WEIBEL, TWO-STREAM, FILAMENTATION, OBLIQUE, BELL, BUNEMAN...WHICH ONE GROWS FASTER?

Published 2009 June 17 © 2009. The American Astronomical Society. All rights reserved.
, , Citation A. Bret 2009 ApJ 699 990 DOI 10.1088/0004-637X/699/2/990

0004-637X/699/2/990

ABSTRACT

Many competing linear instabilities are likely to occur in astrophysical settings, and it is important to assess which one grows faster for a given situation. An analytical model including the main beam plasma instabilities is developed. The full three-dimensional dielectric tensor is thus explained for a cold relativistic electron beam passing through a cold plasma, accounting for a guiding magnetic field, a return electronic current, and moving protons. Considering any orientations of the wave vector allows to retrieve the most unstable mode for any parameters set. An unified description of the filamentation (Weibel), two-stream, Buneman, Bell instabilities (and more) is thus provided, allowing for the exact determination of their hierarchy in terms of the system parameters. For relevance to both real situations and PIC simulations, the electron-to-proton mass ratio is treated as a parameter, and numerical calculations are conducted with two different values, namely 1/1836 and 1/100. In the system parameter phase space, the shape of the domains governed by each kind of instability is far from being trivial. For low-density beams, the ultra-magnetized regime tends to be governed by either the two-stream or the Buneman instabilities. For beam densities equaling the plasma one, up to four kinds of modes are likely to play a role, depending of the beam Lorentz factor. In some regions of the system parameters phase space, the dominant mode may vary with the electron-to-proton mass ratio. Application is made to solar flares, intergalactic streams, and relativistic shocks physics.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Weibel, filamentation, two-stream, Bell, or Buneman instabilities are ubiquitous in astrophysics. They are involved in the physics of solar flares where relativistic electron beams are assumed to lose their energy through beam-plasma instabilities, producing hard X-ray emissions (Karlicky et al. 2008; Karlicky 2009). Filamentation, or Weibel, instabilities could also be responsible for the birth of cosmological magnetic fields (Schlickeiser & Shukla 2003; Schlickeiser 2005; Lazar 2008; Lazar et al. 2009) as unstable particle streams through the intergalactic medium can magnetize an initially un-magnetized system. Such instabilities could also play an important role in explaining the origin of a variety of high-energy photons sources including supernova remnants, active galactic nuclei, gamma-ray bursts, or Pulsar Wind Nebulae (Piran 1999; Gedalin et al. 2002; Piran 2004; Waxman 2006; Stockem et al. 2007). Within some of these systems, it is assumed that cosmic rays are accelerated through shocks (relativistic or not) while the instability generated upstream by their interaction with the interstellar medium provides the magnetic turbulence eventually responsible for synchrotron radiation emissions (Medvedev & Loeb 1999; Silva et al. 2003; Nishikawa et al. 2006; Frederiksen et al. 2004; Milosavljevic & Nakar 2006; Lemoine et al. 2006; Niemiec et al. 2008).

Regardless of the context, the typical structure investigated consists in a beam-plasma system initially both charge and current neutralized. Initial charge neutralization implies the inclusion of positive and negative species, while current neutralization demands at least two streaming species. Streams are often considered as opposed, but they can be parallel when streaming species are of opposite signs. For example, a pair beam does not need any more beams to be current neutral. Finally, accounting for an external magnetic field B0 allows the system to be relevant to a wide class of astrophysical problem. The simplest case consists in a B0 parallel to the flow(s) but normal or oblique orientations have also been considered (Fraix-Burnet & Pelletier 1991; Dieckmann et al. 2008; Bret & Dieckmann 2008a; Sironi & Spitkovsky 2009).

We consider here the generic system formed by a cold electron beam of density nb streaming at initial velocity vb, with Lorentz factor γb = (1 − v2b/c2)−1/2, over a cold electron/proton plasma. The beam current is neutralized by an electronic return current of density np and velocity vp such that nbvb = npvp. Ions are initially at rest with density ni = nb + np. Finally, the model includes a flow-aligned magnetic field B0vb. Even for such a simple system, linear stability analysis is intricate because unstable modes are numerous. The intent of this paper is to clarify this issue and determine the fastest growing mode for any given set of parameters, by implementing an exact model encompassing every basic instability. In order to keep the results tractable, emphasis is on the simplest possible model with the lowest number of free parameters. As explained in the sequel, these requirements demand a beam-plasma system with mobile ions and a guiding magnetic field, while kinetic effects will not be investigated here. A cold fluid model is thus exactly solved, rendering every possible coupling between unstable modes. The dispersion equation arising from the exact dielectric tensor is analyzed without calling on the electrostatic (kE), or the purely electromagnetic (kE), approximations. Such kind of calculation is mandatory if the two-stream or Buneman electrostatic modes are to be described within the very same formalism than the electromagnetic filamentation or Bell-like instabilities.

Let us list the possible unstable modes by progressively "assembling" the system. The first block here is the unmagnetized background plasma with fixed protons. Because we take it cold, it is stable against every kind of perturbations. Note that some temperature anisotropy would make it Weibel unstable (Weibel 1959), the fastest growing modes being found for wave vectors perpendicular to the high-temperature axis (Kalman et al. 1968). Kinetic effects can thus drive this first component unstable even before we add anything else.

Let us now "add" the electron beam and let the plasma electrons establish the return current (protons are still fixed). To simplify the discussion, we consider here a diluted beam with nbnp (the rest of the paper also deals with higher beam densities). As is known, the resulting beam-plasma system is unstable. Perturbations with wave vector kvb are unstable for 0 < k ≲ ωp/vb and define the so-called two-stream Instability (Bohm & Gross 1949). Perturbations having kvb are also unstable for any k and pertain to the filamentation instability.1 Finally, perturbations with k neither parallel nor perpendicular to the flow are also unstable (Faĭnberg et al. 1970; Califano et al. 1997, 1998a, 1998b) and may even be the fastest growing modes in the diluted relativistic beam regime (Bret & Deutsch 2005; Bret et al. 2008).

Let us now give the protons the possibility to move. Every aforementioned unstable mode is still unstable as the two counter-streaming electron beams still interact. But both electron currents can now interact with the protons, giving rise to Buneman unstable modes (Buneman 1959). A closer look at the situation shows that Buneman modes arising from the beam/protons interaction merge with the beam/return-current modes (Bret & Dieckmann 2008b). But the return-current/proton interaction results in unstable Buneman modes reaching their maximum growth rate for k ∼ (np/nbp/vb and kvb. Within the diluted beam regime, nbnp implies unstable modes with much shorter wavelength than the two-stream modes. Indeed, it has been found than the Buneman modes can compete with the two-stream ones if the beam is relativistic enough.

We finally make our system complete by adding a flow-aligned magnetic field. Here again, the new ingredient brings in more unstable modes without necessarily stabilizing the previous ones. To start with, electronic Cyclotron and Upper-Hybrid modes are destabilized, adding new branches to the unstable spectrum for any orientations of k (Godfrey et al. 1975). Also, Alfvén modes resulting from the combination of the vibrating protons in the magnetic field can also be destabilized. These unstable modes were first pointed out by Bell by means of an MHD formalism (Bell 2004, 2005), and their description was later extended through the kinetic one (Zweibel 2002; Reville et al. 2006).

Our simple magnetized beam/plasma system with moving protons in eventually two-stream, filamentation, oblique, Cyclotron, Upper-Hybrid, Buneman, Bell...unstable! As the system is released from equilibrium, every instability is triggered, and the outcome of the linear phase is mostly shaped by the growth of the fastest growing one. This most unstable mode can be found for an oblique wave vector so that the search of the dominant mode requires the implementation of a model capable of describing any unstable mode for any orientation of the wave vector. Note that recent works involving ion beams also evidenced such kind of modes (Niemiec et al. 2008; Ohira & Takahara 2008). We thus now proceed to the elaboration of the simplest possible model incorporating all the aforementioned instabilities.

The unstable modes described here are just too numerous to be detailed one by one. On the other hand, each one needs to be discussed since the present aim is precisely to determine their hierarchy. The full wave vector dependence of each mode is therefore skipped, the focus being set on the most unstable wave vector with the corresponding growth rate. Additionally, modes which have not been found to govern the system for any given sets of parameters are only briefly discussed in order to keep the presentation tractable. Still for clarity, the names of the numerous modes have not been abbreviated. Finally, analytic expressions in terms of the magnetic field parameter ΩB defined by Equations (4) are essentially derived for ΩB > 1, while results are also presented for the un-magnetized regime.

The paper is structured as follow: the analytic model is explained in the next section. Unstable modes found for flow-aligned wave vectors are listed in Section 3 before an overview of the two-dimensional unstable spectrum is given in Section 4. The key results of the article are exposed in Section 5 where the fastest growing mode is determined in terms of the system parameters. In this respect, Figures 5 and 8 can be considered as the main results of the paper. They show which kind of mode governs the linear phase of the beam-plasma system in terms of the parameters. Application is then made to various astrophysical settings in Section 6, before the final discussion and conclusion.

2. ANALYTIC MODEL

The model relies on the cold relativistic fluid equations for the three species involved. The calculation follows the lines of previous ones (Califano et al. 1998b; Kazimura et al. 1998) except that protons are allowed to move while a static flow-aligned magnetic field is accounted for. The basic equations include Maxwell's equations and,

Equation (1)

Equation (2)

where j = b for the electron beam, j = p for the plasma return-current, and j = i for plasma ions (protons). Here, pj = γjmjvj and qj are the momentum and electric charge of specie j.

Although lengthy, the derivation of the dielectric tensor is quite standard. The conservation, Maxwell's, and Euler's equations are first linearized assuming every quantity slightly departs from equilibrium like exp(ık · r − ıωt) where ı2 = −1. Since B0vb, vp while ∑qjnj = ∑qjnjvj = 0, the charge and current neutral equilibrium state considered exactly fulfills the full set of equations. We can write k = (kx, 0, kz) without loss of generality by virtue of the axial symmetry with respect to the beam and magnetic field axis (Godfrey et al. 1975). We choose the z-axis for the direction of the beam and of the magnetic field, having therefore vb = (0, 0, vb), vp = (0, 0, vp), and B0 = (0, 0, B0). The linearized conservation and Euler's equations are first used to express the perturbed total current J1 in terms of the electromagnetic field. The first-order magnetic field is then eliminated through B1 = (c/ω)k × E1 and the resulting expression of the current is inserted into the usual combination of Maxwell–Faraday and Maxwell–Ampère equations, namely

Equation (3)

The electrostatic (k × E1 = 0) or purely electromagnetic approximations (k · E1 = 0) would here result in a simplification of the k × (k × E1) term. Although such options may be useful when focusing on one give mode, the risk at this stage would be to loose information about possible dominant modes which would not fit in. A Mathematica Notebook has been implemented to symbolically compute the tensor (Bret 2007). It is expressed in terms of the dimensionless variables,

Equation (4)

where ω2p = 4πnpq2/me is the electronic background plasma frequency, ωB = |q|B0/mc the non-relativistic electronic cyclotron frequency and me, mi the electron and ion (proton) mass, respectively. The Alfvén velocity vA can be expressed in terms of the variables above as

Equation (5)

The dielectric tensor has the form,

Equation (6)

where z* is the complex conjugate of z. Tensor elements are reported in Appendix A in terms of the dimensionless variables (4). In the limit of zero magnetic field, only the square of the charges appears in the tensor expression. As a consequence, the results in such a case also apply to two counter-streaming pair plasmas (Jaroschek et al. 2005; Ramirez-Ruiz et al. 2007).

The most general expression of the dispersion equation then reads $\det \mathcal {T}=0$ and necessarily encompasses every possible unstable modes. As previously mentioned, we here emphasize the search of the most unstable mode kM of the unstable k spectrum. The function kM can only depend on the beam-to-plasma density ratio α, the beam Lorentz factor γb, the magnetic field strength parameter ΩB, and the mass ratio R. Once the former has been fixed to 1/1836, or any higher value more suited to comparison with PIC simulations (Jaroschek et al. 2004; Spitkovsky 2008a; Karlicky 2009), the most unstable mode can only be a function of (α, γb, ΩB) ∈ [0, 1] × [1, ] × [0, ]. Note that this study is not limited to the diluted beam case so that the return-current flowing at vp = αvb can also reach relativistic velocities. Tables 1 and 2 summarize the results of the next two sections as they display the main unstable modes in the diluted (α ≪ 1) and symmetric (α = 1) regimes, respectively. Mode properties in the magnetized regime are given only for ΩB > 1. A detailed study (at R = 0) of the transition from ΩB = 0 can be found in Godfrey et al. (1975).

Table 1. Main Unstable Modes in the Diluted Beam Regime

Modes Zx Zza Growth Rate
Txx − ıTxy 0 ΩBb $\propto \sqrt{\alpha }R\Omega _B$b
Txx + ıTxy 0 $\frac{1}{2}\alpha /\Omega _B$ ∝αRΩBb
Two-stream 0 1 $\frac{\sqrt{3}}{2^{4/3}}\alpha ^{1/3}/\gamma _b$
Buneman 0 1/α $\frac{\sqrt{3}}{2^{4/3}}R^{1/3}$
ObliqueB0, ΩB = 0 1 $\frac{\sqrt{3}}{2^{4/3}}(\alpha /\gamma _b)^{1/3}$
UHLc, ΩB > 1 ΩBb $\frac{1}{2}(\alpha /\Omega _B)^{1/2}$
Oblique, ΩB > 1 $\sqrt{1+\Omega _B^2}$ $\sqrt{1+\Omega _B^2}$ $\frac{1}{4}\alpha ^{1/3}/\gamma _b$
UHLc, ΩB > 1 $\Omega _B/\gamma _b+\sqrt{1+\Omega _B^2}$ $\frac{1}{2}\alpha ^{1/2}/\Omega _B$

Notes. aFor a given Zx, the growth rate is a function of Zz. The table only mentions the Zz and the growth rate pertaining to the fastest growing mode. Magnetized mode are reported for ΩB > 1. The transition between ΩB = 0 and 1 is described in Godfrey et al. (1975) for R = 0. bBounded by Equation (13). cUHL stands for "upper-hybrid-like."

Download table as:  ASCIITypeset image

Table 2. Main Unstable Modes in the High Density Beam Regime α = 1

Modes Zx Zz Growth Rate
Txx ± ıTxy 0   $\lesssim \sqrt{R}$
TSBa, γbR1/3 < 1/2 0 $\sqrt{3}/2\gamma _b^{3/2}$ 1/2γ3/2b
TSBa, γbR1/3 > 1/2 0 $\sqrt{2R}$ $\frac{\sqrt{3}}{2^{7/6}}R^{1/6}/\gamma _b$
Oblique, ΩB > 1 1/2B0.75b ΩB/2γb 0.5/γ1.15b
UHLb 1 & 2, ΩB > 1 ∼ΩBb 1/ΩB

Notes. aTSB stands for the merged two-stream/buneman mode. bUHL stands for "upper-hybrid-like."

Download table as:  ASCIITypeset image

3. FLOW-ALIGNED WAVE VECTORS

We here detail the case of flow-aligned wave vector instabilities before analyzing the most general situation. Even if these modes do not always govern the system, such an analysis is useful to get familiar with the competing instabilities, and make a bridge with previous investigations. When considering k = (0, 0, kz), the tensor (6) simplifies to,

Equation (7)

so that the dispersion equation simply reads,

Equation (8)

This equation already contains most of the unstable modes already discussed. The Tzz generates electrostatic two-stream and Buneman modes while the others are responsible for the so-called "Bell" unstable electromagnetic modes.

3.1. Electrostatic Modes

The Tzz terms reads,

Equation (9)

The proton contribution stems from the term ∝R/x2, the beam one from the second term with numerator α, and the return-current with γp = (1 − v2p/c2)−1/2 accounts for the last factor. The dispersion equation for these modes is independent of the magnetic field since the resulting electrostatic modes have kE and only generate velocity perturbations along the magnetic field. This equation has been analyzed elsewhere (Bret & Dieckmann 2008b). For a diluted beam, the interaction of the two electron beams produces two-stream unstable modes with maximum growth rate (Mikhailovskii 1974)

Equation (10)

and the interaction of the return current with the proton yields unstable Buneman modes at much smaller wavelength with,

Equation (11)

These expressions change when the density ratio approaches unity because the beam and the return current become symmetric for α = 1 (see Section 3.3).

3.2. Electromagnetic Modes

The unstable electromagnetic modes arise here from the Txx ± ıTxy terms of Equation (8) which read,

Equation (12)

A straightforward calculation from the general dispersion relation for circularly polarized wave with such wave vectors (Ichimaru 1973; Achterberg 1983; Amato & Blasi 2009) yields the very same result. These modes are the present configuration analogs to the ones explained by Bell (2004, 2005). The main difference with the scenarios in which they are usually involved is that they are presently driven by an electronic current rather than by a baryonic one.

Both dispersion equations yield unstable modes here. The most relevant feature with respect to the mode hierarchy is that growth rates are bounded by RΩB, regardless of the parameters involved. This does not imply the growth rate can go to infinity because unstable solutions exist only in a limited region of the phase space (α, γb, ΩB, R).

The Txx − ıTxy term can be partially assessed analytically neglecting the x2 and the return-current factor. The most unstable modes are found at Zz ∼ ΩBb, with xRΩB. The maximum growth rate varies like $\sqrt{\alpha }$, and is always bounded by

Equation (13)

The Txx + ıTxy factor is more involved and has been partially treated numerically. Unstable modes are found here at $Z_z\sim \frac{1}{2}\alpha /\Omega _B$, with maximum growth rates varying like α, and still bounded by Equation (13). Note that these modes are stable for γb ≲ αΩ2B so that instability in the diluted beam regime demands a low Lorentz factor.

Figure 1 displays the growth rate of these four unstable modes in terms of Zz for some parameters triggering them all. One can observe how the relevant spectrum extends over 3 orders of magnitude. This point, and its consequences, are discussed in the conclusion.

Figure 1.

Figure 1. Growth rate δ (ωp units) of the four unstable modes for flow-aligned wave vectors in terms of Zz = kvbp. Parameters are α = 0.1, γb = 20, ΩB = 1 and R = 1/1836. The two electromagnetic modes yield the thin plain and dashed curves from Equation (12) with "−" and "+" respectively. The bold dashed pertains to the two-stream modes and the Buneman modes gives the plain bold one.

Standard image High-resolution image

3.3. Non-diluted Beam Regime

Because we study small perturbations of the system composed by both the beam and the plasma, the present theory is perfectly valid up to α = 1. However, the physical interpretation of the calculations can no longer be derived considering the beam is a perturbation to the isolated background plasma. The system, beam+plasma, has its own proper modes that are quite different from those of the isolated plasma. Taking R = 0 for example, the dispersion equation has two branches exactly given by

Equation (14)

which definitely differs from the cold isolated plasma dispersion relation x = 1, i.e, ω = ωp. For such a symmetric system, the Buneman modes at Zz ∼ 1/α merge with the two-stream modes at Zz ∼ 1, and depending on γbR1/3, the resulting mode may be governed by the electrons or the ions dynamic (see Table 2).

For parallel wave vectors, electrostatic modes have been studied previously (Bret & Dieckmann 2008b), and two different regimes need to be considered depending on the product γbR1/3. For γbR1/3 ≲ 1/2, the fastest growing mode has

Equation (15)

Note that this result is exact for R = 0. In the opposite limit γbR1/3 ≳ 1/2, one has

Equation (16)

Regarding the electromagnetic modes, the two terms Txx ± ıTxy in Equation (8) yield the same growth rate because for α = 1, these two functions are equal through the transformation x → −x. Here again, the growth rate of these modes is always bounded by RΩB, but instability occurs only for γb ≲ 1/R and $\Omega _B\lesssim \sqrt{\gamma _b}$. As a consequence, the most unstable mode for a given couple (γb, ΩB) cannot grow faster than $R\Omega _B<R\sqrt{\gamma _b}<\sqrt{R}$.

Figure 2 displays the growth rate of the two modes just discussed for parameter sets yielding a Bell governed system. Note that such situation demands a highly magnetized system. Although this plot gives the sensation that Bell-like modes can dominate some region of the parameters phase space, they were never found to govern the full two-dimensional spectrum. This example illustrates the importance of accounting for the full unstable k spectrum instead of focusing on a given wave vector orientation. While Bell-like modes grow slower for non-flow-aligned wave vectors, other modes grow faster in oblique directions. As a result, the unstable spectrum ends up governed by the latters, even if the formers can "reign" over the beam axis. This situation basically stems from the necessary smallness of the mass ratio R. Dealing with a proton beam would allow to reach R = 1, where Bell modes should definitely be found governing some portion of the phase space. A similar theory accounting for a proton beam would be needed to describe such situations.

Figure 2.

Figure 2. Growth rate δ (ωp units) for flow-aligned wave vectors in terms of Zz = kvbp. Parameters are α = 1 for ΩB = 10, and 60, with R = 1/1836 and γb = 103. Here, electromagnetic modes (thin lines) can overcome the electrostatic ones (bold line).

Standard image High-resolution image

4. AN OVERVIEW OF THE 2D SPECTRUM

Before we establish the full hierarchy of the unstable modes in terms of the parameters (α, γb, ΩB), let us give an overview of the unstable spectrum in terms of k = (kx, 0, kz). The basic principle here is that the unstable modes previously discussed for kx = 0 remain unstable for some kx ≠ 0. The analytical analysis becomes more involved because all modes are coupled, as the tensor elements Txz and Tyz in Equation (6) no longer vanish. However, the list of unstable modes is easier to grasp in the diluted beam regime where the beam is just a perturbation to the isolated background plasma. We therefore single out the diluted and non-diluted beam regime in our analysis. Furthermore, unstable modes found in an oblique direction have finite or infinite kx. Both kinds of modes can be tackled by different approaches, and are thus studied separately.

4.1. Diluted Beam - Zx =

For the un-magnetized system with infinite ion mass, the resulting spectrum is now well understood, including within the framework of a full kinetic relativistic theory (Bret et al. 2008). There is a continuum of unstable modes bridging the two-stream instability (k = 0) with the filamentation instability (k = 0). The magnetized version of the same system brings two additional oblique branches. For ΩB > 1, the first one is found at Zz ∼ ΩBb, with a maximum growth rate varying like $\frac{1}{2}\sqrt{\alpha /\Omega _B}$. Still for ΩB > 1, the second branch is located at $Z_z \sim \Omega _B/\gamma _b+\sqrt{1+\Omega _B^2}$ with a maximum growth rate varying like $\frac{1}{2}\sqrt{\alpha }/\Omega _B$ (Godfrey et al. 1975). Within the same region of the spectrum, the oblique modes already present at ΩB = 0 near Zz = 1 are now found at $Z_z\sim \sqrt{1+\Omega _B^2}$ with a maximum increment varying like α1/3b. These later modes are commented in Section 4.3. At this stage, the unstable spectrum may look like the one pictured on Figure 3(a) for the parameters specified in caption.

Figure 3.

Figure 3. Two-dimensional unstable spectrum in terms of the reduced wave vector Z for α = 0.1, γb = 2 and ΩB = 3. (a) R = 0. (b) R = 1/100. (c) R = 1/1836.

Standard image High-resolution image

At this junction, some comment is needed to clarify the meaning of "oblique" mode (with small "o"). Figures 3 and 4 make it clear that unstable modes with both kx ≠ 0 and kz ≠ 0 are numerous. Such kind of modes have so far been referred to in the literature as "oblique" (Watson et al. 1960; Niemiec et al. 2008; Ohira & Takahara 2008), "electromagnetic beam-plasma instability" (Califano et al. 1998b), "coupled two-stream Weibel" (Jaroschek et al. 2005), or "Mixed mode" (Frederiksen & Dieckmann 2008). The problem with such labeling is that "oblique modes" are here just too numerous for only one tag. In Tables 1 and 2, as in the rest of the paper, "oblique modes" bridging between two-stream and filamentation instabilities for ΩB = 0 are labeled "obliqueB0." For ΩB > 1, they evolve into what we presently call "oblique" modes (capital "O"), with both kx ≠ 0 and kz ≠ 0, and a growth rate scaling like γ−1b. Finally, modes found reaching their maximum growth rate at finite Zz and Zx = are labeled "upper-hybrid-like" modes after Godfrey et al. (1975).

Figure 4.

Figure 4. Two-dimensional unstable spectrum in terms of the reduced wave vector Z for α = 1, γb = 103, and ΩB = 60. (a) R = 0. (b) R = 1/100. (c) R = 1/1836. With the same parameters than Figure 2, we see here how Bell-like modes around Zz = 0.08 do govern the Zz axis but not all the spectrum. Map (d), with parameters ΩB = 10, γb = 5, and R = 1/1836, clearly evidences the oblique mode explained in Section 4.3 for Zz, Zx ∼ 1.

Standard image High-resolution image

The spectrum is even richer when ions are "allowed" to move, as Buneman and Bell-like modes are triggered. The resulting spectrum is pictured on Figures 3(b) and (c) for R = 1/100 and 1/1836, respectively. An electron-to-proton mass ratio R = 1/100 may be relevant for PIC simulation where the mass ratio is usually incremented from its 1/1836 realistic value in order to speed up the ions dynamic, saving thus computer time. Regarding this last point, it is interesting to note that such a trick is possible because the growth rates associated with the moving ions increase with R, while the related scale length does not (see Equations (11)). This is reflected on Figures 3(b) and (c) where the dominant modes do not migrate with varying R. If such was not the case, some higher mass ratio could demand higher spatial resolution so that the computing time lost by a finer spatial resolution would not necessarily be compensated by the accelerated dynamic. Also on these Figures, the occurrence of the Buneman instability around Zz ∼ 1/α = 10 is clear. The two-stream instability governs here the system even for R = 1/100, but the electron and ion spectrums are almost disconnected form each other only for R = 1/1836. Because of the smallness of the real mass ratio, both the time and length scales of each spectrum are disconnected from each other. In such a case, the overall spectrum is nearly the superposition of each sub-spectrum.

4.2. Non-diluted Beam Regime, α = 1 and Zx =

Without moving ions (R = 0), the most relevant features of the unstable spectrum in this case are oblique resonances at Zz ∼ ΩBb and Zz ∼ ΩB/2γb. Only the first one extends up to Zx = and yields a maximum growth rate δm ∼ 1/ΩB (Bret et al. 2006). The next paragraph is devoted to an overview of the second kind of modes. This large Zx regime can be analytically investigated in the following way. We start deriving the dispersion equation in this limit setting α = 1, γb = γp and Zx in Equation (6). We then use the fact that unstable modes are found with ℜ(ω) ∼ 0 and develop the resulting dispersion equation to the second order in x. In this large Zx limit, the system is found unstable only between,

Equation (17)

and

Equation (18)

These results are exact and imply that filamentation instability is canceled as soon as Zz1 > 0, i.e., $\Omega _B>\beta \sqrt{2\gamma _b}$. For ΩB ≫ βγ1/2b, the unstable range decreases like Zz2Zz1 ∼ β2B. The range of unstable wave vectors becomes therefore increasingly narrow, yielding a quasi-monochromatic unstable spectrum in this limit.

When ion motion is accounted for, part of the stability domain can still be calculated exactly following the same guidelines. With R ≠ 0, one finds

Equation (19)

and

Equation (20)

The maximum growth rate within this range remains close to 1/ΩB, and another range of unstable wave vectors and similar growth rate appears at slightly larger Zzs. These new modes, arising purely from the ion motion, will not be investigated here although they are accounted for in the forthcoming numerical evaluation of modes hierarchy.

Figures 4(a)–(c) shows the two-dimensional spectrum for the parameters chosen for Figure 2 with ΩB = 60. The purpose of this choice is to observe how some oblique modes govern the full spectrum while Bell-like modes dominate the Zz-axis. On Figure 2, the spectrum for ΩB = 60 is clearly governed by these modes at Zz ∼ 0.08. The same modes, together with the way they evolve when leaving the Zz-axis, are perfectly visible on Figure 4(c) and one can check how they quickly stabilize for oblique wave vectors.

4.3. Oblique Modes at Finite Zx

Sections 4.1 and 4.2 discussed unstable modes located at Zx = . The cold fluid approximation typically send some fast growing modes to Zx = , whereas temperature effects tend to stabilize these small wavelengths instabilities by preventing the pinching of too small filaments (Silva et al. 2002). But the occurrence of a magnetic field also triggers some truly locally most unstable oblique modes. The term "locally" means here that the very same mode, as defined by one root of the dispersion equation, reaches its maximum for a finite Zx. For more clarity, Figure 4(d) displays a two-dimensional spectrum with α = 1, while the other parameters have been chosen to single out this mode.

For the diluted beam case, Table 1 mentions a maximum growth rate varying like α1/3b. For the symmetric case α = 1, Table 2 indicates a maximum oblique growth rate ∼0.5/γ1.15b (numerical fit). Such scaling of this finite Zx mode bears important consequences regarding modes hierarchy: a look at Tables 1 and 2 shows that the ultra-magnetized regime tends to stabilize every modes, except this one, the Buneman and the Two Stream. While more modes compete for moderate ΩB, the large ΩB limit is decided among these three. More details on the ultra-magnetized regime are given in Section 5.2.

5. MODES HIERARCHY

We finally turn to the determination of the modes hierarchy in terms of the parameters α, γb, and ΩB. Only the case ΩB = R = 0 has been treated so far (Bret & Deutsch 2005; Bret et al. 2008) evidencing two main features. On the one hand, obliqueB0 modes can govern the diluted beam regime from nb/np ∼ 0.53. For higher density ratios, filamentation may dominate depending on the beam Lorentz factor γb. On the other hand, obliqueB0 modes govern the ultra-relativistic regime unless the system in strictly symmetric with nb = np.

After having calculated symbolically the dielectric tensor using a previously designed Mathematica Notebook, the polynomial dispersion equation (17th degree) has been transferred to MatLab for numerical study. A systematic search of the most unstable modes has thus been conducted in terms of α, γb, and ΩB for R = 1/1836 and 1/100. For each parameter set (α, γb, ΩB), the program returns the Zz and Zx components of the most unstable mode together with the real and imaginary parts of the corresponding complex frequency ωr + ıδ.

The function $\delta _M=\max \lbrace \delta, (k_x,k_z)\in \mathbb {R}^2\rbrace$ is clearly continuous. But the real part ωM corresponding to δM, together with the wave vector ZM defining the mode growing at δM, may well evolve discontinuously as the most unstable mode can perfectly "jump" from one location of the two-dimensional spectrum to another (Bret et al. 2008). For example, the Zz component of the fastest growing mode along the beam axis, switches abruptly from Zz = 1 to 1/α when the Buneman mode overcomes the two-stream one. Indeed, these discontinuities are a way to determine the domains where any given mode governs the system: as long as the parameters evolve in such a way that the very same mode keeps dominating, ZM and ωM are continuous.2 But when the dominant mode changes, one of these functions will suffer some discontinuity. Within the present cold model, the most unstable oblique modes are often "sent" to Zz = so that an oblique–oblique transition will not necessarily trigger some discontinuity of the perpendicular component. But this is neither the case of the parallel component Zz nor of the real part of the frequency ωM. Figure 5 has thus been elaborated exploiting these properties and is commented in the sequel. Here again, and for better clarity, we single out the diluted beam regime α ≪ 1 from the symmetric case α = 1. The main reason for such a structure is that only these two extremes are analytically accessible, whereas the intermediate case definitely requires numerical assistance.

Figure 5.

Figure 5. Frontiers between the domains governed by different modes in terms of α and γb, for various ΩB and R = 1/1836, 1/100. The arrows on the frontiers show how they evolve when ΩB increases. The red crosses show the parameter sets for the solar flares application explained in Section 6.1.

Standard image High-resolution image

5.1. Diluted Beam

Gathering the aforementioned data, the frontiers between the domains governed by different modes are sketched in terms of (γb, α) on Figure 5. Although α is limited to the range [0.1, 1], the lower part of the graph can easily be deduced from the diluted beam regime expressions reported in Table 1. This process is repeated for various ΩBs and R = 1/1836, 1/100.

Starting with ΩB =0, the findings of Bret & Dieckmann (2008b) are confirmed numerically here. Filamentation dominates for large beam-to-plasma density ratios. The frontier visible in the lower-right corner pertains to the oblique/Buneman transition. The growth rate of the un-magnetized oblique competing mode in this case3 reads $\sqrt{3}/2^{4/3}(\alpha /\gamma _b)^{1/3}$ (Faĭnberg et al. 1970). The equation of the border is, therefore, simply

Equation (21)

For α = R and γb = 1, Equations (10) and (11) are identical so that this frontier extends all the way down to α = R. Below this line, Buneman modes govern the linear phase.

From ΩB = 1, the evolution is three-fold. To start with, the two-stream instability governs an increasing weakly relativistic region (lower-left corner) as its magnetized growth rate remains the same, while oblique modes are made less unstable. By virtue of the same kind of effects, Buneman modes gain weight on the lower-right corner because their growth rate too does not vary with the magnetic field. Finally, filamentation domain shrinks as the magnetic field progressively shuts it down (Cary et al. 1981; Stockem et al. 2007). Regarding the border of the two-stream region, a look at the Zx component shows it switches from 0 to when crossing the limit. A comparison between the growth rate values in this region and the expressions gathered in Table 1, shows that the two-stream mode competes with the upper-hybrid-like one located at $Z_z\sim \Omega _B/\gamma _b+\sqrt{1+\Omega _B^2}$. The two-stream/oblique frontier is thus defined for α ≪ 1 by

Equation (22)

Regarding the Buneman/upper-hybrid-like frontier, the growth rate $\sim \frac{1}{2}\sqrt{\alpha }/\Omega _b$ simply decreases with α until the Buneman mode overcomes it. This transition thus occurs for

Equation (23)

Below this intersection, two-stream modes are directly competing with the Buneman ones. This later frontier is thus defined by

Equation (24)

The partition defined by these three equations in the diluted regime is thus pictured on Figure 6 for ΩB larger than 1 but smaller than ∼R−1/3 (see Section 5.2 for this limitation).

Figure 6.

Figure 6. Low α partition of the phase space defined by Equations (22)–(24) for 1 ≲ ΩBR−1/3 (see Section 5.2).

Standard image High-resolution image

For larger values of ΩB, the previous trends amplify. The oblique domain (central region) is bounded toward the low αs because oblique growth rates are scaled like α1/2 (or α1/3), while the Buneman growth rate does not vary with α. The evolution is of course faster with R = 1/100 than 1/1836 because the Buneman grows faster in the second case.

From ΩB ∼ 5.8, the Buneman modes gain more "territory" according to the aforementioned mechanisms. Noteworthily, new modes appear to govern some portion of the phase space at rather high density ratio and moderate Lorentz factor. For R = 1/1836, the inserted spectrum maps on Figure 5 show that a small part of the phase space (blue one) pertains to the oblique mode described in Section 4.3. This "middle region" is analytically involved to explore and such results could only be derived numerically so far. The same mode governs a similar region for R = 1/100, but in that case, another kind of oblique mode (green regions) also intervenes. Why can such modes lead the linear phase for R = 1/100 and not R = 1/1836? The solution lies in the mode coupling already observed in Section 4. Figures 3(a) and (c) show how the Buneman modes interfere with the rest of the spectrum according to the value of R. For R = 1/100, interferences are stronger than for R = 1/1836 because the spectrum is wider. While the beam is diluted (as is the case for these plots), coupling effects are limited because the various modes are well separated in the k space. As the beam-to-plasma density ratio α increases, the Buneman spectrum progressively merges with the rest of the unstable modes. Interferences thus become potentially more intense, and all the more that R is large. It turns out that the green region in Figure 5 arises from a mode coupling which is much less excited for R = 1/1836 than for R = 1/100. We find here that tuning the electron-to-proton mass ratio in order to speed up the system dynamic may bear qualitative consequences by changing the nature of the dominant mode.

5.2. Diluted Beam, Ultra-magnetized Regime

The outcome of the ultra-magnetized regime is relevant for pulsars physics, for example. For diluted beams, an increasing magnetic field progressively stabilizes every mode except the two-stream, the Buneman, and the oblique one (see Table 1). Both two-stream and oblique are scaled like α1/3b, but the pre-factors favor two-stream modes. Competition is eventually between Buneman and two-stream with a frontier defined by Equation (24). The magnitude of ΩB needed to trigger this picture can be derive by setting α = 1 in Equation (23) as if the "triple-point" pictured in Figure 6 was rejected to α = 1. Such criterion places the ultra-magnetized regime beyond

Equation (25)

which fits well what is observed in Figure 5. Let us now detail the symmetric case and detail the mode hierarchy in terms of (γb, ΩB).

5.3. Symmetric Case, α = 1

When α reaches unity, Table 2 indicates that the competition is between the merged two-stream/Buneman, the upper-hybrid-like and the oblique modes. However, a comparison of the growth rates is possible as long as these modes are well separated in the k space. When overlapping, the growth rate interfere, and analytical predictions become very involved. A closer look shows that the region

Equation (26)

is thus defined. This limit lies below the cancellation threshold for the filamentation instability, $\Omega _B=\beta \sqrt{2\gamma _b}$, at least up to γb = 1/R. Indeed, the numerical exploration of this region shows that filamentation governs the system almost as long as it is not canceled. This allows for the tracing of the filamentation domain in Figure 8, together with the other limits explained in this section. As filamentation vanishes, it evolves smoothly into the upper-hybrid-like modes with kZz ≠ 0.

When condition (26) is fulfilled, it is possible to directly compare the growth rate of the three competing modes. The resulting situation is rendered in Figure 7 and varies strongly with R. The origin of such R-sensitivity is the fact that only the two-stream/Buneman growth rate depends on the mass ratio in the large γb limit. While its γb variation is monotonous, the variation of the oblique modes is not, rendering the determination of the hierarchy highly non-trivial for moderately relativistic beam.

Figure 7.

Figure 7. Growth rates δ (ωp units) of the competing modes in terms of γb, for α = 1, R = 1/100 and 1/1836 (numerical evaluation). The merged two-stream/Buneman modes (TSB, in red) varies like γ−3/2b for γbR−1/3, and like γ−1b after. Oblique modes (in blue) vary like γ−1.15b and upper-hybrid-like modes (UHL, black) like Ω−1B. Due to the non-monotonic behavior of oblique modes for moderate γb, the determination of the hierarchy is highly non-trivial within this energy range.

Standard image High-resolution image

Figure 7 shows that for R = 1/100, oblique modes are only allowed to overcome the others as long as ΩB is not too large. Numerically, it is found that for ΩB ≳ 300, they are overcome by two-stream/Buneman regardless of the beam energy. Beyond this threshold, upper-hybrid-like modes govern if

Equation (27)

The resulting hierarchy map is display on Figure 8 and uncover an intriguing bubble-like oblique domain.

Figure 8.

Figure 8. Modes hierarchy for α = 1, plotted for R = 1/100 and 1/1836. Filamentation modes (sometimes referred to as "Weibel") govern almost as long as their are not canceled. The black star and the square pertain to the parameters chosen for the intergalactic streams and relativistic shocks described in Sections 6.2 and 6.3, respectively. The marks have been placed on the γb-axis because the systems they represent are un-magnetized.

Standard image High-resolution image

When decreasing R down to its realistic value 1/1836, Figure 7 shows how the large γb part of the two-stream/Buneman curve is shifted down by virtue of its R1/6b scaling (see Table 2). Even for ΩB = 500, the oblique growth rate (plain blue curve) surpasses the two-stream/Buneman from γb ∼ 5 to ∼200. Whether such a situation holds for any ΩB is an open question, but it has been checked that it does at least up to ΩB = 103. In order to draw the hierarchy map, let us first determine the Lorentz factor for which two-stream/Buneman modes overcome the oblique one. The corresponding value of γb fulfills

Equation (28)

yielding γb ∼ 234 for R = 1/1836. The hierarchy map here is the one pictured in Figure 8 for R = 1/1836. For ΩB > 2 × 2341.15 = 1060, the system is governed with γb increasing by two-stream/Buneman modes, then oblique, then two-stream/Buneman again, then upper-hybrid-like and finally, filamentation. For smaller ΩBs (yet, larger than ∼4), the last two-stream/Buneman step is just skipped, as the system goes directly from the oblique to the upper-hybrid-like regime.

It is therefore not easy to define some ultra-magnetized regimes here. Unlike the diluted beam case where only two modes are left to compete for a reasonably high magnetic field defined by Equation (25), we find here that every mode involved is likely to play a role, regardless of the magnetization. For R = 1/100, one could define a so-called ultra-magnetized regime from the top of the oblique "Bubble" located at ΩB ∼ 300. For larger magnetization, only three modes are left to compete. But such a definition is no longer possible for R = 1/1836.

Finally, let us add a remark regarding the accuracy of the borders calculations. While it is perfectly possible to numerically compute hierarchy maps 5 and 8 with higher definition, such progress is not imperative because two different systems located on each side of a given border are likely to evolve in quite a similar way during the linear phase. Let us assume the border between two mode domains A and B goes through α = α0, in such a way that A dominates for α < α0. For α = α0, the growth rates δA and δB are strictly equal. If α = α0 ± ε, then δA − δB = ∓ε'. If both A and B are excited with the same initial amplitude, the time required for the fastest to overcome the slowest only by a factor e is ∼ω−1p/ε'. Depending on the system considered, this time may well exceed the duration of the linear phase for ε' small enough. It would thus be inappropriate to claim that only A or B is relevant on its respective side of the border. Even if the dominant mode evolution can be discontinuous, the system dynamic stemming from the growth of the entire spectrum should evolve smoothly when crossing a frontier. One can focus on either A or B only far away from the borders so that the precise calculation of its location is not crucial.

6. APPLICATIONS

We finally turn to the determination of the modes hierarchy in terms of the parameters α, γb, and ΩB for the three settings mentioned earlier. The parameters considered in each case are reported in Table 3 and have been chosen after Karlicky (2009), Lazar et al. (2009), and Silva et al. (2003). The respective growth rate maps have been plotted on Figures 9(a)–(c). Although the present theory is limited to the cold case, possible kinetic effects are indicated in each case, and even calculated for the solar flares environment.

Figure 9.

Figure 9. Two-dimensional unstable spectrum in terms of the reduced wave vector Z for parameters mentioned in Table 3. (a) Solar flares. (b) Intergalactic streams. (c) Relativistic shock. The insert on the upper-left figure shows the result of a fully kinetic calculation with R = 0, assuming electronic beam and plasma temperatures of 20 MK.

Standard image High-resolution image

Table 3. Parameters Used for the Applications

Setting nb/np γb ΩB Dominant Mode
Solar Flaresa 1/8 1.34 0, 1 ObliqueB0, Two-stream
Intergalacticb 1 1.005 0 ObliqueB0
Shocksc 1 10 0 Filamentation

Notes. aFrom Karlicky (2009), β = 2/3 with R = 1/1836. bFrom Lazar et al. (2009), β = 1/10 with R = 1/1836. cFrom Silva et al. (2003). Two colliding pair plasmas without background ions, i.e. R = 0.

Download table as:  ASCIITypeset image

6.1. Solar Flares

Here, we perform the calculation for one magnetized and one non-magnetized case. The system location in the parameters phase space is designated by the red crosses on Figure 5, and the growth rate maps are displayed on Figures 9(a). For both scenarios, the Buneman modes are clearly visible at Zz ∼ 8 = 1/α, but cannot compete with the two-stream ones at Zz ∼ 1 because the beam is not relativistic enough. With such a beam-to-plasma density ratio, a much higher relativistic factor is required for the two-stream/Buneman transition.

For the non-magnetized case (upper Figure 9(a)), this weakly diluted beam system is governed by non-magnetized "obliqueB0" modes at Zz ∼ 1. Although not dominant, the filamentation instability4 plays an important role as its growth rate is quite close to the largest one. The introduction of the magnetic field (lower Figure 9(a)) damps filamentation as well as the oblique mode, resulting in a two-stream driven system. Noteworthily, Figure 5 shows that a small variation of the Lorentz factor can trigger a dominant mode transition for ΩB = 1. For R = 1/100, the system is close to a triple-point where two-stream, Buneman and upper-hybrid-like modes grow the same speed. The system behavior in this case is therefore very sensitive to the parameters choice.

A word of caution is needed here before concluding with respect to the solar flares environment. The simulation performed in Karlicky (2009) accounts for an electronic temperatures of 21.4 × 106 K (21.4 MK). Indeed, analysis of soft and hard X-ray flare data indicate that plasma temperatures up to 40 MK are obtained in large flares (Aschwanden 2002). Furthermore, velocity distributions can be highly anisotropic due to stronger heating along the magnetic field lines (Fisk 1976; Miller 1991; Miller et al. 1997). For a mode with wave vector k growing at growth rate ωi, a temperature spread Δv can be neglected providing |Δv−1ik−1 (Faĭnberg et al. 1970). Such a condition simply ensures that during one e-folding time, particles are traveling almost the same path when compared to the wavelength considered. Focusing on the fastest growing obliqueB0 mode for the non-magnetized case (upper Figure 9(a)), we find,

Equation (29)

Accounting for |Δv| = 0.08c (i.e., T = 20 MK), the condition above translates 0.08 ≪ 0.31. The only change when considering the two-stream dominated magnetized case (lower Figure 9(a)) is that the Lorentz factor on the right-hand side has the exponent −1 instead of −1/3. In such a case, the condition reads 0.08 ≪ 0.25. The cold approximation therefore seems reasonable here, although condition (29) is only weakly fulfilled. An interesting consequence of Equation (29) is that the condition is not homogenous over the unstable spectrum. A kinetic theory of the full unstable spectrum is currently under elaboration, together with a rigorous mode-dependent definition of the kinetic-fluid transition (A. Bret & L. Gremillet 2009, in preparation). For a clearer assessment of the accuracy of the cold approximation in the present case, the insert in the upper Figure 9(a) displays a kinetic calculation of the most relevant portion of the unstable spectrum, assuming electronic beam and plasma temperatures of 20 MK. Such a calculation has been performed neglecting ion motion, which is perfectly valid here. One can check how the portion of the unstable spectrum only slightly departs from its cold counterpart, while the largest growth rate switches from 0.3ωp down to 0.22ωp. Note that similar calculation for the magnetized case is yet to be done.

6.2. Intergalactic Streams

The corresponding spectrum is plotted on Figure 9(b). With a small β and a beam-to-plasma density ratio of 1, this case is also sensitive to the parameters (see the star on Figure 8). We find here that non-magnetized "obliqueB0" modes govern the system, though filamentation is far from being shut down. But such a weakly relativistic symmetric system can switch between oblique or filamentation regimes through very small parameter variations. When accounting for kinetic effects, two-stream/Buneman modes can also compete because they are least sensitive to temperature. Indeed, for R = 0, these three modes grow almost exactly the same way for a beam with γb = 1.1 and temperature 100 keV (Bret et al. 2008). Such a result was found for a plasma temperature of 5 keV, but similar conclusions can be drawn with different plasma thermal spread. The discussion at the end of Section 5 is relevant in this case as this system is bordering various frontiers.

6.3. Relativistic Shocks

These kind of structures are currently extensively studied by means of PIC simulations (Medvedev & Loeb 1999; Silva et al. 2003; Nishikawa et al. 2003; Milosavljevic et al. 2006; Chang et al. 2008; Spitkovsky 2008a, 2008b; Martins et al. 2009), due to their role in the Fireball model for gamma-ray bursts (Piran 2000). The typical scenario arising from these studies is two-fold: to start with, two relativistic plasmas shells collide and the resulting instability eventually generates a quasi-steady propagating shock. These plasma shells can be pair plasmas (Silva et al. 2003; Chang et al. 2008), or electron–ion plasmas (Spitkovsky 2008a; Martins et al. 2009). The un-magnetized pair plasma case can be analyzed here setting R = ΩB = 0, and the system is found clearly governed by the filamentation instability (see Figure 9(c)). This conclusion supports the emphasis put on this instability by the aforementioned authors.5 Moreover, such domination is quite robust: as long as nb/np = 1, γb = 10 and ΩB = 0, the only alternative to the filamentation instability are the non-magnetized oblique modes. Accounting for kinetic effects, a transition to such a regime demands a beam temperature of the order 106 keV (Bret et al. 2008). Figure 8 also suggests that some magnetized version of the system with $\Omega _B>\gamma _b\sqrt{2R}$ could trigger a transition to the magnetized oblique regime. But the bigger threat to the domination of filamentation seems to be the beam-to-plasma density ratio. As evidenced by Figure 5, this instability dominates the ultra-relativistic regime only for strictly symmetric systems. At γb = 10 and 100, filamentation governs only for nb/np ≳ 0.6 and 0.85, respectively. The symmetric system hypothesis has been so far related to the simulation of equivalent density colliding shells, within the Fireball framework. Since a slight change in the density ratio may trigger a dominant mode transition, subsequent work should be needed to test the robustness of the shock formation in this respect.

The second stage of the scenario involves particles which are accelerated by the shock (Spitkovsky 2008b), escape it, and interact with the upstream medium generating more instabilities. Considering now ΩB = 0, γb = 100, and α = 10−2, Figure 5 shows that obliqueB0 modes dominate for R = 1/1836 while Buneman modes do so for R = 1/100. A more detailed evaluation of the spectrum should be done from this point, accounting for the non-thermal energy dispersion of the escaping particles, and for the upstream ion temperature. Sorting out this issue may be important because an interaction governed by the purely electrostatic Buneman instability is less likely to feed the magnetic turbulence needed in the Fireball scenario for synchrotron radiation emission. At any rate, the small beam-to-plasma density ratio implied in this second stage should prevent the filamentation instability from playing any prominent role in the linear phase.

7. DISCUSSION AND CONCLUSION

An exact cold fluid model for a relativistic beam-plasma system including a guiding magnetic field and ion motion has been implemented, allowing for a unified description of every possible instability arising within such systems. After briefly reminding, or deriving, the key results for each kind of unstable modes, the hierarchy between them has been established in terms of the beam-to-plasma density ratio, the beam Lorentz factor, and the magnetic field strength, considering electron-to-proton mass ratios of 1/1836 and 1/100.

In the diluted beam regime, the competition is mainly between the two-stream, the Buneman, and the upper-hybrid-like modes, as the hierarchy diagram in terms or (γb, α) adopts a typical "V" shape for a given ΩB (though not too high). Two-stream and Buneman take advantage of the magnetic field which leave both of them unaffected while stabilizing the rest of the spectrum. Additionally, the Buneman modes, arising from the interaction of the electronic return current with the background protons, govern the highly relativistic regime where the two-stream modes are strongly reduced. The ultra-magnetized regime with ΩBR−1/3 is eventually governed by either the two-stream or the Buneman instabilities.

When leaving the diluted beam region, the shape of the hierarchy diagram is far from being trivial, and two different kinds of Zz ≠ 0 modes may dominate. The first ones are those already governing a part of the parameters space from ΩB = 0. In the cold fluid un-magnetized model, they are found at Zz ∼ 1 and Zx = , whereas temperature effects would give them a finite perpendicular component. Under the action of the magnetic field, they continuously evolve to the upper-hybrid-like modes referred to in Table 1, with a growth rate $\propto \sqrt{\alpha }/\Omega _B$. But from ΩB ∼ 4.3 and 3.6 for R = 1/1836 and 1/100, respectively (numerical evaluations), another kind of oblique mode is likely to govern the spectrum. Unlike the previous ones, these modes reach their maximum growth rate for one single wave vector and have the growth rate scaled like α1/3b. With such a γb scaling, their domain shrinks with increasing beam energy, as observed on Figures 5.

The transition is thus made with the symmetric regime α = 1. Here, two-stream and Buneman modes merge and dominate at low Lorentz factor. According to Figure 8, there is no such thing as a simple description of the ultra-magnetized regime. For a large mass ratio R = 1/100 and ΩB > 300, two-stream, upper-hybrid-like, and filamentation modes are likely to dominate, depending of the Lorentz factor. At lower magnetization, oblique modes are likely to play a role for moderate beam energy. At lower mass ratio with R = 1/1836, the hierarchy is even more complex as the oblique "Bubble" found at higher R expands toward the large ΩB regime.

While the ultra-magnetized regime is governed by electrostatic modes with flow aligned wave vector, this is not true for density ratios close to one. Indeed, the later can be determined by virtually any kind of modes. In this respect, diluted systems are simpler to analyze, or simulate, than symmetric ones.

Some modes, like the Buneman's, strongly depend on the electron-to-proton mass ratio R while others do not. As a result, the hierarchy is R-dependent, and lowering the ratio for computational purposes can bring some qualitative, not just quantitative, changes. Figure 5 shows for example that a system with α = 0.1, γb = 20, and ΩB = 1 is governed by oblique modes for R = 1/1836 and Buneman ones for R = 1/100. Equation (24) shows that the nature of the diluted beam ultra-magnetized regime strongly depends on R.

By virtue of its possible discontinuous nature, as explained in Section 5, a change of the dominant mode can have dramatic consequences. One of them can be illustrated through the two-stream/Buneman transition defined by Equation (24) for diluted beams. Because two-stream modes are found near Zz = 1 and Buneman ones near Zz = 1/α, the typical size of the structures generated is ωp/vb on the left-hand side of the border, and (np/nbp/vb on its right side. As long as the beam is not too diluted, these two quantities remain similar. But for a 103 times diluted beam, both kinds of structures differ by three orders of magnitude. For a relativistic symmetric system with α = 1 and $\gamma _b>\frac{1}{2}R^{-1/3}$, a transition from the two-stream/Buneman regime to the upper-hybrid-like one implies a switch from a $Z_z=\sqrt{2R}$ dominant mode to a Zz = ΩBb one which again can span several orders of magnitude.

These transitions have to do with the parallel wave vector component rather than with the perpendicular one. Similar transitions have been recorded in this paper for this later component, but unstable modes located at Zx = acquire a finite normal component as soon as temperatures are accounted for. While kinetic effects may temper the Zx transitions, such should not be the case for the magnitude of the Zz transitions, because parallel components are less sensitive to temperature than normal ones.

Turning now to the astrophysical settings considered here, it is worth stressing that the value of the present theory consists more in pointing what the main instabilities could be, rather that confirming what they actually are (at least within the current models). Considering solar flares of relativistic shocks for example, it has already been checked that the dominant instabilities are those stressed in Section 6. Even in filamentation, or Weibel, modes are sometimes designated instead of oblique ones, these latters are increasingly specifically discussed in the literature, and identified as such (Jaroschek et al. 2005; Niemiec et al. 2008; Ohira & Takahara 2008; Frederiksen & Dieckmann 2008; Kong et al. 2009). But what PIC simulations can hardly do is indicating extensively which kind of modes could take the lead when changing the parameters, including the electron-to-proton mass ratio.

For solar flares, the domination of obliqueB0 modes is quite robust in the non-magnetized case. Although the cold approximation should correctly describe the most unstable mode, given the temperatures involved, a fully kinetic calculation of the spectrum has been done to confirm this point (see upper Figure 9(a)). The situation is quite different when magnetizing the system. For the magnetic parameter considered here, the lower Figure 9(a) clearly show how some upper-hybrid-like modes at Zz ∼ 2 can also play a role. Indeed, Figure 5 shows that for a realistic value a the electron-to-proton mass ratio R = 1/1836, the transition only needs a slightly higher relativistic factor. In such a case, the instability will heat the system both in the parallel and perpendicular direction while the two-stream driven counterpart preferentially heats it along the beam direction (Karlicky 2009). Furthermore, tuning the mass ratio to speed up the dynamics can here bring qualitative consequences, as evidenced in Figure 5. It is thus found that the evolution of a magnetized electron beam in this setting, together with the kind of heating provided, is quite sensitive to the parameters.

Regarding intergalactic streams, the point made in Lazar et al. (2009) is that given the parameters involved, filamentation instability grows faster than the rest of the spectrum. The expected saturation level of the magnetic field thus produced is then found consistent with the measurements. In this respect, Figure 8 shows that the cold version of the system is bordering the two-stream/Buneman-filamentation frontier, so that filamentation "leadership" is eventually not so clear. Admittedly, kinetic effects are important as filamentation is found to dominate through its interaction with some temperature anisotropy, but such growth rate enhancement has been found to operate in other parts of the spectrum as well (Bret et al. 2005). The question remains open as to know how strong filamentation leadership needs to be for the linear phase to saturate accordingly. Such criterion would likely define some parameter window in which Weibel-like instabilities could be responsible for intergalactic magnetic fields.

Finally, the instability-based scenario for shocks formation within the gamma-ray bursts framework has been examined (Fireball model). Beam-plasma instabilities have been found so far to play a key role at two levels. First, they seem to prompt the shock formation itself, as the collisionless encounter of two plasma shells is unstable. Second, once the shock has been formed, it accelerates particles through diffusive Fermi acceleration. Particles escaping the shock upstream interact with the medium and generate instabilities responsible for the magnetic turbulence needed to trigger synchronic radiation emissions. The first unstable system has a beam-to-plasma density ratio close to unity. Indeed, it is necessarily unity in the recent PIC simulations where a beam is reflected against a wall, and eventually interact with itself. For such settings, filamentation instability has been found leading the linear phase, and the shock formation has so far been analyzed accordingly (Medvedev & Loeb 1999; Spitkovsky 2008a, 2008b). But the present calculations suggest that a density ratio slightly smaller than 1 may result in quite a different dominant instability. It would be very interesting to test the robustness of the shock formation scenario in this respect. Once the shock has been formed, instabilities upstream are thus expected to generate magnetic turbulence. We now deal with a highly relativistic diluted beam/plasma system, which should definitely not be governed by the filamentation instability (see Figure 5). Kinetic effects are unlikely to modify this picture, as filamentation modes are usually their first "victims" (A. Bret & L. Gremillet 2009, in preparation). Here again, a parameter window is defined allowing for the development of the best magnetic turbulence generators: Figure 5 clearly shows how obliqueB0, upper-hybrid-like, two-stream, or Buneman modes are likely to shape the linear phase on this second phase. Among these four candidates, the last two are reputed purely electrostatic instabilities, which should thus be avoided. An accurate characterization of the beam escaping the shock upstream will thus be needed to assess the validity, or set limits, to this scenario.

This work has been achieved under projects FIS 2006-05389 of the Spanish Ministerio de Educación y Ciencia and PAI08-0182-3162 of the Consejería de Educación y Ciencia de la Junta de Comunidades de Castilla-La Mancha. Thanks are due to Gustavo Wouchuk for encouraging discussions.

APPENDIX: TENSOR ELEMENTS

The tensor mentioned in Equation (6) can be cast under the from

Equation (A1)

where

Equation (A2)

with,

Equation (A3)

Equation (A4)

Equation (A5)

Equation (A6)

Equation (A7)

Equation (A8)

Equation (A9)

Equation (A10)

Equation (A11)

Footnotes

  • The filamentation instability is singled out from the Weibel one. In its original context, the Weibel instability results from a temperature anisotropy with no drift. Filamentation instability involves relative streaming of various species. In the present context, both instabilities maybe be disconnected as one filamentation stable, or unstable, beam can interact with a Weibel stable, or unstable, plasma (Bret et al. 2004, 2005). They can also interfere with each other (Bret & Deutsch 2006; Lazar et al. 2006; Lazar 2008; Lazar et al. 2009).

  • The roots of a polynomial are continuous functions of its coefficients. See Uherka & Sergott (1977) for a proof.

  • The "obliqueB0" of Table 1.

  • Denoted "Weibel" in Karlicky (2009).

  • Here again, "filamentation" is generally labeled "Weibel" in the literature.

Please wait… references are loading.
10.1088/0004-637X/699/2/990