1: % mn2esample.tex
2: %
3: % v2.1 released 22nd May 2002 (G. Hutton)
4: %
5: % The mnsample.tex file has been amended to highlight
6: % the proper use of LaTeX2e code with the class file
7: % and using natbib cross-referencing. These changes
8: % do not reflect the original paper by A. V. Raveendran.
9: %
10: % Previous versions of this sample document were
11: % compatible with the LaTeX 2.09 style file mn.sty
12: % v1.2 released 5th September 1994 (M. Reed)
13: % v1.1 released 18th July 1994
14: % v1.0 released 28th January 1994
15:
16: \documentclass[useAMS,usenatbib]{mn2e}
17:
18: % If your system does not have the AMS fonts version 2.0 installed, then
19: % remove the useAMS option.
20: %
21: % useAMS allows you to obtain upright Greek characters.
22: % e.g. \umu, \upi etc. See the section on "Upright Greek characters" in
23: % this guide for further information.
24: %
25: % If you are using AMS 2.0 fonts, bold math letters/symbols are available
26: % at a larger range of sizes for NFSS release 1 and 2 (using \boldmath or
27: % preferably \bmath).
28: %
29: % The usenatbib command allows the use of Patrick Daly's natbib.sty for
30: % cross-referencing.
31: %
32: % If you wish to typeset the paper in Times font (if you do not have the
33: % PostScript Type 1 Computer Modern fonts you will need to do this to get
34: % smoother fonts in a PDF file) then uncomment the next line
35: % \usepackage{Times}
36:
37: %%%%% AUTHORS - PLACE YOUR OWN MACROS HERE %%%%%
38: \newcommand{\vect}[1]{\mathbf{#1}} % Vector as bold
39: \newcommand{\dif}[2]{\frac{{\rm d} #1}{{\rm d} #2}}
40: \usepackage{graphicx}
41: \usepackage{epstopdf}
42: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
43:
44: \title[Ultra-High-Energy Cosmic Rays from the Radio Lobes of AGNs]{Ultra-High-Energy Cosmic Rays from the Radio Lobes of AGNs}
45: \author[F. Fraschetti and F. Melia]{F. Fraschetti$^{1,2}$\thanks{E-mail:
46: federico.fraschetti@cea.fr; melia@physics.arizona.edu} and F. Melia$^{3}$\\
47: $^{1}$LUTh, Observatoire de Paris, CNRS-UMR8102 and Universit\'e Paris VII,
48: 5 Place Jules Janssen, F-92195 Meudon C\'edex, France.\\
49: $^2$Laboratoire AIM, CEA/DSM - CNRS - Universit\'e Paris Diderot,
50: Irfu/Service d'Astrophysique, F-91191 Gif sur Yvette C\'edex, France.\\
51: $^{3}$Department of Physics and
52: Steward Observatory, The University of Arizona, Tucson, AZ 85721, USA.}
53:
54: \begin{document}
55:
56: \date{Accepted 2008 September 19. Received 2008 September 19; in original form 2008 July 23}
57:
58: \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2008}
59:
60: \maketitle
61:
62: \label{firstpage}
63:
64: \begin{abstract}
65: \noindent In the past year, the HiRes and Auger collaborations have
66: reported the discovery of a high-energy cutoff in the ultra-high energy
67: cosmic-ray (UHECR) spectrum, and an apparent clustering of the highest
68: energy events towards nearby active galactic nuclei (AGNs). Consensus is
69: building that such $\sim 10^{19}$--$10^{20}$ eV particles are accelerated
70: within the radio-bright lobes of these sources, but it is not yet clear how
71: this actually happens. In this paper, we report (to our knowledge) the
72: first treatment of stochastic particle acceleration in such environments from first
73: principles, showing that energies $\sim 10^{20}$ eV are reached in $\sim
74: 10^6$ years for protons.
75: However, our findings reopen the question regarding whether the high-energy
76: cutoff is due solely to propagation effects, or whether it does in fact
77: represent the maximum energy permitted by the acceleration process itself.
78: \end{abstract}
79:
80: \begin{keywords}
81: cosmic rays -- physical data and processes: acceleration of particles; plasmas; turbulence -- galaxies: active; nuclei
82: \end{keywords}
83:
84: \section{Introduction}
85:
86: Cosmic rays are energetic charged particles traveling throughout
87: the Galaxy and the intergalactic medium under the influence of various physical
88: processes, including deflection by magnetic fields, and collisions with
89: other particles along their trajectory. Their energy spectrum measured at
90: Earth is a steep (roughly power-law) distribution with logarithmic index
91: $\alpha\sim 2.6$--3, extending up to a few times $10^{20}$ eV. Other than
92: their spectrum, these particles are characterized by their angular distribution
93: in the sky, and by their mass composition.
94:
95: A highly significant steepening in the UHECR spectrum was
96: reported by both the HiRes (Abbasi et al. 2008) and \citet{auger2008c}
97: collaborations. (Distinguished from their lower-energy counterparts,
98: UHECRs have energies in excess of 1 EeV $\equiv 10^{18}$ eV.) This result
99: may be a strong confirmation of the predicted Greisen-Zatsepin-Kuzmin (GZK)
100: cutoff due to photohadronic interactions between the UHECRs and low-energy
101: photons in the cosmic microwave background (CMB) radiation (Greisen
102: 1966; Zatsepin \& Kuz'min 1966). Together with the measured low
103: fraction of high-energy photons in the CR distribution, this measurement
104: already rules out so-called top-down models, in which the UHECRs represent the
105: decay products of high-mass dark matter particles created in the early Universe
106: (Semikoz et al. 2007). The measured photon flux is also in conflict with
107: scenarios in which UHECRs are produced by collisions between cosmic strings
108: or topological defects (Bluemer et al. 2008, Auger 2008b). On the other
109: hand, such energetic particles may still be produced via astrophysical
110: acceleration mechanisms (see Torres \& Anchordoqui 2004 and other references cited therein).
111:
112: UHECRs are not detected directly, but through the showers they create in
113: Earth's atmosphere (see, e.g., Melia 2009). Depending on the energy and
114: type of primary particle, the ensuing cascade has characteristics that
115: allow the ground-based observatories to determine not only whether the
116: incoming UHECR is a photon, but also its atomic number. It should be
117: pointed out, however, that a determination of the primaries'
118: composition strongly relies on an extrapolation of current
119: phenomenological hadronic interaction models, so it remains rather
120: uncertain. The \citet{auger2007} data confirmed the dominance
121: of protons in primary cosmic rays, though they also exhibit evidence for
122: a mixed composition extending to energies as high as $\sim$ 50--60 EeV,
123: with a higher atomic number $Z$, up to $Z \sim 26$ (Unger et al. 2007).
124:
125: But the most telling indicator for the possible origin of these UHECRs is
126: the discovery by Auger (Auger 2008a) of their clustering towards nearby
127: ($\sim 75$ Mpc) AGNs along the supergalactic plane. The significance
128: of this correlation has been further strenghtened by a more recent
129: analysis which weights the AGN spatial distribution by their hard
130: X-ray flux (George et al. 2008). This raises at least two questions:
131: (1) How are the UHECRs accelerated to such high energies? and (2) given
132: these nearby sources, is the sharp suppression of UHECRs in the last
133: decade of their observed energies really due solely to the GZK effect,
134: or does it signal a limitation to the acceleration efficiency?
135:
136: Previous attempts at understanding how particles are accelerated to
137: EeV energies and beyond have generally been based on first-order Fermi
138: acceleration (see, e.g., Ostrowski 2008, and other references therein)
139: within shocks created by blast waves like those in supernova remnants
140: (Fatuzzo \& Melia 2003, Crocker et al. 2005). But this process is
141: subject to kinematic restrictions that inhibit the particles from
142: actually reaching ultra-high energies (see, e.g., Nayakshin \& Melia
143: 1998, Gallant et al. 1999). Recent numerical simulations have shown
144: that an increase in the Lorentz factor $\gamma$ of ultra-relativistic
145: shock waves steepens the observed spectrum (Niemiec \& Ostrowski 2006)
146: and reduces its high-energy cutoff.
147:
148: For these reasons, it is not plausible for UHECRs to emerge from astrophysical
149: environments, such as supernova remnants, where first-order processes are dominant
150: so long as the shock velocity is super-Alfv\'enic, because they cannot even contain such
151: high-energy particles (Hillas 1984)---the gyration radius of particles with energy
152: $\sim 10^{20}$ eV for a typical galactic magnetic field
153: is much larger than the size ($<10$ pc) of these structures.
154:
155: On the other hand, a second-order Fermi process (Fermi 1949)
156: can explain observational features not addressed by the first-order process,
157: as in the case of a supernova remnant itself (see Cowsik \& Sarkar 1984).
158: Moreover, stochastic particle acceleration through a gyroresonant
159: interaction with MHD turbulence (a second-order Fermi process; see Fermi 1949) can be
160: very efficient if the Alfv\'en velocity approaches $c$ (Dermer \& Humi 2001).
161: The stochastic acceleration of particles by turbulent plasma waves has already
162: received some attention in the literature (see Liu et al. 2004, 2006, and references
163: cited therein, and Wolfe \& Melia 2006). Indeed, the feasibility of second-order
164: Fermi acceleration in radio galaxies has been demonstrated through the steady
165: re-acceleration of electrons in certain hot spots (Almudena Prieto et al. 2002).
166:
167: Our treatment from first
168: principles, however, avoids many of the previously encountered unknowns and
169: limitations. In this paper, we report (to our knowledge) the first treatment of
170: stochastic acceleration of charged particles in the lobes of radio-bright AGNs
171: by directly computing the trajectory of individual particles. An earlier version
172: of this treatment---for the propagation of charged particles assumed already
173: accelerated at TeV energy through the turbulent magnetic field
174: at the Galactic centre---may be found in Ballantyne et al. 2007,
175: and Wommer, Melia \& Fatuzzo 2008; by contrast, in the present paper
176: both the propagation and acceleration are taken into account. We show
177: that random scatterings (a second-order Fermi process) between the charges
178: and fluctuations in a turbulent magnetic field can accelerate these particles
179: up to ultra-high energies, provided a broad range of fluctuations is present
180: in the system.
181:
182: \section{Description of the model}
183:
184: In our treatment, we follow the three-dimensional motion of {\it individual}
185: particles within a time-varying turbulent magnetic field. By avoiding the use
186: of equations describing statistical averages of the particle distribution, we
187: mitigate our dependence on unknown factors, such as the diffusion coefficient.
188: We also avoid such limitations as the Parker approximation (Padmanhaban 2001)
189: in the transport equation. However, a remaining unknown is the partitioning
190: between turbulent and background fields. For simplicity, we take the minimalist
191: approach and assume that the magnetic energy is divided equally between the two
192: components.
193:
194: Another unknown is the turbulent distribution. For many real astrophysical plasmas,
195: the magnetic turbulence seems to be in accordance with the Kolmogorov spectrum.
196: This is seen, e.g., in the solar wind (Leamon et al. 1998) and through interstellar
197: scintillation (Lee \& Jokipii 1976); a more recent numerical analysis of MHD
198: turbulence confirms the general validity of the Kolmogorov power spectrum (Cho et
199: al. 2003). In addition, renormalization group techniques applied to the analysis of MHD
200: turbulence also favour a Kolmogorov power spectrum (for more details, see
201: Smith et al. 1998, and Verma 2004).
202:
203: We model the radio lobe of an AGN as a sphere of radius ${\mathcal R}$,
204: a second parameter in our simulations.
205: A population of relativistic particles of mass $m$, protons or heavy ions,
206: with an energy $E = \gamma mc^2$, where $\gamma$ is the Lorentz factor,
207: is released in an inner sphere of radius ${\mathcal R'} \sim \alpha {\mathcal R}$.
208: The value of $\alpha$ must be much smaller than 1, otherwise very few particles
209: reach an energy $E >10^{18}$eV. For a small value of $\alpha$, the gyration
210: radius becomes comparable to the size of the acceleration region at $E >10^{18}$eV,
211: and therefore changes in $\alpha$ do not significantly alter the result.
212: For the sake of specificity, we use a value $\alpha \sim 10^{-3}$ in this paper.
213: Once released, the particles propagate through the turbulent field until they
214: escape the accele\-ration region and enter intergalactic space.
215:
216: \subsection{Time varying turbulent field}
217:
218: We use the \citet{gj94} prescription for gene\-rating the turbulent magnetic
219: field. Their principal aim of propagating individual particles through a magnetostatic
220: field was to compute the Fokker-Planck coefficients for a direct
221: comparison with analytic theory. For our purpose, we modify that prescription
222: to include a time-dependent phase factor that allows for temporal variations.
223:
224: The global magnetic field is written as a sum of a background term $\vect{B}_0$,
225: constant and uniform, and a turbulent field varying in space and time (i.e.,
226: as a superposition of Alfv\'en waves).
227:
228: The equation of motion of a relativistic test particle with charge $e$ and mass
229: $m$ moving in an electromagnetic field $F^{\mu \nu}$ is the Lorentz equation
230: (Landau \& Lifchitz 1975, Melia 2001)
231: \begin{equation}
232: mc \frac{d u^\mu}{ds} = \frac{e}{c} F^{\mu \nu} u_\nu
233: \label{geo}
234: \end{equation}
235: (with $\mu = 0, 1, 2, 3$), where $c$ is the speed of light in vacuum, $u^\mu
236: =\left(\gamma, \gamma {\vect{v}}/{c} \right)$
237: is the four-velocity of the particle, $\gamma = 1/ \sqrt{1-(v/c)^2}$
238: is the Lorentz factor, and $s/c$ is the proper time.
239: We calculate the trajectory of the particle in a ma\-gnetic field
240: $\vect{B}(t, \vect{r}) = (mc/e)\vect{\Omega}(t, \vect{r})$
241: as a solution of the space components ($\mu = 1,2,3$) of Equation (1)
242: \begin{equation}
243: \frac{d\vect{u}(t)}{dt} = \delta \vect{\cal E}(t, \vect{r}) + \frac{\vect{u}(t)
244: \times\vect{\Omega}(t,\vect{r})}{\gamma(t)}\;,
245: \label{lorentz}
246: \end{equation}
247: where $t$ is the time in the rest frame of the acceleration region.
248: The quantity $\vect{\Omega}(t, \vect{r})$ in Equation (\ref{lorentz}) is given by
249: \begin{equation}
250: \vect{\Omega}(t, \vect{r}) =
251: \vect{\Omega}_0+\delta \vect{\Omega}(t, \vect{r})\;,
252: \end{equation}
253: where $\vect{\Omega}_0 \equiv (e/mc)
254: \vect{B}_0$, in terms of the background magnetic field $\vect{B}_0$,
255: and $\delta \vect{\Omega}(t, \vect{r})$ is the time-dependent turbulent
256: magnetic field. We ignore any large-scale background electric fields---a
257: reasonable assumption given that currents would quench any such fields
258: within the radio lobes of AGNs. The time variation of the magnetic field,
259: however, induces an electric field $\delta \vect{\cal{E}}(t, \vect{r})
260: \equiv (e/mc) \vect{E}(t, \vect{r})$ according to Faraday's law.
261:
262: The procedure of building the turbulence calls for the random
263: generation of a given number $N$ of transverse waves $\vect{k}_i$, $i=1,..,N$ at every point of
264: physical space where the particle is found, each with a random amplitude, phase
265: and orientation defined by angles $\theta (k_i)$ and $\phi(k_i)$. This form of the
266: fluctuation satisfies $\nabla \cdot \delta \vect{\Omega}(t, \vect{r})=0$. We write
267: \begin{equation}
268: \delta \vect{\Omega}(t, \vect{r}) = \sum_{i=1}^N\Omega(k_i) \hat{\xi}_{\pm} (k_i)
269: e^{\left[i(k_i x' - \omega_i t +
270: \beta(k_i))\right]}\;,
271: \label{omega}
272: \end{equation}
273: where the polarization vector is given by
274: \begin{equation}
275: \hat\xi_{\pm} (k_i)= \cos\alpha(k_i)\hat{\bf y}' \pm i\sin\alpha(k_i)\hat{\bf z}'\;.
276: \end{equation}
277: Given the form in Equation (4) for the turbulence, the electric field $\delta \vect{\cal{E}}(t, \vect{r})$ is given by
278: \begin{equation}
279: \delta \vect{\cal{E}}(t, \vect{r}) =
280: \sum_{i=1}^N\Omega(k_i) \frac{\omega(k_i)}{k_i c} \hat\xi_{\pm} ^E (k_i)
281: e^{\left[i(k_i x' - \omega_i t +
282: \beta(k_i))\right]}\;,
283: \label{electric}
284: \end{equation}
285: with
286: \begin{equation}
287: \hat\xi_{\pm} ^E (k_i)= \pm i\sin\alpha(k_i)\hat{\bf y}' - \cos\alpha(k_i)\hat{\bf z}'\;.
288: \end{equation}
289: The orthonormal primed coordinates ${\bf r'} = (\hat{x}',\hat{y}',\hat{z}')$
290: are related to the lab-frame coordinates ${\bf r} = (\hat{x},\hat{y},\hat{z})$
291: via the rotation matrix $R(\theta,\phi)$, in such a way that for every $k$ the
292: propagation vector is parallel to the $\hat{x}'$ axis. The matrix $R(\theta,\phi)$ is given by
293: \begin{equation}
294: {\bf r'} =
295: \left(
296: \begin{array}{ccc}
297: \cos\theta \; \cos\phi &\cos\theta \; \sin\phi &\sin\theta \\
298: -\sin\phi &\cos\phi &0 \\
299: -\sin\theta \; \cos\phi &-\sin\theta \; \sin\phi &\cos\theta \\
300: \end{array}
301: \right) {\bf r}
302: \;.
303: \end{equation}
304: For each value of $k_i$, there are 5 random numbers: $0<\theta(k_i)<\pi$, $0<\phi(k_i)<2\pi$, $0<\alpha(k_i)<2\pi$,
305: $0<\beta(k_i)<2\pi$ and the sign plus or minus indicating the sense of polarization.
306:
307: Further assumptions are necessary to specify the dispersion relation $\omega=\omega(k_i)$.
308: For every turbulent mode, we use the dispersion relation for transverse non-relativistic
309: Alfv\'en waves (see Kaplan \& Tsytovich 1973 for an extended discussion):
310: $\omega(k_i) = v_A k_i \cos\theta(k_i)$, for $i=1,..,N$,
311: where $v_A = B_0/\sqrt{4\pi m_p n}$ is the non-relativistic
312: Alfv\'en velocity in a medium with background magnetic field $B_0$
313: and number density $n$, $m_p$ the proton mass, and $\theta(k_i)$ is
314: the angle between the wavevector $\vect{k}_i$ and $\vect{B}_0$. This is the condition
315: thought to be valid for the propagation of turbulent modes in a magnetized
316: astrophysical environment, such as the radio lobes of an AGN. The background
317: plasma is assumed to have a background proton number density $n \sim 10^{-4}$
318: cm$^{-3}$, a reasonable value for these environments (Almudena Prieto et al. 2002).
319:
320: The amplitudes of the magnetic fluctuations are assumed to be
321: consistent with Kolmogorov turbulence, so
322: \begin{equation}
323: \Omega(k_i) = \Omega(k_{min}) \left(\frac{k_i}{k_{min}}\right)^{-\Gamma/2}\;,
324: \end{equation}
325: for $i=1,..,N$, where $k_{min}$ corresponds to the longest wavelength of the fluctuations and
326: the index $\Gamma$ of the power spectrum $\Omega^2(k)$ is $5/3$. Finally, the
327: quantity $\Omega(k_{min})$ is computed by requiring that the energy density of
328: the magnetic fluctuations equals that of the background magnetic field:
329: \begin{eqnarray}
330: \lefteqn{
331: S = \sum_{i=1}^N \frac{B^2(k_i)}{8\pi} = \frac{m^2 c^2}{8 \pi e^2} \Omega^2 (k_{min})
332: \sum_{i=1}^N \left(\frac{k_i}{k_{min}}\right)^{-\Gamma}= \frac{{B_0}^2}{8 \pi}.} \nonumber\\
333: \label{sum}
334: \end{eqnarray}
335: We choose N=2400 values of $k$ evenly spaced on a logarithmic scale;
336: i.e., a wavenumber shell with bounds
337: $k_i - k_{i+1}$ holds $k_{i+1} = k_i \times (k_{max}/k_{min})^{1/N}$ values.
338: Considering that the
339: turbulence wavenumber $k$ is related to the turbulent length scale $l$ by $k = 2\pi / l$,
340: we adopt a range of lengthscales from $l_{min} = 10^{-1}\,v_0/\Omega_0$ to $l_{max} =
341: 10^{9}\,v_0/\Omega_0$, where $v_0$ is the initial velocity of the
342: particle and $\Omega_0$ is the initial gyrofrequency in the background magnetic field.
343: Thus the dynamic range covered by $k$ is $k_{max} / k_{min} = l_{max} / l_{min} = 10^{10}$,
344: and our description allows for $240$ transverse modes $k$ per decade.
345: The values of $k_{max}$ and $k_{min}$ fix the magnetic energy equipartition
346: through Equation (10). The value ${k_{min}} ^{-1}$ is proportional through a factor of order $1$
347: to the correlation length of the turbulence (see Ruffert and Melia 1994, and Rockefeller
348: et al. 2004, for examples of how this is generated in the interstellar medium);
349: the value ${k_{max}} ^{-1}$ is the wavelength at which the interaction between
350: the turbulence and most of the particles is the most efficient, so that energy
351: is drained out of $\delta \Omega$.
352:
353: However, since the gyroradius $r_g (E)$ evolves over
354: a large energy interval, the gyroresonant wavenumber $k_{res} (E)$ moves accordingly
355: in such a way that in the global wavenumber interval $(k_{min} - k_{max})$ there is
356: for every $E$ a certain $k_{res} (E)$ fulfilling the resonance condition $r_g (E) k_{res} (E) \sim 1$.
357: Such a $k$ range involves a large computational time,
358: especially if a statistically significant number of particles is to be considered.
359:
360: Since our numerical simulation is not performed by specifying the magnetic turbulence
361: on a computational grid with given cell size $\Delta x$, the choice of $k_{max} = 2
362: \pi / l_{min}$ is not dictated by a fixed spatial resolution (see section \ref{num}
363: for more details). In addition, the result is not affected by spurious effects to the
364: discreteness or the periodicity. As a bypro\-duct, the divergenceless condition
365: $\nabla \cdot \delta \vect{\Omega}(t, \vect{r})=0$ is easily satisfied and does not
366: require an extension of the Godunov solver of the MHD equations for the purpose
367: of ``divergence cleaning'' (Ryu et al. 1998) or a reformulation of the MHD
368: equations including, e.g., divergence-damping terms (Dedner et al. 2002).
369:
370: With this prescription, we construct the turbulent magnetic field at every point
371: of physical space where the particle is found, which we then propagate
372: without taking any time-average along the trajectory. The particles passing through
373: this region are released initially at a random position inside the acceleration
374: zone, which for simplicity is taken to be a sphere of radius ${\mathcal R}$, with
375: a fixed initial velocity $v_0$ pointed in a random direction. The initial value of the
376: Lorentz factor $\gamma_0 = 1/\sqrt{1 - (v_0/c)^2} \simeq 1.015$ is chosen to avoid having
377: to deal with ionization losses for the protons or heavy ions. The particles closest to
378: the edge of the acceleration zone have a higher probability of escaping than
379: those starting farther in, and therefore reach relatively lower energies. In the
380: usual (Fermi) way, this produces (in the highest energy portion of the spectrum)
381: an inverted power-law distribution.
382:
383: \subsection{Energy losses}
384:
385: In principle, energy losses due to synchrotron and inverse Compton processes
386: involving radio and Cosmic Microwave Background (CMB) photons, all of which increase
387: as $\gamma^2$, can significantly limit the maximum energy attainable by a
388: cosmic ray during the acceleration process, given that its Lorentz factor
389: $\gamma$ evolves from $\sim 1$ up to $10^{10}-10^{11}$. For an UHE particle
390: (either a proton or a heavy ion), both the radio and CMB photons will have
391: an energy $\gamma h \nu$ in the centre-of-momentum frame well below the rest
392: energy of the cosmic ray (i.e., $\gamma h \nu << m c^2$, where $m$
393: is the mass of the accele\-rating particle). For the purpose
394: of these estimates, we use a radio frequency $\nu_{radio} = 0.1$ GHz
395: ($h \nu_{radio} \sim 4.2 \times 10^{-7}$ eV) and
396: a CMB frequency corresponding to the peak of the blackbody spectrum,
397: $\nu_{CMB} = 2.821 kT/h = 158$ GHz ($h \nu_{CMB} \sim 6.6 \times 10^{-4}$ eV),
398: where $k$ is the Boltzmann constant,
399: $h$ the Planck constant, and $T = 2.7$ K is the CMB temperature. Consequently,
400: the energy losses due to inverse Compton may be calculated in the Thomson limit.
401: Compare this with the situation for high energy electrons, for which the Thomson
402: condition would not be satisfied even at energies $\gamma m_e c^2 \sim 10^{16}$ eV,
403: requiring in that case the full Klein-Nishina treatment.
404:
405: The propagation of high-energy particles is here mo\-deled in a region of tens of kpc size.
406: Therefore we neglect any effect of the relativistically-narrowed jet on the spatial distribution
407: of the radio background, assumed for simplicity to be isotropic.
408: Since the CMB intensity field is also isotropic, we take these
409: energy losses into account using the following angle-integrated power-loss rate:
410: \begin{equation}
411: -\frac{dE}{dt} = \frac{4}{3} \sigma_T(m) c \gamma^2
412: \left(\frac{B^2}{8\pi} + U_{R} + U_{CMB}\right)\;,
413: \label{loss}
414: \end{equation}
415: where $\sigma_T(m) = 6.6524 \times (m_e/m)^2\, 10^{-25}$ cm$^2$ is the Thomson
416: cross section for a generic particle of mass $m$, which can be a proton or heavy ion,
417: and $B^2/(8\pi) = (2 {B_0}^2)/(8\pi)$ is the total energy density of the
418: magnetic field. The photon energy density $U_{R}$ inside a typical radio
419: lobe is computed as $U_R = L/(4\pi c {\mathcal R}^2)$,
420: where we assume $L$ to be a standard luminosity density corresponding to
421: the Fanaroff-Riley class II of galaxies ($L = 5\times10^{25}$
422: W Hz$^{-1}$ sr$^{-1}$ at $178$ MHz), and ${\mathcal R}$
423: is the size of the spherical acceleration zone.
424: For the CMB, we use $U_{CMB} = a T^4
425: = 4.2 \times 10^{-13}$ erg cm$^{-3}$. In a region where magnetic turbulence
426: is absent or static, a given test particle propagates by ``bouncing" randomly
427: off the inhomogeneities in $\vect{B}$, but its energy remains constant. The
428: field we will model below, however,
429: is comprised of time-varying gyroresonant turbulent waves
430: (see Equation \ref{omega}), and collisions between the test particle and these
431: waves produces a net acceleration (in the lab frame).
432:
433: \begin{figure}
434: \begin{center}
435: \includegraphics[width=9cm]{coord_GJ.eps}
436: \caption{Three-dimensional trajectory of single particle in the turbulent field
437: of \citet{gj94} reproduced with our code. The length scales are in units of the
438: gyration radius, $v_0 / {\Omega_0}$, which remains constant during the propagation.
439: The energy is verified to be constant, as expected, over a time interval
440: $\Delta t = 1000 {\Omega_0}^{-1}$, within a relative error of $10^{-5}$.
441: The background magnetic field $B_0$ is parallel to the $z$ axis and, as found by
442: \citet{gj94}, the diffusion along ${\bf B}_0$ dominates with respect to that across
443: the field.}
444: \label{GJ}
445: \end{center}
446: \end{figure}
447:
448: \begin{figure}
449: \begin{center}
450: \includegraphics[width=9cm]{coord_1part_B-8.eps}
451: \caption{Three-dimensional trajectory of a single particle
452: released at random within the acceleration zone, assumed to be a sphere of
453: radius 50 kpc. The scale on the axes is in units of $10^8 v_0/\Omega_0$,
454: where $v_0/\Omega_0$ is the initial gyroradius.
455: The particle is released with a fixed initial speed $v_0$, but pointed in
456: a randomly chosen direction.
457: The calculation stops when the particle leaves the radio lobe and
458: is injected into the intergalactic medium.}
459: \label{part}
460: \end{center}
461: \end{figure}
462:
463: \section{Numerical code setup}\label{num}
464:
465: In this section we describe the numerical code used to perform the simulations.
466: The Lorentz Equation (\ref{lorentz}) is integrated using a Runge-Kutta $4^{th}$
467: order method (Press et al. 1997) for the system of 6 first-order differential equations
468: \begin{eqnarray}
469: \dif{{\bf x_i}}{t} & = & \frac{c}{\gamma}{\bf u_i}\\
470: \frac{d\vect{u_i}}{dt} & = & \delta \vect{\cal{E}}_i + \frac{[\vect{u}
471: \times\vect{\Omega}]_i}{\gamma}\;,
472: \end{eqnarray}
473: for $i=1,2,3$. The components $\delta \vect{\cal{E}}_i$ and $\vect{\Omega}_i$ are
474: intended to be the real parts of the corresponding complex quantities.
475:
476: The portable random number generator used to produce the turbulence is
477: Knuth's subtractive routine {\it ran3} (Press et al. 1997),
478: with a seed number $I=10^9$. This routine has a relatively
479: short execution time and is suitable to avoid the introduction
480: of unwanted correlations into the numerical computation.
481:
482: There are two approaches to numerically implementing a turbulent magnetic field
483: generated by this method.
484: The first approach (used in Giacalone \& Jokipii 1994) is to calculate the magnetic
485: field at every time step for each particle position.
486: The position is then found by solving the Lorentz Equation (\ref{lorentz}).
487: In the second approach, the magnetic field is generated for a given volume at the beginning of the
488: simu\-lation and then it evolves according to Equation (\ref{omega}).
489: In order to have an acceptable $k$-binning with a dynamical range of $k_{max} / k_{min} = 10^{10}$,
490: one would then need to specify the field at an excessively large number of lattice points.
491: This is not only time-consuming, but also very memory-intensive.
492: So, like Giacalone \& Jokipii (1994), we adopt the former approach. In
493: this way, the magnetic field is generated only where needed, and
494: the overwhelming amount of computer memory required by the second approach is not necessary,.
495: Since the confinement volume is a parameter of the model,
496: the second approach would also require adapting the lattice spacing
497: in order to maintain the same space resolution in physical space.
498:
499: The Runge-Kutta integrator has previously been vali\-dated
500: for the cases of uniform and constant electric and magnetic fields,
501: where the outcome of the simulation can be compared with an analytical solution.
502:
503: Secondly, as a validation test of the code, we reproduced the result of Giacalone
504: \& Jokipii (1994) for
505: the case of a 3D magnetostatic turbulence, by using the same set of para\-meters.
506: We discretize the turbulence in 50 transverse modes $k$,
507: where the values of $k$ are chosen to be evenly spaced in logarithmic scale
508: in the interval of the corresponding length scales from $l_{min} = (1/5)\,v_0/\Omega_0$ to
509: $l_{max} = 10\,v_0/\Omega_0$, where $v_0$ is the initial velocity of the
510: particle and $\Omega_0$ is its gyrofrequency in the background magnetic field.
511: The particles are released in a random initial position
512: with initial velocity randomly oriented but with fixed value $v_0$.
513: In Figure~\ref{GJ}, we present the trajectory of a single particle,
514: the position along the three axes expressed in units of the gyration radius.
515: In this test, the energy of the particle is constant within a relative error of $10^{-5}$
516: over a time interval corresponding to $10^3 \Omega_{0} ^{-1}$.
517:
518: In order to produce the time-dependent turbulent magnetic field considered in this paper,
519: between two successive shufflings of all five random quantities in $\delta \vect{\Omega}$,
520: which are performed every $\Delta t \sim 10^8 - 10^9$s,
521: the particle propagates gyroresonantly with the oscillating turbulence.
522: We verified that a change in the Runge-Kutta time-binning by over one order of magnitude
523: does not produce any systematic numerical effects associated with
524: $\gamma(t)$ and the spectrum. Changing the $k$-binning from $N=1200$ to $N=3000$
525: in Equation \ref{omega} similarly does not noticeably change the resultant
526: $\gamma(t)$ and the spectrum. We chose $N=2400$ which results in a reasonably long
527: computational time. However, a coarser $k$-binning, e.g.,
528: with $10$ modes/decade, could possibly result in a worse determination of the
529: macroscopic indicator as instantaneous spatial diffusion coefficients;
530: the time evolution of the diffusion coefficients is however beyond the scope of the present paper.
531:
532: \begin{figure}
533: \begin{center}
534: \includegraphics[width=8cm]{comp_B-7-9_bw.eps}
535: \caption{Simulated time evolution of the Lorentz factor $\gamma$
536: for a proton propagating through a time-varying turbulent magnetic field.
537: The three curves correspond to three different values of $B_0$: $10^{-7}$,
538: $10^{-8}$, and $10^{-9}$ gauss. The protons are released
539: at an initial random position inside
540: the acceleration zone---a sphere of radius ${\mathcal R} = 50$ kpc---with
541: the same initial speed $v_0$, though pointed in random directions.
542: The proton is followed until it leaves the acceleration zone and enters
543: the intergalactic medium. The acce\-leration timescale $\Delta t$ is inversely
544: proportional to the background field $B_0$. Therefore, as expected, a larger
545: $B_0$ produces a more efficient acceleration. In this example, a proton
546: winding its way through a field $B_0 = 10^{-8}$ gauss attains an energy
547: $E \sim 10^{20}$ eV in approximately $10^6$ years.}
548: \label{gamma2}
549: \end{center}
550: \end{figure}
551:
552:
553: \section{Results and discussion}
554: Figure~\ref{part} (to be compared with Figure~\ref{GJ})
555: shows the trajectory of a single particle released at
556: random within the acce\-leration zone, with the initial speed $v_0$,
557: pointed in random direction, with an ambient magnetic field $B_{0} = 10^{-8}$ gauss.
558: In Figure~\ref{gamma2}, we plot the time evolution of the particle Lorentz factor $\gamma$
559: for three representative values of the background field $B_0$: $10^{-7}$,
560: $10^{-8}$, and $10^{-9}$ gauss. We see the particle undergoing various phases of
561: acceleration and deceleration as it encounters fluctuations in $\vect{B}$.
562:
563: The acceleration of the particle results from the 0-$th$ component of Equation \ref{geo}, which reads
564: \begin{equation}
565: \frac{d\gamma}{dt} = \frac{\delta \vect{\cal E}^i u_i}{\gamma},
566: \label{gamma0}
567: \end{equation}
568: where $\delta \vect{\cal E}^i u_i$ is the scalar product of the electric field and the 3-velocity of the particle.
569: Therefore the acceleration is given by
570: $\gamma (t) = \sqrt{{\gamma_0}^2 + 2 \int_{t_0} ^t \delta \vect{\cal E}^i u_i dt'}$. The integrand function can be strongly time-varying and therefore needs to be computed numerically.
571:
572:
573: As can be seen in Figure~\ref{gamma2}, the acceleration timescale $\Delta t_{acc}$ is inversely
574: proportional to $B_0$ so, as expected, more energetic turbulence accelerates the particles
575: more efficiently, in agreement with what was expected from the non-relativistic
576: Alfv\'en wave theory. Previous studies (Casse et al. 2002) of particle transport through
577: a turbulent magnetic field using the prescription in Giacalone \& Jokipii (1994),
578: compared with a Fast Fourier Transform method,
579: showed that the time of confinement within the jets of FR II galaxies is
580: too small for the particles to attain an energy of $10^{20}$ eV. In our case,
581: the particle acceleration takes place over a much bigger volume
582: (with dimension ${\mathcal R} = 50$ kpc, compatible with the known size of FR-II radio galaxies)
583: and the efficiency is enhanced by the strong temporal variation of the turbulence.
584: We find, in particular, that particles can easily accelerate to UHE on timescales short
585: compared to the age of the radio-lobe structure through a gyroresonant interaction
586: with a magnetic turbulence.
587:
588: In our simulation, the particle acceleration is efficient because it occurs over
589: a wide range of turbulent fluctuations, such that the wave-particle interaction
590: is resonant at all times. Such a distribution is expected if the magnetic
591: energy cascade proceeds (without loss) from the largest spatial scales down
592: to the region where energy dissipation and transfer to the particles becomes
593: most efficient.
594:
595: \begin{figure}
596: \begin{center}
597: \includegraphics[width=8cm]{histo_norm_B1d-8_R.eps}
598: \caption{Calculated differential spectra for 500 protons with $B_0 = 10^{-8}$ gauss
599: and for different values of the size of the acceleration region, assumed to be a sphere of radius spanning
600: the interval ${\mathcal R}=[5-160]$ kpc. The dependence of the energy cutoff on ${\mathcal R}$
601: is evidenced. This result shows that the cutoff in the observed spectral distribution can be due
602: to the competition between two distint effects: propagation through the CMB
603: and intrinsic properties of the accelerator.
604: Moreover, the slope in the region $E > 4 \times 10^{18}$ eV strongly depends
605: on $R$. This diagram supports the view that the steeper CR spectrum
606: below $\log(E/eV) \approx 18.6$
607: likely represents a population of galactic cosmic rays.}
608: \label{histo}
609: \end{center}
610: \end{figure}
611:
612: In Figures \ref{histo} and \ref{histo2} we show the spectral distributions for distinct values of
613: the turbulent energy and size of the acceleration region.
614: In order to produce this result, we followed the
615: trajectory of 500 protons, launched in the manner described above, with different values of the parameters
616: $R$ and $B_0$.
617:
618: We conclude that the observed spectral cutoff (Abbasi et al. 2008, Auger 2008b) can result from
619: the competition of two distint effects: not only the GZK cutoff,
620: namely degradation of primary UHECRs
621: due to the propagation through the CMB, but also, and possibly dominant,
622: intrinsic properties of the source which constrain the process of acceleration.
623:
624: \begin{figure}
625: \begin{center}
626: \includegraphics[width=8cm]{histo_norm_R50kpc_B.eps}
627: \caption{Calculated differential spectra for 500 protons with $R = 50$ kpc
628: and for different values of the turbulent magnetic energy. In this case $B_0$ spans
629: the interval $B_0=1.5\times[10^{-9}-10^{-8}]$ gauss. See the comments in Figure \ref{histo}.}
630: \label{histo2}
631: \end{center}
632: \end{figure}
633:
634: Figure~\ref{gamma3} depicts the calculated differential spectrum in the energy range
635: $\log(E/{\rm eV}) = [18.6-19.5]$. From our sampling of the various physical
636: parameters, we infer that for a radius ${\mathcal R} = 50$ kpc,
637: $B_0$ should lie in the range $[0.5-5]\times10^{-8}$
638: gauss in order to produce UHECRs with the observed distribution shown in
639: this figure.
640:
641:
642:
643: \begin{figure}
644: \begin{center}
645: \includegraphics[width=8cm]{histo_norm_zoom.eps}
646: \caption{Calculated differential spectrum for 1,000 protons in the energy range
647: $\log(E/eV) = [18.6-19.5]$ for the selected parameters
648: $B_0 = 10^{-8}$ gauss and ${\mathcal R} = 50$ kpc.
649: A power-law behaviour with index $-2.6$ in the
650: differential spectrum of protons injected into the intergalactic medium
651: in this model is in agreement with a recent statistical analysis of HiRes observations
652: (Gelmini et al. 2007). This good match supports the view that the steeper CR spectrum
653: below $\log(E/eV) \approx 18.6$
654: represents a different population, possibly associated with the Galaxy itself.}
655: \label{gamma3}
656: \end{center}
657: \end{figure}
658:
659: It is worth emphasizing that this calculation was carried out
660: without the use of several unknown factors often required in approaches
661: solving the hybrid Boltzmann equation to obtain the phase-space distribution
662: function for the particles. In addition, we remark that the acceleration mechanism
663: we have invoked here is sustained over 10 orders of magnitude in particle energy,
664: beginning at $\gamma\sim 1$; the UHECRs therefore emerge naturally from the
665: physical conditions thought to be prevalent within the giant radio lobes of
666: AGNs, without the introduction of any additional exotic physics (for a
667: complete review of the bottom-up models,
668: see Bhattacharjee \& Sigl 2000 and other references cited therein).
669:
670: To provide the possibility of observationally testing the model we have presented here,
671: we show in Figure \ref{el_loss} the temporal evolution of the
672: energy loss rate to first order in $\gamma h \nu / (m_e c^2)$ (Blumenthal \& Gould 1970)
673: for a single electron propagating through the same magnetic turbulence we have used
674: to accelerate the protons and heavier ions. By estimating the flux of UHE protons
675: $\dot N_p$ escaping from one giant radio lobe, under the assumption of neutrality
676: in the source, we can estimate the flux of accelerated electrons $\dot N_e$.
677: In principle, it is therefore possible to estimate the expected radio luminosity
678: from these regions due to this particular acceleration process.
679:
680: \begin{figure}
681: \begin{center}
682: \includegraphics[width=8cm]{synchro_B1-8_1d23_el_bw.eps}
683: \caption{Energy loss rate of a single electron in a turbulent time-varying magnetic field
684: with $B_0 = 10^{-8}$ gauss. In this diagram both the energy losses
685: due to synchrotron and inverse Compton on the
686: Radio and CMB photons in the acceleration region are shown. Unlike the case
687: of protons and heavy ions, the radiation rate for an electron exceeds the
688: acceleration rate in such a way that, for the given $B_0$,
689: in a time of order of $10$ years the electron will have lost all of its energy.}
690: \label{el_loss}
691: \end{center}
692: \end{figure}
693:
694: We should point, however, that though a comparison of our results with the
695: observations supports the viability of this model, our calculations are subject
696: to several factors we have not fully explored here. For example, the observed
697: spectrum may be affected by the cosmological evolution in source density (\citet{bgg05}).
698: However, this omission will not be overly constraining since a likelihood
699: analysis (\citet{gks07}) of the dependence of the observed distribution
700: on input parameters has already shown that, in the case of protons,
701: for $m \sim 0$, where $m$ is the evolution
702: index in the source density, the HiRes observations are compatible with a
703: power-law injection spectrum with index $-2.6$. The analysis of the Auger
704: data seems to confirm this (\cite{auger2008c}).
705: Thus, although source-density evolution may alter
706: our results somewhat, our conclusions will probably not be greatly affected.
707: In a more conservative interpretation, the result presented here provides
708: the injection spectrum from a single source.
709:
710: Second, we have not included the GZK effect for particle energies above 50
711: EeV. This omission becomes progressively more important as the energy approaches
712: $10^{20}$ eV. These refinements, in addition to a more detailed analysis of the
713: composition of primary UHECRs, will be reported in a forthcoming paper.
714: Any discussion concerning the evolution of the instantaneous
715: spatial diffusion coefficients parallel to and perpendicular to the background
716: magnetic field, and on the transition to a diffusive regime, will also be
717: reserved to a future publication.
718:
719: In our approach, we have also neglected the backreaction of the accelerated particles
720: on the turbulent field which might increase the ratio $|\delta \vect{\Omega}| / |\vect{\Omega}|$,
721: and bring about a possible local failure of the assumption of isotropy of the turbulence.
722:
723: We remark that the mechanism of stochastic acceleration presented here may
724: be functioning even for a population of particles, protons or heavy nuclei,
725: pre-accelerated to an initial energy $E \sim 10^{12}- 10^{15}$ eV, e.g.,
726: by multi-shock fronts propagating at super-Alfv\'enic velocity.
727: The corresponding gyroresonant wavenumber range in this case
728: will decrease down to $k_{max}/k_{min} \sim 10^4 - 10^5$.
729:
730: \section{Conclusion}
731:
732: We have shown that a region containing a Kolmogorov (turbulent) distribution
733: of non-relativistic Alfv\'en waves can accelerate particles to ultra-high
734: energies. The physical
735: parameters in these regions are compatible with those believed to
736: be operating in the radio lobes of AGNs. We have discussed the predicted
737: differential spectrum within the parameter space of the model,
738: characterized by the size ${\mathcal R}$ of the acceleration region and the
739: turbulent magnetic energy. Possible tests of this model involve the synchrotron
740: or IC emission by a population of similarly accelerated electrons.
741:
742: As the Auger observatory continues to gather more data, improving on the statistics,
743: our UHECR source identification will continue to get better. Eventually, we should
744: be able to tell how significant the GZK effect really is, and whether the cutoff
745: in the CR distribution is indeed due to propagation effects, or whether it is
746: primarily the result of limitations in the acceleration itself. Given the fact
747: that energies as high as $\sim 10^{20}$ eV may be reached within typical
748: radio lobes, it is possible that both of these factors must be considered in
749: future refinements of this work.
750:
751:
752: \section*{Acknowledgments}
753: {\footnotesize\noindent
754: FF thanks J. Kirk, D. Semikoz, Y. Gallant and A. Markowith for useful and stimulating discussions.
755: The work of FF was supported by CNES (the French Space Agency) and was carried
756: out at Service d'Astrophysique, CEA/Saclay and partially at
757: Laboratory Universe and Theories (LUTh) in
758: Observatoire de Paris-Meudon and at the Center for
759: Particle Astrophysics and Cosmology (APC) in Paris.
760: This research was partially supported by NSF grant 0402502 at the
761: University of Arizona. Part of this work was carried out at
762: Melbourne University and at the Center for Particle Astrophysics
763: and Cosmology in Paris.}
764:
765: \begin{thebibliography}{99}
766:
767: \bibitem[\protect\citeauthoryear{Abbasi et al.}{2008}]{H07}
768: Abbasi R.U. et {\it al.} (HiRes Collaboration), 2008, Phys. Rev. Lett. 100, 101101.
769:
770: \bibitem[\protect\citeauthoryear{Auger}{2007}]{auger2007}
771: The Auger Collaboration, 2007, Science, 318, 939.
772:
773: \bibitem[\protect\citeauthoryear{Auger}{2008a}]{auger2008a}
774: The Auger Collaboration, 2008a, Astropart. Phys., 29, 188.
775:
776: \bibitem[\protect\citeauthoryear{Auger}{2008b}]{auger2008b}
777: The Auger Collaboration, 2008b, Astropart. Phys., 29, 243.
778:
779: \bibitem[\protect\citeauthoryear{Auger}{2008c}]{auger2008c}
780: The Auger Collaboration, 2008c, Phys. Rev. Lett., 101, 061101.
781:
782: \bibitem[\protect\citeauthoryear{Almudena Prieto et al.}{2002}]{abm02}
783: Almudena Prieto, M., Brunetti G., Mack K.-H., 2002, Science, 298, 193.
784:
785: \bibitem[\protect\citeauthoryear{Ballantyne et al.}{2007}]{ball2007}
786: Ballantyne, D. R., Melia, F., Liu, S., and Crocker, R. M., 2007, ApJ
787: Lett., 657, L13.
788:
789: \bibitem[\protect\citeauthoryear{Berezinsky et al.}{2006}]{bgg05}
790: Berezinsky V., Gazizov A., Grigorieva S., 2006, Phys. Rev. D, 74, 043005, 1.
791:
792: \bibitem[\protect\citeauthoryear{Bhattacharjee \& Sigl}{2000}]{bs00}
793: Bhattacharjee P., Sigl G., 2000, Phys. Rep., 327, 109.
794:
795: \bibitem[\protect\citeauthoryear{Bluemer et al.}{2008}]{b08}
796: Bluemer J., for the Auger Collaboration, preprint (astro-ph/0807.4871).
797:
798: \bibitem[\protect\citeauthoryear{Blumenthal \& Gould}{1970}]{bg70}
799: Blumenthal G. R., Gould R. J., Rev. Mod. Phys., 42, 237, 1970.
800:
801: \bibitem[\protect\citeauthoryear{Casse et al.}{2000}]{clp02}
802: Casse F., Lemoine M., Pelletier G., 2002, Phys. Rev D, 65, 023002, 1.
803:
804: \bibitem[\protect\citeauthoryear{Cho et al.}{2003}]{clv01}
805: Cho J., Lazarian A., Vishniac E. T., 2003, ApJ., 595, 812.
806:
807: \bibitem[\protect\citeauthoryear{Cowsik \& Sarkar}{1984}]{cs84}
808: Cowsik R., Sarkar S., 1984, MNRAS, 207, 745.
809:
810: \bibitem[\protect\citeauthoryear{Crocker et al.}{2005}]{crocker05}
811: Crocker R. et {\it al.}, 2005, ApJ, 622, 892.
812:
813: \bibitem[\protect\citeauthoryear{Dedner et al.}{2002}]{d02}
814: Dedner A. et {\it al.}, 2002, J. Comp. Phys., 175, 645.
815:
816: \bibitem[\protect\citeauthoryear{Dermer \& Humi}{2001}]{dh01}
817: Dermer C.D., Humi M., 2001, ApJ, 556, 479.
818:
819: \bibitem[\protect\citeauthoryear{Fatuzzo \& Melia}{2003}]{fatuzzo2003}
820: Fatuzzo, M., Melia, F., 2003, ApJ, 596, 1035.
821:
822: \bibitem[\protect\citeauthoryear{Fermi}{1949}]{f49}
823: Fermi E., 1949, Phys. Rev., 75, 1169.
824:
825: \bibitem[\protect\citeauthoryear{Gallant et al.}{1999}]{gallant99}
826: Gallant Y. A., Achterberg A., Kirk J. G., 1999, A\&AS, 138, 549.
827:
828: \bibitem[\protect\citeauthoryear{Gelmini et al.}{2007}]{gks07}
829: Gelmini G., Kalashev O., Semikoz D. V., 2007, preprint (astro-ph/0702464).
830:
831: \bibitem[\protect\citeauthoryear{George et al.}{2008}]{gfbmt08}
832: George M. R., Fabian A. C. , Baumgartner W. H., Mushotzky R. F., Tueller J., 2008, MNRAS, 388, L59.
833:
834: \bibitem[\protect\citeauthoryear{Giacalone \& Jokipii}{1994}]{gj94}
835: Giacalone J., Jokipii J. R., 1994, ApJ, 430, L137.
836:
837: \bibitem[\protect\citeauthoryear{Giacalone \& Jokipii}{1999}]{GJ99}
838: Giacalone J., Jokipii J. R., 1999, ApJ, 520, 204.
839:
840: \bibitem[\protect\citeauthoryear{Greisen}{1966}]{greisen66}
841: Greisen, K., 1966, Phys. Rev. Lett., 16, 748.
842:
843: \bibitem[\protect\citeauthoryear{Hillas}{1984}]{h84}
844: Hillas A.M., 1984, Annual Review A\&A, 22, 425.
845:
846: \bibitem[\protect\citeauthoryear{Kaplan \& Tsytovich}{1973}]{kt73}
847: Kaplan S.A., Tsytovich V.N., 1973, Plasma Astrophysics, transl. by D. ter Haar, Pergamon Press.
848:
849: \bibitem[\protect\citeauthoryear{Landau \& Lifchitz}{1975}]{ll2}
850: Landau L., Lifchitz E., 1975, Theoretical Physics. Vol. 2: Field Theory, Pergamon Press.
851:
852: \bibitem[\protect\citeauthoryear{Leamon et al.}{1998}]{l98}
853: Leamon R. J., Smith C. W., Ness N. F., Matthaeus W. H., Wong H. K.,1998, J. Geophys. Res., 103, 4775.
854:
855: \bibitem[\protect\citeauthoryear{Lee \& Jokipii}{1976}]{lj76}
856: Lee L. C., Jokipii J. R., 1976, ApJ, 206, 735L.
857:
858: \bibitem[\protect\citeauthoryear{Liu et al.}{2006}]{liu06}
859: Liu S., Melia, F., Petrosian, V., 2006, ApJ, 636, 798.
860:
861: \bibitem[\protect\citeauthoryear{Liu et al.}{2004}]{liu04}
862: Liu S., Petrosian, V., Melia, F., 2004, ApJ Lett., 611, L101.
863:
864: \bibitem[\protect\citeauthoryear{Melia}{2001}]{melia2001}
865: Melia, F., 2001, Electrodynamics, University of Chicago Press.
866:
867: \bibitem[\protect\citeauthoryear{Melia}{2009}]{melia2009}
868: Melia, F., 2009, High-Energy Astrophysics, Princeton University Press.
869:
870: \bibitem[\protect\citeauthoryear{Nayakshin \& Melia}{1998}]{nayak1998}
871: Nayakshin, S., Melia, F., 1998, ApJS, 114, 269.
872:
873: \bibitem[\protect\citeauthoryear{Niemiec \& Ostrowski}{2006}]{no06}
874: Niemiec J., Ostrowski M., 2006, ApJ, 641, 984.
875:
876: \bibitem[\protect\citeauthoryear{Ostrowski}{2008}]{o08}
877: Ostrowski M., 2008, New Astron. Rev. in press, preprint (astro-ph/0801.1339).
878:
879: \bibitem[\protect\citeauthoryear{Padmanhaban}{2001}]{padma}
880: Padmanhaban, T. 2001, Theoretical Astrophysics Vol II: Stars and Stellar
881: Systems, Cambridge University Press.
882:
883: \bibitem[\protect\citeauthoryear{Press et al.}{1997}]{ptvf97}
884: Press W. H., Teukolsky S. A., Vetterling W. T., Flannery B.P., 1997, Numerical Recipes in Fortran 77: The Art of Scientific Computing, Vol. 2, second ed., Cambridge University Press.
885:
886: \bibitem[\protect\citeauthoryear{Rockefeller et al.}{2004}]{rock2004}
887: Rockefeller, G., Fryer, C. L., Melia, F., Warren, M. S., 2004, ApJ, 604, 662.
888:
889: \bibitem[\protect\citeauthoryear{Ruffert \& Melia}{1994}]{rm94}
890: Ruffert M., Melia F., 1994, A\&A, 288, L29.
891:
892: \bibitem[\protect\citeauthoryear{Ryu et al.}{1998}]{rmjf98}
893: Ryu D., Miniati F., Jones T.W., Frank A., 1998, ApJ, 509, 244.
894:
895: \bibitem[\protect\citeauthoryear{Semikoz et al.}{2007}]{semikoz07}
896: Semikoz, D. V., for the Auger Collaboration, preprint (astro-ph/0706.2960).
897:
898: \bibitem[\protect\citeauthoryear{Smith et al.}{1998}]{sw98}
899: Smith L.M., Woodruff S.L., 1998, Ann. Rev. Fluid Mech., 30, 275.
900:
901: \bibitem[\protect\citeauthoryear{Torres \& Anchordoqui}{2004}]{ta04}
902: Torres, D.F. \& Anchordoqui, L.A., 2004, Rep. Prog. Phys., 67, 1663.
903:
904: \bibitem[\protect\citeauthoryear{Unger et al.}{2007}]{u07}
905: Unger M., for the Auger Collaboration, preprint (astro-ph/0706.1495).
906:
907: \bibitem[\protect\citeauthoryear{Verma}{2004}]{v04}
908: Verma M.K., 2004, Phys. Rep., 401, 229.
909:
910: \bibitem[\protect\citeauthoryear{Wolfe \& Melia}{2006}]{wolfe2006}
911: Wolfe, B., Melia, F., 2006, ApJ, 638, 125.
912:
913: \bibitem[\protect\citeauthoryear{Wommer et al.}{2008}]{wommer2008}
914: Wommer, E., Melia, F., and Fatuzzo, M., 2008, MNRAS, 387, 987.
915:
916: \bibitem[\protect\citeauthoryear{Zatsepin \& Kuz'min}{1966}]{zatsepin66}
917: Zatsepin G. T., Kuz'min V. A. 1966, Soviet Journal of Experimental and Theoretical Physics Letters {\bf 4},
918: 78.
919:
920: \end{thebibliography}
921:
922:
923: \label{lastpage}
924:
925: \end{document}