A publishing partnership

Articles

STOCHASTIC ACCELERATION AND THE EVOLUTION OF SPECTRAL DISTRIBUTIONS IN SYNCHRO-SELF-COMPTON SOURCES: A SELF-CONSISTENT MODELING OF BLAZARS' FLARES

, , and

Published 2011 September 7 © 2011. The American Astronomical Society. All rights reserved.
, , Citation A. Tramacere et al 2011 ApJ 739 66 DOI 10.1088/0004-637X/739/2/66

0004-637X/739/2/66

ABSTRACT

The broadband spectral distributions of non-thermal sources, such as those of several known blazars, are well described by a log-parabolic fit. The second-degree term in these fits measures the curvature in the spectrum. In this paper, we investigate whether the curvature parameter observed in the spectra of the synchrotron emission can be used as a fingerprint of stochastic acceleration. As a first approach, we use the multiplicative central limit theorem to show how fluctuations in the energy gain result in the broadening of the spectral shape, introducing a curvature into the energy distribution. Then, by means of a Monte Carlo description, we investigate how the curvature produced in the electron distribution is linked to the diffusion in momentum space. To get a more generic description of the problem we turn to the diffusion equation in momentum space. We first study some "standard" scenarios, in order to understand the conditions that make the curvature in the spectra significant, and the relevance of cooling during the acceleration process. We try to quantify the correlation between the curvature and the diffusive process in the pre-equilibrium stage, and investigate how the transition between the Klein–Nishina and the Thomson regimes, in inverse Compton cooling, determine the curvature in the distribution at equilibrium. We apply these results to some observed trends, such as the anticorrelation between the peak energy and the curvature term observed in the spectra of Mrk 421, and a sample of BL Lac objects whose synchrotron emission peaks at X-ray energies.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

A defining feature of the non-thermal emission from different types of galactic and extragalactic sources is that their spectra are described by a power law (PL) over a broad photon energy range. In several sources, however, their spectra show significant curvature that is typically milder than that expected from an exponential cutoff. In previous papers, Massaro et al. (2004, 2006) discussed the curvature observed in the broadband X-ray spectra of the two well-known HBL (High-energy peaked BL Lac) objects Mrk 421 and Mrk 501. The basic idea was that this curvature was not simply the result of radiative cooling of high energy electrons, responsible for the synchrotron and inverse Compton (IC) emission, but that it was essentially related to the acceleration mechanism. Massaro et al. (2004) showed that curved spectral distributions, in particular log-parabolic (i.e., log-normal) ones, develop when the acceleration probability is a decreasing function of the electron energy. In subsequent works, through the analysis of a large collection of X-ray observations of Mrk 421, Tramacere et al. (2007, 2009) and Tramacere (2007) pointed out that the observed anticorrelation between the peak energy and the curvature measured in the synchrotron spectral energy distribution (SED) could be used as a clear signature of a stochastic component in the acceleration process. Very recently, the log-parabolic law has also been applied to describe the spectral distribution and evolution of some gamma-ray bursts (Massaro & Grindlay 2011).

The principal aim of the present paper is to investigate this scenario by extracting information on the acceleration processes using the curvature parameter measured in the observed synchrotron and IC spectra of Synchro-Self-Compton (SSC) sources. We study the conditions in which the energy distributions of electrons, resulting from stochastic acceleration, can be approximated by a log-parabolic law and how its curvature evolves during their acceleration, and the role of IC cooling. We compare predictions from our theoretical descriptions with the curved spectra of some HBL objects.

In Section 2, we give an intuitive picture, taking into account the effect of random fluctuations on the energy gain of particles and the role these play in determining the spectral curvature, as a consequence of the multiplicative central limit theorem, and compare these results with the analytical solution of the diffusion equation, in the "hard-spheres" approximation. In Sections 3 and 4, we give a more physical description of the problem, using first a Monte Carlo (MC) approach and second by solving numerically the momentum-diffusion equation. We discuss the evolution of the curvature in the electron distribution as a result of momentum diffusion before equilibrium is reached and the role that synchrotron and IC cooling processes play in reaching the equilibrium. In Section 5, we study the peak energy, fluxes, and curvature trends in the SED of both the synchrotron and IC emission, looking for the fingerprints of the stochastic component. In Section 6, we show how our results can reproduce the spectral trends observed in some HBLs, in particular we investigate the relation between the peak energy and the curvature, and between the peak energy and the peak flux. The good agreement between predictions and observed trends confirms that the stochastic acceleration mechanism can play an important role in the physics of blazars' jets and other SSC sources.

2. THE LOG-PARABOLA ORIGIN: ANALYTICAL APPROACH

2.1. Statistical Description

In the statistical picture, the change in energy of the particles at each acceleration step ns is expressed as

Equation (1)

where γ is the Lorentz factor of the particle and ε is the fractional energy gain. Here we investigate the role of fluctuations of ε on the spectral shape of the accelerated particles. With this aim in mind, we express the energy gain fluctuations as

Equation (2)

where the random variable χ has a probability density function with zero mean value (〈χ〉 = 0) and variance σ2χ, and $\bar{\varepsilon }$ represents the systematic energy gain, which we treat as a non-random variable and the probability density function of ε is defined on the range ε ⩾ 0. The particle energy at step ns can be expressed as

Equation (3)

where γ0 is the initial energy of the particle. This equation clearly shows that the final energy distribution (n(γ)  =  dN(γ)/dγ) will result from the product of the random variables εi. The determination of an analytic expression for the distribution resulting from the multiplication of generic random variable is not an easy task (Glen et al. 2004). Using the simplifying assumption that the particles are always accelerated, namely, the acceleration probability, Pa, is set to unity and applying the multiplicative case of the central limit theorem (e.g., Cowan 1998), it is possible to show that the particle energies will be distributed as a log-normal law

Equation (4)

where N0 is the total number of particles, μ = 〈lnγ〉, σ2γ = σ2(lnγ). We can determine these two quantities by taking the logarithm of Equation (3),

Equation (5)

assuming that $\chi _i/\bar{\varepsilon }$ is not large. We obtain for the two parameters μ and σγ

Equation (6)

where we have ignored the covariance terms since we are assuming the energy gain at each acceleration step is independent of the one at the previous step. Remembering that 〈χ〉 = 0, σχ = σε, and ignoring the fourth-order term, we can write

Equation (7)

This equation shows that the variance increases linearly with the number of acceleration steps and is proportional to σε2. Substituting μ and σγ into Equation (4)

Equation (8)

Hereafter we will consider decimal logarithms (log ≡ log10, ce = 1/log10e ≈ 2.3) to make easier a comparison of the curvature results from this paper with those presented in observational papers. Taking the logarithm of Equation (8) and substituting the parameters from Equation (8) we obtain

Equation (9)

where K includes all the constant factors. This is a log-parabolic law with the curvature (second degree in log γ) coefficient given by

Equation (10)

The interesting physical insight of this equation is that the curvature of the particle energy distribution is inversely proportional to the acceleration steps (ns) and to the variance of the energy gain (σ2ε). In the case of Pa < 1, the distribution at step ns will be given by the convolution of different log-normal distributions for each acceleration step, with the distribution at ns broader than that at ns − 1 and containing fewer particles, as already noted in Peacock (1981).

Similar results are obtained considering a constant energy gain but a fluctuating number of acceleration steps. Assuming that after a time t the probability distribution for the number of steps undergone by a particle is given by a Poisson law, it is possible to show that the energy distribution follows a log-parabola whose curvature term depends on the inverse of the mean number of steps multiplied by the duration of the acceleration process.

2.2. Diffusion Equation Approach

The above statistical description provides an intuitive link between the curvature in the energy distribution of accelerated particles and the presence of a randomization process, such as the dispersion in the energy gain or in the number of acceleration steps. However, this approach does not give a complete physical description of the processes responsible for the systematic and stochastic energy gain, ignoring other physical processes, such as the radiative cooling and injection rates, or the acceleration energy dependence, necessary to give a complete description of the particles energy distribution evolution. A physical self-consistent description of stochastic acceleration in a time-dependent fashion can be achieved through a kinetic equation approach. Employing the quasi-linear approximation with the inclusion of momentum-diffusion term (Ramaty 1979; Becker et al. 2006), the equation governing the temporal evolution of n(γ) is

Equation (11)

where Dp(γ, t) is the momentum-diffusion coefficient, DA(γ, t) = (2/γ)Dp(γ, t) is the average energy change term resulting from the momentum-diffusion process, and S(γ, t) = −C(γ, t) + A(γ, t) is an extra term describing systematic energy loss (C) and/or gain (A), and Q(γ, t) is the injection term. In the standard diffusive shock acceleration scenario, there are several possibilities for which one can expect that energy gain fluctuations will occur, due to the momentum-diffusion term. In particular, for the case of a turbulent magnetized medium, the advection of particles toward the shock due to pitch angle scattering may be accompanied by stochastic momentum-diffusion mechanism. In this scenario, particles embedded in a magnetic field with both an ordered (B0) and turbulent (δB) component, exchange energy with resonant plasma waves, and the related diffusion coefficient is determined by the spectrum of the plasma waves. Following the approach of Becker et al. (2006), we describe the energy distribution W(k) in terms of the wave number k = 2π/λ with a PL

Equation (12)

with q = 2 for the "hard-sphere" spectrum, q = 5/3 for the Kolmogorov spectrum, and q = 3/2 for the Kraichnan spectrum, the total energy density in the fluctuations being

Equation (13)

Under these assumptions, the momentum-diffusion coefficient reads (O'Sullivan et al. 2009)

Equation (14)

where βA = VA/c and VA is the Alfvén waves velocity, ρg = pc/qB is the Larmor radius, and λmax is the maximum wavelength of the Alfvén waves spectrum. The acceleration time for particles with Lorentz factor γ, whose Larmor radii resonate with one particular magnetic field turbulence length scale, is dictated by the momentum-diffusion coefficient (Dp) as

Equation (15)

The spatial diffusion coefficient relates to the momentum-diffusion coefficient through the relation, DxDpp2β2A (Skilling 1975), hence the escape time of the particles from the acceleration region of size R depends on the spatial diffusion coefficient through the relation

Equation (16)

The coefficients in Equation (11), and their related timescales, can be expressed as a PL in terms of the Lorentz factor (γ)

Equation (17)

where Dp0 and A0 have the dimension of the inverse of a time. Analytical solutions of the diffusion equation for relativistic electrons have frequently been discussed in the literature since the early work by Kardashev (1962), in particular for the case of the "hard-sphere" approximation. Neglecting the S and Tesc terms in Equation (11), and using a mono-energetic and instantaneous injection (n(γ, 0) = N0δ(γ − γ0)), the solution of the diffusion equation is (Melrose 1969; Kardashev 1962)

Equation (18)

i.e., a log-parabolic distribution, whose curvature term is

Equation (19)

This result is fully consistent with that found in the statistical description; indeed, Equations (18) and (8) have the same functional form in both the statistical and in the diffusion equation scenario, with t playing the role of ns, Dp0 the role of the variance of the energy gain (σ2ε), and Ap0 the role of $\log \bar{\varepsilon }$. Hence we can write

Equation (20)

It is interesting to note that in the case of the "hard-sphere" approximation, the curvature term is simply dictated by the ratio of the diffusive acceleration time (tD) to the evolution time (t).

3. NUMERICAL APPROACH: MONTE CARLO SIMULATION WITH MAGNETIC TURBULENCE

In this section, we demonstrate explicitly how the introduction of energy fluctuations leads to curved spectral distributions of particles. This is carried out using an MC approach.

In our simulations, we considered 105 particles injected into the system with a cold mono-energetic distribution of Lorentz factors, with γ0 = 1. To compare these results with the ones presented in Section 2, we remind the reader that in the MC approach, the duration of the acceleration process t is the equivalent of the number of acceleration steps (ns) used in the statistical picture and that the probability of the particle to be upscattered or downscattered in the MC realizations can be expressed in the statistical approach as P(ε > 1) and P(ε < 1), respectively. The scattering probability of the particles is dictated by the intensity of resonant waves in the turbulent magnetic power spectrum. As a working hypothesis, we assume that particles interact with a turbulent magnetic field whose power spectrum is expressed by Equation (12). In each scattering, the particles have a probability of (1 + βA)/2 of being upscattered and a probability of (1 − βA)/2 of being downscattered. The energy dispersion of the particle due to resonant scattering with Alfvén waves will be 〈ΔE2〉∝(EβA)2t, where E = mec2γ. Using the very good approximation for the variance of the product of n uncorrelated random variables (Goodman 1962)

Equation (21)

and plugging Equation (2) into the equation above, we get

Equation (22)

since E is the particle energy at time t (namely, step ns), we have $E^2=(m_ec^2\gamma _0\bar{\varepsilon }^{n_s})^2$, from which follows

Equation (23)

In the following two sections (Sections 3.1 and 3.2), we study the consequences of the structure in the magnetic turbulence, on the evolution of the particle spectra, following their stochastic acceleration in the turbulent field.

3.1. Hard-sphere Turbulence

Under the "hard-sphere" approximation (q = 2), the spatial diffusion coefficient does not depend on the particle energy, since the exponent of Equation (15) is q − 2 = 0. Only three independent parameters exist in this description: the scattering time, the escape time, and the velocity of the scatterers. The spectra are purely determined by how many scatterings have been able to occur, the velocity of the scatterer, and what fraction of the injected particles has escaped out of the acceleration region. The scattering time relates to the spatial diffusion coefficient by tscatDx/c. Similarly, the resulting acceleration time relates to the spatial diffusion coefficient by taccDx2ActA2A. Thus, for "hard-sphere" turbulence, the scattering and acceleration timescales are independent of the particle energy (since there is equal energy density of scatterers which particles of all energies may resonantly scatter with).

The left panel in Figure 1 shows the resulting instantaneous evolution of spectra for the "hard-sphere" turbulence. The log-parabolic shape is maintained along the entire acceleration process, as shown by the solid lines representing the fit of the MC distributions by means of the law in Equation (9) (Cowsik & Sarkar 1984). The evolution of the curvature parameter, obtained from the r in the log-parabolic fit and plotted in Figure 2 with the red dashed line, clearly shows the trend due to the momentum diffusion, in agreement with the prediction from Equation (19) (blue line in the plot) demonstrating the connection between Dp0, ${\sigma _\varepsilon }/{\bar{\varepsilon }}$, and βA.

Figure 1.

Figure 1. Plots showing the Monte Carlo results. For comparison, the results obtained using an analytic description given in Becker et al. (2006; dashed line, right panel) and a log-parabolic function (dashed line, left panel; Cowsik & Sarkar 1984) are shown.

Standard image High-resolution image
Figure 2.

Figure 2. Curvature parameters of the energy distribution of accelerated electrons shown in Figure 1. In the case of q = 2 (red line), the trend is consistent with the "hard-spheres" prediction (blue line). In the case of Kolmogorov (green line) and Kraichnan (black line) turbulence, the trend predicts larger values compared to the "hard-spheres" prediction and r approaches an asymptotic value dictated by the exponential cutoff in the equilibrium distribution.

Standard image High-resolution image

3.2. Soft Turbulence Spectra

To account for the effects of turbulent magnetic field spectra softer than the "hard-sphere" case, we also consider acceleration in Kolmogorov- and Kraichnan-type turbulence. We have to therefore include a fourth parameter in the MC simulation, in addition to the three considered above: the turbulent field spectral slope q (see Equation (12)).

The right plot in Figure 1 shows the evolution of spectra for the "Kolmogorov" turbulence case. Similar spectra were obtained by Lemoine & Pelletier (2003) and O'Sullivan et al. (2009) who integrated the trajectories of charged particles in a turbulent magnetic field embedded in a fluid. The results are compared to the quasi-linear theory results (Becker et al. 2006; solid lines in Figure 1). We can identify two phases in the temporal evolution. In the first phase, the SEDs are more symmetric and the curvature evolves as in the q = 2 case, while in the second phase they develop a low-energy PL tail. Figure 2 shows that, for the Kolmogorov (green line) and the Kraichnan (black line) turbulence, r is systematically larger than the "hard-sphere" case (red line), and that for t ≳ 2 × tacc, r approaches an asymptotic value (r ≈ 1.2 and r ≈ 1.5 for q = 5/3 and q = 3/2, respectively) ruled by the exponential cutoff in the equilibrium distribution.

4. NUMERICAL APPROACH: DIFFUSION EQUATION WITH STOCHASTIC COMPONENT AND LOSSES

Both the MC approach and statistical description are able to explain the link between the curvature in the energy distribution of accelerated particles and the presence of a stochastic energy gain term. In order to incorporate a more complete description, taking into account the competition between radiative losses and acceleration, and its influence on the curvature, we use the diffusion equation approach, already outlined in Section 2.2, by inserting into Equation (11) a cooling term for the synchrotron and IC radiative losses. Following Moderski et al. (2005) we can write

Equation (24)

where UB = B2/8π is the energy density of the magnetic field, epsilon0 = hν0/mec2 is the IC seed photon energy in units of mec2, nph(epsilon0) is the number density of IC seed photons with the corresponding photon energy density Uph = mec2epsilon0nph(epsilon0)depsilon0. The function fKN results from the analytical integration of the Jones (1968) Compton kernel, fully taking into account Klein–Nishina (KN) effects for an isotropic seed photon field (see Moderski et al. 2005, their Appendix C), and FKN(γ) represents its convolution with the seed photon field. We remark that FKN plays a crucial role in the cooling process, depending both on the IC regime (Thomson (TH) limit for 4γepsilon0 ≪ 1, KN limit for 4γepsilon0 ≫ 1) and on epsilon0nph(epsilon0)∝B2/R2.

Since analytical solutions are possible only for a limited number of cases, to follow the complex dependence of the IC cooling term on nph(epsilon0) in a self-consistent way we must solve the diffusion equation numerically. For this purpose, we further developed the numerical code (Tramacere et al. 2009; Tramacere 2007) used to compute numerically the synchrotron and IC emission and introduced it into the numerical solution of the diffusion equation. In the numerical calculations, we adopted the method proposed by Chang & Cooper (1970) and used the numerical recipe given by Park & Petrosian (1996). This is a finite difference scheme based on the centered difference of the diffusive term, employing weighted differences for the advective term. We use a 5000 point energy grid over the range 1.0 ⩽ γ ⩽ 109, and a time grid that is finely tuned to have a temporal mesh several orders of magnitude smaller than typical cooling and acceleration timescales. The results from our code were compared, when possible, with known analytical solutions and always found good agreement.

4.1. Physical Set-up: the Relations Between Dp and tD with γmax and R

We study the evolution of n(γ) and of the curvature term in a homogeneous spherical geometry, with radius R and an entangled coherent magnetic field B and a turbulent component δB, in the two cases of impulsive and continuous injection with a quasi mono-energetic source function Qinj, t) normalized to have a fixed energy input rate:

Equation (25)

In our approach, we do not distinguish the acceleration region from the radiative one and during the acceleration process we take into account both synchrotron and IC cooling. According to Equation (14), to determine the order of magnitude of Dp we assume 1 ≫ δB/B ≃ 0.1–0.01 and require Alfvén waves to be at least mildly relativistic, with βA ≃ 0.1–0.5, and their maximum wavelength to be much smaller than the accelerator size (λmax < R). To study the effect of IC cooling on the evolution of n(γ), we consider two different sizes of the acceleration region, a compact one (R = 5 × 1013 cm) and a larger one (R = 1 × 1015 cm). With this choice of accelerator size, we set λmax ≈ 1012 cm. We stress that the choice of λmax constrains the accelerative upper limit through ρg < λmax leading to γmax < (λmaxqB)/mec2, since particles with larger ρg (hence larger γ) cannot resonate with shorter wavelengths. Taking into account a coherent magnetic field of the order of 0.1 G and λmax ≈ 1012 cm we found that the purely accelerative efficiency limits the particle energy to γmax ≲ 107.5. In the left panel of Figure 3, we plot tD, given by Equation (17), as a function of λmax, for the case of q = 2, δB/B = 0.1, and βA = 0.5. In this case, the acceleration time is energy independent and for λmax ≈ 1012 cm it will be of the order of tD = 1/Dp0 ≈ 104 s. In the case of q ≠ 2, the acceleration will have an energy dependence given by Equation (17), as shown in the right panel of Figure 3 for the case of q = 3/2. In this section, we focus on the evolution of the curvature as a function of the momentum-diffusion term, and therefore use only the accelerative contributions coming from the diffusion terms (Dp(γ), DA(γ)), neglecting the systematic extra term A(γ). All the parameters and their numerical values are given in Table 1.

Figure 3.

Figure 3. Left panel: the tD acceleration time as a function of λmax, for q = 2, δB/B = 0.1, and βA = 0.5. The vertical lines represent the Larmor radius for γ = 105 (red line), γ = 1.5 × 107 (cyan line), and γ = 108 (orange line). Right panel: the tD acceleration time for the same parameters as in the left panel, for the case of q = 3/2 and as function of γ, for the two different cases of λmax = 3 × 1010 cm (black line) and λmax = 1 × 1015 cm (purple line). The thick black line shows tD, for the case of λmax = 3 × 1010 cm, limited to the highest acceleration energy of the particles constrained by the resonant scattering limit: ρg = λmax.

Standard image High-resolution image

Table 1. Parameters' Values Adopted in the Numerical Solutions of the Diffusion Equation for the Cases Studied in Section 4

Parameter   Impulsive Inj. Cont. Inj.
R (cm) 5 × 1013, 1 × 1015 ... ... ...
B (G) 0.1, 1.0 ... ... ...
Linj (erg s−1) 1039 ... 1037 ...
q   2 3/2 2 3/2
$t_{D_0}=1/D_{P0}$ (s) 1 × 104 1 × 103 1 × 104 1 × 103
Tinj (s) 100 ... 1 × 104 ...
Tesc (R/c) ... 2 ...
Duration (s) 1 × 105 ... ... ...
γinj   10.0 ... 10.0 ...

Download table as:  ASCIITypeset image

4.2. Impulsive Injection

In the left panels of Figure 4 and Figure 5, we plot the evolution of energy distribution n(γ, t) (upper panels) and of γ3n(γ, t) (lower panels) in the case of the impulsive injection without escape, for q = 2, and for two values of R: 1 × 1015 cm (Figure 4) and 5 × 1013 cm (Figure 5). We inject a quasi-monoenergetic electron distribution with γinj ≈ 10. The γ3n(γ, t) representation is useful to compare the results concerning n(γ) presented in this section, with those regarding the synchrotron emission presented in Section 5. We denote by γp the peak energy of n(γ) and by r the curvature evaluated by means of a log-parabolic best fit over a one decade-broad interval centered at γp. γ3p and r3p represent the peak of γ3n(γ) and its curvature, respectively. In the right panels of Figures 4 and 5, we report on the corresponding temporal evolutions of the curvatures under the effect of both momentum diffusion and cooling terms. The solid black line corresponds to t = 0.2 × tacc, where $t_{{\rm acc}}=t_{D_0}$ is the acceleration time due to momentum diffusion. As the time increases, the diffusion term acts on the distribution by means of both DA and Dp. The effect of the latter is to make the distribution broader.

Figure 4.

Figure 4. Left panels: evolution of the particle spectrum with impulsive injection and no escape for the case of R = 1 × 1015 cm and q = 2. Upper panels represent the temporal evolution of n(γ); lower panels represent the temporal evolution of γ3n(γ). Solid lines represent the case of SSC cooling. Red and blue solid lines represent the final state for B = 1.0 G and B = 0.1 G, respectively. Green solid lines represent the temporal evolution, for B = 0.1 G, with step of 0.8 × tD. The dashed lines represent the final stage in the case of only synchrotron cooling. The vertical dot-dashed lines represent the equilibrium energy in the case of only synchrotron cooling. Right panels: evolution of the curvature as a function of $t/t_{D_0}$. Upper panel: curvature r evaluated at γp, for the case of SSC cooling (solid red and blue lines) and for the case of only synchrotron cooling (dashed red and blue lines). The solid green line represents the prediction from Equation (19). Lower panel: the same as in the upper panel, for the curvature r3p evaluated at γ3p (open and filled circles) compared to the case of r (solid lines).

Standard image High-resolution image
Figure 5.

Figure 5. Left panels: the same as in Figure 4, for the case of R = 5 × 1013 cm. Upper right panel: the same as in Figure 4 for the case of R = 5 × 1013 cm. Lower right panel: the evolution of the curvature r3p for the case R = 5 × 1013 cm.

Standard image High-resolution image

One can distinguish three phases: in the first one the energy of particles increases and the curvature parameter decreases following a law rt−1 in agreement with the statistical scenario of Section 2 and with the Equation (19), independent of the magnetic field strength (B = 1.0 G and B = 0.1 G) and of the source size, because the accelerative contribution dominates over the radiative losses; in the second phase, the radiation losses become relevant and the distribution approaches the equilibrium with an increase of the curvature; and in the third phase, the balance between acceleration and radiation losses is established and the curvature reaches a stable value.

The equilibrium distribution reached through stochastic acceleration, is described by a relativistic Maxwellian (Schlickeiser 1985; Stawarz & Petrosian 2008),

Equation (26)

where $f(q,\dot{\gamma })$ is a function depending on the exponent of the diffusion coefficient and on the cooling process and γeq is the Lorentz factor that satisfies the condition tcool(γ) = tacc(γ) and is given by

Equation (27)

with tacc equal to the fastest acceleration timescale among tA, tD, and tDA. In the case of Compton-dominated cooling we have γeq∝(R2/taccB2fKN), while in the case of strong KN regime, or in general for synchrotron-dominated cooling, we have γeq∝(1/taccB2). Using a PL form for the acceleration terms, and in the case of only synchrotron losses (or any cooling process that can be expressed as a PL function of γ), it is possible to give an analytic expression of $f(q,\dot{\gamma })$ (Katarzyński et al. 2006; Stawarz & Petrosian 2008). The expectation for synchrotron and IC/TH cooling process and for q = 2 is $f(q,\dot{\gamma })=3-q=1$. The curvature resulting from a log-parabolic fit over a decade centered on γp is r ≈ 2.5 and r3p ≈ 6.0 in the case of γ3p.

We first discuss the case of R = 1015 cm (Figure 4) with only synchrotron cooling (dashed lines, left panels). In terms of behavior, we note that for the larger value of B (1.0 G; red lines, right panels), the rt trend departs from the purely accelerative one (rt−1; green lines, right panels) early (relative to the B = 0.1 G case; blue lines in the right panels). This happens because the synchrotron equilibrium energy (vertical dot-dashed lines, left panels) is lower in the case of B = 1.0 G. For both values of B, the final values of r are close to the synchrotron equilibrium value of ≈2.5. When IC cooling is also taken into account, the final values of the curvature in n(γ) are r ≈ 2.5 and r ≈ 0.6 for B = 0.1 G and B = 1.0 G, respectively. This difference is due to the different IC cooling regimes for the two cases. To show clearly the complexity of the transition from the TH to the KN regime, and its dependence on R and B, in Figure 6 we plot the ratio $\dot{\gamma }_{{\rm IC}}/\dot{\gamma }_{{\rm Synch.}}$ (solid lines), and Uph/UB (dashed lines), as a function of γ and normalized to unity, for the case of q = 2, for the final step of the temporal evolution. As long as the ratio Uph/UB is close to $\dot{\gamma }_{{\rm IC}}/\dot{\gamma }_{{\rm Synch.}}$, electrons cool in the full TH regime, and C(γ) = C0γ2(UB + Uph). On the contrary, when the electrons radiate in the full KN regime $\dot{\gamma }_{{\rm IC}}/\dot{\gamma }_{{\rm Synch.}}\ll U_{{\rm ph}}/U_B$. In this case, due to the inefficient KN cooling regime we have $\dot{\gamma }_{{\rm Synch.}}\gg \dot{\gamma }_{{\rm IC}}$, and the cooling term is dominated by the synchrotron component: C(γ) ≈ C0γ2UB. In the intermediate cases, it is difficult to estimate analytically the ratio $\dot{\gamma }_{{\rm IC}}/\dot{\gamma }_{{\rm Synch.}}$.

Figure 6.

Figure 6. Normalized ratios of electron cooling rates $\dot{\gamma }_{{\rm IC}}/\dot{\gamma }_{{\rm Synch.}}$ (solid lines), and Uph/UB (dashed lines), as a function of γ for different values of R and B for the case of q = 2 at the final step of the evolution. Left panels: top, case of R = 1015 cm and B = 0.1 G. Bottom, case of R = 1013 cm and B = 0.1 G. Right panels: top, case of R = 1015 cm and B = 1.0 G. Bottom, case of R = 1013 cm and B = 1.0 G.

Standard image High-resolution image

For B = 1.0 G, the SSC equilibrium is reached at γ ≈ 3 × 104 and the SSC cooling occurs between the KN and TH regimes (see the top right panel in Figure 6), hence the value of f is different from unity, as predicted for the case of full IC/TH or synchrotron cooling. When B = 0.1 G, the equilibrium energy is γ ≈ 107 and electrons are in extreme KN cooling (see the top left panel in Figure 6), synchrotron losses are much higher than those due to IC scattering, and again r reaches the previous value of ≈2.5. It is also interesting to note the difference in the trends of rt and r3pt. In the latter case, the trend departs from the purely accelerative regime earlier (see Figure 4, lower right panel) since the electrons with energies close to γ3p are more energetic than those close to γp, and thus have much shorter cooling times.

The results for the compact region (R = 5 × 1013 cm) are plotted in Figure 5. Considering that the injected electron luminosity is the same (see Table 1), we expect a different response from the IC cooling, due to the higher photon density nph(epsilon0). The r evolution for the synchrotron cooling case is similar to the previous case, while for the SSC emission, both for the case of B = 1.0 G and B = 0.1 G, the final value of r is about 2.5. This is due to the larger photon density which moves the IC scattering into the TH regime also for the case of B = 0.1 G (compare bottom left to top left panels in Figure 6), hence n(γ) approaches the solution of Equation (26) in the case of f = 1.

In Figure 7, we show the temporal evolution for the case of q = 3/2 (R = 1.0 × 1015 cm, B = 0.1 G). In this case, contrary to the q = 2 case, the acceleration time tD is energy dependent, hence we study the evolution of r as a function of t/tDinj), where tDinj) is the diffusive acceleration time evaluated at the injection energy γinj. The energy dependence of tD affects the evolution of r, and the shape of the equilibrium distribution, indeed, the rt and r3pt trends show different behavior compared to the case of q = 2. The equilibrium curvature is reached for t ≳ 1 × tDinj), and the two equilibrium curvature values are r ≈ 1.2 and r3p ≈ 3.0, roughly half of those found for the case of q = 2, and in agreement with the result from the MC. We note that the curvature obtained by means of a log-parabolic fit of Equation (26), for the case q = 3/2 (namely, f = 1.5), is r ≈ 3.7. Hence, both the MC and the numerical solution of the diffusion equation give a result different from that predicted by the analytical solution in Equation (26).

Figure 7.

Figure 7. Left panels: evolution of the particle spectrum with impulsive injection and no escape for the case of R = 1 × 1015 cm, B = 1.0, and q = 3/2. Since tD is energy dependent, on the x-axis we plot the ratio t/tDinj), where tDinj) is the diffusive acceleration time evaluated at the injection energy γinj. Green solid lines represent the temporal evolution, for B = 0.1 G, with step of 2.4 × tD0). Right panels: evolution of the curvature r (upper) and r3p (lower).

Standard image High-resolution image

4.3. Continuous Injection

The case of continuous injection (see Figure 8) is more complex. The distribution developes a low-energy PL tail, but a log-parabolic bending, driven by the diffusion, is still present at high energies, hence we evaluate the curvature only at γ3p (i.e., the representation useful to compare it to the synchrotron SED emission). Spectral curvatures are generally milder than the impulsive injection. In the left panel of Figure 8, we plot the rt trend both for the case of impulsive (red lines) and continuous (blue lines) injections, the curvature in the continuous injection case are systematically lower in the pre-equilibrium phases and in the acceleration-dominated stage the trend is again consistent with the "hard-sphere" approximation and statistical approaches. The slope of the electron distribution in the PL tail is ≈1.06, in good agreement with the predicted one ≈1 + tmin-acc/(2tesc) = 1.075, consistent with the results of Katarzyński et al. (2006).

Figure 8.

Figure 8. Left panel: evolution of the particle spectrum for continuous injection, R = 1 × 1015 cm, B = 1.0 G, and q = 2. Right panel: evolution of the curvature r3p.

Standard image High-resolution image

5. EVOLUTION OF THE SPECTRAL PARAMETERS OF SYNCHROTRON AND IC EMISSION

The most relevant parameters describing the SED of SSC sources provided by observations are the peak energies and curvatures of the synchrotron and IC components. We denote these curvature parameters by bs and bc, respectively, and by Es, Ec, and Ss, Sc, we denote the corresponding SED peak energies and flux values. We use νs and νc to indicate the corresponding SED peak frequencies. In the following, we describe the results of the relations between these parameters assuming that electrons are injected into the acceleration region with a quasi mono-energetic spectrum with γinj ≈ 10 and using an injection time of 104 s. We use the same working hypothesis for the momentum-diffusion coefficient as in Section 4.1 and add a systematic acceleration time for the first-order process tA = 1.5 × 103 s, in order to produce Es values ranging between optical and hard X-ray energies. We set the radius of the region at R = 2 × 1015 cm and the same duration for the injection and acceleration processes, namely, 104 s. We varied the other parameters of the model, B, q, and Dp0 to verify how they affect the relation between the observable ones. All the parameters and their variation ranges are summarized in Table 2.

Table 2. Parameters' Values Adopted in the Numerical Solutions of the Diffusion Equation for the Cases Studied in Section 5

Parameter Range
R (cm) 2 × 1015
B (G) [0.01, 1.0]
Linj (erg s−1) 1038
q   [3/2, 2]
tA (s) 1.8 × 103
$t_{D_0}=1/D_{P0}$ (s) [1.5, 25] × 104
Tinj (s) 104
Tesc (R/c) 2.0
Duration (s) 104
γinj   10.0

Download table as:  ASCIITypeset image

A phenomenological approach, based on the δ-function approximation (Tramacere et al. 2007, 2009; Massaro et al. 2006, 2004), is useful to address the expected relation between the curvature parameters and their connections with the peak energies and flux values. According to the standard synchrotron theory (e.g., Rybicki & Lightman (1986)), in the δ-function approximation, the synchrotron SED peak value and the corresponding peak energy can be expressed by the following relations:

Equation (28)

which implies

Equation (29)

where α = 1.5 applies for changes of γ3p leaving constant n3p), α = 2 for variations of B only, and α = 4 when the main driver is δ. For a log-parabolic-shaped n(γ) we have

Equation (30)

and, using the relation bsr/5 (Massaro et al. 2004), or, more precisely, from the analysis presented in Section 4.2, bsr3p/5. It follows that

Equation (31)

The relation between bs and Es is

Equation (32)

with a = 3/5 = 0.6.

The spectral properties of the IC emission are more complex, depending on the transition from the TH to the KN regime (see Massaro et al. 2006 for a detailed discussion). In the former case, the curvature is close to that of the synchrotron emission but systematically smaller due to the energy redistribution by the scattering process. In the transition to the KN regime, the energy of IC photons will approach γmec2, hence the IC spectral shape will reflect that of the high-energy tail of n(γ) and the curvature bc will be closer to that of the electrons. Then, provided the IC scattering happens in the TH regime, the trends involving bc are expected to be similar to those of bs, but systematically show bc < bs. As the KN regime is approached, bc changes differently from bs, converging toward r.

5.1. Temporal Evolution of bs and bc

We compute the evolution of bs and bc, as a function of the time, for the case of $t_{D_0}=1.5\times 10^4$ s, B = 0.1 G, and q = 2, using a temporal mesh of 2 s. We plot in the top left panel of Figure 9 the instantaneous SEDs at steps of 200 s: the solid lines represent the synchrotron and IC SEDs averaged over the full duration of the acceleration process (104 s). As the time is increased, the peak energy of both the synchrotron and IC SEDs moves toward higher energies with a broadening of the spectral distribution. The corresponding evolution of curvature parameters is reported in the top right panel: bs has a trend similar to that of the electron distribution, with bs∝(t/tD0)−α and α ≃ 0.6 (for comparison the cyan solid line represents the r3p/5 trend, as predicted by the S δ-approximation). The trend of bc, as expected, is more complex because of the transition from TH to KN regime. For t/tacc ≲ 0.4, it follows the same trend of bs but with systematically lower values. For t/tacc ≳ 0.4, when the TH–KN transition occurs, bc increases with time, approaching toward the electron curvature r value. This transition starts for values of Es ≈ 5 × 10−3 keV (νs ≈ 1014 Hz) and Ec ≈ 0.05 GeV (νc ≈ 1022 Hz); and the corresponding SEDs are plotted by blue thick-dashed lines in the left panel of Figure 9.

Figure 9.

Figure 9. Left panels: evolution of synchrotron (black dashed lines) and IC (red dashed lines) SEDs, for the case of $t_{D_0}=1.5\times 10^4$ s and q = 2 (top panel), and for the case of tDinj) ≈ 6.3 × 104 s and q = 3/2 (bottom panel). All the other parameters are as reported in Table 2. The solid lines represent the SEDs averaged over the full simulation period, and the blue dashed lines (top panel) represent the SEDs corresponding to the transition from TH to KN regime. Right panels: the temporal evolution of bs (black squares) and bc (red squares) as a function of $t/t_{D_0}$, for the case of q = 2 (top panel), and q = 3/2 (bottom panel). The cyan line (top panel) represents the bs trend predicted for the synchrotron emission in the case of the δ-approximation. The dashed lines (top panel) represent the PL best fit of both bs (purple) and bc (blue) trends.

Standard image High-resolution image

In the bottom panels of Figure 9, we show the case of q = 3/2. The synchrotron curvature quickly approaches the equilibrium value of bs ≈ 0.6, consistent with the equilibrium value r3p ≈ 3.0 discussed in Section 4.2. In this case we do not observe the TH/KN transition in the IC curvature, since the lower values of Es and Ec keep the IC scattering mainly in the TH regime.

5.2. EsSs and Esbs as a Function of Dp0 and q

The other parameter affecting the evolution of the spectral distributions is the diffusion coefficient Dp0 (see Equation (15)), which we assume to vary in the range [1.5 × 104, 2.4 × 105] s−1, studying how the main spectral parameters change. In the top left panel of Figure 10, we plot averaged SEDs for each different value of Dp0. The top right panel shows the trend of bc versus Dp0. As expected, for larger values of Dp0, the curvature measured at the peak energy is smaller. The trend is described by a PL with an exponent of about −0.6 for Dp0 ≲ 2 × 10−5 s−1 and with an exponent of about −0.25 for Dp0 ≳ 2 × 10−5 s−1. This break clearly shows the transition between the TH and KN regimes (marked by a vertical dashed line); indeed it happens for the same values of Dp0 corresponding to the TH/KN transition in both the Dp0bc trend and the Ecbc plot (occurring at Ec ≈ 1 GeV; see the bottom right panel in Figure 10). The break in the Dp0bs trend happens when electrons radiating at Es enter the KN cooling region, hence, due to the lower cooling level (compared to the TH cooling regime, on the left side of the vertical dashed line), the curvature decreases.

Figure 10.

Figure 10. Upper left panel: synchrotron (red lines) and IC (red lines) average SEDs for each different value of $t_{D_0}$ in the range reported in Table 2, with q = 2. Blue points represent the position of ES, C and SS, C. The purple, orange, and green line represent the PL best fit of the ESSS and ECSC trends. Upper right panel: bs and bc, for each average SED in the right panel, as a function of Dp0. Dashed lines represent the PL best fit of the bDp0 trend. Lower left panel: the bsEs trend obtained by means of a log-parabolic best fit of the averaged SEDs plotted in the upper right panel. Lower right panel: same as in the lower left panel, for bcEc.

Standard image High-resolution image

Blue filled circles in the top left panel represent the peak positions for both SED components. For the synchrotron component, according to Equation (29), the exponent α in the case of n3) = const, should be 1.5, while the results of the computations give α = 0.6. This difference is due to the fact that we inject in the mono-energetic initial distribution always the same total power that corresponds to the same number of particles. When the peak energy increases the distribution becomes broader, implying that the same total number of particles is spread over a larger energy interval and the number of particles contributing to the synchrotron peak emission decreases. Consequently, the SsEs trend gets softer compared to the predicted value of 1.5.

We verified quantitatively this effect by computing the trend n3p) versus γ23p, and found a PL relation with an exponent of about 0.98, in nice agreement with the difference between the exponent of 1.5 and that resulting in our simulations. In the bottom panels of Figure 10, we plot bs versus Es (left) and bc versus Ec (right). The ScEc relation can be fitted by a PL (orange line, top left panel in Figure 10) with the same exponent of the EsSs relation, as long as the IC scattering, at Ec and above, happens in TH regime. When the KN suppression becomes relevant (green line, top left panel in Figure 10), the exponent is larger and is close to unity.

The synchrotron trend (the bottom left panel in Figure 10) clearly shows the expected anti-correlation between the peak energy and the spectral curvature, which is well fit by the function given in Equation (32), with a = 0.68, not very different from 0.6, obtained for the δ-function approximation of the synchrotron emission, and assuming that n(γ) has a purely log-parabolic shape. A simple PL fit of the same points returns an exponent −0.14.

We also investigate the role of q on the spectral evolution, setting its variation range to [3/2, 2], i.e., from the Kraichnan to the "hard-sphere" case. The relations between the spectral parameters are very similar to those found in the previous case with SsE0.6s and ScE0.9c. Also in this case, the synchrotron component follows the expectation with a lower curvature for harder turbulence spectra, and the IC trend shows the transition from TH to KN regime. The PL best fit of S(Es) versus Es gives a = 0.88, larger than that obtained for the case of Dp0. In fact, for values of q lower than 2, corresponding to less turbulence and hence diffusion, the curvature gets higher values and the peak energy lower values, compared to the "hard-sphere" case. The PL fit for bs versus Es returns an exponent of −0.16, practically coincident with the previous one, indicating that the average properties of these parameters are the same in both the q and Dp0 cases.

5.3. Es, cSs, c and Es, cbs, c as a Function of B

The magnetic field B drives the radiative losses which affect the evolution of the spectral parameters. In Section 4, we showed that different cooling conditions, and the transition from TH to KN, can determine very different values of γeq for the same acceleration conditions. Assuming that the acceleration timescale is independent of the magnetic field, Equation (27) shows that γeq∝1/B2, implying that, as long as B is small enough to result in γ3p ≪ γeq, the evolution of n(γ) around the peak value is dominated by the acceleration terms, while for values of B resulting in γ3p ≳ γeq the evolution obtains a notable contribution due to cooling. In the top left panel of Figure 11, we plot the averaged SEDs. According to Equations (28) and (29), the synchrotron peak value should scale as Ss∝(Es)2. Indeed, for values of B ≲ 0.2 G we obtain an exponent equal to 2.04, very close to the value found with the δ-approximation. For higher values of the magnetic field, Es is anti-correlated with Ss. This behavior represents a cooling signature due to the decreasing value γeq for increasing B values, with γeq getting closer to γ3p. This is confirmed both by the shape of the synchrotron SEDs and by the bsB plot (top right panel in Figure 11). Indeed, S SEDs for B ≳ 0.2 G exhibit an exponential decay, meaning that the distributions have reached, or are close to reaching, the equilibrium energy. Consistently with the S shape evolution, the bsB relation shows an almost stable value of bs for B ≲ 0.2 G and an increasing trend for B ≳ 0.2 G. This change, in both the SsEs and bsB trends, is interesting and can provide a useful phenomenological tool for understanding the evolution of non-thermal sources. Another interesting feature is shown in the ScEc plot: for B ≲ 0.2 G the IC peak energy is practically constant, as expected in the KN limit from the kinematical limit relating the scattered photons energy to that of the electrons: hνIC ≈ γmec2. In fact, photons at energies ≈Ec are produced in the KN regime and for B ≲ 0.2 G the electron peak energy γ3p is constant, so Ec must also be constant. For B ≳ 0.2 G, γ3p decreases because of cooling, and, accordingly, Ec also decreases. This is another interesting test that can provide a probe for B-driven flares evolving to the KN regime. The Esbs plot in the bottom left panel of Figure 11 confirms the cooling signature discussed above, showing bs uncorrelated with Es as long as γ3p ≪ γeq, and an increasing value of bs with Es almost stable, when γ3p ≳ γeq.

Figure 11.

Figure 11. Same as in Figure 10, for different values of B in the range reported in Table 2.

Standard image High-resolution image

6. SPECTRAL EVOLUTION OF HIGH ENERGY FLARES OF BRIGHT HBL OBJECTS

The previous considerations on the spectral evolution of SSC sources, in which high energy electrons are accelerated in a relatively short timescale by stochastic processes, can be successfully applied to describe the behavior of some bright HBLs objects. These sources are, in fact, characterized by having the synchrotron peak in the UV/X-ray range and the IC peak in γ rays up to TeV energies. Several flares, observed simultaneously in both these ranges, exhibited SEDs very well described by a log-parabolic law, whose parameters, particularly their curvature, are estimated with high accuracy. A similar analysis for low-energy peaked BL Lac objects is much more difficult because the peak of their synchrotron component is typically in the infrared range and the available simultaneous multifrequency data are extremely few. Tramacere et al. (2007, 2009) and Tramacere (2007) pointed out that the observed anticorrelation between Es and bs in the synchrotron SED of Mrk 421 can provide a clear signature of a stochastic component in the acceleration process. In the same analysis, these authors also presented an interesting correlation between Es and Ss. Massaro et al. (2008) found that the Esbs and EsSs trends hold also for a larger sample of 11 HBLs, strengthening the hypothesis that a common accelerative mechanism may drive such physical processes for this class of active galactic nuclei. To give a theoretical framework to these phenomenological relations, we try to reproduce both the Esbs and EsSs relations derived from the data of the aforementioned papers. In the following, we will consider the data of Mrk 421 from Tramacere et al. (2007, 2009) collected over a period of 13 years, and of six HBL objects from Massaro et al. (2008): Mrk 180, Mrk 501, PKS 0548−322, PKS 1959−650, 1H 1426+428, covering a period of about 11 years and including both quiescent and flaring states. The sources from Massaro et al. (2008) were chosen because the data are good enough to safely constrain both curvature and Es values, and because the observed variations of the sample luminosity are compatible with the assumption to be driven by changes of Es.

Following the analysis presented in Section 5, we consider two scenarios in which these trends are driven by the momentum-diffusion term. In the first case, the momentum diffusion changes because of variations of Dp0, due to changes of δB/B or βA, but the turbulence spectrum (q = 2) remains stable. In the second scenario, the turbulence spectrum is variable with q ranging in [3/2, 2]. We use the same method described in Section 5 to compute the averaged SEDs for each value of Dp (or q); computations are performed for three values of the magnetic field B = 0.05, 0.1, and 0.2 G. All the model parameters are summarized in Table 3.

Table 3. Parameters' Values Adopted in the Numerical Solutions of the Diffusion Equation to Reproduce the Observed Trends of the HBLs Reported in Section 6

Parameter   D Trend q Trend
R (cm) 3 × 1015 ...
B (G) [0.05, 0.2] ...
Linj (Esbs trend) (erg s−1) 5 × 1039 ...
Linj (EsLs trend) (erg s−1) 5 × 1038, 5 × 1039 ...
q   2 [3/2, 2]
tA (s) 1.2 × 103 ...
$t_{D_0 }=1/D_{P0}$ (s) [1.5 × 104, 1.5 × 105] 1.5 × 104
Tinj (s) 104 ...
Tesc (R/c) 2.0 ...
Duration (s) 104 ...
γinj   10.0 ...

Download table as:  ASCIITypeset image

The comparison with the data can be affected by an observational bias due to the limited energy range of detectors. In fact, when the peak energy is close to the limits the curvature is not well estimated because one can use only a portion of the parabola below or above the peak. Generally, curvatures lower than the actual ones are obtained. The energy range [0.5, 100.0] keV is the typical spectral window covered by X-ray and hard-X-ray detectors. In our analysis, we used this fixed window to take into account this possible bias in the observed data when Es is variable.

6.1. Esbs Relation

The Esbs trend, and in particular the anticorrelation between these two observables parameters, is the strongest signature of a stochastic component in the acceleration.

In Figure 12, we report the scatter plot in the Esbs plane for the six considered sources. The left panel reports the results obtained by changing the value of Dp0: the green dashed lines describe the trend resulting from a log-parabolic fit of the synchrotron SED over a decade in energy centered on Es; the purple lines represent the same trend obtained by fitting a log-parabola in the fixed spectral window [0.5, 100.0] keV. Both these trends are compatible with the data and track the predicted anticorrelation between Es and bs. Purple data, however, give a better description, hinting that the "window" effect could be a real bias. Each of the three lines was computed for a different value of the magnetic field. It is remarkable that the variation of a single parameter, Dp0, can describe the observed behavior. The dispersion in the data is relevant and can be related to the variation of B (as partially recovered by numerical computation), or by different values of the beaming factor, R, and Linj, during different flares, and for different objects.

Figure 12.

Figure 12. Left panel: the Esbs trend observed for the six HBLs in our sample. The dashed green lines represent the trend reproduced by the stochastic acceleration model, for the parameters reported in Table 3 and for the D trend; the different lines corresponding to three different values of B reported in Table 3. The purple lines represent the trend obtained by fitting the numerically computed SED over a fixed spectral window in the range 0.5–100 keV. Right panel: the same as in the left panel for the case of the q trend.

Standard image High-resolution image

The dot-dashed thick line represents the best fit of the observed data by means of Equation (32), and returns a value of a ≈ 0.6, as expected from theoretical predictions for the case of the δ-approximation, and pure log-parabolic electron distribution. This fitted line is also compatible with the numerical trend shown by the purple lines. Note that the observed curvature values are in the range [0.1, 0.5], corresponding to r3p ∼ [0.5, 3.0]. According to the results presented in Section 4.2, the expected equilibrium curvature in the synchrotron emission, in the full KN or TH regime, and for q = 2, should be of r3p ≈ 6.0 and of r3p ≈ 5.0 in the intermediate regime. In the case of q = 3/2, the equilibrium curvature should be r3p ∼ 3.0. This is perhaps an interesting hint that, both in the flaring and the quiescent states, for q = 2, the distribution is always far from equilibrium. In the case of q = 3/2, only for Es ≲ 1.5 keV is the curvature compatible with the equilibrium (r3p ≃ 3.0, corresponding to bs ∼ 0.6). For larger values of Es, we find again curvature well below the equilibrium value. These results provide a good constraint on the values of the magnetic field B ≲ 0.1 G.

The q-driven trend (right panel) is also compatible with the data, but for values of Es ≲ 1 keV, the Dp0-driven case seems to describe better the observed behavior, but any firm conclusion is not possible because of the dispersion of the data.

6.2. EsLs Trend

As a last benchmark for the stochastic acceleration model, we reproduce the observed correlation between Es and Ss, which follows naturally from the variations of Dp0 and q. Considering that the redshifts of the six considered HBL objects are different, we prefer to use their peak luminosity Ls = SsD2L, where DL is the luminosity distance.4 To account for the different jet power of sources, we considered two data subsets, and we assumed Linj = 5 × 1039 erg s−1 for the first subset (top panels of Figure 13) and Linj = 5 × 1038 for the second (bottom panels of Figure 13). In the left panels of Figure 13, we report the Dp0-driven trend and in the right panels we show the q-driven trend. Solid lines represent the trend obtained by deriving Ls from the log-parabolic best fit of the numerically computed SEDs, centered on Es; dashed lines are the trends obtained by fitting the numerical results in the fixed energy window [0.5, 100] keV.

Figure 13.

Figure 13. Left panels: the EsLs trend observed for six HBLs in our sample; the top panel corresponds to the case of Linj = 5 × 1039 erg s−1; the bottom panel corresponds to the case of Linj = 5 × 1038. The solid black lines represent the trend reproduced by stochastic acceleration model, for the parameters reported in Table 3 and for the D trend, the different lines corresponding to three different values of B reported in Table 3. The dashed lines represent the trend obtained by fitting the numerically computed SED over a fixed spectral window in the range 0.5–100 keV. Right panels: the same as in the left panel for the case of the q trend.

Standard image High-resolution image

Both results give a good description of the observed data and their shapes are similar. Solid lines follow well a PL with an exponent of about 0.6, while the windowed trends (dashed lines) show a break around 1 keV and the exponent below this energy turns to about 1.5. A similar break at the same energy can be noticed in the points of Mrk 421 in the EsSs plot presented by Tramacere et al. (2009), who found an exponent of ∼1.1 and of ∼0.4 below and above 1 keV, respectively. This could again be an indication that the observed values are actually affected by the bias.

7. DISCUSSION

Broadband observations of non-thermal sources have shown that the spectral curvature at the peaks of their SEDs can now be measured with good accuracy. In this paper, we have presented, using different approaches, the relevance of these data for the understanding of the competition between statistical acceleration and radiation losses. First, using a simple statistical approach and MC calculations, we have shown that the log-parabolic energy distribution of the relativistic electron is a good picture in the first phases before equilibrium is reached. In this case, the curvature decreases with time and, therefore, with increasing peak energies. This evolution is confirmed by numerical solutions of the diffusion equation taking properly into account both stochastic acceleration and radiative SSC cooling. The major results can be summarized as follows.

The evolution of the electron energy distributions (Section 4) shows that:

  • 1.  
    In the case of synchrotron and SSC cooling, and for all the values of B and R, as long as the distribution is far from equilibrium, the trend on r is dictated by Dp and is well described by Equation (19).
  • 2.  
    When the distributions approach equilibrium, the value of r is determined by the shape of the equilibrium distribution, which is a relativistic Maxwellian, with the sharpness of the cutoff determined by both q and the IC cooling regime.
  • 3.  
    In the case of q = 2, and for equilibrium energies implying that IC cooling happens either in the TH regime or in the extreme KN regime (IC cooling negligible compared to the synchrotron one), the numerical solution of the diffusion equation follows the analytical prediction (f = 1, that holds for any $\dot{\gamma }\propto \gamma ^2$), and the corresponding equilibrium curvature is r3p ≈ 6.0 (bs ≈ 1.2). In the case of q = 3/2, the equilibrium curvature is r3p ≈ 3.0 (bs ≈ 0.6). These limiting values could be a useful observational test to find cooling-dominated flares with the distribution approaching to the equilibrium.
  • 4.  
    When cooling is in the intermediate regime between TH and KN and for the q = 2 case, the condition f = 1 fails, and the end values of r decrease, strongly depending on the balance between UB and the seed IC photon energy (Uph) numerical computations are necessary to evaluate the right value of r at equilibrium.

The analysis of the spectral evolution of SSC emission (Section 5) shows that:

  • 1.  
    Changes of Dp0 (or q) imply that the curvature and peak energy of the synchrotron emission are anticorrelated; the Esbs trend can be phenomenologically described by Equation (32).
  • 2.  
    The Ecbc trend presents a clear signature of the transition from the TH to the KN regime. In particular, when the IC scattering approaches the KN regime we observe a sharp change in the bc, with a positive correlation with Ec, while in the TH regime the correlation is negative as in the case of the Esbc.
  • 3.  
    The magnetic field plays a relevant role on the cooling process and B-driven variations present relevant differences compared to those due to Dp0 (and q).

In particular, for the B-driven case, we note first that the EsSs correlation follows the prediction of the synchrotron theory and shows the PL relationship with Es∝(Ss)∼2.0. On the contrary, in the case of Dp0 and q changes, we find Es∝(Ss)0.6. Another relevant difference in the B-driven case is the evolution of Sc. For the case of Dp0- and q-driven trends, Sc relates to Ec through a PL with exponent of about [0.7, 0.8]. On the contrary, for the B-driven case with IC scattering in the full KN regime, the value of Ec is almost constant and uncorrelated with Sc (see Figure 11) due to the kinematic limit of the KN regime. Ec starts to decrease when B is enough large to make the cooling process dominant. This is an interesting signature that could be easily checked in the observed data.

The comparison of the Esbs and EsSs trends, obtained through several X-ray observations of six HBL objects spanning a period of many years, with those predicted by the stochastic acceleration model, shows very good agreement. We are able to reproduce these long-term behaviors by changing the value of only one parameter (Dp0 or q). Interestingly, the EsSs relation follows naturally from that between Es and bs. This result is quite robust and hints at a common accelerative scenario acting in the jets of HBLs.

As a last remark, we note that very recently Massaro & Grindlay (2011) also find that in the case of GRBs a Esbs trend similar to that observed in the case of HBL objects. They measured values of the curvature up to 1.0, typically higher than in HBLs. It is interesting to note that the value of 1.0 is close to the limit of ∼1.2 that we predict in the case of distributions approaching the equilibrium in either TH or KN regime for q = 2.

We thank the anonymous referee for providing us with constructive comments and suggestions.

This work has been partially supported by Università di Roma La Sapienza (Dipartimento di Fisica, Gruppo SCAE).

Footnotes

  • We used a flat cosmology model with H0 = 73 km s−1 Mpc−1, Ωmatter = 0.27, and Ωvacuum = 0.73.

Please wait… references are loading.
10.1088/0004-637X/739/2/66