1:
2: % \documentclass{aastex}
3:
4: %% preprint produces a one-column, single-spaced document:
5:
6: \documentclass[12pt,preprint]{aastex}
7:
8: %% preprint2 produces a double-column, single-spaced document:
9:
10: % \documentclass[preprint2]{aastex}
11:
12: \newcommand{\chandra}{{\sl Chandra}}
13: \newcommand{\eg}{e.g.}
14: \newcommand{\etal}{et~al.}
15: \newcommand{\ie}{i.e.}
16: \newcommand{\ltapprox}{\hbox{\raise0.5ex \hbox{$<$}
17: \kern-1.1em \lower0.5ex \hbox{$\sim$}}}
18: \newcommand{\snr}{SN~1006}
19:
20: \slugcomment{ApJ, accepted}
21:
22: \shorttitle{A Curved Synchrotron Spectrum in \snr}
23: \shortauthors{}
24:
25: \begin{document}
26:
27: \title{Evidence of a Curved Synchrotron Spectrum in the Supernova Remnant
28: \snr}
29:
30: \author{G.\ E.\ Allen and J.\ C.\ Houck}
31: \affil{MIT Kavli Institute for Astrophysics and Space Research, Cambridge,
32: MA 02139; gea@space.mit.edu; houck@space.mit.edu}
33:
34: \and
35:
36: \author{S.\ J.\ Sturner}
37: \affil{Astroparticle Physics Laboratory, Code 661, NASA Goddard Space Flight
38: Center, Greenbelt, MD 20771; sturner@swati.gsfc.nasa.gov}
39:
40: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
41:
42: % 1. Abstract
43:
44: \begin{abstract}
45:
46: A joint spectral analysis of some \chandra\ ACIS X-ray data and Molonglo
47: Observatory Synthesis Telescope radio data was performed for 13 small
48: regions along the bright northeastern rim of the supernova remnant \snr.
49: These data were fitted with a synchrotron radiation model. The nonthermal
50: electron spectrum used to compute the photon emission spectra is the
51: traditional exponentially cut off power law, with one notable difference:
52: The power-law index is not a constant. It is a linear function of the
53: logarithm of the momentum. This functional form enables us to show, for the
54: first time, that the synchrotron spectrum of \snr\ seems to flatten with
55: increasing energy. The effective power-law index of the electron spectrum
56: is 2.2 at 1~GeV (\ie, radio synchrotron--emitting momenta) and 2.0 at about
57: 10~TeV (\ie, X-ray synchrotron--emitting momenta). This amount of change in
58: the index is qualitatively consistent with theoretical models of the amount
59: of curvature in the proton spectrum of the remnant. The evidence of
60: spectral curvature implies that cosmic rays are dynamically important
61: instead of being ``test'' particles. The spectral analysis also provides a
62: means of determining the critical frequency of the synchrotron spectrum
63: associated with the highest-energy electrons. The critical frequency seems
64: to vary along the northeastern rim, with a maximum value of
65: $1.1^{+1.0}_{-0.5} \times 10^{17}$~Hz. This value implies that the electron
66: diffusion coefficient can be no larger than a factor of $\sim$4.5--21 times
67: the Bohm diffusion coefficient if the velocity of the forward shock is in
68: the range 2300--5000~km~s$^{-1}$. Since the coefficient is close to the
69: Bohm limit, electrons are accelerated nearly as fast as possible in the
70: regions where the critical frequency is about $10^{17}$~Hz.
71:
72: \end{abstract}
73:
74: \keywords{
75: acceleration of particles ---
76: cosmic rays ---
77: ISM: individual (\snr) ---
78: radiation mechanisms: nonthermal ---
79: supernova remnants ---
80: X-rays: general
81: }
82:
83: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
84:
85: % 2. Introduction
86:
87: \section{Introduction}
88: \label{int}
89:
90: Several observational and theoretical clues support the suggestion that
91: Galactic cosmic rays, up to an energy of 100~TeV \citep{lag83} or more
92: \citep{jok87,vol88,bel01}, are accelerated predominantly in the shocks of
93: supernova remnants. If Galactic supernovae occur at an average rate of one
94: event every 30 years, then an average supernova remnant must transfer about
95: 10\% \citep{dru89} of the initial $10^{51}$~ergs of kinetic energy of the
96: ejecta to cosmic rays. In this case, the cosmic-ray energy density may be
97: large enough to affect the structure of the shock (\ie, cosmic rays may not
98: be mere ``test'' particles).
99:
100: Three consequences of a large cosmic-ray pressure are potentially
101: observable. One consequence is that the ambient material is slowed before
102: it crosses the subshock. This effect reduces the temperature of the shocked
103: gas \citep{che83,ell00,dec00}. An upper limit on the temperature is provided
104: by the well-known relation between the shock speed $v_{s}$ and the postshock
105: temperature of a gas whose ratio of specific heats $\gamma = \frac{5}{3}$:
106: \begin{equation}
107: kT_{i}
108: \le
109: \frac{3}{16} m_{i} v_{s}^2,
110: \label{eqn1}
111: \end{equation}
112: where $k$ is Boltzmann's constant and $m_{i}$ and $T_{i}$ are the mass and
113: immediate postshock temperature, respectively, of particle species $i$.
114: Since kinetic energy is transferred to cosmic rays at the expense of the
115: thermalization of the shocked gas, the equality is appropriate only in the
116: limit that the cosmic-ray energy density is negligible. \cite{hug00}
117: analyzed the transverse motion of the X-ray--emitting material (mostly
118: reverse-shocked oxygen and neon) in the supernova remnant 1E~0102.2$-$7219
119: and inferred a forward-shock velocity $v_{s} =
120: 6200^{+1500}_{-1600}$~km~s$^{-1}$. In this case, the temperature of the
121: shocked protons must be lower than the right-hand side of
122: equation~(\ref{eqn1}) or else the fitted electron temperature will be too
123: low to be explained by Coulomb heating \citep{hug00}. However, the proton
124: temperature is sensitive to the shock velocity ($T \propto v_{s}^{2}$), and
125: a variety of speeds are reported for 1E~0102.2$-$7219. \cite{fla04} report
126: that the \chandra\ High Energy Transmission Grating \ion{Ne}{10} line
127: emission data are consistent with a model that includes radial \ion{Ne}{10}
128: velocities up to $1800 \pm 450$~km~s$^{-1}$. \citet{fin06} analyzed several
129: features in \ion{O}{3} images. The mean transverse velocity of the features
130: is $2000 \pm 200$~km~s$^{-1}$. \citet{eri01} report that the motion of some
131: \ion{O}{3}--emitting material is best described by a radial \ion{O}{3}
132: velocity of about 1800~km~s$^{-1}$. Since these velocity results differ and
133: since none of the measurements provides a direct measure of the velocity of
134: the forward shock, the claim that a significant fraction of the internal
135: energy in 1E~0102.2$-$7219 has been transferred to cosmic rays requires
136: additional support.
137:
138: A second consequence is that the total compression ratio is larger than 4
139: \citep{ell91,ber99}. Since cosmic rays slow the upstream material before it
140: crosses the subshock, the velocity of the shocked material relative to the
141: subshock is reduced. As a result, the contact discontinuity is closer to
142: the subshock than it would be in the absence of a large cosmic-ray pressure.
143: \citet{war05} report that in Tycho's supernova remnant the mean ratio of the
144: radius of the contact discontinuity to the radius of the forward shock is
145: $0.93 \pm 0.02$. This separation corresponds to a total compression ratio $r
146: = 5.1_{-1.0}^{+1.9}$. Similarly, \citet{cas08a} report that for the
147: southeastern rim of \snr\ the ratio varies from a value of
148: $0.91^{+0.03}_{-0.02}$ (\ie, $r = 4.0^{+2.2}_{-0.5}$) between the bright
149: X-ray synchrotron--emitting filaments to a value of 1 (\ie, $r = \infty$)
150: along the filaments. The mean value between the filaments is $0.96 \pm 0.03$
151: (\ie, $r = 9.0^{+25}_{-3.6}$). These authors explore ways in which the
152: location of the contact discontinuity can approach the location of the
153: forward shock, but they cannot explain why these two features curiously
154: appear to be coincident with one another along the filaments.
155: \citet{kse05a}, who also studied \snr, infer a mean compression ratio $r =
156: 5.2$.
157:
158: A third consequence of a large cosmic-ray pressure is that the shock
159: transition region is broadened or ``smeared out'' \citep{ell91,ber99}. Only
160: the subshock has a short transition length. In this case, low-energy cosmic
161: rays, which have relatively small diffusion lengths, experience only a
162: portion of the velocity gradient as they scatter back and forth across the
163: subshock. Higher energy particles, which have relatively large diffusion
164: lengths, experience a larger portion (or all) of the velocity jump. Since
165: the rate of energy gain increases as the velocity difference increases,
166: higher energy particles gain energy faster than lower energy particles. As a
167: result, cosmic-ray spectra do not have power-law distributions. The spectra
168: flatten with increasing energy \citep{bel87,ell91,ber99}. \cite{jon03} and
169: \cite{vin06} report evidence of curvature in the synchrotron spectra of
170: Cas~A and RCW~86, respectively. One possible explanation for such curvature
171: is that the cosmic-ray electrons producing the synchrotron emission have
172: curved spectra.
173:
174: Here we describe the results of an analysis of some radio and X-ray
175: synchrotron data, which suggest that the synchrotron spectrum of \snr\ might
176: be curved. The data, assumptions, and analysis techniques are described in
177: \S\ \ref{dat}. The results of the analysis are discussed in \S\ \ref{dis}.
178: Our conclusions are summarized in \S\ \ref{con}.
179:
180: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
181:
182: % 3. Data and Analysis
183:
184: \section{Data and analysis}
185: \label{dat}
186:
187: X-ray spectra of the filaments along the northeastern rim of \snr\ were
188: obtained by analyzing 68~ks of data from the 2000 July 10--11 {\sl
189: Chandra}\footnote{For more information about the {\sl Chandra X-ray
190: Observatory} refer to the {\sl Chandra} Proposers' Observatory Guide, which
191: is available at http://asc.harvard.edu/proposer/POG/index.html.} observation
192: of the source. The results of previous analyses of the data from this
193: observation were published by \citet{lon03} and \citet{bam03}. During the
194: observation, the telescope was pointed at a location near the middle of the
195: northeastern rim ($\alpha = 15^{\rm h} 03^{\rm m} 51.56^{\rm s}$, $\delta =
196: -41\arcdeg 51\arcmin 18.8\arcsec$; J2000). The X-rays were detected using
197: the Advanced CCD Imaging Spectrometer (ACIS), which includes a $2 \times 2$
198: (ACIS-I) and a $1 \times 6$ (ACIS-S) array of CCDs. Six of these 10
199: detectors (ACIS-I2, -I3, -S1, -S2, -S3, and -S4) were used for the
200: observation. Each $1024\ {\rm pixel} \times 1024\ {\rm pixel}$ CCD has a
201: field of view of $8.4\arcmin \times 8.4\arcmin$. The angular resolution of
202: the {\sl Chandra} mirrors and ACIS varies over the observed portion of \snr\
203: from about $0.5\arcsec$ at the aim point to $20\arcsec$ for a region that is
204: $17.5\arcmin$ off-axis. The on-axis effective area for the mirrors and
205: ACIS-S3 has a maximum of about 720~cm$^{2}$ at 1.5~keV and is greater than
206: 10\% of this value for energies between about 0.3 and 7.3~keV. The
207: fractional energy resolution (FWHM/$E$) between these energies ranges from
208: about 0.4 to 0.03, respectively. The sensitive energy bands and energy
209: resolutions of the other five CCDs used for \snr\ are typically narrower and
210: worse, respectively, than the energy band and resolution of ACIS-S3.
211:
212: The ACIS data were filtered to remove the events that (1) have ${\rm GRADE}
213: = 1$, 5, or 7, (2) have one or more of the STATUS bits set to 1 (except for
214: events that have only one or more of the four cosmic-ray ``afterglow'' bits
215: set), (3) occur on a bad pixel or column, (4) are part of a serial readout
216: streak on ACIS-S4, or (5) occur in the time interval during which the mean
217: background count rate was more than twice the nominal rate (\ie, frames
218: 20515--21294). An image of the 66~ks of ``good'' data is displayed in
219: Figure~\ref{fig1}. The 13, 100~pixel $\times$ 100~pixel ($49 \arcsec \times
220: 49 \arcsec$) boxes are the regions of the bright, northeastern filaments
221: used for the X-ray spectral analysis. The PHA spectra, ARFs, and RMFs for
222: each region were created using version 2.3 of the CIAO\footnote{For more
223: information about the CIAO analysis tools, refer to
224: http://asc.harvard.edu/ciao.} tools {\tt dmextract}, {\tt mkarf}, and {\tt
225: mkrmf}, respectively.
226:
227: The X-ray analysis was performed using data in the energy band 2--7~keV. The
228: higher energy data were excluded because they are dominated by background
229: events. The lower energy data were excluded for three reasons. First, a
230: small, but noticeable, amount of thermal emission is evident at the lower
231: energies. Therefore, inclusion of the lower energy data requires a spectral
232: model that includes a thermal emission component. The extra parameters
233: required to fit the thermal emission make the spectral fits unnecessarily
234: complicated. By ignoring the data below 2~keV, nearly all of the thermal
235: emission is excluded \citep{bam03}. As displayed in Figure~\ref{fig1} (and
236: as originally reported by Koyama et~al.\ 1995), the high-energy nonthermal
237: emission is concentrated along the rim. The second reason is that the ACIS
238: detectors are not well calibrated at energies below about 0.7~keV. The last
239: reason is that the absorption column density can be frozen at some preferred
240: value to simplify the fitting process. Our preferred value is $n_{\rm H} =
241: 6.0 \times 10^{20}$~cm$^{-2}$, which is consistent with the results of
242: \citet{sch96} and \cite{all01}.
243:
244: The radio data used for the spectral analysis include the compilation of
245: flux density measurements listed in Table~\ref{tab1}. These measurements
246: are based on analyses of the emission from the entire supernova remnant. In
247: the cases where no flux density uncertainties are reported, the
248: uncertainties are assumed to be 10\% of the reported flux densities. The
249: results of the spectral analysis are insensitive to the actual fraction used
250: at least for fractions in the range 5\%--20\%. The values listed in
251: Table~\ref{tab1} are plotted in Figure~\ref{fig2}. The dashed line in this
252: figure is the result obtained when only the radio data are fitted with a
253: power-law model: $S(\nu) = 17.6^{+6.3}_{-4.8}[\nu/(1~{\rm
254: GHz})]^{-\alpha}$~Jy, where $\alpha = 0.60^{+0.08}_{-0.09}$. Note that the
255: uncertainties for the power-law index and normalization, which are quoted at
256: the 90\% confidence level, are not independent. This result is an update of
257: and consistent with the result reported by \citet{all01}: $S(\nu) = 17.9 \pm
258: 1.1 [\nu / (1\ {\rm GHz})]^{-\alpha}$~Jy, where $\alpha = 0.57 \pm 0.06$.
259: The dotted line in Figure~\ref{fig2} is a power-law model with the
260: parameters reported by \citet{gre06}: $S(\nu) = 19[\nu/(1~{\rm
261: GHz})]^{-0.6}$~Jy.
262:
263: The radio spectrum for the entire remnant is used to construct radio spectra
264: for each region because there is little evidence of radio spectral
265: variability. For example, a comparison of the images of \citet{ste77} and
266: \citet{gar65} suggests that the slopes of the 0.408--2.7~GHz radio spectra
267: of the northeastern and southwestern rims are about the same. It is possible
268: that the slope may vary along each rim, in which case the best-fit amounts
269: of curvature for individual regions may be inaccurate, but the mean amount
270: of curvature for the 13 regions is relatively insensitive to such
271: variations. Two sets of radio data were used with each X-ray spectrum. Both
272: sets are obtained by multiplying the remnant-integrated flux densities and
273: uncertainties of Table~\ref{tab1} by scaling factors $\zeta$. One set of
274: factors (Tables~\ref{tab2} and \ref{tab3}) is obtained by dividing the
275: 843~MHz flux densities for the 13 regions marked with black squares in
276: Figure~\ref{fig3}, from the Molonglo Observatory Synthesis Telescope (MOST),
277: by the 843~MHz flux density for the entire remnant. These regions are the
278: same ones used to obtain the X-ray spectra (Fig.~\ref{fig1}). The second
279: set of factors (Tables~\ref{tab4} and \ref{tab5}) is obtained in the same
280: way, except that the flux densities for the 13 red squares are used instead
281: of the flux densities for the black squares. The red squares are the regions
282: where the flux densities peak. That is, the flux density along a line from
283: the center of the remnant through the center of a black square peaks at the
284: location of the center of the corresponding red square. Although both sets
285: of flux densities represent local averages because the half-power beamwidth
286: associated with the 843~MHz MOST image ($44'' \times 66''$; see
287: Fig.~\ref{fig3}) is comparable to the size of the regions ($49'' \times
288: 49''$), they are reasonable estimates of the minimum ({\sl black}) and
289: maximum ({\sl red}) flux densities for the 13 regions. Therefore, using
290: these two sets of radio spectra provides a means of obtaining upper
291: (Table~\ref{tab2}) and lower (Table~\ref{tab4}) limits on the amount of
292: curvature for each region.
293:
294: Version 1.4.8 of the spectral fitting package ISIS\footnote{For more
295: information about ISIS, see http://space.mit.edu/cxc/isis.} \citep{hou00}
296: was used to fit the data for each region with a model that includes a
297: synchrotron radiation component for the X-ray and radio data and an
298: interstellar absorption component for the X-ray data (the XSPEC model {\tt
299: wabs}, with the relative abundances of Anders \& Grevesse 1989). Since the
300: 2--7~keV emission is dominated by nonthermal emission, no thermal X-ray
301: component is included.
302:
303: The synchrotron spectra are based on a nonthermal electron spectrum of the
304: form
305: \begin{equation}
306: \frac{dn}{dp}
307: =
308: A \left( \frac{p}{p_{0}} \right)^{-\Gamma + a \log
309: \left( \frac{p}{p_{0}} \right)}
310: {\rm exp} \left( \frac{p_{0} - p}{p_{m}} \right),
311: \label{eqn2}
312: \end{equation}
313: where $n$ is the electron number density, $p = \gamma m v$ ($\gamma$, $m$,
314: and $v$ are the Lorentz factor, rest mass, and velocity of a particle,
315: respectively), $A$ is the number density at $p = p_{0}$ (in units of
316: $p_{0}^{-1}$ cm$^{-3}$), $p_{0} = 1$~GeV~$c^{-1}$ ($c$ is the speed of
317: light), $\Gamma$ is the differential spectral index at $p = p_{0}$, $a$ is
318: the spectral ``curvature,'' the logarithm is base 10, and $p_{m}$ is the
319: exponential cutoff (or ``maximum'') momentum. This form is the same as the
320: standard power law with an exponential cutoff, except for the term $a \log
321: \left( p / p_{0} \right)$. Introduction of this term produces an effective
322: spectral index, $\Gamma_{\rm eff} = \Gamma - a \log \left( p / p_{0}
323: \right)$, that is a linear function of the logarithm of the momentum. For
324: example, at momenta $p$ of $10^{0} p_{0}$, $10^{1} p_{0}$, $10^{2} p_{0}$,
325: and $10^{3} p_{0}$, the effective differential spectral indices are
326: $\Gamma$, $\Gamma - a$, $\Gamma - 2a$, and $\Gamma - 3a$, respectively. If
327: $a > 0$, then the spectrum flattens with increasing momentum. If $a < 0$,
328: then the spectrum steepens with increasing momentum. If $a = 0$, then the
329: spectrum has no curvature and equation~(\ref{eqn2}) reduces to the standard
330: power law with an exponential cutoff. An exponential cutoff is used in
331: equation~(\ref{eqn2}) for simplicity. Note that \citet{ell01},
332: \citet{uch03}, \citet{laz04}, and \cite{zir07a} use a more general form for
333: the cutoff (exp$[- \left( p / p_{m} \right)^{s} ]$, where $s = \case{1}{4}$,
334: $\case{1}{2}$, 1, or 2) and that \citet{pro04} performed a detailed study of
335: the shape of the cutoff for particles experiencing diffusive shock
336: acceleration.
337:
338: The synchrotron spectrum is computed by evaluating the following integral
339: expression to obtain the differential photon flux\footnote{For more
340: information about this model, see \citet{hou06}.} (in SI units):
341: %
342: \begin{equation}
343: \frac{dF}{d(h\nu)}
344: =
345: \frac{V_{S}}{4 \pi d^{2}}
346: \frac{\sqrt{3} e^{3} B}{2 \epsilon_{0} m c h^{2} \nu}
347: \int dp \frac{dn}{dp}
348: \int_{0}^{\pi} d\theta \sin^{2}\!\theta
349: \frac{\nu}{\nu_{c}}
350: \int_{\nu / \nu_{c}}^{\infty} dx K_{5/3}
351: \left( x \right),
352: \label{eqn02}
353: \end{equation}
354: %
355: where $V_{S}$ is the volume of the synchrotron-emitting region, $d$ is the
356: distance of the source from Earth, $e$ is the unit of electric charge, $B$
357: is the total magnetic field strength (not $B_{\perp}$), $\epsilon_{0}$ is
358: the permittivity of free space, $h$ is Planck's constant, $\nu$ is the
359: frequency of an emitted photon, $\theta$ is the pitch angle between the
360: electron momentum and magnetic field vectors, $K_{5/3}$ is an irregular
361: modified Bessel function of the second kind, and
362: %
363: \begin{equation}
364: \nu_{c}
365: =
366: \frac{3 e}{4 \pi m^{3} c^{4}}
367: \left( p^{2}c^{2} + m^{2}c^{4} \right)
368: B \sin \theta
369: \label{eqn03}
370: \end{equation}
371: %
372: is the critical frequency of the synchrotron spectrum produced by an
373: electron with a momentum $p$. The frequency at which a synchrotron power
374: spectrum peaks is $\nu_{p} = 0.286 \nu_{c}$. Hereafter, only the critical
375: frequency $\nu_{c}$ is used.
376:
377: We refer to the critical frequency associated with electrons that have a
378: momentum equal to $p_{m}$ as the ``cutoff critical frequency'' or just the
379: cutoff frequency to emphasize that it is associated with electrons at the
380: cutoff momentum of the electron spectrum [$\nu_{m} \equiv \nu_{c} (p =
381: p_{m})$]. The distribution of pitch angles $\theta$ is expected to be
382: nearly isotropic because the forward-shock speed (2300--5000~km~s$^{-1}$;
383: Laming \etal\ 1996; Dwarkadas \& Chevalier 1998; Ghavamian \etal\ 2002) is
384: much less than the speed of light. In this case, the mean value of $\sin
385: \theta$ in equation~(\ref{eqn03}) is about $\pi / 4$,
386: \begin{equation}
387: \nu_{m}
388: =
389: 1.26 \times 10^{17}
390: \left( \frac{p_{m}}{10\ {\rm TeV}\ c^{-1}} \right)^{2}
391: \left( \frac{B}{100\ \mu{\rm G}} \right)\ {\rm Hz},
392: \label{eqn04}
393: \end{equation}
394: %
395: and
396: %
397: \begin{equation}
398: h \nu_{m}
399: =
400: 0.522
401: \left( \frac{p_{m}}{10\ {\rm TeV}\ c^{-1}} \right)^{2}
402: \left( \frac{B}{100\ \mu{\rm G}} \right)\ {\rm keV}.
403: \label{eqn05}
404: \end{equation}
405:
406: During the spectral fitting process, the right-hand side of
407: equation~(\ref{eqn02}) is computed each time the spectral fitting program
408: changes the parameters of the model. The complete set of spectral
409: parameters includes the spectral index $\Gamma$, the curvature parameter
410: $a$, the cutoff momentum $p_{m}$, the total magnetic field strength $B$, the
411: normalization constant $N_{S} = A V_{S} / 4 \pi d^{2}$, and the absorption
412: column density $n_{\rm H}$. The spectral index is determined by the radio
413: and X-ray data. The curvature is sensitive to the relative radio and X-ray
414: fluxes. Since only two of the three parameters $p_{m}$, $B$, and $N_{S}$
415: are independent, $p_{m}$ was fixed and the cutoff frequency $\left( \nu_{m}
416: \propto p_{m}^{2} B;\ {\rm eq.~[\ref{eqn04}]} \right)$ is reported here
417: instead of the magnetic field strength. This frequency is sensitive to the
418: shape of the X-ray spectrum. The same normalization is used for both the
419: X-ray and radio data for a given region because the nonthermal X-ray and
420: radio emission from the region is assumed to be produced by a common
421: population of electrons. The normalization $N_{S}$ is determined by the
422: radio data. The absorption column density $n_{\rm H}$ was fixed at the
423: value $6.0 \times 10^{20}$~cm$^{-2}$ \citep{sch96,all01}.
424:
425: Four spectral fits were performed for each region. The first pair are
426: identical to one another except that the curvature parameter is a free
427: parameter for one (Table~\ref{tab2}) and is fixed at a value of zero (\ie,
428: no spectral curvature) for the other (Table~\ref{tab3}). In both cases, the
429: radio spectra are normalized to the flux densities in the regions that are
430: cospatial with the X-ray regions (\ie, the regions marked with black squares
431: in Fig.~\ref{fig3}). Since the cospatial radio spectra represent lower
432: limits on the radio fluxes, the amounts of curvature listed in
433: Table~\ref{tab2} are upper limits. The second pair of fits
434: (Tables~\ref{tab4} and \ref{tab5}) is identical to the first pair except
435: that the radio spectra are normalized to the flux densities in the ``peak''
436: regions (\ie, the regions marked with red squares in Fig.~\ref{fig3})
437: instead of the cospatial regions. Since the peak radio spectra represent
438: upper limits on the radio fluxes, the amounts of curvature listed in
439: Table~\ref{tab4} are lower limits. For each of the four sets of fits, the
440: best-fit values of the spectral index vary little from region to region,
441: because the same radio spectral shape was used for each region. The
442: best-fit model for region~6, the region with the largest number of counts,
443: is plotted in Figures~\ref{fig2} and \ref{fig4}.
444:
445: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
446:
447: % 4. Discussion
448:
449: \section{Discussion}
450: \label{dis}
451:
452: It is not immediately obvious whether the cospatial or peak radio fluxes
453: yield results that are more accurate, because the X-ray and radio
454: synchrotron morphologies are not the same. For example, along $49''$-wide
455: strips that pass through the 13 X-ray regions toward the center of the
456: remnant, one finds that the X-ray emission peaks at the locations of the
457: black boxes in Figure~\ref{fig3} and that the radio emission peaks at the
458: locations of the red boxes. There are at least three possible explanations
459: for this difference. One explanation is that the TeV electrons that produce
460: the X-ray synchrotron emission suffer significant synchrotron losses as they
461: diffuse downstream from the shock. These losses deplete the highest-energy
462: end of the electron spectrum, causing the X-ray synchrotron emission to
463: decline in the downstream region. As a result, the X-ray peak is relatively
464: narrow and confined to a region near the shock. Since synchrotron losses are
465: negligible for the GeV electrons that produce the radio emission, the radio
466: emission is not attenuated in the same manner. The radio peaks can be
467: broader and farther from the shock than the X-ray peaks. Nevertheless, there
468: is no reason, on these grounds, to use different radio and X-ray extraction
469: regions, because the radio and X-ray synchrotron emission for a region is
470: produced by the same population of electrons. The radio emission is
471: produced by the GeV portion of the electron spectrum and the X-ray emission
472: is produced by the TeV segment. Synchrotron losses do not introduce a
473: spatial separation between the GeV and TeV electrons.
474:
475: Another explanation for the different X-ray and radio morphologies does
476: involve an apparent (not actual) spatial separation. The X-ray and radio
477: instruments do not yield images that have the same spatial resolution. The
478: size of the spatially dependent point-spread function of the {\sl Chandra}
479: X-ray telescope is much smaller than the size of the $49\arcsec \times
480: 49\arcsec$ boxes and, hence, is neglected. However, the size of the
481: beamwidth for the MOST image ($44\arcsec \times 66\arcsec$) is not
482: negligible, because it is comparable to the size of the boxes. The
483: implications of the difference in the qualities of the X-ray and radio
484: images are illustrated in Figure~\ref{fig5}. The black histogram in the top
485: panel shows the radial X-ray profile for region~6. The red curve in this
486: panel is obtained by using a Gaussian function to smooth the histogram to
487: the beamwidth of the radio instrument. Note that the peak of the red curve
488: is about $18\arcsec$ downstream from the peak of the histogram, which
489: suggests that the radio extraction region should be shifted toward the radio
490: peak by the same amount. Similarly small shifts are obtained for the other
491: 12 regions and for a simple model of a uniformly emitting spherical shell
492: (Fig.~\ref{fig5}, {\sl bottom}).
493:
494: A third explanation is that the bright radio-emitting region has expanded in
495: the 17 intervening years between the time when the MOST data were obtained
496: and the time of the \chandra\ observation. Using the mean radio expansion
497: result of $0.049{\rm \%} \pm 0.014$\%~yr$^{-1}$ \citep{mof93a} suggests that
498: the radio peak has shifted about $7.6'' \pm 2.2''$. To compensate for this
499: expansion, the radio extraction region should be shifted toward the radio
500: peak by the same amount. Since this shift and the apparent shift described
501: above are small compared with the size of the extraction region, the
502: following discussion focuses primarily on the results obtained using the
503: cospatial radio fluxes.
504:
505: \subsection{Curvature}
506: \label{curv}
507:
508: While the values of $\chi^{2}$ per degree of freedom (dof) are acceptable
509: for each of the four sets of fits, a comparison of Tables~\ref{tab2} and
510: \ref{tab3} (and of Tables~\ref{tab4} and \ref{tab5}) reveals that the values
511: are uniformly lower when curvature is included as a free parameter. For
512: example, Figure~\ref{fig6} shows the 1, 2, and 3~$\sigma$ confidence
513: contours for region~6 in the parameter space defined by the electron
514: spectral index $\Gamma$ and the spectral curvature $a$. If the cospatial
515: radio flux density is used, the results for this region suggest that an
516: uncurved spectrum (\ie, the dashed line at $a = 0 $) can be excluded at
517: about the 2.7~$\sigma$ confidence level. Similar results, with varying
518: degrees of statistical significance, are obtained for the other 12 regions.
519: Table~\ref{tab2} lists the probability $P_{\Delta \chi^{2}}$ that the
520: reduction in the value of $\chi^{2}$ when curvature is included as a free
521: parameter is due to chance. The majority of the probabilities are less than
522: 1\%. Some are much less. The probabilities obtained using the F-test
523: ($P_{F}$) are within a factor of 2 of the values listed in Table~\ref{tab2},
524: except for regions 1, 2, 3, 5, 6, and 7, where $P_{F} = 0.058$, 0.0035, $9.9
525: \times 10^{-6}$, 0.0013, 0.0030, and 0.048, respectively. The relatively
526: large discrepancies between $P_{\Delta \chi^{2}}$ and $P_{F}$ for these
527: regions arise because the F-test depends on the value of $\chi^{2}/{\rm
528: dof}$ in addition to the value of $\Delta \chi^{2}$. Note that the values of
529: $\chi^{2}/{\rm dof}$ are significantly less than 1 for regions~1, 2, and 3
530: and that regions~5, 6, and 7 have the largest values of $\chi^{2}/{\rm
531: dof}$. Since the values of $P_{\Delta \chi^{2}}$ are typically larger (\ie,
532: less significant) than the values of $P_{F}$, only the values of $P_{\Delta
533: \chi^{2}}$ are listed in Table~\ref{tab2}. Collectively, these probabilities
534: show that the spectra are significantly better fitted by a curved
535: synchrotron model than an uncurved one.
536:
537: The strength of this evidence depends on the location used to obtain the
538: radio flux density. The radio flux density for a region increases from the
539: cospatial location to the peak location. Since the appropriate location to
540: use lies between these two, the cospatial and peak flux densities represent
541: the lower and upper limits, respectively, on the radio flux density for the
542: region. The amount of curvature depends on the relative X-ray and radio
543: synchrotron fluxes. Therefore, the relatively small cospatial flux density
544: yields an upper limit on the amount of curvature and the peak flux density
545: yields a lower limit. Although the curvature values in Table~\ref{tab4} are
546: lower limits, even these results favor a curved synchrotron model. Yet, as
547: noted above, the upper limit values in Table~\ref{tab2} are expected to be
548: closer to the actual amount of curvature because the appropriate location to
549: use for the radio flux density is much closer to the cospatial location than
550: the peak location.
551:
552: The evidence of curvature cannot be attributed to an instrumental effect. It
553: must be associated with the synchrotron spectrum of the source. For
554: example, the need for curvature could be eliminated if the effective area of
555: the X-ray detector system were too low, but the area would have to be in error
556: by an implausibly large factor of about 4.6. The use of a curved source
557: model is also compelling in the sense that the values of the best-fit
558: indices and curvature parameters are consistent with the expected values.
559: One expectation is that the best-fit index should be consistent with the
560: radio spectral index. That is, the best-fit model should fit the radio
561: spectrum instead of simply intersecting it. The mean value of the index
562: $\Gamma = 2.221 \pm 0.013$ if curvature is included as a free parameter
563: (Table~\ref{tab2}). This value is consistent with the index obtained if only
564: the radio data are fitted with a power-law model ($\Gamma = 2 \alpha + 1 =
565: 2.20^{+ 0.16}_{- 0.18}$; Fig.~\ref{fig2}). Yet this result is not
566: surprising, because the fitted value of the index is determined largely by
567: the shape of the radio spectrum. For comparison, the mean value of the index
568: $\Gamma = 2.031 \pm 0.017$ if an uncurved model is used (Table~\ref{tab3}).
569: This value differs significantly from 2.22 because the uncurved index,
570: unlike the curved index, is sensitive to the relative radio and X-ray
571: fluxes. While this result seems to be closer to the case of having the model
572: simply intersect the radio spectrum, an index of 2.03 is still marginally
573: consistent with the radio data because the radio data have relatively large
574: uncertainties. Otherwise, the radio data might provide a means of
575: discriminating between a curved and an uncurved model.
576:
577: Another expectation is that the value of the curvature parameter should be
578: positive. That is, the spectrum should flatten with increasing energy
579: \citep{bel87,ell91,ber99}. The results in Table~\ref{tab2} show that the
580: best-fit values are greater than zero for all 13 regions. If the spectra
581: are not curved, then the use of a curvature parameter may not be physically
582: meaningful and one might expect roughly half of the values to be negative.
583: If either sign is equally likely, then the chance probability that all 13
584: values have the same sign is 0.00024 $(2^{-12})$.
585:
586: A third expectation is that the highest-energy electrons, the ones that
587: experience the full strength of the shock and that produce the X-ray
588: synchrotron emission, should have an effective spectral index of 2
589: \citep{ber02}. If equation~(\ref{eqn2}) is an accurate representation of the
590: shape of the electron spectrum, then the effective spectral index in the
591: X-ray band $\Gamma_{X} = \Gamma - a \log \left (p_{X} / p_{0} \right)$,
592: where $p_{X}$ is the flux-weighted mean momentum of the electrons that
593: produce the X-ray synchrotron radiation. To compute $p_{X}$ from the fitted
594: value of $\nu_{m}$ requires knowledge of the magnetic field strength
595: (eq.~[\ref{eqn04}]). Since the results of our analysis do not yield an
596: estimate of the field strength, we assume that $p_{X} = 10$~${\rm TeV}\
597: c^{-1}$. (The results are rather insensitive to this assumption as long as
598: $p_{X}$ is within an order of magnitude of 10~${\rm TeV}\ c^{-1}$.) In this
599: case, $\Gamma_{X} = 2.005 \pm 0.027$. This index and the index obtained
600: using the uncurved model ($\Gamma = 2.031^{+0.017}_{-0.016}$) are both
601: consistent with the idea that the highest-energy electrons in \snr\ have a
602: spectral index of 2.0.
603:
604: If the effective spectral index of the X-ray synchrotron--producing
605: electrons $\Gamma_{X} = 2$ (aside from the exponential cutoff), then the
606: expression for the effective spectral index can be inverted to obtain an
607: estimate of the expected amount of curvature as a linear function of the
608: spectral index $\Gamma_{r}$:
609: %
610: \begin{equation}
611: a_{\rm exp}
612: =
613: \frac{\Gamma_{r} - 2}{\log \left( p_{X} / p_{r} \right)}
614: \label{eqn07}
615: \end{equation}
616: %
617: where $\Gamma_{r}$ is the effective spectral index in the radio band and
618: $p_{r}$ is the flux-weighted mean momentum of the electrons that produce the
619: radio synchrotron emission. Although the value of $p_{X} / p_{r}$ is
620: unknown, $p_{X} / p_{r} = \left( \nu_{X} / \nu_{r} \right)^{1/2}$, where
621: $\nu_{X}$ and $\nu_{r}$ are the critical frequencies associated with $p_{X}$
622: and $p_{r}$, respectively. Since the X-ray data used here span the range
623: from 2 to 7~keV, $\nu_{X} \approx 9.0 \times 10^{17}$~Hz, the logarithmic
624: mid-point of this band. Similarly, $\nu_{r} \approx 660$~MHz, the
625: logarithmic mid-point of the range from 86~MHz to 5~GHz (Table~\ref{tab1}).
626: These estimates yield $\log \left( p_{X} / p_{r} \right) = 4.57$. This value
627: is rather insensitive to the uncertainties in $\nu_{X}$ and $\nu_{r}$. For
628: example, a change of a factor of 2 in the ratio $\nu_{X} / \nu_{r}$
629: corresponds to a change of only 3.3\% in $\log \left( p_{X} / p_{r}
630: \right)$. As shown in Figure~\ref{fig6} ({\sl dotted line}), a relation of
631: the form $a_{\rm exp} = \left( \Gamma_{r} - 2 \right) / 4.57$ is consistent
632: with the confidence contours for region~6. This result is remarkable
633: because the contours in Figure~\ref{fig6} are sensitive to the relative
634: X-ray and radio fluxes, while equation~(\ref{eqn07}) has no such dependence.
635: Had the relative normalizations been significantly different, it is unlikely
636: that the contours in Figure~\ref{fig6} would have been consistent with the
637: dotted line. If $\Gamma_{r}$ is accurately represented by the best-fit index
638: (\ie, if $p_{r} \approx 1$~${\rm GeV}\ c^{-1}$) and if the cospatial radio
639: fluxes are used, then $a_{\rm exp} = (2.221 - 2) / 4.57 = 0.048$. This
640: expected value is consistent with the fitted value ($0.054 \pm 0.006$) at
641: the 90\% confidence level. If the peak radio fluxes are used, then $a_{\rm
642: exp} = (2.198 - 2) / 4.57 = 0.043$, which is less consistent with the fitted
643: value ($0.033^{+0.007}_{-0.008}$). Like the expected radio flux, the
644: expected amount of curvature lies between the cospatial and peak values. If
645: the uncertainty associated with the radio flux leads to an uncertainty in
646: the amount of curvature comparable to the statistical uncertainty, then the
647: amount of curvature is most likely $0.05 \pm 0.01$.
648:
649: Not only is the best-fit amount of curvature consistent with
650: equation~(\ref{eqn07}), it is also consistent with the amount of curvature
651: predicted for the proton (not electron) spectrum of \snr. For example,
652: Figure~\ref{fig7} shows the best-fit electron spectra for region~6. The
653: spectra are not curved below $p = 1\ {\rm GeV}\ c^{-1}$, which was true
654: during the fitting process, because the available data poorly constrain the
655: shape of the electron spectrum in this momentum range. The top pair of
656: dotted and dot-dashed curves are predictions described by \cite{ell00} for
657: the proton (not electron) spectrum of \snr. These spectra have been
658: normalized to match the solid black curve at $E = 0.9$~GeV. Since
659: \cite{ell00} computed their spectrum assuming a spectral index $\Gamma =
660: 2.0$, it is not appropriate to compare the solid curve with the upper pair
661: of dotted and dot-dashed curves. However, the lower pair of dotted and
662: dot-dashed curves have been multiplied by the energy-dependent factor $[E /
663: (0.9\ {\rm GeV})]^{-0.2}$ to match the best-fit spectral index $\Gamma =
664: 2.2$. Aside from a possible difference in the assumed value of the cutoff
665: momentum, the lower pair of theoretical curves match the solid black curve
666: remarkably well.
667:
668: The evidence of curvature is based on a set of assumptions. One assumption
669: is that the electron spectrum has the form of equation~(\ref{eqn2}). This
670: functional form excludes the effects of a synchrotron cooling break
671: \citep{vol05}. If the magnetic field strength is as large as 150~$\mu$G,
672: then synchrotron cooling may significantly steepen the highest-energy end of
673: the electron spectrum \citep{kse05a}. Since the fits are sensitive only to
674: the net amount of curvature between radio and X-ray frequencies, inclusion
675: of a cooling break would require a larger amount of curvature (\ie, more
676: flattening than the amount of curvature reported here) to compensate for the
677: spectral steepening of the cooling break. Another assumption is that the
678: shape of the radio spectrum in each of the 13 regions is the same as the
679: shape of the radio spectrum of the entire remnant. If this assumption is
680: invalid, then the best-fit values of the spectral index and curvature may be
681: inaccurate for any given region. Yet, the mean values of the index and
682: curvature are insensitive to such spatial variations.
683: %
684: A third assumption is that the shape of the synchrotron spectrum is uniform
685: inside each of the 13 regions. Some young remnants, such as Cas~A
686: \citep{and96} and Kepler \citep{del02}, exhibit evidence of radio spectral
687: variations from one region to another. Therefore, the combination of the
688: synchrotron spectra of several small-scale features might naturally yield a
689: curved composite spectrum even if the synchrotron spectrum for each feature
690: is just a power law. Small extraction regions were used to minimize this
691: problem.
692: %
693: A fourth assumption is that our estimates of the fraction of the 843~MHz
694: flux from each region are accurate. As discussed above, the size of the
695: beamwidth leads to some uncertainty about the proper radio fluxes to use for
696: each region. Yet, even the worst-case scenario, the results obtained using
697: the peak radio fluxes, requires some spectral curvature.
698: %
699: High-quality radio and X-ray spectra for many, small, spatially resolved
700: features in the remnant would eliminate the need for many of these
701: assumptions. Unfortunately, these kinds of data are not available for \snr.
702:
703: Evidence of curvature in the synchrotron spectra of supernova remnants has
704: been reported previously. \citet{rey92} analyzed the remnant-integrated
705: radio spectra of Kepler, Tycho, and \snr. They find hints of curvature for
706: Tycho and Kepler. Unfortunately, a careful study of the radio emission from
707: Kepler \citep{del02} reveals intrinsic variations in the radio spectral
708: index from region to region. Therefore, it is not possible to dismiss the
709: idea that the evidence of curvature in the remnant-integrated spectrum is
710: the result of the combination of several power-law spectra with different
711: spectral indices. \citet{jon03} minimized this problem for Cas~A by
712: analyzing the radio-to-infrared synchrotron spectra of small, selected
713: features in the remnant. The results of their analysis indicate that the
714: spectra of these regions flatten with increasing energy. We find that their
715: spectra can be fitted with our curved spectral model if the curvature
716: parameter $a = 0.06 \pm 0.01$. It is interesting that this value is
717: consistent with the mean amount of curvature for \snr\ ($0.05 \pm 0.01$).
718: Based on an analysis of X-ray data for a small feature on the northeastern
719: shell of RCW~86, \citet{vin06} report evidence of curvature in the
720: synchrotron spectrum of this remnant. This claim would be strengthened if
721: the data were fitted with a model that includes the spectral index as a free
722: parameter and if there were similar evidence for more than one feature in
723: the remnant.
724:
725: The evidence of curvature suggests that the pressure exerted by nonthermal
726: particles is large enough to modify the structure of the shock. The
727: cosmic-ray pressure is most likely dominated by nuclei, not electrons. For
728: example, analyses of the synchrotron spectrum of \snr\ have been used to
729: estimate the total energy of the nonthermal electrons in the remnant. These
730: estimates, which depend on the assumed value of the mean magnetic field
731: strength, include $5 \times 10^{47}$~ergs (for $B = 150$~$\mu$G; Berezhko
732: \etal\ 2002), $9 \times 10^{47}$~ergs (for $B = 40$~$\mu$G; Allen \etal\
733: 2001), and $7 \times 10^{48}$~ergs (for $B = 3$~$\mu$G; Dyer \etal\ 2001).
734: The remnant was recently observed with the H.E.S.S.\ gamma-ray telescope
735: \citep{aha05a}. If the upper limits on the TeV gamma-ray flux reported by
736: \citet{aha05a} are accurate (cf.\ Tanimori \etal\ 1998), then a combination
737: of these constraints and the measured synchrotron flux can be used to show
738: that the mean magnetic field strength must be at least 25~$\mu$G
739: \citep{aha05a}. In this case, the total energy in cosmic-ray electrons is
740: probably no larger than about $10^{48}$~ergs. Since this energy is a small
741: fraction of the total internal energy ($\sim 10^{51}$~ergs), the nonthermal
742: electron pressure is most likely a small fraction of the ram pressure at the
743: shock. Yet, the total energy in cosmic-ray nuclei, which may be about 2
744: orders of magnitude larger than the total energy in cosmic-ray electrons,
745: could be large enough to modify the structure of the shock.
746:
747: \subsection{Cutoff Frequency and Diffusion Coefficient}
748:
749: In addition to providing evidence of curvature, fits to the synchrotron
750: spectral data provide a measure of the frequency at which the synchrotron
751: spectrum is cut off (eq.~[\ref{eqn04}]). The results for the cutoff
752: critical frequency are listed in Tables~\ref{tab2}--\ref{tab5}. A
753: comparison of Tables~\ref{tab2} and \ref{tab3} (and of Tables~\ref{tab4} and
754: \ref{tab5}) reveals that the frequencies for the curved spectra are somewhat
755: lower than the frequencies for the uncurved spectra because positive
756: curvature causes a synchrotron spectrum to flatten with increasing energy.
757: To compensate, the X-ray synchrotron spectrum is steepened by reducing the
758: cutoff frequency. The solid and dashed black lines in Figure~\ref{fig7}
759: show that the two effects more or less offset one another to produce
760: electron (and, hence, synchrotron) spectra that have nearly the same shape
761: in the X-ray synchrotron--emitting band.
762:
763: \citet{bam03}, \citet{rot04}, and \citet{bam08a} fitted X-ray spectra for
764: the bright rims of \snr\ with models that include the XSPEC synchrotron
765: component {\tt srcut}. Since {\tt srcut} does not include spectral
766: curvature, their results should be compared with the results listed in
767: Table~3. To the extent that the regions used by \citet{bam03} and
768: \citet{rot04} are similar to the regions used here, this comparison reveals
769: that their cutoff frequencies are systematically higher than the frequencies
770: listed in Table~\ref{tab3}. This difference is due, in part, to a difference
771: in the spectral indices. Our best-fit spectral indices ($\Gamma =
772: 1.98$--2.12; Table~\ref{tab3}) are lower than the indices used by Bamba
773: \etal\ (2003; $\Gamma = 2.14$) and Rothenflug \etal\ (2004; $\Gamma = 2.2$).
774: While an index of about 2.2 is appropriate for the radio
775: synchrotron--emitting electrons (\ie, the electrons at momenta $p \sim
776: 1$~GeV~$c^{-1}$), an index of 2.0 seems to better describe the X-ray
777: synchrotron--emitting electrons (\ie, the electrons at momenta $p \sim
778: 10$~TeV~$c^{-1}$). For example, if the electron spectrum is not curved, then
779: the mean index is $2.031 \pm 0.017$ (Table~\ref{tab3}). If curvature is
780: included, then the effective mean index at 10~TeV is $2.005 \pm 0.027$ [\ie,
781: $2.221 \pm 0.013 - (0.054 \pm 0.006) \log(10~{\rm TeV}/1~{\rm GeV})$;
782: Table~\ref{tab2}]. In both cases, the results imply that the index near the
783: cutoff of the electron spectrum is about 2.0, which is consistent with the
784: model of \citet{ber02}. Berezhko \etal\ note that the electron spectrum is
785: steeper at radio synchrotron--emitting momenta than at X-ray
786: synchrotron--emitting momenta because GeV electrons sample only a portion of
787: the shock while the highest-energy electrons ``see'' the full strength of
788: the shock. Although the same argument applies to the results of
789: \citet{bam08a}, who use a spectral index $\Gamma = 2.14$, their best-fit
790: break frequency for the entire northeastern rim is lower than the mean value
791: of $\nu_{m}$ in Table~\ref{tab3} [$\bar{\nu}_{m} = (9.07 \pm 0.98) \times
792: 10^{16}$~Hz]. If the normalization of the synchrotron spectrum is too high,
793: then the best-fit frequency may be reduced. While the normalization
794: reported by Bamba \etal\ (2008; 7.72~Jy at 1~GHz) is approximately correct
795: for the entire northeastern rim, only the flux density of the portion of the
796: rim associated with the very narrow region in which the X-ray synchrotron
797: emission is produced should be used. For these reasons, we believe that the
798: cut-off frequencies reported here are relatively accurate.
799:
800: The frequencies listed in Tables~\ref{tab2} and \ref{tab3} are plotted in
801: Figure~\ref{fig8} as a function of the position angles of the regions. This
802: angle is measured relative to the location of the center of the remnant
803: (Fig.~\ref{fig1}) and is expressed in degrees from north ($0\arcdeg$)
804: through east ($90\arcdeg$). As shown in Figure~\ref{fig8}, there may be an
805: azimuthal variation in the value of the cutoff frequency. A constant value
806: [\ie, the weighted mean value of $\left( 4.98 \pm 0.67 \right) \times
807: 10^{16}$~Hz, 90\% confidence level uncertainties] can be excluded at the
808: 2.5~$\sigma$ confidence level. \citet{rot04} also report evidence of an
809: azimuthal variation in the cutoff frequency. Since the frequency $\nu_{m}
810: \propto p_{m}^{2} B$, these results imply that the cutoff momentum of the
811: electron spectrum and/or the magnetic field strength varies along the
812: northeastern rim. Since it is not possible to independently determine both
813: $p_{m}$ and $B$ using the synchrotron spectral data alone, the present
814: analysis does not provide a means of identifying the cause of a variation in
815: the cutoff frequency.
816:
817: As described in the Appendix, it is possible to use a fitted value
818: of the cutoff critical frequency to constrain the mean electron diffusion
819: coefficient $\bar{\kappa}$. The upper limit on $\bar{\kappa}$, as a fraction
820: of the mean Bohm diffusion coefficient $\bar{\kappa}_{\rm B}$, depends on
821: the forward-shock velocity $u_{1}$, the cutoff critical frequency $\nu_{m}$,
822: and a function $f$ of the compression ratio $r$ and the ratio of the
823: upstream to downstream magnetic field strengths $B_{1} / B_{2}$
824: (eqs.~[\ref{eqna05}]--[\ref{eqna08}]).%
825: %
826: \footnote{The corresponding upper limit on the diffusion length $l_{m}
827: \approx \bar{\kappa} / u_{1}$ \citep{lag83}. Using eqs.~(\ref{eqn03}) and
828: (\ref{eqna04}) yields $l_{m} \propto f u_{1} \nu_{m,2}^{-1/2} B_{2}^{-3/2}$.
829: If $f = 0.1875$, $u_{1} = 2300$--5000~km~s$^{-1}$, $\nu_{m,2} = 10^{17}$~Hz,
830: and $B_{2} > 25$~$\mu$G \citep{aha05a}, then $l_{m} < 0.17$--0.36~pc (\ie,
831: $l_{m} / d < 16''$--34$''$ if $d = 2.2$~kpc;
832: Winkler \etal\ 2003).}%
833: %
834: \ If $u_{1} = 2300$--5000~km~s$^{-1}$ \citep{lam96,dwa98,gha02}, $\nu_{m} =
835: 1.1 \times 10^{17}$~Hz (Table~\ref{tab2}), and $f = 0.1875$ (\ie, $r = 4$
836: and $B_{1}/B_{2} \approx 0$), then $\bar{\kappa} \le (4.5$--$21)
837: \bar{\kappa}_{\rm B}$. If the compression ratio is 5.2 \citep{kse05a}
838: instead of 4, then $f = 0.155$ and $\bar{\kappa} \le (3.7$--$18)
839: \bar{\kappa}_{\rm B}$. If the cutoff frequency is about $10^{18}$~Hz
840: \citep{rot04} instead of $10^{17}$~Hz, then $\bar{\kappa} \le (0.5$--$2.3)
841: \bar{\kappa}_{\rm B}$. In each of these cases, the highest-energy electrons
842: have diffusion coefficients nearly as small as the Bohm coefficient (\ie,
843: are being accelerated about as fast as possible).
844:
845: Similar results for the diffusion coefficient are reported by \citet{bam03},
846: \citet{yam04}, and \citet{par06}. The latter two of these three use
847: equations similar to equation~(\ref{eqna05}) but obtain significantly
848: smaller limits for the diffusion coefficient because they use much larger
849: cutoff critical frequencies ($9.1 \times 10^{17}$ and $2.5 \times
850: 10^{18}$~Hz, respectively, instead of $1.1 \times 10^{17}$~Hz).
851: \citet{bam03} take a different approach. They assume that the width
852: ($\psi_{1} d$) of the part of a filament that is in the upstream region is
853: determined by the diffusion length ($\kappa_{1} / u_{1}$) in this region. In
854: this case,
855: %
856: \begin{eqnarray}
857: \frac{\kappa_{1}}{\kappa_{\rm B,1}}
858: & =
859: & \left( \frac{27 e^{3}}{4 \pi m^{3} c^{4}} \right)^{1/2}
860: \left( \frac{B_{1}^{3} \sin \theta_{1}}{\nu_{m,1}} \right)^{1/2}
861: \psi_{1} d u_{1}
862: \label{eqn08}
863: \\
864: & =
865: & 0.00262
866: \left( \frac{B_{1}}{3\ \mu{\rm G}} \right)^{3/2}
867: \left( \frac{\psi_{1}}{1\ {\rm arcsec}} \right)
868: \left( \frac{d}{1\ {\rm kpc}} \right)
869: \left( \frac{u_{1}}{10^{3}\ {\rm km}\ {\rm s}^{-1}} \right)
870: \left( \frac{\nu_{m,1}}{10^{17}\ {\rm Hz}} \right)^{-1/2},
871: \label{eqn09}
872: \end{eqnarray}
873: %
874: where $\kappa_{1}$, $\kappa_{B,1}$ [$= p c / (3 e B_{1})$], $B_{1}$,
875: $\nu_{m,1}$, and $\psi_{1}$ are the diffusion coefficient, Bohm diffusion
876: coefficient, mean magnetic field strength, cutoff critical frequency, and
877: filament width,%
878: %
879: \footnote{Here $\psi_{1}$ is the actual angular width, not the observed
880: angular width.}
881: %
882: respectively, in the upstream region and $u_{1}$ is the velocity of the
883: forward shock. The numerical constant in equation~(\ref{eqn09}) was
884: computed assuming that the value of $\sin \theta_{1}$ is $\pi/4$, the
885: isotropic mean value. Like equation~(\ref{eqna05}), equation~(\ref{eqn09})
886: is valid only at the momentum $p = p_{m}$. If the value of $B_{1}$ is
887: comparable to a typical interstellar field strength (\eg, 3~$\mu$G), then,
888: for reasonable values of $\psi_{1}$, $d$, $u_{1}$, and $\nu_{m,1}$,
889: $\kappa_{1}$ is uncomfortably small compared with the Bohm diffusion
890: coefficient. There are at least three explanations for this apparent
891: dilemma. One explanation is that, near the shock, $B_{1}$ may be amplified
892: by cosmic-ray streaming \citep{luc00,bel01,bel04}. For example,
893: equation~(15) of \citet{bel01} yields $B_{1} = 30$~$\mu$G (see Ksenofontov
894: \etal\ 2005) and, hence, $\kappa_{1} \sim \kappa_{\rm B,1}$, if the upstream
895: mass density $\rho_{1} = 2.4 \times 10^{-25}$~g~cm$^{-3}$ (\ie, $n_{1} =
896: 0.1$~cm$^{-3}$), the shock velocity $u_{1} = 2300$--5000~km~s$^{-1}$, and
897: the cosmic-ray pressure is about 30\%--14\% of the ram pressure,
898: respectively. Another explanation, which is mentioned by \cite{yam04}, is
899: that the rate of acceleration can exceed the Bohm limit (\ie, $\kappa_{1}$
900: can be less than $\kappa_{\rm B,1}$) if the magnetic field is parallel to
901: the shock (see Jokipii 1987). A third explanation is that the width of a
902: filament represents the size of the region in which the magnetic field is
903: strong \citep{poh05} instead of the diffusion length. In this case,
904: equations~(\ref{eqn08}) and (\ref{eqn09}) are not useful because an
905: assumption upon which they are based is not valid.
906:
907: Independent of the technique used to estimate or constrain the diffusion
908: length, it seems likely that the highest-energy electrons in \snr\ are
909: diffusing close to the Bohm limit. Otherwise, the rate of acceleration is
910: too low for electrons to reach energies high enough to produce the observed
911: X-ray synchrotron radiation. Yet, it is important to note that the limit on
912: the diffusion coefficient described in the Appendix (and the value obtained
913: using eqs.~[\ref{eqn08}] and [\ref{eqn09}]) is only an estimate, for the
914: following reasons:
915:
916: 1. The fitted values of the cutoff frequency are sensitive to the assumed
917: shape of the electron spectrum at momenta near the cutoff in the electron
918: spectrum. The actual functional form of the electron spectrum may be
919: significantly more complicated than equation~(\ref{eqn2}). As an example of
920: the extent to which differences in the shape of the electron spectrum can
921: affect the limit, consider the results obtained using the models with and
922: without curvature. If the curved model is used, then the maximum cut-off
923: critical frequency is $1.1 \times 10^{17}$~Hz (Table~\ref{tab2}) and
924: $\bar{\kappa} \le (4.5$--$21) \bar{\kappa}_{\rm B}$ (for $f = 0.1875$,
925: $B_{1}/B_{2} \approx 0$, and $u_{1} = 2300$--5000~km~s$^{-1}$). If the
926: model does not include curvature, then $\nu_{m} = 3.0 \times 10^{17}$~Hz
927: (Table~\ref{tab3}) and $\bar{\kappa} \le (1.7$--$7.8) \bar{\kappa}_{\rm B}$.
928:
929: 2. There is evidence that the cutoff frequency varies from one region to
930: another in \snr\ (Fig.~\ref{fig8}; Rothenflug \etal\ 2004). Here the
931: frequency used to compute the limit is the largest value in
932: Table~\ref{tab2}. Therefore, our constraint on the diffusion coefficient of
933: \snr\ applies only to the region where the cutoff frequency is largest. The
934: limits for other regions of the remnant are larger.
935:
936: 3. The limit is computed assuming that the electron momentum $p = p_{m}$.
937: Therefore, the limit only applies to electrons at the cutoff momentum [\ie,
938: $\bar{\kappa} = \bar{\kappa}(p = p_{m})$]. An advantage to using this
939: momentum is that the limit does not depend on the functional form of
940: $\bar{\kappa}(p)$.
941:
942: 4. The limit is based on the assumption that the cutoff is due to
943: synchrotron losses. If some other process, such as the escape of electrons
944: from the acceleration region or the length of time over which particles are
945: accelerated, determines the cutoff momentum, then the diffusion coefficient
946: is lower than the right-hand side of equation~(\ref{eqna05}). If this
947: equation yields a limit close to 1, then the results indicate that
948: synchrotron losses are important for electrons that have momenta $p \ge
949: p_{m}$.
950:
951: 5. The value of the function $f$ depends on the unknown compression ratio
952: and the unknown ratio of the upstream to downstream magnetic field strengths
953: (eq.~[\ref{eqna07}]). A value of $f = 0.1875$ (\ie, $r = 4$ and $B_{1} /
954: B_{2} \approx 0$) is used here because the compression ratio is probably at
955: least as large as 4 \citep{blo01,kse05a,war05,cas08a}. If $r > 4$, then $f
956: < 0.1875$ and our upper limit on the diffusion coefficient is overestimated.
957: Likewise, the upper limit is overestimated if $B_{1} / B_{2}$ is
958: significantly larger than zero.
959:
960: 6. Equation~(\ref{eqna05}) was derived assuming an isotropic pitch-angle
961: distribution. While such a distribution is expected if electrons diffuse in
962: the Bohm limit, the actual distribution may be different.
963:
964: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
965:
966: % 5. Conclusions
967:
968: \section{Conclusions}
969: \label{con}
970:
971: We have performed a joint spectral analysis of some {\sl Chandra} ACIS X-ray
972: data and MOST radio data for 13 small regions along the bright northeastern
973: rim of the supernova remnant \snr. The data were fitted with a model that
974: includes a synchrotron emission component. This component is based on an
975: electron spectrum that has a momentum-dependent spectral index. The rate of
976: change in the index for each decade in momentum is a free parameter of the
977: fit. If the assumptions described in \S\ \ref{curv} are valid, then the
978: results of the spectral analysis, which show that the synchrotron spectra of
979: \snr\ are curved, can be interpreted as evidence of curvature in the
980: GeV-to-TeV electron spectra. The mean amount of curvature in the electron
981: spectra is qualitatively consistent with predictions of the amount of
982: curvature in the proton spectrum of \snr\ \citep{ell00}. The best-fit
983: power-law index at 1~GeV (\ie, at radio synchrotron--emitting momenta) is
984: $2.221^{+0.013}_{-0.012}$. Including the effect of curvature, the effective
985: spectral index at about 10~TeV (\ie, at X-ray synchrotron--emitting momenta)
986: is $2.005 \pm 0.027$ (90\% confidence level uncertainties). This effective
987: index is consistent with the predictions of \citet{ber02}.
988:
989: The evidence of curved electron spectra suggests that cosmic rays are not
990: ``test'' particles. The cosmic-ray pressure at the shock is large enough to
991: modify the structure of the shock. Since nonthermal electrons contain only
992: about 0.1\% (\ie, $10^{48}$~ergs) or less of the total internal energy, the
993: results provide indirect evidence of a much more energetic population of
994: cosmic-ray protons. Collectively, the evidence of (1) spectral curvature in
995: Cas~A \citep{jon03}, RCW~86 \citep{vin06}, and \snr, (2) an unusually low
996: electron temperature in 1E~0102.2$-$7219 \citep{hug00}, and (3) a
997: compression ratio greater than 4 in Tycho \citep{war05} and \snr\
998: \citep{cas08a} suggests that efficient particle acceleration may be a common
999: feature of young, shell-type supernova remnants.
1000:
1001: The results of the spectral analysis also determine the ``cutoff critical
1002: frequency'' $\nu_{m}$. This frequency seems to vary from one region to
1003: another (see also Rothenflug \etal\ 2004), which implies that the
1004: exponential cutoff momentum of the electron spectrum and/or the strength of
1005: the magnetic field varies. It is not possible to identify the cause of the
1006: variation using the synchrotron spectral data alone.
1007:
1008: As described in the Appendix, the cutoff frequency can be used to set an
1009: upper limit on the mean diffusion coefficient $\bar{\kappa}$ of the
1010: highest-energy electrons. Aside from the cutoff frequency, this limit
1011: depends (strongly) on the velocity of the forward shock $u_{1}$ and (weakly)
1012: on the compression ratio $r$ and the ratio of the upstream to downstream
1013: magnetic field strengths $B_{1} / B_{2}$. If $\nu_{m} = 1.1 \times
1014: 10^{17}$~Hz, $u_{1} = 2300$--5000~km~s$^{-1}$, $r = 4$, and $B_{1} / B_{2}
1015: \approx 0$, then $\bar{\kappa} < (4.5$--$21) \bar{\kappa}_{\rm B}$, where
1016: $\bar{\kappa}_{\rm B}$ is the mean Bohm diffusion coefficient. This result
1017: implies that at least some of the highest-energy electrons in \snr\ diffuse
1018: close to the Bohm limit (\ie, are accelerated about as fast as possible),
1019: which provides additional support for the idea that Galactic cosmic rays are
1020: predominantly accelerated by the shocks of supernova remnants.
1021:
1022: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1023:
1024: % 6. Acknowledgments
1025:
1026: \acknowledgments
1027:
1028: We gratefully acknowledge the guidance of Don Ellison. This work grew out
1029: of discussions with him regarding the shape of cosmic-ray spectra. We thank
1030: Chuck Dermer, whose encouragement led us to develop a synchrotron model for
1031: a curved electron spectrum. This work benefited substantially from
1032: discussions with Tom Jones and Steve Reynolds and from comments by the
1033: anonymous referee. G.\ E.\ A.\ and J.\ C.\ H.\ are supported by contract
1034: SV3-73016 between MIT and the Smithsonian Astrophysical Observatory. The
1035: Chandra X-Ray Center at the Smithsonian Astrophysical Observatory is
1036: operated on behalf of NASA under contract NAS8-03060.
1037:
1038: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1039:
1040: % 7. Appendix
1041:
1042: \appendix
1043:
1044: \section{Diffusion coefficient}
1045: \label{appa}
1046:
1047: This appendix describes how measurements of or inferences about the shock
1048: velocity and cutoff frequency can be used to place an upper limit on the
1049: electron diffusion coefficient. Since the mean rate of synchrotron losses
1050: cannot exceed the mean rate of energy gains at momenta below the cutoff of
1051: the electron spectrum,
1052: %
1053: \begin{equation}
1054: \left( \frac{dE}{dt} \right)_{\rm acc}
1055: \ge \
1056: - \left( \frac{dE}{dt} \right)_{\rm sync}
1057: \label{eqna01}
1058: \end{equation}
1059: %
1060: in this range. The rate of acceleration at $p = p_{m} \gg mc$ is given
1061: by
1062: %
1063: \begin{equation}
1064: \left( \frac{dE}{dt} \right)_{\rm acc}
1065: =
1066: \frac{p_{m} c}{3} \,
1067: \left( \frac{\kappa_{1}}{u_{1}} + \frac{\kappa_{2}}{u_{2}} \right)^{-1}
1068: \left( u_{1} - u_{2} \right)
1069: \label{eqna02}
1070: \end{equation}
1071: %
1072: (Lagage \& Cesarsky 1983). If the rates of synchrotron losses in the
1073: upstream and downstream regions are weighted by the mean residence times in
1074: these regions [$\Delta t_{1} = 4 \kappa_{1} / (u_{1} v)$ and $\Delta t_{2} =
1075: 4 \kappa_{2} / (u_{2} v)$; Webb \etal\ 1984], then
1076: %
1077: \begin{equation}
1078: - \left( \frac{dE}{dt} \right)_{\rm sync}
1079: =
1080: \frac{e^{4}}{6 \pi \epsilon_{0} m^{4} c^{3}} \,
1081: p_{m}^{2}
1082: \left( \frac{\kappa_{1}}{u_{1}} B_{1}^{2} \sin^{2} \theta_{1} +
1083: \frac{\kappa_{2}}{u_{2}} B_{2}^{2} \sin^{2} \theta_{2} \right)
1084: \left( \frac{\kappa_{1}}{u_{1}} + \frac{\kappa_{2}}{u_{2}}\right)^{-1}
1085: \label{eqna03}
1086: \end{equation}
1087: %
1088: (Blumenthal \& Gould 1970), where $\kappa_{1}$ and $\kappa_{2}$ are the
1089: upstream and downstream electron diffusion coefficients perpendicular to the
1090: shock, $u_{1}$ and $u_{2}$ are the speeds of the upstream and downstream
1091: material relative to the shock, $v$ is the velocity of a particle,
1092: $\epsilon_{0}$ is the permittivity of free space, $B_{1}$ and $B_{2}$ are
1093: the mean upstream and downstream magnetic field strengths, and $\theta_{1}$
1094: and $\theta_{2}$ are the pitch angles between the electron momentum and
1095: magnetic field vectors in the upstream and downstream regions, respectively.
1096: If $\sin \theta$ in equation~(\ref{eqn03}) is replaced by the isotropic mean
1097: value of $\pi / 4$ and if $\sin^{2} \theta_{1}$ and $\sin^{2} \theta_{2}$ in
1098: equation~(\ref{eqna03}) are replaced by the isotropic mean value of
1099: $\case{2}{3}$, then a combination of equations~(\ref{eqn03}),
1100: (\ref{eqna01}), (\ref{eqna02}), and (\ref{eqna03}) yields
1101: %
1102: \begin{eqnarray}
1103: \frac{\bar{\kappa}}{\bar{\kappa}_{\rm B}}
1104: & \le
1105: & \frac{27 \pi \epsilon_{0} m c}{16 e^{2}}
1106: \frac{f u_{1}^{2}}{\nu_{m,2}}
1107: \label{eqna04}
1108: \\
1109: & \le
1110: & 0.936
1111: \left( \frac{f}{0.1875} \right)
1112: \left( \frac{u_{1}}{10^{3}\ {\rm km}\ {\rm s}^{-1}} \right)^{2}
1113: \left( \frac{\nu_{m,2}}{10^{17}\ {\rm Hz}} \right)^{-1},
1114: \label{eqna05}
1115: \end{eqnarray}
1116: where
1117: %
1118: \begin{equation}
1119: \bar{\kappa}
1120: =
1121: \left( \frac{B_{1}^{2}}{u_{1}} \kappa_{1} +
1122: \frac{B_{2}^{2}}{u_{2}} \kappa_{2} \right)
1123: \left( \frac{B_{1}^{2}}{u_{1}} + \frac{B_{2}^{2}}{u_{2}} \right)^{-1},
1124: \label{eqna06}
1125: \end{equation}
1126: %
1127: \begin{equation}
1128: \bar{\kappa}_{\rm B}
1129: =
1130: \left( \frac{B_{1}^{2}}{u_{1}} \kappa_{\rm B,1} +
1131: \frac{B_{2}^{2}}{u_{2}} \kappa_{\rm B,2} \right)
1132: \left( \frac{B_{1}^{2}}{u_{1}} + \frac{B_{2}^{2}}{u_{2}} \right)^{-1},
1133: \label{eqna07}
1134: \end{equation}
1135: %
1136: the upstream and downstream Bohm diffusion coefficients $\kappa_{\rm B,1}$
1137: and $\kappa_{\rm B,2}$ are equal to $p_{m} c / (3 e B_{1})$ and $p_{m} c /
1138: (3 e B_{2})$, respectively, at a momentum $p = p_{m}$, $\nu_{m,2}$ is the
1139: critical frequency of an electron with this momentum in a magnetic field $B
1140: = B_{\rm 2}$,
1141: %
1142: \begin{equation}
1143: f
1144: =
1145: \frac{r - 1}{r \left[ r + \left( B_{1} / B_{2} \right) \right]},
1146: \label{eqna08}
1147: \end{equation}
1148: %
1149: and the compression ratio $r = u_{1} / u_{2}$. The value of 0.1875 in
1150: equation~(\ref{eqna05}) was calculated assuming $r = 4$ and $B_{1} / B_{2}
1151: \approx 0$. If the compression ratio is greater than 4 or $B_{1} / B_{2}$
1152: is significantly larger than zero, then $f < 0.1875$. In no case can $f$ be
1153: larger than 0.25. If $B_{1} = B_{2}$, then equation~(\ref{eqna05}) is
1154: identical to the analogous equation in \cite{sta06}.
1155: Equation~(\ref{eqna05}) is similar to equation~(22) of \cite{aha99a},
1156: equation~(12) of \cite{laz04}, equation~(A.4) of \cite{yam04}, and
1157: equation~(22) of \cite{par06}, except these authors use the peak frequency
1158: instead of the critical frequency. The latter three also use $\sin \theta =
1159: 1$ instead of $\pi / 4$.
1160:
1161: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1162:
1163: % 8. References
1164:
1165: \begin{thebibliography}{}
1166:
1167: \bibitem[Aharonian \etal(2005)]{aha05a}
1168: Aharonian, F.\ A., \etal\
1169: 2005,
1170: \aap,
1171: 437
1172: 135
1173:
1174: \bibitem[Aharonian \& Atoyan(1999)]{aha99a}
1175: Aharonian, F.\ A., \& Atoyan, A.\ M.\
1176: 1999,
1177: \aap,
1178: 351,
1179: 330
1180:
1181: \bibitem[Allen \etal(2001)]{all01}
1182: Allen, G.\ E., Petre, R., \& Gotthelf, E.\ V.\
1183: 2001,
1184: \apj,
1185: 558,
1186: 739
1187:
1188: \bibitem[Anders \& Grevesse(1989)]{and89}
1189: Anders, E., \& Grevesse, N.\ 1989, Geochim.\ Cosmochim.\ Acta, 53, 197
1190:
1191: \bibitem[Anderson \& Rudnick(1996)]{and96}
1192: Anderson, M.\ C., \& Rudnick, L.\
1193: 1996,
1194: \apj,
1195: 456,
1196: 234
1197:
1198: \bibitem[Bamba et~al.(2003)]{bam03}
1199: Bamba, A., Yamazaki, R., Ueno, M., \& Koyama, K.\ 2003, \apj, 589, 827
1200:
1201: \bibitem[Bamba et~al.(2008)]{bam08a}
1202: Bamba, A., \etal\
1203: 2008,
1204: \pasj,
1205: 60,
1206: S153
1207:
1208: \bibitem[Bell(1987)]{bel87}
1209: Bell, A.\ R.\
1210: 1987,
1211: \mnras,
1212: 225,
1213: 615
1214:
1215: \bibitem[Bell(2004)]{bel04}
1216: ---------.\
1217: 2004,
1218: \mnras,
1219: 353,
1220: 550
1221:
1222: \bibitem[Bell \& Lucek(2001)]{bel01}
1223: Bell, A.\ R., \& Lucek, S.\ G.\
1224: 2001,
1225: MNRAS,
1226: 321,
1227: 433
1228:
1229: \bibitem[Berezhko \& Ellison(1999)]{ber99}
1230: Berezhko, E.\ G., \& Ellison, D.\ C.\
1231: 1999,
1232: \apj,
1233: 526,
1234: 385
1235:
1236: \bibitem[Berezhko \etal(2002)]{ber02}
1237: Berezhko, E.\ G., Ksenofontov, L.\ T., \& V\"{o}lk, H.\ J.\
1238: 2002
1239: \aap,
1240: 395,
1241: 943
1242:
1243: \bibitem[Blondin \& Ellison(2001)]{blo01}
1244: Blondin, J.\ M., \& Ellison, D.\ C.\
1245: 2001,
1246: \apj,
1247: 560,
1248: 244
1249:
1250: \bibitem[Blumenthal \& Gould(1970)]{blu70}
1251: Blumenthal, G.\ R., \& Gould, R.\ J.\
1252: 1970,
1253: Rev.\ Mod.\ Phys.,
1254: 42,
1255: 237
1256:
1257: \bibitem[Cassam-Chenai et~al.(2008)]{cas08a}
1258: Cassam-Chenai, G., Hughes, J.\ P., Reynoso, E.\ M., Badenes, C., \&
1259: Moffett, D.\
1260: 2008,
1261: \apj,
1262: 680,
1263: 1180
1264:
1265: \bibitem[Chevalier(1983)]{che83}
1266: Chevalier, R.\ A.\
1267: 1983,
1268: \apj,
1269: 272,
1270: 765
1271:
1272: \bibitem[Decourchelle \etal(2000)]{dec00}
1273: Decourchelle, A., Ellison, D.\ C., \& Ballet, J.\
1274: 2000,
1275: \apj,
1276: 543,
1277: L57
1278:
1279: \bibitem[DeLaney \etal(2002)]{del02}
1280: DeLaney, T., Koralesky, B., Rudnick, L., \& Dickel, J.\ R.\
1281: 2002,
1282: \apj,
1283: 580,
1284: 914
1285:
1286: \bibitem[Drury \etal(1989)]{dru89}
1287: Drury, L.\ O'C., Markiewicz, W.\ J., \& V\"{o}lk, H.\ J.\
1288: 1989,
1289: \aap,
1290: 225,
1291: 179
1292:
1293: \bibitem[Dwarkadas \& Chevalier(1998)]{dwa98}
1294: Dwarkadas, V.\ V., \& Chevalier, R.\ A.\
1295: 1998,
1296: \apj,
1297: 497,
1298: 807
1299:
1300: \bibitem[Dyer et~al.(2001)]{dye01}
1301: Dyer, K.\ K., Reynolds, S.\ P., Borkowski, K.\ J., Allen, G.\ E., \&
1302: Petre, R.\
1303: 2001,
1304: \apj,
1305: 551,
1306: 439
1307:
1308: \bibitem[Ellison \etal(2000)]{ell00}
1309: Ellison, D.\ C., Berezhko, E.\ G., \& Baring, M.\ G.\ 2000, \apj, 540, 292
1310:
1311: \bibitem[Ellison \& Reynolds(1991)]{ell91}
1312: Ellison, D.\ C., \& Reynolds, S.\ P.\
1313: 1991,
1314: \apj,
1315: 382,
1316: 242
1317:
1318: \bibitem[Ellison \etal(2001)]{ell01}
1319: Ellison, D.\ C., Slane, P., \& Gaensler, B.\ M.\
1320: 2001,
1321: \apj,
1322: 563,
1323: 191
1324:
1325: \bibitem[Eriksen \etal(2001)]{eri01}
1326: Eriksen, K.\ A., Morse, J.\ A., Kirshner, R.\ P., \& Winkler, P.\ F.\
1327: 2001,
1328: in AIP Conf.\ Proc.\
1329: 565,
1330: Young Supernova Remnants,
1331: ed.\ S.\ S.\ Holt \& U.\ Hwang
1332: (Melville, NY: AIP),
1333: 193
1334:
1335: \bibitem[Finkelstein \etal(2006)]{fin06}
1336: Finkelstein, S.\ L., \etal\
1337: 2006,
1338: \apj,
1339: 641,
1340: 919
1341:
1342: \bibitem[Flanagan \etal(2004)]{fla04}
1343: Flanagan, K.~A, Canizares, C.~R., Dewey, D., Houck, J.~C., Fredericks,
1344: A.~C., Schattenburg, M.~L., Markert, T.~H., \& Davis, D.~S.\
1345: 2004,
1346: \apj,
1347: 605,
1348: 230
1349:
1350: \bibitem[Gardner \& Milne(1965)]{gar65}
1351: Gardner, F.\ F., \& Milne, D.\ K.\
1352: 1965,
1353: AJ,
1354: 70,
1355: 754
1356:
1357: \bibitem[Ghavamian \etal(2002)]{gha02}
1358: Ghavamian, P., Winkler, P.\ F., Raymond, J.\ C., \& Long, K.\ S.\
1359: 2002,
1360: \apj,
1361: 572,
1362: 888
1363:
1364: \bibitem[Green(2006)]{gre06}
1365: Green D.\ A.\
1366: 2006,
1367: A Catalogue of Galactic Supernova Remnants (ver.\ 2006 April; Cambridge:
1368: Cavendish Lab.), http://www.mrao.cam.ac.uk/surveys/snrs
1369:
1370: \bibitem[Houck \& Allen(2006)]{hou06}
1371: Houck, J.\ C., \& Allen, G.\ E.\
1372: 2006,
1373: \apjs,
1374: 167,
1375: 26%--39
1376:
1377: \bibitem[Houck \& DeNicola(2000)]{hou00}
1378: Houck, J.\ C., \& DeNicola, L.\ A.\
1379: 2000,
1380: in ASP Conf.\ Ser.\ 216,
1381: Astronomical Data Analysis Software and Systems IX,
1382: ed.\ N.\ Manset, C.\ Veillet, \& D.\ Crabtree
1383: (San Francisco: ASP),
1384: 591
1385:
1386: \bibitem[Hughes \etal(2000)]{hug00}
1387: Hughes, J.\ P., Rakowski, C.\ E., \& Decourchelle, A.\
1388: 2000,
1389: \apjl,
1390: 543,
1391: L61
1392:
1393: \bibitem[Jokipii(1987)]{jok87}
1394: Jokipii, J.\ R.\
1395: 1987,
1396: \apj,
1397: 313,
1398: 842
1399:
1400: \bibitem[Jones \etal(2003)]{jon03}
1401: Jones, T.\ J., Rudnick, L., DeLaney, T., \& Bowden, J.\
1402: 2003,
1403: \apj,
1404: 587,
1405: 227
1406:
1407: \bibitem[Ksenofontov \etal(2005)]{kse05a}
1408: Ksenofontov, L.\ T., Berezhko, E.\ G., \& V\"{o}lk, H.\ J.\
1409: 2005,
1410: \aap,
1411: 443,
1412: 973
1413:
1414: \bibitem[Koyama \etal(1995)]{koy95}
1415: Koyama, K., Petre, R., Gotthelf, E.\ V., Hwang, U., Matsuura, M., Ozaki,
1416: M., \& Holt, S.\ S.\
1417: 1995,
1418: Nature,
1419: 378,
1420: 255
1421:
1422: \bibitem[Kundu(1970)]{kun70}
1423: Kundu, M.\ R.\
1424: 1970,
1425: \apj,
1426: 162,
1427: 17
1428:
1429: \bibitem[Lagage \& Cesarsky(1983)]{lag83}
1430: Lagage, P.-O., \& Cesarsky, C.\ J.\
1431: 1983,
1432: \aap,
1433: 125,
1434: 249
1435:
1436: \bibitem[Laming \etal(1996)]{lam96}
1437: Laming, J.\ M., Raymond, J.\ C., McLaughlin, B.\ M., \& Blair, W.\ P.\
1438: 1996,
1439: \apj,
1440: 472,
1441: 267
1442:
1443: \bibitem[Lazendic \etal(2004)]{laz04}
1444: Lazendic, J.\ S., Slane, P.\ O., Gaensler, B.\ M., Reynolds, S.\ P.,
1445: Plucinsky, P.\ P., \& Hughes, J.\ P.\
1446: 2004,
1447: \apj,
1448: 602,
1449: 271
1450:
1451: \bibitem[Long \etal(2003)]{lon03}
1452: Long, K.\ S., Reynolds, S.\ P., Raymond, J.\ C., Winkler, P.\ F., Dyer,
1453: K.\ K., Petre, R.\
1454: 2003,
1455: \apj,
1456: 586,
1457: 1162
1458:
1459: \bibitem[Lucek \& Bell(2000)]{luc00}
1460: Lucek, S.\ G. \& Bell, A.\ R.\
1461: 2000,
1462: \mnras,
1463: 314,
1464: 65
1465:
1466: \bibitem[Milne(1971)]{mil71}
1467: Milne, D.\ K.\ 1971, Australian.\ J.\ Phys., 24, 757
1468:
1469: \bibitem[Milne \& Dickel(1975)]{mil75}
1470: Milne, D.\ K., \& Dickel, J.\ R.\
1471: 1975,
1472: Australian.\ J.\ Phys.,
1473: 28,
1474: 209
1475:
1476: \bibitem[Moffett \etal(1993)]{mof93a}
1477: Moffett, D.\ A., Goss, W.\ M., \& Reynolds, S.\ P.
1478: 1993,
1479: \aj,
1480: 106,
1481: 1566
1482:
1483: \bibitem[Parizot \etal(2006)]{par06}
1484: Parizot, E., Marcowith, A., Ballet, J., \& Gallant, Y.\ A.\
1485: 2006,
1486: \aap,
1487: 453,
1488: 387
1489:
1490: \bibitem[Pohl \etal(2005)]{poh05}
1491: Pohl, M., Yan, H., \& Lazarian, A.\
1492: 2005,
1493: \apjl,
1494: 626,
1495: L101
1496:
1497: \bibitem[Protheroe(2004)]{pro04}
1498: Protheroe, R.\ J.\
1499: 2004,
1500: Astropart.\ Phys.,
1501: 21,
1502: 415
1503:
1504: \bibitem[Reynolds \& Ellison(1992)]{rey92}
1505: Reynolds, S.\ P., \& Ellison, D.\ C.\
1506: 1992,
1507: \apjl,
1508: 399,
1509: L75
1510:
1511: \bibitem[Reynolds \& Gilmore(1986)]{rey86}
1512: Reynolds, S.\ P., \& Gilmore, D.\ M.\ 1986, AJ, 92, 1138
1513:
1514: \bibitem[Roger et~al.(1988)]{rog88}
1515: Roger, R.\ S., Milne, D.\ K., Kesteven, M.\ J., Wellington, K.\ J., \&
1516: Haynes, R.\ F.\
1517: 1988,
1518: \apj,
1519: 332,
1520: 940
1521:
1522: \bibitem[Rothenflug \etal(2004)]{rot04}
1523: Rothenflug, R., Ballet, J., Dubner, G., Giacani, E., Decourchelle, A., \&
1524: Ferrando, P.\
1525: 2004,
1526: \aap,
1527: 425,
1528: 121
1529:
1530: \bibitem[Schaefer(1996)]{sch96}
1531: Schaefer, B.\ E.\
1532: 1996,
1533: \apj,
1534: 459,
1535: 438
1536:
1537: \bibitem[Stage \etal(2006)]{sta06}
1538: Stage, M.\ D., Allen, G.\ E., Houck, J.\ C., \& Davis, J.\ E.\
1539: 2006,
1540: Nature Phys.,
1541: 2,
1542: 614
1543:
1544: \bibitem[Stephenson \etal(1977)]{ste77}
1545: Stephenson, F.\ R., Clark, D.\ H., \& Crawford, D.\ F.\
1546: 1977,
1547: \mnras,
1548: 180,
1549: 567
1550:
1551: \bibitem[Tanimori \etal(1998)]{tan98}
1552: Tanimori, T., \etal\
1553: 1998,
1554: \apjl,
1555: 497,
1556: L25
1557:
1558: \bibitem[Uchiyama \etal(2003)]{uch03}
1559: Uchiyama, Y., Aharonian, F.\ A., \& Takahashi, T.\
1560: 2003,
1561: \aap,
1562: 400,
1563: 567
1564:
1565: \bibitem[Vink \etal(2006)]{vin06}
1566: Vink, J., Bleeker, J., van der Heyden, K., Bykov, A., Bamba, A., \
1567: \& Yamazaki, R.\
1568: 2006,
1569: \apjl,
1570: 648,
1571: L33
1572:
1573: \bibitem[V\"{o}lk \etal(2005)]{vol05}
1574: V\"{o}lk, H.\ J., Berezhko, E.\ G., \& Ksenofontov, L.\ T.\
1575: 2005,
1576: \aap,
1577: 433,
1578: 229
1579:
1580: \bibitem[V\"{o}lk \& Biermann(1988)]{vol88}
1581: V\"{o}lk, H.\ J., \& Biermann, P.\ L.\
1582: 1988,
1583: \apjl,
1584: 333,
1585: L65
1586:
1587: \bibitem[Warren \etal(2005)]{war05}
1588: Warren, J.\ S., \etal\
1589: 2005,
1590: \apj,
1591: 634,
1592: 376
1593:
1594: \bibitem[Webb \etal(1984)]{web84}
1595: Webb, G.\ M., Drury, L.\ O'C., \& Biermann, P.\
1596: 1984,
1597: \aap,
1598: 137,
1599: 185
1600:
1601: \bibitem[Winkler \etal(2003)]{win03}
1602: Winkler, P.\ F., Gupta, G., \& Long, K.\ S.\ 2003, \apj, 585, 324
1603:
1604: \bibitem[Winkler \& Long(1997)]{win97}
1605: Winkler, P.\ F., \& Long, K.\ S.\ 1997, \apj, 491, 829
1606:
1607: \bibitem[Yamazaki \etal(2004)]{yam04}
1608: Yamazaki, R., Yoshida, T., Terasawa, T., Bamba, A., \& Koyama, K.\
1609: 2004,
1610: \aap,
1611: 416,
1612: 595
1613:
1614: \bibitem[Zirakashvili \& Aharonian(2007)]{zir07a}
1615: Zirakashvili, V.\ N., \& Aharonian, F.\
1616: 2007,
1617: \aap,
1618: 465,
1619: 695
1620:
1621: \end{thebibliography}
1622:
1623: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1624:
1625: % 9. Figures
1626:
1627: \plotone{f1b.eps}
1628: \figcaption[f1b.eps]{
1629: %
1630: Color-coded ACIS image of the northeastern rim of \snr. Red, green, and blue
1631: correspond to the energy bands 0.4--1, 1--2, and 2--7~keV, respectively.
1632: Most of the emission at energies greater than 1~keV is concentrated in the
1633: whitish filaments along the rim. The edges of the six square CCDs used to
1634: observe the remnant are evident. The relatively high brightnesses of the CCD
1635: with the circle on it and the CCD in the lower left corner are due to
1636: enhanced low-energy sensitivities for these two devices. The circle at
1637: $\alpha = 15^{\rm h} 03^{\rm m} 51.56^{\rm s}$ and $\delta = -41\arcdeg
1638: 51\arcmin 18.8\arcsec$ (J2000) indicates the location of the nominal aim
1639: point of the telescope. The asterisk at $\alpha = 15^{\rm h} 02^{\rm m}
1640: 51.7^{\rm s}$ and $\delta = -41\arcdeg 56\arcmin 33.0\arcsec$ is the
1641: location of the center of the supernova remnant \citep{win97}. The 13
1642: squares are the X-ray spectral extraction regions.
1643: %
1644: \label{fig1}}
1645:
1646: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1647:
1648: \newpage
1649:
1650: \plotone{f2.eps}
1651: \figcaption[f2.eps]{
1652: %
1653: Radio spectrum for the entire supernova remnant \snr. {\sl Top}: The data
1654: points (Table~\ref{tab1}) and three models. The dashed line is the best-fit
1655: power law, $S(\nu) = 17.6^{+6.3}_{-4.8}[\nu/(1~{\rm GHz})]^{-\alpha}$~Jy,
1656: where $\alpha = 0.60^{+0.08}_{-0.09}$. The dotted line is a power law using
1657: Green's (2006) spectral parameters [$S(\nu) = 19 [\nu/(1~{\rm
1658: GHz})]^{-0.6}$~Jy]. The solid line is the best-fit (curved) synchrotron
1659: model for region~6 divided by the factor 0.00175 (see Table~\ref{tab2}).
1660: {\sl Bottom}: Differences between the data points and the solid line,
1661: divided by the uncertainties in the data points.
1662: %
1663: \label{fig2}}
1664:
1665: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1666:
1667: \plotone{f3.eps}
1668: \figcaption[f3.eps]{
1669: %
1670: An 843~MHz MOST image of \snr, courtesy of R.\ Roger \citep{rog88}. The
1671: ellipse at lower right indicates the half-power beamwidth. The 13 black
1672: squares are the locations of the X-ray spectral extraction regions. These
1673: regions were used to obtain the ``cospatial'' radio fluxes
1674: (Tables~\ref{tab2} and \ref{tab3}). The red squares are the locations used
1675: to obtain the peak radio fluxes (Tables~\ref{tab4} and \ref{tab5}). The
1676: ratios of the flux densities of the black and red regions to the total flux
1677: density are listed in the $\zeta$ columns of Tables~\ref{tab2}--\ref{tab5}.
1678: The bright spot at about $\alpha = 15^{\rm h} 04^{\rm m} 04^{s}$ and $\delta
1679: = -41\arcdeg 55\arcmin 48\arcsec$ is produced by an extragalactic source
1680: \citep{rey86}.
1681: %
1682: \label{fig3}}
1683:
1684: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1685:
1686: \plotone{f4.eps}
1687: \figcaption[f4.eps]{
1688: %
1689: ACIS spectrum for region~6 (see Fig.~\ref{fig1}) of the bright northeastern
1690: rim of \snr. {\sl Top}: Sum of the source and background spectra ({\sl data
1691: points}) and the sum of the best-fit model and background spectra ({\sl
1692: solid line}). The model displayed here includes spectral curvature as a
1693: free parameter (see Table~\ref{tab2}). {\sl Bottom}: Differences between
1694: the data points and the solid line, divided by the uncertainties in the data
1695: points.
1696: %
1697: \label{fig4}}
1698:
1699: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1700:
1701: \plotone{f5.eps}
1702: \figcaption[f5.eps]{
1703: %
1704: X-ray emission profiles along a $49 \arcsec$-wide strip passing through
1705: region~6 toward the center of \snr. The black histogram in the top panel is
1706: the 2--7~keV ACIS data. For comparison, the black line in the bottom panel
1707: is a model based on the assumption that the emission is produced in a
1708: uniformly emitting shell that is $6 \arcsec$ wide. The red curves are the
1709: data and model smoothed to the resolution of the MOST image. (The horizontal
1710: red bars are the 1~$\sigma$ widths of the Gaussian smoothing function.) The
1711: reduction in spatial resolution causes the peak to shift downstream, but the
1712: offset is small compared with the width of the extraction region. (The blue
1713: lines are the boundaries of region~6.) Therefore, it seems unlikely that the
1714: offsets between the X-ray and radio peaks (see Fig.~\ref{fig3} and
1715: Tables~\ref{tab4} and \ref{tab5}) are due to the different X-ray and radio
1716: resolutions.
1717: %
1718: \label{fig5}}
1719:
1720: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1721:
1722: \plotone{f6.eps}
1723: \figcaption[f6.eps]{
1724: %
1725: The 1, 2, and 3~$\sigma$ confidence contours for region~6 (see
1726: Fig.~\ref{fig1}) in the parameter space defined by the electron spectral
1727: index $\Gamma$ and electron spectral curvature $a$ (see eq.~[\ref{eqn2}]).
1728: The solid black and dashed red contours are the results obtained using the
1729: cospatial (Table~\ref{tab2}) and peak (Table~\ref{tab4}) radio fluxes,
1730: respectively. The plus signs indicate the best-fit values of the index and
1731: curvature. The dotted black line is the expected relationship between
1732: $\Gamma$ and $a$ (eq.~[\ref{eqn07}]). The dashed black line is the line
1733: along which the electron spectrum is not curved (\ie, $a = 0$). For
1734: region~6, a positive curvature (\ie, a result above the dashed line) is
1735: favored at about the 2.7 and 1.6~$\sigma$ confidence levels for the black
1736: and red contours, respectively.
1737: % As described in the text, the black dotted line has the expected slope of
1738: % the linear relationship between the index and curvature parameters.
1739: %
1740: \label{fig6}}
1741:
1742: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1743:
1744: \plotone{f7.eps}
1745: \figcaption[f7.eps]{
1746: %
1747: Best-fit electron number density spectra for region 6 (see Fig.~\ref{fig1}).
1748: The solid and dashed black lines are the curved (Table~\ref{tab2}) and
1749: uncurved (Table~\ref{tab3}) models, respectively, using the cospatial radio
1750: flux. The solid red line is the curved model using the peak radio flux
1751: (Table~\ref{tab4}). The upper dotted and dot-dashed curves are the models
1752: for the proton (not electron) spectrum of \snr\ presented by \cite{ell00}.
1753: These lines are plotted only for momenta $p > mc$ (\ie, the range of momenta
1754: to which our fits are sensitive) and are normalized to the solid black line
1755: at a kinetic energy of 0.9~GeV. As described in the text, the lower dotted
1756: and dot-dashed curves are the same pair of models multiplied by $[E / (0.9\
1757: {\rm GeV})]^{-0.2}$. The amounts of spectral curvature in the solid lines
1758: are consistent with the amounts of curvature in the lower dotted and
1759: dot-dashed curves. From left to right, the three bracketed energy bands
1760: contain the electrons that are primarily responsible for the observed
1761: thermal bremsstrahlung, radio synchrotron, and X-ray synchrotron emission.
1762: %
1763: \label{fig7}}
1764:
1765: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1766:
1767: \plotone{f8.eps}
1768: \figcaption[f8.eps]{
1769: %
1770: Cutoff frequency as a function of position angle (measured counterclockwise
1771: from north through east). From right to left, the 13 pairs of points are
1772: for regions 1--13, respectively (see Fig.~\ref{fig1}). The black data points
1773: and 90\% confidence intervals are the results obtained using a curved
1774: electron spectrum and the cospatial radio fluxes (Table~\ref{tab2}). The
1775: red data points, which have similar confidence intervals, are the results
1776: obtained using the peak radio fluxes (Table~\ref{tab4}). The dashed line is
1777: the weighted mean value ($4.98 \times 10^{16}$~Hz) of the black points. The
1778: corresponding 90\% confidence level interval [(4.31--$5.65) \times
1779: 10^{16}$~Hz] lies between the two dotted lines. A constant value for the
1780: cutoff frequency can be excluded at the 2.5~$\sigma$ confidence level.
1781: %
1782: \label{fig8}}
1783:
1784: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1785:
1786: % 10. Tables
1787:
1788: \clearpage
1789: \input{tab1.tex}
1790:
1791: \clearpage
1792: \input{tab2.tex}
1793:
1794: \clearpage
1795: \input{tab3.tex}
1796:
1797: \clearpage
1798: \input{tab4.tex}
1799:
1800: \clearpage
1801: \input{tab5.tex}
1802:
1803: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1804:
1805: % 11. Finish
1806:
1807: \end {document}
1808:
1809: