1: \documentclass[12pt,preprint]{aastex}
2: %\documentclass[onecolumn]{emulateapj}
3: %\usepackage{graphicx}
4: %\usepackage{natbib}
5: %\bibliographystyle{apj}
6:
7: \newcommand{\eqb}{\begin{equation}}
8: \newcommand{\eqe}{\end{equation}}
9:
10: \begin{document}
11: %
12: \title{Induced scattering of short radio pulses}
13: %
14: \author{Yuri Lyubarsky}
15: \affil{Department of Physics, Ben-Gurion University, Beer-Sheva,
16: Israel}
17: %
18:
19: \begin{abstract}
20: Effect of the induced Compton and Raman scattering on short,
21: bright radio pulses is investigated. It is shown that when a
22: single pulse propagates through the scattering medium, the
23: effective optical depth is determined by the duration of the pulse
24: but not by the scale of the medium. The induced scattering could
25: hinder propagation of the radio pulse only if close enough to the
26: source a dense enough plasma is presented. The induced scattering
27: within the relativistically moving source places lower limits on
28: the Lorentz factor of the source. The results are applied to the
29: recently discovered short extragalactic radio pulse.
30: \end{abstract}
31: %
32: \keywords{plasmas -- radiation mechanisms:nonthermal --
33: scattering}
34:
35: \section{Introduction}
36: Induced scattering could significantly affect radiation from
37: sources with high brightness temperatures. The induced Compton
38: scattering may be relevant in pulsars (Wilson \& Rees 1978;
39: Lyubarskii \& Petrova 1996, Petrova 2004a,b, 2007a,b), masers
40: (Zeldovich, Levich \& Sunyaev 1972; Montes 1977) and radio loud
41: active galactic nuclei (Sunyaev 1971; Coppi, Blandford \& Rees
42: 1993; Sincell \& Coppi 1996). The induced Raman scattering is
43: considered as the most plausible mechanism of eclipses in
44: binary pulsars (Eichler 1991; Gedalin \& Eichler 1993; Thompson
45: et al. 1994; Luo \& Melrose 1995) and was also invoked to place
46: constraints on the models of pulsars (Lyutikov 1998; Luo \&
47: Melrose 2006) and models of intraday variability in compact
48: extragalactic sources (Levinson \& Blandford 1995).
49:
50: Macquart (2007) used the induced Compton and Raman scattering
51: in order to place limits on the observability of the prompt
52: radio emission predicted (Usov \& Katz 2000; Sagiv \& Waxman
53: 2002; Moortgat \& Kuijpers 2005) to emanate from gamma-ray
54: bursts. Recent discovery of an enigmatic short extragalactic
55: radio pulse (Lorimer et al 2007) demonstrates that very high
56: brightness temperature transients do exist in nature. In this
57: paper, we address the induced scattering of short bright radio
58: pulses. First we study the induced Compton and Raman scattering
59: in the plasma surrounding the source. The central point is that
60: due to the non-linear character of the process, the effective
61: optical depth is determined not by the scale of the scattered
62: medium but by the width of the pulse provided that the pulse is
63: short in the sense that duration of the pulse is less than the
64: light travel time in the scattered medium. For this reason,
65: short enough pulses could propagate through the interstellar
66: medium, contrary to Macquart's claim. The induced scattering
67: could hinder propagation of a high brightness temperature pulse
68: only close enough to the source if the density of the ambient
69: plasma is large enough; here we find the corresponding
70: observability conditions. We also address the induced
71: scattering within the relativistically moving source and show
72: that transparency of the source implies a lower limit on the
73: Lorentz factor of the source. We apply the general results to
74: the short extragalactic radio pulse discovered by Lorimer et
75: al. (2007).
76:
77: \section{Induced Compton scattering}
78: The kinetic equation for the induced Compton scattering in the
79: non-relativistic plasma is written as (e.g. Wilson 1982)
80: \eqb
81: \frac{\partial n(\nu,\mathbf{\Omega})}{\partial t}+
82: c(\mathbf{\Omega\cdot\nabla})n(\nu,\mathbf{\Omega})=
83: \frac{3\sigma_T}{8\pi}N \frac{h}{m_ec}n(\nu,\mathbf{\Omega})
84: \int(\mathbf{e\cdot e}_1)^2
85: (1-\mathbf{\Omega\cdot\Omega}_1)\frac{\partial
86: \nu^2n(\nu,\mathbf{\Omega}_1)}{\partial\nu}d\mathbf{\Omega}_1;
87: \label{kinComp}\eqe
88: where $n(\nu,\mathbf{\Omega})$ is the photon occupation number of
89: a beam in the direction $\mathbf{\Omega}$, $N$ the electron number
90: density, $\mathbf{e}$ the polarization vector. The induced
91: scattering rate is proportional to the number of photons already
92: available in the final state therefore the scattering initially
93: occurs within the primary emission beam where the radiation
94: density is high. However, when the primary beam is narrow, as is
95: anyway the case at large distance from the source, the recoil
96: factor $1-\mathbf{\Omega\cdot\Omega}_1$ makes the scattering
97: within the beam inefficient; then the scattering outside the beam
98: dominates because according to Eq.(\ref{kinComp}), even weak
99: isotropic background radiation (created, e.g., by spontaneous
100: scattering) grows exponentially so that the energy of the
101: scattered radiation becomes eventually comparable with the energy
102: density in the primary beam.
103:
104: In this and the next sections, we study the induced scattering
105: outside of the source therefore we can assume that the
106: scattering angle is larger than the small angle subtended by
107: the primary radiation. In this case, the occupation number of
108: the scattered photons varies according to the equation
109: \eqb \frac
110: 1n\frac{dn}{dt}=\frac{3\sigma_T}{8\pi} \frac{cN}{m_e}
111: (\mathbf{e\cdot e}_1)^2
112: (1-\cos\theta)\frac{\partial}{\partial\nu}\left(\frac
113: F{\nu}\right);
114: \label{kinComp1} \eqe
115: where $F=c^{-2}h\int\nu^3nd\mathbf{\Omega}$ is the local radio
116: flux density of the primary radiation, $\theta$ the scattering
117: angle. Solution to this equation is written as
118: \eqb
119: n=n_0\exp{\tau_C};
120: \eqe
121: where $n_0$ is the background photon density, $\tau_C$ the
122: effective optical depth determined by the integral along the
123: scattered ray,
124: $\mathbf{r}=\mathbf{r}_0+c\mathbf{\Omega}(t-t_0)$, as
125: \eqb
126: \tau_C=\int \frac{3\sigma_T}{8\pi} \frac{cN}{m_e}
127: (\mathbf{e\cdot e}_1)^2
128: (1-\cos\theta)\frac{\partial}{\partial\nu}\left(\frac
129: F{\nu}\right) dt.
130: \label{tau}\eqe
131: The intensity of the scattered radiation increases
132: exponentially provided the photon spectrum of the primary beam,
133: $F/\nu$, has a positive slope. Therefore the induced scattering
134: is the most efficient just below the spectral maximum. If the
135: radiation with a decreasing spectrum is detected, one can find
136: the observability condition substituting the frequency
137: derivative in Eq.(\ref{tau}) by $F/\nu^2$ at the observed
138: frequency because the stimulated scattering rate thus estimated
139: is lower than that near the spectral maximum. As the brightness
140: temperature of the primary beam many orders of magnitude
141: exceeds the brightness temperature of the background radiation,
142: the fraction of the scattered photons remains small until
143: $\tau_C$ reaches a few dozens. As a simple criterion for the
144: observability of the primary radiation (the condition that the
145: induced scattering does not affect the primary radiation), one
146: can use the condition $\tau_C<10$.
147:
148: In order to check this condition, one can substitute the
149: undisturbed primary flux into Eq.(\ref{tau}). Let a radio pulse of
150: the duration $\Delta t$ propagate radially from the source; then
151: the primary flux can be presented in the form
152: \eqb
153: F=\left(\frac{D}r\right)^2F_{\rm obs}
154: \Theta\left(\frac{ct-r}{c\Delta t}\right);
155: \label{flux}\eqe
156: where $D$ is the distance to the source, $r$ the distance from
157: the source to the scattering point, the function $\Theta(x)$
158: describes the shape of the pulse. Below we adopt the simplest
159: rectangular form: $\Theta(x)=1$ at $0<x<1$ and $\Theta(x)=0$
160: otherwise. Note that we can ignore the transverse structure of
161: the pulse because the most efficient is the backscattering so
162: that the scattered ray interacts only with the radiation
163: emitted in the same direction.
164:
165: Note also that even though Eq.(\ref{flux}) assumes that the
166: pulse structure is attributed to the intrinsic time variation
167: of the source, the same structure arises if pulsed radiation is
168: generated by a narrow beam sweeping across the observer. In
169: this case, the radiation field has a shape
170: $\Theta\left[(ct-r-r_0\varphi)/c\Delta t\right]$, which is
171: reduced to Eq.(\ref{flux}) at $\varphi=\it const$. However, one
172: should take into account that in this paper, we assume that the
173: pulse is {\it single} in the sense that the distance between
174: pulses is larger than the scale of the scattering medium. If
175: this condition is not fulfilled, the scattered ray could pass
176: through a few pulses and then the induced scattering occurs as
177: in the steady radiation field with the intensity equal to the
178: average intensity of the source. Therefore the results of this
179: paper should be applied only to true single events like radio
180: emission from gamma-ray bursts or giant pulses from pulsars,
181: which are rare enough to be considered as isolated phenomena.
182:
183: Let a seed ray be launched in the point $r_0$ at the time
184: $t_0=cr_0$ just when the pulse reached this point. Due to
185: induced scattering of the photons from the pulse, the intensity
186: of the ray grows exponentially while the ray remains within the
187: zone illuminated by the pulse. Below we assume that the pulse
188: is narrow enough, $c\Delta t\ll r_0$. In order to find the
189: amplification factor of the seed ray, one should find the
190: effective optical depth (\ref{tau}), which could be presented
191: as
192: \eqb
193: \tau_C=\frac{3\sigma_T}{8\pi} \frac{cNF_{\rm
194: obs}}{m_e\nu^2}\left(\frac{D}{r_0}\right)^2Z
195: \label{tau1}\eqe
196: where
197: \eqb
198: Z=\int (1-\cos\theta)\left(\frac{r_0}r\right)^2
199: \Theta\left(\frac{ct-r}{c\Delta t}\right)dt
200: \label{int}\eqe
201: is the integral along the ray. For the estimates, we take
202: $\mathbf{e\cdot e}_1=1$. Let the ray be directed at the angle
203: $\theta_0$ to the radial direction at the initial point. Then the
204: scattering angle, $\theta$, and the distance from the source, $r$,
205: at the time $t$ could be found from the laws of sines and cosines
206: for the triangle in Fig. 1
207: \eqb
208: \frac{c(t-t_0)}{\sin(\theta_0-\theta)}=\frac{r_0}{\sin \theta};
209: \label{sin}\eqe \eqb
210: r^2=r_0^2+c^2(t-t_0)^2+2r_0c(t-t_0)\cos\theta_0.
211: \label{cos}\eqe
212: Eliminating $r$ and $t$, one can present the integral
213: (\ref{int}) as
214: \eqb
215: Z=\frac{r_0}{c\sin\theta_0}\int_{\theta_{\rm
216: min}}^{\theta_0}(1-\cos\theta)\theta d\theta
217: =\frac{r_0}c\frac{\theta_0-\theta_{\rm
218: min}-\sin\theta_0+\sin\theta_{\rm min}}{\sin\theta_0};
219: \eqe
220: where $\theta_{\rm min}$ is determined from the condition that
221: the function $\Theta$ vanishes, $r=c(t-\Delta t)$. This
222: condition, together with Eqs.(\ref{sin}) and (\ref{cos}),
223: yields the equation for $\theta_{\rm min}$
224: \eqb
225: \frac{\tan\theta_0}{\tan\theta_{\rm min}}-1=\frac{\Delta t}
226: {2t_0\cos\theta_0} \frac{2-\Delta t}{t_0(1-\cos\theta_0)-\Delta
227: t}.
228: \eqe
229: Taking into account that $\theta_0-\theta_{\rm min}\ll 1$ at
230: $\Delta t\ll t_0$, one gets
231: \eqb
232: \theta_{\rm min}=\left\{\begin{array}{ll}\theta_0-\frac{c\Delta
233: t\sin\theta_0}{r_0(1-\cos\theta_0)};
234: & \theta_0>\sqrt{\frac{2c\Delta t}{r_0}}; \\
235: 0 & \theta_0<\sqrt{\frac{2c\Delta t}{r_0}}.\end{array}\right.
236: \eqe
237: If $\theta_0<\sqrt{2c\Delta t/r_0}$, the scattered ray remains
238: within the illuminated area till infinity therefore
239: $\theta_{\rm min}=0$ in this case. Finally one finds
240: \eqb
241: Z=\left\{\begin{array}{ll} \Delta t\left(1-\frac{2c\Delta
242: t}{r_0\theta_0^2}+\frac{4c^2\Delta t^2}{4r_0^2\theta_0^4}\right);
243: & \theta_0>\sqrt{\frac{2c\Delta t}{r_0}}; \\
244: \frac{r_0}{6c}\theta_0^2 & \theta_0<\sqrt{\frac{2c\Delta
245: t}{r_0}}.\end{array}\right.
246: \eqe
247: One sees that the amplification factor is the same for all the
248: scattered rays launched at not too small angles,
249: $\theta_0\gg\sqrt{2c\Delta t/r_0}$. This is because decreasing
250: of the scattering rate with decreasing angle (due to the recoil
251: factor $1-\cos\theta$ in the scattering rate) is compensated by
252: increasing of the time the scattered ray spends within the
253: illuminated area. The rays launched at the angles
254: $\theta_0\lesssim\sqrt{2c\Delta t/r_0}$ spend within the
255: illuminated area the time $t-t_0\gtrsim r_0/c$; then the
256: amplification factor decreases because of decreasing of the
257: primary radiation density with the distance. Of course if the
258: amplification factor is large, it is the backscattered
259: radiation that takes the whole energy of the primary beam
260: because the backward scattering is the fastest.
261:
262: Substituting $Z=\Delta t$ into Eq.(\ref{tau1}) one can now
263: estimate the effective optical depth to the induced scattering;
264: numerically one gets
265: \eqb
266: \tau_C=0.24\frac{N_6\Delta t_{\rm s}F_{\rm obs, Jy}}{\nu_{\rm
267: GHz}^2}\left(\frac{D_{8}}{r_{-3}}\right)^2;
268: \eqe
269: where $\Delta t_{\rm s}$, $F_{\rm obs, Jy}$ and $\nu_{\rm GHz}$
270: are measured in units shown in the index, $D=10^8D_8$ pc,
271: $N=10^6N_6$ cm$^{-3}$, $r_0=10^{-3}r_{-3}$ pc. One sees that
272: the induced scattering is negligible in the interstellar medium
273: however it could become significant in dense enough environment
274: close enough to the source. For example a massive star could be
275: a progenitor of the gamma-ray burst; then the emission
276: propagates through the relic stellar wind. In this case the
277: plasma density falls off as
278: \eqb
279: N_6=0.03\frac{\dot{M}_{-5}}{V_3 r_{-3}^2};
280: \label{wind}\eqe
281: where $\dot{M}=10^{-5}\dot{M}_{-5}$ $M_{\bigodot}\cdot$yr$^{-1}$
282: is the mass loss rate, $V=10^3V_3$ km$\cdot$s$^{-1}$ the wind
283: velocity. The condition $\tau_C<10$ places the lower limit on the
284: radius beyond which the radio pulse could propagate:
285: \eqb
286: r_{-3}>0.16\left(\frac{D_{8}}{\nu_{\rm
287: GHz}}\right)^{1/2}\left(\frac{\Delta t_{\rm s}F_{\rm obs,
288: Jy}\dot{M}_{-5}}{V_3}\right)^{1/4}.
289: \label{windC}\eqe
290:
291: \section{Induced Raman scattering}
292:
293: The high intensity radio beam could be scattered by emitting
294: Langmuir waves. The energy and momentum conservation in this
295: three-wave process require that
296: \eqb
297: \nu_1=\nu+\nu_p;\qquad \mathbf{k}_1=\mathbf{k}+\mathbf{q};
298: \label{cons}\eqe
299: where $\nu_p=\sqrt{e^2N/(\pi m_e)}$ is the plasma frequency,
300: $\mathbf{q}$ the wave vector of the plasma wave. In the case
301: $\nu\gg\nu_p$ one can neglect the frequency shift of the
302: scattered wave; then one finds
303: \eqb
304: \mathbf{q}_{\pm}=\pm\frac{\omega}c
305: (\mathbf{\Omega}_1-\mathbf{\Omega}).
306: \label{qu}\eqe
307: Here the sign $+$ is associated with a plasmon emitted by the
308: photon $\mathbf{k}_1$ and the sign $-$ with a plasmon emitted
309: by the photon $\mathbf{k}$. Because of Landau damping, only
310: plasmons with large enough phase velocities could survive; this
311: places a limit on the scattering angle (Thompson et al. 1994).
312: Namely, choosing the allowable range of the plasmon wavevectors
313: from the condition that the Landau damping time exceeds the
314: period of the plasma wave, $q\lambda_D<0.27$, where
315: $\lambda_D=\sqrt{k_BT/(4\pi e^2N)}$ is the Debye length, one
316: finds from Eq. (\ref{qu}) that the backscattering is possible
317: only if $\nu<\nu_L=90N_6^{1/2}T_6^{-1/2}$ MHz. In the case
318: $\nu\gg\nu_L$, the maximum angle of scattering is
319: \eqb
320: \theta_{\rm max}=2\nu_L/\nu.
321: \label{maxangle}\eqe
322:
323: The kinetic equations for the occupation numbers of photons and
324: plasmons are written as (Thompson et al. 1994)
325: \eqb
326: \frac{\partial n(\nu,\mathbf{\Omega})}{\partial
327: t}+c(\mathbf{\Omega\cdot\nabla})n(\nu,\mathbf{\Omega})
328: \label{kinRamanph}\eqe
329: $$=\frac{3\sigma_T}{8\pi}N \frac{h\nu}{m_ec}\frac{\nu}{\nu_p}
330: \int (\mathbf{e\cdot e}_1)^2(1-\mathbf{\Omega\cdot\Omega}_1)
331: [n_{q}+n_{-q}]
332: [n(\nu,\mathbf{\Omega}_1)-n(\nu,\mathbf{\Omega})]d\mathbf{\Omega}_1;
333: $$\eqb
334: \frac{\partial n_{\pm q}}{\partial t}+v_{\rm
335: g}((\mathbf{q}/q)\cdot\nabla)n_{\pm q}
336: \label{kinRamanpl}\eqe
337: $$=\frac{3\sigma_T}{8\pi}N
338: \frac{h\nu}{m_ec}\frac{\nu}{\nu_p}\int (\mathbf{e\cdot e}_1)^2
339: \{n(\nu,\mathbf{\Omega}_1)n(\nu,\mathbf{\Omega})\pm n_{\pm
340: q}[n(\nu,\mathbf{\Omega}_1)-n(\nu,\mathbf{\Omega})]\}d\mathbf{\Omega}_1
341: -2\kappa n_{\pm q};
342: $$
343: where $n_q$ is the plasmon occupation number, $\kappa$ the
344: plasmon amplitude damping rate, $v_{\rm
345: g}=3q\lambda_D\sqrt{k_BT/m_e}$ the plasmon group velocity. The
346: last is small in the non-relativistic plasma therefore one can
347: neglect the spatial transfer of plasmons.
348:
349: As in the previous section, we assume that the primary
350: radiation subtends the angle smaller than the scattering angle
351: (\ref{maxangle}); then the scattering occurs outside the
352: primary beam because the scattering within the beam is
353: suppressed by the factor $1-\mathbf{\Omega\cdot\Omega}_1$. As
354: in the previous section, we will find the observability
355: condition demanding that the amplification factor of a weak
356: background radiation due to the Raman scattering does not
357: become exponentially large. One should stress that the Raman
358: scattering does not necessary hinders propagation of the
359: radiation even if the effective optical depth is large because
360: the scattering angle (\ref{maxangle}) may be small. Then the
361: radiation beam just widens and a special analysis is necessary
362: in order to figure out how much parameters of the emerged
363: radiation are affected. An example of such an analysis is given
364: in sect. 5. Here we just find the effective optical depth to
365: the Raman scattering.
366:
367: Assuming that the primary pulse has the form (\ref{flux}) and
368: that the intensity of the scattering radiation is small as
369: compared with the primary radiation, one reduces the kinetic
370: equations (\ref{kinRamanph}) and (\ref{kinRamanpl}) to the form
371: \eqb
372: \frac{\partial n}{\partial t}+c\cos\theta\frac{\partial
373: n}{\partial r}
374: =S\left(\frac{r_0}r\right)^2(1-\cos\theta)(n_q+n_{-q})
375: \Theta\left(\frac{ct-r}{c\Delta t}\right);
376: \label{kinRph}\eqe
377: \eqb
378: \frac{\partial n_{\pm q}}{\partial t} =S\left[(n\pm n_{\pm
379: q})\left(\frac{r_0}r\right)^2-\alpha n_{\pm q}
380: \right]\Theta\left(\frac{ct-r}{c\Delta t}\right);
381: \label{kinRpl}\eqe
382: where
383: \eqb
384: S=\frac{3\sigma_T}{8\pi}\frac{cNF_0}{m_e\nu\nu_p}
385: \left(\frac{D}{r_0}\right)^2(\mathbf{e\cdot e}_1)^2 ;\qquad
386: \alpha=\frac{F_{\kappa}}{F_{\rm obs}};\qquad
387: F_{\kappa}=\frac{16\pi m_e\nu\nu_p}{3\sigma_TcN(\mathbf{e\cdot
388: e}_1)^2}\left(\frac{r_0}{D}\right)^2\kappa.
389: \eqe
390: The plasmon decay rate due to electron-ion collisions is
391: $\kappa=0.032N_6T_6^{-3/2}\,{\rm s}^{-1}$; then
392: \eqb
393: F_{\kappa}=2.2\times 10^{-3}\frac{N_6^{1/2}\nu_{\rm
394: GHz}}{T_6^{3/2}} \left(\frac{r_{-3}}{D_8}\right)^2\,\rm Jy.
395: \label{Fkappa}\eqe
396: We assume that before the pulse arrives, some weak background
397: radiation preexists in the medium therefore the boundary
398: conditions may be written as
399: \eqb
400: n\vert_{r=ct}=n_0;\qquad n_q\vert_{r=ct}=n_{q0}.
401: \label{boundary}\eqe
402:
403: The factor $(r_0/r)^2$ in the right-hand side of Eqs.
404: (\ref{kinRph}) and (\ref{kinRpl}) arises due to decreasing of
405: the primary radiation flux (\ref{flux}) with the distance. It
406: was shown in the previous section that if the scattering angle
407: is not too small, $\theta_0\gg\sqrt{2c\Delta t/r_0}$, the
408: scattered ray remains within the illuminated area only during
409: the time $t-t_0\ll r_0/c$; then the factor $(r_0/r)^2$ may be
410: substituted by unity. Taking into account that the maximal
411: scattering angle is given by Eq.(\ref{maxangle}), this
412: condition is written as
413: \eqb
414: \frac{N_6r_{-3}}{T_6\Delta t_{\rm s}\nu^2_{\rm GHz}}\gg 6\times
415: 10^{-4}.
416: \label{condition}\eqe
417: In this case, Eqs. (\ref{kinRph}) and (\ref{kinRpl}) are easily
418: solved. Namely, transforming the variables
419: \eqb
420: v=Sct;\qquad u=S(ct-r);
421: \eqe
422: one comes to the set of equations
423: \eqb \frac{\partial
424: n}{\partial v}+(1-\cos\theta)\frac{\partial n}{\partial
425: u}=(1-\cos\theta)(n_q+n_{-q})\Theta\left(\frac{u}{cS\Delta
426: t}\right);
427: \eqe
428: \eqb
429: \frac{\partial n_{\pm q}}{\partial v}+\frac{\partial n_{\pm
430: q}}{\partial u}=\left[n-(\alpha\mp 1)n_{\pm
431: q}\right]\Theta\left(\frac{u}{cS\Delta t}\right);
432: \eqe
433: with the boundary conditions at the point $u=0$. As both
434: coefficients of the equations and the boundary conditions are
435: independent of $v$, the solution is also independent of $v$
436: therefore one finally gets a simple set of ordinary differential
437: equations at the segment $0<u<cS\Delta t$:
438: \eqb
439: \frac{d n}{du}=n_q+n_{-q}; \qquad \frac{d n_{\pm
440: q}}{du}=n-(\alpha\mp 1)n_{\pm q}.
441: \label{set}\eqe
442: The boundary conditions are: $n(0)=n_0$, $n_{\pm q}(0)=n_{q0}$.
443:
444: Partial solutions to these equations have a form $\exp(su)$ where
445: $s$ obeys the characteristic equation
446: \eqb
447: s^3+2\alpha s^2+(\alpha^2-3)s-2\alpha=0.
448: \label{charact}\eqe
449: Simple solutions are found in the two limiting cases, namely
450: when one can neglect the decay of plasmons, $\alpha=0$, and
451: when the decay is strong, $\alpha\gg 1$ (these limits
452: correspond to the conditions that the primary radiation flux is
453: well above or well below the limiting flux (\ref{Fkappa}),
454: correspondingly). In the limit $\alpha=0$ the solution to
455: Eqs.(\ref{set}) is
456: \eqb
457: n=\frac 13 [n_0+2n_0\cosh\sqrt{3}u
458: +2\sqrt{3}n_{q0}\sinh\sqrt{3}u];
459: \label{solution1}\eqe\eqb
460: n_{\pm q}=\frac 13[\mp n_0+(3n_{q0}\pm n_{0})\cosh\sqrt{3}u
461: +\sqrt{3}(n_{0}\pm n_{q0})\sinh\sqrt{3}u].
462: \eqe
463: In the limit $\alpha\gg 1$ the solution is
464: \eqb
465: n=n_0\exp\left(\frac 2{\alpha}u\right);
466: \label{solution2}\eqe\eqb
467: n_{\pm q}=\frac{n_0}{\alpha}\exp\left(\frac
468: 2{\alpha}u\right)+\left(n_{q0}-\frac{n_0}{\alpha}\right)\exp[(\pm
469: 1-\alpha)u].
470: \eqe
471: The intensity of the scattered radiation grows until $u=u_{\rm
472: max}=S\Delta t$ so the effective optical depth to the Raman
473: scattering may be estimated as $\tau_R=S\Delta t$ in the case
474: $\alpha\ll 1$ and $\tau_R=S\Delta t/\alpha$ in the opposite
475: limit. Numerically one gets
476: \eqb
477: \tau_R=29\frac{N_6^{1/2}\Delta t_{\rm s}F_{\rm obs}}{\nu_{\rm
478: GHz}}\left(\frac{D_8}{r_{-3}}\right)^2\left\{\begin{array}{ll}1;
479: & F_{\rm obs}\gg F_{\kappa}; \\
480: F_{\rm obs}/F_{\kappa}; & F_{\rm obs}\ll F_{\kappa}.
481: \end{array}\right.
482: \label{tauR}\eqe
483:
484: As in the case of the induced Compton scattering, the condition
485: for the Raman scattering to remain negligible may be written as
486: $\tau_R<10$. One can see again that the scattering in the
487: interstellar medium is negligible. Assuming that the emission
488: is generated within the stellar wind of the progenitor star
489: (see Eq.\ref{wind}), one obtains that the Raman scattering
490: could be neglected if the radio pulse was emitted at the
491: distance
492: \eqb
493: r_{-3}>0.8D_8^{2/3}\left(\frac{\Delta t_{\rm s}F_{\rm
494: obs,Jy}}{\nu_{\rm GHz}}\right)^{1/3}
495: \left(\frac{\dot{M}_{-5}}{V_3}\right)^{1/6}\left\{\begin{array}{ll}1;
496: & F_{\rm obs}\gg F_{\kappa}; \\
497: (F_{\rm obs}/F_{\kappa})^{1/3}; & F_{\rm obs}\ll F_{\kappa}.
498: \end{array}\right.
499: \label{windR}\eqe
500: from the source.
501:
502: This result was obtained under the condition (\ref{condition}),
503: i.e. if the scattering angle is not too small and the
504: interaction of the scattered ray with the primary pulse occurs
505: at the scale small enough that one can neglect decreasing of
506: the primary radiation flux with radius. Therefore we neglected
507: the factor $(r_0/r)^2$ in the right-hand side of
508: Eqs.(\ref{kinRph}) and (\ref{kinRpl}). In the opposite limit,
509: the scattered ray remains within the illuminated zone for a
510: long time however due to decreasing of the radiation flux, only
511: the region $r-r_0\sim r_0$ contributes to the effective optical
512: depth (cp. Eq.(13) and discussion after). As the solutions
513: (\ref{solution1}) and (\ref{solution2}) are valid at
514: $r-r_0<r_0$, one can find the amplification factor substituting
515: into these solutions $u_{\rm max}$ corresponding to the radius
516: $r=2r_0$. It follows from the scattering geometry (see Fig.1
517: and Eq.(\ref{cos})) that $u_{\rm max}=Sr_0\theta_0^2/(4c)$.
518: Substituting $\theta_0$ by the maximal scattering angle of
519: Eq.(\ref{maxangle}), one gets finally the estimate for the
520: effective optical depth at the condition opposite to that of
521: Eq.(\ref{condition})
522: \eqb
523: \tau_R=2.4\times 10^4\frac{N_6^{3/2}F_{\rm
524: obs,Jy}D_8^2}{T_6\nu^3_{\rm
525: GHz}r_{-3}}\left\{\begin{array}{ll}1;
526: & F_{\rm obs}\gg F_{\kappa}; \\
527: F_{\rm obs}/F_{\kappa}; & F_{\rm obs}\ll F_{\kappa}.
528: \end{array}\right.
529: \eqe
530: Note that within the range of applicability of this formula, it
531: gives the optical depth smaller than Eq.(\ref{tauR}).
532:
533: \section{Induced scattering within a relativistic source}
534: If a high brightness temperature radio pulse is generated in a
535: relativistic source, one can restrict parameters of the source
536: considering stimulated emission within it. Let a radio pulse
537: come from a relativistically hot plasma moving with the Lorentz
538: factor $\Gamma$. In the comoving frame, the radiation could be
539: considered as isotropic; then the kinetic equation for the
540: induced Compton scattering could be written as (Melrose 1971)
541: \eqb
542: \frac 1{n(\nu')}\frac{\partial n(\nu')}{\partial t'}=\frac
543: 3{16}\sigma_T\frac{h\nu'}{m_ec}N'
544: \int d\gamma\frac{\partial}{\partial\gamma'}\left(\frac{f(\gamma')}{\gamma'^2}\right)
545: \int_0^{\infty}\frac{d\nu'_1}{\nu'_1}\left(1-\frac{\nu'_1}{\nu'}\right)
546: g\left(\frac{\nu'_1}{\nu'}\right)n(\nu'_1);
547: \label{eq_melrose}\eqe
548: where $f(\gamma')$ is the electron distribution function
549: normalized as $\int f(\gamma')d\gamma'=1$; the primed
550: quantities are measured in the comoving frame. The kernel $g$
551: is approximated as (correcting a typo in Melrose's paper)
552: \eqb
553: g(x)=\left\{\begin{array}{ll}4x^2;
554: & (2\gamma')^{-2}\le x\le 1; \\
555: 4x; & 1\le x\le 4\gamma'^2; \\
556: 0; & {\rm otherwise}.\end{array}\right.
557: \eqe
558: The right-hand side of Eq.(\ref{eq_melrose}) is the induced
559: scattering rate; it should be compared with the rate of photon
560: escape from the source, $c/l'$, where $l'$ is the characteristic
561: size of the emitting region. If the emitting plasma moves, as is
562: typically the case, radially from the origin, one should also take
563: into account that the plasma density decreases in the proper frame
564: with the rate $c\Gamma/r$. Then the condition that the induced
565: scattering does not affect the emerged radiation is written as
566: \eqb
567: \frac 1{n(\nu')}\frac{\partial n(\nu')}{\partial t'}
568: <\max\left(\frac c{l'},\frac{c\Gamma}{r}\right).
569: \label{ind_cond}\eqe
570:
571: In order to estimate the stimulated scattering rate, let us
572: assume that the particle distribution is Maxwellian
573: %with the thermal spread in Lorentz factors $\Gamma_T$
574: \eqb
575: f(\gamma)=\frac{\gamma^2}{2\gamma_T^3}\exp\left(-\frac{\gamma}{\gamma_T}\right)
576: \eqe
577: and that the radiation spectrum has a form
578: \eqb
579: I(\nu')=I_0\left\{\begin{array}{ll}(\nu'/\nu'_0)^a;
580: & \nu'<\nu'_0; \\
581: (\nu'/\nu'_0)^{-b}; & \nu'>\nu'_0;\end{array}\right.
582: \eqe
583: where $I(\nu')=h\nu'^3n(\nu')/c^2$ is the radiation intensity,
584: $a>2$, $b>0$. The frequency of the photon decreases in the
585: course of stimulated scattering in the isotropic medium
586: therefore the right-hand side of Eq.(\ref{eq_melrose}) is
587: positive for $\nu'<\nu'_0$. Upon integrating one gets
588: \eqb
589: \frac 1{n(\nu')}\frac{\partial n(\nu')}{\partial t'}= \frac
590: 3{8}\frac{\sigma_TN'cI_0}{m_e\gamma_T^3\nu'_0\nu'}
591: \Phi\left(\frac{\nu'_0}{4\gamma_T^2\nu'}\right);
592: \eqe $$
593: \Phi(x)=\frac{a+b}{(a-1)(b+1)}e^{-\sqrt{x}}+\frac{x^{-(a-1)}}{a-1}\int_0^{\sqrt{x}}y^{2(a-1)}e^{-y}dy
594: $$\eqb
595: -\frac{x^{b+1}}{b+1}\int_{\sqrt{x}}^{\infty}y^{-2(b+1)}e^{-y}dy
596: =\left\{\begin{array}{ll}(\frac{a+b}{(a-1)(b+1)} & x\ll 1;
597: \\ \frac{\Gamma(2a-1)}{a-1}x^{-(a-1)}; &
598: x\gg 1;\end{array}\right.
599: \eqe
600: where $\Gamma(x)$ is the gamma-function. One sees that the
601: scattering rate is maximal at $\nu'\sim\nu'_0/(2\gamma_T)^2$, the
602: exact value depending on $a$ and $b$. Substituting $\nu'=\nu'_0/(2\gamma_T)^2$ and $\Phi=1$,
603: one gets an estimate of the induced scattering rate:
604: \eqb
605: \frac 1{n}\frac{\partial n}{\partial t'}= \frac
606: 32\frac{\sigma_TN'cI_0}{m_e\gamma_T\nu'^2_0}.
607: \label{rate}\eqe
608:
609: In order to check the observability condition (\ref{ind_cond}),
610: one should substitute $\nu'_0=\nu_{\rm obs}/\Gamma$ into the
611: formula (\ref{rate}) for the induced scattering rate and
612: express $I_0$ via the observed flux. If the source size is
613: small so that the proper light travel time, $l'/c$, is less
614: than the proper expansion time, $r/(c\Gamma)$, the source
615: radiates within the angle $1/\Gamma$ and the luminosity may be
616: expressed via the observed flux as $L=(\pi/2)F_{\rm
617: obs}\nu_{\rm obs}D^2\Gamma^{-4}$. On the other hand, the
618: luminosity, which is the relativistic invariant, could be
619: calculated in the proper frame as $L=8\pi^2l'^2\nu'_0I_0$
620: (assuming the source is spherical in the comoving frame). This
621: yields
622: \eqb
623: I_0=\frac{F_{\rm obs}D^2}{16\pi\Gamma^3 l'^2};\qquad l'<\frac
624: r{\Gamma}.
625: \eqe
626: In the opposite case $l'>r/\Gamma$, one can imagine a radially
627: expanding plasma radiating forward so that the local radiation
628: flux is $F\nu=4W'c\Gamma^2$, where $W'=4\pi I_0\nu'_0/c$ is the
629: radiation density in the comoving frame. Then one can write
630: \eqb
631: I_0=\frac{F_{\rm obs}D^2}{16\pi\Gamma r^2};\qquad l'>\frac
632: r{\Gamma}.
633: \eqe
634: Now the observability condition (\ref{ind_cond}) is written as
635: \eqb
636: \frac{3\sigma_TN'F_{\rm obs}D^2}{32\pi\gamma_Tm_e\nu_{\rm
637: obs}^2}<\left\{\begin{array}{ll} \Gamma l'; & l'<r/\Gamma;
638: \\ r; & l'>r/\Gamma.\end{array}\right.
639: \label{observ_rel}\eqe
640: For any specific radiation model, one can check the
641: observability condition substituting parameters of the emitting
642: plasma in Eq. (\ref{observ_rel}).
643:
644: For a rather general preliminary estimate, one can express the
645: plasma density in the source via the fraction $\zeta$ of the
646: plasma energy radiated in the pulse. Only a small fraction of
647: the plasma energy could typically be radiated in the radio band
648: so that one can expect $\zeta\ll 1$ however one can not exclude
649: a priori a larger $\zeta$ (and even $\zeta>1$ for a Poynting
650: dominated source). We will see that this uncertainty is
651: compensated by a very weak dependence of the result on $\zeta$.
652: If the source is small, $l'<r/\Gamma$, the total radiated
653: energy is estimated as $ E_{\rm rad}=\pi D^2\nu_{\rm obs}
654: F_{\rm obs}\Delta t_{\rm obs}/\Gamma^2$ whereas the total
655: plasma energy in the source is $E_{\rm pl}=4\pi
656: l'^3N'\gamma_Tmc^2\Gamma$, where $m=m_e$ in the
657: electron-positron plasma and $m=m_p$ in the electron-ion
658: plasma. In the opposite limit, $l'>r/\Gamma$, one should
659: compare the plasma energy density, $\varepsilon_{\rm
660: pl}=3mc^3N'\gamma_T\Gamma^2$, with the radiation energy
661: density, $\varepsilon_{\rm rad}=F_{\rm obs}\nu_{\rm
662: obs}(D/r)^2$. Now one can write
663: \eqb
664: N'=\frac{F_{\rm obs}\nu_{\rm obs}D^2}{mc^2\gamma_T\zeta}
665: \left\{\begin{array}{ll} \Delta t/(4l'^3\Gamma^3); &
666: l'<r/\Gamma;
667: \\ 1/(3cr^2\Gamma^2); & l'>r/\Gamma.\end{array}\right.
668: \eqe
669: Then the observability condition is written as
670: \eqb
671: \frac{\sigma_TF_{\rm obs}^2D^4}{\zeta\gamma_T^2m_emc^2\nu_{\rm
672: obs}\Gamma^2}<\left\{\begin{array}{ll} 128\pi l'^4/(3c\Delta
673: t); & l'<r/\Gamma;
674: \\ 32\pi r^3; & l'>r/\Gamma.\end{array}\right.
675: \eqe
676: The light travel time arguments imply that $l'<c\Delta t\Gamma$
677: if $l'<r/\Gamma$ and $r<c\Delta t\Gamma^2$ in the opposite
678: limit. Taking this into account, one finally finds that the
679: observability condition implies a lower limit on the Lorentz
680: factor of the source:
681: \eqb
682: \Gamma>100\frac{F_{\rm
683: obs,Jy}^{1/4}D^{1/2}_8}{\zeta^{1/8}\gamma_T^{1/4}(m/m_e)^{1/8}(\Delta
684: t_{\rm s})^{3/8}\nu_{\rm GHz}^{1/8}}.
685: \label{Lorents}\eqe
686:
687: Note that as the radiation within the source is nearly
688: isotropic, the induced scattering is important only if it
689: affects the spectrum of the radiation. In the case of the
690: induced Compton scattering in the relativistically hot plasma,
691: the photon frequency decreases $\sim 4\gamma_T^2$ times already
692: in a single scattering therefore the observability condition
693: for the induced Compton scattering is the condition that the
694: source is just transparent with respect to this process. The
695: frequency change in the Raman scattering is small therefore the
696: corresponding observability condition is less restrictive.
697:
698: \section{Implications for the observed short extragalactic pulse}
699:
700: Let us apply the obtained general observability conditions to
701: the enigmatic radio pulse recently found by Lorimer et al.
702: (2007) in a pulsar survey at the frequency 1.4 GHz. The
703: duration of the pulse was $\Delta t\le 5$ ms, the energy in the
704: pulse $F_{\rm obs}\Delta t=0.15\pm 0.05$ Jy$\cdot$s. The
705: dispersion measure is an order of magnitude larger than the
706: expected contribution from the Milky Way and moreover, no
707: galaxy was found at the position of the source. This lead
708: Lorimer et al. to conclude that the source of the pulse is on
709: cosmological distance; they give a very rough estimate $D\sim
710: 500$ Mpc. Origin of this pulse is obscure; Popov \& Postnov
711: (2007) argue, on the statistical grounds, that this event could
712: be related to a hyperflare from an extragalactic soft gamma-ray
713: repeater.
714:
715: Substituting the parameters of the pulse into Eqs. (\ref{windC}),
716: one concludes that if the pulse was generated within a stellar
717: wind, the induced Compton scattering places the lower limit on
718: the emission radius, $r>6\cdot 10^{14}(\dot M_{-5}/V_3)^{1/4}$ cm.
719: A stronger limit is imposed by the Raman scattering; Eq.
720: (\ref{windR}) yields
721: \eqb
722: r>5\cdot 10^{15}\left(\frac{\dot M_{-5}}{V_3}\right)^{1/6}\left(\frac D{500 \rm Mpc}\right)^{2/3}
723: \rm cm.
724: \label{rad}\eqe
725: One should note that according to Eq.(\ref{maxangle}), the angle of the Raman
726: scattering is small in this case, $\theta_{\rm max}=0.024(V_3/\dot
727: M_{-5})^{1/6}T_6^{-1/2}$, so that the Raman scattering does not
728: hinder propagation of the pulse. However, scattering even by this
729: small angle implies the temporal smearing of the pulse above the
730: observed limit unless the pulse was initially collimated within
731: the angle $\vartheta<2.4\cdot 10^{-4}(\Delta t/5 {\rm ms})^{1/2}$.
732: In the last case, the initial radiation flux in the pulse should
733: have been $(\vartheta/\theta_{\max})^2=10^4\Delta t T_6(\dot
734: M_{-5}/V_3)^{1/3}$ times larger than that estimated above under
735: the no scattering assumption; then the induced scattering would
736: imply even stronger constraints. Therefore in any case, the lower
737: limit (\ref{rad}) for the emission radius is robust provided the
738: pulse was generated within the relic stellar wind of the
739: progenitor star.
740:
741: The observed pulse was definitely generated within relativistic
742: plasma. The induced Compton scattering within the source place a
743: limit on the Lorentz factor of the emitting plasma. Substituting
744: the observed parameters of the pulse into Eq. (\ref{Lorents}), one
745: gets
746: \eqb
747: \Gamma>3800\gamma_T^{-1/4}\left(\frac{\Delta t}{5 \rm ms}\right)^{-5/8}
748: \left(\frac{m_e}{\zeta m}\right)^{1/8}\left(\frac D{500 \rm Mpc}\right)^{1/2}.
749: \eqe
750:
751: \section{Conclusions}
752:
753: In this paper, we have analyzed the effect of induced Compton
754: and Raman scattering on propagation of a short bright radio
755: pulse. The work was motivated by predictions that such pulses
756: could accompany gamma-ray bursts (Usov \& Katz 2000; Sagiv \&
757: Waxman 2002; Moortgat \& Kuijpers 2005) and by a recent
758: discovery of a single extragalactic radio pulse (Lorimer et al.
759: 2007). Macquart (2007) claimed that induced scattering in the
760: interstellar medium strongly limits the observability of high
761: brightness temperature transients.
762: %In his analysis, he used limits on the brightness temperature of
763: %the radiation obtained for the steady sources. However, the
764: However he ignored two fundamental properties of the process.
765: First of all, the induced scattering occurs only if the
766: scattered ray remains within the zone illuminated by the
767: scattering radiation. In the case of a single short pulse, the
768: effective optical depth is determined by the duration of the
769: pulse but not by the scale of the scattering medium. Therefore
770: a short enough pulse could propagate freely through the
771: interstellar medium. The second important property is that in
772: the presence of a powerful radiation beam, even a weak
773: background radiation grows exponentially via the induced
774: scattering of the beam photons. If the primary beam is narrow,
775: the scattering outside the beam dominates because the
776: scattering within the beam is suppressed by the recoil factor
777: $1-\mathbf{\Omega\cdot\Omega}_1$ in the scattering rate.
778: Outside the source, the radiation subtends a small solid angle
779: therefore the induced scattering in the surrounding medium
780: occurs outside the beam. In this case, the effective optical
781: depth depends not on the brightness temperature and the angle
782: subtended by the primary radiation, which could not be found
783: separately without model assumptions, but only on the radiation
784: flux, which is a directly observable quantity. We demonstrated
785: that the induced scattering in the surrounding medium could
786: hinder the escape of a bright short pulse only if the source is
787: embedded into a dense medium, like stellar wind. We estimated a
788: limiting radius beyond which the pulse could propagate.
789:
790: One should stress that these estimates assume a single pulse.
791: If a sequence of pulses (e.g., pulsar emission) propagates in
792: the medium with the characteristic scale exceeding the distance
793: between the pulses, the induced scattering occurs as if the
794: emission was steady with the average radiation flux. On the
795: other hand, our results could be applied to giant pulses from
796: pulsars, which are rare enough to be considered as isolated
797: phenomena.
798:
799: We have also analyzed induced scattering within the
800: relativistically moving source. Transparency of the source is
801: determined by the radiation intensity and by amount of plasma
802: within the source. %The last is limited from below by demanding
803: %that the total radiated energy does not exceed the total plasma
804: %energy unless the flow is Poynting dominated. This yields
805: Introducing a fraction $\zeta$ of the plasma energy emitted in
806: the pulse, we found a lower limit on the Lorentz factor of the
807: source, which turned out to be very weakly dependent on
808: $\zeta$.
809:
810: \acknowledgements
811: This work was supported by the German-Israeli Foundation for
812: Scientific Research and Development.
813:
814: \begin{thebibliography}{99}
815:
816: \bibitem[]{791} Coppi, P., Blandford, R. D., \& Rees, M. J. 1993,
817: \mnras, 262, 603
818: \bibitem[]{793} Gedalin, M., Eichler, D., 1993, \apj, 406, 62
819: \bibitem[]{794} Eichler, D., 1991, \apj, 370, L27
820: \bibitem[]{795} Levinson, A., Blandford, R., 1995, \mnras, 274, 717
821: \bibitem[]{796} Lorimer, D. R.; Bailes, M.; McLaughlin, M. A.;
822: Narkevic, D. J.; Crawford, F. 2007, Sci, 318, 777
823: \bibitem[]{798} Luo, Q., \& Melrose, D. B. 1995, \apj, 452, 346
824: \bibitem[]{799} Luo, Q., \& Melrose, D. B. 2006, \mnras, 371, 1395
825: \bibitem[]{800} Lyubarskii, Yu. E. \& Petrova, S. A.
826: 1996, AstL 22, 399
827: \bibitem[]{802} Lyutikov, M. 1998, \mnras, 298, 1198
828: \bibitem[]{803} Macquart, J.-P., 2007, \apj, 658, L1
829: \bibitem[]{804} Melrose, D.~B. 1971, \apss, 13, 56
830: \bibitem[]{805} Montes, C. 1977, \apj, 216, 329
831: \bibitem[]{806} Moortgat, J., Kuijpers, J. 2005, gr-qc/0503074
832: \bibitem[]{807} Petrova, S. A. 2004a, \aap, 417, L29
833: \bibitem[]{808} Petrova, S. A. 2004b, \aap, 424, 227
834: \bibitem[]{809} Petrova, S. A. 2007a, arXiv0709.3173
835: \bibitem[]{810} Petrova, S. A. 2007b, arXiv0709.4578
836: \bibitem[]{811} Popov, S. B. \& Postnov, K. A., 2007,
837: arXiv:0710.2006
838: \bibitem[]{813} Sagiv, A., Waxman, E. 2002, \apj, 574, 861
839: \bibitem[]{814} Sincell, M.W., \& Coppi, P. S. 1996, \apj, 460, 163
840: \bibitem[]{815} Sunyaev, R.A. 1971, Sov. Astron., 15, 190
841: \bibitem[]{816} Thompson, C., Blandford, R.~D., Evans, C.~R.,
842: Phinney, E.~S., 1994, \apj, 422, 304
843: \bibitem[]{818} Usov, V.~V.; Katz, J.~I. 2000, \aap, 364, 655
844: \bibitem[]{819} Wilson, D.~B., 1982, \mnras, 200, 881
845: \bibitem[]{820} Wilson, D.~B. and Rees, M.~J., 1978, \mnras, 185,
846: 297
847: \bibitem[]{822} Zel'dovich, Ya. B., Levich, E. V., \& Sunyaev, R.
848: A. 1972, Sov. Phys. JETP, 35, 733
849: \end{thebibliography}
850:
851: \clearpage
852: \begin{figure*}
853: \includegraphics[scale=0.6]{f1.eps}
854: \caption{Geometry of the scattering. The primary pulse propagates
855: radially from the point O. Just when it arrives at the point A (at
856: the time $t_0$), a seed ray is launched at the angle $\theta_0$.
857: At the time $t$, the seed ray arrives at the point S where it
858: makes the angle $\theta$ with the radial direction.}
859: \end{figure*}
860: \end{document}
861: