A publishing partnership

Articles

RELATIVISTIC PAIR BEAMS FROM TeV BLAZARS: A SOURCE OF REPROCESSED GeV EMISSION RATHER THAN INTERGALACTIC HEATING

and

Published 2014 May 2 © 2014. The American Astronomical Society. All rights reserved.
, , Citation Lorenzo Sironi and Dimitrios Giannios 2014 ApJ 787 49 DOI 10.1088/0004-637X/787/1/49

0004-637X/787/1/49

ABSTRACT

The interaction of TeV photons from blazars with the extragalactic background light produces a relativistic beam of electron-positron pairs streaming through the intergalactic medium (IGM). The fate of the beam energy is uncertain. By means of two- and three-dimensional particle-in-cell simulations, we study the nonlinear evolution of dilute ultra-relativistic pair beams propagating through the IGM. We explore a wide range of beam Lorentz factors γb ≫ 1 and beam-to-plasma density ratios α  ≪  1, so that our results can be extrapolated to the extreme parameters of blazar-induced beams (γb  ∼  106 and α  ∼  10−15, for powerful blazars). For cold beams, we show that the oblique instability governs the early stages of evolution, but its exponential growth terminates—due to self-heating of the beam in the transverse direction—when only a negligible fraction ∼(α/γb)1/3  ∼  10−7 of the beam energy has been transferred to the IGM plasma. Further relaxation of the beam proceeds through quasi-longitudinal modes, until the momentum dispersion in the direction of propagation saturates at Δpb, ∥bmec  ∼  0.2. This corresponds to a fraction ∼10% of the beam energy—irrespective of γb or α—being ultimately transferred to the IGM plasma (as compared to the heating efficiency of ∼50% predicted by one-dimensional models, which cannot properly account for the transverse broadening of the beam). For the warm beams generated by TeV blazars, the development of the longitudinal relaxation is suppressed, since the initial dispersion in beam momentum is already Δpb0, ∥bmec ≳ 1. Here, the fraction of beam energy ultimately deposited into the IGM is only ∼α γb  ∼  10−9. It follows that most of the beam energy is still available to power the GeV emission produced by inverse Compton up-scattering of the cosmic microwave background by the beam pairs.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

With the current generation of Čerenkov telescopes, hundreds of TeV sources have been discovered. By far, the extragalactic TeV sky is dominated by blazars: jets from galactic centers beaming their emission toward our line of sight. The TeV photons from distant blazars cannot travel cosmological distances, since they interact with the extragalactic background light (EBL), producing electron–positron pairs. Studies of the attenuated ∼100 GeV−TeV light from distant blazars can therefore provide constraints on the strength of the EBL (e.g., Aharonian et al. 2006; Abdo et al. 2010).

The produced electron-positron pairs form a relativistic beam moving in the direction of the incident TeV photons. It is usually assumed that the energy of the pair beam is lost via inverse Compton (IC) scattering off the cosmic microwave background (CMB). As a result, the TeV radiation will be reprocessed into the GeV band (Neronov & Semikoz 2009). While cooling, the pairs gyrate around the magnetic fields intergalactic medium (IGM). Depending of the field strength and length scale, the GeV emission may form an extended source, show characteristic delays with respect to the TeV flux, or be strongly suppressed. These effects make combined GeV–TeV studies a useful probe of the IGM fields (e.g., Neronov & Vovk 2010; Tavecchio et al. 2010; Dermer et al. 2011; Dolag et al. 2011; Taylor et al. 2011; Takahashi et al. 2012; Vovk et al. 2012).

Recently, it has been proposed that the destiny of the blazar-induced beams may be different. As they stream through the IGM plasma, the electron–positron pairs are expected to trigger collective plasma instabilities (as opposed to binary Coulomb collisions, that are negligible, as discussed by Miniati & Elyiv 2013). For the parameters relevant to blazar-induced beams (i.e., dilute and ultra-relativistic), the fastest growing mode is the electrostatic oblique instability (e.g., Fainberg et al. 1970; Bret et al. 2010b), whose linear growth rate can exceed the IC cooling rate by several orders of magnitude (Broderick et al. 2012). Assuming that the instability keeps growing at the linear rate until all the beam energy is deposited into the IGM, the beam energy loss will be dominated by collective beam-plasma instabilities, rather than IC cooling. In this case, the blazar TeV emission would not be reprocessed down to multi-GeV energies, thus invalidating the IGM field estimates based on the GeV–TeV flux (Broderick et al. 2012, 2013). In addition, as a result of the beam relaxation, a substantial amount of energy would be deposited into the IGM. This "volumetric heating" can have dramatic consequences for the thermal history of the IGM (Chang et al. 2012; Pfrommer et al. 2012; Puchwein et al. 2012). However, blazar-induced heating should produce an inverted temperature-density relation in the IGM, which is not supported by the observations (Rudie et al. 2012).

While plasma instabilities could, in principle, be fast enough to thermalize the pair beam, their nonlinear stages are far more complicated than what linear dispersion analysis predicts (e.g., the studies by Miniati & Elyiv 2013 and Schlickeiser et al. 2012b, 2013 reached opposite conclusions regarding the ultimate fate of blazar-induced beams). The nature of the fastest growing instability can change as the beam-plasma system evolves, due to the suppression of temperature-sensitive modes as the beam heats up (e.g., Bret et al. 2008). Also, beam-plasma instabilities can saturate at very small amplitudes, in particular for the extremely dilute beams produced by TeV blazars (e.g., Thode & Sudan 1975; Thode 1976).

In this work, we use first-principles particle-in-cell (PIC) simulations in two and three dimensions to study the nonlinear stages and saturation of the instabilities generated as the blazar-induced pair beams propagate through the IGM. The nonlinear effects of the beam-plasma interaction are hard to capture with analytical tools, and they require fully kinetic simulations. We explore a wide range of beam Lorentz factors γb  ≫  1 and beam-to-plasma density ratios α  ≪  1, so that our results can be extrapolated to the extreme parameters of blazar-induced beams (γb  ∼  106 and α  ∼  10−15, for the most powerful blazars). We find that, for ultra-relativistic dilute beams that start with a negligible thermal spread, electrostatic beam-plasma instabilities can deposit ∼10% of the beam energy into the background electrons. However, if the beam is born with a significant momentum dispersion (as expected for blazar-induced beams), the fraction of energy going into IGM heating is as small as ∼α γb  ∼  10−9, for typical blazars. We conclude that most of the beam energy is still available to power the GeV emission produced by IC up-scattering of the CMB. This lends support to the IGM magnetic field estimates that employ the combined GeV–TeV signature of distant blazars.

The paper is organized as follows. In Section 2, we derive the typical parameters of blazar-induced pair beams in the IGM. In Section 3, we describe the setup of our PIC simulations, whose results are presented in Section 4. In particular, in Section 4.1, we focus on one representative choice of beam parameters (in the regime of dilute ultra-relativistic cold beams) and we describe the complete evolution of the beam-plasma unstable system, from the early exponential phase up to the nonlinear stages. In Section 4.2, we discuss the dependence of our findings on the beam Lorentz factor, the beam-to-plasma density contrast and the beam temperature. For the convenience of readers uninterested in the kinetic details of beam-plasma instabilities, in Section 4.3 we summarize our results in application to blazar-induced beams. Finally, in Section 5, we assess the implications of our findings for the thermal history of the IGM and the detection of reprocessed GeV emission from powerful TeV blazars.

2. PHYSICAL PARAMETERS OF BLAZAR-DRIVEN BEAMS

In this section, we summarize the physical parameters of blazar-induced beams, including the density contrast to the IGM, the beam Lorentz factor and velocity spread. We present order-of-magnitude estimates, and we refer to Schlickeiser et al. (2012a) and Miniati & Elyiv (2013) for a more detailed analysis of the beam distribution function, and its dependence on the spectrum of the EBL and of the blazar TeV emission.

Blazar photons of energy Eγ  ∼  10 TeV travel a distance of Dγγ  ≃  80 KEBL (Eγ/10 TeV)−1 Mpc before they interact with the EBL and produce electron-positron pairs (Neronov & Semikoz 2009). Here, KEBL  ∼  1 accounts for uncertainties in the intensity of the EBL, with models predicting 0.3 ≲ KEBL ≲ 3 for 0.1 TeV ≲ Eγ ≲ 10 TeV (e.g., Aharonian 2001).4 Each particle moves along the direction of the incident TeV photon, and it carries about half of the photon energy, so the beam Lorentz factor is γb  ≃  107(Eγ/10 TeV). Assuming that plasma instabilities in the IGM do not appreciably affect the beam propagation (an assumption that is correct a posteriori, as we demonstrate in this work), the pairs travel a distance of dIC  ≃  100 (Eγ/10 TeV)−1 kpc before cooling by IC scattering off the CMB (leading to many photons of energy ∼100 (Eγ/10 TeV)2 GeV per original TeV photon).

For a powerful blazar with TeV isotropic equivalent luminosity Lγ  ≃  1045Lγ, 45 erg s−1 (e.g., Ghisellini et al. 2010), the number density nb of the beam pairs is set by the balance of the pair production rate with the energy loss rate (dominated by IC cooling), which gives

Equation (1)

where the factor of dIC/Dγγ ≪ 1 accounts for the rapid energy loss of the pairs due to IC.5 If the number density in the IGM is nIGM  ∼  10−7 cm−3 (but it may be a factor of several smaller in cosmological voids, which dominate the cosmic space at z  ∼  0), the density ratio between the streaming pairs and the background plasma is

Equation (2)

The density contrast α and the beam Lorentz factor γb are the two crucial parameters determining the plasma physics of the beam-IGM interaction. For blazar-induced beams, we expect that they should vary in the range α  ∼  10−18–10−15 and γb  ∼  106–107, respectively.

As discussed by Broderick et al. (2012), collective beam-plasma effects can be relevant only if many beam pairs are present within a sphere of radius equal to the wavelength of the most unstable mode. As we show below, the scale of the fastest growing modes is 2π ce, where $c/\omega _e = \sqrt{m_{e} c^{2}/4 \pi e^{2} n_{e}} \,{\simeq} 1.5\times 10^9\, (n_e/10^{-7}{\mathrm{\;cm^{-3}}})^{-1/2}{\mathrm{\;cm}}$ is the plasma skin depth of the IGM electrons (with number density ne = nIGM/2). A sphere of skin-depth radius contains (2πce)3nb  ∼  107 (α/10−16)(ne/10−7 cm−3)−1/2 beam particles. For α  ∼  10−18–10−15, we find that collective phenomena always play a role in the evolution of blazar-induced beams.

Another important parameter is the dispersion in beam momentum at birth. Since the pair-creation cross section peaks slightly above the threshold energy, the pairs are born moderately warm, with a comoving temperature of kBTb ∼ mec2.6 Moreover, since the EBL and the blazar TeV spectra are broad, the beam energy distribution will extend over a wide range of Lorentz factors, so that the effective beam temperature can be as large as kBTb/mec2  ∼  5–10 (Miniati & Elyiv 2013). In Section 4.2.2, we explore the role of thermal effects on the nonlinear evolution of blazar-induced beams.

3. SIMULATION SETUP

We investigate blazar-driven plasma instabilities in the IGM by means of fully kinetic PIC simulations. We employ the three-dimensional (3D) electromagnetic PIC code TRISTAN-MP (Spitkovsky 2005), which is a parallel version of the publicly available code TRISTAN (Buneman 1993), that was optimized for handling ultra-relativistic flows (see, e.g., Sironi & Spitkovsky 2011 and Sironi et al. 2013 for studies of ultra-relativistic collisionless shocks using TRISTAN-MP). We initialize a relativistic dilute pair beam that propagates along $+ {\boldsymbol{\hat{\boldsymbol x}}}$ through an unmagnetized electron-proton plasma (with the realistic mass ratio mp/me = 1836). The simulations are performed in the frame of the background plasma, i.e., of the IGM. No background magnetic field is assumed, so the electric and magnetic fields generated by beam-plasma instabilities will grow from noise.

To follow the beam-plasma evolution to longer times with fixed computational resources, we mainly utilize 2D computational domains in the xy plane. In Section 4.1, we compare 2D and 3D runs, and we show that 2D simulations can capture most of the relevant 3D physics. In the case of 2D simulations with the beam lying in the simulation plane, only the in-plane components of the velocity, current and electric field, and only the out-of-plane component of the magnetic field are present. The simulation box is periodic in all directions, i.e., we assume that the spatial scale of the gradients (e.g., in the IGM density) is much larger than the box length. By choosing a periodic domain, we simulate the bulk of the beam-plasma system, rather than the "head" of the pair beam.

The background plasma consists of cold electrons and protons, with initial electron temperature kBTe/mec2  ≃  10−8. We have tested that higher temperatures of the background electrons do not significant change the development of the relevant instabilities, as long as the electron temperature is non-relativistic, in agreement with Bret et al. (2005). The background protons are allowed to move, but we obtain similar results when the protons are treated as a static (i.e., infinitely massive) charge-neutralizing background. It follows that nonlinear Landau damping,7 which would be artificially suppressed for infinitely massive protons, does not seem to play a major role in the evolution of the beam-plasma system, for the parameters explored in our study.

The beam consists of electron-positron pairs propagating with Lorentz factor γb along the $+ {\boldsymbol{\hat{\boldsymbol x}}}$ direction. To our knowledge, our PIC simulations are the first to address the evolution of an electron-positron beam. All of the previous studies have focused on the case of an electron beam propagating through an electron-proton plasma, with the background electrons moving opposite to the beam to compensate for the beam current (e.g., Dieckmann et al. 2006b; Bret et al. 2008; Kong et al. 2009). Any instability triggered by the relative drift between the background electrons and protons (e.g., the Buneman 1958 instability) will then be absent in our setup, where the pair beam carries no net current.

The beam-to-plasma density ratio α and the beam Lorentz factor γb expected for blazar-induced pairs streaming through the IGM (see Section 2) cannot be directly studied with PIC simulations. However, by performing dedicated experiments with a broad range of α and γb (in the regime α ≪ 1 and γb ≫ 1 of ultra-relativistic dilute beams), we can extrapolate the relevant physics to the extreme parameters expected in the IGM. We vary the beam Lorentz factor from γb = 3 up to γb = 1000, and the density contrast from α = 10−1 down to α = 10−3.8 For numerical convenience, the density ratio between the beam and the plasma is established by initializing the same number of beam and plasma computational particles, with the beam particles having a weight α. We have tested that, by choosing a different weight (yet, keeping the same physical density contrast), our results do not change. In addition to studying the dependence on γb and α, we also compare the evolution of cold beams (with comoving temperature at initialization kBTb/mec2  ≃  10−4) with the case of warm beams, up to the limit of mildly relativistic thermal spreads kBTb/mec2  ∼  1 most relevant for blazar-induced beams.

The results presented below have been extensively tested for convergence. We typically employ 50 particles per computational cell for the background plasma (25 electrons and 25 protons), and the same number for the beam particles (if each carries a weight α). However, we have tested that our 2D results are the same when using up to 256 particles per cell, for both the beam and the plasma. We resolve the skin depth ce of the background electrons with 8 computational cells, but we have tested that our 2D results do not change when using 12 or 16 cells per skin depth.9 In 2D runs, the simulation plane is typically a square with 1024 cells (∼125 ce) on each side, but we have checked that our results do not substantially change when employing a larger box that is 500 ce long (in the direction of beam propagation) and 250 ce wide. In 3D, we employ a box with 512 cells (∼67.5 ce) in the transverse direction and 1024 cells (∼125 ce) along the longitudinal direction.10 To capture the linear and nonlinear stages of the beam-plasma evolution, we follow the system up to unprecedentedly long times, in 2D up to ωet  ∼  105, or equivalently ∼1.75 × 106 timesteps, and in 3D up to ωet  ∼  4 × 104, or ∼7 × 105 timesteps.

The number of beam particles is kept constant during the evolution of the beam-plasma system, since the photon–photon interactions that would introduce fresh electron–positron pairs are extremely rare on the timescales covered by our simulations. Also, we neglect IC cooling of the beam pairs, since it is irrelevant over the timespan of our runs. This implies that the total energy in our periodic beam-plasma system should be constant over time. However, explicit PIC codes do not conserve energy to machine precision. We track the energy conservation in our runs, and we find that at late times it is still better than 1%. This makes our estimates of the amount of beam energy transferred to the plasma electrons (of the order of ∼10%) extremely robust, for the beam parameters explored in this work.

Finally, we remark that in all PIC codes a numerical heating instability arises when cold relativistic plasma propagates for large distances over the numerical grid. The instability, known as "numerical Čerenkov," results from the coupling of electromagnetic waves with spurious beam mode aliases on the computational grid (Dieckmann et al. 2006a; Godfrey & Vay 2013; Xu et al. 2013). The numerical Čerenkov instability might artificially slow down the beam, even in the absence of physical beam-plasma instabilities. We have assessed that the results reported below arise from a physical instability (as opposed to the numerical Čerenkov mode), by comparing our beam-plasma simulations with the artificial case of a beam that propagates through the computational grid in the absence of any background plasma. The beam evolution in the two cases is dramatically different, which proves that the beam energy loss that we discuss below arises from the physical interaction of the beam with the background plasma, rather than from the numerical Čerenkov instability.

4. RESULTS

In this section, we explore the linear and nonlinear evolution of ultra-relativistic dilute pair beams by means of 2D and 3D PIC simulations. In Section 4.1, we describe the different stages of evolution of the beam-plasma system, for a representative choice of beam parameters in the regime of ultra-relativistic dilute beams (γb = 300, α = 10−2 and negligible beam thermal spread at initialization). In Section 4.2, we investigate the dependence of our results—in particular, of the fraction of beam energy transferred to the background electrons—on the beam-to-plasma density contrast, the beam Lorentz factor and the beam temperature at birth. The reader that is not interested in the kinetic details of the beam-plasma interaction might proceed to Section 4.3, where we extrapolate the findings of our PIC simulations to the extreme parameters of blazar-induced beams.

4.1. The Linear and Nonlinear Evolution of Ultra-relativistic Dilute Cold Beams

In this section, we follow the evolution of a cold beam with γb = 300 and α = 10−2. We start with the analysis of the linear phase, and then we investigate the nonlinear relaxation. We find that the exponential phase of the oblique mode (which is the fastest growing instability for dilute ultra-relativistic beams) terminates due to self-heating of the beam in the direction transverse to the beam motion. At the end of the oblique phase, only a minor fraction ∼(α/γb)1/3 of the beam energy has been deposited into the background electrons. Further evolution of the beam is governed by quasi-longitudinal modes, which operate on a timescale that is much longer (at least two orders of magnitude) than the oblique growth. At the end of the quasi-longitudinal phase, the dispersion of beam momentum in the longitudinal direction saturates at Δpb, ∥bmec  ∼  0.2, which corresponds to a fraction of ∼10% of beam energy transferred to the plasma.

4.1.1. The Oblique Exponential Phase

The evolution of the beam-plasma system during the oblique phase is presented in panels (a)–(d) of Figure 1. The beam is set up with a small thermal spread (kBTb/mec2  ≃  10−6), so that the oblique instability initially proceeds in the reactive regime, i.e., all the beam particles are in resonance with each harmonic of the packet of unstable modes, and the instability is the strongest. This is opposed to the kinetic regime, in which the beam velocity spread is considerable. Here, a number of unstable modes with a broad spectrum in phase velocity will be excited, with only a small number of beam particles being in resonance with each mode. This results in a slower growth, as compared to the reactive regime.

Figure 1.

Figure 1. Temporal evolution of the beam-plasma interaction, from the 2D simulation of a cold beam with γb = 300 and α = 10−2. We follow the evolution of the system through the exponential phase of the oblique mode (panels (a)–(d)) until the relaxation stage (panels (e)–(h)). Panels (a) and (e): fraction of beam kinetic energy transferred to the plasma electrons (orange), to the longitudinal and transverse electric fields (red and blue, respectively), and to the transverse magnetic fields (green). In panel (a), the dotted orange line shows the growth rate of the oblique mode expected from linear dispersion analysis. Panels (b) and (f): temporal evolution of the momentum dispersion of the beam (solid) and plasma (dashed) electrons, along the beam (red) or transverse to the beam (blue). The momenta are in units of mec. Panels (c), (d), and (g): 2D plot of the longitudinal electric field E, in units of $\sqrt{{{{8 \pi \gamma _b n_b m_e c^2}}}}$. The electric field is shown at three different stages of evolution, as marked by the red arrows at the bottom of panels (a) and (e). Panel (h): 2D plot of the transverse magnetic field B, in units of $\sqrt{{{{8 \pi \gamma _b n_b m_e c^2}}}}$, at the time marked by the green arrow at the bottom of panel (e).

Standard image High-resolution image

In Figure 1, we confirm that the oblique instability is the fastest growing mode for ultra-relativistic dilute beams. The reactive phase of the instability governs the evolution of the system for $t\,{\lesssim}\, t_{\rm OBL}\,{\simeq}\,600\,\omega _e^{-1}$, where tOBL is marked as a dash-dotted vertical blue line in panels (a) and (b). Here, $\omega _e=\sqrt{4\pi e^2 n_e/m_e}$ is the plasma frequency of the background electrons. The fastest mode grows on a scale ∼2πce, where ce is the electron skin depth, and its wavevector is inclined at ∼45° with respect to the beam propagation. This is apparent in the 2D structure of the longitudinal electric field in Figure 1(c), as well as in the 2D plots of the transverse electric and magnetic fields (not shown).11 The oblique mode is also captured in 3D simulations, as shown in the 3D structure of the longitudinal electric field of Figure 2(a).

Figure 2.

Figure 2. 3D structure of the longitudinal electric field E, from the 3D simulation of a cold beam with γb = 300 and α = 10−2. The electric field is in units of $\sqrt{{{{8 \pi \gamma _b n_b m_e c^2}}}}$. The three snapshots are taken at the same times of panels (c), (d), and (g) in Figure 1, and they show that the 3D physics of the relevant electrostatic beam-plasma instabilities can be correctly captured by our 2D runs. In the 3D simulations, the temporal evolution of the fraction of beam kinetic energy transferred to the background electrons and to the electromagnetic fields (not shown here), as well as the time evolution of the beam and plasma momentum dispersions (still not shown here), closely follow the 2D results presented in Figures 1(a), (b), (e), and (f). The equivalence of 2D and 3D results for cold beams is indeed expected on analytical grounds (e.g., Bret et al. 2010b).

Standard image High-resolution image

The growth rate of the oblique instability in the reactive regime is (e.g., Fainberg et al. 1970)

Equation (3)

where the different dependence on k and k is related to the fact that for relativistic beams the transverse inertia is much smaller than the longitudinal inertia (by a factor of $\gamma _b^2$), so that the modes transverse to the beam are the easiest to be excited (for an intuitive physical description, see Nakar et al. 2011).12 From the pattern in Figure 1(c), we infer $k_\perp \,{\sim}\,k_{\parallel }\,{\sim}\,k/\sqrt{2}$, so the growth rate of the fastest growing oblique mode will be

Equation (4)

which nicely agrees with our results.13 In fact, in Figure 1(a) we show that the fraction of beam kinetic energy deposited into the background electrons (orange line), into the longitudinal (red) and transverse (blue) electric fields, and into the transverse magnetic field (green) all grow at the rate predicted by Equation (4) for $t\lesssim 600\,\omega _e^{-1}$ (dotted orange line in Figure 1(a)).

The oblique mode is quasi-electrostatic, i.e., roughly kEk (e.g., Bret et al. 2010b). Since the angle between the wavevector and the beam is ≳ 45°, in agreement with analytical expectations (e.g., Bret et al. 2010b), it follows that the electric field component transverse to the beam is slightly larger than the longitudinal component, i.e., kk implies that EE. This explains the small difference between the red and the blue lines in Figure 1(a). Also, since the mode is quasi-electrostatic, the magnetic component will be sub-dominant relative to the electric fields. In agreement with the analytical considerations of Lemoine & Pelletier (2010), we find that B  ∼  2 δOBLE. Given that δOBL ≪ 1 for ultra-relativistic dilute beams, it follows that B ≪ E.

For electrostatic modes in the linear phase, it is expected that the amount of kinetic energy lost by the beam should be equally distributed between electric fields and plasma heating (e.g., Thode 1976). This explains why the energy epsilone of the background electrons in the exponential phase (orange line in Figure 1(a) at $t\lesssim t_{\rm OBL} \,{\simeq}\,600\,\omega _e^{-1}$) is comparable to the energy in electric fields (epsilonE, ∥ and epsilonE, ⊥, respectively, red and blue lines in Figure 1(a)). We have verified that the fraction of beam energy transferred to the background protons is negligible as compared to the plasma electrons, so the evolution of the protons will be ignored hereafter.

Since E  ∼  E during the exponential phase of the oblique mode, both the plasma and the beam are heated quasi-isotropically, so that the momentum spreads in the longitudinal and transverse directions are nearly identical (compare the red and blue lines for ttOBL in Figure 1(b); dashed lines refer to the plasma, solid lines to the beam). The transverse momentum spread of the beam can be related to the transverse electric field E via the Lorentz force

Equation (5)

where we have assumed that the characteristic timescale is set by the oblique growth rate (i.e., Δt−1  ∼  δOBL ωe) and that the magnetic force is negligible compared to the electric force (in fact, B/E  ∼  2 δOBL ≪ 1).

Since the oblique mode is heating up the beam in the transverse direction (solid blue line in Figure 1(b)), the exponential growth at the reactive rate ωOBL will necessarily terminate, when the assumption of a cold beam required by the reactive approximation becomes invalid. It is well known that the system will transition from the reactive phase to the kinetic phase when the beam velocity dispersion ${\boldsymbol {\Delta v_b}}$ reaches (e.g., Fainberg et al. 1970; Bret et al. 2010b)

Equation (6)

namely, when the beam, due to its velocity spread, can move across one wavelength of the most unstable mode during the growth time of the reactive instability. In this case, most of the beam particles will lose resonance with the unstable mode, and the instability will transition from the reactive to the kinetic regime. For ultra-relativistic beams with isotropic momentum dispersions (in fact, Figure 1(b) shows that Δpb, ⊥  ∼  Δpb, ∥ during the oblique reactive phase), the transverse velocity spread Δvb, ⊥/c  ∼  Δpb, ⊥bmec is much larger than the longitudinal spread $\Delta v_{b,\parallel }/c\,{\sim}\,(\Delta v_{b,\perp }/c)^2+\Delta p_{b,\parallel }/\gamma _b^3m_e c$. Since k  ∼  k  ∼  ωe/c, Equation (6) above reduces to

Equation (7)

where Δpb, OBL is the expected transverse dispersion in beam momentum at the end of the reactive oblique phase. The threshold in momentum dispersion Δpb, OBL can also be recast as a limit in beam temperature (e.g., Bret et al. 2010a). We confirm that the reactive phase of the oblique mode terminates at $t\,{\sim}\,t_{\rm OBL}\,{\sim}\,600\,\omega _e^{-1}$, when the beam transverse momentum reaches the threshold Δpb, OBL in Equation (7) (which is shown as a horizontal dash-dotted blue line in Figure 1(b)).

We can now derive the expected fraction of beam kinetic energy transferred to the plasma electrons and to the electromagnetic fields at the end of the oblique reactive phase. By setting Δpb, ⊥ = Δpb, OBL in Equation (5), we find that the fraction of beam energy converted into transverse electric fields at t  ∼  tOBL is

Equation (8)

Since the oblique mode is quasi-electrostatic (i.e., E  ∼  E), it follows that epsilonE, ⊥  ∼  epsilonE, ∥. Moreover, for electrostatic modes, the fraction epsilone of the beam kinetic energy converted into plasma heating is comparable to the energy in electric fields (e.g., Thode 1976), so epsilone  ∼  epsilonE, ⊥  ∼  epsilonE, ∥. Finally, since B  ∼  2 δOBLE, the magnetic energy fraction will be $\epsilon _{B,\perp }\,{\sim}\,\sqrt{27}\,\delta _{\rm OBL}^3/8$. We have extensively verified that the expected scalings of the efficiency parameters epsilone, epsilonE, ⊥, epsilonE, ∥ and epsilonB, ⊥ with respect to δOBL are in agreement with the results of our simulations, across the whole range of beam Lorentz factors and density contrasts we have explored (see the various curves in Figure 1(a) at t  ∼  tOBL, and also Section 4.2.1).

For ttOBL, the evolution of the oblique mode will proceed in the kinetic (rather than reactive) regime. The kinetic oblique mode is indeed responsible for the peak in the electric field energy observed at ωet  ∼  1000 in Figure 1(a) (red and blue lines), which produces a moderate increase in the fraction of beam energy transferred to the background electrons (orange line in Figure 1(a) at ωet  ∼  1000). In this phase, the 2D structure of the longitudinal electric field in Figure 1(d) shows that the wavevector of the kinetic oblique mode is oriented at ∼20° relative to the beam propagation (as compared to the ∼45° angle observed during the reactive phase, see Figure 1(c)). A similar pattern is shown in the 3D plot of Figure 2(b). As expected, the increase in the transverse momentum dispersion has suppressed the modes having kk, which are most sensitive to transverse temperature effects (e.g., Bret et al. 2010b).

For a beam with initial transverse velocity dispersion Δv0, ⊥, the growth rate of the kinetic oblique mode for k ≲ ωe/c is (e.g., Breiˇzman & Ryutov 1971)

Equation (9)

At the end of the reactive oblique phase, Equation (7) prescribes that Δv0, ⊥/c = Δpb, OBLbmec  ∼  δOBL, so that the growth rate of the kinetic oblique mode will be ωk∝δOBL ωe, i.e., it will have the same scalings with α and γb as the reactive oblique mode (here, we have neglected factors of the order of unity).

We have explicitly verified that the peak in electric fields at ωet  ∼  1000 is due to the kinetic oblique mode, by performing a dedicated simulation in which at ωet  ∼  800 (i.e., shortly after the end of the reactive stage) we reset by hand the electromagnetic fields and the plasma temperature to their initial values (i.e., no seed fields and kBTe/mec2  ≃  10−8), yet we retain the beam momentum distribution that results self-consistently from the reactive oblique phase. In this setup, we find that the fastest growing mode has the same 2D pattern as in Figure 1(d) and its growth rate scales as ∝δOBL ωe. This confirms that the peak in electric fields at ωet  ∼  1000 (Figure 1(a)) is indeed associated to the kinetic oblique instability.

The growth of the kinetic oblique mode terminates due to self-heating of the beam, in analogy to the reactive oblique phase. In particular, the growth in the kinetic oblique phase cannot be sustained beyond the point where, due to the self-excited electric fields, the beam momentum dispersion in the transverse direction exceeds the initial value γbΔv0, ⊥me. At this point, the expression for the growth rate in Equation (9) becomes clearly invalid. This happens when

Equation (10)

which leads to the same scaling as in Equation (8), if we take Δv0, ⊥/c = Δpb, OBLbmec  ∼  δOBL, as appropriate for the beam velocity dispersion at the end of the reactive oblique phase. In summary, apart from factors of the order of unity, the electric fields at the end of the kinetic oblique phase will saturate at a level similar to the reactive oblique stage (compare the two peaks of electric energy in Figure 1(a) at ωet  ∼  600 and ωet  ∼  1000). It follows that the kinetic oblique instability will increase the fraction of beam kinetic energy transferred to the plasma electrons only by a factor of the order of unity, as compared to the reactive oblique mode (see the orange line in Figure 1(a), for ωet ≳ 1000).

4.1.2. The Longitudinal Relaxation Phase

After the saturation of the oblique instability, further evolution of the beam-plasma system proceeds via quasi-longitudinal modes, as shown in the 2D plot of the longitudinal electric field in Figure 1(g), as well as in the 3D pattern of Figure 2(c). Being quasi-longitudinal, these modes are insensitive to thermal spreads in the direction perpendicular to the beam, so they can grow even after the end of the kinetic oblique phase.

The quasi-longitudinal oscillations shown in Figure 1(g) are a characteristic signature of the quasi-linear relaxation of the beam (see, e.g., Grognard 1975; Lesch & Schlickeiser 1987; Schlickeiser et al. 2002; Pavan et al. 2011). In the quasi-linear relaxation, the beam generates longitudinal Langmuir waves (see the peak in E at ωet  ∼  104 in Figure 1(e)), which scatter the beam particles and heat the background plasma (see the increase in the electron thermal energy shown by the orange line at 104 ≲ ωet ≲ 1.5 × 104 in Figure 1(e)). Since EE (compare the red and blue curves in Figure 1(e) at ωet  ∼  104), the background electrons will be heated preferentially in the direction of motion of the beam (see the increase in Δpe, ∥ at 104 ≲ ωet ≲ 1.5 × 104 in Figure 1(f)). For the same reason, the quasi-linear relaxation is accompanied by a substantial increase in the beam momentum spread along the direction of propagation (red solid line in Figure 1(f), showing the growth of Δpb, ∥ at 5 × 103 ≲ ωet ≲ 1.5 × 104). The spread in the parallel beam momentum is associated to the formation of phase space holes, that result from the trapping of beam particles by the longitudinal electric oscillations (e.g., O'Neil et al. 1971; Thode & Sudan 1975).

The quasi-linear relaxation occurs on a timescale much longer than the exponential oblique phase. Within the range of beam Lorentz factors and density contrasts probed by our simulations, we find that the characteristic relaxation time τR is at least two orders of magnitude longer than the exponential growth time of the oblique instability $\tau _{\rm OBL}=\omega _{\rm OBL}^{-1}$, in agreement with previous 1D simulations (Grognard 1975; Pavan et al. 2011). This emphasizes the importance of evolving our PIC simulations to sufficiently long times to capture the physics of the quasi-linear relaxation (see Section 4.2 for further details).

The quasi-linear modes broaden the beam momentum spectrum in the longitudinal direction up to the point where Δpb, ∥bmec  ∼  0.2 (see the red solid line in Figure 1(f), saturating at Δpb, ∥/mec  ∼  0.2 γb  ∼  60). This in agreement with the so-called Penrose's criterion, stating that the beam will be stable to electrostatic modes only when the longitudinal dispersion in momentum approaches the initial beam Lorentz factor (e.g., Buschauer & Benford 1977). From Δpb, ∥bmec  ∼  0.2, it follows that at the end of the relaxation phase, a fraction ∼10% of the beam energy has been transferred to the background electrons (see the orange line in Figure 1(e) at ωet ≳ 2 × 104). In Section 4.2.1 we demonstrate that, irrespective of the beam Lorentz factor or the beam-to-plasma density contrast, a generic by-product of the relaxation of cold ultra-relativistic dilute beams is the conversion of ∼10% of their energy into plasma heating.14

The quasi-linear relaxation significantly affects the shape of the beam and plasma longitudinal momentum spectrum, as shown in Figure 3. As a result of the quasi-longitudinal relaxation, the plasma distribution at late times (ωet ≳ 2 × 104) develops a pronounced high-energy tail (at 3 ≲ p/mec ≲ 102), that bridges the main thermal peak of the plasma electrons (at p/mec  ∼  0.5) with the beam particles (that populate the isolated high-energy bump at p/mec  ∼  300 in Figure 3). This high-energy component in the background electrons, which contains a significant amount of energy, is present only in the forward direction (i.e., along the beam propagation). In the backward direction, the spectrum of plasma electrons (red dotted line in Figure 3, at ωet = 5 × 104) is compatible with a Maxwellian.

Figure 3.

Figure 3. Temporal evolution of the longitudinal momentum spectrum pdN/dp, for a beam-plasma system with γb = 300 and α = 10−2. The momentum is in units of mec. The beam, which can be identified with the isolated peak at p  ∼  300, stops evolving after ωet ≳ 1.5 × 104, in agreement with Figure 1(f). The beam spectrum at late times does not relax to the so-called "plateau" distribution ${dN}/{dp}_\parallel \propto p_\parallel ^0$, which is indicated as a black dotted line. We find that the beam approaches the plateau distribution only if we artificially inhibit the evolution of the beam transverse momentum, i.e., we force the beam relaxation to proceed in a quasi-1D configuration (red dashed line). As a result of the beam relaxation, the plasma develops a high-energy tail in the forward direction. For comparison, the plasma distribution in the backward direction (i.e., opposite to the beam) is shown as a dotted red line.

Standard image High-resolution image

During the quasi-linear relaxation, the beam spectrum evolves from a quasi-monoenergetic distribution into a broad bump (from the black to the red curve at p/mec  ∼  300 in Figure 3). In agreement with Figure 1(f) (red solid line), most of the evolution occurs at ωet ≲ 1.5 × 104, whereas the beam spectrum at longer times is remarkably steady (we have followed the system up to ωet  ∼  1.5 × 105, finding no further signs of evolution).

We point out that the beam momentum spectrum at late times does not approach the so-called "plateau" distribution ${dN}/{dp}_\parallel \propto p_\parallel ^0$ (indicated as a black dotted line in Figure 3), which is believed to be the ultimate outcome of the beam relaxation in 1D (e.g., Grognard 1975; Schlickeiser et al. 2002; Dieckmann et al. 2013). As opposed to earlier 1D claims, in our 2D and 3D simulations we find that the beam longitudinal relaxation leads to a momentum spectrum that is harder than the plateau distribution, yet the beam-plasma system appears stable.15 In turn, the fact that the beam spectrum is harder than ${dN}/{dp}_\parallel \propto p_\parallel ^0$ explains why the amount of beam energy transferred to the plasma is only ∼10% (it should be ∼50% for a plateau distribution extending up to γbmec, see Thode & Sudan 1975).

We argue that the transverse dispersion in beam momentum, which could not be properly captured in previous 1D studies, prevents the longitudinal beam spectrum from relaxing to the plateau distribution (see Appendix B for further details). Our claim is supported by the following experiment. At the end of the kinetic oblique phase (ωet  ∼  2000), we artificially set the beam transverse dispersion to be Δpb, ⊥/mec ≪ 1 (for comparison, the self-consistent evolution in Figure 1(b) yields Δpb, ⊥/mec  ∼  10 at the end of the kinetic oblique phase, see the solid blue line). Also, in the subsequent evolution, we inhibit any growth in the transverse beam momentum. In this setup, in which any transverse dispersion effects are artificially neglected, the beam relaxation should proceed as in 1D. So, it is not surprising that, in agreement with previous 1D studies, at late times the beam relaxes to the plateau distribution (red dashed line in Figure 3). In other words, we find that the relaxation to the plateau distribution is not a general result of the multi-dimensional evolution of ultra-relativistic beams, but it only occurs when the beam transverse dispersion stays sufficiently small, so the system is quasi-1D. We now provide a qualitative explanation of this important aspect of the physics of beam-plasma interactions, deferring a more detailed analytical treatment to a future publication.

As we further discuss in Appendix B, the beam relaxation produces a plateau distribution only if Δpb, ⊥/mec ≪ 1, due to the following argument. Complete stabilization of the beam-plasma system is achieved when the longitudinal velocity spread of the ultra-relativistic beam reaches Δvb, ∥/c  ∼  1 (more precisely, when the velocity spread is comparable to the beam speed). The spread in longitudinal velocity includes contributions from both the longitudinal and the transverse momentum dispersions:

Equation (11)

where the second term on the right hand side is absent in the case of 1D relaxation. It follows that the transverse momentum spread can appreciably modify the relaxation process only if $\Delta p_{b,\perp }/m_e c\gtrsim \sqrt{\Delta p_{b,\parallel }/\gamma _b m_e c}\,{\sim}\,1$, where we have used that Δpb, ∥bmec  ∼  0.2 at the end of the relaxation phase. Then, the fact that in the case studied in Figure 1(b) the transverse beam dispersion after the oblique phase is Δpb, ⊥/mec  ∼  10 explains why the beam relaxation cannot lead to a plateau distribution. In Appendix B, we provide further evidence that the relaxation to a plateau distribution requires Δpb, ⊥/mec ≪ 1.

4.1.3. The Magnetic Field Growth

In the previous subsections, we have primarily focused on the electrostatic character of the growing modes, which determines the coupling efficiency between the beam energy and the plasma thermal energy. Here, we comment on the generation of magnetic fields associated with the evolution of the beam-plasma system.

As shown in Figure 1(a), the growth of the quasi-electrostatic oblique mode (both in the reactive and in the kinetic regime) is accompanied by a minor magnetic component. In Section 4.1.1, we have estimated that the fraction of beam kinetic energy transferred to the magnetic fields at the end of the oblique phase is $\epsilon _{B,\perp }\,{\sim}\,\delta _{\rm OBL}^3$, apart from factors of the order of unity. Since δOBL ≪ 1 for ultra-relativistic dilute beams, the magnetic fields generated by the oblique instability are generally unimportant.

At the end of the relaxation phase, the beam and the plasma are highly anisotropic, with the longitudinal momentum spread much larger than the transverse one (see Figure 1(f) at ωet  ∼  1.5 × 104). As a result, the system is prone to the Weibel instability (e.g., Weibel 1959; Yoon & Davidson 1987; Sakai et al. 2000; Silva et al. 2002, 2003; Jaroschek et al. 2005), which generates the transverse magnetic field pattern shown in Figure 1(h).16 As a result of the Weibel instability, the magnetic field energy increases (see the green line in Figure 1(e), at 104 ≲ ωet ≲ 2 × 104), and the beam and plasma anisotropy is reduced by increasing the transverse momentum spread (see the solid and dashed blue lines at 104 ≲ ωet ≲ 2 × 104 in Figure 1(f)).

The Weibel instability is predominantly magnetic, so it does not mediate any significant exchange of energy from the beam to the plasma electrons, and the heating efficiency epsilone does not increase beyond the value epsilone  ∼  10% attained at the end of the relaxation phase. However, the Weibel instability might be a promising source for the generation of magnetic fields, as discussed by Schlickeiser et al. (2012b). However, the evolution of the magnetic filaments shown in Figure 1(h) can only be captured with large-scale 3D simulations (e.g., Bret et al. 2008; Kong et al. 2009). A detailed 3D investigation of the strength and scale of the magnetic fields resulting from ultra-relativistic dilute pair beams from TeV blazars will be presented elsewhere.

4.2. Dependence on the Beam Parameters

In this section, we discuss the dependence of the beam-plasma evolution on the beam parameters. In Section 4.2.1, we consider the case of cold beams, and we show that the quasi-longitudinal relaxation leads to a beam momentum spread along the direction of motion Δpb, ∥bmec  ∼  0.2, regardless of the beam Lorentz factor γb or the beam-to-plasma density contrast α. In turn, this implies that a fraction ∼10% of the beam energy is converted into heat of the background plasma, irrespective of γb or α.

In Section 4.2.2, we discuss the effect of the initial beam thermal spread on the efficiency of the beam-to-plasma energy transfer. We find that if the initial dispersion in longitudinal momentum satisfies Δpb0, ∥bmec ≳ 0.2 (as typically expected for blazar-induced beams), the fraction of beam energy deposited into the background plasma is much smaller than ∼10%.

4.2.1. Cold Beams

In Figure 4, we show how the evolution of the beam-plasma system depends on the beam Lorentz factor (that we vary from γb = 3 up to γb = 1000) and on the beam-to-plasma density contrast (from α = 3 × 10−2 down to α = 3 × 10−3). Most of the previous studies have focused on moderately relativistic electron beams (γb = 3–6) with α = 10−1 (e.g., Gremillet et al. 2007; Bret et al. 2008; Kong et al. 2009). Here, we investigate the case of ultra-relativistic dilute electron-positron beams, as appropriate for blazar-induced beams.

Figure 4.

Figure 4. Dependence of the beam-plasma evolution on the beam Lorentz factor (from γb = 3 in black up to γb = 1000 in red, in each panel; see the legend in panels (d)–(f)) and on the beam-to-plasma density contrast (α = 3 × 10−2 for the leftmost column, α = 10−2 for the middle column, and α = 3 × 10−3 for the rightmost column). Panels (a)–(c): fraction of the beam kinetic energy transferred to the plasma electrons (in the inset, a zoom-in on the earliest phases of evolution). Panels (d)–(f): beam momentum dispersion in the longitudinal direction, normalized to the initial beam momentum (i.e., Δpb, ∥bmec). Panels (g)–(i): beam (solid) and total (dashed) momentum spectrum pdN/dp in the longitudinal direction, at the time indicated with the dotted black lines in the upper two rows. In panels (g)–(i), the slope ${dN}/{dp}_\parallel \propto p_\parallel ^0$ expected for the plateau distribution is shown as a dotted black line.

Standard image High-resolution image

For each choice of γb and α, we follow the beam-plasma system from the oblique phase until the quasi-linear relaxation (typically, up to ωet  ∼  105). We initialize a cold beam with thermal spread kBTb/mec2  ≃  10−4, so that the oblique instability initially proceeds in the reactive regime, for the range of γb and α covered by our simulations. In the reactive phase, we confirm that the fastest growing mode has a wavevector oriented at ∼45° to the beam direction of propagation. The growth rate is in excellent agreement with Equation (4). As described in Section 4.1.1, the exponential phase of the reactive oblique mode terminates due to self-heating of the beam in the transverse direction. At the end of the reactive phase, we find that the fractions of beam kinetic energy transferred to the plasma electrons, to the electric fields and to the magnetic fields scale, respectively, as epsilone∝δOBL, epsilonE, ⊥  ∼  epsilonE, ∥∝δOBL and $\epsilon _{B,\perp }\propto \delta _{\rm OBL}^3$, as in Section 4.1.1.

The reactive oblique phase is followed by the kinetic oblique phase. We find that the characteristic wave pattern in the kinetic oblique phase (see Figure 1(d) in 2D and Figure 2(b) in 3D) appears in the evolution of all the beam-plasma systems that we present in Figure 4. In the temporal evolution of the fraction of beam energy converted into plasma heating, the kinetic oblique mode is responsible for the additional increase that is seen in the most relativistic cases (γb ≳ 100) after the end of the reactive oblique phase (see the insets in the top row of Figure 4). Regardless of γb or α, we find that the kinetic oblique phase deposits only a fraction ∼δOBL of the beam kinetic energy into the background electrons, i.e., comparable to the reactive oblique phase (see Section 4.1.1).

The long-term evolution of the system is controlled by the quasi-longitudinal relaxation, which operates on a timescale τR  ≳  102 τOBL, where $\tau _{\rm OBL}=\omega _{\rm OBL}^{-1}$ is the characteristic e-folding time of the oblique mode. In Figure 4, the quasi-linear relaxation governs the growth in the plasma thermal energy (top row) and in the beam parallel momentum spread (middle row) occurring at ωet ≳ 5000. In the regime γb ≫ 1, the quasi-linear relaxation terminates when the beam momentum spread in the longitudinal direction reaches Δpb, ∥bmec  ∼  0.2, regardless of γb or α (middle row in Figure 4). Correspondingly, the fraction of beam energy transferred to the plasma saturates at epsilone  ∼  10% (top row in Figure 4).

Mildly relativistic beams with moderate density contrasts deviate from such simple scalings, for the following reason. The quasi-linear relaxation will not operate if the beam dispersion at the end of the oblique phase is already Δpb, ∥bmec ≳ 0.2. According to Equation (7), the beam spread at the end of the oblique phase is Δpb, ∥  ∼  Δpb, ⊥  ∼  δOBLγbmec, so that the quasi-linear relaxation will be suppressed if δOBL  ∼  (α/γb)1/3 ≳ 0.2, i.e., for mildly relativistic beams with moderate α.

A similar argument explains why a plateau distribution in the longitudinal momentum spectrum (bottom row in Figure 4) is established only for relatively small γb. As we have argued in Section 4.1.2, a transverse spread Δpb, ⊥/mec ≳ 1 prevents the beam relaxation to the plateau distribution (shown as a dotted black line in the bottom row of Figure 4). At the end of the oblique phase, Δpb, ⊥/mec  ∼  γb δOBL, so that the beam will relax to the plateau distribution only if $\gamma _b \delta _{\rm OBL}\,{\sim}\,(\gamma _b^2 \alpha)^{1/3}\ll 1$. Clearly, this constraint is hardest to satisfy for highly relativistic beams, which explains why, at fixed α, the momentum spectrum of more relativistic beams shows stronger deviations from the plateau distribution.

4.2.2. Hot Beams

In the previous subsection, we have assumed that the beam is born with a negligible thermal spread. Here, we discuss how the results presented above for cold beams will be modified by temperature effects. We describe separately the role of thermal spreads in the development of the oblique instability, and in the relaxation phase.

As we have anticipated in Section 4.1.1, the oblique instability will proceed in the kinetic (rather than reactive) regime if the initial beam dispersion in the transverse direction is Δv0, ⊥/c ≳ δOBL.17 The growth rate in the kinetic regime is reported in Equation (9). The plasma thermal energy grows exponentially, until the self-generated electric fields increase the transverse dispersion in beam velocity beyond the initial value Δv0, ⊥. At this point, the exponential growth at the rate in Equation (9) will necessarily terminate. From the Lorentz force applied to the beam particles, we find that this will happen when

Equation (12)

which results in a fraction epsilonE, ⊥  ∼  δk of the beam energy transferred to the electric fields at the end of the kinetic phase (δk ≡ ωke is defined in Equation (9)). Since the kinetic oblique mode is an electrostatic instability, the fraction of beam energy converted into heat will also be epsilone  ∼  δk. Moreover, due to the fact that ωk  ≲  ωOBL—they are comparable only if Δv0, ⊥/c  ∼  δOBL, i.e., at the boundary between reactive and kinetic regimes—we expect the kinetic oblique mode to be less efficient in heating the plasma electrons, as compared to the reactive phase. For beams with γb = 3–10 and α = 10−3–10−4, we have indeed verified with PIC simulations (not presented here) that the exponential growth of the kinetic oblique mode terminates at smaller epsilone for larger values of Δv0, ⊥, in good agreement with the expected scaling $\epsilon _e\propto \Delta v_{0,\perp }^{-2}$ (at fixed γb and α).18

Regardless of the character of the oblique mode (reactive or kinetic), the long-term evolution of the beam-plasma system is controlled by the quasi-linear relaxation. In Figure 5, we show how the relaxation phase is affected by a finite beam temperature Tb. For the set of beam parameters employed in Figure 5b = 300 and α = 10−2), the oblique phase is expected to occur in the reactive regime, as long as the beam comoving temperature is non-relativistic. In fact, for all the choices of Tb presented in Figure 5 (in the plot, Tb is in units of mec2/kB), the early increase in the heating efficiency epsilone proceeds at the reactive rate ωOBL (compare the curves in the inset of Figure 5(a) with the dotted red line, that scales with the oblique growth rate). Also, the kinetic oblique instability, which is responsible for the further growth in epsilone at ωet  ∼  1200 (see the inset in Figure 5(a)), does not show any dependence on temperature, in the range 10−6kBTb/mec2 ≲ 0.3 explored in Figure 5.

Figure 5.

Figure 5. Temporal evolution of a beam-plasma system with γb = 300 and α = 10−2, for different beam comoving temperatures at initialization (as shown by the legend in panel (b), where the beam temperature is in units of mec2/kB). Panel (a): fraction of the beam kinetic energy deposited into the background electrons, with the inset showing the evolution at early times. Panel (b): temporal evolution of the beam longitudinal momentum spread, in units of γbmec. Panel (c): beam (solid) and total (dashed) momentum spectra in the longitudinal direction, at the time indicated in panels (a) and (b) with the vertical black dotted line. In panel (c), the dotted oblique line shows the slope expected for a plateau distribution ${dN}/{dp}_\parallel \propto p_\parallel ^0$.

Standard image High-resolution image

The beam temperature has profound effects on the quasi-linear relaxation phase for the beam parameters employed in Figure 5. For cold beams (kBTb/mec2 ≲ 10−4, yellow and red lines in Figure 5), the relaxation phase does not depend on the beam temperature. In agreement with the results presented in Section 4.2.1, the longitudinal spread in the beam momentum increases during the relaxation stage until Δpb, ∥bmec  ∼  0.2, as shown in Figure 5(b). This corresponds to a fraction epsilone  ∼  10% of the beam kinetic energy being converted into plasma heating (Figure 5(a)). Similar conclusions hold for moderate beam temperatures (kBTb/mec2 = 3 × 10−2, green line), whereas the quasi-linear relaxation is suppressed if the beam temperature at birth is such that the initial beam spread Δpb0, ∥bmec ≳ 0.2 (cyan, blue, and black lines in Figure 5(b)). In this case, the quasi-linear phase does not mediate any further increase in the heating efficiency epsilone, beyond the early oblique phase (see the black line in Figure 5(a)).19 In short, if the initial momentum spread is Δpb0, ∥bmec ≳ 0.2, the plasma heating efficiency epsilone stays fixed at the value epsilone  ∼  δOBL attained at the end of the reactive oblique phase—or epsilone  ∼  δk, if the oblique instability proceeds in the kinetic regime.

The longitudinal momentum spectrum in Figure 5(c) clarifies why the quasi-linear relaxation is suppressed for hot beams. As we have discussed in Section 4.1.2, the relaxation is mediated by Langmuir waves excited by the beam. The beam particles that are slightly faster than the wave tend to transfer energy to the wave, and thus excite it, while those that are slightly slower than the wave tend to receive energy from it, and thus damp the wave. It follows that a mode is unstable if the number of beam particles moving slightly faster than the wave exceeds that of those moving slightly slower. More precisely, the instability will be stronger for a harder slope dlog N/dlog p of the longitudinal momentum spectrum at momenta ≲ γbmec (this is the relevant momentum scale, since the phase velocity of the most unstable mode is slightly smaller than the beam speed). For a sharply peaked beam (i.e., with kBTb/mec2 ≪ 1), the slope dlog N/dlog p will be extremely hard, and the excitation of the Langmuir modes that mediate the relaxation process will be most efficient. As a result of the quasi-linear relaxation, the beam particles will be scattered down to lower energies, and when the beam momentum spread exceeds Δpb, ∥bmec ≳ 0.2 the slope of the momentum spectrum below the peak becomes too shallow (see the red and yellow curves in Figure 5(c) at p/mec ≲ 300), and the quasi-linear relaxation terminates. If the beam temperature at birth is too hot (i.e., such that Δpb0, ∥bmec  ≳  0.2), the initial slope of the momentum spectrum at p ≲ γbmec is already too shallow to trigger efficient excitation of Langmuir waves, and the quasi-relaxation process is suppressed.

In Figure 5(c), we further support this argument by showing that in the latest stages of evolution, the shape of the beam momentum spectrum below γbmec (in Figure 5(c), at 30 ≲ p/mec ≲ 300), is nearly independent of the beam temperature at birth. For initially cold beams (yellow and red lines), the quasi-linear relaxation increases the beam momentum spread over time (see Figure 5(b)), until the beam spectrum relaxes to the broad bump shown in Figure 5(c). At this point, the beam is stable, although it has not relaxed to the so-called plateau distribution (as we explain in Section 4.1.2, the relaxation to the plateau distribution requires Δpb, ⊥/mec ≪ 1, which is not satisfied here). For initially hot beams, the initial momentum spectrum below γbmec resembles the final momentum distribution of initially cold beams. This explains why hot beams starting with a longitudinal momentum spread Δpb0, ∥bmec ≳ 0.2 will not experience the quasi-linear relaxation phase.

So far, we have discussed the effect of thermal spreads on the beam relaxation, assuming that the beam spectrum at birth is a drifting Maxwellian. However, the conclusions derived above hold for more complicated beam distributions. In particular, we have performed a set of PIC simulations, assuming that the beam spectrum at birth is a power law ${dN}/{dp}_\parallel \propto p_\parallel ^{-2}$ for ppminmec. For blazar environments, such a beam spectrum is expected as a result of intense IC cooling off the CMB.

If the beam power-law distribution has a negligible spread in the transverse direction (which is typically not the case for blazar-induced beams, see Section 4.3), then the oblique instability will proceed in the reactive regime. Apart from factors of the order of unity, we find that the exponential growth rate is ∼ (α mec/pmin)1/3, similar to the case of mono-energetic beams. As we have discussed above, the effectiveness of the quasi-linear relaxation—which ultimately determines whether the heating efficiency epsilone can reach ∼10%—is determined by the spread in the beam longitudinal spectrum below the peak, i.e., at ppmin. If the beam spectrum has a sharp low-energy cutoff at pmin, then the quasi-linear relaxation does operate, and the beam deposits ∼10% of its energy into the background electrons. However, if the low-energy end of the beam distribution is broader, the quasi-linear relaxation will be inhibited, and the amount of beam energy transferred to the plasma will be much smaller, in agreement with the results in Figure 5. We find that if the beam longitudinal spectrum below pmin can be modeled as a power law ${dN}/{dp}_\parallel \propto p_{\parallel }^s$, the quasi-linear relaxation is suppressed if s ≲ 3, with the case s = 0 corresponding to the plateau distribution.

4.3. Implications for Blazar-induced Beams

We now analyze the implications of our findings for the evolution of blazar-induced beams in the IGM. As we have discussed in Section 3, the Lorentz factors and density contrasts of blazar-driven beams (γb  ∼  106–107 and α  ∼  10−18–10−15) cannot be directly studied with PIC simulations. However, by performing a number of experiments with a broad range of γb ≫ 1 and α ≪ 1, we have been able to assess how the relaxation of ultra-relativistic dilute beams depends on the beam Lorentz factor and the beam-to-plasma density ratio. Our results can then be confidently extrapolated to the extreme parameters of blazar-induced beams.

We have demonstrated that the oblique instability, which governs the earliest stages of evolution of ultra-relativistic dilute beams, proceeds in the kinetic regime if the initial velocity dispersion in the direction transverse to the beam is Δv0, ⊥/c ≳ δOBL, where δOBL ≡ ωOBLe  ∼  (α/γb)1/3 is the growth rate of the reactive oblique mode in units of the plasma frequency $\omega _e=\sqrt{4\pi e^2 n_e/m_e}\,{\simeq}\,20\, (n_e/10^{-7}\ \rm cm^{-3})^{1/2}$ rad s−1 of the IGM electrons. Since the beam pairs are born with a mildly relativistic thermal spread in the center-of-momentum frame of the photon-photon interaction (see Section 2), the transverse velocity spread in the IGM frame will be Δv0, ⊥/c ≳ 1/γb. If follows that the oblique instability will proceed in the kinetic regime if γbδOBL ≲ 1, which is typically satisfied for blazar-induced beams (γbδOBL  ∼  0.01–1). The oblique instability will grow at the kinetic rate ωk  ∼  ωe(cv0, ⊥)2α/γb in Equation (9), which for Δv0, ⊥/c  ∼  1/γb, as appropriate for blazar-induced beams, reduces to ωk  ∼  γb α ωe. For the parameters of blazar-induced beams, this is a factor of

Equation (13)

larger than the IC cooling rate c/dIC, where dIC  ≃  100(γb/107)−1 kpc is the IC cooling length computed in Section 2. In agreement with Broderick et al. (2012) and Schlickeiser et al. (2012b), we find that the kinetic oblique instability has ample time to grow, before the beam loses energy to IC emission. Furthermore, the typical lifetime of blazar activity of ∼107 years ≫dIC/c is sufficient for the instability to operate.20

Due to self-heating of the beam in the transverse direction (see Sections 4.1 and 4.2.2), the exponential phase of the kinetic oblique instability terminates when only a minor fraction of the beam energy has been transferred to the IGM electrons. As we have argued in Section 4.2.2, the heating efficiency of blazar-induced beams at the end of the kinetic oblique phase is only

Equation (14)

It follows that even though the oblique instability can grow faster than the IC cooling time, its efficiency in heating the plasma electrons is extremely poor.

As we have emphasized in Section 4.1.2, a larger amount of beam energy (up to ∼10%) can be deposited into the plasma electrons by the quasi-linear relaxation phase. We find that the quasi-linear relaxation occurs on a timescale much longer than the growth time of the oblique instability, $\tau _{\rm R}\gtrsim 10^2 \omega _k^{-1}$. Yet, since the oblique growth rate is much faster than the IC cooling rate, see Equation (13), the quasi-linear relaxation should have enough time to operate before the beam energy is lost to IC emission. Here, we are conservatively neglecting the possibility that τR might be much longer than ${\sim }10^2\, \omega _k^{-1}$ for more extreme beam parameters, as a result of nonlinear plasma processes—most importantly nonlinear Landau damping by the thermal ions—that reduce the strength of the electric fields available for the quasi-linear relaxation phase (Schlickeiser et al. 2012b; Miniati & Elyiv 2013).21

In Section 4.2.2, we have demonstrated that the quasi-linear relaxation can occur only if the beam momentum spectrum along the longitudinal direction is sufficiently narrow. More precisely, we find that if the beam distribution peaks at γbmec (in the case of a power-law tail, this would be the low-energy cutoff), the quasi-linear relaxation can operate only if the parallel momentum spread below the peak satisfies Δpb0, ∥bmec ≲ 0.2. This constraint is hard to fulfill by blazar-induced beams. Since the pair-creation cross section peaks slightly above the threshold energy, the pairs are born moderately warm, with a comoving temperature kBTb/mec2  ∼  1. This corresponds to a longitudinal momentum spread at birth (measured in the IGM frame) of Δpb0, ∥bmec  ∼  1, which is already prohibitive for the development of the quasi-linear relaxation. Moreover, the momentum dispersion of blazar-induced beams will be even larger, since the spectrum of both the EBL and the blazar TeV emission are usually modeled as broad power laws (e.g., Miniati & Elyiv 2013 estimate that kBTb/mec2  ∼  5–10 at birth).

In summary, the shape of the beam momentum distribution below the peak is essential to predict the amount of beam energy deposited into the IGM. If the beam spectrum at birth were to have Δpb0, ∥bmec ≲ 0.2 (which is unlikely to be the case, but see Schlickeiser et al. 2012b), then the quasi-linear relaxation would transfer epsilone  ∼  10% of the beam energy to the IGM. Even in this optimistic case, we remark that the heating efficiency would reach at most ∼10%, so ∼90% of the energy would still remain in the beam. At the end of the quasi-linear relaxation phase, the beam spectrum will be harder than the so-called plateau distribution, since the transverse dispersion in beam velocity does not meet the requirement Δv0, ⊥/c ≪ 1/γb for relaxation to the plateau spectrum. In the more realistic case Δpb0, ∥bmec ≳ 0.2 (e.g., as resulting from the Monte Carlo model by Miniati & Elyiv 2013), the quasi-linear relaxation will be inhibited, resulting in a lower efficiency of IGM heating—with a firm lower limit being the heating fraction at the end of the oblique phase, see Equation (14).

4.3.1. Comparison with Earlier Studies

We now compare our numerical work with earlier analytical studies of the relaxation of blazar-induced beams in the IGM. As we have emphasized in Section 3, our PIC simulations are the first to address the evolution of dilute ultra-relativistic electron–positron beams, as appropriate for blazar-induced beams in the IGM. All of the previous PIC studies (e.g., Dieckmann et al. 2006b; Gremillet et al. 2007; Bret et al. 2008; Kong et al. 2009) have focused on mildly relativistic electron beams.

In agreement with Broderick et al. (2012) and Schlickeiser et al. (2012a), we find that the fastest growing instability for blazar-induced beams propagating through the IGM is the oblique mode. If the pair beam were to be initially cold, the oblique instability would evolve at the reactive rate in Equation (4), with wavevector oriented at ∼45° relative to the beam (see Schlickeiser et al. 2012a). However, the initial transverse spread in beam momentum is large enough such that the oblique mode evolves at the kinetic rate in Equation (9), as argued by Broderick et al. (2012) and Miniati & Elyiv (2013). We remark that the transverse beam spread does affect the growth of the oblique mode, but it will not impact the evolution of longitudinal waves, as found by Schlickeiser et al. (2013). However, since the early evolution of blazar-induced beams is controlled by the oblique (rather than longitudinal) mode, transverse thermal effects are indeed important for the beam-plasma interaction at early times.

The evolution of blazar-induced beams at late times is governed by nonlinear plasma processes, whose efficiency depends sensitively on the assumed beam distribution. Schlickeiser et al. (2012b) and Miniati & Elyiv (2013) attempted to describe the nonlinear relaxation of blazar-induced beams, reaching opposite conclusions regarding the ultimate fate of the beam energy. Schlickeiser et al. (2012a) assumed that the beam momentum distribution can be modeled as a delta function (i.e., they approximated the beam as mono-energetic and uni-directional), and they found that the relaxation phase occurs much faster than the IC cooling losses. From this, they argued that more than 50% of the beam energy can be transferred to the IGM plasma. Miniati & Elyiv (2013) reached the opposite conclusion, when accounting for the effect of the finite transverse momentum spread of blazar-induced beams, as derived from a detailed Monte Carlo model of the electromagnetic cascade. They found that the beam energy is radiated by IC off the CMB well before the relaxation phase.

With our PIC simulations, we find that the relaxation phase occurs on a much longer timescale than the exponential oblique growth, at least by two orders of magnitude. However, it might be delayed even more for more extreme beam parameters, as suggested by both Schlickeiser et al. (2012b) and Miniati & Elyiv (2013). Even under the conservative assumption that the relaxation phase is faster than the IC cooling time, this does not imply that all of the beam energy is ultimately deposited into the IGM plasma. In short, the relaxation process being faster than the IC losses does not guarantee that it will also be efficient in heating the IGM electrons. Under the unrealistic assumption that blazar-driven beams are born with a small longitudinal momentum spread Δpb0, ∥bmec ≲ 0.2, we find that only ∼10% (rather than ∼50%–100%, as assumed by Broderick et al. (2012)) of the beam energy is transferred to the plasma. This should be compared to the heating efficiency of ∼50% predicted by 1D models, that cannot properly account for the transverse broadening of the beam. For the realistic spread Δpb0, ∥bmec  ∼  1 of blazar-induced beams, the coupling between the beam and the plasma will be much less efficient, with the heating efficiency as low as epsilone  ≃  10−9b/107)(α/10−16).

5. SUMMARY AND DISCUSSION

The interaction of TeV photons from distant blazars with the EBL produces ultra-relativistic electron-positron pairs. The resulting pair beam is unstable to the excitation of beam-plasma instabilities in the unmagnetized IGM. The ultimate fate of the beam energy is uncertain, and it is hard to capture with analytical tools. By means of 2D and 3D PIC simulations, we have investigated the linear and nonlinear evolution of ultra-relativistic dilute electron-positron beams. We have performed dedicated experiments with a broad range of beam Lorentz factors γb = 3–1000 and beam-to-plasma density contrasts α = 10−3–10−1, so that our results can be extrapolated to the extreme parameters of blazar-induced beams (γb = 106–107 and α = 10−18–10−15).

We find that the earliest stages of evolution of ultra-relativistic dilute beams are governed by the oblique instability. For cold beams, the oblique mode proceeds in the so-called reactive regime, where all the beam particles are interacting with the unstable waves, and the coupling between the beam and the plasma is most efficient. In this case, the instability grows at the reactive rate ωOBL = δOBLωe, where δOBL  ∼  (α/γb)1/3 and $\omega _e=\sqrt{4\pi e^2 n_e/m_e}\,{\simeq}\,20\,(n_e/10^{-7}{\mathrm{\;cm^{-3}}})^{1/2} {\mathrm{\;rad\, s^{-1}}}$ is the plasma frequency of the IGM electrons. However, blazar-induced beams are not cold. The initial spread in transverse velocity is Δv0, ⊥/c  ∼  1/γb ≳ δOBL, so that the oblique mode proceeds in the kinetic (rather than reactive) regime. Here, the spectrum of unstable modes is broader, with only a few beam particles being in resonance with each mode. The growth rate is ωk = δk ωe  ∼  ωe(cv0, ⊥)2α/γb, which is smaller than the corresponding rate ωOBL of cold beams, yet large enough such that the kinetic oblique instability has ample time to grow, before the beam loses energy to IC emission by scattering off the CMB.

On the other hand, the oblique instability growing faster than the IC losses does not guarantee that it will also be efficient in heating the IGM electrons. Due to self-heating of the beam in the transverse direction, we find that the exponential phase of the kinetic oblique instability terminates when only a minor fraction of the beam energy has been transferred to the IGM plasma. At the end of the kinetic oblique phase, the heating efficiency of IGM electrons is extremely poor, with epsilone  ∼  δk  ≃  10−9b/107)(α/10−16).

Additional transfer of energy from the beam to the plasma occurs at later times (with a delay of two or more orders of magnitude, relative to the oblique growth time), and it is mediated by the quasi-linear relaxation process. Here, the beam generates longitudinal electrostatic waves, which scatter the beam particles—thus broadening the beam momentum spectrum in the longitudinal direction—and heat the IGM electrons. The quasi-linear relaxation can operate only if the beam momentum spectrum along the longitudinal direction is sufficiently narrow. More precisely, we find that if the beam distribution peaks at γbmec, the quasi-linear relaxation requires the parallel momentum spread below the peak to be Δpb0, ∥bmec  ≲  0.2. In this case, a fraction epsilone  ∼  10% of the beam energy is transferred to the plasma (rather than ∼50%–100%, as assumed by Broderick et al. 2012).

The constraint Δpb0, ∥bmec  ≲  0.2 on the longitudinal dispersion at birth is hard to fulfill by blazar-induced beams. Since the pair-creation cross section peaks slightly above the threshold energy, the pairs are born moderately warm, with a comoving temperature kBTb/mec2  ∼  1. This corresponds to a longitudinal momentum spread at birth (measured in the IGM frame) of Δpb0, ∥bmec  ∼  1, which is already prohibitive for the development of the quasi-linear relaxation. Moreover, the momentum dispersion of blazar-induced beams might be even larger, since the spectrum of both the EBL and the blazar TeV emission are usually modeled as broad power laws (e.g., Miniati & Elyiv 2013 estimate that kBTb/mec2  ∼  5–10 at birth). For Δpb0, ∥bmec  ∼  1, the quasi-linear relaxation will be suppressed, which results in a much lower efficiency of IGM heating—with a firm lower limit being the heating fraction at the end of the oblique phase epsilone  ∼  δk ≪ 1.

At the end of the relaxation phase, the beam and plasma distributions are highly anisotropic, with the longitudinal momentum spread much larger than the transverse one. As a result, the system is prone to the Weibel instability (e.g., Weibel 1959; Yoon & Davidson 1987; Sakai et al. 2000; Silva et al. 2002, 2003; Jaroschek et al. 2005), which relaxes the beam and plasma anisotropy by generating transverse magnetic fields. The Weibel instability is predominantly magnetic, so it does not mediate any further exchange of energy from the beam to the plasma electrons. However, it might be a promising source for the generation of magnetic fields in the IGM, as discussed by Schlickeiser et al. (2012b). A multi-dimensional PIC investigation of the strength and scale of the magnetic fields resulting from blazar-induced beams will be presented elsewhere.

Our results have important implications for the ultimate fate of the energy of blazar-induced beams. As we have argued above, most of the beam energy is still available for IC interactions with the CMB. Therefore, the Fermi non-detection of the IC scattered GeV emission around TeV blazars can be reliably used to probe the strength of the EBL and of the IGM magnetic fields. In particular, the lower bounds on the IGM field strength derived by various authors (e.g., Neronov & Vovk 2010; Tavecchio et al. 2010) are still valid, despite the fast growth of beam-plasma instabilities in the IGM. However, when computing the expected IC signature, one should take into account that beam-plasma instabilities tend to broaden the beam distribution function. As a result, the reprocessed GeV emission will be spread over a wider frequency range (and consequently, with smaller flux), as compared to the case of mono-energetic beams.

The fraction of beam energy deposited into the IGM might have important cosmological implications, as argued by Chang et al. (2012) and Pfrommer et al. (2012). By assuming that all of the beam energy is transferred to the IGM electrons by beam-plasma instabilities, they showed that blazar heating could dominate over photo-heating in the low-redshift evolution of the IGM, by almost one order of magnitude (yet, this does not seem to be supported by the observations, see Rudie et al. 2012). If the heating efficiency of blazar-induced beams were ∼10%, rather than ∼100% as they assumed, blazar heating could still contribute as much as photo-heating to the thermal history of the IGM. However, we expect the heating efficiency of blazar-induced beams to be epsilone ≪ 1, for the following reasons.

  • 1.  
    As we have argued above, blazar-induced beams are born with a significant longitudinal spread in momentum, Δpb0, ∥bmec  ∼  1. In this case, the quasi-linear relaxation process, which mediates efficient transfer of the beam energy to the IGM electrons up to epsilone  ∼  10%, will be suppressed, and the heating fraction remains epsilone  ∼  δk ≪ 1.
  • 2.  
    Even if the relaxation process were to be operating, the beam relaxation timescale might be much longer than the IC loss time, due to the competing effect of nonlinear Landau damping by the thermal protons, as argued by Miniati & Elyiv (2013; but see Schlickeiser et al. 2012b) for opposite conclusions assuming a different shape of the beam distribution function).
  • 3.  
    Density inhomogeneities associated with cosmic structure induce loss of resonance between the beam particles and the excited plasma oscillations, strongly inhibiting the growth of the unstable modes (Miniati & Elyiv 2013).
  • 4.  
    A large-scale IGM field might spread a mono-energetic uni-directional beam in the transverse and longitudinal directions, thus inhibiting the efficiency of the beam-plasma interaction.22 In the oblique phase, a sufficiently strong magnetic field might trigger the transition to the kinetic regime, if during the growth time of the reactive instability ($\sim \omega _{\rm OBL}^{-1}$) it deflects the beam velocity sideways by more than the threshold ∼δOBLc between the reactive and kinetic regimes. This happens if $\omega _{\rm B}/\omega _e\gtrsim \delta _{\rm OBL}^2$, where ωB = eBbmec is the Larmor frequency of the beam particles in the IGM fields. The limit on the IGM field strength is
    Equation (15)
    which is generally realized for IGM fields (e.g., Neronov & Semikoz 2009).As regards to the quasi-linear relaxation, IGM fields are expected to change the evolution of the quasi-linear process if during the characteristic quasi-linear growth time τR the field deflects the beam such that the longitudinal dispersion in momentum reaches Δpb, ∥bmec ≳ 0.2. This can be recast as ωBτR ≳ 1. A weaker constraint can be derived by using that the quasi-linear relaxation plays a role only if the IC cooling time is dIC/c ≳ τR. By setting ωBdIC/c ≳ 1, we obtain a limit on the field strength that inhibits the quasi-linear relaxation
    Equation (16)
    which might be satisfied by IGM fields (e.g., Neronov & Semikoz 2009).For these reasons, we conclude that blazar-induced beams are not likely to play a major role in the thermal history of the IGM.

We thank A. Bret, M. Dieckmann, and the anonymous referee for insightful comments. L.S. is supported by NASA through Einstein Postdoctoral Fellowship grant number PF1-120090 awarded by the Chandra X-ray Center, which is operated by the Smithsonian Astrophysical Observatory for NASA under contract NAS8-03060. The simulations were performed on XSEDE resources under contract No. TG-AST120010, and on NASA High-End Computing (HEC) resources through the NASA Advanced Supercomputing (NAS) Division at Ames Research Center.

APPENDIX A: 1D SIMULATIONS OF RELATIVISTIC DILUTE BEAMS

In the main body of the paper, we have presented our results on the relaxation of ultra-relativistic dilute beams, by employing 2D and 3D simulations. Here, we discuss how the physics of the beam-plasma evolution differs, when performing 1D simulations.

In Figure 6, we compare our 2D results (red line) with two selected 1D simulations (black and green curves), that differ in the orientation of the beam relative to the simulation box. For the black line, the beam is aligned with the simulation domain, whereas the two directions form an angle of θbox = 45° for the green line. Since the fastest growing oblique mode for ultra-relativistic dilute cold beams is oriented at ∼45° relative to the beam propagation (see Section 4.1.1), the 1D box at an angle θbox = 45° with respect to the beam should correctly capture the evolution of the oblique mode. This is confirmed by the inset in Figure 6, which shows that the exponential growth in the heating efficiency epsilone proceeds at the expected rate ωOBL of Equation (4) both in 2D (red line) and in the 1D simulation with θbox = 45° (green).

Figure 6.

Figure 6. Comparison between our 2D simulation (red line) with two 1D simulations (black and green lines), such that the computational box is oriented at two different angles relative to the beam velocity. The beam has γb = 100 and α = 3 × 10−3. For the black line, the 1D computational box is oriented along the beam, i.e., the simulation only selects the unstable modes whose wavevector is parallel to the beam. For the green line, the 1D domain forms an angle θbox = 45° with the beam velocity, so that only the oblique modes can be captured by the 1D simulation. The inset shows the evolution of the heating efficiency epsilone at early times.

Standard image High-resolution image

On the other hand, the 1D simulation with θbox = 45° cannot correctly capture the relaxation phase, which is mediated by quasi-longitudinal modes. In fact, the heating efficiency epsilone in the 1D box with θbox = 45° does not significantly change after the end of the oblique phase. In contrast, as a result of the quasi-longitudinal relaxation, in 2D the heating fraction epsilone increases at 104 ≲ ωet ≲ 5 × 104 up to the saturation value epsilone  ∼  10%.

The quasi-linear relaxation, being driven by longitudinal modes, can be described in 1D with a simulation box oriented along the beam (black line in Figure 6). For a 1D box with θbox = 0° (black line), the quasi-linear relaxation controls the beam evolution at 2 × 104 ≲ ωet ≲ 3 × 104. At late times, the heating efficiency saturates at epsilone  ∼  20% (in agreement with Thode & Sudan 1975), which is twice as large as compared to the analogous 2D case. As anticipated in Section 4.1.2, this is related to the role of transverse spreads in the relaxation of ultra-relativistic beams. In multi-dimensions, the longitudinal velocity spread Δvb, ∥ required to terminate the relaxation phase can be achieved either by decelerating the beam in the longitudinal direction (with a fractional energy loss Δpb, ∥bmec), or by deflecting the beam sideways (which gives a transverse spread Δpb, ⊥, but no significant energy loss). The contribution of the two terms is presented in Equation (11). A 1D box with θbox = 0° can only capture beam-aligned modes, which cannot change the transverse spread Δpb, ⊥, so that the second term in Equation (11) does not contribute. It follows that the same Δvb, ∥ will be attained in 1D by a larger Δpb, ∥bmec (and so, higher epsilone), relative to its 2D counterpart. This explains why in 1D (for beam-aligned boxes) the heating fraction epsilone is a factor of a few larger than in 2D (compare black and red lines in Figure 6).

In summary, beam-aligned 1D simulations tend to overestimate the fraction of beam energy deposited into the background electrons by the relaxation phase. Most importantly, they cannot properly model the early evolution of the beam-plasma system, which is mediated by oblique modes. Rather, the exponential phase in 1D simulations with θbox = 0° will necessarily proceed at the two-stream growth rate $\omega _{\rm TS}=\gamma _b^{2/3}\omega _{\rm OBL}$, which is indicated as a dotted black line in Figure 6. In short, the multi-dimensional physics of the beam-plasma evolution cannot be properly captured by 1D simulations.

The difference between 1D and 2D simulations is also presented in Figure 7, where we discuss the dependence of our results on the transverse size of the computational domain, from L = 0.125 ce (1D simulation) to our standard choice L = 125 ce. We also confirm that our 2D results are the same when doubling the box size in the transverse direction (compare the red lines for L = 125 ce with the yellow lines for L = 250 ce).

Figure 7.

Figure 7. Temporal evolution of the beam-plasma interaction, as a function of the box size L in the direction transverse to the beam (instead, L = 125 ce along the beam in all cases). The beam has γb = 100 and α = 10−2. Panel (a): temporal evolution of the heating efficiency epsilone, with the inset showing the evolution at early times. Panel (b): temporal evolution of the beam longitudinal momentum spread, in units of γbmec. Panel (c): the solid lines show the beam momentum spectrum in the longitudinal direction, at the time indicated in panels (a) and (b) with the vertical black dotted line. For three selected cases (L = 0.125 ce, 125 ce, and 250 ce), we also plot the total (beam plus plasma) momentum spectra with dashed lines. In panel (c), the dotted oblique line shows the slope expected for a plateau distribution ${dN}/{dp}_\parallel \propto p_\parallel ^0$.

Standard image High-resolution image

For 1D boxes aligned with the beam, the two-stream instability governs the exponential growth of epsilone at early times (compare the black solid and dotted lines in Figures 7(a) and (b)). The oblique mode can operate only if the transverse size of the box is L ≳ 2.5 ce, as shown by the fact that the green line in the inset of Figure 7(a) grows at the oblique rate ωOBL indicated by the dotted red line. For a box with L = 0.625 ce (blue line), the oblique phase mediates the growth of epsilone at early times (ωet ≲ 3000), yet at a rate smaller than ωOBL, whereas the exponential stage of the two-stream mode emerges at later times (ωet  ∼  7000) with the expected rate ωTS (see the blue line in Figures 7(a) and (b)).

In agreement with Figure 6, we find that the quasi-linear relaxation in 2D proceeds in a similar way as in 1D, apart from the fact that the dispersion in longitudinal momentum at late times is smaller for larger box widths, as shown in Figure 7(b). In turn, this is related to the shape of the beam momentum distribution at the end of the quasi-linear relaxation phase. As shown in Figure 7(c), the longitudinal momentum spectrum in 1D simulations relaxes to the plateau distribution ${dN}/{dp}_\parallel \propto p_\parallel ^0$ (indicated as a dotted black line in Figure 7(c)), whereas in 2D the beam spectrum below the peak stays harder than the plateau distribution. As we have argued in Section 4.1.2 (see Appendix B for further details), the difference between our 1D and 2D results is ultimately related to the transverse spread in beam momentum, which is larger in 2D than in 1D.

APPENDIX B: RELAXATION TO THE PLATEAU DISTRIBUTION

In Section 4.1.2, we have argued, based on Equation (11), that the longitudinal velocity spread Δvb, ∥ required to terminate the relaxation process can be sourced not only by a longitudinal spread in momentum Δpb, ∥, but also by a transverse spread Δpb, ⊥. In 1D, the latter is absent, whereas in 2D it is the interplay of the two that determines the shape of the beam momentum spectrum at late times. In particular, this explains why in 2D the beam does not necessarily relax to the plateau distribution, which instead is a general outcome of the beam-plasma evolution in 1D configurations.

Figure 8 clarifies the role of the transverse momentum spread for the relaxation to the plateau distribution. Using 2D simulations, we perform the following experiment. Right after the end of the oblique phase, we artificially reset the transverse momentum spread Δpb, ⊥ to the values indicated in the legend. Also, in the course of the subsequent evolution, we inhibit its evolution by hand. We find that if we artificially set Δpb, ⊥/mec ≪ 1, the beam relaxation proceeds as in 1D, and the longitudinal momentum spectrum at late times relaxes to the plateau distribution (compare the black solid and dotted lines in Figure 8). In contrast, if Δpb, ⊥/mec ≳ 1 the quasi-linear relaxation does spread the beam momentum in the longitudinal direction, but not enough to approach the plateau distribution. We have verified that the threshold Δpb, ⊥/mec  ∼  1 for relaxation to the plateau distribution holds irrespective of the beam Lorentz factor or the beam-to-plasma density contrast.

Figure 8.

Figure 8. Longitudinal momentum spectrum at late times (ωet = 1.5 × 105), for a beam-plasma system with γb = 1000 and α = 10−2. We perform the following experiment: after the end of the oblique phase, we artificially reset the transverse spread in beam momentum to the values indicated in the legend (there, Δpb, ⊥ is in units of mec), and we inhibit by hand its evolution. For the case indicated with the yellow line, we allow the beam to evolve without constraints (i.e., this would correspond to the self-consistent evolution of the beam-plasma system), which results in Δpb, ⊥/mec  ∼  20 at late times. It is apparent that relaxation to the plateau distribution (indicated as a dotted black line) requires Δpb, ⊥/mec ≪ 1.

Standard image High-resolution image

As a result of the growth of the oblique mode, the transverse dispersion of cold beams at the end of the oblique phase approaches Δpb, ⊥/mec  ∼  γbδOBL. For the beam parameters employed in Figure 8b = 1000 and α = 10−2), this would give Δpb, ⊥/mec  ∼  20. This explains why the beam spectrum plotted as a yellow solid curve in Figure 8, which corresponds to the self-consistent evolution of the beam-plasma system (i.e., the beam transverse dispersion is not constrained by hand), does not approach the plateau distribution at late times.

Footnotes

  • Throughout the paper, we neglect the dependence on cosmological redshift. Strictly speaking, our results apply to z  ∼  0, but they can be easily generalized to arbitrary redshifts, provided that one makes additional assumptions about the redshift evolution of the EBL and of the blazar luminosity function.

  • If plasma instabilities were to dominate the energy loss of the beam, dIC should be replaced by the beam thermalization length.

  • The comoving beam temperature Tb corresponds to a beam transverse velocity dispersion Δv0, ⊥ (in the IGM frame) such that $\Delta v_{0,\perp }/c\,{\sim}\,\sqrt{k_{\rm B} T_b /\gamma _b^2 m_e c^2}$ if kBTb/mec2 ≪ 1, whereas Δv0, ⊥/c  ∼  kBTbbmec2 if kBTb/mec2 ≫ 1.

  • Nonlinear Landau damping is the interaction between thermal protons in the IGM and the beat of two Langmuir waves. This would scatter the waves to higher phase velocities, where they cannot interact efficiently with the beam. As discussed by Miniati & Elyiv (2013), nonlinear Landau damping can strongly inhibit the transfer of the beam energy to the IGM plasma.

  • Beams with more extreme parameters (in particular, with α ≲ 10−3) will take longer to evolve, and at that point the fact that explicit PIC codes do not conserve energy to machine precision (see below) might impact the reliability of our results.

  • The speed of light in the simulations is 0.45 cells/timestep, so that the temporal resolution is $\delta t=0.05625\,\omega _e^{-1}$.

  • 10 

    Hereafter, "longitudinal" and "transverse" will be relative to the beam direction of motion.

  • 11 

    We remind that for 2D runs with the beam lying in the simulation plane, only the in-plane components of the electric field, and the out-of-plane component of the magnetic field are present.

  • 12 

    In Equation (3), the factor of two that multiplies α is related to our definition of α = nb/nIGM = nb/2 ne.

  • 13 

    From Equation (3), one would expect that in the ultra-relativistic limit γb ≫ 1, the fastest growing mode should have kk, rather than k  ∼  k as we find. In reality, modes with kk are efficiently suppressed due to self-heating of the beam in the transverse direction, as we show in Equation (6). It follows that the fastest growing mode lies at ∼45° of the beam direction.

  • 14 

    This is smaller than the heating efficiency of ∼30% reported by Thode & Sudan (1975) using 1D simulations. In Appendix A, we demonstrate that the transfer of beam energy to plasma electrons is indeed less efficient in 2D, as compared to the 1D case studied by Thode & Sudan (1975).

  • 15 

    We have extensively checked that this result is numerically solid. We have confirmed our conclusions by using a larger number of computational particles per cell (up to 256), a larger 2D box (up to four times as large, in each direction), and a finer spatial resolution (up to  ce = 16, instead of the usual value  ce = 8).

  • 16 

    While the quasi-linear relaxation can also be captured with 1D simulations (although in 1D one would unphysically neglect the effect of the beam transverse spread), the growth of the Weibel modes necessarily requires multi-dimensional simulations.

  • 17 

    The velocity dispersion Δv0, ⊥ can be recast as a comoving beam temperature kBTb/mec2  ∼  (γbΔv0, ⊥/c)2, where we have assumed kBTb/mec2 ≪ 1 (i.e., non-relativistic temperatures).

  • 18 

    The condition $\Delta v_{0,\perp }/c\,{\sim}\,\sqrt{k_{\rm B} T_b /\gamma _b^2 m_e c^2}\,{\gtrsim}\, \delta _{\rm OBL}$, as required for the kinetic regime, together with the assumption of a quasi-monoenergetic beam (i.e., with comoving beam temperature kBTb/mec2 ≲ 1), constrains $\gamma _b\delta _{\rm OBL}\,{\sim}\,(\gamma _b^2\alpha)^{1/3}\lesssim 1$, i.e., the kinetic regime can be best probed by low-γb beams with α ≪ 1.

  • 19 

    Since the relaxation phase is quasi-longitudinal, the same conclusions hold in 1D. We have confirmed that regardless of the nature of the fastest growing mode (oblique in 2D, longitudinal in 1D), the quasi-linear relaxation is suppressed if the beam spread at birth is such that Δpb0, ∥bmec ≳ 0.2, both in 1D and in 2D.

  • 20 

    However, jet variability may be fast enough to compete with the instability growth time. Our analysis focuses on the steady or long-term average TeV emission.

  • 21 

    As we have anticipated in Section 3, nonlinear Landau damping by the thermal ions does not seem to play a major role in the evolution of the beam-plasma system, for the parameters explored in our study. Most likely, this stems from the fact that the beam energy density in our simulations is much greater than the initial thermal energy density of the background plasma, as compared to blazar-induced beams.

  • 22 

    For simplicity, we neglect the fact that the presence of a large-scale field might change the nature of the fastest growing instabilities, as compared to the unmagnetized case explored in this work.

Please wait… references are loading.
10.1088/0004-637X/787/1/49