0803.0805/ms.tex
1: %% AASTeX v5.x LaTeX 2e macros.
2: 
3: %% Produce a one-column, single-spaceddocument.
4: 
5: %\documentclass[apj]{emulateapj}
6: \documentclass[12pt,preprint]{aastex}
7: 
8: %% manuscript produces a one-column, double-spaced document:
9: 
10: %\documentclass[manuscript]{aastex}
11: 
12: %% preprint2 produces a double-column, single-spaced document:
13: 
14: %\documentclass[preprint2,longabstract]{aastex}
15: 
16: %Bibliographie suivant le style A&A
17: 
18: \usepackage{natbib}
19: \usepackage{color} % texte en couleur -> \textcolor{red}{texte}
20: \usepackage{ulem}  % texte barre -> \st{texte}; vague -> \uwave{texte}
21: \usepackage{psfrag}
22: 
23: \bibpunct{(}{)}{;}{a}{}{,}
24: 
25: \newcommand{\modif}{\bf }
26: 
27: \slugcomment{Accepted for publication in ApJ -- February 21, 2008}
28: 
29: \shorttitle{Azimuthal variations of the CR ion acceleration at the BW of SN 1006} 
30: 
31: \shortauthors{Cassam-Chena{\"\i} et al.}
32: 
33: %=====================================================================
34: 
35: \begin{document}
36: 
37: \title{Morphological evidence for azimuthal variations of the cosmic ray ion
38: acceleration at the blast wave of SN 1006}
39: 
40: \author{\mbox{Gamil Cassam-Chena{\"\i}\altaffilmark{1}, John P.
41: Hughes\altaffilmark{1}, Estela M. Reynoso\altaffilmark{2,3}, 
42: Carles Badenes\altaffilmark{4}, David Moffett\altaffilmark{5}}}
43: 
44: \altaffiltext{1}{Department of Physics and Astronomy, Rutgers
45: University, 136 Frelinghuysen Road, Piscataway, NJ 08854-8019;
46: \texttt{chenai@physics.rutgers.edu, jph@physics.rutgers.edu}} 
47: \altaffiltext{2}{Instituto de Astronom\'ia y F\'isica del
48: Espacio (IAFE), CC 67, Sucursal 28, 1428 Buenos Aires, Argentina}
49: \altaffiltext{3}{Departamento de F\'isica, Facultad de Ciencias Exactas y
50: Naturales, Universidad de Buenos Aires, Argentina}
51: \altaffiltext{4}{Department of Astrophysical Sciences, Princeton University,
52: Peyton Hall, Ivvy lane, Princeton, NJ 08544-1001}
53: \altaffiltext{5}{Department of Physics, Furman University, 
54: 3300 Poinsett Highway, Greenville, SC 29613} 
55: 
56: \begin{abstract}
57: Using radio, X-ray and optical observations, we present evidence for
58: morphological changes due to efficient cosmic ray ion
59: acceleration in the structure of the southeastern region of the
60: supernova remnant SN 1006. SN 1006 has an apparent bipolar morphology
61: in both the radio and high-energy X-ray synchrotron emission. In the
62: optical, the shock front is clearly traced by a filament of Balmer
63: emission in the southeast. This optical emission enables us to trace
64: the location of the blast wave (BW) even in places where the synchrotron
65: emission from relativistic electrons is either absent or too weak to
66: detect. The contact discontinuity (CD) is traced using images in the
67: low-energy X-rays (oxygen band) which we argue reveals the
68: distribution of shocked ejecta. We interpret the azimuthal variations
69: of the ratio of radii between the BW and CD plus the X-ray and radio
70: synchrotron emission at the BW using CR-modified hydrodynamic models.
71: We assumed different azimuthal profiles for the injection rate of
72: particles into the acceleration process, magnetic field and level of
73: turbulence.  We found that the observations are consistent with a
74: model in which these parameters are all azimuthally varying, being
75: largest in the brightest regions.
76: \end{abstract}
77: 
78: \keywords{ acceleration of particles --- cosmic rays --- shock waves
79: --- ISM: individual (SN 1006) --- supernova remnants }
80: 
81: 
82: %*******************************************
83: \section{Introduction\label{sect-intro}}
84: %*******************************************
85: 
86: Collisionless shocks in young supernova remnants (SNRs) are thought to
87: be responsible for the production and acceleration of the bulk of
88: Galactic cosmic rays (CRs)  at least up to the ``knee'' ($\sim 3$ PeV)
89: of the CR spectrum \citep[e.g.,][]{bev07}. The theoretical mechanism
90: is believed to be the first order Fermi acceleration, also known as
91: diffusive shock acceleration (DSA), where electrons, protons and other
92: ions scatter back and forth on magnetic fluctuations or Alfv\'en wave
93: turbulence through the velocity discontinuity associated with the
94: shock \citep[][and references therein]{joe91, mad01}.  The presence of
95: the turbulence is a key ingredient in the process.  The higher the
96: turbulence level, the higher the scattering rate and the faster the
97: acceleration proceeds, allowing particles to reach higher energies.
98: The fast acceleration implies also that the CR pressure becomes
99: significant very quickly so that the shock structure is highly
100: nonlinear \citep[e.g.,][]{elb00, ba07_NL}. The back-reaction of the
101: particles on the shock hydrodynamics increases the level of magnetic
102: turbulence and the injection rate in turn, i.e., the accelerated
103: particles create the scattering environment by themselves
104: \citep{be78a, blo78, vob03}. Moreover, the production of relativistic
105: particles (extracted from the thermal plasma) along with their escape
106: from the shock system increases the overall density compression ratio
107: \citep{bee99, bl02}.
108: 
109: This theoretical picture leads to several observational consequences.
110: The intensity of the synchrotron emission from shock-accelerated
111: electrons will be larger as the acceleration becomes more efficient
112: since both the number of relativistic particles and intensity of the
113: magnetic field must increase. In the region of very efficient
114: acceleration, the X-ray synchrotron emission from the highest energy
115: electrons will be concentrated in the form of thin sheets behind the
116: blast wave due to synchrotron cooling in the high turbulent magnetic
117: field amplified at the shock whereas the radio synchrotron emission
118: will be much broader in comparison \citep{elc05}. Finally, the
119: morphology of the interaction region between the blast wave (BW) and
120: the ejecta interface or contact discontinuity (CD) will become much
121: thinner geometrically as the acceleration becomes more efficient due
122: to the change in the compressibility of the plasma \citep{dee00,
123: eld04}. Because shocks are believed to put far more energy into ions
124: than electrons, this last point, if observed, would provide direct
125: evidence for the acceleration of CR ions at the BW.
126: 
127: The detection of $\gamma$-rays with a neutral pion-decay signature
128: resulting from the interaction of shock-accelerated protons with the
129: ambient matter would provide evidence for CR ion acceleration as well.
130: The best chance to see a clear pion-decay signal is when a SNR
131: interacts with a dense medium, the prototype case being RX
132: J1713.7--3946 \citep{cad04b}. Very high energy $\gamma$-rays have been
133: detected from this remnant \citep{aha07_RXJ1713} and also from a few
134: others \citep[e.g.,][]{aha07_RXJ0852,hol07_RCW86}. However, the
135: pion-decay signal is generally difficult to extract from these
136: observations because there are other processes efficiently producing
137: $\gamma$-rays but involving CR electrons, namely bremsstrahlung and
138: Inverse-Compton scattering off the 2.7 K cosmic microwave background
139: and Galactic photons. The nature of the underlying process leading to
140: $\gamma$-rays and therefore the nature of the emissive particles is
141: subject to intense debate \citep[e.g.,][]{bev06_RXJ1713, pom06,
142: uca07, pl08}. Besides, in a large number of SNRs where bright and
143: geometrically thin X-ray synchrotron-emitting rims are observed (e.g.,
144: Keper, Tycho, SN 1006), ambient densities are often very low leaving
145: open the question of CR ion production in these SNRs \citep[see
146: also][]{kaw08}.
147: 
148: The idea that the gap between the BW and CD allows quantifying the
149: efficiency of CR ion acceleration was proposed by \cite{de05} in the
150: case of Tycho, Kepler and Cas A. In Tycho, a remnant where thin X-ray
151: synchrotron-emitting rims are observed all around with little
152: variations in brightness, \citet{wah05} showed that the BW and CD are
153: so close to each other that they cannot be described by standard
154: hydrodynamical models (i.e., with no CR component), hence providing
155: evidence for efficient CR ion acceleration in this SNR.  In a later
156: study, \citet{cah07} showed that the X-ray and to some extent the
157: radio properties of the synchrotron emitting rim as well as the gap
158: between the CD and BW were consistent with efficient CR ion
159: acceleration modifying the hydrodynamical evolution of Tycho.
160: 
161: Ideally, one would like to find a SNR where both efficient and
162: inefficient particle acceleration take place so that differential
163: measurements can be made. This offers a great advantage for the
164: astrophysical interpretation since a number of uncertainties (e.g.,
165: distance and age) can be eliminated in the comparison.  Furthermore,
166: by directly observing a difference in the gap between the BW and CD in
167: regions of efficient and inefficient acceleration, one obtains a
168: direct, model-independent, confirmation that CRs do indeed modify the
169: hydrodynamical evolution of the BW. SN 1006 provides a valuable
170: laboratory in that regard since the gradual variations of the
171: intensity of the radio and X-ray synchrotron emission along the BW
172: point to underlying variation of CR ion acceleration.  The task
173: of measuring the ratio of radii between the BW and CD in
174: the regions of inefficient and efficient CR acceleration will be
175: challenging. Nonetheless, SN 1006 is one of the rare objects (perhaps
176: even the unique one) where this can be accomplished as we explain below.
177: 
178: In this paper, we focus our attention on the poorly studied
179: southeastern (SE) quadrant of SN 1006, i.e., along the shock front
180: lying between the northeastern and the southwestern
181: synchrotron-emitting caps. Our first goal is to measure the ratio of
182: radii between the BW and CD there. We can fairly easily follow the BW
183: in the bright synchrotron-emitting caps (presumably the regions of
184: efficient CR acceleration) and the BW can still be traced from the
185: H$\alpha$ emission in the region where both the radio and X-ray
186: synchrotron emissions are either too faint to be detected or absent
187: (presumably the regions of inefficient CR acceleration). As for the
188: CD, it can be traced from the thermal X-ray emission associated with
189: the shocked ejecta. Our second objective is to measure the azimuthal
190: variations of the synchrotron brightness in the radio and several
191: X-ray energy bands.  With these observational key results in hand, we
192: use CR-modified hydrodynamic models of SNR evolution to constrain the
193: azimuthal variations of the acceleration parameters, i.e., the
194: injection rate of particles in the acceleration process, strength of
195: the magnetic field, particle diffusion coefficient and maximum energy
196: of electrons and ions.
197: 
198: This paper is organized as followed. In \S \ref{sect-sn1006}, we
199: describe the basic characteristics of SN 1006. In \S \ref{sect-data},
200: we present the data used in our study. In \S \ref{sect-results}, we
201: present the key observational results in SN 1006. In \S
202: \ref{sect-models-vs_obs}, we try to relate hydrodynamical models with
203: those results in the context of CR-unmodified and -modified shocks. 
204: Finally, we discuss our results (\S \ref{sect-discussion}) and present
205: our conclusion (\S \ref{sect-conclusion}).
206: 
207: %*******************************************
208: \section{Basic characteristics of SN 1006\label{sect-sn1006}}
209: %*******************************************
210: 
211: SN 1006 was a thermonuclear supernova (SN) widely seen on Earth in the
212: year 1006 AD. More than a thousand year later, the remnant from this
213: explosion is a huge shell of $30\arcmin$ angular size.
214: 
215: The synchrotron emission is detected from the radio up to the X-rays
216: and dominates in two bright limbs -- the northeast (NE) and southwest
217: (SW) -- where several thin ($20\arcsec$ width or so) rims/arcs are
218: running at the periphery, sometimes crossing each other \citep{bay03,
219: lor03, rob04}. Those apparent ripples are most likely the result of
220: the projection of undulating sheets associated with the shock front.
221: The radio emission is well correlated with the nonthermal X-rays
222: \citep{reg86, lor03}. Both show an abrupt turn-on with coincident
223: edges in the radial directions, but the X-ray emission has more
224: pronounced narrow peaks. This points to a scenario in which the rims
225: are limited by the radiative losses\footnote{Note however that such a
226: model is incapable of reproducing the observed sharp turn-on of radio
227: synchrotron emission at SN 1006's outer edge, as in Tycho
228: \citep{cah07}.} in a highly turbulent and amplified magnetic field
229: \citep{bek02, bek03, elc05, ba06}.  Finally, the observed synchrotron
230: morphology in SN 1006 is best explained if the bright limbs are polar
231: caps with the ambient magnetic field parallel to the shock velocity
232: \citep{rob04}. In the caps, the maximum energy reached by the
233: accelerated particles, as well as their number, must be higher than
234: elsewhere \citep{rob04}.
235: 
236: In contrast to the nonthermal X-ray emission, the very faint thermal
237: X-ray emission seems to be distributed more or less uniformly
238: \citep{rob04}. This is best seen in the oxygen band (0.5-0.8 keV). It
239: is, however, difficult to separate the thermal emission from the
240: nonthermal emission in the caps. Moreover, it is not clear whether the
241: thermal X-rays should be attributed to the shocked ambient medium
242: \citep{yak07} or the shocked ejecta \citep{lor03}, but both the
243: over-solar abundances required to fit the X-ray spectra in the inner
244: northwest (NW) and NE parts of the SNR and the clumpiness on a
245: $30\arcsec$ to $1\arcmin$ scale of the low-energy X-ray emission favor
246: an ejecta origin. The SE region seems to be slightly
247: different than the rest of the SNR in terms of ejecta composition and
248: clumpiness too. There is, in particular, evidence for the presence of
249: cold \citep{wil05} and reverse-shock heated \citep{yak07} iron. In
250: terms of dynamics, there is evidence for ejecta extending
251: to/overtaking the BW in the NW \citep{lor03, vil03b_SN1006, rak07}.
252: 
253: In the optical, the remnant has a very different morphology.  Deep
254: H$\alpha$ imaging reveals very faint Balmer emission around almost the
255: entire periphery with a clear arc or shock front running from east to
256: south \citep{wil97b_SN1006, wig03}. Interestingly, no synchrotron
257: emission is detected in X-ray in this southeast (SE) region whereas
258: extremely faint (and highly polarized) synchrotron emission is
259: detected in the radio \citep{reg93}. Besides, none of the optical
260: emission is of synchrotron origin. There is also a clear Balmer line
261: filament in the NW of SN 1006 which has been observed in great detail
262: in the optical \citep{ghw02_SN1006, sog03, rak07}, ultraviolet
263: \citep{rab95, kor04, haf07} and X-ray bands \citep{lor03,
264: vil03b_SN1006, acb07}, providing diagnostics for the shock speed ($V_s
265: \sim 2400-3000\: \mathrm{km/s}$), ion-electron thermal equilibration
266: at the shock ($T_e/T_p \le 0.07$) and preshock ambient density ($0.25
267: \: \mathrm{cm}^{-3} \le n_0 \le 0.4 \: \mathrm{cm}^{-3}$). Optical
268: proper motions \citep{wig03} combined with the shock speed estimate
269: \citep{ghw02_SN1006, hem07} along this NW filament leads to a distance
270: to the SNR between $1.8$ and $2.3$ kpc. However, because the remnant
271: is interacting with a denser medium in the NW \citep{mog93, dug02},
272: the shock speed and the ambient density, in particular, might be
273: respectively higher and lower elsewhere \citep{acb07}. For instance,
274: the broadening of the oxygen \citep{vil03b_SN1006} and silicon
275: \citep{yak07} X-ray emission-lines suggest a shock speed $V_s \ge 4000
276: \: \mathrm{km/s}$ (although these measurements were carried out at
277: different locations in SN 1006). Such high velocities are compatible
278: with recent proper motion measurements based on radio data
279: \citep{moc04, rec08}. Moreover, the lack of observed TeV $\gamma$-ray
280: emission \citep{aha05_SN1006} sets an upper limit on the ambient
281: density of $n_0 \le 0.1 \: \mathrm{cm}^{-3}$ \citep{ksb05} and the
282: high latitude ($\sim 500 \: \mathrm{pc}$ at a distance of $2$ kpc) of
283: SN 1006 suggests that $n_0 \sim 0.03-0.04 \: \mathrm{cm}^{-3}$
284: \citep{fe01_ISM}.  
285: 
286: %*******************************************
287: \section{Data\label{sect-data}}
288: %*******************************************
289: 
290: \subsection{X-ray\label{subsect-data-xray}}
291: 
292: We used the most recent \textit{Chandra} data of SN 1006 (ObsId 3838
293: and ObsId 4385 up to 4394) which were obtained in April 2003 with the
294: ACIS-I imaging spectrometer in timed exposure and very faint data
295: modes. Eleven pointings were necessary to cover the entire extent of
296: the remnant. The final image is shown in Figure
297: \ref{fig_SN1006_3color_xray}. The X-ray analysis was done using CIAO
298: software (ver. 3.4). Standard data reduction methods were applied for
299: event filtering, flare rejection, gain correction. The final exposure
300: time amounts to $\sim 20$ ks per pointing.
301: 
302: \subsection{Radio\label{subsect-data-radio}}
303: 
304: Radio observations were performed in 2003 (nearly at the same
305: time as the X-ray observations) with both the Australia Telescope
306: Compact Array (ATCA) and the Very Large Array (VLA).
307: 
308: The ATCA consists of an array of six 22-m antennae that attain a
309: maximum baseline of 6 km in the East-West direction.  The ATCA
310: observations occurred on three separate occasions in antenna
311: configurations that optimize high spatial resolution: January 24 in
312: its 6-km B configuration for 12 hours, March 3 in its 6-km A
313: configuration for 12 hours, and June 12 in its 750-m C configuration
314: for 7 hours. Observations were made using two 128-MHz bands (divided
315: into 32 channels each) centered at 1384 and 1704 MHz.  The source PKS
316: 1934--638 was used for flux density and bandpass calibration, while PKS
317: 1458--391 was used as a phase calibrator.  The total integration time
318: on SN 1006 was over 1500 minutes.
319: 
320: The VLA consists of 27 25-m antennae arranged in a ``Y'' pattern.
321: Observations with the VLA were carried out during 4 hours on January
322: 25th in its hybrid CnD configuration (arranged to maximize spatial
323: resolution when observing southern declination sources).  Observations
324: were made using two 12.5-MHz channels centered at 1370 and 1665 MHz.
325: 3C286 and 1451--400 were observed for flux and phase calibration.  The
326: total VLA integration time on SN 1006 was 140 minutes.
327: 
328: Data from all of the observations were combined, then uniformly
329: weighted during imaging to minimize the effects of interferometric
330: sidelobes.  The final image, shown in Figure \ref{fig_SN1006_radio},
331: has a resolution of $\sim 6\arcsec \times 9\arcsec$ and an
332: off-source rms noise of $20 \: \mu\mathrm{Jy/beam}$ near the rims.
333: Since the angular size of SN 1006 is comparable to the primary beam of
334: both instruments (half a degree), corrections were applied to the
335: image to recover the lost flux. After primary beam corrections, the
336: total flux density was recovered to within $< 5\%$ of the expected value. 
337: With primary beam correction, noise increases in the faint SE portion.
338: For the study of the azimuthal variations of the radio emission at the shock, 
339: we use non beam corrected data (\S \ref{subsubsect-azimu-synch-emiss}).
340: 
341: \subsection{Optical\label{subsect-data-optical}}
342: 
343: We used the very deep H$\alpha$ image presented by \cite{wig03}.  This
344: image taken in June 1998, that is 5 years before the X-ray data, is shown
345: in Figure \ref{fig_SN1006_Halpha}.
346: To compare the optical and X-ray images, we had to correct for the
347: remnant's expansion. In the optical, proper motions of $0.280\pm0.008
348: \: \arcsec \: \mathrm{yr}^{-1}$ were measured (from April 1987 to June
349: 1998) along the NW rim where thin nonradiative Balmer-dominated
350: filaments are seen \citep{wig03}. In the radio, an overall expansion
351: rate of $0.44\pm0.13 \: \arcsec \: \mathrm{yr}^{-1}$ was measured
352: \citep[from May 1983 to July 1992,][]{mog93}, although higher values
353: have been measured using more recent observations \citep{moc04,
354: rec08}. This value does not include the NW rim where the optical
355: filaments are observed because there is simply little or no radio
356: emission there.  The observed expansion rate is clearly higher in the
357: radio than in the optical and is consistent with the picture in which
358: the BW encounters a localized and relatively dense medium in the NW. 
359: For simplicity, when using radial profiles, we use the value of
360: $0.40\arcsec \: \mathrm{yr}^{-1}$, so that the optical profiles are
361: shifted by $2\arcsec$ to be compared with the X-ray profiles. As we
362: show below, this correction is negligible compared to other
363: uncertainties when trying to locate the fluid discontinuities.
364: 
365: %*******************************************
366: \section{Key observational results\label{sect-results}}
367: %*******************************************
368: 
369: \subsection{Fluid discontinuities\label{sub-sect-radii-SE}}
370: 
371: In this section, we determine the location of the fluid
372: discontinuities -- i.e., the BW and CD -- in the SE quadrant of SN
373: 1006. Our goal is to measure the gap between the BW and CD
374: as a function of azimuth.
375: 
376: To trace the location of the BW, we use the H$\alpha$ image of SN 1006
377: which shows very faint filaments of Balmer emission around much of the
378: periphery of the remnant (Fig.~\ref{fig_SN1006_all}, \textit{top-right
379: panel}).  In the southeastern quadrant, the H$\alpha$ emission follows
380: a nearly circular arc.  We extracted radial profiles by azimuthally
381: summing over $4^{\circ}$ wide sectors starting from $158^{\circ}$
382: (from west) to $302^{\circ}$ (in the clockwise direction).  The center
383: of these profiles was determined to provide the best match to the
384: curvature of the H$\alpha$ rim.  The coordinates of the center are
385: $(\alpha_{\mathrm{J2000}}, \delta_{\mathrm{J2000}}) =
386: (15^{\mathrm{h}}02^{\mathrm{m}}56.8^{\mathrm{s}},-41^\circ 56\arcmin
387: 56.6\arcsec)$. This center is very close to the geometrical center of
388: the SNR, which, given its overall circularity, may not be too far from
389: the overall expansion center which has not yet been determined. In
390: practice we found it very difficult to determine the location of the
391: optical rim based purely on a numerical value (e.g., contour value or
392: enhancement factor above the local background level) because of the
393: presence of faint stellar emission, poorly subtracted stars, and faint
394: diffuse H$\alpha$ emission across the image.  So, we identified
395: locations where the H$\alpha$ emission increased slightly in the
396: radial profiles by eye and then further checked (again visually) the
397: corresponding radii on the optical image. We associated generous
398: uncertainties with the radii derived from this procedure: $12 \arcsec$
399: outward and $24 \arcsec$ inward (our process tended to overestimate
400: the radii, hence the asymmetric errors).
401: 
402: To trace the location of the CD, we use the X-ray image in the
403: low-energy band ($0.5-0.8$ keV) which contains most of the oxygen
404: lines (Fig. \ref{fig_SN1006_all}, \textit{bottom-left panel}). While
405: it is not clear whether the oxygen emission comes from the shocked
406: ambient medium or the shocked ejecta, we consider here an ejecta
407: origin. We discuss and justify this assumption below (\S
408: \ref{subsect-oxygen-ejecta-or-ISM}). Another issue comes from the fact
409: that the X-ray emission in the oxygen band may contain some nonthermal
410: contribution in the bright limbs. The three-color composite X-ray
411: image (Fig. \ref{fig_SN1006_3color_xray}) reveals that the oxygen emission
412: (red color) is in fact fairly different from the nonthermal X-rays
413: (white color), and appears notably clumpier even in the synchrotron
414: limbs (see the eastern limb for instance). This tells us that we can
415: still use the oxygen emission to trace the CD in the bright limbs, at
416: least in the azimuthal range we selected ($158^{\circ}-302^{\circ}$)
417: where the synchrotron emission is not completely overwhelming.  Radial
418: profiles were extracted from a flux image, summing over $1^{\circ}$
419: wide sectors, with the same center and azimuthal range as for the
420: optical. In the radial profiles, we selected the radii for which the
421: brightness becomes larger than $1.5 \times 10^{-5}$
422: ph/cm$^2$/s/arcmin$^2$.
423: 
424: Figure \ref{fig_radius_Ha_xray} summarizes the above-described
425: measurements and shows the azimuthal variations of the radii as
426: determined from the H$\alpha$ emission (\textit{black lines}) and
427: low-energy ($0.5-0.8$ keV, \textit{red lines}), mid-energy ($0.8-2.0$
428: keV, \textit{green lines}) and high-energy ($2.0-4.5$ keV,
429: \textit{blue lines}) X-ray emissions. The cross-hatched lines
430: indicate the location of the bright synchrotron limbs. They correspond
431: to the places where the contour of the outer high-energy X-ray
432: emission (\textit{blue lines}) can be determined. Between the two
433: synchrotron caps (angles between $200^\circ$ and $270^\circ$), there
434: is a clear gap between the BW and CD which is easily seen in Figure
435: \ref{fig_SN1006_all} (\textit{bottom-left} panel).  
436: In the bright limbs, however, the BW and CD are
437: apparently globally coincident.  Fingers of ejecta (indicated by the
438: \textit{stars}) are even sometimes visible clearly ahead of the
439: H$\alpha$ emission. In these places, the H$\alpha$ emission might not
440: be the best tracer of the BW location. The mid-energy X-ray emission
441: (\textit{green lines}) shows that these fingers are also present.  It
442: is not clear whether this emission traces the BW or, again, the ejecta
443: (mostly silicon).  The high-energy X-ray emission (\textit{blue
444: lines}) cannot be used, nor a X-ray spectral analysis, to answer this
445: point because of the low number of X-ray counts in this region.
446: 
447: Figure \ref{fig_Rs_o_Rc} shows the ratio of radii for the H$\alpha$
448: and low-energy X-ray emission, which to first approximation
449: corresponds to the ratio of radii between the BW and CD, $R_{\rm
450: BW}/R_{\rm CD}$.  With increasing azimuthal angles the ratio of radii
451: increases from values near unity (within the northeastern
452: synchrotron-emitting cap) to a maximum of
453: $R_{\mathrm{BW}}/R_{\mathrm{CD}} \simeq 1.10^{+0.02}_{-0.04}$ before
454: falling again to values near unity (in the southwestern
455: synchrotron-emitting cap). Over the entire azimuthal region where the
456: synchrotron emission is faint, the azimuthally averaged ratio of radii
457: is $R_{\rm BW}/R_{\rm CD} \simeq 1.04\pm0.03$. In the regions within
458: the synchrotron rims, $R_{\rm BW}/R_{\rm CD} \simeq 1.00$.
459: 
460: \subsection{Synchrotron emission at the blast wave\label{subsubsect-azimu-synch-emiss}}
461: 
462: In this section, we measure the azimuthal variations of the
463: synchrotron flux extracted at the BW and we investigate how these
464: variations compare when measured at different frequencies.
465: 
466: For that purpose, we use radio and X-ray data (Fig.
467: \ref{fig_SN1006_all}, \textit{left panels}).  We extracted the flux
468: in a $30\arcsec$-wide region behind the BW along the SE rim. To define
469: the BW location, we used the well-defined radii obtained from the
470: high-energy X-ray image for the northeastern ($\theta \leq 200^\circ$)
471: and southwestern ($\theta \geq 270^\circ$) regions while, in the
472: remaining SE quadrant ($200^\circ < \theta < 270^\circ$), we used the
473: radii determined from the H$\alpha$ image since no synchrotron
474: emission is detected there.
475: 
476: In Figure \ref{fig_flux_vs_angle_radio_xrays} (\textit{top panel}), we
477: show the azimuthal variations of the projected brightness in the radio
478: (1.5 GHz, \textit{black lines}), and several X-ray energy bands:
479: $0.5-0.8$ keV (\textit{red lines}), $0.8-2.0$ keV (\textit{green
480: lines}) and $2.0-4.5$ keV (\textit{blue lines}). The radio brightness
481: was multiplied by $2 \times 10^{-5}$ to make it appear on the same
482: plot with the X-ray data.  When moving from the northeastern or
483: southwestern synchrotron-emitting caps toward the faint SE region, we
484: see that the projected brightness gradually decreases in both radio
485: and X-ray bands. The radio, medium- (\textit{green lines}) and
486: high-energy (\textit{blue lines}) X-rays show mostly the variations of
487: the synchrotron emission, while the low-energy X-rays (\textit{red
488: lines}) show mostly the variations of the thermal (oxygen) emission,
489: but may contain some nonthermal contribution as well.
490: 
491: In Figure \ref{fig_flux_vs_angle_radio_xrays} (\textit{bottom panel}),
492: we show the same azimuthal variations but rescaled roughly to the same
493: level in the synchrotron-emitting caps. The logarithmic scale shows
494: that the brightness contrast between the bright caps and faint SE
495: region is of order $20$ to $50$ in the radio and X-rays. In fact, this
496: value might even be larger since the radio and high-energy X-ray
497: fluxes in the faint SE quadrant are compatible with no emission.  We
498: note that the azimuthal variations of the radio synchrotron emission
499: are much less pronounced than those in the X-rays. 
500: 
501: %*******************************************
502: \section{Relating hydrodynamical models to the observations\label{sect-models-vs_obs}}
503: %*******************************************
504: 
505: In the previous section, we have found evidence for azimuthal
506: variations of the separation between the BW and CD and synchrotron
507: brightness just behind the BW (\S \ref{sect-results}). Both the
508: separation and the brightness are correlated: the brighter the BW, the
509: smaller the separation. We measured in particular an unexpectedly small
510: separation between the BW and CD ($\sim 1.04$) in the SE quadrant of
511: SN 1006 where little to no synchrotron emission is detected.  These
512: measurements depend on the assumption that the oxygen emission comes
513: from the shocked ejecta.  After justifying this approach below (\S
514: \ref{subsect-oxygen-ejecta-or-ISM}), we first investigate the
515: different scenarios that could potentially lead to a small ratio
516: of radii between the BW and CD assuming no CR acceleration at the
517: shock (\S \ref{subsect-no-CR-accel}). Finally, we interpret the
518: observed azimuthal variations of the ratio of radii and synchrotron
519: brightness using CR-modified hydrodynamic models (\S
520: \ref{subsect-CR-accel}).
521: 
522: \subsection{Does the oxygen come from the ejecta?\label{subsect-oxygen-ejecta-or-ISM}}
523: 
524: A strong assumption made in the measurement of the ratio of radii
525: between the BW and CD in SN 1006 is that the thermal X-ray emission
526: (from the oxygen) is associated with the shocked ejecta.  
527: 
528: There is, however, still the possibility that the oxygen emission is
529: associated with the shocked ambient medium. \cite{yak07} have shown
530: that the integrated X-ray spectrum of the SE quadrant can be described
531: by a combination of three thermal plasmas in non-equilibrium
532: ionization and one power-law component.  One of the thermal
533: components, assumed to have solar abundances and therefore associated
534: with the shocked ambient medium, was able to produce most of the
535: observed low-energy X-rays (in particular the K$\alpha$ lines from O
536: \textsc{vii}, O \textsc{viii} and Ne \textsc{ix}).  The two other
537: thermal components, with non-solar abundances, were able to reproduce
538: most of the K$\alpha$ lines of Si, S and Fe, and were attributed to
539: the shocked ejecta.
540: 
541: There are several arguments from an examination of the high resolution
542: \textit{Chandra} images, however, that favor an ejecta origin for the
543: low-energy X-ray emission.  The three-color composite X-ray image
544: (Fig.~\ref{fig_SN1006_3color_xray}) shows that the spectral character
545: of the X-ray emission in the SE quadrant (where there is little to no
546: synchrotron emission) does not change appreciably as a function of
547: radius behind the BW, which would be expected if the \cite{yak07}
548: picture were correct. The lack of virtually any X-ray emission between
549: the BW and CD can be explained by the expected low density of the
550: ambient medium \citep[$\sim 0.03-0.04$ cm$^{-3}$][]{fe01_ISM} which
551: can reduce both the overall intensity and the level of line emission
552: (due to strong nonequilibrium ionization effects). If the soft X-ray
553: emission were from the shocked ambient medium its brightness should in
554: principle rise gradually behind the BW as a result of the projection
555: of a thick shell onto the line-of-sight \citep{wah05}, but the radial
556: profile of the low-energy X-ray emission (not shown) in fact turns on
557: rather quickly at the periphery.
558: 
559: Another argument in favor of an ejecta origin is the observed clumpiness
560: of the oxygen emission throughout the SE region. There is also the
561: protuberances seen right at the edge of SN 1006, which are suggestive
562: of Rayleigh-Taylor hydrodynamical instabilities that are expected at
563: the CD (see Fig. \ref{fig_SN1006_3color_xray}). Note that the outer edge of
564: the long, thin filament of X-ray emission in the NW -- which overlaps
565: the brightest H$\alpha$ emission and therefore is likely to be at least
566: partly due to shocked ambient medium -- is much smoother than the edge of the
567: remnant in the SE. Finally, high emission measures of oxygen in the
568: shocked ejecta at young dynamical SNR ages are compatible with most
569: thermonuclear explosion models. Essentially, all 1-D and 3-D
570: deflagration models, and all delayed detonation and pulsating delayed
571: detonation models have oxygen in the outer layers (but not prompt
572: detonations or sub-Chandrasekhar explosions) \citep{bab03}.
573: 
574: 
575: \subsection{The small gap in the faint SE region\label{subsect-no-CR-accel}}
576: 
577: We have presented evidence in \S \ref{sub-sect-radii-SE} for a small
578: ratio of radii $R_{\mathrm{BW}}/R_{\mathrm{CD}} \sim 1.04$ between the
579: BW and CD in the SE quadrant of SN 1006.  This is also the location
580: where the radio and X-ray nonthermal emission are either very weak or
581: undetected.  The most straightforward explanation for this lack of
582: synchrotron emission is an absence of efficient CR acceleration at the
583: BW here. In the following sections, we first investigate whether pure
584: 1-D hydrodynamical models for SNR evolution (i.e., without CR
585: acceleration) can reproduce such a small ratio of radii.  We consider
586: the role of ambient density, ejecta profile, and explosion energy on
587: the size of the gap between CD and BW.  Then we consider the effects
588: of 3-D projection on the hydrodynamically unstable CD.
589: 
590: \subsubsection{Predictions from standard hydrodynamical models\label{subsubsect-hydro-simu-no-CRs}}
591: 
592: To understand how a small gap can be produced in a region where there
593: is apparently no evidence for efficient particle acceleration, we used
594: standard one dimensional (1-D) spherically symmetric numerical
595: hydrodynamical simulations that follow the interaction of the ejecta
596: with the ambient medium. These simulations do not include any CR
597: component. 
598: 
599: Key inputs to these simulations are the initial density profiles of
600: the ejecta and ambient medium. In the following, we consider a uniform
601: density in the ambient medium and two different initial ejecta density
602: profiles: exponential and power-law. The exponential form has been
603: shown to be most representative of explosion models for thermonuclear
604: SNe \citep{dwc98}, but we also present results with a power-law
605: distribution to quantify the impact of the shape or compactness of the
606: ejecta profile on $R_{\mathrm{BW}}/R_{\mathrm{CD}}$.
607: 
608: In general, an exponential profile is expected for thermonuclear SNe
609: because the explosion is driven by the continuous release of energy
610: from the burning front as it propagates through the star, while the
611: power-law profile (more compact) is expected for core collapse SNe
612: because the explosion is driven by a central engine (core bounce), and
613: the shock loses energy as it propagates through the star
614: \citep{mam99}. Power law profiles with $n=7$ have been used to
615: represent 1-D deflagration models \citep[in particular, model W7
616: from][]{not84}, while exponential profiles are more adequate to
617: represent delayed detonation models \citep{dwc98}. When comparing
618: these two analytical profiles, it is worth noting that they transmit
619: momentum to the ISM in a different manner. The more compact power law
620: profile is a more efficient piston for ISM acceleration, and will lead
621: to smaller gaps than the less compact exponential profile
622: \citep{dwc98, bab03}.
623: 
624: The hydrodynamical simulations provide the radii of the BW and CD at
625: any time for a given ejecta profile, ambient medium density, and
626: explosion energy. In Figure \ref{fig_n0_vs_ratio}, we quantify the
627: impact of these three contributions on the ratio of radii
628: $R_{\mathrm{BW}}/R_{\mathrm{CD}}$. In particular, we plot the ratio
629: obtained for a wide range of ambient medium density ($0.001 \:
630: \mathrm{cm}^{-3} \leq n_0 \leq 30 \: \mathrm{cm}^{-3}$) using 1-D
631: exponential (EXP) and powerlaw (PL7) ejecta profiles with different
632: kinetic energies. The cross-hatched domains define the range of ratio
633: of radii\footnote{In Figure \ref{fig_n0_vs_ratio}, it is important to
634: keep in mind that the maximum value of the ratio found in Tycho comes
635: from a presumably efficient particle acceleration shock region while
636: in SN 1006, it comes from an inefficient particle acceleration
637: region.} observed in SN 1006 and Tycho
638: ($R_{\mathrm{BW}}/R_{\mathrm{CD}} \leq 1.10$) and range of typical
639: ambient densities ($0.01 \: \mathrm{cm}^{-3} \leq n_0 \leq 0.06 \:
640: \mathrm{cm}^{-3}$ in SN 1006; $0.1 \: \mathrm{cm}^{-3} \leq n_0 \leq
641: 0.6 \: \mathrm{cm}^{-3}$ in Tycho) consistent with constraints derived
642: from the observations \citep{acb07, cah07}.  It is clear that the 1-D
643: hydrodynamical simulations are unable to reproduce a ratio of radii as
644: small as the one observed in both SNRs.  We note however that such a
645: comparison between the observations and the models is not
646: straightforward since our measurements are likely to be affected by
647: projection and other effects as we detail next.
648: 
649: \subsubsection{Three dimensional projection effects\label{subsubsect-geom-proj}}
650: 
651: One of the limitations of our numerical simulations is the one
652: dimensionality.  Hydrodynamical simulations in 2-D or 3-D show that
653: hydrodynamical (Rayleigh--Taylor) instabilities project pieces of
654: ejecta ahead of the 1-D CD.  In these simulations, the outermost
655: pieces of ejecta reach half way of the gap between the 1-D CD and BW
656: \citep[e.g.,][]{chb92, wac01}. Therefore, because of such protrusions,
657: the line-of-sight projected CD radius, $\hat{R}_{\mathrm{CD}}$, will
658: appear larger than the true average CD radius, $R_{\mathrm{CD}}$.  On
659: the other hand, because the BW is not as highly structured as the CD
660: interface (it is not subject to Rayleigh--Taylor instabilities), it is
661: reasonable to assume that the true average BW radius,
662: $R_{\mathrm{BW}}$, is very close to the projected value,
663: $\hat{R}_{\mathrm{BW}}$.  It follows that the ratio of projected radii
664: between the BW and CD, $\hat{R}_{\mathrm{BW}} /
665: \hat{R}_{\mathrm{CD}}$, will be smaller than the true average ratio,
666: $R_{\mathrm{BW}} / R_{\mathrm{CD}}$. Quantifying the effects due to
667: projection is of considerable importance for interpreting the ratio of
668: radii between the BW and CD in the context of particle
669: acceleration\footnote{It is not even clear how the 1-D CR-modified
670: hydrodynamical results on the ratio of radii need to be adjusted given
671: multi-D Rayleigh--Taylor instability effects. Preliminary studies on
672: this have been done by \cite{ble01} but this work does not fully
673: quantify all relevant effects.}.
674: 
675: There are several attempts aimed at estimating the projection
676: correcting factor, $\xi$, where $\xi$ is defined as
677: $\hat{R}_{\mathrm{CD}} = (1 + \xi) \: R_{\mathrm{CD}}$. Based on the
678: projection with ejecta protrusions determined from a power-spectrum
679: analysis done at the observed CD in Tycho, \cite{wah05} estimated an
680: amount of bias of $\xi_{\mathrm{proj}} \simeq 6\%$. Based on the
681: projection of a shell of shocked ejecta with protrusions calculated in
682: a 2-D hydrodynamical simulations, \cite{dw00} and \cite{wac01} found a
683: slightly larger value of $\xi_{\mathrm{proj}} \simeq 10\%$. Taking
684: $\xi=10\%$, the predicted ratio of projected radii,
685: $\hat{R}_{\mathrm{BW}} / \hat{R}_{\mathrm{CD}} = R_{\mathrm{BW}} /
686: R_{\mathrm{CD}} \: / \: (1 + \xi)$, becomes $1.18 / (1+0.1) \simeq
687: 1.07$, where $1.18$ is the ratio of radii found at an age of 1000
688: years using 1-D power-law ejecta profiles assuming a kinetic energy of
689: the explosion of $10^{51}$ ergs and an ambient medium density of $0.03
690: \: \mathrm{cm}^{-3}$. Using exponential ejecta profiles, we find
691: $\hat{R}_{\mathrm{BW}} / \hat{R}_{\mathrm{CD}} =1.14$ because they
692: produce higher ratio of radii (i.e., $R_{\mathrm{BW}} /
693: R_{\mathrm{CD}}=1.25$). In either case (power-law or exponential
694: ejecta profiles), the average ratio (resp. 1.07 or 1.14) is still
695: larger than the average value of 1.04 measured in the SE quadrant of
696: SN 1006 (cf. \S \ref{sub-sect-radii-SE}). 
697: 
698: There are several possibilities to explain such discrepancy between
699: the models and the observations. For instance, the value of
700: $\xi_{\mathrm{proj}}$ could be larger in 3-D than in 2-D. Indeed,
701: models show that hydrodynamical instabilities can grow considerably
702: faster (by $\sim 30\%$) and penetrate further in 3-D than in 2-D
703: \citep{kaa00}. There can also be a certain amount of inhomogeneity in
704: the ejecta density distribution. Spectropolarimetric observations of
705: thermonuclear SNe do show that ejecta are clumpy on large scales
706: \citep{lel05}. Besides, \cite{wac01} have shown that high density
707: clump traveling with high speed in the diffuse SN ejecta can reach and
708: perturb the BW. This requires a high density contrast of a factor 100.
709: The density contrast is certainly not as high in SN 1006 because
710: otherwise we would see it in the low-energy X-ray and/or in the
711: optical image, but may be non-negligible. According to the previous
712: calculation, the ejecta clumping is required to contribute an
713: additional $\xi_{\mathrm{clump}} \sim 3-10\%$ factor (so that $\xi =
714: \xi_{\mathrm{proj}}+\xi_{\mathrm{clump}}$) in order to make the
715: observed and predicted ratios of radii agree. Finally, although we
716: could not come up with a reasonable solution, we cannot exclude the
717: possibility of a complex 3-D geometry in the SE quadrant of SN 1006
718: where the outermost extent of the shocked ejecta would not be directly
719: connected to the shock front emission in the hydrodynamical sense
720: (i.e., the CD and BW would corresponds to different parts of the
721: remnant seen in projection).
722: 
723: \subsection{Azimuthal variations of the CR ion acceleration\label{subsect-CR-accel}}
724: 
725: In this section, we try to explain both the observed azimuthal
726: variations of the ratio of radii between the BW and CD (\S
727: \ref{sub-sect-radii-SE}) and synchrotron flux at the BW (\S
728: \ref{subsubsect-azimu-synch-emiss}) in SN 1006 in the context of
729: CR-modified shock hydrodynamics where the efficiency of CR
730: acceleration is varying along the BW.
731: 
732: In the following, we use CR-modified hydrodynamic models of SNR
733: evolution which allow us for a given set of acceleration parameters to
734: calculate the ratio of radii, $R_{\mathrm{BW}}/R_{\mathrm{CD}}$, and
735: the synchrotron flux, $F_{\nu}$, at any frequency $\nu$. Relevant
736: parameters are the injection rate of particles into the DSA process,
737: the magnetic field strength and the particle diffusion coefficient (or
738: equivalently the turbulence level). Our goal will not consist in
739: finding the exact variations of these parameters along the shock by
740: fitting the data, but rather to provide a qualitative description of
741: the variations based on the theory to verify whether the model
742: predictions are consistent with the observations or not.
743: 
744: \subsubsection{Conceptual basis for modeling\label{subsect-basis-modelling}}
745: 
746: Here, we explicate the various assumptions concerning the spatial
747: distribution of the ambient density and magnetic field before running
748: our CR-modified hydrodynamic models. The knowledge of the remnant's
749: environment is important since it affects its evolution and also its
750: emission characteristics.
751: 
752: First, we assume that the ambient density is uniform. This is
753: suggested by the quasi-circularity of the H$\alpha$ filament in the SE
754: of SN 1006 and a reasonable assumption for remnants of thermonuclear
755: explosion \citep{bah07}. Second, we assume that the ambient magnetic
756: field direction is oriented along a preferred axis. In SN 1006, this
757: would correspond to southwest-northeast axis which is the direction
758: parallel to the Galactic plane \citep[following][]{rob04}. The
759: magnetic field lines are then parallel to the shock velocity in the
760: brightest part of the synchrotron caps in SN 1006 and perpendicular at
761: the equator. The angle between the ambient magnetic field lines and
762: the shock velocity will determine the number of particles injected in
763: the DSA process, i.e., the injection rate. The smaller the angle, the
764: larger the injection rate \citep{elb95, vob03}.
765: 
766: With this picture of an axisymmetry around the magnetic field
767: orientation axis in mind, the azimuthal conditions at the BW are
768: essentially spatially static, the evolution is self-similar and hence
769: temporal evolution of gas parcels can be followed with a 1-D
770: spherically symmetric code. We will run such a code with specific
771: initial conditions at each azimuthal angle along the BW (viewing the
772: SNR in a plane containing the revolution axis of the magnetic field).
773: We implicitly assume a radial flow approximation in our approach,
774: i.e., that each azimuthal zone is sufficiently isolated from the
775: others so that they evolve independently.
776: 
777: We run the self-similar models assuming a power-law density profile in
778: the ejecta with an index of $n=7$, an ejected mass and kinetic energy
779: of the ejecta of 1.4 $M_{\odot}$ and $10^{51}$ ergs, respectively,
780: which are standard values for thermonuclear SNe. We assume that the
781: SNR evolves into an interstellar medium whose  density is $n_0 = 0.03$
782: cm$^{-3}$ and pressure 2300 K cm$^{-3}$. For such a low density, the
783: use of the self-similarity is well justified as shown in Figure
784: \ref{fig_n0_vs_ratio}: the ratio of radii
785: $R_{\mathrm{BW}}/R_{\mathrm{CD}}$ obtained with a 1-D purely numerical
786: hydrodynamic simulation with the same initial conditions remains
787: indeed more or less constant even until the age of 1000 yrs
788: (\textit{solid yellow lines}). Other input parameters will be
789: specified later. First, we describe the CR
790: acceleration model used with the hydrodynamical model.
791: 
792: \subsubsection{CR acceleration model\label{CR-hydro-model}}
793: 
794: The models consist of a self-similar hydrodynamical calculation
795: coupled with a nonlinear diffusive shock acceleration model, so that
796: the back-reaction of the accelerated particles at the BW is taken into
797: account \citep{dee00}. For a given injection rate of protons,
798: $\eta_{\mathrm{inj}}$, far upstream magnetic field,  $B_{\mathrm{u}}$,
799: and diffusion coefficient, $D$, the shock acceleration model
800: determines the shock jump conditions and the particle spectrum from
801: thermal to relativistic energies \citep{bee99}. This allows us to
802: compute the ratio of radii and, by following the downstream evolution
803: of the particle spectrum associated with each fluid element, the
804: synchrotron emission within the remnant \citep{cad05}.
805: 
806: {The CR proton spectrum at the BW is a piecewise power-law
807: model with an exponential cutoff at high energies:
808: \begin{equation} \label{fp}
809: f_{\mathrm{p}}(E) = a \:  E^{-\Gamma(E)} \:
810:  \exp \left( -
811:  E / E_{\mathrm{p,max}} \right),
812: \end{equation}
813: where $a$ is the normalization, $\Gamma$ is the power-law index which
814: depends on the energy $E$, and $E_{\mathrm{p,max}}$ is the maximum
815: energy reached by the protons. Typically three distinct energy regimes
816: with different $\Gamma$ values are assumed \citep{bee99}.  The
817: normalization, $a$, is given by:
818: \begin{equation}\label{a}
819: a =  \: \frac{n_{\mathrm{inj}} \: q_{\mathrm{sub}}}{4 \: \pi \: p_{\mathrm{inj}}^3}
820: \end{equation}
821: where $n_{\mathrm{inj}} = \eta_{\mathrm{inj}} \: n_0 \
822: (r_{\mathrm{tot}}/r_{\mathrm{sub}} )$ is the number density of gas
823: particles injected in the acceleration process, with
824: $r_{\mathrm{tot}}$ and $r_{\mathrm{sub}}$ being the overall density
825: and subshock compression ratios, $q_{\mathrm{sub}} = 3 \;
826: r_{\mathrm{sub}} / (r_{\mathrm{sub}} - 1)$ and $p_{\mathrm{inj}}$ is
827: the injection momentum. Here, $p_{\mathrm{inj}} = \lambda \:
828: m_{\mathrm{p}} \: c_{s2}$ where $\lambda$ is a parameter (here fixed
829: to a value of 4) which encodes all the complex microphysics of the
830: shock and $c_{s2}$ is the sound speed in the shock-heated gas
831: immediately downstream from the subshock \citep[see \S 2.2
832: in][]{bee99}. We note that because there is a nonlinear reaction on
833: the system due to the injection, the parameter $\eta_{\mathrm{inj}}$
834: should be in fact related to $p_{\mathrm{inj}}$ \citep[see \S 5
835: in][for more details]{blg05}.
836: 
837: The CR electron spectrum at the BW is determined by assuming a certain
838: electron-to-proton density ratio at relativistic energies,
839: $K_{\mathrm{ep}}$, which is defined as the ratio between the electron
840: and proton distributions at a regime in energy where the protons are
841: already relativistic but the electrons have not yet cooled radiatively
842: \citep[e.g.,][]{elb00}. In the appropriate energy range, the CR
843: electron spectrum is then:
844: \begin{equation} \label{fe}
845: f_{\mathrm{e}}(E) = a \: K_{\mathrm{ep}} \: E^{-\Gamma(E)} \:
846:  \exp \left( - 
847:  E / E_{\mathrm{e,max}} \right),
848: \end{equation}
849: where $E_{\mathrm{e,max}}$ is the maximum energy reached by the
850: electrons, which could be eventually lower than that of protons
851: ($E_{p,\mathrm{max}}$) through synchrotron cooling of electrons
852: depending on the strength of the post-shock magnetic field.
853: $K_{\mathrm{ep}}$ is left as a free parameter. We will obtain
854: typically $K_{\mathrm{ep}} \sim 10^{-4}-10^{-2}$.
855: 
856: The maximum energies $E_{p,\mathrm{max}}$ and $E_{e,\mathrm{max}}$
857: contain information on the limits of the acceleration. They are set by
858: matching either the acceleration time to the shock age or to the
859: characteristic time for synchrotron losses, or by matching the
860: upstream diffusive length to some fraction, $\xi_s$, of the shock
861: radius, whichever gives the lowest value. We set $\xi_s=0.05$,
862: a value that allows to mimic the effect of an expanding spherical
863: shock  as compared to the plane-parallel shock approximation assumed
864: here \citep[see][]{be96, elb00}. A fundamental parameter in
865: determining the maximum energy of particles is the diffusion
866: coefficient, $D$, which contains information on the level of
867: turbulence and encodes the scattering law \citep{pam06}. For the sake
868: of simplicity, we assumed the Bohm regime for all particles (i.e.,
869: diffusion coefficient proportional to energy). We allow deviations
870: from the Bohm limit (which corresponds to a mean free path of the
871: charged particles equal to the Larmor radius, which is thought to be
872: the lowest possible value for isotropic turbulence) via the parameter
873: $k_0$ defined as the ratio between the diffusion coefficient, $D$, and
874: its Bohm value, $D_{\mathcal{B}}$. Hence $k_0 \equiv D /
875: D_{\mathcal{B}}$ is always $\geq 1$ and $k_0=1$ corresponds to the
876: highest level of turbulence. The smaller the diffusion coefficient (or
877: $k_0$), the higher the maximum energy (the maximum energy scales as
878: $1/\sqrt{k_0}$ in the radiative loss case and as $1/k_0$ in the age-
879: and escape-limited cases).
880: 
881: \subsubsection{Heuristic models}
882: 
883: In order to make predictions for the azimuthal variations of the ratio
884: of radii and synchrotron flux, we must make some assumptions about the
885: azimuthal variations of the input parameters used in our CR-modified
886: hydrodynamical models. Relevant input parameters are the injection
887: rate ($\eta_{\mathrm{inj}}$), the upstream magnetic field
888: ($B_{\mathrm{u}}$), the electron-to-proton ratio at relativistic
889: energies ($K_{\mathrm{ep}}$) and the level of turbulence or diffusion
890: coefficient relative to the Bohm limit ($k_0$).
891: 
892: To understand the influence of these four basic input parameters, we
893: begin first with a heuristic discussion to set the stage for later
894: detailed and more physical models (\S \ref{subsubsect-good-model}).
895: In Figure \ref{fig_heuristic_model_profiles}, we show three models
896: where we vary the two most important parameters, $\eta_{\mathrm{inj}}$
897: and $B_{\mathrm{u}}$, with azimuthal angle ($K_{\mathrm{ep}}$ is held
898: constant and $k_0$ is set to 1 at all angles). Model 1 (\textit{left
899: panels}) has $\eta_{\mathrm{inj}}$ varying from $10^{-5}$ to
900: $10^{-3}$, while $B_{\mathrm{u}}$ is held constant at the value $25 \;
901: \mu\mathrm{G}$. Model 2 (\textit{middle panels}) has constant
902: $\eta_{\mathrm{inj}} = 10^{-4}$ and $B_{\mathrm{u}}$ varying from $3
903: \; \mu\mathrm{G}$ to $25 \;\mu\mathrm{G}$. Model 3  (\textit{right
904: panels}) is a combination of models 1 and 2 and has
905: $\eta_{\mathrm{inj}}$ varying from $10^{-5}$ to $10^{-3}$ and
906: $B_{\mathrm{u}}$ varying from $3 \; \mu\mathrm{G}$ to $25
907: \;\mu\mathrm{G}$.
908: 
909: Note that in all these models, the functional form of the profiles
910: when $\eta_{\mathrm{inj}}$ and $B_{\mathrm{u}}$ vary is rather
911: arbitrary and just used for illustration purposes (for completeness we
912: note that the angular range shown corresponds to the data extraction
913: region in Fig.~\ref{fig_flux_vs_angle_radio_xrays}). Moreover, the
914: upstream magnetic field, $B_{\mathrm{u}}$ (\textit{second panels,
915: solid lines}), is allowed to vary with azimuthal angle because we
916: implicitly assume that it can be significantly amplified by the
917: CR-streaming instability. Our model does not include self-consistently
918: the magnetic field amplification believed to occur at SNR shocks, but
919: is provided with a simple compression.  Assuming that the magnetic
920: turbulence is isotropic ahead of the shock, the immediate post-shock
921: magnetic field, $B_{\mathrm{d}}$ (\textit{dashed lines}), will then be
922: larger than upstream by a factor $r_B \equiv B_{\mathrm{d}} /
923: B_{\mathrm{u}} = \sqrt{(1 + 2 \: r_{\mathrm{tot}}^{2})/3}$. This
924: relation is assumed at each azimuthal angle\footnote{Here, the
925: compression does not depend on the angle between the upstream magnetic
926: field and shock velocity as considered in \cite{re98} or \cite{orb07}.
927: Instead we assume that after the amplification process, the magnetic
928: field becomes entirely turbulent, so its originally ordered character
929: has been lost.}.
930: 
931: We show the predicted profiles of the ratio of radii between the BW
932: and CD (\textit{third panels}) and synchrotron brightness projected
933: along the line-of-sight in the radio and different X-ray energy bands
934: (\textit{bottom panels}). We explain later how those brightness
935: profiles were precisely constructed (\S \ref {subsubsect-good-model}).
936: Clearly, model 2 does not predict the expected correlation between the
937: ratio of radii and the synchrotron brightness and hence can be
938: rejected immediately\footnote{Looking at the BW to CD ratio in
939: model 2, it naively seems surprising to have a less modified shock
940: (i.e., higher BW to CD ratio) where the magnetic field is larger.
941: This effect is due to the presence of Alfv\'en heating in the
942: precursor \cite[see][]{bee99}, although there is debate about whether
943: this is the most efficient mechanism for turbulent heating in the
944: precursor \citep[see][]{amb06}.}. On the other hand, models 1 and 3
945: lead to the expected correlation: the smaller the BW/CD ratio, the
946: brighter the synchrotron emission. Comparing models 1 or 3 with model
947: 2 allows us to see that only an azimuthal variation of the injection
948: rate leads to the appropriate variation in the ratio of BW to CD radii
949: (\textit{third panels}). Comparing model 1 with model 3 shows the
950: impact of the magnetic field variations through synchrotron losses on
951: the azimuthal profiles of the X-ray synchrotron emission
952: (\textit{bottom panels}).
953: 
954: Although models 1 and 3 seem to describe the key observational
955: constraints fairly well, they suffer weaknesses as regards their
956: underlying astrophysical premises.  In model 1, it is difficult to
957: justify why the magnetic field should be enhanced in the region of
958: inefficient particle acceleration ($\eta_{\mathrm{inj}} \leq 6 \times
959: 10^{-5}$).  The model proposes a post-shock magnetic field there of
960: $\sim 80 \: \mu\mathrm{G}$, when a value of $10-15 \: \mu\mathrm{G}$ would
961: more likely reflect the value of the compressed ambient field.  In
962: model 3, it is difficult to understand why the turbulence level would
963: saturate (i.e., $k_0 = 1$) everywhere. Hence, we are led to develop
964: more elaborated models where $k_0$ and/or $K_{\mathrm{ep}}$ are
965: allowed to vary with azimuthal angle. A variation of either $k_0$ or
966: $K_{\mathrm{ep}}$ will not significantly affect the ratio of BW to CD
967: radii, but will only modify the profiles of the synchrotron emission.
968: In fact, a model with varying $K_{\mathrm{ep}}$ can be already
969: excluded because this quantity changes the radio and X-ray synchrotron
970: intensities in the same way while the observed radio and X-ray
971: profiles (Fig.~\ref{fig_flux_vs_angle_radio_xrays}) vary with azimuth
972: in different ways. Varying $k_0$ (as we show in \S \ref
973: {subsubsect-good-model}) does allow us to modify the relative
974: azimuthal profiles.
975: 
976: \subsubsection{Can CR-modified models provide $R_{\mathrm{BW}}/R_{\mathrm{CD}}=1.00$?\label{ratio-unity}}
977: 
978: Although CR-modified hydrodynamical simulations predict a ratio of
979: radii $R_{\mathrm{BW}}/R_{\mathrm{CD}}$ smaller than standard
980: hydrodynamical simulations, they still predict a lower limit on that
981: ratio that is strictly larger than unity. For instance, Figure
982: \ref{fig_heuristic_model_profiles} (\textit{third-right panel}) shows
983: that the modeled ratio of radii where the synchrotron emission is
984: strong is $\sim 1.13$. This value was obtained assuming an injection
985: rate of $10^{-3}$ and an ambient magnetic field of 25 $\mu$G. Keeping
986: all the same inputs, but reducing the magnetic field value to 3
987: $\mu$G, we obtain the lowest possible value of
988: $R_{\mathrm{BW}}/R_{\mathrm{CD}}=1.06$.
989: 
990: On the other hand, the observations of SN 1006 show that the ratio of
991: radii is of order 1.00 in the synchrotron-emitting caps, a value that
992: is not possible to obtain from our CR-modified model under the DSA
993: framework.  However, as already mentioned before in our discussions of
994: the small gap between the BW and CD in the region of inefficient
995: acceleration (\S \ref{subsect-no-CR-accel}), we need to consider
996: projection effects due to hydrodynamical instabilities at the CD which
997: can reduce the gap by $6-10\%$.  This does not fully account for the
998: difference with the model and, therefore, we may need to invoke
999: another effect, such as clumping of the ejecta, to fully explain the
1000: offset between the observation and models.  It is important to note
1001: that in both the regions of efficient and inefficient CR acceleration,
1002: we require essentially the same numerical factor to bring the modeled
1003: ratio of radii (for the appropriate model in each case) into agreement
1004: with the observed ratio. In the following, we introduce an ad-hoc
1005: rescaling of the ratio of radii by decreasing the modeled values by a
1006: constant factor ($\sim 13\%$) at all azimuthal angles. By doing this,
1007: we implicitly assume that this factor contains the total contribution
1008: of effects that are unrelated to the CR acceleration process. The
1009: numerical value used depends, of course, on the particular SN
1010: explosion model used. For example, the factor would need to be
1011: increased slightly in the case of an exponential ejecta profile (cf.,
1012: Fig.~\ref{fig_n0_vs_ratio}).
1013: 
1014: \subsubsection{Toward a good astrophysical model\label{subsubsect-good-model}}
1015: 
1016: In Figure \ref{fig_model_k0_var_profiles}, we present in more detail
1017: what  we believe is a reasonable astrophysical model. This model
1018: assumes that $\eta_{\mathrm{inj}}$ increases from a value of $5 \times
1019: 10^{-5}$ to a value of $10^{-3}$ while at the same time
1020: $B_{\mathrm{u}}$ increases from a value of $3 \;\mu\mathrm{G}$ to a
1021: value of $20 \; \mu\mathrm{G}$ and $k_0$ increases from an arbitrary
1022: value of 100 to 1 ($K_{\mathrm{ep}}$ is kept constant at all angles).
1023: The azimuthal variations of these parameters is consistent with the
1024: picture of DSA where, once the magnetic field fluctuations at the
1025: shock exceed the background ISM fluctuations, any low injection rate
1026: leads to growth of the turbulent magnetic field in the upstream
1027: region, which in turns leads to an increase of the injection rate
1028: \citep{vob03}. The acceleration efficiency, defined as the
1029: fraction of total energy flux crossing the shock that goes into
1030: relativistic particles \citep[see \S 3.3 in][]{bee99}, is $51\%$ in
1031: the case with $\eta_{\mathrm{inj}} = 10^{-3}$ and $7\%$ in the case
1032: with $\eta_{\mathrm{inj}} = 5 \times 10^{-5}$.
1033: 
1034: Note that the form of the azimuthal profiles of $\eta_{\mathrm{inj}}$
1035: and $B_{\mathrm{u}}$ is again arbitrary (here a squared-sine
1036: variation) and just used for illustration purpose.  Our goal is not to
1037: find the exact form of the profiles based on an accurate fit of the
1038: model predictions to the observations, but rather to describe
1039: qualitatively how the acceleration parameters vary along the BW in SN
1040: 1006. The extremum values for $\eta_{\mathrm{inj}}$ were chosen so
1041: that the predicted variations of the ratio of radii
1042: $R_{\mathrm{BW}}/R_{\mathrm{CD}}$ roughly matches the observations. A
1043: maximum (minimum) value of $\eta_{\mathrm{inj}}$ between $10^{-4}$ and
1044: $10^{-2}$  (resp. $10^{-5}$ and $5 \times 10^{-5}$) will not
1045: significantly affect our results concerning the profile of
1046: $R_{\mathrm{BW}}/R_{\mathrm{CD}}$, but differences will be seen in the
1047: absolute intensity in the synchrotron emission. The maximum value of
1048: $B_{\mathrm{u}}$ was adjusted so that it yields an immediate postshock
1049: value, $B_{\mathrm{d}} \simeq 80 \: \mu{\mathrm{G}}$ (\textit{panels
1050: e}), consistent with the value derived from the thickness of the X-ray
1051: synchrotron-emitting rims in SN 1006, under the assumption that this
1052: thickness is limited by the synchrotron losses of the highest energy
1053: electrons \citep[e.g.,][]{ba06}. Finally, we adjusted the minimum
1054: value of $B_{\mathrm{u}}$ in the faint region so that the predicted
1055: and observed radio synchrotron flux are roughly the same; here we show
1056: profiles with a value of $B_{\mathrm{u}} = 3 \: \mu{\mathrm{G}}$,
1057: i.e.,  the typical ISM value.
1058: 
1059: In Figure~\ref{fig_model_k0_var_profiles} (\textit{panel g}), we show
1060: the predicted azimuthal profile of ratio of radii between the BW and
1061: CD (\textit{dashed lines}). The ratio of radii was normalized
1062: (\textit{solid lines}) so that it roughly equals 1.00 in the region of
1063: efficient CR acceleration as in the observations. This normalization
1064: procedure (which we justified in \S \ref{ratio-unity}) works
1065: remarkably well. The variations of the ratio of radii reflects nothing
1066: but the variations of the overall density compression ratio,
1067: $r_{\mathrm{tot}}$ (\textit{panel d}). For low injection rates
1068: ($\eta_{\mathrm{inj}} \leq 6 \times 10^{-5}$), the compression ratio
1069: comes close to the value obtained in the test-particle case (i.e.,
1070: $r_{\mathrm{tot}} = 4$) and as the injection rate increases, the
1071: plasma becomes more compressible with $r_{\mathrm{tot}} \sim 6$.
1072: However, once the magnetic field has been considerably amplified
1073: ($B_{\mathrm{u}} \geq 10 \; \mu\mathrm{G}$), $r_{\mathrm{tot}}$ starts
1074: to slightly decrease due to the heating of the gas by the Alfv\'en
1075: waves in the precursor region \citep[see][]{bee99}.
1076: 
1077: In Figure~\ref{fig_model_k0_var_profiles} (\textit{panel h}), we show
1078: the azimuthal profiles of the synchrotron emission in the radio and
1079: several X-ray energy bands. To obtain these profiles, we first
1080: computed the radial profile of the synchrotron flux (calculated in a
1081: given energy band) at each azimuthal angle. Then, we projected the
1082: flux along the line-of-sight and finally extracted the projected flux
1083: in $30\arcsec$-wide regions behind the BW (assuming a distance of 2
1084: kpc) as we did in the observations of SN 1006 (\S
1085: \ref{subsubsect-azimu-synch-emiss}). In this procedure, we assumed a
1086: spherically symmetric distribution for the emissivity.  This is a
1087: reasonable assumption as long as we restrict the analysis close to the
1088: BW, considering that the bright limbs in SN 1006 are in fact polar
1089: caps, i.e., where the emissivity distribution is axisymmetric.
1090: Finally, we fixed the radio flux in the synchrotron caps to be the
1091: same as in the observations by adjusting the electron-to-proton
1092: density ratio at relativistic energies, $K_{\mathrm{ep}}$.  This is
1093: obtained for a value of $K_{\mathrm{ep}} = 7 \times 10^{-4}$.
1094: 
1095: Relaxing the assumption of the Bohm diffusion (i.e., $k_0$) at each
1096: azimuthal angle significantly modifies the azimuthal profiles of the
1097: synchrotron flux in the X-ray band  (while the radio profile is
1098: unchanged) compared to the model~3 of Figure
1099: \ref{fig_heuristic_model_profiles} in which $k_0$ was always equal to
1100: 1.  This is because an increase of $k_0$ decreases the maximum energy
1101: of electrons, $E_{e,\mathrm{max}}$. Since the synchrotron emission is
1102: very sensitive to the position of the high-energy cutoff in the
1103: electron distribution (i.e., $E_{e,\mathrm{max}}$), all profiles do
1104: not show anymore a plateau in the region of efficient CR acceleration
1105: but are now gradually decreasing (\textit{panel h}). Besides, the
1106: synchrotron brightness in the X-ray bands now decreases faster than in
1107: the radio (\textit{panel i}). This is roughly consistent with what is
1108: observed in SN 1006. We note that in this model, the maximum energy of
1109: both protons and electrons (\textit{panel f}) decrease rather quickly
1110: as we move toward regions of inefficient CR acceleration as opposed to
1111: the previous model~3 with $k_0=1$ where the maximum energy of
1112: electrons remained approximately constant (not shown). Overall, our
1113: results from this model are consistent with those of \cite{rob04}
1114: which were based on a different approach (i.e., measure of the
1115: azimuthal variations of the cutoff-frequency all around SN 1006) using
1116: \textit{XMM-Newton} observations.
1117: 
1118: 
1119: \subsubsection{Other azimuthal profiles for $\eta_{\mathrm{inj}}$, $B_{\mathrm{u}}$ and $k_0$}
1120: 
1121: We have found that a simple model in which the injection rate of
1122: particles in the acceleration process ($\eta_{\mathrm{inj}}$),
1123: amplified magnetic field ($B_{\mathrm{u}}$) and level of turbulence
1124: ($k_0$) were gradually increasing around the shock front was able to
1125: provide predictions in a good agreement with the observations for the
1126: azimuthal variations of the ratio of radii and radio and X-ray
1127: synchrotron fluxes. The input parameters and their azimuthal
1128: variations were chosen based on the physical picture where the
1129: injection of particles and amplification of the magnetic turbulence
1130: are coupled.
1131: 
1132: One can imagine however more complicated scenario and hence more
1133: complicated input profiles for the acceleration parameters
1134: ($\eta_{\mathrm{inj}}$, $B_{\mathrm{u}}$, $k_0$). For instance, the
1135: coupling between the injection and the turbulence which occurs at the
1136: beginning of the acceleration process may end at some point when the
1137: amplitude of the magnetic field fluctuations become so high that their
1138: further growth is prevented by strong dissipation processes
1139: \citep{vob03}. The injection, magnetic field and level of turbulence
1140: may in fact saturate (over some azimuthal range) as we reach regions of
1141: very efficient CR acceleration.  Another example of scenario can be
1142: imagined if the shock velocity is not uniformly distributed along the
1143: shock (as opposed to what we have assumed in our models) due to a
1144: asymmetrical explosion. Such a scenario could explain why the bright
1145: synchrotron-emitting caps in SN 1006 seem to have larger radii than
1146: the faint regions. If so, the injection could be larger in the region
1147: of higher velocity while the level of turbulence and magnetic field
1148: could potentially saturate there. In general, these new scenarios will
1149: not lead to a significant modification of the profile of the ratio of
1150: radii. However, we do expect some changes in the azimuthal profiles of
1151: the radio and X-ray synchrotron flux. They will generally be flatter
1152: in the regions of very efficient CR acceleration.
1153: 
1154: %*******************************************
1155: \section{Discussion\label{sect-discussion}}
1156: %*******************************************
1157: 
1158: \subsection{Polar cap morphology\label{sub-morphology}}
1159: 
1160: The model in which the injection rate of particles, magnetic field and
1161: level of turbulence are all varying as a function of azimuth was
1162: successful in reproducing the overall azimuthal variations of the
1163: radio and X-ray synchrotron flux observed along the BW of SN 1006
1164: (Fig.~\ref{fig_model_k0_var_profiles}). In fact, it is actually
1165: possible to build a 2-D projected map of the radio and X-ray
1166: synchrotron morphology comparable to the radio and X-ray images of SN
1167: 1006 (Figs. \ref{fig_SN1006_3color_xray} and \ref{fig_SN1006_radio}).
1168: This requires us to know the three-dimensional distribution of the
1169: emissivity. Here, we consider polar caps as suggested by \cite{rob04}.
1170: 
1171: To simplify the calculation, we consider that the radial profiles of
1172: the emissivity obtained from the different values of injection,
1173: magnetic field and turbulence level have an exponential form
1174: characterized by a maximum emissivity at the shock and width $\Delta
1175: R$ over which the emissivity decreases (see Appendix
1176: \ref{app-projection}). We also fixed the BW radius, 
1177: $R_{\mathrm{BW}}$, to be the same at each azimuthal angle where the
1178: injection and magnetic field vary, although it slightly depends on the
1179: CR acceleration efficiency. These are good approximations.
1180: 
1181: In the case of the radio synchrotron emission,  the width $\Delta R$
1182: does not depend on the azimuthal angle (from the pole to the equator),
1183: and we have typically $\Delta \tilde{R} \equiv \Delta R /
1184: R_{\mathrm{BW}} = 0.01$. The projected radio morphology is shown in
1185: Figure \ref{fig_projection} (\textit{top-right panel}). It has a
1186: bipolar limb-brightened morphology as is observed in SN 1006. 
1187: 
1188: In the case of the X-ray synchrotron emission, the width $\Delta R$ is
1189: expected first to increase (starting from the equator) as particles
1190: reach higher energies and the emission builds up, and then to slightly
1191: decrease when the effect of radiative losses starts to become
1192: important due to the larger post-shock magnetic field as we move
1193: toward the pole. Because the brightness is much lower at the equator
1194: than at the pole, a reasonable approximation is to consider the width
1195: constant with typically $\Delta \tilde{R} = 2 - 5 \times 10^{-3}$
1196: (\textit{bottom-left panel}) or even a slightly increasing width from
1197: the pole to the equator (\textit{bottom-right panel}). A slightly
1198: increasing width will not change the overall morphology. It still
1199: produces a thin rim whose projected width is decreasing starting from
1200: the pole to the equator. In Figure \ref{fig_projection}, we have
1201: assumed that the emissivity contrast at the BW between the pole and
1202: the equator is always a factor 100. In fact, because the contrast
1203: between the pole and equator is larger in the X-rays than in the
1204: radio, we will obtain a similar morphology but with the size of the
1205: caps slightly reduced in X-ray. In other words, the X-ray synchrotron
1206: emitting rims become geometrically thinner and of smallest azimuthal
1207: extent as we go to higher energy. All these effects would be however
1208: slightly attenuated if we had included the instrumental effects
1209: (particularly the PSF), but overall this is consistent with what is
1210: observed in SN 1006.
1211: 
1212: \subsection{CR acceleration in a partially ionized medium\label{sect-neutral-med}}
1213: 
1214: When trying to delineate the BW, we found that the X-ray
1215: synchrotron-emitting rims and the diffuse filament of H$\alpha$
1216: emission were coincident over some little azimuthal range, notably for
1217: instance at the edge of the eastern cap where the synchrotron emission
1218: starts to turn on (see Fig. \ref{fig_SN1006_3color_xray}). SN 1006 is
1219: one of the rare remnants where this characteristic is clearly
1220: observed, although this likely happens in the Tycho SNR too.
1221: 
1222: In general, H$\alpha$ emission from non-radiative shocks is believed
1223: to arise when the blast wave encounters partially ionized gas. This
1224: points to the presence of neutral atoms in the ambient medium around
1225: SN 1006 \citep{ghw02_SN1006}. Theoretical studies have shown that in
1226: such medium, the scattering Alfv\'en waves should be damped,
1227: henceforth quenching the acceleration of high energy particles
1228: \citep{drd96}. In fact, recent theoretical studies suggest that the
1229: magnetic field could be turbulently amplified even in such medium
1230: \citep{byt05}. The fact that we observe the H$\alpha$ and synchrotron
1231: X-rays to be coincident gives support to this statement. Other
1232: observations of H$\alpha$ emission coexisting with synchrotron X-rays
1233: would be very interesting in that regard.
1234: 
1235: \subsection{Comparison between SN 1006 and Tycho}
1236: 
1237: Like SN 1006, the Tycho SNR is believed to be the remnant of a
1238: thermonuclear explosion and, as such, is expected to develop an
1239: approximately exponential ejecta density profile and to evolve in a
1240: uniform ambient medium. Measurements of the gap between BW and CD have
1241: been done in the Tycho SNR using \textit{Chandra} observations
1242: \citep{wah05}. We comment on the Tycho results in the light of our
1243: results.
1244: 
1245: Because the thin and bright X-ray synchrotron-emitting rims
1246: concentrated at the BW are observed all around Tycho, in contrast to
1247: SN 1006, the BW's location can be easily determined. In addition,
1248: because the X-ray emission from the shocked ejecta is dominant in
1249: Tycho, in contrast to SN 1006, it is also fairly easy to determine the
1250: CD's location. The observed azimuthally-averaged
1251: ratio of radii between the BW and CD was found to be $\sim 1.04$ in
1252: Tycho \citep{wah05}. After correcting, approximately, for effects due
1253: to projection of the highly structured CD, the ratio of radii became
1254: $\sim 1.075$.  This is still a lot smaller than the (unprojected)
1255: ratio of $\sim 1.25$ derived using standard one-dimensional
1256: hydrodynamical simulations with no CR acceleration, at the current age
1257: of Tycho ($430$ yrs).  It was then concluded that efficient CR ion
1258: acceleration was occurring around nearly the entire BW of Tycho.
1259: 
1260: More recently, \citet{cah07} applied a CR-modified self-similar
1261: hydrodynamic model to the radio and X-ray observations of the Tycho
1262: SNR using the observed ratio of BW to CD radii as an important input
1263: constraint.  Using reasonable values for the hydrodynamical parameters
1264: (ambient medium density, SN explosion energy, ejected mass), DSA
1265: parameters (injection efficiency, magnetic field, diffusion
1266: coefficient, electron-to-proton ratio at relativistic energies), and
1267: the distance, a good description of the observational properties at
1268: the blast wave of Tycho was obtained. This detailed study provides a
1269: self-consistent/coherent picture of efficient CR ion acceleration.
1270: 
1271: Regarding SN 1006, we have some good news and some bad (compared to
1272: the case of Tycho, where CR-modified models seem to describe the data
1273: well).  The good news is that there is a significant variation in the
1274: ratio of BW to CD radii as a function of azimuthal angle that is
1275: correlated with the varying intensity of the X-ray and radio
1276: synchrotron emission.  Where the gap is large the synchrotron emission
1277: is faint (or not detected) and where the gap is small the emission is
1278: bright.  This is consistent with an azimuthal variation in the effects
1279: of CR-modification to the remnant dynamics.  The bad news is that in
1280: regions of presumably efficient acceleration (i.e., the bright rims),
1281: the ratio of BW to CD radii approaches small values, very close to 1.
1282: The puzzle is that our CR-modified hydrodynamical models are unable to
1283: produce ratio values less than 1.06, while the specific scenarios
1284: presented above produce ratio values closer to $1.12$. In order
1285: to make progress in our study we decided to introduce an ad-hoc
1286: rescaling of the ratio of radii by decreasing the modeled values at
1287: all angles by a constant factor of $\sim 12\%$.  This procedure
1288: works remarkably well and suggests to us that something else,
1289: unrelated to the CR acceleration process (such as 3-D projection
1290: effects and ejecta clumping), may be responsible for the smaller than
1291: expected gap between the BW and CD in SN 1006.
1292: 
1293: %*******************************************
1294: \section{Conclusion\label{sect-conclusion}}
1295: %*******************************************
1296: 
1297: Using a combination of radio, optical (H$\alpha$) and X-ray images, we
1298: have located the positions of the BW and CD radii along the
1299: southeastern sector of SN 1006. We found that with increasing
1300: azimuthal angles the ratio of radii between the BW and CD increases
1301: from values near unity (within the northeastern synchrotron-emitting
1302: cap) to a maximum of about 1.10 before falling again to values near
1303: unity (in the southwestern synchrotron-emitting cap). These variations
1304: reflect changes in the compressibility of the plasma attributed to
1305: variations in the efficiency of the BW at accelerating CR ions and
1306: give strong support to the overall picture that SNR shocks produce
1307: some fraction of Galactic CRs.  However, at the present time we do not
1308: have a detailed explanation for the apparent overall smallness of the
1309: measured ratios of radii in SN 1006: the minimum value predicted by
1310: our CR-modified self-similar dynamical models is 1.06.  In this study
1311: we simply rescaled the ratio of radii by a constant factor
1312: (independent of azimuthal angle) of $\sim 12\%$ with the expectation
1313: that some process, other than CR acceleration itself, was responsible
1314: for driving the edge of the ejecta closer to the BW all along the rim
1315: of SN 1006. The lack of a definitive astrophysical explanation for
1316: this discrepancy is a significant factor limiting our ability
1317: to understand the CR acceleration process in SN 1006. Further research
1318: into the effects of hydrodynamical instabilities at the CD and ejecta
1319: clumping (two possible explanations for the small BW/CD ratio) would
1320: incidentally provide new insights into the acceleration process.
1321: 
1322: In addition to the azimuthal variations of the ratio of radii between
1323: the BW and CD we also interpreted the variations of the synchrotron
1324: flux (at various frequencies) at the BW using CR-modified hydrodynamic
1325: models.  We assumed different azimuthal profiles for the injection
1326: rate of particles in the acceleration process, magnetic field and
1327: turbulence level.  We found the observations to be consistent with a
1328: model in which these quantities are all azimuthally varying, being the
1329: largest in the brightest regions. Overall this is consistent with the
1330: picture of diffusive shock acceleration. In terms of morphology, we
1331: found that our model was generally consistent with the observed
1332: properties of SN 1006, i.e., a bright and geometrically-thin
1333: synchrotron-emitting rim at the poles and very faint synchrotron
1334: emission at the equator and in the interior.  In the model, the X-ray
1335: synchrotron-emitting rims are geometrically thinner and of smallest
1336: azimuthal extent than the radio rims, which is in broad agreement with
1337: observations.  This is because the most energetic electrons
1338: accelerated at the blast wave lose energy efficiently in the amplified
1339: post-shock magnetic field.  Based on this picture, it would be worth
1340: trying to measure the azimuthal variation of the magnetic field
1341: strength (from the X-ray rim widths) and thereby gain further insight
1342: into the amplification process.
1343: 
1344: 
1345: \acknowledgments It's G.C.-C.'s pleasure to acknowledge
1346: \mbox{J.~Ballet}, \mbox{A.~Decourchelle} and \mbox{D.~C.\ Ellison} for
1347: previous enlightening discussions and work. G.C.-C.\ would like also
1348: to thank \mbox{F.~Winkler} for providing the precious H$\alpha$ image.
1349:  C.B.\ thanks John Blondin for making the VH-1 code available (this
1350: code was used to produce Figure \ref{fig_n0_vs_ratio}). E.M.R.\ is
1351: member of the Carrera del Investigador Cient\'ifico of CONICET,
1352: Argentina. E.M.R.\ was supported by grants PIP-CONICET 6433,
1353: ANPCyT-14018 and UBACYT A055 (Argentina). C.B.\ was supported by NASA
1354: through Chandra Postdoctoral Fellowship Award Number PF6-70046 issued
1355: by the Chandra X-ray Observatory Center, which is operated by the
1356: Smithsonian Astrophysical Observatory for and on behalf of NASA under
1357: contract NAS8-03060. Financial support was also provided by Chandra
1358: grants GO3-4066X and GO7-8071X to Rutgers, The State University
1359: of New Jersey.
1360: 
1361: 
1362: %===========================================================
1363: 
1364: \appendix
1365: 
1366: \section{Projection along the line-of-sight in the case of polar caps\label{app-projection}}
1367: 
1368: Let $r$ be the distance to the center $O$ of a sphere of radius $R_s$,
1369: $\phi$ the azimuthal angle in the $(O\vec{x},O\vec{z})$ plane with $0
1370: \leq \phi \leq 2 \pi$, and $\theta$ the latitude with $- \pi / 2 \leq
1371: \theta \leq \pi / 2$.  The projection of a
1372: spherical emissivity distribution $\mathcal{E}(r,\theta,\phi)$ onto
1373: the plane $(O\vec{x},O\vec{y})$ results in a two-dimensional
1374: brightness distribution:
1375: \begin{equation}\label{brightness}
1376: \mathcal{B}(x,y) = 2 \: \int_{0}^{\ell} \mathcal{E}(r,\theta,\phi) \: dz,
1377: \end{equation}
1378: where $\ell^2 = x_s^2 - x^2$, and $0 \leq x \leq x_s \equiv R_s \:
1379: \cos \theta$ and $0 \leq y \leq R_s$.  At any point of coordinate
1380: $(x,y,z)$, the emissivity is $\mathcal{E}(x,y,z) =
1381: \mathcal{E}(r,\theta,\phi)$ where $r^2 = x^2 + y^2 + z^2$ and $\theta
1382: = \arcsin \left( y / r \right)$ and $\phi = \arccos( x / \sqrt{x^2 +
1383: z^2} )$.  When the emissivity has a symmetry of revolution around the
1384: $O\vec{y}$ axis (e.g., polar caps), there is no dependency on $\phi$.
1385: 
1386: Let us consider a sphere where the emissivity $\mathcal{E}$ is
1387: radially decreasing from the maximum $\mathcal{E}_{\mathrm{max}}$ at
1388: the surface of the sphere with a characteristic width $\Delta R$. In
1389: the case of an exponential decrease and cylindrical symmetry, we have:
1390: \begin{equation}\label{emissivity}
1391:   \mathcal{E}(r,\theta,\phi) = \mathcal{E}_{\mathrm{max}}(\theta) \:
1392:   \exp \left( \frac{r - R_s}{\Delta R(\theta)} \right).
1393: \end{equation}
1394: In the case of polar caps, $\mathcal{E}_{\mathrm{max}}$ will
1395: be larger at the poles ($\theta=\pi / 2$ and $\theta = - \pi / 2$).
1396: 
1397: In Figure \ref{fig_projection}, we show the projected morphology
1398: obtained with the emissivity spatial distribution given by Eq.
1399: (\ref{emissivity}) in the range $0 \leq \theta \leq \pi / 2$ assuming
1400: different values and angular dependencies for the width $\Delta
1401: \tilde{R} \equiv \Delta R / R_s$. In those plots, a squared-sine
1402: variation was assumed for ${\log}_{10} \left(
1403: \mathcal{E_{\mathrm{max}}} \right)$, having the value $0.0$ at the
1404: pole and $-2.0$ at the equator. Decreasing the value at the equator
1405: produces similar rims but with a lower azimuthal extent (i.e.,
1406: smaller polar caps).
1407: 
1408: %===========================================================
1409: 
1410: \newpage
1411: 
1412: %++++++++++++++++++++++++++++++++++++++++++++++++++++
1413: 
1414: % Bibliographie a cet endroit ( style A&A )
1415: 
1416: \bibliographystyle{aa} % A&A style
1417: 
1418: \bibliography{ms} % ne surtout pas mettre le .bib !!!
1419: 
1420: %+++++++++++++++++++++++++++++++++++++++++++++++++++++
1421: 
1422: %############################################## 
1423: 
1424: %          FIGURES
1425: 
1426: %##############################################
1427: 
1428: \clearpage
1429: 
1430: \begin{figure*}[t] 
1431: \centering
1432: \includegraphics[width=14cm]{f1.eps}
1433: \caption{Three-color \textit{Chandra} image of SN 1006: 0.5-0.8 keV
1434: (\textit{red}), 0.8-2 keV (\textit{green}) and 2-4.5 keV
1435: (\textit{blue}). The regions dominated by synchrotron emission
1436: from high-energy relativistic electrons appears in white and are
1437: naturally associated with the BW. The regions dominated by the
1438: thermal emission from the shocked gas (mostly oxygen) appears in red
1439: and are most likely associated with the ejecta. Note that the oxygen
1440: is found slightly ahead of the BW in some places (e.g., east
1441: and south).  Point sources have been removed here so that only the
1442: diffuse emission from the remnant is visible. Images were background
1443: subtracted, corrected for vignetting and slightly smoothed.  The
1444: intensity scaling is square-root with a maximum fixed at $5.0 \times
1445: 10^{-4}$, $5.0 \times 10^{-4}$ and $2.0 \times 10^{-4}$
1446: ph/cm$^2$/s/arcmin$^2$ for the red, green and blue bands,
1447: respectively.} 
1448: \label{fig_SN1006_3color_xray} 
1449: \end{figure*}
1450: 
1451: \begin{figure*}[t] 
1452: \centering
1453: \includegraphics[width=14cm]{f2.eps}
1454: \caption{Radio image of SN 1006 at 1.5 GHz. The
1455: intensity scaling is square-root with a maximum fixed at $1.0$ mJy$/$beam.
1456: The bright elongated source in the western rim is a background radio galaxy.} 
1457: \label{fig_SN1006_radio} 
1458: \end{figure*}
1459: 
1460: \begin{figure*}[t] 
1461: \centering
1462: \includegraphics[width=14cm]{f3.eps}
1463: \caption{H$\alpha$ image of SN 1006.} 
1464: \label{fig_SN1006_Halpha} 
1465: \end{figure*}
1466: 
1467: \begin{figure*}[t] 
1468: \centering
1469: \includegraphics[width=8cm]{f4a.eps}
1470: \includegraphics[width=8cm]{f4b.eps}
1471: \includegraphics[width=8cm]{f4c.eps}
1472: \includegraphics[width=8cm]{f4d.eps}
1473: \caption{Images of SN 1006 in several energy bands. \textit{Top-left}:
1474: Radio image at 1.5 GHz. The maximum intensity was fixed at $0.6$
1475: mJy$/$beam. \textit{Top-right}: H$\alpha$ image with the contour
1476: errors associated with the shock front seen in the southeast
1477: (\textit{red lines}). \textit{Bottom-left}: \textit{Chandra} X-ray
1478: image in the oxygen band (0.5-0.8 keV). The maximum intensity was
1479: fixed at $5.0 \times 10^{-4}$ ph/cm$^2$/s/arcmin$^2$.
1480: \textit{Bottom-right}: \textit{Chandra} image in the mid-energy X-ray
1481: band (0.8-2.0 keV). The maximum intensity was fixed at $6.0 \times
1482: 10^{-4}$ ph/cm$^2$/s/arcmin$^2$. Both X-ray images were background
1483: subtracted, corrected for vignetting and slightly smoothed. In the
1484: radio and X-ray images, we show the contour derived from the H$\alpha$
1485: emission (\textit{red lines}). This contour lies within the contour
1486: errors shown in the H$\alpha$ image. In all the images, the intensity
1487: scaling is linear.}
1488: \label{fig_SN1006_all} 
1489: \end{figure*}
1490: 
1491: \begin{figure*}[t] 
1492: \centering
1493: \includegraphics[width=13cm]{f5.eps} 
1494: \caption{Azimuthal variation of the outer H$\alpha$ emission
1495: (\textit{black lines}) and X-ray emission as obtained in several X-ray
1496: energy bands: 0.5-0.8 keV (\textit{red lines}), 0.8-2 keV
1497: (\textit{green lines}) and 2-4.5 keV (\textit{blue lines}). Regions
1498: where the X-ray synchrotron emission is dominant are indicated by
1499: cross-hatched yellow lines.  In those regions, the fingers
1500: (\textit{stars}) indicate the presence of shocked ejecta found at or
1501: even slightly ahead of the BW. The X-ray contours correspond
1502: to places where the brightness becomes larger than $1.5 \times
1503: 10^{-5}$ ph/cm$^2$/s/arcmin$^2$ in the $0.5-0.8$ keV band, $0.6 \times
1504: 10^{-5}$ ph/cm$^2$/s/arcmin$^2$ in the $0.8-2$ keV band and $0.4
1505: \times 10^{-5}$ ph/cm$^2$/s/arcmin$^2$ in the $2-4.5$ keV band. Note
1506: that the radii measurements are quite sensitive to the contour values,
1507: but the range of variations is contained between the red and blue
1508: lines.}
1509: \label{fig_radius_Ha_xray} 
1510: \end{figure*}
1511: 
1512: \begin{figure*}[t] 
1513: \centering
1514: \includegraphics[width=13cm]{f6.eps} 
1515: \caption{Ratio of the outer H$\alpha$ radius (\textit{black lines} in
1516: Fig.  \ref{fig_radius_Ha_xray}) to the 0.5-0.8 keV X-ray radius
1517: (\textit{red lines} in Fig. \ref{fig_radius_Ha_xray}). Regions where
1518: the X-ray synchrotron emission is dominant are indicated by
1519: cross-hatched yellow lines. Error bars include uncertainties on the
1520: radius derived from the (expanded) H$\alpha$ image and that
1521: done in the low-energy X-rays choosing different
1522: contour values ($2.5 \pm 1.0 \times 10^{-5}$
1523: ph/cm$^2$/s/arcmin$^2$). The large error bars between
1524: $200^\circ-205^\circ$ are due to a low exposure there.}
1525: \label{fig_Rs_o_Rc} 
1526: \end{figure*}
1527: 
1528: \begin{figure*}[t] 
1529: \centering
1530: \includegraphics[width=13cm]{f7a.eps}
1531: \includegraphics[width=13cm]{f7b.eps}
1532: \caption{\textit{Top}: Azimuthal variations of the brightness at the
1533: BW of SN 1006 at several frequencies/energies: 1.5 GHz (\textit{black
1534: lines}), 0.5-0.8 keV (\textit{red lines}), 0.8-2 keV (\textit{green
1535: lines}) and 2-4.5 keV (\textit{blue lines}). The brightness in each
1536: azimuthal bin was calculated between the BW radius and the radius
1537: $30\arcsec$ behind.  For comparison, we show the level of the radio
1538: rms noise level and X-ray background located $1\arcmin$ outside the BW
1539: and extracted from a $30\arcsec$-wide box (\textit{dashed lines}).
1540: \textit{Bottom}: Same but with the curves rescaled roughly to the same
1541: level in the brightest regions and shown with a log scale.}
1542: \label{fig_flux_vs_angle_radio_xrays} 
1543: \end{figure*}
1544: 
1545: \begin{figure*}[t] 
1546: \centering 
1547: \includegraphics[width=13cm]{f8.eps} 
1548: \caption{Ratio of radii between the BW and CD,
1549: $R_{\mathrm{BW}}/R_{\mathrm{CD}}$, as obtained from 1-D numerical
1550: hydrodynamical simulations (assuming no CR acceleration at the BW) and
1551: calculated for a large range of ambient density, $n_0$, for a fixed
1552: SNR age (\textit{dashed lines}: 400 yrs; \textit{solid lines}: 1000
1553: yrs).  We show curves obtained with a 1-D exponential (EXP) ejecta
1554: profile with a kinetic energy of the explosion $E_{51} \equiv E /
1555: 10^{51}$ ergs $=1.0$ (\textit{red lines}), $E_{51} = 0.8$
1556: (\textit{blue lines}), $E_{51} = 1.4$ (\textit{green lines}) and a 1-D
1557: powerlaw (PL7) ejecta profile ($n=7$) with $E_{51} = 1.0$
1558: (\textit{yellow lines}). The cross-hatched domains correspond to the
1559: intersection of the range of ratio of radii observed in SN 1006 and
1560: Tycho ($\leq 1.10$, \textit{vertical solid lines}) and range of
1561: ambient medium density typically encountered in their vicinity
1562: (\textit{horizontal dotted lines}). The 1-D hydrodynamical simulations
1563: are clearly unable to predict a ratio of radii as small as the one
1564: observed in both SNRs. For completeness, we indicate by a
1565: filled circle the value of the ambient density at which the simulated
1566: SNR radius matches the one observed for each model. An angular BW
1567: radius of $14.6\arcmin$ (resp. $256\arcsec$) and a distance to the SNR
1568: of $2.0$ kpc ($2.8$ kpc) were assumed for SN 1006 (resp. Tycho),
1569: resulting in a physical BW radius of $\sim 8.5$ pc (resp. $\sim 3.5$
1570: pc). Larger values for the physical radius (due for instance to a
1571: larger distance to the remnant) would correspond to lower ambient
1572: densities.}
1573: \label{fig_n0_vs_ratio} 
1574: \end{figure*}
1575: 
1576: \begin{figure*}[t] 
1577: \centering 
1578: \includegraphics[width=13cm]{f9.eps} 
1579: \caption{Various models assuming different azimuthal profiles for the
1580: injection rate, $\eta_{\mathrm{inj}}$ (\textit{top}), and far
1581: upstream magnetic field, $B_{\mathrm{u}}$ (\textit{second, solid lines}). Each
1582: model predicts azimuthal variations of the ratio of radii between the
1583: BW and CD, $R_{\mathrm{BW}}/R_{\mathrm{CD}}$ (\textit{third}), and
1584: synchrotron brightness (projected along the line-of-sight) at different 
1585: frequencies (\textit{bottom}).}
1586: \label{fig_heuristic_model_profiles} 
1587: \end{figure*}
1588: 
1589: \begin{figure*}[t] 
1590: \centering \includegraphics[bb=0 60 720 720,clip,width=18cm]{f10.eps}
1591: \caption{\textit{Left column:} azimuthal profiles assumed for the
1592: injection efficiency, $\eta_{\mathrm{inj}}$ (\textit{panel a}), far
1593: upstream magnetic field, $B_{\mathrm{u}}$ (\textit{panel b}) and
1594: diffusion coefficient normalized to the Bohm-limit value, $k_0$
1595: (\textit{panel c}).  \textit{Middle column:} model profiles of the
1596: overall density ($r_{\mathrm{tot}}$) and magnetic field ($r_{B}$)
1597: compression ratios (\textit{panel d}), immediate post-shock value of
1598: the magnetic field, $B_{\mathrm{d}}$ (\textit{panel e}) and maximum
1599: energies (\textit{panel f}).  \textit{Right column:} predicted
1600: (\textit{thick} lines) and observed (\textit{thin} histograms)
1601: azimuthal profiles for the ratio of radii between the BW and CD,
1602: $R_{\mathrm{BW}}/R_{\mathrm{CD}}$ (\textit{panel g, solid lines}) and
1603: line-of-sight projected synchrotron brightness at several frequencies
1604: (\textit{panels h and i}).}
1605: \label{fig_model_k0_var_profiles} 
1606: \end{figure*}
1607: 
1608: \begin{figure*}[t] 
1609: \centering 
1610: \includegraphics[width=6cm]{f11a.eps} 
1611: \includegraphics[width=6cm]{f11b.eps} 
1612: \includegraphics[width=6cm]{f11c.eps} 
1613: \includegraphics[width=6cm]{f11d.eps} 
1614: \caption{Projected morphology predicted in the case of polar caps (see
1615: Appendix \ref{app-projection}) with $\Delta \tilde{R}=5 \times 10^{-2}$
1616: (\textit{top-left}), $\Delta \tilde{R} = 10^{-2}$ (\textit{top-right}),
1617: $\Delta \tilde{R} = 5 \times 10^{-3}$ (\textit{bottom-left}), $\Delta \tilde{R}$
1618: linearly varying from $2 \times 10^{-3}$ at the pole to $8 \times
1619: 10^{-3}$ at the equator (\textit{bottom-right}).  
1620: The radio and the high-energy X-ray
1621: morphology predicted by the model shown in Figure
1622: \ref{fig_model_k0_var_profiles} correspond roughly here to the
1623: top-right and bottom-right panels, respectively. In
1624: all panels, the intensity scaling is square-root.}
1625: \label{fig_projection} 
1626: \end{figure*}
1627: 
1628: %############################################## 
1629: 
1630: %          TABLES
1631: 
1632: %##############################################
1633: 
1634: 
1635: 
1636: \end{document}
1637: