained, multi-scale models of the collisionless plasma turbulence generation must be used in addition to the microscopic treatment afforded by PIC/hybrid simulations (see, e.g., Blandford & Eichler 1987; Malkov & Drury 2001; Amato & Blasi 2006; Vladimirov et al. 2008; Kang 2013; Reville & Bell 2013).

Different approaches to model the multi-scale nature of DSA in strong non-relativistic shocks are currently underway. Kinetic semi-analytic models (see, e.g., Malkov & Drury 2001; Völk et al. 2005; Amato & Blasi 2006; Zirakashvili & Aharonian 2010; Bykov et al. 2013b; Lee et al. 2013; Schure & Bell 2013; Reville & Bell 2013) perform self-consistent calculations once the injection rate of background particles into the Fermi mechanism is parameterized. These models use various approximations for the particle diffusion coefficient and the MFA. The background plasma is typically modeled with a diffusion–convection equation which incorporates the magnetic field structure. Some recent time-dependent models have been presented by Kang et al. (2012), Zirakashvili & Ptuskin (2012), and Schure & Bell (2013).

Another approach combines standard MHD equations to describe the background plasma, including the CR current jCR × B term producing the non-resonant hybrid (NRH; i.e., Bell) instability, with a Vlasov–Fokker–Planck equation describing the CR distribution function (see Bell et al. 2013 and references therein).

Here, we use a Monte Carlo technique which allows broadband modeling of NL DSA incorporating both resonant and non-resonant (short- and long-wavelength) current-driven instabilities. This is a plane-parallel, steady-state model extending earlier work by, for example, Jones & Ellison (1991), Ellison et al. (1996), Vladimirov et al. (2006), and Vladimirov et al. (2008, 2009). The model gives a consistent account of the amplification of mesoscopic scale fluctuating magnetic fields interacting with superthermal CRs and the NL feedback of particles and fields on the bulk flow is derived from basic energy-momentum conservation laws. By mesoscale, we mean magnetic turbulence with scales between the very short-scale Weibel-type instabilities that determine the subshock structure and are seen in PIC simulations, and the very large-scale turbulence, e.g., Raleigh–Taylor instabilities, seen in hydro or MHD simulations (e.g., Giacalone & Jokipii 2007; Ferrand et al. 2012; Warren & Blondin 2013). Weibel and Raleigh–Taylor instabilities are not included in our model.

In contrast to semi-analytic models based on the diffusion–advection equation, which make the diffusion approximation and require an independent injection parameter, thermal particle injection in the Monte Carlo model is achieved self-consistently once the particle mean free path is specified (see Section 2.7.4). Particles from thermal energies to energies sufficient to capture the essential NL effects expected from CR production in young SNRs, as well as the magnetic turbulence interacting with these particles, are included self-consistently.

The source of the mesoscopic resonant and non-resonant turbulence in DSA is the anisotropy of the distribution function of the energetic particles. Since the Monte Carlo method does not assume near isotropic distributions, no approximation is made concerning the CR distribution. The CR pressure gradient and current are obtained to all orders in the shock precursor. Due to computational limits, more restrictive approximations may be necessary in models based on the Vlasov–Fokker–Planck equation (e.g., Bell et al. 2013).

We calculate the spectra of the magnetic turbulence using the quasi-linear growth rates described in Section 2.5. The growth rates are derived from a general dispersion relation given by Bykov et al. (2013a). This single dispersion relation accounts for the three main CR-driven instabilities: Bell's short-scale instability (Bell 2004, 2005), the resonant streaming instability (see, for example, Achterberg 1981; Blandford & Eichler 1987), and the long-wavelength, ponderomotive instability (see Bykov et al. 2011b, 2013a; Schure et al. 2012).

The Monte Carlo simulation solves the full DSA/ΔB/smooth–shock problem iteratively. Within this process, we calculate the local growth rates at any iteration using the mean magnetic field and the CR-current derived at the previous iteration. Importantly, between any two iterations ΔB/B < 1 so we converge to an amplified magnetic field $\sqrt{\Delta B(x)^{2}} \,{\gg}\, B_0$, i.e., with rms-amplitudes much larger than the initial magnetic field B0, without violating the quasi-linear approximation (see Section 2.8).8

The spectrum of the mesoscopic magnetic fluctuations depends critically on cascading, i.e., the transfer of turbulent energy from long to short wavelengths (e.g., Vladimirov et al. 2008). If MFA is strong, and local turbulent cascading parallel to the mean field is suppressed, as expected in MHD turbulence, then the turbulence spectrum will contain one or more discrete peaks (see Vladimirov et al. 2009; Bykov et al. 2011a; and Section 2.8 below). We show, however, that critical aspects of DSA, such as the particle distribution function, are relatively insensitive to cascading. This effect is discussed in detail in Section 3.1.

In what follows we emphasize the complex, coupled nature of the problem necessitated by the efficient overall acceleration, the wide dynamic range of the CR distribution function, and the equally broadband, self-generated turbulence produced simultaneously by several instabilities. To highlight these effects, we consider acceleration efficiencies where more than 50% of the shock ram kinetic energy is placed in accelerated particles; combined in trapped and escaping CRs. While trapped and escaping CRs are produced together from the shock ram pressure dissipation they have different properties. The escaping CRs contribute immediately to the galactic CR population while the fate of the trapped CRs depends on the shock evolution, which we do not address in our steady-state, plane geometry model. The trapped particles will eventually leave the SNR but only after losing some fraction of their energy to work done on the magnetic field and expanding plasma. Our model is designed to understand the physical effects of efficient DSA in the local vicinity of an extended supernova shock.

Of particular importance is our treatment of the scattering center speed, vscat(x), relative to the bulk plasma speed, u(x). This is determined from basic conservation considerations without assuming Alfvén waves and we find |vscat| to be significantly below what Alfvén waves would predict. Finally, a limited parameter survey is given and we discuss the parameters that determine the maximum CR energy a given shock can produce as well as various observational consequences related to the magnetic turbulence spectra we calculate.

2. MODEL

The basic elements of our plane-parallel, steady-state, Monte Carlo method have been discussed in a number of previous papers (e.g., Jones & Ellison 1991; Ellison et al. 1996, 2013). Support for its general validity comes from detailed comparisons with in situ spacecraft observations of particle acceleration at heliospheric shocks (e.g., Blandford & Eichler 1987; Ellison et al. 1990b; Baring et al. 1997), as well as from direct comparison with hybrid plasma simulations (Ellison et al. 1993). While early versions of the Monte Carlo code assumed simple forms for the particle diffusion, we continue here an extensive generalization of the code to include magnetic turbulence generation and self-consistent diffusion (e.g., Vladimirov et al. 2006, 2008, 2009).

The Monte Carlo code includes the following main elements.

  • 1.  
    thermal leakage injection self-consistently coupled to DSA where some fraction of shock-heated thermal particles re-cross the subshock and gain additional energy;
  • 2.  
    shock-smoothing where the pressure from superthermal particles in the shock precursor slows and heats the incoming plasma upstream of the viscous subshock;
  • 3.  
    CR escape at an upstream free escape boundary (FEB);
  • 4.  
    a determination of the overall shock compression ratio Rtot, taking into account escaping CRs and the modification of the equation of state from the production of relativistic particles;
  • 5.  
    fluctuating magnetic fields simultaneously calculated from resonant, short-, and long-wavelength instabilities generated from the CR current and pressure gradient in the shock precursor;
  • 6.  
    momentum and position dependent particle diffusion calculated from the self-generated magnetic fields;
  • 7.  
    a consistent determination of the local scattering center speed relative to the bulk plasma; and,
  • 8.  
    NL spectral energy transfer (i.e., cascading) and dissipation of wave energy into the background plasma. All of these processes are coupled together making a reasonably consistent, albeit complicated, model.

In all that follows, the bulk plasma flow is in the positive x-direction with speed u(x) (see Figure 1). The unmodified (i.e., far upstream) shock speed is u0 and the downstream plasma speed, in the shock frame, is u2 = u0/Rtot. In modified shocks, there is a well-defined subshock compression Rsub = u1/u2, where u1 is the plasma speed immediately upstream of the viscous subshock. Everywhere, the subscript 0 (2) implies far upstream (downstream) values.

Figure 1.

Figure 1. Dashed (red) curves show the results for an unmodified shock with Rtot ≃ 4 (Model UM). The solid (black) curves (Model A) show the self-consistent result where the momentum and energy fluxes are conserved across the shock. For this example, where all three instabilities are active, the self-consistent compression ratio is Rtot ≃ 11.3 and ∼40% of the energy flux is lost at the FEB at x = −108rg0, where rg0mpu0c/(eB0) ≃ 5.6 × 10−9 pc. Both models have u0 = 5000 km s−1, n0 = 0.3 cm−3, and B0 = 3 μG. Note the split log-linear x-axis.

Standard image High-resolution image

2.1. Small Angle Scattering

The position and momentum dependent scattering mean free path, λmfp(x, p), of a particle is determined from the local diffusion coefficient D(x, p). A particle moves for a time δt and then scatters elastically and isotropically in the local frame through a small angle $\theta _{\rm scat}\le \sqrt{6 \delta t/t_c}$, where tc = λmfp/vp  ≫  δt (see Ellison et al. 1990a). At this new x-position, the particle will have changed momentum because of the converging bulk flow. For the next δt, λmfp is updated from the new local D(x, p) and the scattering process is repeated: the particle executes a random walk on a six-dimensional sphere in space and momentum in an ever changing magnetic and bulk flow background. This continues until the particle leaves the shock either by convecting far downstream or escaping at the upstream FEB.

2.2. Thermal Leakage Injection

A well-defined, essentially discontinuous subshock with a subshock compression ratio Rsub < Rtot is always present in our smooth-shock solutions.9 As thermal particles injected upstream cross the subshock and scatter in the downstream flow they gain enough energy so vp  >  u2 for some fraction of the population depending on the shock Mach number. By virtual of random scatterings in the downstream region, some vp  >  u2 particles will re-cross the subshock into the precursor, gain additional energy, and enter the first-order Fermi acceleration process. This simple form of thermal leakage injection assumes the subshock is transparent and ignores the possible existence of a cross-shock potential or other effects from large amplitudes waves that may occur at the subshock (see, for example, Kato & Takabe 2010 for PIC results showing the shock transition region).

The injection scheme requires no superthermal seed particles and is entirely defined within the Monte Carlo scattering assumptions of Section 2.1. Since the subshock strength and the downstream flow speed (via Rtot) are determined self-consistently with the global shock properties, the injection rate is coupled to DSA through the conservation conditions we discuss below. No "injection parameter" is needed to model injection from the thermal population.

2.3. Mass–Energy–Momentum Conservation

In our steady-state, plane-parallel shock, the conservation of mass flux is given by

Equation (1)

where ρ(x) is the plasma density and ρ0u0 is the far upstream mass flux. The momentum flux conservation is given by

Equation (2)

where $\Phi ^{\rm part}_P(x)$ is the particle momentum flux, Pw(x) is the momentum flux carried by the magnetic waves, and ΦP0 is the far upstream momentum flux, i.e., upstream from the FEB where the interstellar magnetic field is BISM.10

Separating the contributions from the thermal and accelerated particles we have

Equation (3)

where Pth(x) is the thermal particle pressure and Pcr(x) is the accelerated particle pressure. A particle is "accelerated" if it has crossed the subshock more then once and even though we use the subscript "CR," the vast majority of accelerated particles will always be non-relativistic. Of course, if the acceleration is efficient, a large fraction of the pressure may be in relativistic particles.

The energy flux conservation law is

Equation (4)

where $\Phi ^{\rm part}_E(x)$ and Fw(x) are the energy fluxes in particles and magnetic field correspondingly, and ΦE0 is the energy flux far upstream. Taking into account particle escape at an upstream FEB, this can be re-written as

Equation (5)

where Fth(x) is the internal energy flux of the background plasma, Fcr(x) is the energy flux of accelerated particles, and Qesc is the energy flux of particles that escape at the upstream FEB (note that Qesc is defined as positive even though CRs escape moving in the negative x-direction).

As seen in Figure 6 in Ellison et al. (1990b) or Figure 8 in Vladimirov et al. (2006) or Figure 2 in Caprioli & Spitkovsky (2013a), to give just three examples, the separation between "thermal" particles and "accelerated" particles in a shock undergoing DSA is not necessarily well defined. Furthermore, energy exchange between the thermal and superthermal populations is certain to occur through non-trivial wave-particle interactions. Nevertheless, the bulk of the plasma mass will always be in quasi-thermal background particles and the internal energy flux of this background plasma can be expressed as

Equation (6)

where γg = 5/3 is the adiabatic index of the background plasma.

2.4. Magnetic Turbulence Generation and Dissipation

We calculate the magnetic turbulence generated by the CR current, Jcr, and pressure gradient, dPcr/dx, which simultaneously drive resonant, short-, and long-wavelength instabilities.

2.4.1. Energy Balance Equations

The spectral energy density of the magnetic fluctuations W(x, k), where Wdk is the amount of energy in the wavenumber interval dk per unit spatial volume, can be calculated including cascading. The energy balance equation is

Equation (7)

where Π(x, k) is the flux of magnetic energy through k-space toward larger k, and G(x, k) and ${\cal L}(x,k)$ are the spectral energy growth and the dissipation rates, respectively. The turbulence energy flux and pressure are given by

Equation (8)

and

Equation (9)

where the energy flux and pressure from the CR-current driven fluctuations are given in Equations (10) and (11) below. Equation (7) differs from Equation (1) in Vladimirov et al. (2009) in that we now explicitly include adiabatic compression of the magnetic turbulence energy density.

The energy flux and pressure from the CR-current driven fluctuations that are included in Equations (3) and (5) are defined as

Equation (10)

and

Equation (11)

For the case of CR-driven modes in a highly conducting plasma with frozen-in magnetic fields, the velocity and field amplitudes of a harmonic perturbation are connected through the relation

Equation (12)

where φ(k), with a complicated derivation, can be determined from the dispersion Equation (20) given below and the mode polarizations. For simplicity, we only present results for φ(k) = 1 (as is the case for Alfvén modes), to be compared with those in Vladimirov et al. (2009), but we have verified that the resulting spectra of magnetic fields and particles do not vary significantly for the range 0.1 < |φ(k)| < 10.

Integrating Equation (7) from kmin to kmax we obtain

Equation (13)

The limits kmin and kmax are chosen to be outside the range containing a maximum scale determined by MFA and a minimum scale determined by dissipation. For these limits, the Π(x, k) term in Equation (7) vanishes.

For Kolmogorov-type cascading, the energy flux is redistributed over the energy scale but the total energy flux density, integrated over wavenumbers is conserved. Therefore, the total dissipation rate at any position x is

Equation (14)

We note that Kolmogorov-type cascade models have been successful at explaining the spectra of locally isotropic, incompressible turbulence in non-conducting fluids observed in experiments and simulations (e.g., Biskamp 2003). However, it is uncertain how spectral energy transfer operates in a collisionless shock precursor with a CR current strong enough to modify the MHD modes, and in the presence of strong magnetic turbulence. In weak MHD turbulence, cascading was shown to be anisotropic (e.g., Goldreich & Sridhar 1997): harmonics with wavenumbers transverse to the mean magnetic field experience a Kolmogorov-like cascade, while the cascade in wavenumbers parallel to the mean field is suppressed.

2.5. Growth Rates of CR-driven Instabilities

Self-generated magnetic turbulence is a key ingredient for DSA and the production of galactic CRs (e.g., Bell 1978; Blandford & Eichler 1987; Jones & Ellison 1991; Malkov & Drury 2001; Parizot et al. 2006; Bykov et al. 2012; Schure et al. 2012). Most analytical models that calculate MFA (e.g., Bell 2004; Pelletier et al. 2006; Amato & Blasi 2009; Bykov et al. 2011b, 2013a) assume the following "homogeneous" form for the CR distribution function in the local rest frame of the upstream flow:

Equation (15)

where ncr is the concentration of CRs, u0 is the shock velocity, p is the particle momentum in the plasma frame, μ = cos θ, and θ is the particle pitch angle, i.e., the angle between p and the magnetic field assumed to lie in the x-direction. The CR distribution function, N(p), is normalized as $\int _{p_{\rm min}}^{p_{\rm max}} N(p) p^2 dp = 1$, where pmin and pmax are the minimum and maximum particle momenta in the CR distribution at the upstream position. In analytic or semi-analytic treatments, ncr and u0 are normally assumed to be position independent and N(p) is often assumed to be a power law.

In the Monte Carlo code, the particle distribution function contains more information than represented by Equation (15). It is calculated directly in the modified shock precursor with a varying ncr(x) and u(x), rather than constant values, and no approximations are made restricting the particle anisotropy. Instead of f0(p), we can use

Equation (16)

to calculate the local growth rates. Here, Jcr(x, p) is the differential CR current, vp is the particle velocity in the local frame, and N(x, p) is determined self-consistently with the shock structure and will not be a power law in the shock precursor. The full CR current (for now we consider only protons) is

Equation (17)

and this is used to calculate the CR-driven instabilities. Note that Jcr(x, p) only contains superthermal particles and pmin depends on x, increasing as the observation position moves further upstream from the subshock.

At any x-position in the precursor, we use local Monte Carlo values for ncr(x), u(x), and Jcr(x) to calculate the dispersion relation, ignoring effects from spatial gradients in our "local homogeneous" approximation. Specifically, the growth rates for the three CR-driven instabilities are derived from the dispersion relation for modes propagated along the mean magnetic field in the magnetized background plasma. This dispersion relation, which includes the short-wavelength Bell instability (i.e., Bell 2004), was derived in Bykov et al. (2011b, 2013a), and is

Equation (18)

where ρ is the background plasma density, and the ± signs, here and in the equations that follow, correspond to the two circularly polarized modes. The mean magnetic field, Bls(x, k), in the dispersion relations is defined as the sum of the long-wavelength (i.e., large-scale "ls") harmonics plus the ambient homogeneous field, B0, i.e.,

Equation (19)

The definitions for A(x, k) (Equation (23)) and k* (Equation (29)) are given below and αt and κt are defined in Section 5.1 of Bykov et al. (2011b).

The solution to Equation (18) is

Equation (20)

In Equations (18) and (20),

Equation (21)

Equation (22)

Equation (23)

Equation (24)

In these equations, $v_a = B_{\rm ls}(x,k)/\sqrt{4 \pi \rho (x)}$, z = kcp/(eBls), a is the collision parameter equal to the ratio of the gyroradius of a CR particle to its mean free path determined by scattering on magnetic fluctuations, and kc = 4π|Jcr(x)|/[cBls(x, k)] is Bell's critical wavenumber at position x in the precursor.

The effective, self-generated, magnetic field is given by

Equation (25)

or equivalently,

Equation (26)

The total field is Btot = Bls(x, kmax) so $B_{\rm tot}^2 = B_{\rm eff}^2 + B_0^2$. As mentioned above, we assume the ambient interstellar medium (ISM) field consists of a uniform component B0 = 3 μG, and a turbulent part generated from the background Kolmogorov turbulence assumed to have a strength such that BKolm = 3 μG. Therefore, for the ISM, Btot ≃ 4.2 μG.

2.5.1. Long-wavelength Instability

The plasma fluctuations created by Bell's short-wavelength instability influence the plasma dynamics and this effect is modeled with the ponderomotive coefficients, αt and κt, in Equations (21) and (22).11 These ponderomotive effects result in the long-wavelength plasma instability (LWI) developed by Bykov et al. (2011b, 2013a).

The ponderomotive coefficients are determined by the mean square of the short-scale magnetic field fluctuations produced by Bell's instability, i.e.,

Equation (27)

and

Equation (28)

where ξ is the dimensionless mixing length of the short-scale turbulence as defined in Bykov et al. (2011b), va is the Alfvén speed which is the characteristic speed of the medium, and NB, the dimensionless amplitude of the magnetic fluctuations, is defined below. While NB is determined in the Monte Carlo simulation from the CR distribution, ξ is not. In the simulations reported here, we take ξ = 5 but note that our results are only weakly dependent on ξ.

To achieve a solution, we set the ponderomotive coefficients in the dispersion Equation (20) equal to zero for wavenumbers k > k*, where k* is determined by the effective resonant condition

Equation (29)

At any x-position we set NB(x) = 0 for k*(x) > kc(x), where there is no wave growth from Bell's mode, and we set

Equation (30)

for k*(x) < kc(x, k*), where Bell's instability operates. The mode growth rates, Γ(x, k), in the turbulence energy balance Equation (13), where G is the energy growth rate (see Equation (58) below), are connected to the roots of the dispersion equation as

Equation (31)

where the 2 accounts for the fact that the energy in turbulence is proportional to the square of the amplitude of ΔB. At each x-position we choose the mode with the maximum value of Im[ω(x, k)], if positive.

Because we use the full, anisotropic CR distribution function in Equation (18) and the substitutions that follow, the dispersion Equation (20) simultaneously accounts for the CR-pressure driven resonant streaming instability, and the two CR-current driven instabilities. To our knowledge, this is the first attempt to combine these instabilities in a consistent, broadband shock model.12

2.5.2. Dissipation and Fluxes of Particles and Waves

The self-generated turbulence is assumed to suffer viscous dissipation at a rate proportional to k2, and the dissipated energy is pumped directly into the thermal particle background. The background plasma energy balance is governed by the plasma compression and the turbulence dissipation rate and obeys the equation

Equation (32)

If the heating of the background plasma by the turbulence dissipation is weak (i.e., L(x)  ∼  0), Equations (1), (6), and (32) result in the adiabatic compression of the background plasma.

The results of Vladimirov et al. (2008), which included only the resonant CR-streaming instability for MFA with a variable dissipation rate, demonstrated that even a modest rate of turbulence dissipation can significantly increase the precursor temperature (see Figure 15 below) and that this, in turn, can increase the rate of injection of thermal particles (see also Vladimirov et al. 2009). However, the NL feedback of these changes on the shock structure tend to cancel so that the spectrum of high-energy particles is only modestly affected. As described in Vladimirov et al. (2009), we only apply dissipation to our models with cascading.

The relation between the CR pressure gradient and the CR energy flux Fcr, can be written as

Equation (33)

where vscat(x) is the scattering center velocity measured in the local frame.13 This definition is a position dependent generalization of the position independent "wave frame" velocity introduced by Skilling (1971) for CR interactions with hydromagnetic waves. An important feature of the CR-current driven modes is that, in the most interesting cases with Im[ω(k, x)]  >  kva, the scattering center velocity must be substantially less than the Alfvén speed va in order to achieve energy conservation. This is true whether B0 or Bls is used to calculate va and is discussed in detail in Section 2.6.

As follows from Equation (10), the energy flux ${\cal F}_w(x,k)$ may span the range from W(x, k) to 2 W(x, k), for φ(k) between 0 and 1. In our Monte Carlo simulations, we use φ(k) = 1 but our results are not sensitive to φ(k). For resonantly generated Alfvén modes, McKenzie & Voelk (1982) derived

Equation (34)

and

Equation (35)

which corresponds to φ(k) = 1, since the kinetic and magnetic energy densities are exactly equal for Alfvén modes.

2.6. Effective Scattering Center Velocity

The turbulence produced by shock accelerated particles can move relative to the bulk plasma and this movement must be self-consistently included when determining the NL shock structure. If the turbulence is assumed to be Alfvén waves, the scattering center speed can be taken to be the Alfvén speed with the far upstream field, $v_{a0}(x) = B_0 / \sqrt{4 \pi \rho (x)}$, and the turbulence can be calculated in Equation (13) in a straightforward fashion using

Equation (36)

to model the wave growth.

Equation (36) was obtained by McKenzie & Voelk (1982) assuming (1) that the resonant wave-CR particle interactions are quasi-linear with λ∝1/W(x, k), and (2) that the magnetic turbulence is dominated by quasi-linear modes with growth rates Γ(x, k) ≪ kva0 (for simplicity we only consider modes propagating parallel to the mean magnetic field), (3) that the mode growth rate is a linear function of the isotropic part of the distribution function f(x, p). One ambiguity with this approach is the choice of B. It is typically chosen as either B0 or some fraction of the amplified magnetic field at x.

In considering Equation (36), however, it must be noted that non-resonant, CR current-driven modes (e.g., Bell 2004; Bykov et al. 2011b; Schure et al. 2012; Bykov et al. 2013a) have their fastest growth rates when Γ(x, k)  >  kva0 and these modes dominate the magnetic fluctuation spectra. This point can be illustrated in a simple way. Consider just the growth rate of Bell's instability without the LWI. Then, the solution to Equation (18) is

Equation (37)

where K(k, Jcr) is determined by the CR current. This equation can be obtained directly from Equation (20) by setting αt = κt = 0, the two parameters responsible for the LWI. To get Equation (36), two conditions must hold. The first is that $\left|K\right|\ll v_{a}^{2}k^{2}$ and therefore the square root in Equation (37) can be expanded as

Equation (38)

The second condition is that Γ(x, k) is a linear function of the distribution function f(x, p). Note that G(x, k) = Γ(x, k)W(x, k) is generally assumed in our approach.

While both of these conditions can be fulfilled in the case of weakly growing Alfvén-like turbulence, weak growth is not expected if the CR current is as large as believed to occur in young SNRs. If the CR current is large, and non-resonant current driven instabilities are important, then $\left|K\right|\,{\gg}\, v_{a0}^{2}k^{2}$ and Equation (37) simplifies to

Equation (39)

that is, ω is proportional to the square of K rather than proportional to K. While K(k, Jcr) is a linear function of f(x, p) (if the diffusion approximation is valid), it is clear from Equation (39) that the wave growth term G(x, k) is a NL function of f(x, p) for large CR currents.

We believe this reevaluation of the self-generation of turbulence by non-resonant modes is fundamentally important. The non-resonant turbulence has a character quite different from Alfvén waves14 and we find that Equation (36), regardless of the choice of B, does not allow for momentum and energy conserving solutions. With the Monte Carlo technique, we obtain consistent solutions by simply generalizing Equation (36) to

Equation (40)

assuming there is a single, position dependent, effective scattering speed, and including vscat(x) in our iterative scheme. Equation (40) is derived using Equations (1), (3), (5), (13), (33), and the equation of state of the background plasma.

At each iteration, the ith + 1 value of the scattering center velocity, $v_{\rm scat}^{(i+1)}(x)$, is obtained from

Equation (41)

where Γ, W, and dPcr/dx are all determined in the Monte Carlo simulation in the ith iteration. With Equation (41) there is no need to associate vscat with the Alfvén speed. In the upstream region, accelerated particles are propagated through the shock assuming the scattering center velocity is u(x) + vscat(x), while the velocity is u(x) for the background thermal particles. Downstream from the shock, we take vscat = 0.

An essential and unique element of our calculation is that when the integral in Equation (40), with vscat determined from Equation (41), is used to replace the integral in the energy balance Equation (7), we make use of the full anisotropic CR distribution function including the pressure gradient and the CR current. Our derivation of vscat(x) accounts for the anisotropic magnetic modes with dispersive properties (phase and group velocities) determined by both the background plasma and the CR angular and momentum distributions. As we show below (see Figure 18), while the fastest growing CR-driven modes are highly anisotropic, their phase and group velocities are typically strongly sub-Alfvénic with vscat(x) ≪ u(x) for all x.

Significantly, even though vscat(x) may be small, it has a strong influence on particle acceleration and must be taken into account to determine a consistent shock structure in the Monte Carlo model. Despite its wide use for many years, we caution that replacing the wave-growth term in Equation (13) with Equation (36) is a poor approximation when short- and long-wavelength instabilities are taken into account.

2.7. Particle Mean Free Paths

The Monte Carlo code determines λmfp(x, p) using various analytic approximations and the locally averaged magnetic field. As indicated in Figures 6 and 7, we define different regimes for determining λmfp starting with thermal particles (see Vladimirov et al. 2009 for additional details).

2.7.1. Thermal Particles

Thermal particles enter the simulation upstream of the subshock at a position which depends on the plasma flow velocity gradient and then propagate toward the subshock. Far upstream, and during the first iteration when the shock is unmodified and before MFA has generated turbulence in the precursor, these thermal particles with momenta pth experience relatively weak turbulence and have a relatively large λmfp = rg, where rgpthc/(eBls).

It subsequent iterations, when strong turbulence exists in the precursor, the λmfp for low-energy particles can become considerably smaller. At this point, we take into account the fact that the transport of low-energy particles may not be diffusive but can be governed by turbulent advection of particles frozen into large-scale turbulent plasma vortexes (see Bykov & Toptygin 1993 for more details).

2.7.2. Vortex Advection for Low-energy Particles

The Monte Carlo simulation describes turbulence on mesoscopic scales. At the low end of this mesoscopic range, turbulence, particularly if produced with a quasi-power-law spectrum by cascading, may result in low-energy particles having mean free paths smaller than the scale of vortexes that are expected to develop. In this case, the low-energy particles can be "trapped" by the vortexes and execute non-diffusive transport. We have developed a transport model that mimics the essential physics for vortex advection in and near the viscous subshock layer and this is included in our model.

Following the discussion in Vladimirov (2009), we assume that low-energy particles can be confined by resonant scattering and trapped within turbulent plasma vortexes of different scales. In this case, the transport of these particles on scales greater than the correlation length of the turbulence (i.e., greater than the largest turbulent harmonics), is governed by the turbulent advection of the vortexes rather than by resonant diffusion of individual particles. This vortex trapping mimics how low-energy particles would interact with local displacements and distortions of a shock front moving through the turbulent ISM. If the Bohm diffusion coefficient is small enough, particles can be considered to be "anchored" to a small section of the shock front and it was shown by Bykov (1982) that there is a "diffusion regime" in space and time of the average displacement of a section of the subshock surface. The low-energy particles anchored to this section have an effective transport that is diffusive.

Our recipe for the turbulent transport of low-energy particles can be summarized as follows: the particle diffusion coefficient due to turbulent vortex advection, Dvor(x), is momentum independent and determined by

Equation (42)

where Uvor, the typical speed of turbulent motions with correlation length lvor, is estimated from

Equation (43)

Here, $k_{\rm vor}(x)=2\pi l_{\rm vor}^{-1}(x)$ and kmax is determined by the turbulence dissipation mechanism. For concreteness we take

Equation (44)

and determine the mean free path from vortex motion from

Equation (45)

This "convective" diffusion coefficient is indicated with the label "1" in Figures 6 and 7.

For "trapped" low-energy particles, Dvor(x) can be much greater than the resonant scattering coefficient. Low-energy CRs in the vicinity of the subshock have a high frequency of scattering and because of this they are tied to the large-scale vortexes. Their transport is dominated by the convection of the turbulence instead of microscopic scattering off waves. At every small-angle-scattering event, we compare the mean free path from magnetic fluctuations, λ(x, p) (discussed in Section 2.7.4 below), to λvor(x). The larger of the two is used to propagate the low-energy particles. This will modify the injection process but injection will still be self-consistently determined in the Monte Carlo simulation as the shock structure adjusts to conserve momentum and energy. Our results are not strongly dependent on the scattering assumptions made for low-energy particles.

2.7.3. Particle Scattering by Short-scale Fluctuations

The wave number kres associated with resonant interactions of particles with momentum p is given by

Equation (46)

Our model includes the short-scale turbulence produced by CR-driven instabilities where the wave number of the short-scale modes kss  ≫  kres and the modes are defined by

Equation (47)

For particles with gyroradii rg = cp/(eBls) much larger than the scale-length of the short-scale turbulence, the mean free path produced by the short-scale modes is

Equation (48)

where rss = cp/(eBss),

Equation (49)

and

Equation (50)

It has been shown that the short-scale scattering regime with λssp2 holds even for large amplitude magnetic field fluctuations (see, for example, Jokipii 1971; Toptygin 1985). This mode is particularly important because it dominates particle scattering for the highest energy CRs a given shock can produce.

2.7.4. Effective Particle Mean Free Path

The total effective scattering mean free path is

Equation (51)

where

Equation (52)

The diffusion regimes included in Equation (52) are consistent with those seen in numerical simulations of particle transport in strong magnetic turbulence (e.g., Casse et al. 2002; Marcowith et al. 2006; Reville et al. 2008).

For low-energy particles we define a transition momentum ptran = ftranmpu0 where ftran > 1 is a free parameter and λpic(x, p) is the Bohm diffusion length, defined here by15

Equation (53)

For particles with momenta p > ptran, the mean free path is determined by

Equation (54)

In the simulations reported here, we take ftran = 3.0.

The mean free path due to quasi-resonant scattering is

Equation (55)

and lcor is defined as $\min [L_{\rm ls}(x,p),k_{\rm ls}^{-1}(x,p)]$ where

Equation (56)

Here kls is the maximum wavenumber satisfying the relation $ k_{\rm ls}W(x,k_{\rm ls})>\int _{k_{\rm min}}^{k_{\rm ls}}W(x,k^{^{\prime }})dk^{^{\prime }}$, for kls < kres. If $ k_{\rm ls}W(x,k_{\rm ls})<\int _{k_{\rm min}}^{k_{\rm ls}}W(x,k^{^{\prime }})dk^{^{\prime }}$ for any kls in the range kmin < kls < kres, then kls = kmin.

The parameter ftran is required because the Monte Carlo technique does not model production of the short-scale turbulence responsible for formation of the viscous subshock, i.e., that produced by the Weibel instability. While PIC simulations can do this, they cannot yet cover the wide dynamic range needed to produce the high-energy CRs responsible for the large-scale turbulence and cascading that produce the vortex transport we described in Section 2.7.2. Parameterizing the transition between subshock and vortex advection is currently the best way to describe the transport of superthermal particles until they reach p  >  ptran and Equation (52) can be used. We note that our results are sensitive to ptran for low shock speeds (u0  ∼  1000 km s−1) but become less sensitive as u0 increases.

2.8. Monte Carlo Iterative Solution

The coupled components of our steady-state, plane-shock simulation are solved iteratively. Given that the particle acceleration process generates full particle spectra f(x, p), CR currents, and CR pressure gradients, at all positions relative to the subshock, we can write the spectral energy density at the ith iteration, W(i)(x, k), as16

Equation (57)

Equation (57) is a discretized version of Equation (7) where we assume that the small magnetic turbulence increment between any two iterations can be estimated using the quasi-linear growth rate of the turbulence. Then, the magnetic turbulence growth rate

Equation (58)

is derived using the CR-current and the mean magnetic field $B_{\rm ls}^{(i-1)}(x,k)$ (which is defined by Equation (19)) with W(i − 1)(x, k) derived on the ith − 1 iteration. This approach is somewhat similar to mean field models used in the statistical theory of ferromagnetism (e.g., Kittel 1976), and while it neglects some correlation effects and we assume randomization of the field direction, we contend it overcomes the major limitations of quasi-linear theory. We use the quasi-linear approximation but we only apply Equation (58) between iterations where $\Delta B^{(i-1)} < B_{\rm ls}^{(i-1)}$. This allows us to slowly increase Bls to values Bls  ≫  B0 which are currently beyond any exact simulation method without violating ΔB < Bls in any one iteration step.

For a given set of shock parameters, we start with an unmodified shock with compression ratio, RRH, determined by the Rankine–Hugoniot conditions, vscat(x) = 0, and λth = rg, th = pthc/(eB0), where pth is the thermal particle momentum. The initial magnetic turbulence is taken to be

Equation (59)

Thermal particles are injected far upstream and diffusively accelerated, leaving the shock by convecting far downstream or escaping at the upstream FEB. After the first iteration, f(x, p) is determined, along with Jcr(x) and dPcr(x)/dx, and these are used to calculate Γ(x, k) and W(x, k) for the next iteration. From these we determine Fw(x), Pw(x), D(x, p), and vscat(x). We include cascading and energy dissipation which transfers energy from W(x, k) to the background plasma influencing the subshock strength and particle injection.

The momentum, $\Phi ^{\rm ({i})MC}_P(x)$, and energy, $\Phi ^{\rm ({i})MC}_E(x)$, fluxes from all particles are calculated directly in the Monte Carlo simulation for the ith iteration. To these particle fluxes we add the wave components $P_w^{(i)}(x)$ and $F_w^{(i)}(x)$ and check to see if the total momentum and energy fluxes (i.e., Equations (3) and (5)) are conserved to within some limit at all x. If the fluxes are not conserved in the ith iteration, we use Equation (3) in the form

Equation (60)

to predict the shock speed profile, u(i + 1)(x), for the ith + 1 iteration.

When relativistic particles are produced and/or high-energy particles escape at an upstream FEB, the overall compression, Rtot, will increase above RRH and Rtot must be found by iteration simultaneously with u(x) and vscat(x). A consistent solution, within some statistical uncertainty, will conserve momentum and energy fluxes at all x, including a match to the escaping energy flux, Qesc in Equation (5). Despite the complexity of this system, with several processes all coupled nonlinearly, we are able to obtain unique modified shock solutions covering a wide dynamic range, with self-consistent injection, MFA, and an overall compression ratio consistent with particle escape.

3. RESULTS

In our solutions, the bulk flow speed, u(x), overall compression ratio, Rtot, and CR induced magnetic turbulence are calculated such that the momentum and energy fluxes are conserved across the shock, as shown by the solid (black) curves in Figure 1 (Model A in Table 1). These fluxes include the magnetic field contributions and account for the escaping energy flux at the FEB. The drop in energy flux seen in the solid curve at x ∼ −108rg0 in the bottom left panel of Figure 1 is a direct measure of the energy flux escaping at the FEB.

Table 1. Model Parameters

Model Inst.a u0 n0 LFEBb MS: MA Rtot:Rsub Pcr, 2c Pw, 2c Qescd Beff, 2 T2
(km s−1) (cm−3) (pc) (%) (%) (%) (μG) (106 K)
UMe ... 5000 0.3 0.56 430: 420 4.0: 4.0 ... ... ... 3 490f
A B, L 5000 0.3 0.56 430: 420 11.3: 2.85 73 9 37 540 30f
B B, L, C 5000 0.3 0.56 430: 420 13.3: 2.52 80 3.5 47 330 20f
C B, C 5000 0.3 0.56 430: 420 12.5: 2.37 79 3.5 42 340 20f
D B 5000 0.3 0.56 430: 420 10.2: 2.84 69 10 35 570 35f
E B, L 2500 0.3 0.11 210: 210 11.7: 2.81 76 6 41 220 6f
F B, L 5000 0.3 0.11 430: 420 11.0: 2.83 72 10 37 560 30f
G B, L 10000 0.3 0.11 850: 840 8.3: 2.85 61 12 23 1250 220f
H B, L 20000 0.3 0.11 1700: 1670 6.93: 2.87 52 14 15 2600 1400f
I B, L 5000 1 0.11 430: 760 10.3: 2.88 69 11 36 1100 35f
J B, L 5000 3 0.11 430: 1300 9.2: 2.92 65 11 29 1900 45f
K B, L 5000 10 0.11 430: 2400 8.65: 3.01 62 13 25 3700 50f
L B, L 5000 30 0.11 430: 4200 8.58: 3.10 60 14 23 6500 55f
M B, L 5000 0.3 0.34 430: 420 11.0: 2.83 72 9 39 520 30f
N B, L 1000 0.3 0.11 8.5: 84 6.78: 1.45 66 1.5 12 41 3.3g
P B, L 30000 0.3 0.11 2560: 2500 6.33: 2.85 48 13 11 3900 3800f
Q B, L, C 1000 0.3 0.11 8.5: 84 6.72: 1.33 66 0.3 19 19 3.5g
R B, L, C 2500 0.3 0.11 21: 210 10.4: 1.94 75 1.0 31 90 7.9g
S B, L, C 5000 0.3 0.11 43: 420 11.9: 2.16 79 3 44 315 24g
T B, L, C 10000 0.3 0.11 85: 840 10.9: 2.34 75 5.3 39 820 110g
U B, L, C 20000 0.3 0.11 170: 1700 9.15: 2.43 68 8 28 2000 710g
V B, L, C 30000 0.3 0.11 260: 2500 8.0: 2.47 63 8.5 23 3100 2200g
W B, L 30000 0.3 0.11 260: 2500 6.33: 2.85 48 13 11 3900 3800g
X B, L 2500 0.3 0.11 21: 210 10.5: 2.13 75 5 39 200 7.3g
Y B, L 5000 0.3 0.11 43: 420 10.5: 2.6 72 8.4 37 520 29g
Z B, L 10000 0.3 0.11 85: 840 8.31: 2.79 61 12.4 23 1250 215g
AA B, L 20000 0.3 0.11 170: 1700 6.84: 2.83 52 13 14 2600 1400g

Notes. For all models, B0 = BKolm = 3 × 10−6 G and we note that the derived results for all runs have a statistical uncertainty of ≲ 5%. aThe letter B stands for Bell's instability, L stands for the long-wavelength instability, and C indicates that cascading is included. All models include the resonant instability. bThe distance LFEB is determined from a given number, N, of rg0 such that LFEB ≃ 1.1 × 10−9N(u0/1000 km s−1)(3 μG/B0) pc. cPercent of far upstream momentum flux, ΦP0. dPercent of far upstream energy flux, ΦE0. eModel UM is an unmodified shock where energy and momentum are not conserved. fUpstream temperature, T0 = 104 K. gUpstream temperature, T0 = 106 K.

Download table as:  ASCIITypeset image

The dashed (red) curves in Figure 1 (Model UM in Table 1) show the shock structure and fluxes for the same input parameters without shock smoothing, adjustment of Rtot, or MFA. In the UM case, Rtot = RRH ≃ 4, versus Rtot ≃ 11.3 in the NL case, and the momentum and energy fluxes are ∼100 times above the conserved values throughout much of the shock. These quantitative results, of course, depend on the efficiency of DSA which, in turn, depends on the details of our model, i.e., on the injection scheme, the MFA description, and the calculation of the scattering mean free path from the turbulence. Nevertheless, it is essential to note that any consistent description of efficient DSA with a diffusion coefficient that is an increasing function of momentum must produce a shock structure similar to the solid (black) curves shown in Figure 1.

In the self-consistent solution, the shock structure develops a distinct subshock with compression ratio Rsub < Rtot, as indicated by u1 in the top panel of Figure 1. The subshock is responsible for most of the heating of the ambient material and in this case, Rsub = u1/u2 ≃ 2.85. In order to have a consistent solution with efficient DSA, the plasma heating must be reduced compared to the UM case, as reflected by Rsub < Rtot.

3.1. Magnetic Field Fluctuations Spectra

Our model gives a thorough accounting of the magnetic fluctuation amplification produced simultaneously by the three CR-current instabilities as discussed in Section 2.5. In Figure 2 we show the converged flow speed profile along with the CR current Jcr, CR pressure gradient dPcr/dx, CR pressure Pcr, and the effective magnetic field Beff that results from MFA. Note that Beff is obtained from Equation (26) and does not include the homogeneous part of the ISM field. The dashed (red) curves (Model B in Table 1) include cascading while the solid (black) curves (Model A) do not. From the (d) and (e) panels it can be seen that cascading reduces the large-scale magnetic field by about a factor of two for x ≳ −106rg0 and causes the CR pressure to increase by ∼10% over the same range. In both cases, the shock structure (panel a) adjusts so momentum and energy are conserved through the shock.

Figure 2.

Figure 2. Top panel (a) shows the bulk flow speed as in Figure 1, panel (b) shows the CR current, panel (c) shows dPcr/dx in cgs units, panel (d) shows the CR pressure, and the bottom panel (e) shows the effective magnetic field derived from Equation (25). The dashed (red) curves include cascading (Model B) while the solid (black) curves (Model A) do not. Both models have u0 = 5000 km s−1, n0 = 0.3 cm−3, and B0 = 3 μG.

Standard image High-resolution image

In Figure 3 we show Γ(x, k)/(rg0) versus krg0 as calculated from Equation (20) for Model A. Here, Γ(x, k) = 2 Im[ω(x, k)] is the growth rate of magnetic energy of the modes. The curves are calculated at different positions in the precursor going from the FEB at x = −108rg0 (a) to just upstream of the subshock position at x = −10−4rg0 (g). It is instructive to compare this figure with Figure 1 in Bykov et al. (2011b) where the growth rates for the three instabilities are plotted separately for a fixed set of parameters. In the Monte Carlo code, the combined turbulence growth rate is determined self-consistently from the three instabilities at each position in the shock precursor and the instantaneous growth rate varies widely as a function of position and wave number.

Figure 3.

Figure 3. Instability growth rate at different positions in the shock precursor for our nonlinear Model A. We plot Γ/(krg0) vs. krg0 to reduce the scale spread and the labels indicate the x-positions where the curves are calculated: (a) 108, (b) 107, (c) 106, (d) 1.9 × 103, (e) 5.2, (f) 1.9 × 10−3, and (g) 10−4, all in units of −rg0. Compare this with Figure 1 in Bykov et al. (2011b).

Standard image High-resolution image

In addition to modifying the MFA, the efficiency of turbulence cascade through k-space strongly influences the spectrum of mesoscopic magnetic fluctuations. Unfortunately, the nature of turbulence cascade is not well understood. Local turbulence cascade parallel to the mean field can be suppressed in MHD turbulence (e.g., Goldreich & Sridhar 1997; Biskamp 2003; Sahraoui et al. 2006). We assume this to justify our models without cascade where we set Π(x, k) = 0 in Equation (7). Both the Bell and long-wavelength instabilities have maximum growth rates along the local mean magnetic field (e.g., Bykov et al. 2013a). We leave the more difficult issue of anisotropic cascading to future work.

In Figures 4 and 5 we show the turbulence, with and without cascading, calculated upstream at the FEB (dotted, blue curves), in the precursor at 1% of the distance to the FEB (dashed, red curves), and downstream from the shock (solid, black curves).17 As expected, cascading produces large differences in the turbulence at short wavelengths but the longest wavelengths are much less affected. Also shown in Figures 4 and 5 is the effect of the LWI. The top panels show the turbulence without the LWI (NB = 0), the middle panels show the case where resonant, Bell, and the LWI are combined, and the bottom panels show a long-wavelength comparison in the precursor at x = −0.01 LFEB. The comparison at x = −0.01 LFEB in the bottom panels shows that the LWI broadens the spectral peak and shifts it toward larger scales. The LWI enhances the longest wavelength turbulence by at least an order of magnitude with or without cascading.

Figure 4.

Figure 4. Turbulence spectra for the example shown in Figure 2 with turbulence cascading. In the top two panels, the solid (black) curve is calculated downstream from the subshock, the dashed (red) curve is calculated in the shock precursor at x = −0.01 LFEB = −106rg0, and the dotted (blue) curve is calculated at the FEB. In the top panel (Model C), the LWI is not included while in the middle panel (Model A) it is. In all cases, the resonant instability is included. In the bottom panel, the cases with and without the LWI are compared at x = −0.01 LFEB.

Standard image High-resolution image
Figure 5.

Figure 5. Turbulence spectra for the example shown in Figure 2 without turbulence cascading. Other than cascading, all aspects of the figure are similar to Figure 4. The model without the LWI is Model D, while that with the LWI is Model A. The turbulence surrounding the solid dot in the middle panel corresponds roughly to the diffusion coefficient at the solid dot in the bottom panel of Figure 7.

Standard image High-resolution image

In the top panel of Figure 6 we show the shock frame, downstream proton phase-space distributions, multiplied by p4/(mpc), for the four cases shown in Figures 4 and 5. The bottom panel shows the downstream diffusion coefficient, D2, for these cases where D2, or λ(x, p), is determined from Equation (52). Figure 7 is a similar plot calculated in the shock precursor at x = −0.01 LFEB = −106rg0. The regions indicated by numbers in Figures 6 and 7 have particular characteristics. The low momentum region 1 is where vortex convection dominates and D2 is independent of p (Equation (45)). For the upstream position shown in Figure 7, low-energy accelerated particles do not reach this position so there is no CR pressure gradient or current to produce turbulence on short scales. The highest energy region 4, is dominated by short-scale fluctuations (Equation (48)) and D2p2. The intermediate range 2–3 is where quasi-resonant processes dominate and the p dependence of D2 depends on the interplay of Bls, k, and W in Equation (55). However, since Bls depends on W and k through Equation (19) there is no simple one-to-one correspondence between the turbulence shown in Figures 4 and 5 and D2.

Figure 6.

Figure 6. Phase-space distributions, [p4/(mpc)]f2(p), and diffusion coefficients for the four examples shown in Figures 4 and 5 calculated downstream from the subshock. The color coding is the same for each panel and the labeled regions in the bottom panel are discussed in the text. Note that the resonant instability is active in all examples. Here and elsewhere the particle distribution is anisotropic and calculated in the shock frame and the plotted f2(p) is averaged over all angles and normalized such that $n(x)=4 \pi \int _0^\infty f(x,p)p^2dp =n(x)$ is the local number density. The models are black (D), red (C), green (A), and blue (B).

Standard image High-resolution image
Figure 7.

Figure 7. Phase-space distributions, [p4/(mpc)]f(x, p), and diffusion coefficients for the four examples shown in Figures 4 and 5 calculated in the shock precursor at x = −0.01 LFEB. The color coding, labels, and normalization of f(x, p) are the same as in Figure 6 and the resonant instability is active for all examples. The models are black (D), red (C), green (A), and blue (B). Note that the cold incoming beam is present in the precursor but not shown in the upper panel. The solid dot in the bottom panel corresponds to the solid dot in the middle panel of Figure 5.

Standard image High-resolution image

Nevertheless features such as the flattening of D2 between regions 2 and 3 can be understood in general terms. In going from low to high p, the resonant k decreases causing Bls to decrease (Equation (19)). This, combined with the increase (or slow decrease) in kW as k decreases causes λ(x, p) (Equation (52)) to increase slowly. Here, in region 3, D2 increases less rapidly then p, i.e., slower than the Bohm limit. In contrast, in region 2, as p increases, k decreases, kW flattens out and D2 increases faster than p2. The transition in D2 from 2 to 3 indicated with a solid dot in the bottom panel of Figure 7 corresponds roughly to the turbulence at the position indicated by the solid dot in the middle panel of Figure 5.

While the magnetic fluctuation spectra with and without cascading are very different in the short-scale regime, as shown in Figures 4 and 5, the corresponding DS particle spectra, shown in Figure 6, are quite similar. There is, however, a clear increase in pmax when the LWI operates with the Bell and resonant instabilities. This is shown in Figure 8, where pmax increases by about a factor of two when the LWI is included. This coupling between the short-wavelength modes from Bell's instability (much shorter than the CR gyroradius) and the long-wavelength modes from the LWI (much longer than the CR gyroradius) highlights the need for simulations to cover a wide dynamic range in order to capture this essential physics.

Figure 8.

Figure 8. High momentum portion of the downstream phase-space distributions, [p4/(mpc)]f2(p), shown in Figure 6. The resonant instability is active for all examples. The models are black (D), red (C), green (A), and blue (B).

Standard image High-resolution image

3.2. Particle Spectra

In Figure 9 we show downstream particle spectra where the resonant, Bell, and long-wavelength instabilities are active for shocks with varying speed (top panel), varying ambient density (middle panel), and varying LFEB (bottom panel. For all plots there is no cascading, T0 = 104 K, B0 = 3 μG, the ISM turbulence is such that BKolm = 3 μG, and for the top two panels the FEB is at the same physical distance from the subshock, i.e., LFEB ≃ −0.11 pc.

Figure 9.

Figure 9. Downstream, shock frame proton distribution functions calculated without cascading for the models as indicated. In the top panel, the shock speed, u0, is varied; in the middle panel, the ambient density, n0, is varied; and in the bottom panel, the FEB, LFEB, is varied. Parameters held constant are shown in each panel. In all cases, T0 = 104 K and B0 = 3 μG. The solid dots indicate where p4f2(p) is a maximum and defines pmax.

Standard image High-resolution image

In contrast to the top panel in Figure 6, where the downstream distribution functions vary little with changes in the turbulence generation and cascading, the spectra in Figure 9 vary importantly with u0, n0, and LFEB. For a given physical distance to the FEB (top panel), the maximum CR momentum, pmax, increases with u0. We indicate the trend in pmax with a solid dot placed at the maximum in p4f(p). This trend results from the fact that the upstream diffusion length scales as 1/u0 so a fixed LFEB increases in terms of particle gyroradii as u0 increases. The top panel also shows that the thermal peak and the minimum in the p4f(p) distribution increase with u0. It is significant that the minimum in p4f(p) occurs well above mpc in all cases.

The middle and bottom panels of Figure 9 show that the normalization of f2(p) scales as n0 and pmax scales both with LFEB and n0. From the three sets of simulations in Figure 9 (and additional runs not shown for clarity), we obtain the scaling relation

Equation (61)

or, assuming that LFEB is some fraction of the shock radius Rsk, one obtains $p_{\rm max}\propto n^\delta _0 \, u_{0} R_{\rm sk}$. Notably, we find a rather weak dependence of pmax on n0 of δ ∼ 0.25 (middle panel of Figure 9). This result, for a quantity critical for all shock applications, is in contrast with the scaling expected if one assumes D(pmax)/u0Rsk and that D(pmax) depends on the proton gyroradius rg(pmax) in the effective, downstream, amplified magnetic field, Beff, 2. Since MFA depends on the ram pressure, it has been suggested that $B_{\rm eff,2}\propto \sqrt{n_{0}}\, u_{0}$ (see, e.g., Ptuskin et al. 2010; Schure & Bell 2013). In this case, pmax would scale as $\sqrt{n_{0}}\, u_{0}^2\, L_{\rm FEB}$, a stronger dependence on both the upstream density and the shock velocity then we find with our self-consistent simulations.

In regards to the postshock turbulent magnetic field, Beff, 2, we find the following dependence on the far upstream density and shock velocity (see Figures 10 and 12)

Equation (62)

At u0 ≳ 5000 km s−1, the efficiency of MFA, defined as Pw, 2P0, saturates at roughly 10%–15% of the far upstream ram pressure (see Figure 11). This implies $P_{w,2}\propto u_0^2$, or $B_{\rm eff,2}^2 \propto u_0^2 n_0$, corresponding to θ  ∼  1. At lower shock velocities, we find Pw, 2P0u0 (i.e., $P_{w,2}\propto u_0^3$), and therefore $B_{\rm eff,2}^2/n_0 \propto u_0^3$, i.e., θ ∼ 1.5. At shock velocities below ∼1000 km s−1, the efficiency of MFA drops to about a percent of the ram pressure.

Figure 10.

Figure 10. Downstream amplified field, Beff, 2, vs. ambient density. The models running left to right are F, I, J, K, and L. The dashed (red) line shows the function as indicated.

Standard image High-resolution image
Figure 11.

Figure 11. Downstream pressure in turbulence vs. shock speed for Models (running left to right) N, X, Y, Z, AA, and W without turbulence cascade and Models Q, R, S, T, U, and V with cascade. The dashed (red) line approximates the behavior for low shock speeds.

Standard image High-resolution image

In Figure 12 we compare our results with the scaling relation determined in Caprioli & Spitkovsky (2014) using hybrid simulations. Their Equation (2) (with our notation) is

Equation (63)

and it is clear from Figure 12 that, while the Monte Carlo result matches Equation (63) at u0 = 1000 km s−1 (i.e., MA = 84), it has a very different scaling at higher shock speeds and higher Alfvén Mach numbers. It is important to understand the reasons for this difference, which must stem from the different assumptions and parameters for the two simulations, since any modeling of a real object requires such scalings. Our Alfvén Mach numbers run from about 84 to 2500, whereas the hybrid results are for MA ≲ 100. This is significant because MFA in shocks with Alfvén Mach numbers below about 30 is dominated by the resonant CR instability (e.g., Amato & Blasi 2009; Caprioli & Spitkovsky 2014). In our high Mach number results, non-resonant instabilities dominate. Apart from different magnetizations, there are different assumptions made for the magnetic fluctuation spectra of the incoming plasma. In the Monte Carlo modeling, the initial spectrum of magnetic fluctuations is Kolmogorov, as determined by Equation (59), whereas the initial fluctuations in the hybrid simulations are determined by numerical noise.

Figure 12.

Figure 12. Comparison of our Models (from left to right) N, X, Y, Z, AA, and W with Equation (2) in Caprioli & Spitkovsky (2014). Note that $P_{\rm cr}^{\prime }=P_{\rm cr}/(\rho _0 u_0^2)$ plotted in the y-axis is the normalized CR pressure.

Standard image High-resolution image

It is significant that the dependence of the post-shock turbulent magnetic field can be tested with SNR observations. The compilation by Vink (2012; see his Figure 39) can be understood if the scaling $P_{w,2}\propto u_0^3$ (i.e., $B_{\rm eff,2}^2 \propto u_0^3 n_0$) holds at shock velocities below ∼104 km s−1 and changes to $P_{w,2}\propto u_0^2$ (i.e., $B_{\rm eff,2}^2 \propto u_0^2 n_0$) at higher shock velocities. We find that this scaling holds for both upstream temperatures T0 = 104 and 106 K, indicating that it is not a sonic Mach number effect, at least for fairly large MS. We also find that it is only weakly dependent on the far upstream mean field B0, indicating only a weak Alfvén Mach number dependence. In addition, the scaling does not depend strongly on LFEB.

Apart from the magnetic field–shock velocity scalings, high spatial resolution Chandra X-ray observations of Tycho's SNR reported by Eriksen et al. (2011) have revealed coherent synchrotron structures which may be related to amplified magnetic fields (e.g., Bykov et al. 2011a; Malkov et al. 2012). The presence of long-wavelength peaks in magnetic turbulence spectra, which are apparent in Figures 5 and 17 (bottom panel), may be tested with high resolution synchrotron images. We shall discuss the modeling of synchrotron images elsewhere.

Other scalings include the acceleration efficiency, as indicated by the escaping CR energy flux Qesc,18 the DS CR pressure Pcr, 2, the DS wave pressure Pw, 2, and the shock compression ratios, Rtot and Rsub. In Figures 13 and 14 we show these scalings for shocks without cascading and for LFEB ≃ −0.11 pc. We find that increasing n0 from 0.3 to 30 cm−3 (Figure 13) results in a decrease of Rtot from ∼11 to ∼8.6 with a corresponding increase in Rsub from ∼2.8 to ∼3.1. In Figure 14 we see that as u0 and the shock strength increase, Rtot and Qesc first increase and then decrease. Since LFEB is set at a fixed physical distance for these examples, the size of the shock acceleration site decreases with increasing u0 resulting in a lower pmax, as shown in the top panel of Figure 9. The decrease in pmax dominates the increase in shock strength causing Qesc to decrease when u0 ≳ 2500 km s−1.

Figure 13.

Figure 13. Top panel shows the total and subshock compression ratios and the bottom panel shows the acceleration efficiency as given by the DS pressure in trapped CRs Pcr, 2, the DS pressure in turbulence Pw, 2 (both as fractions of ΦP0), and the fraction of upstream energy flux escaping at the upstream FEB Qesc vs. ambient density for Models (from left to right) F, I, J, K, and L.

Standard image High-resolution image
Figure 14.

Figure 14. Top panel shows the total and subshock compression ratios as a function of shock speed for fixed n0 and physical distance to the FEB. In the bottom panel, various quantities are shown as labeled. The CR and wave pressures are calculated downstream from the subshock as fractions of ΦP0 and Qesc is given as the fraction of ΦE0. The models, running left to right, are N, X, Y, Z, AA, and W.

Standard image High-resolution image

3.2.1. Cascading and Thermodynamic Properties

In Figure 15 we show the effects of cascading and the smooth shock structure on the background plasma temperature. The dashed (blue) curves are for an unmodified shock (Model UM), the solid (black) curves are for a modified shock without cascading (Model A), and the dot-dashed (red) curves are for a modified shock with cascading (Model B). As in Vladimirov et al. (2009), without cascading the dissipation term, L, in Equation (7) is set to zero. When cascading is included, we assume viscous dissipation such that ${\cal L}= v_a k^2 k_d^{-1} W$ (e.g., Vainshtein et al. 1993). The wavenumber, kd, is identified with the inverse of the thermal proton gyroradius: $k_d(x) = eB_{\rm ls}(x,k_d)/[c\sqrt{m_pk_B T(x)}]$, where T(x) is the local gas temperature determined from the heating induced by L, as described in Vladimirov et al. (2008).

Figure 15.

Figure 15. Top panel compares the shock structure for an unmodified shock (Model UM, dashed blue curve), a modified shock without cascading (Model A, solid black curve), and a modified shock with cascading (Model B, dot-dashed red curve) In the bottom panel, the temperature structure is shown using the same line notation.

Standard image High-resolution image

Without cascading, or in the UM shock, the precursor temperature remains within a factor of three of T0 until a sharp increase occurs at the subshock. In contrast, cascading heats the precursor substantially so T(x) ≳ 100T0 well in front of the subshock which may produce observable consequences. The downstream temperature is dramatically reduced in the NL cases compared to the UM shock and T2 is slightly less with cascading than without.

In Figure 16 we compare the downstream temperature for a different set of shocks with and without turbulence cascading as a function of u0. It is clear that T2 is not very sensitive to cascading despite the large effect on the precursor temperature, although the difference is larger for larger u0. In the top panel of Figure 17, we show DS proton spectra for u0 = 3 × 104 km s−1. The fluxes at the high-energy end of the spectra are somewhat sensitive to the turbulent cascading, while the maximum proton energy is not. As we have emphasized earlier, NL effects damp changes that might otherwise be expected from the large differences in the self-generated turbulence (bottom panel). If the acceleration is as efficient as we show here, conservation considerations force the shock to adjust such that the particle distribution functions cannot vary substantially. In contrast to the turbulence for the u0 = 5 × 103 km s−1 shock shown in Figure 4, the spectrum with cascading for u0 = 3 × 104 km s−1 shows a clear long-wavelength peak (dashed red curve in the bottom panel of Figure 17).

Figure 16.

Figure 16. Downstream temperatures with (dashed curve) and without (solid curve) cascading. The models without cascading, running left to right, are N, X, Y, Z, AA, and W. Those with cascading are Q, R, S, T, U, and V.

Standard image High-resolution image
Figure 17.

Figure 17. Top panel shows the downstream phase-space distributions and the bottom panel shows the downstream turbulence for Models W and V. These models have u0 = 3 × 104 km s−1 and LFEB ≃ 0.11 pc.

Standard image High-resolution image

3.3. Scattering Center Velocity

As described in Section 2.6, we calculate the scattering center velocity, vscat(x), from conservation considerations without assuming any specific form for the turbulence, in particular, without assuming Alfvén waves. This macroscopic approach guarantees that the energy in the growing magnetic turbulence, which produces the CR scattering, is taken from the free energy of the anisotropic CR distribution.

In contrast, if the scattering center velocity is calculated in a microscopic, quasi-linear approach, one has to weight the phase velocities of the modes with the anisotropy of their wave-vector distributions. In the simplest case of the resonant CR streaming instability, the amplified modes are assumed to be Alfvén modes with wave-vectors aligned against the CR pressure gradient in the shock precursor. The situation is much more complicated for the dominant Bell and long-wavelength nonresonant instabilities. For these, no simple Alfvén-wave-like assumption is adequate.

In Figure 18 we show vscat(x) from Model A along with the mean flow speed, u(x), and the Alfvén speeds derived with the upstream field B0 and with the local amplified field Bls(x, kmax). It is seen that vscat(x) is everywhere well below u(x) so there is no sizeable CR spectral softening due to a finite scattering center velocity. The Monte Carlo vscat(x) exceeds the Alfvén speed calculated with B0 in most of the precursor, but is well below the Alfvén speed determined by the amplified Bls(x, kmax) except in the far upstream region near the FEB. If va[Bls(x, kmax)] gives the scattering center speed, strong MFA implies significant softening of the CR spectrum since the effective compression ratio for DSA will be less than Rtot. Such softening has been discussed by Zirakashvili & Ptuskin (2008), Blasi (2013), and Kang et al. (2013).

Figure 18.

Figure 18. Scattering center velocity vscat(x) (blue line) derived from our Model A for a shock velocity u0 = 5 × 103 km s−1, far upstream density n0 = 0.3 cm−3, far upstream magnetic field B0 = 3 μG, and LFEB ≃ 0.56 pc. We also show the bulk flow velocity u(x) (black line) and the Alfvén velocities calculated with the amplified magnetic field Bls(x, kmax) (green line) and the initial field B0 (red line). Note that vscat and the Alfvén speeds are directed upward so we plot the negative values.

Standard image High-resolution image

Morlino & Caprioli (2012) have suggested that a large scattering speed resulting from MFA and rapid Alfvén waves in DSA with CR acceleration efficiencies ≃ 20%, could produce CR spectra steeper than dN/dEE−2. If so, this might address the issue of the steep CR spectrum derived from Fermi observations of Tycho's SNR (see also Caprioli 2012; Slane et al. 2014). However, in general, the situation is not this simple.

Zirakashvili et al. (2014) recently showed that to explain the available multi-wavelength observations of Cas A, NL DSA (with a total efficiency ≳25%) of electrons, protons, and oxygen ions, by both the forward and reverse shocks, must be invoked. Particle injection in their kinetic models is an adjustable parameter and was varied independently for the forward and reverse shocks to fit the multi-wavelength observations. The model results in a variety of spectral shapes—hard spectra for oxygen ions accelerated in the reverse shock and soft for proton spectra from the forward shock with spectral indexes above 2, as inferred for Tycho. Furthermore, soft CR spectra might be expected for quasi-perpendicular shocks where the angle between the shock normal and the ambient magnetic field ∼90° (e.g., Schure & Bell 2013). Another uncertainty occurs for DSA in partially ionized plasmas where the shock compression is reduced by the neutral return flux (e.g., Blasi et al. 2012; Blasi 2013). We cite these cases to emphasize that the interpretation of gamma-ray emission from DSA in young SNRs requires careful modeling of complicated systems and is not yet fully developed.

4. DISCUSSION AND CONCLUSIONS

We have presented a comprehensive, NL model of MFA in shocks undergoing DSA. The magnetic turbulence responsible for scattering particles is calculated self-consistently from the free energy in the anisotropic, position dependent, distributions of those particles. For the first time, we simultaneously include turbulence growth from the resonant CR streaming instability together with the non-resonant short- and long-wavelength, CR-current-driven instabilities. From the magnetic turbulence, we determine the particle diffusion coefficient with a set of assumptions that depend on the particle momentum. Our plane-parallel, steady-state model includes shock modification and thermal particle injection and, for the parameters assumed here, results in efficient acceleration with Qesc up to ∼50% of the incoming energy flux and large downstream CR pressures. We have explored the implications of this efficient acceleration with a limited parameter survey.

Despite the approximations required, we believe this is the most general description of NL DSA yet presented for the following reasons.

  • 1.  
    The Monte Carlo technique has a wide dynamic range and can follow particles from injection at thermal energies (∼1 eV) to escape at PeV energies. This range is currently far greater than can be achieved with PIC simulations or hydrodynamic turbulence calculations. Since no assumption of near-isotropy is required, both the injection of thermal particles and the escape of the highest energy CRs can be treated self-consistently with a determination of the shock structure. Modeling the feedback between thermal and PeV particles is essential in high Mach number shocks typical of young SNRs since CRs near pmax can contain a large fraction of the shock ram kinetic energy.
  • 2.  
    Our iterative model is the first to include the combined growth rates for resonant and non-resonant CR-driven instabilities and to derive these from the highly anisotropic CR distributions in the self-consistently determined shock precursor. The instabilities produce highly amplified magnetic fluctuations which feed on one another making a consistent treatment essential.
  • 3.  
    The iterative technique also provides a way to calculate the scattering center speed vscat(x), directly from energy conservation without assuming any particular form for the magnetic turbulence. This approach is different from all previous treatments of the effects of a finite scattering center speed on DSA. Since the CR current modifies the dispersion properties of the magnetic modes and results in fast non-resonant instabilities in the shock precursor, there is no reason to assume the self-generated turbulence in NL shocks is well described as Alfvén waves. We do not assume this and believe ours is the first attempt at a general determination of vscat since the resonant Alfvén wave instability was introduced for NL DSA by McKenzie & Voelk (1982).

Other non-test-particle techniques that are currently being used to study NL DSA and MFA with instabilities in addition to the resonant streaming instability are MHD calculations or plasma simulations, either full-particle PIC or hybrid.

Three-dimensional MHD calculations have been performed by Bell et al. (2013; see also Bell et al. 2011; Schure & Bell 2013 and references therein) where the CRs are modeled kinetically and the CR current responsible for the NRH (i.e., Bell) instability is included self-consistently. Particular attention is given to the escaping CRs in this work and, based on estimates for the escaping CR current needed to produce enough NRH turbulence to confine high-energy CRs, the model predicts the maximum CR energy without scaling by the age or size of the accelerator (see, for example, Blasi et al. 2007 for earlier work on determining pmax taking into account NL effects). Due to the computational requirements of the three-dimensional MHD simulation, however, the CR momentum range is restricted to a factor of order 10–100 and large shock speeds (c/5) are assumed.

As is well known, plasma simulations have a great advantage over other techniques because, in principle, they provide a full, self-consistent calculation of the shock formation, particle injection and acceleration, and magnetic turbulence generation. This comes with severe computational requirements imposed by directly following particles in a self-generated magnetic field where the relevant length and time scales are set by the mircoscopic plasma parameters. For a hybrid simulation, the basic length scale is the ion inertial length, cpi, and the basic time scale is the ion gyro-period tg = mpc/(eB0).

Recently, Caprioli & Spitkovsky (2014; see also Caprioli & Spitkovsky 2013a, 2013b) have presented hybrid simulations of large, high Mach number, parallel shocks. They see some evidence for the LWI and suggest this may be from escaping CRs. As we noted in our discussion of Figure 12, however, we see a very different MFA scaling from their Equation (2). While it is not clear what causes this, the basic assumptions and scales of the two simulations are extremely different. For the time-dependent PIC simulations of Caprioli & Spitkovsky (2014), for a background field B0 = 3 μG, the largest dimension in a two-dimensional box was Lmax = 4 × 105cpi ≃ 5 × 10−6 pc, the longest run time was tmax = 2500tg ≃ 3 × 10−3 yr, and the momentum range was less than three decades. For the run with Lmax = 4 × 105cpi, the transverse size was 1000 ωpi ≃ 10−8 pc. In comparison, the steady-state Monte Carlo shocks used to generate Figure 12 had LFEB ≃ 0.11 pc, followed CR acceleration for the equivalent of hundreds of years, and had a momentum range of ∼8 decades. Regardless of the computational limits of the plasma simulations, they contain critical plasma physics that is not modeled with the Monte Carlo or MHD techniques making it important to meaningfully match the "small-scale" plasma simulation results to the "large-scale" Monte Carlo results to obtain a consistent calculation over a dynamic range large enough in both space and time to model SNRs.

The results we have presented are all for highly efficient DSA since our main goal is to study the effects of MFA from the three CR-driven instabilities in such cases. While global CR acceleration efficiencies on the order of ≃ 10% are generally assumed for SNRs, and efficiencies on this order are obtained by simpler parameterized NL models (e.g., Ellison et al. 2012), local efficiencies could be far higher (e.g., Völk et al. 2003). There is also indirect evidence from multi-wavelength observations of some SNRs for high efficiencies (see, for example, Hughes et al. 2000; Reynolds 2008; Helder et al. 2012). We note that a direct observation of CR acceleration efficiency ≃ 25% was obtained at the small, weak quasi-parallel Earth bow shock (Ellison et al. 1990b) and we see no fundamental reason, either from theory or SNR observations, that restricts the intrinsic acceleration efficiency to values well below what we model at large, strong, SNR shocks, at least in some circumstances.

The efficiency of DSA, at least in terms of the downstream Pcr, is potentially observable through the widths of Balmer lines from young SNRs in partially ionized material. This gives an estimate of the electron temperature and a model dependent estimate of the ion temperature and CR pressure can be obtained in association with X-ray observations and a measurement of the shock velocity. We note that neutrals may play an important role in the energy balance of shocks with velocities below ∼3000 km s−1 propagating in partially ionized media (see Helder et al. 2012; Blasi 2013; Bykov et al. 2013c for reviews). Here, we assumed fully ionized plasmas and did not attempt to model the effects of neutrals. To directly compare the pressure in the trapped downstream CR particles with that derived from Balmer line observations, one should account for the effects of neutrals and the realistic geometry of the postshock flow in a NL DSA model.

Of course the spectral shape is an additional fundamental prediction of DSA and concave spectra as hard as we show may present a problem in this regard, as mentioned above in our discussion of vscat. We caution that, as is generally the case, we make a diffusion approximation for CR propagation.19 However, Bell et al. (2011) showed that diffusive propagation may break down for CRs in oblique shocks. They found that higher order anisotropies, that appear in a non-diffusion model, may result in harder spectra at quasi-parallel shocks and softer spectra at quasi-perpendicular shocks. It is important to point out that departures from standard diffusive propagation may be important in NL shocks and such propagation may modify the spectral shape. We shall consider such departures in a separate paper.

The authors thank the referee for useful comments. D.C.E. acknowledges support from NASA grant NNX11AE03G. A.M.B. was partially supported by the Russian Academy of Sciences OFN 15 Program. S.M.O. was partially supported by the Russian Academy of Sciences Presidium Program, RBRF grant for young scientists 14-02-31721.

Footnotes

  • Of course, escape can occur at any boundary but for concreteness we only consider upstream escape from the shock precursor.

  • In PIC simulations, both electrons and ions are followed while in hybrid simulations the ions are followed but the electrons are treated as a charge neutralizing fluid. Since hybrid simulations do not follow the plasma on electron time scales the computational requirements are much less.

  • The ion inertial length is cpi, where $\omega _{\rm pi}=\sqrt{4 \pi e^2 n_e/m_p}$ is the proton plasma frequency, c is the speed of light, e is the electronic charge, ne is the electron number density, and mp is the proton mass.

  • To go from B0 = 3 μG to B  >  300 μG with ΔB/B ≲ 0.1 requires ∼50 iterations, a number easily accommodated within the Monte Carlo procedure.

  • As mentioned above, we do not attempt to model very short-scale Weibel-type instabilities.

  • 10 

    In our scenario, the ISM field consists of two components. There is a homogeneous part, B0, and a turbulent part, BKolm, where the ISM turbulence is assumed to have a Kolmogorov spectrum. The total $B_{\rm ISM}^2 = B_0^2 + B_{\rm Kolm}^2$.

  • 11 

    If αt = κt = 0, Equation (18) gives the standard current driven resonant and Bell instabilities.

  • 12 

    We note that we do not include the ion-acoustic instability (e.g., Drury & Falle 1986; Malkov & Drury 2001) which may be important for plasma heating in the precursor.

  • 13 

    Note that in the precursor vscat will always be negative, that is directed upstream in the local frame.

  • 14 

    We note that the turbulence found in PIC and hybrid simulations (e.g., Kato & Takabe 2010; Caprioli & Spitkovsky 2014), as well as turbulence generated near interplanetary shocks (e.g., Baring et al. 1997; Kajdič et al. 2012), is not necessarily well described as Alfvén waves.

  • 15 

    The subscript "pic" suggests that λpic can be determined from PIC simulations.

  • 16 

    While the energy flux of escaping CRs is included self-consistently, we do not calculate here the turbulence generated by this flux. In efficient DSA, the escaping flux at the upstream FEB is produced by the highest energy CRs and the turbulence generated by these CRs may significantly influence the maximum CR energy the shock can produce. Work to account for the turbulence generated by escaping CRs is in progress.

  • 17 

    Except for the small bump produced by escaping CRs, the dotted (blue) curves at the FEB are essentially the background turbulence assumed in the simulation. The far upstream turbulence may well be modified by the escaping CR flux (e.g., Schure & Bell 2014) but we do not consider that here.

  • 18 

    The escaping CR energy flux in Figures 13 and 14 and in Table 1 are defined in the shock rest frame. The escaping energy flux in the far upstream (i.e., ISM for SNRs) frame is also meaningful. These two fluxes are very close for shocks with velocities below ∼104 km s−1, but the difference can become important for higher speeds. At u0 = 3 × 104 km s−1, the escaping flux in the ISM rest frame is ∼ 30% higher than in the shock frame.

  • 19 

    While the Monte Carlo model does not rely on a diffusion equation to describe the evolution of the CR distribution particles are propagated with an assumed exponential pathlength distribution typical of ordinary diffusion.

Please wait… references are loading.
10.1088/0004-637X/789/2/137