Heuristic Description of Perpendicular Diffusion of Energetic Particles in Astrophysical Plasmas

Published 2019 August 19 © 2019. The American Astronomical Society. All rights reserved.
, , Citation A. Shalchi 2019 ApJL 881 L27 DOI 10.3847/2041-8213/ab379d

Download Article PDF
DownloadArticle ePub

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

2041-8205/881/2/L27

Abstract

A heuristic approach for collisionless perpendicular diffusion of energetic particles is presented. Analytic forms for the corresponding diffusion coefficient are derived. The heuristic approach presented here explains the parameter a2 used in previous theories in order to achieve agreement with simulations and its relation to collisionless Rechester & Rosenbluth diffusion. The obtained results are highly relevant for applications because previously used formulas are altered significantly in certain situations.

Export citation and abstract BibTeX RIS

1. Introduction

A problem of the utmost importance is the interaction between electrically charged particles and magnetized plasmas. It plays a significant role in a variety of physical systems ranging from fusion devices, over the solar wind, to the shock fronts of supernova explosions. In all of those scenarios energetic particles experience scattering due to complicated interactions with turbulent magnetic fields. Some early work was done based on perturbation theory also known as quasi-linear theory (see, e.g., Jokipii 1966), but in general this approach fails. Some more heuristic arguments but also systematic theories have been developed focusing on electron heat transport in fusion plasmas where collisions are assumed to play a significant role (see, e.g., Rechester & Rosenbluth 1978; Kadomtsev & Pogutse 1979; Krommes et al. 1983). In space plasmas such as the solar wind or the interstellar medium, on the other hand, collisions are absent and, thus, it was concluded that the aforementioned approaches are not applicable. The assumption of exponential field line separation was also questioned (see, e.g., Matthaeus et al. 2003). In the context of astrophysical plasmas, however, one still finds perpendicular diffusion in most cases as shown via test-particle simulations (see, e.g., Giacalone & Jokipii 1999; Qin et al. 2002), but it remained unclear what the mechanisms behind this type of transport are. Progress has been achieved due to the development of the nonlinear guiding center theory (Matthaeus et al. 2003), the unified nonlinear transport (UNLT) theory of Shalchi (2010), as well as its time-dependent generalization (Shalchi 2017). Within diffusive UNLT theory the perpendicular diffusion coefficient is given by

Equation (1)

where F(k, k) = (vk)2/(3κk2). The solution of this equation depends on the spectral tensor Pnm describing the magnetic fluctuations, the parallel mean free path λ = 3κ/v, the particle speed v, and the mean field B0. Asymptotic solutions of Equation (1) and the importance of the Kubo number (Kubo 1963) defined via K = (δBx)/(B0), depending on the parallel and perpendicular bendover scales and as well as the turbulent magnetic field δBx, have been discussed in Shalchi (2015). Equation (1) shows good agreement with most test-particle simulations, in particular with those performed for three-dimensional turbulence with small and intermediate Kubo numbers. Furthermore, Equation (1) contains quasi-linear theory as well as the nonlinear theory of field line random walk (FLRW) developed by Matthaeus et al. (1995). In Shalchi (2017) time-dependent UNLT theory has been derived, which is represented by

Equation (2)

with the parallel correlation function given by $\xi ({k}_{\parallel },t)=({v}^{2}/3)({\omega }_{+}{e}^{{\omega }_{+}t}-{\omega }_{-}{e}^{{\omega }_{-}t})/({\omega }_{+}-{\omega }_{-})$ where ${\omega }_{\pm }=-v/{(2{\lambda }_{\parallel })\pm {[{v}^{2}/(2{\lambda }_{\parallel })}^{2}-{v}^{2}{k}_{\parallel }^{2}/3]}^{1/2}$. Equation (1) can be derived from Equation (2) by employing a diffusion approximation. Furthermore, the theory explains why diffusion is restored and this is entirely due to transverse complexity becoming important. Due to the exponential factor in Equation (2), this means that diffusion is obtained if $\langle {({\rm{\Delta }}x)}^{2}\rangle \geqslant 2{{\ell }}_{\perp }^{2}$. However, there are at least two remaining problems in the theory of perpendicular diffusion. First, there is a discrepancy between theory and simulations in the large Kubo number regime that was previously balanced out by using the factor a2 (see Equations (1) and (2)) and by setting a2 = 1/3 (see Matthaeus et al. 2003). Furthermore, the question remains what the physics behind collisionless perpendicular diffusion is. This Letter provides an answer to both questions.

2. The Three Rules of Perpendicular Diffusion

We now formulate rules allowing us to derive formulas for the perpendicular diffusion coefficient without employing systematic theories. Those rules are:

  • 1.  
    Perpendicular transport is only controlled by three effects, namely, parallel transport, the random walk of magnetic field lines, as well as transverse complexity. The last of these three effects leads to the particles getting scattered away from the original magnetic field lines they were tied to.
  • 2.  
    We assume that the bendover scales and , the integral scales L and L, the ultra-scale LU, as well as the Kolmogorov scale LK are finite and nonzero. Furthermore, the parallel motion is assumed to be ballistic at early times and thereafter turns into a diffusive motion described by the parallel diffusion coefficient κ. The FLRW is initially ballistic and becomes diffusive for larger distances. In this case it is described by the field line diffusion coefficient κFL which depends on some of the aforementioned scales.
  • 3.  
    In order to obtain normal diffusion, the particles need to leave the original magnetic field lines they followed. This happens as soon as transverse complexity becomes significant corresponding to1
    Equation (3)
    What the perpendicular diffusion coefficient is depends solely on the state of parallel and field line transport at the time particles start to satisfy condition (3).

3. The Perpendicular Diffusion Coefficient

In the following we construct the perpendicular diffusion coefficient κ based on the three rules formulated above. We shall derive eight cases that are summarized in Table 1. As demonstrated, there are four different routes to perpendicular diffusion as listed in Table 2.

Table 1.  The Eight Cases of Perpendicular Transport

Case Parallel Motion Field Lines $\langle {({\rm{\Delta }}x)}^{2}\rangle \geqslant 2{{\ell }}_{\perp }^{2}$ Perpendicular Transport Diffusion Coefficient Described by UNLT Theory
1 Ballistic Ballistic No Ballistic ${d}_{\perp }(t)=\tfrac{{v}^{2}}{3}\tfrac{\delta {B}_{x}^{2}}{{B}_{0}^{2}}t$ Yes
2 Ballistic Ballistic Yes Double-ballistic diffusion ${\kappa }_{\perp }=\sqrt{\tfrac{2}{3}}v{{\ell }}_{\perp }\tfrac{\delta {B}_{x}}{{B}_{0}}$ Yes
3 Ballistic Diffusive No FLRW Limit ${\kappa }_{\perp }=\tfrac{v}{2}{\kappa }_{\mathrm{FL}}$ Yes
4 Ballistic Diffusive Yes FLRW Limit ${\kappa }_{\perp }=\tfrac{v}{2}{\kappa }_{\mathrm{FL}}$ Yes
5 Diffusive Ballistic No Fluid Limit ${\kappa }_{\perp }=\tfrac{\delta {B}_{x}^{2}}{{B}_{0}^{2}}{\kappa }_{\parallel }$ Yes
6 Diffusive Ballistic Yes Fluid Limit ${\kappa }_{\perp }=\tfrac{\delta {B}_{x}^{2}}{{B}_{0}^{2}}{\kappa }_{\parallel }$ Yes
7 Diffusive Diffusive No Compound Subdiffusion ${d}_{\perp }(t)={\kappa }_{\mathrm{FL}}\sqrt{\tfrac{{\kappa }_{\parallel }}{2t}}$ Only for small Kubo numbers
8 Diffusive Diffusive Yes CLRR Limit ${\kappa }_{\perp }={\left(\tfrac{{\kappa }_{\mathrm{FL}}}{{{\ell }}_{\perp }}\right)}^{2}{\kappa }_{\parallel }$ Only for small Kubo numbers

Note. The results are compared with limits contained in time-dependent UNLT theory represented by Equation (2).

Download table as:  ASCIITypeset image

Table 2.  The Four Routes to Perpendicular Diffusion

Route Final State Diffusion Coefficient
$1\to 2$ Double-ballistic diffusion ${\kappa }_{\perp }=\sqrt{\tfrac{2}{3}}v{{\ell }}_{\perp }\tfrac{\delta {B}_{x}}{{B}_{0}}$
$1\to 3\to 4$ FLRW Limit ${\kappa }_{\perp }=\tfrac{v}{2}{\kappa }_{\mathrm{FL}}$
$1\to 5\to 6$ Fluid Limit ${\kappa }_{\perp }=\tfrac{\delta {B}_{x}^{2}}{{B}_{0}^{2}}{\kappa }_{\parallel }$
$1\to 7\to 8$ CLRR Limit ${\kappa }_{\perp }={\left(\tfrac{{\kappa }_{\mathrm{FL}}}{{{\ell }}_{\perp }}\right)}^{2}{\kappa }_{\parallel }$

Note. In the first three cases perpendicular transport starts as ballistic motion that then turns into a diffusive motion. In the fourth case the ballistic motion is followed by a subdiffusive regime and thereafter diffusion is restored.

Download table as:  ASCIITypeset image

3.1. The Field Line Random Walk Limit

First, we assume that the random walk of magnetic field lines is diffusive in the scenario of interest2

Equation (4)

If we assume that there are no collisions and no pitch-angle scattering, we can set z = vμt where we used the pitch-angle cosine μ. Combining this with Equation (4) and averaging over μ yields $\langle {({\rm{\Delta }}x)}^{2}\rangle =v{\kappa }_{\mathrm{FL}}t$, and thus

Equation (5)

corresponding to the FLRW limit. This limit is stable because if condition (3) is met, it does not alter the transport. This case is highly relevant in the limit of long parallel mean free paths corresponding to high particle energies (see Figures 1 and 2).

Figure 1.

Figure 1. Results for a spectral tensor based on the critical balance condition of Goldreich & Sridhar (1995). For δB/B0 = 1 the field line diffusion coefficient is in this case κFL = 0.38. Shown are the simulations (dots) of Jokipii (2011), the result of diffusive UNLT theory for a2 = 1 (solid line), the CLRR limit (dashed line) as given by Equation (9), the FLRW limit (dotted line) as given by Equation (5), and the composite formula (gray line) as given by Equation (20).

Standard image High-resolution image
Figure 2.

Figure 2. Results for a spectral tensor based on the NRMHD model of Ruffolo & Matthaeus (2013). For δB/B0 = 1 the field line diffusion coefficient is in this case κFL = 0.23. Shown are the simulations (dots) of Shalchi & Hussein (2014), the results of diffusive UNLT theory for a2 = 1/3 and a2 = 1 (solid lines), the CLRR limit (dashed line) as given by Equation (9), the FLRW limit (dotted line) as given by Equation (5), and the composite formula (gray line) as given by Equation (20).

Standard image High-resolution image

3.2. Compound Subdiffusion

If there is strong pitch-angle scattering the parallel motion is diffusive meaning that

Equation (6)

Assuming that field lines are diffusive and particles follow field lines, we can combine Equations (4) and (6) to find

Equation (7)

The running perpendicular diffusion coefficient is then

Equation (8)

corresponding to subdiffusive transport. However, diffusion will be restored as soon as condition (3) is satisfied as discussed in the next paragraph. For slab turbulence, on the other hand, this condition is never satisfied due to ${{\ell }}_{\perp }=\infty $, and thus we find compound subdiffusion as the final state of perpendicular transport.

3.3. The Collisionless Rechester & Rosenbluth Regime

We now assume that diffusion is restored as soon as the particles scatter away from their original field lines. This happens as soon as the condition (3) is satisfied. We also assume that this happens after the particles travel the distance LK in the parallel direction leading to ${\kappa }_{\perp }/{\kappa }_{\parallel }=\langle {({\rm{\Delta }}x)}^{2}\rangle /\langle {({\rm{\Delta }}z)}^{2}\rangle ={{\ell }}_{\perp }^{2}/{L}_{K}^{2}$. In order to eliminate LK we use the field line diffusion coefficient ${\kappa }_{\mathrm{FL}}=\langle {({\rm{\Delta }}x)}^{2}\rangle /(2| z| )={{\ell }}_{\perp }^{2}/{L}_{K}$ yielding

Equation (9)

Alternatively, one can replace the scale therein so that

Equation (10)

in agreement with Equation (8) of Rechester & Rosenbluth (1978) as well as Equation (4) of Krommes et al. (1983). The quantity LK is either called the Kolmogorov–Lyapunov length or just the Kolmogorov length (see, e.g., Krommes et al. 1983). However, here LK is not an exponentiation length but a characteristic distance along the mean field at which transverse complexity becomes significant. Furthermore, Equations (9) and (10) were obtained without assuming collisions, and thus we call this result the collisionless Rechester & Rosenbluth (CLRR) limit. One can also obtain this by using a slightly different derivation. We assume that we find compound subdiffusion until the particles satisfy condition (3), which happens at the diffusion time td so that Equations (7) and (8) become $2{{\ell }}_{\perp }^{2}=2{\kappa }_{\mathrm{FL}}\sqrt{2{\kappa }_{\parallel }{t}_{d}}$ as well as ${\kappa }_{\perp }={\kappa }_{\mathrm{FL}}\sqrt{{\kappa }_{\parallel }/(2{t}_{d})}$. Combining the latter two equations in order to eliminate td yields again Equation (9). In order to evaluate this further, we consider two subcases, namely, small and large values of the Kubo number, respectively. For small Kubo numbers the field line diffusion coefficient is given by the quasi-linear limit

Equation (11)

With this form Equation (9) becomes

Equation (12)

in agreement with the scaling obtained from the diffusive UNLT theory in Shalchi (2015). Furthermore, we find

Equation (13)

For large Kubo numbers, on the other hand, we have

Equation (14)

with the ultra-scale LU. Equation (14) is either called the nonlinear or Bohmian limit and is similar compared to the field line diffusion coefficient obtained by Kadomtsev & Pogutse (1979). Therewith, Equation (9) becomes

Equation (15)

and the Kolmogorov scale is ${L}_{K}=({{\ell }}_{\perp }^{2}{B}_{0})/({L}_{U}\delta {B}_{x})$.

3.4. The Fluid Limit

Let us now assume that parallel transport is diffusive but magnetic field lines are still ballistic when the particles start to satisfy condition (3). Then we can derive $\langle {({\rm{\Delta }}x)}^{2}\rangle =\langle {({\rm{\Delta }}z)}^{2}\rangle \delta {B}_{x}^{2}/{B}_{0}^{2}\,=\,2{\kappa }_{\parallel }t\delta {B}_{x}^{2}/{B}_{0}^{2}$, and thus

Equation (16)

which Krommes et al. (1983) called the fluid limit.

3.5. The Initial Free-streaming Regime

The simplest case is obtained for the early times when parallel and field line transport are ballistic. In this case

Equation (17)

so that

Equation (18)

corresponding to ballistic perpendicular transport. However, this is not a stable regime since we only find this type of transport before condition (3) is met.

3.6. Double-ballistic Diffusion

We now consider a scenario where the transport is still ballistic when the particles start to satisfy condition (3). Therefore, we use Equations (17) and (18) to derive $2{{\ell }}_{\perp }^{2}={v}^{2}\delta {B}_{x}^{2}{t}_{d}^{2}/(3{B}_{0}^{2})$ as well as ${\kappa }_{\perp }={v}^{2}\delta {B}_{x}^{2}{t}_{d}/(3{B}_{0}^{2})$. Combining the latter two equations leads to

Equation (19)

A similar result can be derived from Equation (2) by assuming a ballistic perpendicular motion.

3.7. Timescale Arguments

In order to determine which case is valid for which scenario, one needs to explore at which time a certain process takes place. In the parallel direction particles need to travel a parallel mean free path in order to get diffusive, and thus t = λ/v = 3κ/v2. In the following we focus on the case of short λ. For small Kubo numbers the field lines become diffusive for $| z| \approx {{\ell }}_{\parallel }$ and the corresponding time is ${t}_{{\rm{}}{\rm{F}}{\rm{L}}}\approx {{\ell }}_{\parallel }^{2}/(2{\kappa }_{\parallel })$. Then, on the other hand, if we assume that condition (3) is satisfied while the field lines are still ballistic, we have $2{{\ell }}_{\perp }^{2}=2{\kappa }_{\parallel }{t}_{\mathrm{Fluid}}\delta {B}_{x}^{2}/{B}_{0}^{2}$. For ${t}_{\parallel }\lt {t}_{\mathrm{Fluid}}\lt {t}_{{\rm{}}{\rm{F}}{\rm{L}}}$ the final state is the fluid limit because then we find that parallel transport becomes diffusive first and then we meet condition (3). If, on the other hand, ${t}_{\parallel }\lt {t}_{{\rm{}}{\rm{F}}{\rm{L}}}\lt {t}_{\mathrm{Fluid}}$ the field lines become diffusive before condition (3) is met. This means that we find compound subdiffusion first. At even later times condition (3) is eventually met and diffusion is restored. The corresponding diffusion coefficient is then the CLRR limit. Using the formulas for the times discussed above, this means that we find CLRR diffusion for ${\lambda }_{\parallel }^{2}\ll {{\ell }}_{\parallel }^{2}\ll {{\ell }}_{\parallel }{L}_{K}$ where the Kolmogorov length LK is given by Equation (13). Thus, for λ ≪  we either find the fluid limit or CLRR diffusion. If additionally  ≫ LK we find the fluid limit, but for  ≪ LK we get CLRR diffusion. It follows from Equation (13) that LK/ ≈ K−2 ≫ 1 meaning that for small Kubo numbers we always find CLRR diffusion. For large Kubo numbers similar considerations can be made.

3.8. A Composite Formula

A problem of the heuristic approach is that the obtained formulas are only valid in asymptotic limits. Since the two most important cases are CLRR and FLRW limits, we propose for the perpendicular mean free path defined via λ = 3κ/v the formula

Equation (20)

Equation (20) was chosen so that for ${\lambda }_{\parallel }\to 0$ we obtain Equation (9) and for ${\lambda }_{\parallel }\to \infty $ we get Equation (5). Note that Equation (20) does not contain the fluid limit given by Equation (16), and thus it has some limitations.

3.9. Further Comments

The results obtained here are sometimes not comparable to previous results. First of all there are cases such as slab or two-dimensional (2D) turbulence. In the former case condition (3) is never satisfied leading to compound subdiffusion as the final state. In the 2D case parallel transport is not diffusive (see, e.g., Arendt & Shalchi 2018) violating the second rule. In some work (see, e.g., Matthaeus et al. 2003; Shalchi et al. 2004) a flat spectrum at large scales was used for the 2D modes. For this type of spectrum the ultra-scale is not finite, also violating the second rule. In order to determine the form of κFL, we have used the Kubo number. However, in some turbulence models (see, e.g., Goldreich & Sridhar 1995) there is only one scale, and thus the Kubo number becomes K = δBx/B0, often called the Alfvénic Mach number. The arguments presented above are still valid.

4. Comparison between Theory and Simulations

As a first example we consider two-component turbulence with dominant 2D modes. For a well-behaving spectrum, Shalchi & Weinhorst (2009) have derived

Equation (21)

requiring q > 1 for the energy range spectral index and 1 < s < 2 for the inertial range spectral index. With the parameter a2 included, nonlinear theories provide in the limit of short parallel mean free paths and 2D turbulence (see, e.g., Shalchi et al. 2004 and Zank et al. 2004)

Equation (22)

According to the heuristic approach we expect CLRR diffusion in the considered parameter regime. Comparing Equations (22) and (15) yields a = LU/ and using Equation (21) for the ultra-scale gives us a2 = (s − 1)/(q − 1). Previously it was often assumed that s = 5/3 and q = 3 (see, for instance, Arendt & Shalchi 2018) leading to a2 = 1/3. Although it was already stated in Matthaeus et al. (2003) that a2 = 1/3 is needed to achieve agreement between theory and simulations, in the current Letter we found for the first time an explanation for this value. It has to be noted that this result was obtained for a specific form of the spectrum. Alternative spectra and the associated turbulence scales have been discussed in Matthaeus et al. (2007). For some of those spectra one obtains an ultra-scale larger than the bendover scale. In such cases, however, one would expect that the diffusion coefficient is close to the fluid limit, and thus a2 ≈ 1.

Two further examples are shown in Figures 1 and 2, respectively. A spectral tensor based on the critical balance condition of Goldreich & Sridhar (1995) was used in the simulations of Sun & Jokipii (2011). Figure 1 compares diffusive UNLT theory and the heuristic approach presented in the current paper with those simulations. As shown, the FLRW and CLRR limits have to be understood as asymptotic limits. Figure 2 visualizes the comparison for a spectral tensor based on the noisy reduced MHD (NRMHD) model of Ruffolo & Matthaeus (2013) showing very good agreement.

The heuristic arguments presented in this Letter cannot substitute systematic theories due to the lack of accuracy in the general case. The remaining step is to further improve UNLT theory so that the factor a2 is no longer needed. This should lead to a complete systematic theory for perpendicular transport in space plasmas.

Footnotes

  • It is assumed here that is the scale at which transverse complexity becomes significant. In principle, this could be a different scale such as the integral scale L. The use of the bendover scale, however, is motivated by time-dependent UNLT theory (see Section 1).

  • Note that the diffusion coefficient κFL has length units.

Please wait… references are loading.
10.3847/2041-8213/ab379d