1: \documentclass[12pt,preprint]{aastex}
2: %\documentclass[manuscript]{aastex}
3: %%\documentclass[A4, referee]{aa}
4: %\usepackage{psfig}
5: %\usepackage{rotating}
6: %\usepackage{txfonts}
7: %%\usepackage{graphicx}
8: %\usepackage{natbib}
9: %
10: \def\M21{\hbox{Mrk\,421} }
11: \def\xmm{\hbox{$XMM-Newton$}}
12: \def\etal{et al.}
13: \def\ie{i.e., }
14: \def\me{m_{\rm e}}
15: \def\flux{erg\,cm$^{-2}\,s^{-1}$}
16: \def\mlum{$erg\,s^{-1}\,Hz^{-1}$}
17: \def\ltsima{$\; \buildrel < \over \sim \;$}
18: \def\simlt{\lower.5ex\hbox{\ltsima}} % < over ~
19: %
20: \def\gtsima{$\; \buildrel > \over \sim \;$}
21: \def\simgt{\lower.5ex\hbox{\gtsima}} % > over ~
22: %
23: \newcommand{\eg}{e.g.,\ }
24: %
25: \newcommand{\eqref}[1]{(\ref{#1})}
26: \newcommand{\dsfrac}[2]{\displaystyle{\frac{#1}{#2}}} \def\difd{{\rm
27: d}} \def\zb{z_{\rm b}} \def\zf{z_{\rm f}} \def\yb{y_{\rm b}}
28: %
29: \newcommand{\SF}[2]{{\bf S}#1-{\bf F}#2}
30: \def\pmag{p_{\rm mag}}
31: \def\epsth{\epsilon_{\rm th}}
32: \def\epsmag{\epsilon_{\rm mag}}
33: \def\gad{\gamma_{\rm ad}}
34: \def\BP{B^{\prime}}
35: %
36: \newcommand{\BG}[1]{{\bf B}#1}
37: \newcommand{\TG}[1]{{\bf T}#1}
38:
39:
40:
41: \shorttitle{Annomalous shear layers in relativistic jets}
42: \shortauthors{Aloy \& Mimica}
43:
44:
45: \begin{document}
46:
47: \title{Observational Effects of Anomalous Boundary Layers in
48: Relativistic Jets}
49:
50: \author{M.A.\,Aloy\altaffilmark{1} %, J.\,Ferrand\altaffilmark{1}
51: and P.\,Mimica\altaffilmark{1}}
52: %
53: %\offprints{MAA, e-mail: Miguel.A.Aloy@uv.es}
54: \affil{Departamento de Astronom\'{\i}a y Astrof\'{\i}sica,
55: Universidad de Valencia, 46100 Burjassot, Spain}
56: \email{MAA, e-mail: Miguel.A.Aloy@uv.es}
57: %
58: \begin{abstract}
59: %
60: Recent theoretical work has pointed out that the transition layer
61: between a jet an the medium surrounding it may be more complex than
62: previously thought. Under physically realizable conditions, the
63: transverse profile of the Lorentz factor in the boundary layer can
64: be non-monotonic, displaying the absolute maximum where the flow is
65: faster than at the jet spine, followed by an steep fall
66: off. Likewise, the rest-mass density, reaches an absolute minimum
67: (coincident with the maximum in Lorentz factor) and then grows until
68: it reaches the external medium value. Such a behaviour is in
69: contrast to the standard monotonic decline of the Lorentz factor
70: (from a maximum value at the jet central spine) and the
71: corresponding increase of the rest-mass density (from the minimum
72: reached at the jet core). We study the emission properties of the
73: aforementioned anomalous shear layer structures in kiloparsec-scale
74: jets aiming to show observable differences with respect to
75: conventional monotonic and smooth boundary layers.
76: %
77: \end{abstract}
78: %
79: \keywords{galaxies: jets: general; X-rays: general ---
80: radiation mechanisms: non-thermal; acceleration of
81: particles --- methods: numerical; MHD}
82:
83: %\titlerunning{Annomalous shear layers}
84: %\authorrunning{Ferrand, Mimica \& Aloy}
85: %\maketitle
86:
87:
88: %--------------------------------------------------------
89: \section{Introduction}
90: \label{sec:intro}
91: %--------------------------------------------------------
92:
93: The interaction of extragalactic jets with their environment leads,
94: under rather general circumstances, to the stratification of the
95: beam of the jet in the direction normal to its velocity. The
96: morphology of FR I sources (e.g., M87, \citealp{owe89}; 3C 31,
97: \citealp{lai96}; 3C 296, \citealp{hac97}) has been explained in
98: terms of a jet whose dynamics is dominated by a boundary shear
99: layer. In such layers the emissivity in radio and optical maps peaks
100: and, in some cases (e.g., Mrk 501; \citealp{caw93}), the spectrum is
101: rather flat, suggesting that the acceleration of non-thermal
102: particles takes place right at the boundary region. Further
103: evidence in favor of such stratification is provided by radio
104: polarization measurements, which indicate that, towards the jet
105: boundary, the magnetic field is highly ordered and parallel to the
106: jet axis \citep{per99}.
107:
108: A radial stratification of jets in FR~II radio galaxies has also
109: been observed \citep{SBB98,caw93}. The edge brightening found in
110: 3C353 has been interpreted by \cite{SBB98} as due to the Doppler
111: hiding of the emission of the central spine of the jet,
112: suggesting that most of the observed radiation comes from the jet
113: boundary layer. On the other hand, the {\it rails of low
114: polarization} found by \cite{SBB98} close the the jet boundary
115: indicate that the magnetic field in the layer is either axial or
116: toroidal but not radial. Even at parsec scale, the FR~II radio jet
117: of 1055+018 exhibits a transverse structure consistent with a {\it
118: spine - shear boundary layer} jet morphology \citep{ARW99}.
119:
120: Also sources which are in the borderline dividing FR~I and FR~II
121: sources show evidences of radial jet stratification. A couple of
122: examples of such sources are 3C~15 and Cen~A. In the first case the
123: jet is generally narrower in the optical than in the radio
124: \citep{Dulwichetal07}, and simple spine-sheath model may account for
125: the polarization angle differences seen in the optical and radio
126: data. For Cen~A, the recent X-ray data \citep{Kataokaetal06} can
127: also be properly explained by a stratified jet model with a radially
128: decreasing velocity field.
129:
130: The aforementioned observational evidence along with some others
131: (e.g., \citealp{chi00}), suggests that the velocity in some jet
132: boundary layers is smaller (but still relativistic) than the
133: velocity of the beam itself. On the other hand, laboratory
134: experiments show that, almost unavoidably, turbulent interaction
135: layers may develop as jets propagate into a viscous medium (e.g.,
136: \citealt{BR74}). These laboratory shear layers display a radial
137: velocity profile roughly monotonic in which the velocity of the jet
138: core smoothly decreases until it vanishes at the external medium.
139:
140: From a theoretical point of view, the Reynolds number of the jet
141: flow may reach values of $\sim 10^{23}$ and, hence, it seems
142: unquestionable that such flows will quickly develop a turbulent
143: boundary layer which spreads into the flow and leads to the
144: entrainment and acceleration of ambient gas
145: \citep{DeYoung93}. However, the nature and the amount of viscosity
146: in relativistic jets is still largely unknown. Most probably, the
147: effective viscosity in the lateral interaction between the jet an
148: the external medium is a mixture of the turbulent eddy viscosity,
149: magnetic viscosity (based on the finite size of the Larmor radius;
150: see, e.g., \citealp{Baan80}) and {\it cosmic-ray} viscosity (i.e.,
151: the viscosity originated due to the diffusion of the momentum
152: carried by energetic particles; see, e.g., \citealp{EJM88}).
153: Whichever the origin of the viscosity is, the boundary layer itself
154: results from the nonlinear growth of Kelvin-Helmholtz (KH)
155: instabilities on the jet surface. The fact that relativistic jets
156: are prone to these instabilities was demonstrated theoretically in
157: the pioneer work of \cite{TS76}. Later, many other papers have
158: explored the stability of relativistic jets with respect to KH
159: instabilities in the vortex-sheet approximation (e.g.,
160: \citealt{FTZ78,Hardee79}). \cite{FTZ82} and \citet{bir91} presented
161: the first attempts to study (in the linear regime) the influence of
162: a finite thickness shear layer in classical and relativistic jets,
163: respectively. Very recently, \cite{Hardee07} has also considered the
164: stability properties (in the linear regime) of spine-sheath
165: magnetized, relativistic jets in which the magnetic field is
166: parallel to the flow velocity. The development of KH modes in
167: sheared relativistic (slab) flows from the linear (analytic) regime
168: to the non-linear regime by means of relativistic hydrodynamic
169: simulations has been carried forward by
170: \cite{Peruchoetal07}. Alternative approaches to the linear analysis
171: of the stability of stratified relativistic jets and sheared
172: relativistic jets have been performed by \cite{HS96}, \cite{Urpin02}
173: \cite{alo02} and \cite{MK07}, respectively.
174:
175: Further support in favor of radial jet stratification as a
176: consequence of the interaction between a jet and its environment is
177: provided by three-dimensional hydrodynamic simulations of
178: relativistic, large scale jets \citep{alo99a,alo99b,alo99c,alo00}
179: and two-dimensional simulations of relativistic magnetized jets
180: \citep{LA05}. These simulations show that radially stratified jets
181: are developed out of initially uniform beams. These numerical works
182: display also a rather smooth transition layer with decreasing
183: transverse velocity towards the jet boundary and high specific
184: internal energy. It must be noted that the viscosity in such
185: simulations is of numerical origin, and mimics only qualitatively
186: actual viscous flows.
187:
188: In addition to the existence of {\it environmental} reasons to
189: produce a radial flow stratification in astrophysical jets, this
190: effect can also be a natural consequence of the jet formation
191: process. Jet launching from accretion disks may directly lead to a
192: certain mass flux profile and/or magnetic flux profile within the
193: jet. This possibility has been verified by means of axisymmetric MHD
194: simulations both in the classical \citep{PCO06,Fendt06}, as well as
195: in the general relativistic \citep{MN07} regime.
196:
197:
198: \cite{AR06} show that, under physically plausible conditions (see
199: Sect.~\ref{sec:RMHD_boost}), the lateral structure of the
200: interaction layer of a relativistic flow can be richer than
201: previously thought. This is also the case in relativistic magnetized
202: jets \citep{Mizuetal08}. Due to a purely relativistic phenomenon
203: (the conversion of specific enthalpy into bulk Lorentz factor across
204: a flow discontinuity), the jet velocity may increase right at the
205: contact discontinuity (CD) separating the jet and the external
206: medium. Beyond the CD the velocity decreases steeply to zero in the
207: radial direction. The growth of the velocity or, equivalently, of
208: the Lorentz factor of the beam is associated to the development of a
209: rarefaction wave that emerges form the CD where the density and the
210: pressure decrease with respect to their corresponding values at the
211: jet spine. The profiles exhibited by the hydrodynamic variables in
212: these layers are non-monotonic, in contrast with the smooth
213: monotonic profiles that have been typically discussed in the
214: literature. Thereby, we will refer to them as {\it anomalous} or
215: {\it AR} shear layers.
216:
217: As mentioned above, jet shear layers, represent natural sites for
218: particle acceleration, providing high-energy cosmic rays and
219: influencing the dynamics of relativistic jets in extragalactic radio
220: sources by forming cosmic-ray cocoons \citep{ost00}. The efficiency
221: of the acceleration process in these turbulent shear layers depends
222: on the particle mean free path and on the velocity
223: structure. \citeauthor{SO02} (2002; SO02 hereafter) performed a
224: through study of both the acceleration processes acting at jet
225: boundaries and the resulting observational effects. They former
226: authors considered the possibility that turbulent standard boundary
227: layers of kiloparsec scale jets may substantially contribute to the
228: radiative output of the jet. In the present work we will closely
229: follow SO02 working hypothesis, however we will replace the
230: monotonic shear layer kinematic structure they assumed by an
231: anomalous one as suggested by \cite{AR06}. The aim being to show
232: whether AR-shear layers imprint any distinctive feature in the
233: radiation produced by the jet as compared with standard boundary
234: layers. We show in this paper that, indeed, the radiative output of
235: jets bound by anomalous shear layers significantly differs from that
236: of ordinary boundary layers. Thus observations may confirm or rule
237: out the existence of such anomalous shear layers in kiloparsec scale
238: jets and, consistently, this may be used to constrain hardly known
239: physical parameters in these regions.
240:
241: The plan of the paper is to first show [\S~\ref{sec:RMHD_boost}] the
242: influence of magnetic fields in the profiles of the physical
243: variables in anomalous boundary layers. In this way we extend the
244: results of \cite{AR06} and \cite{Mizuetal08} to the adequate
245: parameter range. In Sect.~\ref{sec:model} we summarize the basic
246: model developed by SO02 and adapt it to account for the kinematic
247: differences when applied to anomalous shear layers whose profiles
248: will be discussed in
249: Sect.~\ref{sec:RMHD_boost}. Section~\ref{sec:SED} provides the
250: spectral energy distribution obtained for the different models of
251: kinematic shear layers shown in the previous section. We discuss our
252: results in Sect.~\ref{sec:discussion} and sum them up in
253: Sect.~\ref{sec:summary}.
254:
255:
256:
257: %--------------------------------------------------------
258: \section{Magnetohydrodynamic boosting in relativistic jet boundary
259: layers}
260: \label{sec:RMHD_boost}
261: %--------------------------------------------------------
262:
263: The dynamics of a relativistic magnetohydrodynamic jet in an external
264: medium can be treated as the motion of two fluids, one of which (the
265: jet) is much hotter and at higher (or equal) pressure than the other
266: one (the external medium), and is moving with a large tangential
267: velocity with respect to the cold, slowly moving fluid. Furthermore,
268: the jet can be magnetized while the external medium magnetic field is
269: much smaller than that of the jet. \cite{AR06} found that if the
270: specific enthalpy of the jet is sufficiently large, a net conversion
271: of internal to kinetic energy can be produced through a purely
272: relativistic channel. The reason for its exclusively relativistic
273: character is that, the evolution of contact or tangential
274: discontinuities (like, e.g., the one that separates laterally the jet
275: from the external medium), is governed by a genuine coupling between
276: the specific enthalpy and the Lorentz factor at both sides of the
277: discontinuity which does not exist in Newtonian
278: (magneto-)hydrodynamics. Such mechanism yields a substantial
279: hydrodynamic boost along the boundary layer between the jet and the
280: external medium. Basically, the boost along the jet lateral boundaries
281: is due to the work done by the external medium on an overpressured,
282: relativistic jet. \cite{Mizuetal08} extended the validity of the
283: results of \cite{AR06} to the case in which the jet is magnetized and
284: conclude that the presence of a jet magnetic field (either poloidal or
285: toroidal) enlarges the boost that might be obtained by purely
286: hydrodynamic means. The magnetohydrodynamic boost is larger for jets
287: carrying toroidal than poloidal magnetic fields.
288:
289: The present analysis is aimed to show what is the qualitative
290: variation of the Lorentz factor and of the magnetic field across a
291: jet boundary layer. Therefore, following \cite{AR06}, we model the
292: interaction between the jet and the external medium as a one
293: dimensional Riemann problem in Cartesian coordinates. Furthermore, we
294: restrict to the case in which the magnetic field is perpendicular to
295: the flow speed in the jet. As demonstrated by \cite{AR06} and
296: \cite{Mizuetal08} this simple model retains the basic features of the
297: phenomenon. Certainly, a more detailed model would require three
298: dimensional relativistic magnetohydrodynamic simulations.
299:
300: We solve exactly the Riemann problem set by the jump in the conditions
301: between the jet and the external medium assuming that the magnetic
302: field is perpendicular to the velocity field. For this purpose we use
303: the method devised by \cite{Romeroetal05}. The two uniform ``left''
304: (jet) and ``right'' (external medium) states which set the proposed
305: Riemann problems possess different and discontinuous
306: magnetohydrodynamical properties: the rest-mass density $\rho$, the
307: total pressure $p=p_{\rm mag}+p_{\rm gas}$ ($p_{\rm mag}$ and $p_{\rm
308: gas}$ being the magnetic and the gas pressure, respectively), the
309: components of the velocity normal $v^{\rm n}$ and tangential $v^{\rm
310: t}$ to the initial discontinuity, and the magnetization ${\tilde
311: \beta}:=p_{\rm mag}/p_{\rm gas}$. In the following, we will use for
312: the quantities in the jet and in the external medium the subscripts
313: ``L'' and ``R'', respectively.
314:
315: Like in the purely hydrodynamic case, depending on the conditions of
316: the left and right states two qualitatively and quantitatively
317: different solutions develop (Fig.~\ref{fig:sl_structure_magnetic}). On
318: the one hand, a pattern of waves can be formed, which we indicate as
319: ${\cal _{\leftarrow}\!S C S_{\!\rightarrow}}$ (Fig.\@
320: \ref{fig:sl_structure_magnetic} left panel), where ${\cal
321: _{\leftarrow}\!S}$ refers to the shock propagating towards the left
322: sweeping up the jet, ${\cal S_{\!\rightarrow}}$ to the shock moving
323: towards the right crossing the external medium and ${\cal C}$ to the
324: contact discontinuity between the two. On the other hand, the pattern
325: can change for sufficiently large velocities parallel to the jet axis
326: (tangential) and in this case the inward-moving shock is replaced by a
327: rarefaction wave thus producing a ${\cal _{\leftarrow}\!R C
328: S_{\!\rightarrow}}$ pattern (Fig.~\ref{fig:sl_structure_magnetic}
329: right panel). Once again, we point out that the change in the wave
330: pattern does only happen in relativistic (magneto-)hydrodynamics. Only
331: in the relativistic regime the evolution of a Riemann problem depends
332: upon the components of the velocity parallel to the initial
333: discontinuity or, equivalently, upon the Lorentz factor at both sides
334: of the initial discontinuity. Furthermore, only in the relativistic
335: regime there is a coupling between the momentum and energy fluxes that
336: depends upon the specific enthalpy of initial left and right states.
337:
338: We notice that the variation of the magnetization in the intermediate
339: left state is less than a few percent when a ${\cal _{\leftarrow}\!R C
340: S_{\!\rightarrow}}$ solution forms (see
341: Fig.~\ref{fig:sl_structure_magnetic} right panel). In case a ${\cal
342: _{\leftarrow}\!S C S_{\!\rightarrow}}$ pattern develops, the
343: magnetization in the shocked left intermediate state ${\tilde \beta}^*_{_{\rm
344: L}}$ grows due to the shock compression, but, in any case,
345: ${\tilde \beta}^*_{_{\rm L}} \simlt (2 - 3)\times{\tilde \beta}_{_{\rm L}}$.
346:
347:
348: Although \cite{Mizuetal08} have already explored the influence of a
349: dynamically relevant magnetic field in the jet on the aforementioned
350: boost, they restricted their study to only four cases and to extremely
351: hot jets ($p_{_{\rm gas,L}}/\rho_{_{\rm L}} \ge 10^5$), which is
352: appropriate for gamma-ray burst jets (see, e.g.,
353: \citealt{alo00b,Alo05,AO07,Birkletal07}). Here we consider a parameter
354: space more adequate for kiloparsec scale jets. More precisely, we
355: consider {\it warm} or {\it cold} jets with a ratio $p_{\rm
356: gas,L}/\rho_{_{\rm L}} \le 10^2$ and a magnetization ${\tilde
357: \beta}_{_{\rm L}}$, which is varied between $0.1$ (slightly
358: magnetized jet) and $10^4$ (Poynting flux dominated jet)\footnote{In
359: \cite{Mizuetal08} a maximum value of ${\tilde \beta}_{_{\rm L}}=1.8$
360: is considered in their model MHDB.}. The left state will have
361: comparable but larger total pressure than the external medium
362: $p_{_{_{\rm L}}} = 10 p_{_{\rm R}}$, will be under dense $\rho_{_{\rm
363: L}} / \rho_{_{\rm R}} = 10^{-5}$, with a fixed bulk Lorentz factor
364: $\Gamma_{_{\rm L}} \equiv [1 - (v^t_{_{\rm L}})^2 - (v^n_{_{\rm
365: L}})^2]^{-1/2}=10$. We choose to fix the external medium (right
366: state) to be cold ($p_{_{\rm R}}/\rho_{_{\rm R}} \simeq 10^{-4}$),
367: non-magnetized and non-moving. Precisely, we take for the external
368: medium $p_{_{\rm R}}=10^{-6}/\rho_{_{\rm ext}} c^2$, $\rho_{_{\rm R}}
369: = 10^{-2}\rho_{_{\rm ext}}$, $v^n_{_{\rm R}} = v^t_{_{\rm R}} =
370: {\tilde \beta}_{_{\rm L}} = 0$, where $\rho_{_{\rm ext}}$ is a
371: normalization constant which allows us to be scale-free. We point out
372: that with this parametrization we provide physically plausible
373: conditions for both the jet and the external medium. For convenience,
374: hereafter we will assume $c = \rho_{_{\rm ext}} = 1$, unless stated
375: otherwise.
376:
377: We explore the resulting Riemann solution as we vary the component of
378: the jet velocity normal to the shear layer\footnote{For a relativistic
379: jet it is expected that such a component is $v^n_{_{\rm L}} \simeq 0
380: \ll v^t_{_{\rm L}} \sim 1$.} and the jet magnetization ${\tilde
381: \beta}_{_{\rm L}}$. We point out that models with ${\tilde \beta}
382: \ge 100$ are actually rather cold, since $p_{\rm gas,L}/\rho_{_{\rm
383: L}} \ll 1$ (e.g., for ${\tilde \beta} \ge 10^4$, $p_{\rm
384: gas,L}/\rho_{_{\rm L}} \simeq 10^{-2}$). According to the most
385: broadly accepted view, even MHD-generated jets may become matter
386: dominated at kiloparsec scale (e.g., \citealp{BL94,FO04}). However,
387: there are others who advocate jet models which are basically
388: electromagnetic entities at such large scales (e.g.,
389: \citealp{Blandford02,Blandford03}). In view of these two possible
390: extremes, we have covered the ${\tilde \beta}$-parameter space with
391: jet models which are both matter dominated and Poyinting-flux
392: dominated at kiloparsec scale. In Fig.~\ref{fig:betal_dep} we show the
393: value of the Lorentz factor reached at the state left to the contact
394: discontinuity $\Gamma^*_{_{\rm L}}$. This intermediate state is made
395: out of boosted or deboosted jet matter depending respectively on
396: whether a ${\cal _{\leftarrow}\!R C S_{\!\rightarrow}}$ or a ${\cal
397: _{\leftarrow}\!S C S_{\!\rightarrow}}$ solution develops.
398:
399: Figure~\ref{fig:betal_dep} illustrates how a magnetohydrodynamic
400: boost, where $\Gamma^*_{_{\rm L}} > \Gamma_{_{\rm L}}$, exists below
401: a critical value of the normal velocity ($v^n_{_{\rm L}}\simeq 0.026$
402: in the case considered here) whereby a ${\cal _{\leftarrow}\!R C
403: S_{\!\rightarrow}}$ solution develops. The critical value of
404: $v^n_{_{\rm L}}$ is independent of the jet magnetization (note the
405: crossing of all the solutions at the point where $\Gamma^*_{_{\rm L}}
406: = \Gamma_{_{\rm L}}=10$).
407:
408: The magnetohydrodynamic boost is larger if the magnetization of the
409: jet is increased: while for a poorly magnetized jet (${\tilde
410: \beta}_{_{\rm L}}=0.1$) the boost may increase the bulk Lorentz
411: factor of the layer by $\sim 10\% - 50\%$, for a strongly magnetized
412: jet (${\tilde \beta}_{_{\rm L}}\ge 10^2$) the Lorentz factor increase
413: can be $\ge 100\%$. Increasing the magnetization beyond ${\tilde
414: \beta}_{_{\rm L}}\ge 10^4$ does not produce larger boosts (the
415: effect saturates). Below ${\tilde \beta}_{_{\rm L}} < 0.1$ the results
416: are almost indistinguishable from the case ${\tilde \beta}_{_{\rm
417: L}}=0.1$ and we have decided not to include more lines in the
418: Fig.~\ref{fig:betal_dep} for the sake of readability.
419:
420: The bottom line of this parametric study is that for conditions
421: realizable in kiloparsec scale jets, one may find an increase of the
422: Lorentz factor in the transition layer between the jet and the
423: external medium ranging from $(1.5 - 2) \times\Gamma_{_{\rm L}}$,
424: while the magnetization in that layer is basically the same as in the
425: jet if a ${\cal _{\leftarrow}\!R C S_{\!\rightarrow}}$ pattern
426: occurs. Furthermore, since only the strength of the
427: magnetohydrodynamic boost, but not its existence, depends on the
428: topology of the field (see \citealt{Mizuetal08}), we point out that it
429: also happens in the presence of randomly oriented magnetic fields.
430:
431:
432: %--------------------------------------------------------
433: \section{The physical model}
434: \label{sec:model}
435: %--------------------------------------------------------
436:
437: Our model follows very closely that devised by SO02. We assume that
438: non-thermal particles can be accelerated at the boundary layers of
439: relativistic jets (e.g., \citealp{ost90,ost98,ost00,RD04}), thereby
440: producing high-energy cosmic rays. As mentioned in
441: Sect.~\ref{sec:intro}, the main goal of this paper is to replace the
442: prototype shear layer assumed in SO02 by an anomalous shear layer of
443: the type suggested by \cite{AR06} and whose basic features have been
444: outlined in Sect.~\ref{sec:RMHD_boost}. At the most basic level, this
445: means to replace the kinematic shear layer structure assumed by SO02,
446: namely, a jet sheath where the bulk Lorentz factor ($\Gamma$)
447: decreases monotonically (Fig.~\ref{fig:sl_structure}a), by a boundary
448: layer where the Lorentz factor develops a spike in which it is larger
449: than at the jet core (Fig.~\ref{fig:sl_structure}b). We shall assume
450: that the magnetic field is uniform in the whole shear layer, which is
451: compatible with the results of our previous section and a reasonable
452: assumption in our simple model.
453:
454: The working hypothesis made by SO02 are also valid in our case since,
455: the only difference between monotonic and anomalous layers is simply
456: kinematic (see Sect.~\ref{sec:kinematic}). Thus, following SO02, the
457: maximum electron Lorentz factor which can be obtained by the
458: combination of acceleration of particles along the shear layer and the
459: radiative cooling of such particles is
460: \begin{equation}
461: \gamma_{\rm eq} \approx 4 \cdot 10^{8} \, V_{8} \, B_{\mu G}^{-1 / 2} ,
462: \label{eq:gamma_eq}
463: \end{equation}
464: where $B_{\mu G}$ is the magnetic field in microgauss and $V_8\equiv
465: V_{\rm A}/10^8\,$cm\,s$^{-1}$ is the Alfv\'en speed $V_{\rm A}$ in
466: units of $10^8\,$cm\,s$^{-1}$. For kiloparsec scale jets, typical
467: values of these two parameters are $V_8\simeq 1$ and $B_{\mu G}\sim
468: 10$, which yields $\gamma_{\rm eq} \simeq 10^8$.
469:
470: We assume that the resulting spectral energy distribution of the
471: electrons accelerated at the layer consists of a low energy power law
472: $n_{\rm e}(\gamma) \propto \gamma^{-\sigma}$, with $\sigma = 2$,
473: finished with a high-energy component modeled as a nearly mono
474: energetic peak at $\gamma = \gamma_{eq}$ due to the pile-up of
475: accelerated particles caused by losses (see Fig.~1 of SO02). In terms
476: of the Dirac delta function, $\delta$, at energies above the injection
477: energy, $\gamma > \gamma_0$, the considered quasy-stationary, averaged
478: spectrum of electrons can be cast in the form
479: \begin{equation}
480: n_{\rm e}( \gamma) = a \, \gamma^{- \sigma} \, \exp( - \gamma / \gamma_{\rm eq}) + b \, \delta( \gamma - \gamma_{\rm eq}) ,
481: \label{eq:n(gamma)}
482: \end{equation}
483: where $a$ and $b$ are normalization parameters whose values ($a\simeq
484: 10^{-7}\,$cm$^{-3}$, $b \simeq 10^{-14}\,$cm$^{-3}$) result from
485: assuming equipartition between the magnetic field energy density
486: $u_{_{\rm B}} \equiv B^2We / 8 \, \pi$ and the energy densities of both
487: electron spectral components:
488: %
489: \begin{equation}
490: \int_{\gamma_0}^\infty (\gamma \, mc^2) \, a \, \gamma^{-2} \, \exp(
491: - \gamma / \gamma_{\rm eq}) \, d \gamma = \int_{\gamma_0}^\infty (\gamma
492: \, mc^2) \, b \, \delta( \gamma - \gamma_{\rm eq}) \, d \gamma = {1 \over
493: 2} \, u_{_{\rm B}} .
494: \label{eq:normalization}
495: \end{equation}
496:
497: We point out that in SO02, the power-law part of the electron spectrum
498: was also endowed with an exponential cut-off instead of with a
499: Heaviside $\Theta$ function in spite of their Eqs.~6 and 7. The
500: $\Theta$ function was used in their work only to compute analytic
501: estimates, not for the numerical integrations necessary to calculate
502: their spectral energy distributions \citep{Stawarz}.
503:
504: Assuming equipartition between the particle and magnetic field energy
505: in relativistic sheared jets has been proved to be a rather solid
506: theoretical assumption (see, e.g., \citealt{Urpin06}). However, in
507: some cases the values of $a$ and $b$ derived from
508: Eq.~\ref{eq:normalization} and based on equipartition arguments may
509: yield only lower bounds of the true values. \cite{KS05} argue that
510: powerful jets in quasars and FR II objects can be far from the minimum
511: energy condition, and the field strength very likely exceeds the
512: equipartition value. On the other hand, upper limits to the inverse
513: Compton radiation of the jet in M\,87 imposed by HESS and HEGRA
514: Cerenkov Telescopes also indicate that the magnetic field cannot be
515: weaker than the equipartition value \citep{Stawarzetal05}.
516:
517: %--------------------------------------------------------
518: \subsection{Radiative processes in boundary layers of kpc-scale jets}
519: \label{sec:radiation}
520: %--------------------------------------------------------
521: %
522: % Synchrotron radiation
523: %
524:
525: The relevant radiative processes taking place at the boundary layer of
526: a relativistic, kpc-scale jet are, on the one hand, the synchrotron
527: (syn) emission of the spectral family represented by
528: Eq.~\ref{eq:n(gamma)} and, on the other hand, their inverse Compton
529: (IC) cooling due to the interaction with the previously produced
530: synchrotron photons (synchrotron self Compton -SSC-) or with seed
531: photons of the cosmic microwave background (external Compton
532: -EC-). These radiative processes are the dominant for large scale
533: relativistic ($\Gamma>2$) jets at distances from the galactic nucleus
534: $z> 10\,$kpc.
535:
536: Following SO02, we assume that the magnetic field is randomly oriented
537: in the shear layer\footnote{Note that anomalous layers can also form
538: if the magnetic field is randomly oriented, see
539: Sect.~\ref{sec:RMHD_boost}.}. Additionally, we neglect synchrotron
540: self-absorption and, therefore, we limit the computation of the
541: resulting spectra to frequencies $\nu\geq 10^{10}\,$Hz.
542:
543: Restricted to the case of an isotropic electron distribution
544: $n_{\rm e}(\gamma)$ the synchrotron emissivity averaged over a randomly
545: oriented magnetic field $B$ can be computed from
546: %
547: \begin{equation}
548: j_{\rm syn}(\nu) = { \sqrt{3} e^{3} B \over mc^{2}} \int R \left( {\nu
549: \over c_{1} \gamma^{2}} \right) \, {n_{e}(\gamma) \over 4 \pi}
550: \, d\gamma ,
551: \label{eq:j_syn}
552: \end{equation}
553: %
554: where $c_1 = 3 e B \, / \, 4 \pi m c$ and $R(x)$ is a combination of
555: the modified second order Bessel functions \citep{cru86} as
556: %
557: \begin{equation}
558: R(x) = {x^2 \over 2} \, K_{4/3} \left({x \over 2} \right) \, K_{1/3}
559: \left({x \over 2} \right) - 0.3 \, {x^3 \over 2} \left[ K_{4/3}^2
560: \left({x \over 2} \right) - K_{1/3}^2 \left({x \over 2} \right)
561: \right] .
562: \label{eq:R(x)}
563: \end{equation}
564:
565: %
566: % IC (SSC) radiation
567: %
568: The IC photon emissivity in the source frame%
569: %
570: \footnote{In our case the source frame is a cylindrical layer of the
571: boundary region where $\Gamma$ is assumed to be uniform.},
572: $\dot{n}_{\rm ic} (\epsilon, \Omega)$, at a certain photon energy
573: $\epsilon$ (in units of the electron rest mass, $m_{\rm e}c^2$,
574: $\epsilon \equiv h \nu / m_{\rm e}c^2$) and scattering direction
575: $\Omega$, is given by
576: %
577: \begin{equation}
578: \dot{n}_{\rm ic}(\epsilon,\Omega) = c \int d\epsilon_i \oint d\Omega_i
579: \int d\gamma \oint d\Omega_{\rm e} \, (1- \beta \, \cos \psi) \, \sigma \,
580: n_i(\epsilon_i,\Omega_i) \, n_{\rm e}(\gamma,\Omega_{\rm e}) ,
581: \label{eq:n_ic}
582: \end{equation}
583: %
584: where $n_i (\epsilon_i, \Omega_i)$ is the seed photon number density
585: as a function of the incident photon energy $\epsilon_i$ and the
586: photon direction $\Omega_i$,
587: %(where the subscripts $i, ic$ are replaced
588: %by $i=syn$, $ic = ssc$ for SSC emission, and $i=cmb$, $ic = ec$, for
589: %EC emission)
590: $n_{\rm e} (\epsilon_{\rm e}, \Omega_{\rm e})$ is the electron energy
591: distribution, $\psi$ is the angle between the electron and the
592: incident photon directions, $\gamma=(1-\beta)^{1/2}$ is the electron
593: Lorentz factor corresponding to a velocity $v_{\rm e}=\beta c$ and
594: $\Omega_{\rm e}=(\cos^{-1}{\mu_{\rm e}}, \phi_{\rm e})$ its direction
595: in the plasma source frame (e.g., \citealt{der95}). The rest-frame
596: emissivity in CGS units is $j_{\rm ic}(\nu,\Omega) = h \, \epsilon \,
597: \dot{n}_{\rm ic} (\epsilon, \Omega)$.
598:
599: In the calculation of the emissivity due to IC process (SSC or EC), we
600: limit ourselves to the Thompson regime, in which the scattering cross
601: section $\sigma \equiv \sigma_{_{\rm T}}=6.65\times 10^{-25}\,$cm$^2$ is
602: independent of the seed photon energy, which is an adequate
603: approximation in the regime $\gamma \epsilon_i < 1$.
604:
605: Following \cite{der95}, the scattered photons will be beamed along the
606: direction of the electron motion (i.e. $\Omega\approx\Omega_{\rm e}$)
607: and will emerge after scattering with an average final energy
608: $\epsilon\approx\frac{4}{3}\gamma^2\epsilon_{\rm syn}$. We assume as
609: SO02 that the synchrotron radiation and the electron distribution are
610: isotropic and, therefore, $n_i(\epsilon_i,\Omega_i) = n_{\rm
611: syn}(\epsilon_{\rm syn})/4\pi$ and $n_{\rm e}(\gamma,\Omega_{\rm e})
612: = n_{\rm e}(\gamma)/4\pi$. Then, in Eq.~\ref{eq:n_ic} one can use
613: $\sigma = \sigma_{_{\rm T}} \delta(\Omega-\Omega_{\rm e}) \delta
614: \left(\epsilon-(4/3)\gamma^2\epsilon_{\rm syn}\right)$ in order to
615: obtain the following expression of $\dot{n}_{_{\rm
616: SSC}}(\epsilon,\Omega)$:
617: %
618: \begin{equation}
619: \dot{n}_{_{\rm SSC}}(\epsilon,\Omega) =
620: \frac{c\sigma_{_{\rm T}}}{4\pi}\int_{0}^{\infty} {d\epsilon_{\rm
621: syn}}\int_{\gamma_0}^{\infty} {d\gamma}n_{\rm syn}(\epsilon_{\rm
622: syn}) n_{\rm
623: e}(\gamma)\delta\left(\epsilon-\frac{4}{3}\gamma^2\epsilon_{\rm syn}\right).
624: \label{eq:n_ssc}
625: \end{equation}
626: %
627: Using Eqs.~\ref{eq:n(gamma)},~\ref{eq:n_ssc}, the relation $j_{_{\rm
628: SSC}}(\nu,\Omega) = h\epsilon \dot{n}_{_{\rm
629: SSC}}(\epsilon,\Omega)$ and the optically thin synchrotron
630: intensity $I_{\rm syn}(\nu_{\rm syn})\equiv j_{\rm
631: syn}l=(ch/4\pi)\epsilon_{\rm syn}n_{\rm syn}(\epsilon_{\rm syn})$,
632: where $l$ is the emitting region linear size, the integral form of the
633: synchrotron self-Compton emissivity $j_{_{\rm SSC}}(\nu)$ is written
634: as
635: %
636: \begin{equation}
637: j_{_{\rm SSC}}(\nu)=\sigma_{_{\rm T}} \int_{\gamma_{_{\rm low}}}^{\infty} {j_{\rm syn}\left(\frac{3\nu}{4\gamma^2}\right)l\frac{4\gamma^2}{3}n_{\rm e}(\gamma)d\gamma},
638: \label{eq:j_ssc}
639: \end{equation}
640: \noindent
641: where $\gamma_{_{\rm low}}={\rm max}(\gamma_0, \frac{3}{4}\epsilon)$ is
642: the lower limit imposed by the restriction $\gamma \epsilon_{\rm
643: syn}<1$, which guaranties that we stay within the Thompson regime.
644:
645: %
646: % CMB radiation
647: %
648: The CMB radiation is strongly anisotropic in the source frame which is
649: moving at a bulk Lorentz factor $\Gamma$, providing an external source
650: of seed photons which move opposite to the jet direction, i.e., $\psi
651: \simeq \pi$. Thereby, $(1-\beta \cos \psi)=(1+\mu_{\rm e})$, where
652: $\Omega_{\rm e}=(\cos^{-1}{\mu_{\rm e}}, \phi_{\rm e})$ is the
653: direction of the colliding electron in the comoving frame of the shear
654: layer, similarly to what we made for the case of SSC radiation. In the
655: Thomson regime $\sigma=\sigma_{_{\rm T}}\delta(\Omega-\Omega_{\rm
656: e})\delta \left[ \epsilon-\gamma^2\epsilon_{\rm cmb}(1+\mu_{\rm
657: e})\right]$, since the scattered photon energy is
658: $\epsilon=\gamma^2\epsilon_{\rm cmb}(1+\mu_{\rm e})$, where
659: $\epsilon_{\rm cmb}$ is the energy of the incident CMB photon in units
660: of the electron rest-mass. Hence, Eq.~\ref{eq:n_ic} can be cast as
661: \begin{equation}
662: \dot{n}_{_{\rm EC}}(\epsilon,\Omega)=\frac{c\sigma_{_{\rm T}}}{4\pi}(1+\mu)\int_{0}^{\infty}
663: {d\epsilon_{\rm cmb}}\int_{\gamma_0}^{\infty}
664: {d\gamma}n_{\rm cmb}(\epsilon_{\rm cmb})n_{\rm e}(\gamma)\delta\left[\epsilon-\gamma^2\epsilon_{\rm cmb}(1+\mu)\right]\ .
665: \label{eq:n_ec}
666: \end{equation}
667:
668: %
669: In order to compute the EC emissivity produced by the interaction of
670: CMB photons with the power-law part of the electron distribution, we
671: model the CMB field as a monochromatic radiation with photon energies
672: equal to the average value of a thermal black body spectrum
673: \citep{BG70} but blue-shifted by a factor $\Gamma$, i.e.,
674: $\epsilon_{\rm cmb} = \Gamma \langle \epsilon_{\rm cmb}^{\ast} \rangle \equiv 2.7
675: \times \Gamma \, k \, T^{\ast} / \me c^2$, and a number density
676: $n_{\rm cmb} = \Gamma \, u_{\rm cmb}^{\ast} / (\langle \epsilon_{\rm
677: cmb}^{\ast} \rangle \me c^2)$. The value that we adopt for the energy
678: density of the CMB radiation in the rest frame of the CMB is $u_{\rm
679: cmb}^{\ast}=4\cdot 10^{-13}\,$erg\,cm$^{-3}$. With these
680: approximations, the EC emissivity associated to the power-law part of
681: the electron distribution can be computed analytically
682: %
683: \begin{eqnarray}
684: j_{_{{\rm EC},1}}(\nu) &=&
685: \displaystyle{\frac{c \sigma_{_{\rm T}} a}{8\pi} u_{\rm cmb}^{\ast}
686: \left(\frac{h}{2.7kT^\ast} \right)^{1/2}
687: [(1+\mu)\Gamma]^{3/2}
688: \exp{\left[-\left(\frac{\epsilon}{\Gamma (1+\mu)
689: \langle \epsilon_{\rm cmb}^\ast \rangle \gamma_{\rm eq}^2} \right)^{1/2}
690: \right]}\nu^{-1/2}\:\:} \nonumber \\
691: & &\mbox{if }
692: \gamma_0^2 \le \displaystyle{
693: \frac{h\nu}{2.7kT^\ast\Gamma(1+\mu)}}\, .
694: \label{eq:j_ec_1}
695: \end{eqnarray}
696:
697: For the monoenergetic electron peak component, we use the exact black
698: body spectrum transformed to the source frame. In this case, the
699: expression of $n_{\rm cmb}(\epsilon_{\rm cmb})$ in the boundary layer
700: comoving frame is (see Eq.~2.58 of \citealp{BG70})
701: %
702: \begin{equation}
703: n_{\rm cmb}(\epsilon_{\rm cmb})=8\pi \me
704: \left(
705: \frac{\me c^2\epsilon_{\rm cmb}}{h}
706: \right)^2
707: \left[ \exp{ \left( \displaystyle\frac{\Gamma \me c^2\epsilon_{\rm cmb}}{kT^*}
708: \right) } - 1
709: \right]^{-1}\, ,
710: \label{eq:n_cmb}
711: \end{equation}
712: where $k$ is the Boltzmann constant. Substituting Eq.~\ref{eq:n_cmb}
713: and $n_{\rm e}(\gamma) = b\delta(\gamma - \gamma_{\rm eq})$ into
714: Eq.\ref{eq:n_ec}, taking into account that $j_{_{\rm
715: EC}}(\nu,\Omega)=h\epsilon \dot{n}_{_{\rm EC}}(\epsilon,\Omega)$
716: and the restriction imposed by the Thompson regime ($\epsilon_{\rm
717: cmb} \gamma_{\rm eq} < 1$), one obtains
718: \begin{eqnarray}
719: j_{_{{\rm EC},2}}(\nu)&=&
720: \displaystyle{\frac{2 h \sigma_{_{\rm T}} b}
721: {(1+\mu)^2 c^2 \gamma_{\rm eq}^6}
722: \left(\exp{\left(\frac{\Gamma h \nu}{(1+\mu)kT^\ast}\right)}-1 \right)^{-1}
723: \nu^3} \nonumber \\
724: & & \mbox{ if }
725: \nu \le \displaystyle{
726: \frac{(1+\mu) \gamma_{\rm eq} m_{\rm e} c^2}{h} }\, .
727: \label{eq:j_ec_2}
728: \end{eqnarray}
729:
730: The total EC emissivity is the sum of Eqs.~\ref{eq:j_ec_1} and
731: \ref{eq:j_ec_2}, i.e.,
732: \begin{equation}
733: j_{_{\rm EC}}(\nu) = j_{_{{\rm EC},1}}(\nu) + j_{_{{\rm EC},2}}(\nu)\, .
734: \label{eq:j_ec}
735: \end{equation}
736:
737: %--------------------------------------------------------
738: \subsection{Kinematic model of the shear layer}
739: \label{sec:kinematic}
740: %--------------------------------------------------------
741:
742: We model a large scale jet as a uniform cylindrical spine of radius
743: $R_{\rm c}$ with a Lorentz factor $\Gamma_{\rm c}$ followed by a
744: transition layer where the Lorentz factor changes linearly with radius
745: until $r=R_{\rm j}$, where the flow velocity either becomes zero
746: (similar to the SO02 model) or reaches a maximum from which it
747: abruptly decreases to zero (Figs.~\ref{fig:sl_structure},
748: \ref{fig:scheme}). For completeness, we also consider a boundary layer
749: where the Lorentz factor is uniform. More explicitly
750: %
751: \begin{equation}
752: \Gamma_{i}(r) = \left\{ \begin{array}{ll}
753: \Gamma_{\rm c} &\mbox{$r \leq R_{\rm c}$} \\
754: \Gamma_{{\rm j}, i} + (\Gamma_{\rm c} - \Gamma_{{\rm j}, i})
755: \frac{R_{\rm j} - r}{R_{\rm j} - R_{\rm c}} &\mbox{$R_{\rm c} < r
756: \leq R_{\rm j}$} \\
757: 1 &\mbox{$r>R_{\rm j}$}
758: \end{array} \right. ,
759: \label{eq:Gamma}
760: \end{equation}
761: %
762: where we take $\Gamma_{\rm c}=10$ for our reference models. The
763: subscript $i=1$ labels the case of a monotonic shear layer
764: ($\Gamma_{{\rm j},1}=1$), $i=2$ represents the case of an anomalous
765: shear layer ($\Gamma_{{\rm j},2}=15$) and $i=3$ denotes the case of a
766: jet where the boundary layer is uniform and has the same Lorentz
767: factor as the core ($\Gamma_{{\rm j},3}=10$). The kinematic
768: idealization of the case corresponding to an anomalous layer is a
769: rough prototype of the structure displayed by the Lorentz factor when
770: a ${\cal _{\leftarrow}\!R C S_{\!\rightarrow}}$ solution develops
771: (Fig.~\ref{fig:sl_structure_magnetic}). We neglect the variation of
772: the magnetic field across the boundary layer because it is rather
773: small (see Sect.~\ref{sec:RMHD_boost}) and we assume that the magnetic
774: field is uniform and equal to $10\,\mu$G.
775:
776: In order to properly normalize the emissivity properties of the
777: different jet models, we choose that all the shear layers transport
778: the same kinetic power as the monotonically decreasing one with a jet
779: radius $R_{{\rm j},1} = 2\,$kpc and a density and pressure equal to
780: the values chosen for the left state of the models shown in
781: Sect.~\ref{sec:RMHD_boost}. This means that we adjust the external jet
782: radius of the other two cases to $R_{{\rm j},2} = 1.25\,$kpc and
783: $R_{{\rm j},3} = 1.38\,$kpc. Since, typically, for extragalactic jets,
784: $\rho_{_{\rm ext}}\sim 10^{-27}-10^{-24}\,$g\,cm$^{-3}$
785: \citep{Ferrari98}, the fluxes of mass and of magnetic field in the jet
786: axial direction are
787: \begin{eqnarray}
788: \Phi &=& 3\cdot 10^{38} \left(\frac{B}{10^{-4}\,{\rm G}}\right)
789: \left(\frac{R_{\rm j}}{1\,{\rm kpc}}\right)^2 \,{\rm G\,cm}^{-2} \\
790: {\dot M} &=& 9\cdot10^{25}\left(\frac{R_{\rm c}}{1\,{\rm
791: kpc}}\right)^2 \left(\frac{\rho_{_{\rm ext}}}{10^{-24}\,{\rm
792: g\,cm}^{-3}}\right)\, {\rm g\,s}^{-1}\, .
793: \end{eqnarray}
794:
795: For each cylindrical shell in the shear layer of radius $r$ and moving
796: with a bulk Lorentz factor $\Gamma(r)$, we compute the local source
797: frame emissivities for the radiative processes considered in
798: Sect.~\ref{sec:model} (Eqs.~\ref{eq:j_syn}, \ref{eq:j_ssc} and
799: \ref{eq:j_ec}). Under the assumption that the radiating electrons are
800: distributed uniformly in the boundary layer (for a comoving observer),
801: the emissivity of the sub-layer at the distance $r$ from the jet axis
802: can be computed as
803: %
804: \begin{equation}
805: j^{\ast}_I (\nu^{\ast}, \theta^{\ast}, r) =
806: {\cal D}^2(r,\theta^{\ast}) \, j_I\left( \nu = {\nu^{\ast} \over
807: {\cal D}(r,\theta^{\ast} )}, \, \mu = { \mu^{\ast} - \beta(r)
808: \over 1 - \beta(r) \, \mu^{\ast}} \right) ,
809: \label{eq:j^ast}
810: \end{equation}
811: %
812: where ${\cal D}(r, \theta^{\ast}) \equiv \left[ \Gamma(r) (1 -
813: \beta(r) \, \mu^{\ast})\right]^{-1}$ is the Doppler factor
814: associated to a sub-layer moving with the Lorentz factor $\Gamma(r) =
815: \left[ 1 - \beta^{2}(r) \right]^{-1/2}$ at an angle $\theta^{\ast}
816: \equiv \cos^{-1} \mu^{\ast}$ with respect to the line of sight. The
817: subscript $I =$ syn, SSC or EC specifies the radiative process.
818:
819: The observed flux density from the boundary layer volume $V^{\ast}$
820: appropriately modified by the flow-beaming patterns is
821: %
822: \begin{equation}
823: S^{\ast} (\nu^{\ast}, \theta^{\ast}) =
824: d^{-2} \int_{R_{\rm c,i}}^{R_{\rm j,i}} dV^{\ast} \, j^{\ast}(\nu^{\ast}, \theta^{\ast},
825: r),
826: \label{eq:S(nu,theta)}
827: \end{equation}
828: %
829: $d$ being the distance to the observer (we take $d=10\,$Mpc unless
830: stated otherwise). $dV^{\ast}=2\pi \Delta_{\rm z}r dr$ corresponds to a
831: volume element of the shear layer with an observed length $\Delta_{\rm
832: z}$, which we assume to be $\Delta_{\rm z}=1\,$kpc. The integrals
833: involved in the evaluation of Eq.~\ref{eq:S(nu,theta)} are performed
834: numerically and, in order to compute the SSC contribution to the flux
835: density, we estimate the SSC emissivity (Eq.~\ref{eq:j_ssc}) by taking
836: $l=(R_{{\rm j},i} - R_{{\rm c},i})/\sin{\theta^{\ast}}$.
837:
838:
839: %--------------------------------------------------------
840: \section{Spectral energy distribution of different models of boundary
841: layers for kpc-scale jets}
842: \label{sec:SED}
843: %--------------------------------------------------------
844:
845: Our main goal is to outline the observational differences between the
846: emission of large-scale jets laterally endowed by different types of
847: shear layer. The observed spectral energy distribution of the
848: radiation emitted by different kinematic boundary layers in a
849: prototype kpc-scale jet is show in Fig.~\ref{fig:SED_AR-SO} for two
850: different viewing angles $\theta^\ast = 1^\circ$ and $\theta^\ast =
851: 60^\circ$. We neglect both the absorption of very high energy (VHE)
852: gamma-rays during the propagation to the observer and any contribution
853: from a separate non-thermal electron population accelerated at shocks
854: happening at the jet core. The choice of a small and large viewing
855: angle is motivated by the fact that for small viewing angles,
856: $\theta^\ast \simlt 4 \simeq 1/\Gamma_{{\rm j},2}$, the larger Doppler
857: boosting provided to the emitted radiation by the larger Lorentz
858: factor in AR layers than in monotonic layers makes the total flux
859: density of the former case also slightly larger
860: (Fig.~\ref{fig:SED_AR-SO} left panel). The differences in the total
861: flux density are smaller in the synchrotron dominated part of the
862: spectrum than in the EC dominated one. At larger angles
863: (Fig.~\ref{fig:SED_AR-SO} right panel), the flux density predicted for
864: standard layers is 2 to 5 orders of magnitude larger than for uniform
865: or AR layers due to the Doppler hiding of the later two cases. The
866: differences are very important in the EC dominated part of the
867: spectrum (i.e., in the gamma-ray regime).
868:
869: Both, anomalous and uniform boundary layers generate a rather similar
870: spectral pattern, particularly at small viewing angles. As we observe
871: the jet at larger values of $\theta^\ast$ (Fig.~\ref{fig:SED_AR-SO}
872: right panel) the smallest Lorentz factor of the uniform case provides
873: a smaller Doppler hiding which results into a larger flux density
874: than in case of AR layers.
875:
876: Considering specifically the flux density produced by anomalous layers
877: (Fig.~\ref{fig:SED_AR-thetas}), we note two main differences with the
878: standard case studied by SO02. First, the separation of the SSC and EC
879: peaks corresponding to IC emission of the power-law distribution of
880: the electrons (Eq.~\ref{eq:n(gamma)}) is larger in the former case.
881: For jets with AR layers the frequency separation between the EC1 and
882: SSC1 peaks grows with increasing viewing angle from less than one
883: order of magnitude ($\theta^\ast\sim 5^\circ$) to more than two orders
884: ($\theta^\ast\sim 90^\circ$). In case of jets endowed by standard
885: layers (Fig.~\ref{fig:SED_uniform-thetas}), the frequency separation
886: between the EC1 and SSC1 peaks also grows with the viewing angle, but
887: much less than in the previous case: the frequency separation between
888: the two peaks is smaller than a factor 8 at $\theta^\ast=90^\circ$. If
889: a sufficiently large number of TeV photons were detected, this
890: difference could be used to distinguish between jets with standard and
891: anomalous layers. However, we repeat here the cautionary note of SO02:
892: the spectral energy distributions (SEDs) computed in this paper
893: correspond to the most optimistic scenario, with a highly relativistic
894: jet spine and very efficient acceleration creating electrons with
895: large Lorentz factors up to $\gamma_{\rm eq}\sim 10^8$. Either smaller
896: values of $\Gamma_{\rm c}$ or $\gamma_{\rm eq}$ would reduce the
897: observed very high energy gamma-ray flux and shift the corresponding
898: peaks to lower energies.
899:
900: The second most important difference between AR and standard radiating
901: boundary layers is that the exponential decay of the synchrotron spectrum
902: is shifted towards smaller frequencies (delving deeper into the very
903: soft X-ray regime). This has a big impact in the effective X-ray
904: spectral index as we will discuss below.
905:
906: %
907: % Jet to counter jet brightness ratio.
908: %
909: As noted by \cite{kom90} and confirmed by SO02, extragalactic jets
910: with a monotonic boundary layer may show a jet-to-counterjet
911: radio-to-optical brightness ratio ($S^{\ast}_{\rm syn} (\theta^{\ast})
912: / S^{\ast}_{\rm syn} (\pi - \theta^{\ast})$) which is smaller than if
913: the jet were uniform. We find that such an assertion has to be
914: modified for jets limited by AR-boundary layers. In the latter case,
915: at small viewing angles (namely, $\theta^\ast\simlt 5^\circ$ for the
916: considered parametrization) the jet/counterjet asymmetry is $\sim 2 -
917: 3$ times larger than the one corresponding to a uniform jet with the
918: same kinetic power (Fig.~\ref{fig:jet2counterjet}). At larger viewing
919: angles, the brightness asymmetry turns out to be almost
920: indistinguishable from the case of an equivalent uniform jet (see in
921: Fig.~\ref{fig:jet2counterjet} how the solid and dashed lines
922: practically overlap each other for $\theta^\ast \simgt 20^\circ$ if
923: $\Gamma_{\rm c}=10$ or for $\theta^\ast \simgt 50^\circ$ in case
924: $\Gamma_{\rm c}=3$). Certainly, since the brightness asymmetry is a
925: result of the Doppler boosting (hiding) of the jet (counterjet), the
926: smaller is the value of the maximum Lorentz factor in the anomalous
927: shear layer ($\Gamma_{{\rm j},2}$), the smaller will be the maximum
928: asymmetry at small $\theta^\ast$. If the Lorentz factor of the jet
929: spine is small, the jet-to-counterjet brightness ratio reduces in
930: absolute value (see the grey lines corresponding to $\Gamma_{\rm
931: c}=3$, in Fig.~\ref{fig:jet2counterjet}).%
932: %
933: \footnote{The relative increase of the Lorentz factor in an anomalous
934: shear layer is rather independent on the Lorentz factor of the jet
935: (left state). Therefore, since we have chosen a parametrization for
936: our prototype jet in which $\Gamma_{{\rm j},2}/\Gamma_{\rm c}=1.5$,
937: for the case of $\Gamma_{\rm c}=3$ we take $\Gamma_{{\rm
938: j},2}=4.5$.}
939: %
940: However, even for the relatively small value of $\Gamma_{\rm c}=3$ the
941: jet-to-counterjet brightness ratio of the anomalous shear layer case
942: is larger than the one corresponding to the standard case at
943: $\Gamma_{\rm c}=10$ independent of the viewing angle (e.g., at
944: $\theta^\ast=0^\circ$, $\log{\left[S^{\ast}_{\rm syn} (0) /
945: S^{\ast}_{\rm syn} (\pi)\right]} \simeq 4.35$ for the AR model,
946: while $\log{\left[S^{\ast}_{\rm syn} (0) / S^{\ast}_{\rm syn}
947: (\pi)\right]} \simeq 4.25$). If there were a handle on the value
948: of the Lorentz factor of the jet, this difference might allow us to
949: discriminate observationally between jets with anomalous and standard
950: shear layers. However, given the similarity of the jet-to-counterjet
951: brightness ratio between jets with anomalous layers and jets with {\it
952: top-hat} profiles (uniform), it might be practically impossible to
953: disentangle from this unique value whether the jet is shielded by an
954: anomalous layer or by no layer at all.
955:
956: Considering the results obtained for standard boundary layers, SO02
957: concluded that the mildly relativistic velocities inferred from the
958: observed brightness asymmetries at tens of kiloparsec scales, may
959: correspond to a slower boundary layer and not necessarily to the faster
960: jet core, which might be highly relativistic at the observed
961: distances. If anomalous shear layers may happen in actual jets,
962: slower spines are preferred and relatively moderate maximum shear
963: layer Lorentz factors might also be invoked to explain the observed
964: jet-to-counterjet asymmetries.
965:
966: %
967: % X-ray bumb emission
968: %
969: In our models, the X-ray radiated power of kpc-scale jets bounded by
970: standard layers is dominated by the high-energy electron bump spectral
971: component, implying that the resulting synchrotron X-ray flux lies
972: above the extrapolated radio-to-optical continuum (see SO02 and
973: Figs.~\ref{fig:SED_AR-SO}, \ref{fig:SED_uniform-thetas}). This is also
974: true for jets endowed by anomalous boundary layers (see
975: Fig.~\ref{fig:SED_AR-thetas} at $\nu^\ast \simeq 10^{17}\,$Hz). For
976: both types of layers, the spectral slope at X-ray frequencies may be
977: rather different from that of the power-law at radio-to-optical
978: frequencies, and the difference depends also on the viewing
979: angle. This implies that the effective X-ray spectral index
980: $\alpha_{\rm X, eff}(\theta^\ast)$ computed between $h\nu^\ast_1 =
981: 1\,$keV and $h\nu^\ast_2 = 5\,$keV as
982: %
983: \begin{equation}
984: \alpha_{\rm X, eff}(\theta^{\ast}) = { \log \left[ S^{\ast}_{\rm X}
985: (\nu^{\ast}_1, \theta^{\ast}) \, / \, S^{\ast}_{\rm X} (\nu^{\ast}_2,
986: \theta^{\ast}) \right] \over \log \left[ \nu^{\ast}_2 \, / \,
987: \nu^{\ast}_1 \right] } ,
988: \label{eq:alphaX}
989: \end{equation}
990: %
991: can be significantly distinct. We point out that $S^\ast_{\rm
992: X}(\nu^{\ast}, \theta^{\ast})$ is computed in this paper as the
993: value of the flux density calculated using Eq.~\ref{eq:S(nu,theta)}
994: instead of SO02 who only take $S^\ast_{\rm X}(\nu^{\ast},
995: \theta^{\ast}) \propto \int r dr {\cal D}^2(r,\theta^\ast)
996: R(x^\ast_{\rm eq})$, with $R(x^\ast_{\rm eq})$ given by
997: Eq.~\ref{eq:R(x)}, and $x^\ast_{\rm eq}= \nu^\ast/c_1\gamma^2_{\rm eq}
998: {\cal D}(r,\theta^\ast)$.
999:
1000: For highly relativistic jet cores ($\Gamma_{\rm c}=10$),
1001: Fig.~\ref{fig:alphaX} (black lines) shows that the most noticeable
1002: difference between jets with anomalous or uniform shear layers and
1003: jets with standard limiting boundaries, is the fact that the effective
1004: X-ray spectral index of the jet $\alpha_{\rm X, eff}^{\rm j}$ is
1005: larger than the corresponding to the counterjet $\alpha_{\rm X,
1006: eff}^{\rm cj}$ for jet inclinations $\theta^\ast \simgt
1007: 65^\circ$. For standard boundary layer jets $\alpha_{\rm X, eff}^{\rm
1008: j} \leq \alpha_{\rm X, eff}^{\rm cj}$, $\forall \theta^\ast$ holds
1009: (i.e., the counter jet has a steeper X-ray continuum as compared to
1010: the jet spectrum). Furthermore, jets with uniform and AR layers show a
1011: much larger%
1012: %
1013: \footnote{Particularly, in the range $\theta^\ast \in
1014: [60^\circ,90^\circ]$, $\alpha_{\rm X, eff, (Uniform, AR)}^{\rm j}$
1015: is 6 to 7 times larger than $\alpha_{\rm X, eff, standard}^{\rm
1016: j}$.}
1017: %
1018: jet and counter jet effective X-ray spectral index than jets with
1019: standard layers if the inclination is $\theta^\ast \simgt 30^\circ$.
1020: Indeed, $\alpha_{\rm X, eff}^{\rm j}$ in the former two types of
1021: boundary layers is even larger than the $\alpha_{\rm X, eff}^{\rm cj}$
1022: corresponding to jets with monotonic layers for viewing angles
1023: $\theta^\ast \simgt 30^\circ$.
1024:
1025: The large values of the effective X-ray spectral index in jets with
1026: uniform and AR layers is due to the fact that the decay of the
1027: spectrum after the synchrotron peak crosses the X-ray band and moves
1028: towards optical frequencies for inclinations $\theta^\ast \simgt
1029: 60^\circ$ (Fig.~\ref{fig:SED_AR-thetas}). Consistently, the flux
1030: density at 5\,keV decays abruptly and, thus $\alpha_{\rm X, eff}$
1031: grows.
1032:
1033: In case of jets with more moderate spine velocities ($\Gamma_{\rm
1034: c}=3$; Fig.~\ref{fig:alphaX}, grey lines), there is no viewing angle
1035: for which $\alpha_{\rm X, eff}^{\rm j} > \alpha_{\rm X, eff}^{\rm
1036: cj}$, independent on the shear layer model. However, it remains true
1037: that the counterjet X-ray spectral index of jets bearing uniform or AR
1038: layers is much larger than the one corresponding to standard boundary
1039: layer jets (there is a factor of 2 to 4 difference between them).
1040:
1041: Taking together the results for models with jet spines flowing at
1042: $\Gamma_{\rm c}=3$ and 10, it turns out that the range of variability
1043: of both $\alpha_{\rm X, eff}^{\rm j}$ and $\alpha_{\rm X, eff}^{\rm
1044: cj}$ is larger for jets flanked by uniform and AR layers than for
1045: jets with standard boundary layers. Interestingly, the large-scale
1046: X-ray emission observed from several radio-loud AGNs exhibits rather
1047: different spectral characteristics. For instance, X-ray spectra of the
1048: known jets in quasars are very flat (e.g., $\alpha_{\rm X} \sim 0.23$
1049: for 3C 207, \citealp{bru01}; $\alpha_{\rm X} \sim 0.5$ for PKS 1127,
1050: \citealp{sie02}; $\alpha_{\rm X} \sim 0.8$ for 3C 273 and PKS
1051: 0637). The low effective spectral index is consistent with small jet
1052: inclinations (c.f. Fig.~\ref{fig:alphaX}, inset) independent of the
1053: shear layer model adopted. On the other hand, jets in radio galaxies
1054: tend to display relatively steep X-ray spectra, $\alpha_{\rm X}>1.0 -
1055: 1.5$ (\citealt{hac01} for 3C 66B, \citealt{wor01} for B2 0206 and B2
1056: 0755). These values correspond to large inclination angles on
1057: Fig.~\ref{fig:alphaX}, and/or lower jet Lorentz factors for jets with
1058: standard boundary layers. But, if uniform or AR emitting layers are
1059: considered, the viewing angles needed to account for such values of
1060: the X-ray spectral index are smaller ($\theta^\ast \in [20^\circ,
1061: 25^\circ]$ if $\Gamma_{\rm c}=10$, or $\theta^\ast \in [40^\circ,
1062: 55^\circ]$ if $\Gamma_{\rm c}=3$). Very recently, \cite{hac07} have
1063: reported values of the X-ray spectral index $\alpha_{\rm X}>2.5$ at
1064: distances of $\simeq 4\,$kpc from the nucleus in Cen A. These large
1065: values of $\alpha_{\rm X}$ are not within reach of jet models flanked
1066: by standard layers (the maximum of $\alpha_{\rm X, eff}^{\rm j}$ is
1067: 1.84 at $\theta^\ast=90^\circ$ for the model with $\Gamma_{\rm
1068: c}=10$). Nonetheless, jets with uniform or AR boundaries may yield
1069: $\alpha_{\rm X, eff}^{\rm j}>2.5$ for $\theta^\ast>32^\circ$
1070: ($\theta^\ast>59^\circ$) if $\Gamma_{\rm c}=10$ ($\Gamma_{\rm
1071: c}=3$). Indeed, considering that the inclination of Cen A is $\sim
1072: 60^\circ - 70^\circ$ \citep{SDK94}, our results suggest that, if a
1073: uniform or an AR emitting boundary layer were responsible for the
1074: emission of Cen A at distances from the nucleus $\simgt 4\,$kpc, small
1075: values of the spine Lorentz factor ($\Gamma_{\rm c}\simeq 3$) are
1076: preferred to explain the large values of $\alpha_{\rm X}$ observed at
1077: such scales.
1078:
1079: % counterjet X-ray spectral index
1080: Although very recently deep {\it Chandra} of observations of Cen A
1081: \citep{hac07} have provided us with spectra of some counterjet X-ray
1082: features, this is not the usual case for large scale
1083: counterjets. There are, however, some estimates of the lower limit of
1084: the X-ray brightness asymmetry, like e.g., for the 3C 66B ($> 25$,
1085: \citealt{hac01}), Pictor A ($> 15$, \citealt{wil01}; $6^{+12}_{-2}$
1086: \citealt{HC05}), or PKS 1127-145 ($> 5$, \citealt{sie02}). Our model
1087: of AR or uniform radiating boundary layer tends to favor larger X-ray
1088: brightness asymmetries between the jet and the counterjet. The reason
1089: being that the X-ray emission of the jet is dominated by the
1090: synchrotron radiation from the monoenergetic, hard ($\gamma=10^8$)
1091: electron component, while the counterjet emission is dominated by the
1092: EC radiation of the power-law part of the electron spectrum.
1093:
1094:
1095: %--------------------------------------------------------
1096: \section{Discussion}
1097: \label{sec:discussion}
1098: %--------------------------------------------------------
1099:
1100: SO02 chose the parameters of their radiating boundary layers such that
1101: they exhibit clearly the main characteristics of their model. We do
1102: not repeat here their discussion. Instead, we focus on the discussion
1103: of the choice of the new parameters of our model.
1104:
1105: SO02 anticipated that the form of the $\Gamma(r)$ radial profile and
1106: of the related spatial variation of the acceleration efficiency can
1107: significantly influence the beaming pattern and the intensity of the
1108: boundary layer emission. We actually have confirmed this point by
1109: replacing the monotonically decaying profile of $\Gamma(r)$ by an
1110: idealization of the Lorentz factor profile that results from AR
1111: boundaries. This results in a different relative normalization of the
1112: SSC and EC components presented in
1113: Figs.~\ref{fig:SED_AR-SO}-\ref{fig:SED_uniform-thetas}.
1114:
1115: The uniform layer model could be considered as a limiting case of AR
1116: layers when $\Gamma_{\rm j,i} \rightarrow \Gamma_{\rm c}$. Thus, when
1117: looking at the plots of the jet-to-counterjet ratio
1118: (Fig.~\ref{fig:jet2counterjet}) or the X-ray effective spectral index
1119: (Fig.~\ref{fig:alphaX}), any model including an AR layer with
1120: $\Gamma_{\rm j,i} \in [10,15]$ (if $\Gamma_{\rm c}=10$; black lines in
1121: Figs.~\ref{fig:jet2counterjet} and \ref{fig:alphaX}) or $\Gamma_{\rm
1122: j,i} \in [3,4.5]$ (if $\Gamma_{\rm c}=3$; grey lines in
1123: Figs.~\ref{fig:jet2counterjet} and \ref{fig:alphaX}), would display a
1124: graph located between the boundaries set by the solid and dashed lines
1125: of the corresponding figures. Since the area enclosed by these two
1126: line types (solid and dashed) is rather small, it turns out that our
1127: choice of the value of $\Gamma_{\rm j,i}=1.5\Gamma_{\rm c}$ does not
1128: significantly influence the results. More precisely, if for the
1129: considered values of $\Gamma_{\rm c}$ (3 or 10) we would have picked
1130: up any other value of $\Gamma_{\rm j,i} \in [\Gamma_{\rm c},
1131: 1.5\Gamma_{\rm c}]$, neither the value of the jet-to-counterjet
1132: brightness ratio, nor $\alpha_{\rm X}$, nor even the spectrum would
1133: have changed appreciably. Since, keeping all other physical conditions
1134: fixed, a decrease of ${\tilde \beta}_{_{\rm L}}$ yields smaller values of
1135: $\Gamma_{\rm j,i}$, the results we obtain shall be qualitatively valid
1136: both for magnetized or non-magnetized jets.
1137:
1138: The jet-to-counterjet effective X-ray spectral index asymmetry and its
1139: angular dependence (Fig.~\ref{fig:alphaX}) depends on which of the two
1140: spectral components at X-ray frequencies dominates: synchrotron
1141: radiation of the monoenergetic, high-energy electron component or the
1142: EC radiation from low energy electrons ($\gamma \sim 100$). In its
1143: turn, the dominance of any of these two processes depends on the exact
1144: shape of the electron spectrum at the highest energies, on the viewing
1145: angle, on the jet Lorentz factor and on
1146: $\Gamma(r)$. Figure~\ref{fig:alphaX} serves as an example of the
1147: comparative behavior of different boundary layer models (in which all
1148: the parameters are the same except $\Gamma(r)$).
1149:
1150:
1151: %--------------------------------------------------------
1152: \subsection{Summary}
1153: \label{sec:summary}
1154: %--------------------------------------------------------
1155:
1156: In this work, we compare the observational signatures imprinted by
1157: different kinematic models of shear layers of kiloparsec scale jets on
1158: the radiation of ultrarelativistic electrons accelerated at such
1159: boundary layers. The (simple) radiative model of a jet boundary layer
1160: follows that of SO02. Alternative (more sophisticated) models to
1161: compute the boundary layer radiation from multizone models (see, e.g.,
1162: \citealt{MA04,MA05,MA07}) will be considered elsewhere. The most
1163: important difference among distinct kinematic models is the radial
1164: profile of the Lorentz factor $\Gamma(r)$ across the shear layer. We
1165: have considered three different profiles. The first one is the usually
1166: assumed for large scale jets, namely, a monotonically decaying Lorentz
1167: factor profile (as in SO02). The second one is motivated by the
1168: results of \cite{AR06} and \cite{Mizuetal08}, namely, a non-monotonic
1169: profile where the Lorentz factor reaches a maximum at the limit
1170: between the jet and the external medium (AR layer). For completeness,
1171: the case of a uniform jet with a sharp edge is considered. The last
1172: case can be regarded as a limit of the AR profile when the maximum
1173: Lorentz factor in the boundary layer equals the Lorentz factor of the
1174: jet spine. Along the way, we have conveniently developed the results
1175: of \cite{AR06} and \cite{Mizuetal08} to include the effects of
1176: dynamically important magnetic fields in the parameter range which is
1177: adequate for large scale jets. From such development we have obtained
1178: and idealized kinematic model of anomalous boundary layers in
1179: magnetized extragalactic jets.
1180:
1181: Not surprisingly, we find that the jet-to-counter jet brightness ratio
1182: at radio frequencies is larger for jets flanked by AR layers than for
1183: uniform or standard layers with the same jet core Lorentz factor. This
1184: results from the fact that the differences in the radial profile of
1185: the velocity field determine the beaming pattern of the boundary layer
1186: emission. Thereby, this effect has to be taken into account when the
1187: jet bulk Lorentz factor is inferred from the jet-to-counterjet
1188: brightness asymmetry.
1189:
1190: The differences between AR and uniform boundary layer jets are small
1191: in the framework of our simple model. In practice, these small
1192: differences might render their observational distinction very
1193: difficult. Comparing these two models with standard radiating boundary
1194: layers the differences are much larger. Several independent clues can
1195: be used to distinguish (observationally) among them:
1196: %
1197: \begin{enumerate}
1198: \item If a sufficiently large number of TeV photons were detected, it
1199: would be possible to distinguish between jets with standard and
1200: anomalous layers by looking at the separation of the SSC and EC
1201: peaks. The separation between the two peaks is larger for jets with
1202: AR or uniform layers.
1203: %
1204: \item A large jet-to-counter jet brightness ratio (about two orders of
1205: magnitude larger than in standard boundary layer jets) is expected
1206: for jets bounded by AR or uniform layers.
1207: %
1208: \item For large viewing angles the effective X-ray spectral index is much
1209: larger for AR or uniform jets than for jets with standard boundary
1210: layers.
1211: %
1212: \item There can exist an inversion of the jet and counterjet X-ray
1213: spectral indices for jets seen at large viewing angles. However, for
1214: most inclination angles and moderate bulk Lorentz factors of the jet
1215: core, the effective X-ray spectral index of the jet is much smaller
1216: than that of the counterjet. Indeed, if there is a hint on the bulk
1217: Lorentz factor of the jet (e.g., because superluminal proper motions
1218: are detected), the jet-to-counterjet spectral X-ray index ratio may
1219: tell us whether the radiating layer is uniform, standard or AR. We
1220: note that for large Lorentz factors of the jet core ($\Gamma_{\rm c}
1221: \simlt 8$), the largest ratio $\alpha_{\rm X, eff}^{\rm j} /
1222: \alpha_{\rm X, eff}^{\rm cj}$ corresponds to jets with uniform
1223: layers, whilst the smaller one should be identified with standard
1224: boundary layers. For more moderate core Lorentz factor, the ratio
1225: $\alpha_{\rm X, eff}^{\rm j} / \alpha_{\rm X, eff}^{\rm cj}$ is
1226: larger for jets flanked by AR layers than for jets with uniform or
1227: standard layers. In spite of these facts, the differences in the
1228: effective spectral X-ray index between models with uniform and with
1229: AR layers are rather small and, they might render their
1230: observational distinction very difficult unless a very careful
1231: analysis of the observational data is performed.
1232: %
1233: \end{enumerate}
1234:
1235:
1236: \acknowledgements
1237: We kindly acknowledge the work of Jerome Ferrand during his
1238: internship at the Departamento de Astronom\'{\i}a y
1239: Astrof\'{\i}sica. We also thank L. Stawarz for very useful
1240: discussions and helpful hints. MAA is a Ram\'on y Cajal Fellow of
1241: the Spanish Ministry of Education and Science. He also acknowledges
1242: partial support from the Spanish Ministry of Education and Science
1243: (AYA2004-08067-C03-C01, AYA2007-67626-C03-01, CSD2007-00050). PM is
1244: at the University of Valencia with a European Union Marie Curie
1245: Incoming International Fellowship (MEIF-CT-2005-021603). The authors
1246: thankfully acknowledge the computer resources, technical expertise
1247: and assistance provided by the Barcelona Supercomputing Center -
1248: Centro Nacional de Supercomputaci\'on.
1249:
1250:
1251: %\appendix
1252:
1253:
1254:
1255: \begin{thebibliography}{}
1256: %
1257: \bibitem[Aloy \etal(1999a)]{alo99a} Aloy, M.A., Ib\'a\~nez, J.M., Mart\'{\i},
1258: J.M., M\"uller, E. 1999, \apj\, Supl. Ser., 122, 151
1259: %
1260: \bibitem[Aloy \etal(1999b)]{alo99b} Aloy, M.A., Ib\'a\~nez, J.M.,
1261: Mart\'{\i}, J.M., G\'omez, J.L., \& M\"uller, E. 1999, \apjl, 523,
1262: L125
1263: %
1264: \bibitem[Aloy \etal(1999c)]{alo99c} Aloy, M.A., Pons, J.A.,
1265: Ib\'a\~nez, J.M. 1999, Comp. Phys. Comm., 120, 115
1266: %
1267: \bibitem[Aloy \etal(2000a)]{alo00} Aloy, M.A., G\'omez, J.L.,
1268: Ib\'a\~nez, J.M., Mart\'{\i}, J.M., \& M\"uller, E. 2000, \apjl,
1269: 528, L85
1270: %
1271: \bibitem[Aloy, \etal(2000b)]{alo00b} Aloy, M.A., M\"uller, E.,
1272: Ib\'a\~nez, J.M., Mart\'{\i}, J.M., \& MacFadyen, A. 2000, \apjl,
1273: 531, L119
1274: %
1275: \bibitem[Aloy, \etal(2002)]{alo02} Aloy, M.A., Ib\'a\~nez, J.M.,
1276: Miralles, J.A., \& Urpin, V. 2002, \aap, 396, 693
1277: %
1278: \bibitem[Aloy, \etal(2003)]{alo03} Aloy, M.A., Mart\'{\i}, J.M.,
1279: G\'omez, J.L., Agudo, I., M\"uller, E., \& Ib\'a\~nez, J.M. 2003,
1280: \apjl, 585, L109
1281: %
1282: \bibitem[Aloy, Janka \& M{\"u}ller (2005)]{Alo05} {Aloy} M.~A.,
1283: {Janka} H.-T., {M{\"u}ller} E., 2005, A\&A, 436, 273
1284: %
1285: \bibitem[Aloy \& Rezzolla(2006)]{AR06} Aloy, M.A., \& Rezzolla,
1286: L. 2006, \apjl, 640, L115
1287: %
1288: \bibitem[Aloy \& Obergaulinger (2007)]{AO07} {Aloy} M.~A.,
1289: Obergaulinger, M., 2007, Revista Mexicana de Astronom\'{\i}a y
1290: Astrof\'{\i}sica, 30, 96
1291: %
1292: \bibitem[{{Attridge}, {Roberts} \& {Wardle}(1999)}]{ARW99}
1293: Attridge, J.M., Roberts, D.H., \& Wardle, J.F.C. 1999, \apjl,
1294: 518, L87
1295: %
1296: \bibitem[Baan(1980)]{Baan80} Baan, W.A. 1980, \apj, 239, 433
1297: %
1298: \bibitem[Begelman \& Li(1994)]{BL94} Begelman, M.C., Li, Z.-Y. 1994,
1299: \apj, 426, 269
1300: %
1301: \bibitem[Birkinshaw(1991)]{bir91} Birkinshaw, M. 1991, \mnras, 252, 505
1302: %
1303: \bibitem[{Birkl} et~al.(2007)]{Birkletal07} {Birkl}, R., {Aloy},
1304: M.~A., {Janka}, H.~-Th., \& {Mueller}, E. 2007, \aap, 463, 51
1305: %
1306: \bibitem[Blandford(2002)]{Blandford02} Blandford, R.G. 2002, in ESO
1307: Astrophysics Symposia. Eds. M. Gilfanov, R. Sunyaev, \&
1308: E. Churazov. Springer-Verlag, p. 381
1309: %
1310: \bibitem[Blandford(2003)]{Blandford03} Blandford, R. 2003, Ap\&SS,
1311: 288, 155
1312: %
1313: \bibitem[Blumenthal \& Gould(1970)]{BG70} Blumenthal, G.R., Gould,
1314: R.J. 1970, Rev. Mod. Phys., 42, 237
1315: %
1316: \bibitem[{{Brown} \& {Roshko}(1974)}]{BR74}
1317: Brown, G.~L. \& Roshko, A. 1974, J. Fluid Mech., 64(4), 775
1318: %
1319: \bibitem[Brunetti et al.(2001)]{bru01} Brunetti, G., Bondi, M.,
1320: Comastri, A., Setti, G. 2001, \aap, 381, 795
1321: %
1322: \bibitem[Cawthorne, \etal(1993)]{caw93} Cawthorne, T.V., Wardle,
1323: J.F.C., Roberts, D.H., \& Gabuzda, D.C. 1993, \apj, 416, 519
1324: %
1325: \bibitem[Chiaberge et al.(2000)]{chi00} Chiaberge, M., Celotti, A.,
1326: Capetti, A., \& Ghisellini, G. 2000, \aap, 358, 104
1327: %
1328: \bibitem[Crusius \& Schlickeiser(1986)]{cru86} Crusius, A., \&
1329: Schlickeiser, R. 1986, \aap, 164, L16
1330: %
1331: \bibitem[Dermer(1995)]{der95} Dermer, C.D. 1995, \apjl, 446, L63
1332: %
1333: \bibitem[De Young(1993)]{DeYoung93} De Young, D.~S. 1993,
1334: \apj, 405, L13
1335: %
1336: \bibitem[Dulwich, \etal(2007)]{Dulwichetal07} Dulwich, F., Worrall,
1337: D.M., Birkinshaw, M., Padgett, C.A., Perlman, E.S. 2007, \mnras,
1338: 374, 1216
1339: %
1340: \bibitem[Earl, Jokipii \& Morfill(1988)]{EJM88} Earl, J.A., Jokipii,
1341: J.R., Morfill, G. 1988, \apjl, 331, L91
1342: %
1343: \bibitem[Fendt(2006)]{Fendt06} Fendt, C. 2006, \apj, 651, 272
1344: %
1345: \bibitem[Fendt \& Ouyed(2004)]{FO04} Fendt, C., Ouyed, R. 2004, \apj,
1346: 608, 378
1347: %
1348: \bibitem[Ferrari(1998)]{Ferrari98} Ferrari, A. 1998, ARA\&A, 36, 539
1349: %
1350: \bibitem[Ferrari, Trussoni \& Zaninetti(1978)]{FTZ78} Ferrari, A.,
1351: Trussoni, E., \& Zaninetti, L. 1978, \aap, 64, 43
1352: %
1353: \bibitem[Ferrari, Trussoni \& Zaninetti(1982)]{FTZ82} Ferrari, A.,
1354: Trussoni, E., \& Zaninetti, L. 1982, \mnras, 198, 1065
1355: %
1356: \bibitem[Hanasz \& Sol(1996)]{HS96} Hanasz, M., \& Sol, H. 1996,
1357: \aap, 315, 355
1358: %
1359: \bibitem[Hardcastle, \etal(1997)]{hac97} Hardcastle, M.J., Alex\&er,
1360: P., Pooley, G.G., \& Riley, J.M. 1997, \mnras, 288, L1
1361: %
1362: \bibitem[Hardcastle et al.(2001)]{hac01} Hardcastle, M.J., Birkinshaw,
1363: M., \& Worrall, D.M. 2001, \mnras, 326, 1499
1364: %
1365: \bibitem[Hardcastle \& Croston(2005)]{HC05} Hardcastle, M.J., Croston,
1366: J.H. 2005, \mnras, 363, 649
1367: %
1368: \bibitem[Hardcastle et al.(2007)]{hac07} Hardcastle, M.J., Kraft,
1369: R.P., Sivakoff, G.R., et al. 2007, \apj, 670, L81
1370: %
1371: \bibitem[Hardee(1979)]{Hardee79} Hardee, P.E. 1979, \apj, 234, 47
1372: %
1373: \bibitem[Hardee(2007)]{Hardee07} Hardee, P.E. 2007, \apj, 664, 26
1374: %
1375: \bibitem[Kataoka \& Stawarz(2005)]{KS05} Kataoka, J., \& Stawarz,
1376: L. 2005, \apj, 622, 797
1377: %
1378: \bibitem[Kataoka, \etal(2006)]{Kataokaetal06} Kataoka, J., \& Stawarz,
1379: L., Aharonian, F., Takahara, F., Ostrowski, M., Edwards, P.G. 2006,
1380: \apj, 641, 158
1381: %
1382: \bibitem[Komissarov(1990)]{kom90} Komissarov, S.S. 1990, SvAL, 16, 284
1383: %
1384: \bibitem[Laing(1996)]{lai96} Laing, R.A. 1996, in {\it `Energy
1385: Transport in Radio Galaxies \& Quasars', ASP Conference Series
1386: vol. 100}, eds. P.E. Hardee, A.H. Bridle \& J.A. Zensus, San
1387: Francisco
1388: %
1389: \bibitem[Leismann \etal(2005)]{LA05} Leismann, T., Ant\'on, L., Aloy, M. A.,
1390: M\"uller, E., Mart\'i, J. M., Miralles, J. A., Ib\'a\~nez, J. M.
1391: 2005, A\&A, 436, 503
1392: %
1393: \bibitem[McKinney \& Narayan(2007)]{MN07} McKinney, J.C., Narayan,
1394: R. (2007), \mnras, 375, 513
1395: %
1396: \bibitem[Meliani \& Keppens(2007)]{MK07} Melinani, Z., Keppens,
1397: R. 2007, \aap, 475, 785
1398: %
1399: \bibitem[Mimica \etal(2004)]{MA04} Mimica P., Aloy M-A., M\"uller E.,
1400: Brinkmann W. 2004, A\&A, 418, 947
1401: %
1402: \bibitem[Mimica \etal(2005)]{MA05} Mimica P., Aloy M-A., M\"uller E.,
1403: Brinkmann W. 2005, A\&A, 441, 103
1404: %
1405: \bibitem[Mimica \etal(2007)]{MA07} Mimica P., Aloy M-A., M\"uller
1406: E. 2007, A\&A, 466, 93
1407: %
1408: \bibitem[Mizuno et al.(2008)]{Mizuetal08} Mizuno, Y., Hardee,
1409: P., Hartmann, D.H., Nishikawa, K.-I., \& Zhang, B. 2008, \apj, 672, 72
1410: %
1411: \bibitem[Ostrowski(1990)]{ost90} Ostrowski, M. 1990, \aap, 238, 435
1412: %
1413: \bibitem[Ostrowski(1998)]{ost98} Ostrowski, M. 1998, \aap, 335, 134
1414: %
1415: \bibitem[Ostrowski(2000)]{ost00} Ostrowski, M. 2000, \mnras, 312, 579
1416: %
1417: \bibitem[Owen, Hardee \& Cornwell(1989)]{owe89} Owen, F.N., Hardee,
1418: P.E., \& Cornwell, T.J. 1989, \apj, 340, 698
1419: %
1420: \bibitem[Perlman, \etal(1999)]{per99} Perlman, E.S., Biretta, J.A.,
1421: Fang, Z., Sparks, W.B., \& Macchetto, F.D. 1999, \aj, 117, 2185
1422: %
1423: \bibitem[Perucho, \etal(2007)]{Peruchoetal07} {Perucho}, M., {Hanasz}, M.,
1424: {Mart{\'{\i}}}, J.~M., \& {Miralles}, J.~A. 2007, Phys. Rev. E, 75, 056312
1425: %
1426: \bibitem[Pudridtz, Rogers \& Ouyed(2006)]{PCO06} Pudridtz, R.E., Rogers,
1427: C.S., Ouyed, R. 2006, \mnras, 365, 1131
1428: %
1429: \bibitem[Rieger \& Duffy(2004)]{RD04} Rieger, F.~M., \& Duffy,
1430: P. 2004, \apj, 617, 155
1431: %
1432: \bibitem[Romero et al.(2005)]{Romeroetal05} Romero, R., Mart\'{\i},
1433: J.M., Pons, J.A., Ib\'a\~nez, J.M., Miralles, J.A. (2005), J. Fluid
1434: Mech. 544, 323
1435: %
1436: \bibitem[Siemiginowska et al.(2002)]{sie02} Siemiginowska, A.,
1437: Bechtold, J., Aldcroft, T.L., Elvis, M., Harris, D.A., Dobrzycki,
1438: A. 2002, \apj, 570, 543
1439: %
1440: \bibitem[Skibo, Dermer \& Kinzer(1994)]{SDK94} Skibo, J.B., Dermer,
1441: C.D., Kinzer, R.L. 1994, ApJ, 426, L23
1442: %
1443: \bibitem[Stawarz \& Ostrowski(2002)]{SO02} Stawarz, \L ., \&
1444: Ostrowski, M. 2002, \apj, 578, 763 (SO02)
1445: %
1446: \bibitem[Stawarz et al.(2005)]{Stawarzetal05} Stawarz, \L .,
1447: Siemiginowska, A., Ostrowski, M., \& Sikora, M. 2005, \apj, 626, 120
1448: %
1449: \bibitem[Stawarz(2007)]{Stawarz} Stawarz, \L . 2007, private
1450: communication.
1451: %
1452: \bibitem[Swain, Bridle \& Baum(1998)]{SBB98} Swain, M.R., Bridle,
1453: A.H., \& Baum, S.A. 1998, \apjl, 507, L29
1454: %
1455: \bibitem[Turland \& Scheuer(1976)]{TS76} Turland, B.~D., \& Scheuer,
1456: P.~A.~G. 1976, \mnras, 176, 421
1457: %
1458: \bibitem[Urpin(2002)]{Urpin02} Urpin, V. 2002, \aap, 385, 14
1459: %
1460: \bibitem[Urpin(2006)]{Urpin06} Urpin, V. 2006, \aap, 455, 779
1461: %
1462: \bibitem[Wilson et al.(2001)]{wil01} Wilson, A.S., Young, A.J.,
1463: Shopbell, P.L. 2001, \apj, 547, 740
1464: %
1465: \bibitem[Worrall et al.(2001)]{wor01} Worrall, D.M., Birkinshaw, M.,
1466: \& Hardcastle, M.J. 2001, \mnras, 326, L7
1467: %
1468: \end{thebibliography}
1469:
1470: \clearpage
1471: \begin{figure}
1472: \plotone{f1.eps}
1473: \caption{{\it Left panel:} Profiles of the Lorentz factor (solid line)
1474: and of the laboratory frame magnetic field (dashed line) resulting
1475: from a prototypical Riemann problem yielding a $_{\leftarrow}{\cal
1476: SCS}_{\rightarrow}$ pattern. The initial left and right states are
1477: $(p_{_{\rm L}},\rho_{L},v^n_{_{\rm L}},\Gamma_{_{\rm L}},{\tilde \beta}_{\rm
1478: L})=(10^{-5},10^{-7},0.9,10,0.1)$ and $(p_{\rm
1479: R},\rho_{R},v^n_{\rm R},\Gamma_{\rm R},{\tilde \beta}_{\rm
1480: R})=(10^{-6},10^{-2},0,1,0.1)$, respectively. Note that no boost
1481: is produced to the left of the contact discontinuity located at
1482: $r\simeq 1.147$. {\it Right panel:} Same as the left panel, but for
1483: a prototype $_{\leftarrow}{\cal RCS}_{\rightarrow}$ pattern. The
1484: initial left and right states are $(p_{_{\rm L}},\rho_{L},v^n_{_{\rm
1485: L}},\Gamma_{_{\rm L}},{\tilde \beta}_{\rm L})=(10^{-5},10^{-7},0,10,100)$
1486: and $(p_{\rm R},\rho_{R},v^n_{\rm R},\Gamma_{\rm R},{\tilde \beta}_{\rm
1487: R})=(10^{-6},10^{-2},0,1,0.1)$, respectively. Note the boost (i.e., $\Gamma^*_{_{\rm L}} > \Gamma_{_{\rm L}}$)
1488: reached by the rarefied jet to the left of the contact discontinuity
1489: at $r\simeq 1.003$. In both cases the initial discontinuity is set
1490: at $r=1$ and the solution is shown after an arbitrary amount of time
1491: (we recall that the solution of the Riemann problem is
1492: self-similar).
1493: \label{fig:sl_structure_magnetic}}
1494: \end{figure}
1495:
1496: \clearpage
1497: \begin{figure}
1498: \epsscale{0.82}
1499: \plotone{f2.eps}
1500: \caption{Lorentz factor reached at the uniform state left to the
1501: contact discontinuity $\Gamma^*_{_{\rm L}}$ as a function of the
1502: normal velocity in the left state $v^n_{_{\rm L}}$. The right
1503: state is held fixed and is given by $(p_{_{\rm R}}, \rho_{_{\rm
1504: R}}, v^n_{_{\rm R}}, \Gamma_{_{\rm R}},{\tilde \beta}_{_{\rm R}}) =
1505: (10^{-6}, 10^{-2}, 0, 1, 0.1)$. The left state has fixed values in
1506: the total pressure, the rest-mass density and the Lorentz Factor
1507: $(p_{_{\rm L}}, \rho_{_{\rm L}}, \Gamma_{_{\rm L}})=(10^{-5},
1508: 10^{-7}, 10)$, while the magnetization ${\tilde \beta}_{_{\rm L}}$ is
1509: varied for each curve as indicated by the different labels (see
1510: inset). The solid lines refer to the $_{\leftarrow}{\cal
1511: RCS}_{\rightarrow}$ pattern, and the dashed lines refer to a
1512: $_{\leftarrow}{\cal SCS}_{\rightarrow}$ pattern. The inset shows a
1513: zoom of the solutions for very small values of $v^n_{_{\rm L}}$,
1514: which are the most expected ones for {\it standard} jet flows.
1515: \label{fig:betal_dep}}
1516: \end{figure}
1517:
1518: \clearpage
1519:
1520: \begin{figure}
1521: \plotone{f3.eps}
1522: \caption{Radial profile of the Lorentz factor across the jet showing
1523: the kinematic difference in the structure of an standard, monotonic
1524: shear layer (panel a) and an anomalous shear layer structure (panel
1525: b).
1526: \label{fig:sl_structure}}
1527: \end{figure}
1528:
1529: \clearpage
1530:
1531: \begin{figure}
1532: \epsscale{0.45}
1533: \plotone{f4.eps}
1534: \caption{Schematic view of the jet model.
1535: \label{fig:scheme}}
1536: \end{figure}
1537:
1538: \clearpage
1539: \begin{figure}
1540: \plotone{f5.eps}
1541: \caption{Predicted spectral energy distributions of the radiation
1542: generated in several boundary layers of a kpc-scale jet at
1543: $\theta^\ast=1^\circ$ and $\theta^\ast=60^\circ$ assuming
1544: $\gamma_{\rm eq}=10^8$. The blue lines corresponds to the same
1545: monotonic shear layer profile as in SO02, the black lines to a
1546: uniform boundary layer and the red lines to an anomalous one (see
1547: Sect.~\ref{sec:kinematic}). The different contributions to the
1548: spectrum are labeled above the corresponding lines. The labels 1 and
1549: 2 refer to the components associated to the power-law and to the
1550: monoenergetic parts of the electron distribution
1551: (Eq.~\ref{eq:n(gamma)}), respectively.
1552: \label{fig:SED_AR-SO}}
1553: \end{figure}
1554:
1555: \clearpage
1556: \begin{figure}
1557: \plotone{f6.eps}
1558: \caption{Predicted spectral energy distributions of the radiation
1559: generated by an AR boundary layer ($\Gamma_{\rm c}=10$,
1560: $\Gamma_{{\rm j},2}=20$) of a kpc-scale jet assuming $\gamma_{\rm
1561: eq}=10^8$, for observing angles
1562: $\theta^\ast=5^\circ, 30^\circ, 60^\circ$ and $90^\circ$. Indices
1563: 1 and 2 denote the spectral components associated to the power-law
1564: isotropic energy distribution and to the monoenergetic part,
1565: respectively. Absorption of VHE gamma-rays during the propagation
1566: to the observer is neglected.
1567: \label{fig:SED_AR-thetas}}
1568: \end{figure}
1569:
1570: \clearpage
1571: \begin{figure}
1572: \plotone{f7.eps}
1573: \caption{Same as Fig.\ref{fig:SED_AR-thetas} but for a standard
1574: boundary layer ($\Gamma_{\rm c}=10$, $\Gamma_{{\rm j},1}=1$).
1575: \label{fig:SED_uniform-thetas}}
1576: \end{figure}
1577:
1578: \clearpage
1579:
1580: \begin{figure}
1581: \plotone{f8.eps}
1582: \caption{Jet-to-counterjet flux density ratio,
1583: $S^{\ast}(\theta^{\ast}) / S^{\ast}(\pi - \theta^{\ast})$ at radio
1584: frequencies, as a function of the viewing angle
1585: $\theta^{\ast}$. Black and grey lines correspond to values
1586: $\Gamma_{\rm c}=10$ and $\Gamma_{\rm c}=3$, respectively --the
1587: values of $\Gamma_{\rm c}$ are provided near the respective
1588: curves--. The black (grey) solid line corresponds to the radiation
1589: from an anomalous boundary shear layer with the radial profile of
1590: Eq.~\ref{eq:Gamma} and $\Gamma_{{\rm j},1}=15$ ($\Gamma_{{\rm
1591: j},1}=4.5$). The dashed-doted line corresponds to the radiation
1592: from a monotonic boundary layer with the radial profile of
1593: Eq.~\ref{eq:Gamma} and $\Gamma_{{\rm j},2}=1$. The dashed line
1594: corresponds to the model with a uniform boundary layer with a
1595: Lorentz factor $\Gamma_{{\rm j},3}=10$. \label{fig:jet2counterjet}}
1596: \end{figure}
1597:
1598: \clearpage
1599:
1600: \begin{figure}
1601: \epsscale{0.85}
1602: \plotone{f9.eps}
1603: \caption{Expected effective X-ray spectral index of a jet (j) and of a
1604: counterjet (cj) boundary layer emission, $\alpha_{\rm X,
1605: eff}(\theta^\ast)$, as a function of the viewing angle
1606: $\theta^\ast$. Solid, dash-dotted and dashed lines correspond to
1607: jets with anomalous, standard and uniform shear layers, respectively
1608: (Eq.~\ref{eq:Gamma}). Black (grey) lines correspond to jets whose
1609: core moves with a Lorentz factor $\Gamma_{\rm c}=10$ ($\Gamma_{\rm
1610: c}=3$). The inset shows a zoom of the jet $\alpha_{\rm X, eff}$
1611: for the smallest viewing angles $\theta^\ast$. \label{fig:alphaX}}
1612: \end{figure}
1613: %
1614: \end{document}
1615: