1: \documentclass[12pt,preprint]{aastex}
2: %%%%%%%%%%%%%%%%%%%%% macro definitions %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: \newcommand \be {\begin{equation}}
4: \newcommand \ee {\end{equation}}
5: \newcommand \Angstrom {\,{\rm \AA}}
6: \newcommand \eV {\,{\rm eV}\,}
7: \newcommand \K {\,{\rm K}}
8: \newcommand \ab {{\,{\rm a}}}
9: \newcommand \g {\,{\rm g}}
10: \newcommand \m {\,{\rm m}}
11: \newcommand \mum {\,\mu{\rm m}}
12: \newcommand \mic {\,\mu{\rm m}}
13: \newcommand \cm {\,{\rm cm}}
14: \newcommand \km {\,{\rm km}}
15: \newcommand \s {\,{\rm s}}
16: \newcommand \erg {\,{\rm erg}}
17: \newcommand \kms {\,{\rm km \, s}^{-1}}
18: \newcommand \pc {\,{\rm pc}}
19: \newcommand \yr {\,{\rm yr}}
20: \newcommand \Myr {\,{\rm Myr}}
21: \newcommand \nH {n_{\rm H}}
22: \newcommand \xH {x_{\rm H}}
23: \newcommand \urad {u_{\rm rad}}
24: \newcommand \Qabs {Q_{\rm abs}}
25: \newcommand \Qpr {Q_{\rm pr}}
26: \newcommand \simlt {\lesssim}
27: \newcommand \simgt {\gtrsim}
28: \newcommand \gtsim {\gtrsim}
29: \newcommand \ltsim {\lesssim}
30: \newcommand \E {\erg \cm^{-3}}
31: \newcommand \vej {v_{\rm ej}}
32:
33:
34: %------- delete following for submission to ApJ --------
35: %\pagestyle{myheadings}
36: % today's date
37: %use number register 200 for ``decade''
38: %\countdef\decade=200
39: %\decade=0
40: %\advance\decade by \year
41: %\advance\decade by -2000 %to suppress two leading digits of yearb
42: %\countdef\hours=201
43: %\hours=0
44: %\advance\hours by \time
45: %\divide\hours by 60
46: %\countdef\mins=202
47: %\mins=0
48: %\advance\mins by \hours
49: %\multiply\mins by 60
50: %\multiply\hours by 100
51: %\countdef\miltime=203
52: %\miltime=0
53: %\advance\miltime by \hours
54: %\advance\miltime by \time
55: %\advance\miltime by -\mins
56: %\def\today{\number\decade.\number\month.\number\day.\number\miltime}
57: %\markright{\today: DRAFT}
58:
59: \shorttitle{Extra-Solar Dust}
60: \shortauthors{Murray Weingartner \& Capobianco}
61:
62: \begin{document}
63:
64: \title{On the Flux of Extra-Solar Dust in Earth's Atmosphere} %\\
65: % {\small DRAFT: \today}}
66:
67: \author{N. Murray\altaffilmark{1}, Joseph C. Weingartner \& P. Capobianco\altaffilmark{2}}
68: \affil{CITA, 60 St. George Street, University of
69: Toronto, Toronto, ON M5S 3H8, Canada}
70: \email{murray@cita.utoronto.ca, weingart@cita.utoronto.ca}
71: \altaffiltext{1}{Canada Research Chair in Astrophysics}
72: \altaffiltext{2}{Current address: Department of Physics and Astronomy,
73: University of Victoria, Elliott Building, 3800 Finnerty Rd.,
74: Victoria, BC, V8P 1A1, Canada}
75:
76: \begin{abstract}
77: Micron size extrasolar dust particles have been convincingly detected
78: by satellites. Larger extrasolar meteoroids (5-35 microns) have most
79: likely been detected by ground based radar at Arecibo and New
80: Zealand. We present estimates of the minimum detectable particle sizes
81: and collecting areas for both radar systems. We show that particles
82: larger than $\sim10\micron$ can propagate for tens of parsecs through
83: the interstellar medium, opening up the possibility that ground based
84: radar systems can detect AGB stars, young stellar objects such as T
85: Tauri stars, and debris disks around Vega-like stars. We provide
86: analytical and numerical estimates of the ejection velocity in the
87: case of a debris disk interacting with a Jupiter mass planet. We
88: provide rough estimates of the flux of large micrometeoroids from all
89: three classes of sources. Current radar systems are unlikely to detect
90: significant numbers of meteors from debris disks such as Beta
91: Pictoris. However, we suggest improvements to radar
92: systems that should allow for the detection of multiple examples of
93: all three classes.
94: \end{abstract}
95:
96: \keywords{}
97:
98:
99: \section{\label{sec:intro} INTRODUCTION}
100:
101: In astronomy, as in everyday life, most of our information comes to us
102: in the form of electromagnetic radiation. Some astronomical systems
103: also emit solid particles, which could yield valuable information if
104: detected. The dust detectors on the {\it Ulysses} and {\it Galileo}
105: spacecraft have yielded preliminary information on the flux of
106: interstellar grains passing through the solar system (Frisch et
107: al.~1999; Landgraf et al.~2000). Due to the limited area of the
108: collecting surfaces, of order $200\cm^2$, these authors were unable to
109: report fluxes for grains with masses $\gtsim 10^{-10} \g$. These
110: low-mass grains cannot be traced to their point of origin, since
111: interstellar influences (e.g., gas drag, radiation pressure, magnetic
112: fields) rapidly adjust their velocities.
113:
114: Particle fluxes probably decline dramatically as the grain mass increases,
115: necessitating dust detectors with much larger collecting areas.
116: Meteor-tracking radar facilities can be used to detect grains
117: originating beyond the solar system, if the grain's initial velocity can
118: be inferred with sufficient accuracy. The effective collecting area in
119: this case could be $\gtsim 10^4 \, {\rm km}^2$.
120:
121: Recently, Baggaley and coworkers reported the detection of extra-solar
122: grains using the Advanced Meteor Orbit Radar, or AMOR (Baggaley et
123: al.~1994; Baggaley 2000). Baggaley (2000) finds a ``discrete source'' with
124: an angular diameter of $\sim 30\arcdeg$, which he tentatively
125: associates with $\beta$ Pictoris. Inspection of his Fig. (2) suggests
126: the presence of a broad, band-like feature as well. Baggaley does not
127: provide any estimate of the particle fluxes.
128:
129: Our goal here is to identify likely sources of extra-solar grains and
130: estimate expected fluxes at Earth, as a function of the grain size.
131: We consider three types of objects that could potentially yield
132: significant fluxes: young main sequence stars, asymptotic giant branch
133: stars (AGB stars), and young stellar objects (YSOs).
134:
135: For a given source of interest, the flux at Earth depends
136: on the following three factors: 1. The ``dust luminosity'' of the source,
137: i.e., the rate at which grains are emitted. We also need
138: to know whether or not the emission is isotropic. If not, then we need
139: to know the orientation of the source. 2. The distance between the source
140: (at the time when the grains were ejected) and the Sun (now). 3. The
141: probability that the grain survives the trip and is not deflected on its way.
142: Although deflected grains can be detected at Earth and will contribute to
143: the general dust background, they do not reveal their source.
144:
145: Define the ``specific dust luminosity'' $L_{v,a}(t, \vej, a)$ by
146: \be
147: L(t) = \int d\vej \, da \, L_{v,a}(t, \vej, a)~~~,
148: \ee
149: where $t$ is the age of the source,
150: $L(t)$ is the dust luminosity, $\vej$ is the speed at
151: which grains are ejected, and $a$ is the grain radius.\footnote{Throughout
152: this paper we make the simplifying assumption that the grains are spheres.}
153: In order to calculate the dust flux at Earth, it is most convenient to
154: consider the reference frame in which the source is stationary. We assume
155: that, in this frame, large grains simply travel radially outward; i.e.,
156: we ignore the Galactic potential and gravitational interactions with
157: individual stars, as well as any other influences that could deflect the
158: grains (see \S \ref{sec:magnetic_deflection}).
159: The number density of grains at distance $d$ from the source
160: is given by $n(t, d) = \int d\vej \, da \, n_{v,a}(t, \vej, a, d)$,
161: where
162: \be
163: n_{v,a}(t, \vej, a, d) = \frac{L_{v,a}(\vej, a, t_{\rm ej} = t-d/\vej)}
164: {4 \pi d^2 \vej} f_{\rm beam} \, f_{\rm survive}(\vej, a, d)~~~;
165: \ee
166: $t_{\rm ej}$ is the age of the source when the grains are ejected.
167: The factor $f_{\rm beam}$ accounts for anisotropic emission from the
168: source (we assume that it is independent of $\vej$ and $a$) and
169: $f_{\rm survive}(\vej, a, d)$ is the fraction of the grains that survive the
170: trip out to distance $d$ without being destroyed or significantly
171: deflected. The particle flux at Earth at the present time is given by
172: $F(t) = \int d\vej \, da \, F_{v,a}(t, \vej, a, d)$, where
173: \be
174: \label{eq:specific_flux}
175: F_{v,a}(t, \vej, a, d) = n_{v,a}(t, \vej, a, d) v_{d, \sun} =
176: \frac{L_{v,a}(\vej, a, t_{\rm ej} = t - d/\vej)}{4 \pi d^2} \frac{v_{d, \sun}}
177: {\vej} f_{\rm beam} \, f_{\rm survive}(\vej, a, d)~~~.
178: \ee
179: The velocity of the dust grain with respect to the Sun is given by
180: \be
181: \label{eq:v_dust_Sun}
182: \vec{v}_{d, \sun} = \vec{v}_{\ast, \odot} - \vej \hat{r}_{\ast, \sun}~~~,
183: \ee
184: where $\vec{v}_{\ast, \sun}$ is the velocity of the source with respect
185: to the Sun and $\hat{r}_{\ast, \sun}$ is the unit vector pointing from the
186: Sun to the source.
187:
188: In \S \ref{sec:detectable_fluxes}, we estimate the limiting particle
189: flux that could potentially be detected with radar facilities.
190: In \S \ref{sec:survive}, we consider the mechanisms that can prevent
191: a grain from reaching us and estimate the (size-dependent) distance that a
192: grain can travel and still reveal its source. We estimate observable
193: dust fluxes from young main sequence stars, AGB stars, and YSOs in
194: \S \S \ref{sec:Vega-like}, \ref{sec:AGB}, and \ref{sec:YSOs},
195: respectively. We discuss our results in \S \ref{sec:Discussion}, and
196: summarize our conclusions in \S \ref{sec:conclusions}.
197:
198: \section{\label{sec:detectable_fluxes} DETECTABLE FLUXES}
199: \subsection{Satellite Measured Fluxes}
200: There have been several claimed detections of extrasolar meteors, both
201: from satellite detectors and ground based radar. The most convincing
202: detections are those of the Ulysses and Galileo satellites. Frisch et
203: al. (1999) and Landgraf et al. (2000) show that the flux of
204: $m\approx6\times10^{-13}\g$ extrasolar particles is given by
205: $mf_m\approx10^{-9}\cm^{-2}\s^{-1}$. Here $f_mdm$ is the flux of
206: particles with masses between $m$ and $m+d m$.\footnote{ The cited
207: authors use the cumulative flux in a logarithmic mass interval,
208: $f_{\log_{10} m}d(\log m)=f_mdm$. The two fluxes are related by
209: $mf_m=f_{\log_{10} m}/\ln10$. } At the low mass end of the mass
210: distribution (below $m\approx6\times10^{-13}\g$) the particles are subject to
211: strong perturbations from the solar wind and its associated magnetic
212: field, so that the measured flux is not representative of the flux of
213: small particles outside the heliosphere. Above $m\approx 6\times10^{-13}\g$,
214: they find $mf_m\propto m^{-1.1}$; there are
215: roughly equal masses of particles in every logarithmic mass bin.
216:
217: For purposes of extrapolation we will occasionally assume that the
218: cumulative flux of
219: interstellar particles of mass $m$ is
220: %
221: \be \label{eq:observed flux}%$
222: f_{\log_{10} m}\approx3\times10^{-9}
223: \left({6\times10^{-13}\g\over m}\right)^{1.1}
224: \cm^{-2}\s^{-1},
225: \ee %$
226: %
227: for $m>6\times10^{-13}\g$.
228:
229: We employ two other differential fluxes. The first is $f_a$, where
230: $f_ada$ is the flux of particles with radii between $a$ and $a+da$;
231: $f_a$ has units $\cm^{-2}\s^{-1}\cm^{-1}$. In many cases one assumes
232: that $f_a$ follows a power law distribution, $f_a\propto
233: a^{-\gamma}$. This corresponds to $f_m\propto m^{-\alpha}$, with
234: $\alpha=(\gamma+2)/3$. We also use the differential mass flux,
235: $m^2f_m$, with units $\g\cm^{-2}\s^{-1}$.
236:
237: In the asteroid literature one encounters the Dohnanyi (1969) law,
238: $f_m\propto m^{-\alpha}$, with $\alpha=11/6$, corresponding to
239: $\gamma=7/2$. The Dohnany law represents a singular, steady state
240: solution to a set of equations describing a closed system of colliding
241: bodies. Asteroids do not follow this relation\cite{Ivezic}. The
242: same scaling appears in the interstellar dust literature, in the MRN
243: dust model (Mathis, Rumpl, \& Nordsieck 1977). In these models the
244: mass in large particles diverges in proportion to $\sqrt{a}$, or as
245: $m^{1/6}$. The observed scaling (eqn. \ref{eq:observed flux}) is
246: $\alpha\approx2.1$, slightly larger than the Dohnanyi value, so that
247: the mass in large particles is finite. We note that particles in
248: debris disks or in the outflows of AGB stars need not be distributed
249: according to either law; in fact the findings below suggest that they are not
250: well described by a power law with $\gamma=7/2$.
251:
252: \subsection{Radar Fluxes}
253: Ground based radar at Arecibo \citet{2002ApJ...567..323M} and in New
254: Zealand (AMOR) \citet{TBS96, Bag00} have also reported
255: detections of extrasolar meteors. When a meteoroid enters the Earth's
256: atmosphere, air molecules ablate and ionize material from the
257: meteoroid. The free electrons created by this ablation reflect radio
258: waves, a fact exploited by radar aficionados.
259:
260: The size of individual meteoroids detected by radar systems can be
261: inferred by three different methods; by the radar power reflected from
262: a Fresnel zone of the meteor trail (\S \ref{sec:Fresnel}), by the
263: power reflected from the meteor head (\S \ref{sec:head}), or by the
264: deceleration of the meteoroid as it ablates in the Earth's atmosphere
265: (also in \S \ref{sec:head}).
266:
267: The first two size estimates rely on the power of the reflected radar
268: signal. For radar micrometeors the relevant reflection is
269: coherent. The wavelengths employed are comparable to or larger than
270: the initial width $r_0$ of the ionization trail, so the radar is
271: sensitive to the electron line density $q$ rather than the space
272: density. The reflected power depends on the transmitted power, the
273: distance between the radar and the meteoroid, the mass and velocity of
274: the meteoroid, and the (uncertain) ionization efficiency. Here we
275: quantify this relation. A good introduction to the subject of radar
276: meteors is given by \citet{McKinley61}.
277:
278: \subsubsection{Meteoroid properties\label{sec:Meteor properties}}
279: The radii of radar meteoroids range from $a\sim3\mum$ to
280: $\sim40\mum$. For purposes of illustration, we consider meteoroids
281: with radius $a=10\mum$, density $\rho=3\g\cm^{-3}$, and made of
282: silicates such as Forsterite, Mg$_2$SiO$_4$. Our default meteoroid has a
283: mean molecular weight of $140$ grams per mole, and a vaporization
284: temperature of order 2000K. We assume that the energy per bond is
285: $\sim2$ eV. This is equivalent to a heat of ablation of
286: $\zeta\sim10^{11}$ ergs per gram.
287:
288: The mass of our default meteoroid is $m=1.3\times10^{-8}\g$. The mean
289: atomic weight is $\mu\approx20m_p$, where $m_p=1.67\times10^{-24}\g$
290: is the mass of a proton, while the total number of atoms in the meteoroid
291: is $N=m/\mu\approx4\times10^{14}$. The total binding energy of the
292: meteoroid is
293: %
294: \be \label{binding}%$
295: E_{B}\equiv\zeta m\approx 1300 \left({a\over 10\mic}\right)^3
296: \left({2\, {\rm eV}\over{\rm bond}}\right)
297: \left({\rho\over 3\g\cm^{-3}}\right){\, \rm ergs}.
298: \ee %$
299: %
300:
301: A typical meteor is highly supersonic, with velocity $v\sim40\km/\s$
302: at the top of the atmosphere, so the kinetic energy of the meteoroid is
303: %
304: \be %$
305: KE\approx 10^5\left({a\over 10\mic}\right)^3
306: \left({v\over 40\km\s^{-1}}\right)^2
307: \left({\rho\over 3\g\cm^{-3}}\right){\,\rm ergs},
308: \ee %$
309: %
310: much larger than the binding energy.
311:
312: \subsubsection{The ionization trail}
313:
314: As the meteoroid passes through the atmosphere, it collides with air
315: molecules and is ablated. Because the kinetic energy per meteoroid atom
316: is in excess of $100$ eV, while the ionization potential is of order
317: 10eV (depending on the atomic species), collisions between ablated
318: atoms and air molecules may ionize either or both particles. The
319: number of ions produced by each meteoroid atom is denoted by $\beta$ in
320: the meteor literature. It is electrons from these ionized atoms that
321: reflect the radar signal.
322:
323: If the mass loss rate of a meteoroid traveling at velocity $v$ is
324: $dm/dt$, the number of ions produced per centimeter along the flight
325: path is
326: %
327: \be \label{q}%$
328: q=-{\beta\over \mu v}{dm\over dt}.
329: \ee %$
330: %
331: The rate at which the meteoroid ablates depends on the ambient air
332: density, which increases exponentially as the meteoroid descends
333: through the atmosphere. In other words, the mass loss rate increases
334: rapidly with time. The pressure scale height $H_p$ of the atmosphere
335: is roughly constant, since $H_p=kT/(\mu_{air} g)$ and the air
336: temperature $T\approx200{\rm \,K}$. The line density $q$ reaches a
337: maximum just before the meteor vanishes, with most of the matter being
338: deposited in the last scale height of its path. Taking $-vdt=H_p$, we
339: find
340: %
341: \be \label{q approx}%$
342: q={3\over2}{\beta m\over \mu H_p},
343: \ee %$
344: %
345: where the factor of $3/2$ comes from the more accurate calculation
346: outlined in the appendix.
347:
348: The value of $\beta$ depends strongly on velocity $v$, but the range
349: of $v$ for typical extrasolar meteors at Arecibo or AMOR is
350: $20-60\km\s^{-1}$, with most near $40\km\s^{-1}$. Jones (1997) and
351: Jones \& Halliday (2001) examine $\beta$ using a combination of
352: laboratory experiments and observational data. For iron atoms they
353: find $\beta\approx0.6(v/40\km\s^{-1})^{3.12}$ (with
354: $v<60\km\s^{-1}$). For more realistic compositions they expect $\beta$
355: to be a factor of $5$ smaller. For a meteoroid with mean atomic weight
356: $\mu=20m_p$ (as for Forsterite),
357:
358: %
359: \be %$
360: q\approx6\times10^7\left({\beta\over 0.1}\right)
361: \left({a\over10\mic}\right)^3\left({\rho\over3\g\cm^{-3}}\right)
362: \left({6\km\over H_p}\right)\cm^{-1}.
363: \ee %$
364: %
365:
366: The height at which the meteoroid ablates can be estimated by finding the
367: height at which the binding energy of the meteoroid decreases most
368: rapidly. The rate of change of the binding energy is
369: %
370: \be \label{ablation height} %$
371: \zeta{dm\over dt}= -\Lambda \cdot\pi a^2\cdot{1\over2}\rho_av^3,
372: \ee %$
373: %
374: where $\rho_a$ is the mass density of the atmosphere at the height of
375: the meteoroid and $\Lambda$ is the fraction of the kinetic energy flux that goes
376: toward ablating the meteor. Taking the derivative of this expression
377: with respect to time and setting the result equal to zero yields the
378: relation
379: %
380: \be %$
381: -{2\over3m}{dm\over dt}={1\over\rho_a}{d\rho_a\over dt}+{3\over
382: v}{dv\over dt}.
383: \ee %$
384: %
385: We show in the appendix that the last term is typically smaller than
386: either of the other two, or at best comparable, so we ignore it. Using
387: equation (\ref{ablation height}), and setting $d\rho_a/dt=\rho_a
388: v/H_p$, we find that the meteoroid ablates when the atmospheric density is
389: %
390: \be %$
391: \rho_a^*=\rho {4\over\Lambda}\left({a\over H_p}\right)
392: \left({\zeta\over v^2}\right)
393: %$
394: \ee
395: %
396: Using typical values, we find
397: %
398: \be \label{eqn: density}%$
399: \rho_a^*\approx1.3\times10^{-9}\left({\rho\over 3\g\cm^{-3}}\right)
400: \left({10^{-1}\over \Lambda}\right)\left({a\over 10\mum}\right)
401: \left({\zeta\over10^{11}\cm^2\s^{-2}}\right)\left({40\km\s^{-1}\over v}\right)^2
402: \left({6\km\over H_p}\right)
403: \ee %$
404: %
405:
406: Using an MSIS model atmosphere\footnote{Available at
407: http://nssdc.gsfc.nasa.gov/space/model/models/msis.html} we find that
408: this density occurs at a height of $96\km$, comparable to
409: the observed height at both AMOR and Arecibo.
410:
411: On the other hand, radar data from Jodrell Bank and from Ottawa show
412: that radar meteoroids detected by those systems also ablate at heights
413: near $95\km$ (McKinley 1961), despite the fact that they have line
414: densities $q\approx10^{14}\cm^{-1}$, six orders of magnitude larger than the
415: Arecibo or AMOR meteoroids traveling at the same velocity. These larger
416: meteoroids are still much smaller than the mean free path, so that
417: aerodynamic effects seem unlikely to explain the difference;
418: nevertheless, something must differ between the two classes of
419: objects, else the more massive Jodrell Bank meteoroids would be ablated
420: five scale heights below the less massive meteoroids discussed here. One
421: possibility that has been suggested is that the more massive meteoroids
422: (those seen at Jodrell Bank and Ottawa) fragment into smaller pieces,
423: which are then rapidly ablated. The observed heights of the Arecibo
424: and AMOR events suggest that for micrometeoroids that do not fragment
425: $\Lambda\approx 0.1$ and $\zeta\approx10^{11}\cm^2\s^{-2}$.
426:
427: We stated above that the initial width $r_0$ of the ionization trail
428: was comparable to or smaller than the radar wavelength. We estimate
429: $r_0$ as follows. The cross-section of an atom is roughly $\sigma=\pi
430: d^2$, with $d\approx2\times10^{-8}\cm$; at 95km the MSIS atmosphere has
431: $T=171K$, and a mean free path of $l\approx37\cm$. However, the
432: typical meteoroid atom has an atomic mass about twice that of a nitrogen
433: atom. In a collision between a meteoroid atom and a nitrogen molecule,
434: the momentum will not be shared equally between the two N atoms; the
435: binding energy of the $N_2$ is much less than the kinetic energy of
436: the meteoroid atom. Neglecting the momentum carried away by the slower of
437: the two rebounding N atoms, we find that the final velocity of the
438: meteoroid atom is about $2/3$ of its initial velocity for a head on
439: collision. After $n$ such collisions, the velocity is
440: $v_n=(2/3)^nv_0$. Setting $v_n$ equal to twice the thermal velocity at
441: $95\km$, we find that the meteoroid atom is thermalized after $n\sim10$
442: collisions. Thus $r_0\approx\sqrt{n}\,l$, or about $100\cm$. A strict
443: upper limit to the thermalization time is $\tau_{th}=l/v$, about
444: $10\mu\s$, much shorter than the time between radar pulses. Using
445: different arguments, \cite{Manning} and \cite{Bronshten} obtain
446: similar results. After this initial rapid diffusion, ordinary thermal
447: diffusion sets in and the trail radius slowly increases.
448:
449: The initial electron density is
450: %
451: \be %$
452: n_e=q/\pi
453: r_0^2\approx2\times10^3\cm^{-3}\left({\beta\over 0.1}\right)
454: \left({a\over10\mic}\right)^3\left({6\km\over H_p}\right)
455: \left({\rho\over 3 \g\cm^{-3}}\right)\cm^{-3}
456: \ee %$
457: %
458:
459: The plasma frequency $\nu_p \approx 0.39 {\rm\,MHz}$ is
460: % = (4\pi n_ee^2/m_e)^{1/2}/2\pi
461: well below the frequency employed in the radar systems, so the radar
462: beam will penetrate through the trail, ensuring that all the electrons
463: will reflect. The average distance between electrons is
464: $n^{-1/3}\approx 0.1\cm$, much less than the wavelength of either AMOR
465: or Arecibo, so the reflected emission will be coherent in both
466: systems.
467:
468: The thermal diffusion mentioned above eventually causes the trail
469: radius to exceed the wavelength of the radar. This diminishes the
470: visibility of the trail, particularly for trails formed at great
471: heights, or for short wavelength radar. A recent discussion of this
472: 'height ceiling' can be found in Campbell-Brown \& Jones (2003).
473:
474:
475:
476: \subsubsection{Minimum detectable meteoroid size at AMOR\label{sec:Fresnel}}
477: We proceed to calculate the power collected by the radar receiver at
478: AMOR. The relevant properties of the AMOR radar (as
479: well as those for the Arecibo setup) are listed in Table 1. An
480: electron a distance $R$ from the radar sees a flux $P_TG_T/4\pi R^2$,
481: assuming the radar transmits a power $P_T$ and has an antenna gain
482: $G_T$. The gain is defined as $4\pi$ steradians (the solid angle into
483: which an isotropic emitter radiates) divided by the solid angle of the
484: radar beam; the latter is roughly $(\lambda/D)^2$, where $D$ is the
485: radius or length of the antenna. Hence the effective area of the
486: antenna is related to the gain by $A_{eff}=G_T\lambda^2/4\pi$. The
487: electron scatters the incident wave, emitting a power
488: $(3/2)P_TG_T\sigma_T/4\pi R^2$ back toward the radar; the factor $3/2$
489: arises because Thompson scattering is that of a Hertzian dipole rather
490: than isotropic. The receiving antenna (which in the case of Arecibo
491: is the same as the transmitting antenna) captures a fraction
492: $A_{eff}/4\pi R^2$ of this scattered power. The power received at the
493: antenna from a single electron is
494: %
495: \be \label{power} %$
496: \Delta P_e = P_T\left({3G_TG_R\over 128\pi^3 }\right)
497: \left({\sigma_T\lambda^2\over R^4}\right).
498: \ee %$
499: %
500: We have allowed for the possibility that the gain for the receiver and
501: transmitter are different, as is indeed the case at AMOR.
502:
503: The amplitude of the
504: electric field at the receiver is
505: %
506: \be \label{amplitude} %$
507: \Delta A_e=\sqrt{2r\Delta P_e},
508: \ee %$
509: %
510: where $r$ is the impedance of the receiver.
511:
512:
513:
514: The instantaneous signal amplitude received from a meteor trail is found by summing
515: over all the electrons along the track:
516: %
517: \be %$
518: A(t)=\int_{s_1}^{s_2}\left(2r\Delta P_e\right)^{1/2}q(s)\sin(\omega t-2kR)ds,
519: \ee %$
520: %
521: where $\omega$ is the angular frequency of the radar, and
522: $k=2\pi/\lambda$. The total distance from the radar to the trail and
523: back is $2R$, whence the factor $2$ in the argument of the sine. We
524: measure $s$ along the trail starting from the point $X$ on the trail
525: at which $q$ reaches its maximum. The distance between the radar and
526: point $X$ is denoted by $R_0$, while the angle between the line of
527: sight and the meteor trail at $X$ is denoted by $\pi/2+\theta$.
528:
529: Near $X$,
530: %
531: \be %$
532: R\approx R_0\left[1+{s\over R_0}\theta+{1\over2}\left(s\over R_0\right)^2\right],
533: \ee %$
534: %
535: where we assume $\theta<<1$. Define $\chi=\omega t-2kR_0$, and
536: $z=2s/\sqrt{R_0\lambda}$.
537: Then
538: %
539: \be %$
540: A(t)=\sqrt{2r\Delta P_e}q{\sqrt{R_0\lambda}\over 2}
541: \int_{z_1}^{z_2}\sin\left[\chi-2\pi\sqrt{R_0\over\lambda}\theta z-
542: {\pi z^2\over2}\right]dz
543: \ee %$
544: %
545: When $\theta<<\sqrt{\lambda/R_0}$ the integral reduces to a Fresnel
546: integral, which is an oscillating quantity of order unity. The peak
547: power at AMOR is then
548: %
549: \be \label{AMOR power final}%$
550: P_R\approx P_T {3G_TG_R\over256\pi^3}
551: \left({\lambda\over R_0}\right)^3q^2\sigma_T.
552: \ee %$
553: %
554: When $\theta\ga\sqrt{\lambda/R_0}$, the integral cuts off at $z\la 1/4$.
555: In other words, the AMOR
556: detector is sensitive only to micrometeor trails within an angle
557: $\theta\approx\sqrt{\lambda/ R_0}$ of the perpendicular to the
558: line of sight.
559:
560:
561:
562:
563: The minimum detectable line density for AMOR is
564: %
565: \be \label{q min}%$
566: q_{min}={1\over \sqrt{\sigma_T}}\left(P_R\over P_T\right)^{1/2}
567: \left(256\pi^3\over 3G_TG_R\right)^{1/2}\left({R_0\over
568: \lambda}\right)^{3/2}.
569: \ee %$
570: %
571: The minimum detectable power at AMOR is $P_n=1.6\times 10^{-13}$
572: Watts, the transmitted power is $P_T=100$kW, while the gains are
573: $G_T=430$ and $G_R=130$. The typical range is $R_0\approx200\km$,
574: leading to
575: %
576: \be %$
577: q_{min}\approx7\times10^8\cm^{-1}.
578: \ee %$
579: %
580: Using equation (\ref{q approx}), the minimum detectable mass is
581: %
582: \be %$
583: m_{min}\approx2\times10^{-7}\g,
584: \ee %$
585: %
586: corresponding to a minimum radius of
587: %
588: \be %$
589: a_{min}\approx25\left({3\g\cm^{-3}\over\rho}\right)^{1/3}
590: \left({0.1\over\beta}\right)^{1/3}
591: \left({R_0\over200\km}\right)^{2/3} \mu m.
592: \ee %$
593: %
594:
595: We are now in a position to estimate the collecting area $A_{coll}$ of
596: the AMOR detector (as a meteor detector, not as a radar). The
597: collecting area is the product of the sensitivity weighted geometric
598: area $A_G$ imaged by the radar and the fraction of solid angle
599: $\sqrt{\lambda/R_0}$ to which it is sensitive;
600: %
601: \be %$
602: A_{coll}\approx \sqrt{\lambda\over R_0}A_G.
603: \ee %$
604: %
605: We define the sensitivity weighted geometric area as
606: %
607: \be %$
608: A_G\equiv\int {F(m)\over F(m(R_{min}))}R\,dR\,d\phi .
609: \ee %$
610: %
611: Here $F(m)$ is the flux of particles of the minimum detectable mass
612: $m=m_{min}(R)$; the latter is a function of $R$, the distance between
613: the radar and the meteor trail. We assume that $m_{min}\propto
614: q_{min}\propto R^{3/2}$, where the last scaling comes from equation
615: (\ref{q min}), and that $F(m)\sim m^{-1}$ as an analytically simpler
616: version of (\ref{eq:observed flux}). With these assumptions,
617: %
618: \be %$
619: A_G\approx 2R_{min}^2\Delta\phi{\left[\sqrt{R_{max}/R_{min}}-1\right]}
620: \ee %$
621: %
622: The AMOR beam has a width of $\Delta \phi \approx3^\circ$, and
623: extends from $12^\circ$ to $30^\circ$ above the horizon. We further
624: assume that all the meteors are detected at heights of
625: $\sim90\km$. This implies that the lower and upper limits to the
626: integration are $R_{min}=165\km$ and $R_{max}=430\km$. We
627: find $A_G\approx10^{13}\cm$. The collecting
628: area of the radar system is then
629: %
630: \be %$
631: A_{col}\approx 8\km^2.
632: \ee %$
633: %
634:
635: From Fig. 1b in Baggaley (2000), we estimate that in 4 years of
636: continuous operation AMOR detected $\sim10^4$ extrasolar meteors, so
637: we estimate the flux as
638: %
639: \be %$
640: mf_m\left(m=2\times10^{-7}\g\right)\approx 4\times10^{-16}
641: \cm^{-2}\s^{-1}.
642: \ee %$
643: %
644: From the same paper, we estimate that the number of extrasolar meteors
645: detected from Baggaley's ``point source'' to be about about $200$,
646: leading to a flux $50$ times smaller. Both fluxes are plotted in
647: Fig. (\ref{fig:flux}), along with flux estimates from Ulysses and
648: Galileo, and from Arecibo.
649:
650: Fig. (\ref{fig:mass flux}) shows the same data as a differential mass
651: flux. In this plot the bulk of the mass flux is represented by the
652: highest point, making it easy to see that most of the mass of
653: interstellar meteoroids reaching the inner solar system comes in the form
654: of $\sim 0.3\micron$ objects. On the same figure we have plotted the
655: size distribution found by Kim \& Martin (1995), normalized to a
656: hydrogen number density of $0.1$ and assuming that half the metals are
657: in the form of dust grains.
658:
659: \subsubsection{Minimum detectable meteoroid size at Arecibo \label{sec:head}}
660: The minimum detectable meteoroid size for Arecibo can be calculated in a
661: similar manner, starting from equations (\ref{power}) and
662: (\ref{amplitude}). Arecibo uses a much shorter wavelength than AMOR,
663: $70\cm$ rather than $1145\cm$. Arecibo also has a much narrower beam
664: than AMOR, $\sim500\m$ versus $\sim10\km$. It is thus insensitive to trails
665: traveling across the beam, in contrast to AMOR. Due to the fact that
666: trails observed perpendicular to the line of sight have long coherence
667: lengths, AMOR is sensitive to fairly small meteoroids despite the rather
668: modest power and antenna gains employed. Arecibo, by contrast, is
669: sensitive primarily to vertical trails, which have short coherence
670: lengths (a quarter of a wavelength, less than $20\cm$). This is
671: compensated for by the much larger gains, and by the larger
672: transmitted power of the Arecibo radar.
673:
674: The short wavelength employed makes the Arecibo radar sensitive to
675: diffusion of electrons across the meteoroid trail. The Arecibo radar beam
676: is roughly vertical, with a half opening angle of $1/6^\circ$, so at a
677: height of $100\km$ it is about $500$ meters wide. Typical Arecibo
678: meteor trails are a few kilometers long, so they are within a tenth of
679: a radian or so of the vertical. Thus the ion trail left by an Arecibo
680: meteoroid diffuses primarily horizontally, since the density along the
681: trail varies on kilometer scales. As shown above, the initial trail
682: width is of order $100\cm$, but this width increases over time due to
683: diffusion. The diffusion coefficient $D\sim c_sl\sim3\times
684: 10^5\cm^2\s^{-1}$, where $c_s\approx3\times10^4\cm\s^{-1}$ is the
685: sound speed, and $l\approx10\cm$ is the mean free path of air
686: molecules. In the horizontal direction the Fresnel length is
687: $L_F\equiv\sqrt{R_0\lambda}\approx 3\times10^4\cm$, much larger than
688: the initial beam width $r_0\approx100\cm$. The time to diffuse
689: horizontally over a Fresnel zone is $L_F^2/D\approx3000\s$, so
690: diffusion is irrelevant for a vertical trail. However, for a trail
691: $0.1$ radians off the vertical, the time to diffuse $\lambda/4$ in the
692: vertical direction is about 13 ms. This will limit the time the beam
693: remains visible to the radar at a given height to about 13 ms for a
694: typical meteor. This is consistent with the trail lengths observed by
695: Mathews et al. (1997).
696:
697: We have already noted that the meteor trail is much narrower than the
698: Fresnel length of the Arecibo radar; then all the
699: electrons in a patch of length of order $\lambda/2$ along the trail
700: radiate coherently. Suppose that $N_{ch}$ is the number of coherently
701: emitting electrons in such a patch. The
702: amplitude of the electric field at the receiver from a single patch is
703: %
704: \be \label{Arecibo Amplitude}%$
705: A\approx\sqrt{2r\Delta P_e}N_{ch}.
706: \ee %$
707: %
708: The vertical resolution of the Arecibo antenna is $L\approx 150\m$, so
709: that there are $N_p\sim L/\lambda\approx200$
710: patches radiating in a resolution length. The signal observed at
711: Arecibo varies on a length scale much shorter than the atmospheric
712: scale height, indicating that the mass loss rate fluctuates
713: rapidly. We assume that these fluctuations occur on scales smaller
714: than $L$, so that the electric field from each patch adds
715: incoherently.
716: The power received by the detector is
717: %
718: \be \label{Arecibo power} %$
719: P_{\rm Arecibo}=P_T\left({3G_TG_R\over128\pi^3}\right)
720: \sqrt{L\over\lambda}\left({\sigma_T\lambda^2\over R_0^4}\right)N_{ch}^2.
721: \ee %$
722: %
723: Since the coherence length along the (assumed vertical) trail is of
724: order $\lambda$, the number of coherently scattering electrons
725: is simply related to the line density,
726: %
727: \be %$
728: N_{ch}\approx q\cdot\lambda/2.
729: \ee %$
730: %
731: The expression for the power becomes
732: %
733: \be \label{Arecibo power final}%$
734: P=P_T\left(3G_TG_R\over512\pi^3\right)
735: \sqrt{L\over\lambda}\left({\lambda^4\over R_0^4}\right)
736: q^2\sigma_T.
737: \ee %$
738: %
739:
740:
741: For a given transmitted power, wavelength and antenna gain, this is smaller than
742: the equivalent expression (\ref{AMOR power final}) by the ratio
743: %
744: \be %$
745: {\sqrt{\lambda L}\over R_0}\approx10^{-4},
746: \ee %$
747: %
748: using values for Arecibo. Worse, the power scales as $\lambda^3$ (or
749: as $\lambda^{3.5}$ for Arecibo), so the $\sim15$ times shorter
750: wavelength employed by Arecibo reduces the received power by another
751: factor of $\sim3000$. However, these two factors are more than
752: compensated for by the much larger gains of the Arecibo antenna. The
753: net effect is that despite its higher power and much larger gain, the
754: Arecibo radar can detect meteoroids that are a factor of seven to ten
755: smaller in radius than the AMOR setup, but no smaller.
756:
757: Define the signal to noise ratio $SNR\equiv P/P_{n}$, where
758: $P_n$ is the noise power for the Arecibo receiver. Using this definition, and
759: solving for $q$,
760: %
761: \be \label{eqn:Arecibo q}%$
762: q={1\over \sqrt{\sigma_T}}\left({P_n\over P_T}\right)^{1/2}
763: \left({512\pi^3\over 3G_TG_R}\right)^{1/2}
764: \left({\lambda\over L}\right)^{1/4}
765: \left({R_0\over \lambda}\right)^2
766: SNR^{1/2}
767: \ee %$
768: %
769:
770: Using the values in Table 1, we find
771: $q\approx10^7\cm^{-1}$. Using equation (\ref{q approx}) we find
772: %
773: \be %$
774: m\approx3\times10^{-9}\left({0.1\over\beta}\right)
775: \left({R_0\over100\km}\right)^2 \left({150\m\over L}\right)^{1/4}SNR^{1/2}g.
776: \ee %$
777: %
778: The corresponding meteoroid radius is
779: %
780: \be \label{radar size}%$
781: a \approx6\left({3\g\cm^{-3}\over\rho}\right)^{1/3}
782: \left({0.1\over\beta}\right)^{1/3}
783: \left({R_0\over100\km}\right)^{2/3}
784: \left({150\m\over L}\right)^{1/12}
785: SNR^{1/6}\mu m.
786: \ee %$
787: %
788:
789: A firm lower limit for the size of the meteoroids seen at Arecibo is
790: found by noting that the energy per meteoroid atom is $\sim100$ eV. Since
791: the ionization potential of either atoms or molecules is typically of
792: order $10$ eV, a single meteoroid atom can free at most $10$
793: electrons. Taking $\rho=3\g/\cm^3$ and $\beta=10$, we find
794: $a_{min}\approx1.3\micron$.
795:
796: From equation (\ref{Arecibo power final}), and noting that
797: $G_R\lambda^2=const.$, we see that, as long as $\lambda$ is larger
798: than the initial radius $r_0$ of the trail, $P\propto
799: \lambda^1.5$. Some Arecibo meteors are seen at both 430 MHz and 50 MHz
800: (Meisel, private communication). This suggests that the volume
801: occupied by the reflecting electrons has linear dimensions smaller
802: than $\lambda/2\approx35\cm$.
803:
804: For example, it might be thought that the meteoroid itself captures
805: enough electrons to scatter the radar beam. However, this is not
806: possible. The number of charges required to reflect the beam is given
807: by equation (\ref{Arecibo power}), except that in this case there is
808: only one emitting region (the meteor) so the factor $\sqrt{L/\lambda}$
809: should be dropped. We find $N_{ch}\approx10^9$; the meteoroid is charged
810: to a (negative) voltage of $V=4.8\times10^5(3\micron/a)$
811: volts. Charging the meteoroid to such a high voltage leads to rapid
812: electron emission by quantum tunneling, also known as field emission
813: (Fowler \& Nordheim 1928). The discharge current (in statamps per
814: square centimeter) is given by
815: %
816: \be %$
817: I={e\over2\pi h}{(\mu/\phi)^{1/2}\over\phi+\mu}E^2e^{-4\pi \phi^3/2/3E},
818: \ee %$
819: %
820: where $E\approx V/a$ is the electric field
821: near the surface of the meteor, $\phi$ is the work function, and $\mu$
822: is the Fermi energy. Both of the latter have values of order $5$ eV.
823:
824: We make the optimistic assumption that the meteoroid gains an electron
825: for every atom it collides with in the atmosphere, $I_{in}=e n^*_a v$,
826: also in statamps per square centimeter ($n^*_a$ is the number density
827: of air molecules at the height where the meteoroid ablates, see equation
828: \ref{eqn: density}). Setting these two currents equal, we find the
829: equilibrium field $E\approx3.6\times10^7\ {\rm V}\cm^{-1}$ or a
830: maximum voltage of $V=10^4(a/3\micron)$. The number of charges on the meteor
831: cannot exceed $N_{ch, max}\approx2\times10^7(a/3\micron)^2$, a factor
832: of $30$ too small to explain the Arecibo observations. The meteor head
833: signal must be due to electrons in the atmosphere around or trailing
834: the meteor.
835:
836: One may also estimate $a$ dynamically, if one can measure the rate of
837: deceleration of the meteoroid due to atmospheric drag;
838: %
839: \be %$
840: m{dv\over dt}=-\Gamma\rho_av^2A_m,
841: \ee %$
842: %
843: where $\Gamma$ is the drag coefficient, and $A_m$ is the area of the
844: meteor.
845: The radar measures $dv/dt$, $v$, and the height of the meteor. Using a
846: standard atmospheric model, the density $\rho_a$ can be determined,
847: For micrometeoroids, which have a size less than the mean free path for
848: collisions, $\Gamma=1$, so the ratio $m/A_m$ can be determined
849: directly from measured quantities. \cite{JMMZ} refer to this ratio as
850: the ballistic parameter, $BP$. We assume spherical meteoroids, so
851: $BP\approx 4\rho a/3$. Then a simple estimate for $a$ is
852: %
853: \be %$
854: a={3\over4}{{\rm BP}\over \rho}
855: \ee %$
856: %
857: \cite{JMMZ} find a range of ballistic parameters, ranging down to
858: $10^{-3}\g\cm^{-2}$ or slightly lower, leading to
859: %
860: \be \label{BP size}%$
861: a_{min}\approx2.5\left({3\g\cm^{-3}\over
862: \rho}\right)\left({BP\over10^{-3}\g\cm^{-2}}\right)\mu m.
863: \ee %$
864: %
865: This is a factor 2 smaller than the radius estimate based on received
866: radar power, or a factor of 8 smaller when considering the meteor
867: mass. For the meteors with the smallest BP, the size drops to $1\mu m$
868: or less. It is very difficult to see how the power reflected from such
869: a low mass meteoroid could be detected at Arecibo, suggesting that the
870: micrometeoroids have a density closer to $1\g\cm^{-3}$ than to
871: $3\g\cm^{-3}$. In fact, if we set $\rho=1$, we find $a\sim8.5\mu m$ and
872: $a\sim 7.5\mu m$ from the power and BP estimates, respectively.
873:
874: The fluxes observed at Arecibo are given by Meisel et al. (2002); they
875: are also shown on Figure 1.
876:
877: \section{\label{sec:survive} PROPAGATION OF LARGE GRAINS THROUGH THE ISM}
878:
879: We have seen that AMOR can only detect grains with masses
880: $\gtsim 2 \times 10^{-7} \g$, corresponding to a grain radius
881: $a \approx 25 \micron$ for silicate dust (mass density $\rho \approx
882: 3.5 \g \cm^{-3}$). Similarly, Arecibo can detect particles with
883: $a\gtsim 6\micron$. Thus, we will restrict our attention to ``large''
884: (by interstellar dust standards) grains, with $a \gtsim 10 \micron$.
885:
886: For such large grains in the interstellar medium (ISM), radiation pressure
887: can be ignored. The force due to radiation pressure is
888: $F_{\rm rad} = \pi a^2 \langle Q_{\rm pr} \rangle \Delta u_{\rm rad}$,
889: where $\langle Q_{\rm pr} \rangle$ is the radiation pressure efficiency
890: factor averaged over the interstellar radiation field (ISRF) and
891: $\Delta u_{\rm rad}$ is the energy density in the
892: anisotropic component of the ISRF.
893: Adopting the ISRF for the solar neighborhood (Mezger, Mathis, \&
894: Panagia 1982; Mathis, Mezger, \& Panagia 1983) and a 10\% anisotropy
895: (Weingartner \& Draine 2001b),
896: $\langle Q_{\rm pr} \rangle \approx 1$ for $a \gtsim 10 \micron$ and
897: $\Delta u_{\rm rad} \approx 8.64 \times 10^{-14} \E$. The time interval
898: required for radiation pressure to change a silicate
899: grain's velocity by
900: $1 \kms$ (assuming the anisotropy direction remains the same as the grain
901: moves through space) is thus given by $\Delta t \approx 5 \times 10^8 \yr
902: (a/10 \micron)$. For large grains in the diffuse ISM, the forces resulting
903: from the asymmetric photon-stimulated ejection of electrons and adsorbed
904: atoms are even less effective than the radiation pressure (Weingartner \&
905: Draine 2001b), so these forces can be neglected as well.
906:
907: The following three influences can be significant for grains with
908: $a \gtsim 10 \micron$: 1. the drag force, 2. grain destruction following
909: impacts with interstellar grains or high-energy gas atoms, and 3. the
910: magnetic force.
911:
912: We will consider the propagation of grains through two idealized phases of
913: the ISM, the cold and warm neutral media (CNM and WNM, respectively). We
914: also consider the Local Bubble (LB), i.e., the large volume of low-density,
915: ionized gas that surrounds the Sun (see, e.g., Sfeir et al.~1999). Our
916: adopted values for the gas temperature $T_{\rm gas}$, H number density $\nH$,
917: and electron fraction $x_e \equiv n_e/\nH$ ($n_e$ is the electron number
918: density) in these three environments are
919: given in Table \ref{tab:phases}. In each case we adopt the ISRF for the
920: solar neighborhood.
921:
922: The drag and magnetic forces both depend on the grain's electric charge,
923: which is set by a balance between the accretion of electrons from the gas
924: versus photoelectric emission and proton accretion. For large enough grains,
925: the rates of each of these processes are proportional to the grain area;
926: thus, the electric potential $\Phi$ of large grains is independent of the
927: size $a$. Using the charging algorithm from Weingartner \& Draine (2001c),
928: we find that silicate grains with $a \gtsim 10 \micron$ charge to
929: (0.15 V, 0.74 V, 0.94 V) in the (CNM, WNM, LB).\footnote{Here, and throughout
930: this paper, we take silicate optical properties from Li \& Draine 2001.}
931:
932: \subsection{The Drag Force}
933:
934: When a grain's motion is highly supersonic, the hydrodynamic drag (i.e., the
935: drag due to direct impacts of gas atoms and ions) does not depend on the gas
936: temperature. In this case, the grain's speed decreases by a factor of $e$
937: once it has encountered its own mass in gas. The sound speed is
938: $0.81 \kms$ ($6.2 \kms$) in the CNM (WNM), assuming that the He number
939: density $n_{\rm He} = 0.1 \nH$.
940: We expect grain speeds as low as $\sim 10 \kms$,
941: so deviations from the highly supersonic limit can be expected in the
942: WNM. In addition to the hydrodynamic drag, there is also the Coulomb drag,
943: due to long-range electric interactions between the charged grains and
944: gas-phase ions. The Coulomb drag contributes significantly to the total
945: drag in the WNM for drift speeds $\ltsim 10 \kms$.
946: Using the approximate drag expressions from Draine \& Salpeter (1979),
947: we find that the actual drag force (hydrodynamic plus Coulomb)
948: in the WNM exceeds the hydrodynamic drag force calculated in the
949: highly supersonic limit by only a factor of $\approx 2$ at a drift speed
950: of $10 \kms$. For regions outside of the LB, we will adopt the highly
951: supersonic limit and assume that drag
952: limits the distance a grain can travel to its velocity $e$-folding distance
953: $D_{\rm drag}$, given by
954: \be
955: D_{\rm drag} = 650 \pc \left( \frac{\rho}{3.5 \g \cm^{-3}} \right)
956: \left( \frac{a}{10 \micron} \right) \left( \frac{\nH}{1 \cm^{-3}}
957: \right)^{-1}~~~.
958: \ee
959: Near the Sun, the average H density near the Galactic midplane is
960: $\nH \approx 1 \cm^{-3}$ (Whittet 1992).
961:
962: In the LB, the sound speed is $\approx 100 \kms$; thus, the highly subsonic
963: limit may apply. In this case, the drag force is proportional to the grain
964: speed. Using the Draine \& Salpeter (1979) drag expressions, we find that
965: the Coulomb drag is negligible and that
966: \be
967: D_{\rm drag} = 7.8 \, {\rm kpc} \left( \frac{\rho}{3.5 \g \cm^{-3}} \right)
968: \left( \frac{a}{10 \micron} \right) \left( \frac{\nH}{1 \cm^{-3}}
969: \right)^{-1} \left( \frac{T_{\rm gas}}{10^6 \K} \right)^{-1/2}
970: \left( \frac{v_0}{10 \kms} \right)~~~,
971: \ee
972: where $v_0$ is the grain's initial speed. Since the maximum extent of the
973: LB is $\approx 250 \pc$ (Sfeir et al.~1999) and $\nH \approx 5 \times 10^{-3}
974: \cm^{-3}$, drag in the LB is always insignificant, whether the subsonic or
975: supersonic limit applies.
976:
977: \subsection{\label{sec:magnetic_deflection} The Magnetic Force}
978:
979: Charged grains spiral around magnetic field lines with a gyroradius
980: \be
981: \label{eq:gyroradius}
982: r_B = 17 \pc \left( \frac{\rho}{3.5 \g \cm^{-3}}\right)
983: \left( \frac{\Phi}{0.5 \, {\rm V}}\right)^{-1}
984: \left( \frac{B}{5 \mu {\rm G}}\right)^{-1}
985: \left( \frac{v}{10 \kms}\right)
986: \left( \frac{a}{10 \micron}\right)^2~~~.
987: \ee
988: Although the magnetic field strength $B$ has not been measured in the Local
989: Bubble, $B \sim 5 \mu G$ just outside the LB, with the random component
990: dominating the ordered component (Heiles 1998). Thus, the magnetic force
991: can significantly deflect grains with $a \approx 10 \micron$. Since
992: $r_B \propto a^2$, the importance of the magnetic deflection rapidly
993: decreases with grain size.
994:
995: \subsection{Grain Destruction}
996:
997: When a grain travels through the ISM, it is subjected to various agents
998: of destruction, including: 1. Sputtering, in which gas-phase ions strike
999: the grain and remove grain surface atoms, which then enter the gas. It is
1000: useful to distinguish between thermal sputtering (due to thermal motion
1001: of the gas atoms) and non-thermal sputtering (due to the motion of the
1002: grain with respect to the gas). 2. Shattering and vaporization in
1003: grain-grain collisions. 3. For ices, sublimation and photodesorption.
1004:
1005: Thermal and non-thermal sputtering have been extensively discussed by
1006: Tielens et al.~(1994). Applying their analysis, we find that sputtering
1007: of grains with $a \gtsim 10 \micron$ can be neglected in all environments
1008: of interest.
1009:
1010: The physics of grain-grain collisions has been treated extensively by
1011: Tielens et al.~(1994) and Jones et al.~(1996). When the relative
1012: speed exceeds $\approx 3 \kms$ (see Table 1 in Jones et al.~1996), a crater
1013: forms on the larger grain (the target). A portion of the evacuated mass
1014: remains on the grain as a crater lip. The rest of the crater mass is
1015: removed, both as vapor and shattering fragments. The shattering fragments
1016: dominate the vapor, and the largest fragment is typically somewhat larger
1017: than the smaller impacting grain (the projectile). When the relative
1018: speed exceeds $\sim 100 \kms$ (depending on the target and projectile
1019: materials, see Table 1 in Jones et al.~1996), the larger grain is
1020: entirely disrupted, and the largest shattering fragment can be a substantial
1021: fraction of the target size. For our purposes, the cratering regime is
1022: generally applicable.
1023:
1024: Jones et al.~(1996) give simple approximations for the (material-dependent)
1025: ratio of the crater mass ($M_c$) to the projectile mass ($M_{proj}$).
1026: Interstellar dust is thought to be dominated by two populations:
1027: silicates and carbonaceous (e.g., graphite) grains. For a head-on
1028: collision between a silicate target
1029: and either a silicate or graphite projectile, the Jones et al.~(1996)
1030: results can be approximated by
1031: \be
1032: \label{eq:crater_mass}
1033: \frac{M_c}{M_{proj}} \approx 18 \left(\frac{v}{10 \kms}\right) + 29 \left(
1034: \frac{v}{10 \kms} \right)^2~~~, \ee where $v$ is the grain-grain
1035: relative speed.
1036: This empirical approximation reproduces detailed calculations, using equation
1037: (1) and Table 1 from Jones et al. (1996), to within 10\% when $5 \le v \le
1038: 100 \km \s^{-1}$.
1039: %%The above approximation is accurate to within
1040: %%$\approx 10\%$ for $5 \le v \le 100 \kms$.
1041: The crater mass is
1042: approximately 4.2 (2.2) times bigger for a graphite (ice) target. We
1043: will assume that equation (\ref{eq:crater_mass}) applies to all
1044: collisions and that the entire crater mass is ejected. This is a
1045: conservative assumption, since the crater mass should actually be
1046: smaller for oblique collisions, and a portion of the crater mass will
1047: form a lip. On the other hand, we underestimate destruction if target
1048: grains are made of less resilient material than silicate or if the
1049: grains are fluffy.
1050:
1051: Since the mass in dust is $\approx 0.011$ times the mass in H (in all
1052: gas-phase forms) in the ISM (see, e.g., Weingartner \& Draine 2001a), the
1053: distance a grain can travel before it loses half its mass is given by
1054: \be
1055: \label{eq:D_dust}
1056: D_{\rm dest} \approx 40.2 \, {\rm kpc} \left( \frac{\rho}{3.5 \g \cm^{-3}}
1057: \right) \left( \frac{a}{10 \micron} \right) \left( \frac{\nH}{1 \cm^{-3}}
1058: \right)^{-1} \left[ 18 \left(\frac{v}{10 \kms} \right) + 29
1059: \left(\frac{v}{10 \kms} \right)^2 \right]^{-1}~~~,
1060: \ee
1061: yielding $D_{\rm dest} \approx 860 \pc$ for $v = 10 \kms$ (and with other
1062: canonical parameter values as in eq.~\ref{eq:D_dust}).
1063:
1064: \subsection{The Survival Probability}
1065:
1066: In Figure \ref{fig:f_survive}, we plot $D_{\rm drag}$, $r_B$, and
1067: $D_{\rm dest}$ versus $v$ for four grain sizes, covering the range of
1068: interest. Due to the expected dramatic decrease in flux with $a$
1069: (e.g., see eq.~\ref{eq:flux_at_earth} below), we are not interested in
1070: $a \gtsim 100 \micron$. We adopt $\rho = 3.5 \g \cm^{-3}$,
1071: $\nH = 1 \cm^{-3}$, $\Phi=0.5 \, {\rm V}$, and $B = 5 \, \mu G$, and assume
1072: supersonic grain speeds, for these plots. For all of
1073: the considered grain sizes, magnetic deflection dominates the other two
1074: processes when $v = 10 \kms$, while destruction dominates when
1075: $v > \,$a few to several$\, \times 10 \kms$. The drag force is never dominant.
1076: The detailed analysis of the dynamics of a charged
1077: grain in a region with both ordered and random magnetic field components
1078: is beyond the scope of this paper. Thus, we will simply assume that
1079: \be
1080: f_{\rm survive} = \cases{1 &, $d < \min(r_B, D_{\rm dest})$\cr
1081: 0 &, $d > \min(r_B, D_{\rm dest})$\cr}~~~.
1082: \ee
1083: The grain velocity used in evaluating $r_B$ and $D_{\rm dest}$ should be
1084: taken with respect to the ambient ISM. However, for simplicity, we will
1085: use the velocity with respect to the Sun.
1086:
1087: For Local Bubble conditions, magnetic deflection dominates destruction
1088: for grains with $10 < a < 100 \micron$ when $v < 100 \kms$.
1089: For example, for $a=10 \micron$, $D_{\rm dest} = r_B$ when
1090: $v=311 \kms$ and $r_B = 281 \pc$ (which exceeds the maximum extent of
1091: the LB). For $a=100 \micron$, $D_{\rm dest} = r_B$ when
1092: $v=143 \kms$ and $r_B = 13 \, {\rm kpc}$.
1093:
1094:
1095: \section{\label{sec:Vega-like} DUST FROM YOUNG MAIN SEQUENCE STARS}
1096:
1097: IRAS observations of several young main
1098: sequence stars revealed emission at 60 and $100 \micron$ substantially in
1099: excess of the emission from the star's photosphere (e.g., Aumann et
1100: al.~1984, Gillett 1986). This ``Vega phenomenon'' (named for the first
1101: example to be observed) was
1102: attributed to circumstellar dust, which absorbs the star's optical radiation
1103: and re-emits in the infrared. Shortly following this discovery, Smith \&
1104: Terrile (1984) observed the optical light scattered by the dust around
1105: $\beta$ Pictoris and found that the grains lie in an edge-on disk. More
1106: recently, disks have been imaged (in the infrared and sub-mm)
1107: around a handful of other stars (e.g., Holland et al.~1998, Schneider
1108: et al.~1999). For reviews of the Vega phenomenon and circumstellar
1109: dust disks, see Backman \& Paresce (1993), Lagrange, Backman, \&
1110: Artymowicz (2000), and Zuckerman (2001).
1111: It is not yet clear what fraction of Vega-like
1112: stars posses dust disks, but if these stars also posses planets, then
1113: dynamical interactions can eject the grains from the system.
1114: (Radiation pressure also removes small grains, but these are too small to
1115: be traced back to their source.)
1116:
1117:
1118: The most massive disks, which orbit the youngest stars, have optical
1119: depths in the near IR of order $10^{-2}$ to $10^{-3}$. The bulk of
1120: this optical depth is contributed by grains with size
1121: $a_0\sim\lambda/2\pi$, where $\lambda$ is the wavelength, typically
1122: $1-10\micron$ (see below).
1123:
1124: \subsection{Gravitational ejection of small particles}
1125: Jupiter mass planets are seen around $5-10\%$ of nearby solar type
1126: stars in radial velocity surveys. These surveys are not sensitive to
1127: planets in orbits larger than $\sim5$ AU, such as Jupiter, so the
1128: fraction of solar type stars with such planets is likely to be
1129: substantially higher. Observations of debris disks also hint at the
1130: presence of planets (Scholl, Roques \& Sicardy 1993; Wilner et
1131: al. 2002). This suggests that gravitational interactions between a
1132: massive planet and dust particles in debris disks are a natural means for
1133: producing interstellar meteoroids.
1134:
1135: We estimate the ejection velocity of large ($25\micron$) dust grains
1136: interacting with a Jupiter mass planet. We neglect collisions with
1137: other dust grains, an assumption which we justify at the end of the
1138: calculation, as well as radiation pressure. We follow the derivation
1139: of \"Opik (1976); a good general introduction to the two body problem
1140: can be found in Murray and Dermott (1999).
1141:
1142: We assume that the planet, of mass $M_p$, orbits the star, of mass
1143: $M_*$, on a circular orbit of semimajor axis $\ab_p$. A test mass (the
1144: dust particle) also orbits the star. The test mass is subject to the
1145: gravity of both the planet and the star, but the planet is assumed to
1146: be immune to the gravity of the test particle. When the test particle
1147: is far from the planet it follows a roughly Keplerian orbit around the
1148: star. When the test particle is very close to the planet, it also
1149: follows a roughly Keplerian orbit, but this time around the
1150: planet. While it is close to the planet, the test particle experiences
1151: a tidal force from the star, with a magnitude given by
1152: %
1153: \be %$
1154: F_*\approx {GM_*m_g\over \ab_p^2} {r\over \ab_p},
1155: \ee %$
1156: %
1157: where $r$ is the distance between the test particle and the planet,
1158: $G$ is the gravitational constant, $M_*$ is the mass of the central
1159: star, $\ab_p$ is the semimajor axis of the planet's orbit, and $m_g$ is
1160: the mass of the dust grain. The Hill radius is the distance $r$ at
1161: which this tidal force equals the gravitational force of the planet
1162: on the test particle, $GM_pm_g/r^2$, or
1163: %
1164: \be \label{eqn: Hill radius}%$
1165: r_H\equiv\left(M_p\over 3M_*\right)^{1/3}\ab_p,
1166: \ee %$
1167: %
1168: where the factor of $3$ is included for historical reasons.
1169:
1170: We make use of the \"Opik approximation; while the test body is inside
1171: the Hill sphere of the planet, we assume that the motion is described
1172: by the two body problem in the frame revolving with the planet. This
1173: involves ignoring both the tidal force of the star and the Coriolis
1174: force associated with the motion of the planet around the star. The
1175: acceleration due to the Coriolis force is
1176: %
1177: \be %$
1178: a_{Cor}\approx \Omega_p V,
1179: \ee %$
1180: %
1181: where $\Omega_p$ is the mean motion of the planet (the angular speed
1182: of the planet's revolution about the star) and $V$ is the
1183: velocity of the test particle in the rotating frame. A rough estimate
1184: for $V$ is $v_p$, the Keplerian velocity of the planet, so the
1185: Coriolis force is larger than the tidal force by a factor $\ab_p/r_H\sim
1186: \mu_p^{-1/3}$ at the Hill radius; in this section $\mu_p\equiv M_p/M_*$.
1187:
1188: \subsubsection{Inside the Hill sphere}
1189: We assume that the test particle orbits the star rather than the
1190: planet. Hence the particle follows a hyperbolic orbit relative to the
1191: planet during the close encounter. The hyperbola is specified by its
1192: eccentricity $\bar e$ and the transverse semimajor axis $\bar {\ab}$
1193: (note that the semimajor axis of the dust particle $\bar \ab$ should not be
1194: confused with the radius $a$ of the dust particle). Barred elements
1195: are calculated relative to the planet. Note that $\bar e>1$ and $\bar
1196: \ab<0$. Alternately we can specify the specific energy $\bar
1197: E=-GM_p/2\bar \ab$ and specific angular momentum $\bar L=\sqrt{GM_p\bar
1198: \ab(1-\bar e^2)}$ relative to the planet. The energy is positive, so the
1199: particle would have a finite velocity as it traveled to infinity, if
1200: we ignore the effects of the star. This velocity is known as the
1201: velocity at infinity; we denote it by $U$.
1202:
1203: The approximations described above allow us to find an analytic relation
1204: between the angle of deflection resulting from the encounter and the
1205: eccentricity of the (hyperbolic) orbit of the test body around the
1206: planet. We then relate the eccentricity to the periapse distance
1207: $\bar q$ between the planet and the test particle, and
1208: the velocity at infinity.
1209:
1210: The specific energy and angular momentum of a test mass interacting
1211: with a planet of mass $M_p$ are given by
1212: %
1213: \be \label{eqn: two body E} %$
1214: \bar E={1\over 2 }V^2-{GM_p \over \bar r}
1215: \ee %$
1216: %
1217: and
1218: %
1219: \be \label{eqn: two body L} %$
1220: \bar L=\bar r^2{d\theta\over dt}.
1221: \ee %$
1222: %
1223: In these expressions $G$ is the gravitational constant, ${\bf V}$ is
1224: the velocity of the test mass relative to the planet, and $\bar r$ is
1225: the distance between the two bodies. The angle $\theta$ is measured
1226: from a fixed line (which we choose to be the apsidal line, which
1227: connects the star and the planet at the point of closest approach) and
1228: the line joining the two bodies.
1229:
1230:
1231: The distance $\bar r$ is given by
1232: %
1233: \be \label{eqn: radius} %$
1234: \bar r={\bar p \over {1+\bar e\cos\theta}}.
1235: \ee %$
1236: %
1237:
1238: The quantity $\bar p\equiv \bar L^2/GM_p$ is called the semilatus
1239: rectum. The periapse distance (at closest approach) is denoted by $\bar
1240: q$ and is given by $\bar q=\bar p/(1+\bar e)$.
1241:
1242: The semilatus rectum $\bar p$ is related to $\bar e$ and $\bar \ab$ by
1243: %
1244: \be \label{eqn: semilatus}%$
1245: \bar p=\bar \ab(1-\bar e^2)
1246: \ee %$
1247: %
1248: Note that $\bar p>0$, and that $\bar q=\ab(1-e)>0$. Using the definition
1249: of $\bar \ab$, we can write the specific energy in terms of $\bar p$ and
1250: $\bar e$:
1251: %
1252: \be %$
1253: \bar E={GM_p\over 2\bar p}(\bar e^2-1).
1254: \ee %$
1255: %
1256: From eqn. (\ref{eqn: two body E})
1257: %
1258: \be %$
1259: \bar E={1\over2}\bar v_q^2-{GM_p\over \bar q}={1\over2}U^2,
1260: \ee %$
1261: %
1262: where $\bar v_q$ is the speed at closest approach, and $U$ is the
1263: speed at infinity.
1264:
1265: From equation (\ref{eqn: radius}), as $r\to\infty$, the angle $\theta$ tends to
1266: %
1267: \be %$
1268: \theta_0=-\arccos{1\over \bar e}
1269: \ee %$
1270: %
1271: Referring to Figure (\ref{fig:scatter}), and denoting the angle of
1272: deflection by $\gamma$, simple geometry gives
1273: %
1274: \be %$
1275: \sin{\gamma\over 2}=\sin(\theta_0-{\pi\over2})=-\cos\theta_0={1\over
1276: \bar e}.
1277: \ee %$
1278: %
1279: Since $\bar e=1-(\bar q/\bar \ab)$, we find
1280: %
1281: \be \label{eqn: gamma}%$
1282: \sin{\gamma\over 2}=\left[1+{U^2\bar q\over GM_p}\right]^{-1}
1283: \ee %$
1284: %
1285:
1286: It is convenient to write this in terms of the impact parameter
1287: $s$. From conservation of angular momentum, $sU=\bar q \bar
1288: v_q$. We find
1289: %
1290: \be %$
1291: \bar q=\sqrt{\left(GM_p\over U^2\right)^2+s^2}-{GM_p\over U^2}
1292: \ee %$
1293: %
1294: Using this in eqn. (\ref{eqn: gamma}) we find
1295: %
1296: \be \label{eqn: scattering angle}%$
1297: \sin{\gamma\over 2}=\left[1+{U^4 s^2\over G^2M_p^2}\right]^{-1/2}
1298: \ee %$
1299: %
1300:
1301:
1302: \subsubsection{Relating $U$ to the orbital elements $\ab$, and $e$ of the
1303: test particle}
1304:
1305: We will assume that the impact parameter $s$ is uniformly distributed
1306: between $0$ and $r_H$. However, we still need to know $U$, the
1307: speed at infinity, in the frame rotating with the planet. We can
1308: find $U$ in terms of the semimajor axis $\ab$ and eccentricity $e$ of
1309: the test particle, where the orbital elements are now calculated
1310: relative to the star. The relations between $E$, $L$, $\ab$, $e$, and
1311: $p$ are as given above, but with the star playing the role of the
1312: central mass, so that $M_p$ is replaced by $M_*$, the mass of the
1313: star. Since the test particle is bound to the star $E<0$, $\ab>0$, and
1314: $e<1$, at least until the scattering event that ejects the test
1315: particle. We will restrict our attention to the case where the planet
1316: and the test particle orbit in the same plane.
1317:
1318: We start by dividing the particle velocity (relative to an inertial
1319: frame centered on the center of mass) into a radial $v_r$ and a
1320: transverse $v_t$ part. From the definition of angular momentum, the
1321: transverse velocity at the time of the close encounter is
1322: %
1323: \be \label{eqn: v t}%$
1324: v_t=L/r\approx v_p\sqrt{{\ab\over \ab_p}(1-e^2)},
1325: \ee %$
1326: %
1327: where $v_p=\sqrt{GM_*/\ab_p}$ is the Keplerian velocity of the planet,
1328: and we have made use of the fact that the star-particle distance
1329: $r\approx \ab_p$ (it can differ by $r_H$) during the encounter.
1330:
1331: The magnitude of the total velocity $v$ is found from the expression for
1332: the energy,
1333: %
1334: \be \label{eqn: E}%$
1335: E={1\over2}v^2-{GM_*\over \ab_p}=-{GM_*\over 2\ab},
1336: \ee %$
1337: %
1338: where we have again set $r=\ab_p$ in the first equality. Solving
1339: equations (\ref{eqn: v t}) and (\ref{eqn: E}) for
1340: $v_r$,
1341: %
1342: \be %$
1343: v_r\approx v_p\left[2-{\ab_p\over \ab}-{\ab\over \ab_p}(1-e^2)\right]^{1/2}.
1344: \ee %$
1345: %
1346: Note that if the orbits just cross, $\ab(1-e)/\ab_p=1$, $1+e=2-\ab_p/\ab$, and
1347: $v_r=0$; the collision occurs at periapse. Then
1348: $U^2=(v_t-v_p)^2\approx v_p^2$. However, the orbit of the test
1349: particle need only pass through the Hill sphere, so that
1350: $\ab(1-e)/\ab_p\approx 1\pm r_H/\ab_p\approx 1\pm \mu_p^{1/3}$, and the radial
1351: velocity $v_r\approx\sqrt{2}\mu_p^{1/6}v_p$, which for a Jupiter mass
1352: planet is about $0.45 v_p$. If there are multiple planets in the
1353: system, the test particle periapse need not be comparable to $\ab_p$, in
1354: which case $v_r$ could be slightly larger than $v_p$. The transverse
1355: velocity $v_t$ is always of order $v_p$.
1356:
1357: To find $U$, we transform to the planet frame, which entails
1358: subtracting $v_p$ from $v_t$:
1359: \begin{eqnarray}
1360: U_r = v_r & \approx &\sqrt{2}\mu_p^{1/6}v_p\\
1361: U_t = v_t-v_p & \approx &v_p\left[\sqrt{{\ab\over \ab_p}(1-e^2)}-1\right]\\
1362: U^2 & \approx & v_p^2\left[2\mu_p^{1/3} + (3-2\sqrt{2})\right].
1363: \label{eq:U}
1364: \end{eqnarray}
1365:
1366: \subsubsection{The change in energy}
1367: In the frame rotating with the planet, the close encounter simply
1368: rotates the test particle velocity by the angle $\gamma$, without
1369: changing its magnitude $U$. The components of the scattered velocity
1370: are
1371: \begin{eqnarray}
1372: U_r' & = & U_r\cos\gamma - U_t\sin\gamma\\
1373: U_t' & = & U_r\sin\gamma + U_t\cos\gamma.
1374: \end{eqnarray}
1375:
1376: Transforming back to the inertial frame, the square of the new
1377: velocity is given by
1378: %
1379: \be %$
1380: v'^2=U_r'^2+(U_t'+v_p)^2=U_r^2+U_t^2+2v_p(U_r\sin\gamma+U_t\cos\gamma)+v_p^2.
1381: \ee %$
1382: %
1383: The change in specific energy
1384: %
1385: \be %$
1386: \Delta E \equiv {(v'^2-v^2)\over2} \approx v_p
1387: \left[U_r\sin\gamma + U_t(1-\cos\gamma)\right]
1388: \ee %$
1389: %
1390: The radius $r$ does not change
1391: appreciably during the encounter.
1392: Using the relations between $U$ and $v$,
1393: %
1394: \be %$
1395: \Delta E = v_p\left[v_r\sin\gamma
1396: +(v_t-v_p)(1-\cos\gamma)\right].
1397: \ee %$
1398: %
1399:
1400: \subsubsection{The small scattering angle approximation}
1401: Equation (\ref{eq:U}) shows that $U^2\approx 0.4v_p^2$ for test
1402: particles with $\ab>>\ab_p$ scattering off Jupiter mass planets. From
1403: equation (\ref{eqn: scattering angle}), we find
1404: %
1405: \be \label{eqn: small angle}%$
1406: \gamma\approx 2GM_p/(sU^2) = 2{\root 3\of 3}\mu_p^{2/3}
1407: \left({v_p\over U}\right)^2\left(r_H\over s\right).
1408: \ee %$
1409: %
1410: %%This small scattering angle approximation holds for
1411: %
1412: %%\be %$
1413: %%s>>{\root 3\of 3}\mu_p^{2/3}\left(v_p\over U\right)^2r_H
1414: %%=\mu_p \left(v_p\over U\right)^2a_p.
1415: %%\ee %$
1416: %
1417: The small angle approximation is valid for $s\gtsim s_\gamma$, where
1418: %
1419: \be %$
1420: s_\gamma\approx\mu_p \left(v_p\over U\right)^2\ab_p
1421: \ee %$
1422: %
1423:
1424: There is another constraint on the minimum value of the impact
1425: parameter $s$, namely that the test particle does not physically
1426: collide with the planet. This defines $s_{min}$, the impact parameter
1427: for which the periapse $\bar q=r_p$, where $r_p$ is the radius of the
1428: planet. For $\ab>>\ab_p$, $s_{min}\approx r_p\sqrt{2\mu_p(v_p/U)^2 \ab_p/r_p}$. At 5AU
1429: $s_{min}\approx 7 r_p$ for a planet with Jupiter's mass and
1430: radius. Thus non-collisional close encounters occur if the impact
1431: parameter satisfies
1432: %
1433: \be %$
1434: 7r_p\lesssim s\lesssim r_H,
1435: \ee %$
1436: %
1437: For a Jupiter mass planet at 5 AU, $s_\gamma\approx 25 r_p$,
1438: about four times larger than the minimum impact parameter. For the
1439: same values of $M_p$ and $\ab_p$, $r_H\approx 750r_p$. In other words,
1440: the small angle approximation is valid for almost any impact parameter
1441: that does not lead to a physical collision.
1442:
1443: Assuming $\gamma<<1$, the change in energy due to a single close
1444: encounter for a test particle with $\ab>>\ab_p$ is
1445: %
1446: \be \label{eqn: delta E}%$
1447: \Delta E(\mu_p,\ab_p,s)\approx 2^{3/2}{\root 3\of 3}\,\mu_p^{5/6}
1448: \left(v_p\over U\right)^2
1449: \left(r_H\over s\right)v_p^2.
1450: \ee %$
1451: %
1452:
1453:
1454:
1455: \subsubsection{The ejection velocity}
1456:
1457: We are now in position to estimate the typical ejection velocity of a
1458: test particle interacting with a Jupiter mass planet. The particle
1459: must have an energy satisfying $-\Delta E<E<0$ when the last encounter
1460: occurs. On average it will emerge with an energy $E\approx \Delta E/2$
1461: after the encounter, so it is ejected with a velocity
1462: %
1463: \be \label{eqn: kick}%$
1464: v_\infty(\mu_p,\ab_p,s)\approx2\mu_p^{5/12}\left(v_p\over U\right)\sqrt{r_H\over s}v_p.
1465: \ee %$
1466: %
1467: Note that the ejection velocity scales as the $1/2$ power of the
1468: planet's semimajor axis $\ab_p$. The scaling with the mass ratio $\mu_p$
1469: is complicated by the appearance of the Safronov number $U/v_p$. From
1470: equation (\ref{eq:U}), the Safronov number is independent of $\mu_p$ for
1471: $\mu_p<<6\times10^{-4}$, but scales as $\mu_p^{1/6}$ for larger
1472: $\mu_p$. Hence the ejection velocity will scale as $\mu_p^{5/12}$ for
1473: Saturn mass planets, and as $\mu_p^{1/4}$ for planets significantly
1474: more massive than Jupiter. If the planet has a substantial
1475: eccentricity, as many extrasolar planets do, then $v_t\approx v_p$ and
1476: the ejection velocity will scale as $\mu_p^{1/3}$.
1477:
1478: For a Jupiter mass planet at $5$ AU from a solar mass star,
1479: %
1480: \be \label{eqn: Jupiter kick}%$
1481: v_\infty(\mu_p,\ab_p,s)\approx 2\left(\mu_p \over10^{-3}\right)^{1/3}
1482: \left(5{\,\rm AU}\over \ab_p\right)^{1/2}
1483: \left(v_p\over U\right)
1484: \sqrt{r_H\over s}{\km/\s}.
1485: \ee %$
1486: %
1487: Let $x=s/r_H$; then the cross section for an encounter with impact
1488: parameter between $s$ and $s+ds$ is $2\pi sds=2\pi r_H^2\, xdx$. We
1489: average over $x$ to find the mean escape velocity. We integrate from
1490: $x_{\gamma}\equiv s_{\gamma}/r_H \approx \mu_p^{2/3}<<1$ to $s=\alpha
1491: r_H$, where $\alpha$ is a dimensionless constant of order unity, and
1492: $s_{\gamma}$ is the minimum value of $s$ for which equation (\ref{eqn:
1493: small angle}) is valid. Doing the integration,
1494: %
1495: \be %$
1496: \left<v_\infty\right>={\int_{x_\gamma}^\alpha \pi x v_\infty(x) dx\over
1497: \int_{x_\gamma}^\alpha \pi x dx}\approx {4v_\infty(r_H)\over3\sqrt{\alpha}},
1498: \ee %$
1499: %
1500: where we have neglected terms of order $\mu_p v_\infty(r_H)$; we use the
1501: notation $v_\infty(r_H)=v_\infty(\mu_p, \ab_p, s=r_H)$.
1502: The rms escape velocity is calculated in a similar manner. We find
1503: %
1504: \be \label{eqn: v_rms}%$
1505: v_{rms}\approx \sqrt{2\over \alpha}v_\infty(r_H),
1506: \ee %$
1507: %
1508: so the spread in escape velocities is similar to the mean escape velocity.
1509:
1510: For those rare
1511: encounters with $s_{min}<s<s_\gamma$, the scattering angle is of order
1512: unity, and the velocity kick experienced by the particle is larger
1513: than the estimate in eqn. (\ref{eqn: kick}). We take
1514: $\sin\gamma\approx\cos\gamma\approx 1/\sqrt{2}$, so
1515: %
1516: \be %$
1517: \Delta E\approx v_p^2,
1518: \ee %$
1519: %
1520: and
1521: %
1522: \be %$
1523: v_\infty\approx v_p,
1524: \ee %$
1525: %
1526: much larger than the small angle limit. About 1\% of the particles
1527: will be ejected with velocities of tens of kilometers per second.
1528:
1529: \subsubsection{Ejection time scale compared to collisional timescale}
1530: So far we have assumed that the grains are ejected before they suffer
1531: sufficient collisions with other dust particles to alter their orbits
1532: substantially. We are now in a position to check this assumption. We
1533: start with an estimate of the dust-dust collision time. The optical
1534: depth $\tau$ for the most massive debris disks is of order
1535: $0.001$. This optical depth is due to particles with radii
1536: $a\approx\lambda/2\pi$, where $\lambda$ is the wavelength at which the
1537: disk is observed. Since the disks are usually detected by their IR
1538: excesses, the dust particles responsible for the optical depth have
1539: $a\approx0.3\micron$ or smaller. We are interested in larger ejected
1540: particles, with $a_{eject}=25\micron$. These larger particles will
1541: suffer a collisions with a particle of size $a_{target}$ after roughly
1542: $1/\tau(a_{target})$ orbits, where $\tau(a_{target})$ is the optical
1543: depth to particles of size $a_{target}$.
1544:
1545: Using the scaling $mf_m\propto m^{-1}$, the optical depth in particles
1546: of size $a_{target}$, as viewed at a wavelength of order or smaller
1547: than $a_{target}$, is
1548: %
1549: \be %$
1550: \tau(a_{target})=\left(0.3\micron\over a_{target}\right)\tau(a=0.3\micron).
1551: \ee %$
1552: %
1553:
1554: The typical number of orbits required for our test (ejected) particles
1555: to sweep up their own mass in smaller particles, and hence have their
1556: orbits altered substantially, or to strike a larger particle, is
1557: %
1558: \be %$
1559: N \approx
1560: {1\over\tau(a=0.3\micron)}
1561: \Bigg\{
1562: \begin{array}{cc}
1563: \left({a_{eject}\over 0.3\micron}\right)
1564: \left(a_{eject}\over a_{target}\right)^2, & a_{target}\le a_{eject}\\
1565: \left({a_{target}\over 0.3\micron}\right), & a_{target}\ge a_{eject}
1566: \end{array}
1567: \phantom{\}}
1568: \ee %$
1569: %
1570: orbital periods.
1571:
1572: Thus the most efficient way to alter the orbit of a particle is to
1573: collide with another particle of the same size, assuming that the mass
1574: is logarithmically distributed, and assuming that gravity plays no
1575: role in the collision. The typical number of orbits for a $25\micron$
1576: size particle to collide with a particle of its own mass is
1577: $\sim1/\tau(a=25\micron)\approx7\times10^4$.
1578:
1579:
1580: Dust-dust collisions can also shatter the grains. The binding energy
1581: for a target grain of mass $m$ is $\zeta m$. A collision with a grain of mass
1582: $m_c$ moving with relative velocity $v$ will shatter the target if the
1583: kinetic energy exceeds the binding energy of the target. Allowing for
1584: the possibility that only a fraction $f_{KE}$ of the kinetic energy is
1585: available to disrupt the target, the critical mass needed to shatter
1586: the target is
1587: %
1588: \be %$
1589: m_c={2\over f_{KE}}{\zeta\over e^2v_p^2}m.
1590: \ee %$
1591: %
1592: In deriving this result we assume that the relative velocity is
1593: $ev_p$, corresponding to the random velocity of a grain suffering
1594: close encounters with a massive planet of semimajor axis $\ab_p$. The
1595: eccentricity of the target grain undergoes a random walk in $e$,
1596: starting at zero and evolving to $e=1$, $e(t)\approx \sqrt{D_e t}$, so
1597: we take $e=2/3$. The minimum mass needed to shatter the target grain
1598: is then $m_c\approx 0.5(\ab_p/5 {\rm AU})m/f_{KE}$, roughly a mass equal
1599: to that of the target.
1600:
1601: How long does it take for the typical particle to be
1602: ejected? The test particles undergo a random walk in energy, with a
1603: step size given by equation (\ref{eqn: delta E}). The number of
1604: collisions needed to random walk from $E=-GM/2\ab_p=-v_p^2/2$ to $E=0$ is
1605: $N\approx 3^{2/3}/16\mu_p^{4/3}$. The probability of a close encounter on
1606: each periapse passage is
1607: %
1608: \be %$
1609: P\approx {\pi r_H^2\over(2\pi \ab_p r_H)}\approx{1\over 2}
1610: \left({\mu_p\over 3}\right)^{1/3}.
1611: \ee %$
1612: %
1613: The total number of orbits up to ejection is
1614: %
1615: \be %$
1616: N_{eject}\approx{3\over8}\mu_p^{-5/3}\approx4\times10^4
1617: \left(10^{-3}\over\mu_p\right)^{5/3}.
1618: \ee %$
1619: %
1620: We conclude that particles larger than about $25\micron$ will be
1621: ejected before their orbits are significantly altered by collisions
1622: with smaller or similar size dust grains.
1623:
1624: \subsubsection{Numerical Results}
1625: We have carried out numerical integrations of test particles in the
1626: gravitational field of one or more massive planets orbiting a star. We
1627: use the publicly available SWIFT (Levison \& Duncan 1994)
1628: integration package, which is based on the Wisdom \& Holman (1991)
1629: symplectic integration scheme. We used the rmvs3 integrator, in order to
1630: integrate through close encounters between the test particles and the
1631: planets. All integrations were for 100 million years, with a timestep
1632: chosen to be small enough to resolve periapse passage for the
1633: estimated most extreme test particle orbits, or to resolve the
1634: perijove passage for planet grazing close encounters (taken to be at
1635: $2r_p$), whichever was smaller. If a particle passed within $2r_p$ is
1636: was deemed to have collided with the planet and was removed from the
1637: integration.
1638:
1639: The initial eccentricities and inclinations (measured from the
1640: planet's orbital plane) of the test particles were generally set to
1641: $0.1$ and $0.087$ radians respectively, although we tried runs with
1642: other values. The final ejection velocities did not depend strongly on
1643: the initial $e$ and $i$. The test particles were given semimajor axes
1644: ranging from $0.5$ to $1.5\ab_p$ in single planet cases, since in those
1645: cases particles starting at larger distances from the planet were
1646: typically not ejected by the time the integrations were halted.
1647:
1648: Figure (\ref{fig:swift}) shows the result of a numerical integration
1649: of $\sim 600$ test particles in the gravitational field of Jupiter and
1650: the Sun. Our analytic calculations should be a good approximation to
1651: this case, since $e_J=0.048$. The test particles were started with a
1652: range of semimajor axes between $x$ and $y$, with initial
1653: eccentricities $e\sim0.05$ and inclinations $i\sim0.05$ radians. The
1654: figure shows the velocities at which the particles were ejected,
1655: including the (small) correction for their finite distance from the
1656: Sun when the integration was stopped. The stopping criterion was that
1657: the test particle have a positive energy relative to the Sun and a
1658: semimajor axis larger than $100$AU. The mean escape velocity is
1659: $\sim1\km\s^{-1}$, slightly lower than predicted by equation
1660: (\ref{eqn: Jupiter kick}).
1661:
1662: We found that the final inclinations were small, typically within $10$
1663: degrees of the orbital plane of the planet, but with a significant
1664: minority of particles ejected at inclinations up to $30$ degrees.
1665:
1666: We checked the scaling of $v_{ej}$ with $\ab_p$ by varying the size of
1667: the planet's orbit over a decade. A least squares fit to the mean
1668: ejection velocity as a function of $\ab_p$ gives $v_{ej}\sim \ab_p^{-0.5}$
1669: with an error of about $0.05$ in the exponent. This is consistent with
1670: the predicted slope of $-0.5$ given by equation (\ref{eqn: kick}). We
1671: also checked the scaling with $M_p$, finding $v_{ej}\sim M_p^{0.28}$,
1672: compared with the prediction exponent of $1/4$. In summary, equations
1673: (\ref{eqn: kick}) and (\ref{eqn: Jupiter kick}) appear to be a good
1674: description of the ejection process.
1675:
1676: The numerical integrations employing multiple Jupiter mass planets
1677: gave similar results, with the test particle ejection velocity
1678: determined by the most massive of the planets in the integration. The
1679: same was true of integrations involving a single Jupiter mass planet
1680: together with several 1-10 Earth mass bodies.
1681:
1682: These results suggest that equations (\ref{eqn: Jupiter kick}) and
1683: (\ref{eqn: v_rms}) can be used to describe the ejection velocity of
1684: small particles in systems with one or more Jupiter mass planets.
1685:
1686:
1687:
1688: \subsection{Dust Luminosity}
1689:
1690: In order to estimate the (age-dependent)
1691: dust luminosity of a Vega-like star, we must
1692: first estimate the number of grains in the disk. Spangler et al.~(2001)
1693: have measured IR excesses for stars in several nearby clusters. Their
1694: results are consistent with the following relation: dust disk mass
1695: $\propto$ (age)$^{-2}$. We normalize this relation by estimating the
1696: (size-dependent) number of grains in the disk around $\beta$
1697: Pictoris, with an age $\approx 12 \Myr$ (Zuckerman et al.~2001).
1698:
1699: We assume that the grains around $\beta$ Pic are located at a distance
1700: $D = 100 {\rm AU}$ from the star
1701: and that they have a size distribution $dN/da = C a^{-3.5}$, where
1702: $N(a)$ is the number of grains with size $\le a$. This is the equilibrium
1703: size distribution that results when mass is redistributed among grain sizes
1704: via shattering in collisions (Dohnanyi 1969). We obtain the normalization
1705: constant $C$ by setting the emission at $800 \micron$ equal to the observed
1706: value of $F_{\nu} = (115 \pm 30) {\rm mJy}$ (Zuckerman \& Becklin 1993).
1707: The specific flux is given by
1708: \be
1709: \label{eq:F_nu}
1710: F_{\nu} = \frac{\pi}{D^2} \int da \frac{dN}{da} \Qabs a^2 B_{\nu}[T(a)]~~~,
1711: \ee
1712: where the absorption cross section is $\Qabs \pi a^2$, $B_{\nu}(T)$ is the
1713: Planck function, and the grain temperature $T(a)$ is determined by
1714: equating the absorption and emission rates. For $\beta$ Pic, we take
1715: effective temperature $T_{\rm eff} = 8200 \K$, luminosity
1716: $L = 8.7 L_{\odot}$, and radius $R = 1.47 R_{\odot}$ (Crifo et al.~1997).
1717: Grains with $a \gtsim 1 \cm$ contribute little to $F_{\nu}(800 \micron)$.
1718: Terminating the integral in equation (\ref{eq:F_nu}) at $a=1 \cm$
1719: yields $C = 4.65 \times 10^{25} \cm^{2.5}$ for silicate grains, implying
1720: a disk mass of $1.4 \times 10^{27} \g$ ($\approx 19$ lunar masses, 0.23
1721: Earth masses) in
1722: grains with $a \le 1 \cm$. If we assume graphite (water ice) rather than
1723: silicate composition, then the estimated mass increases by a factor of
1724: 1.6 (4.6). Since pure water ice grains are very unlikely, our mass
1725: estimate is robust against variations in compositions. If the grains are
1726: located $30 \pc$ ($500 \pc$) rather than $100 \pc$ from the star, then the
1727: estimated mass decreases by a factor of 2 (increases by a factor of 3).
1728: Thus, though the dust actually will be distributed over a range of distances
1729: from the star, this does not seriously affect the estimate of the total
1730: dust mass.
1731:
1732: Our result can be compared with the analysis of Li \& Greenberg (1998),
1733: who constructed a detailed dust model to account for the emission (at
1734: multiple wavelengths) from the dust around $\beta$ Pic, using fluffy grains.
1735: In their model, the total dust mass needed to produce the observed
1736: emission is $\sim 2 \times 10^{27} \g$, lying mostly beyond 100 AU from
1737: the star.
1738:
1739: It is interesting to note that estimates of the dust mass in circumstellar
1740: disks typically assume that the grains are in the Rayleigh limit, i.e.,
1741: that $a \ll \lambda$, the wavelength of the observation (see, e.g., the
1742: review by Zuckerman 2001). In this regime, $\Qabs \propto a$. In the
1743: opposite regime ($a \gg \lambda$), $\Qabs \sim 1$. Suppose $a_0$ is
1744: the minimum $a$ for which $\Qabs \sim 1$, for fixed $\lambda$. Then,
1745: $\Qabs \sim a/a_0$ in the Rayleigh limit. The opacity $\kappa$
1746: (absorption cross section per unit mass) of the emitting grains is
1747: the crucial quantity for estimating dust mass from emission;
1748: $\kappa = 3 \Qabs / 4 a \rho$. In the Rayleigh limit,
1749: $\kappa \sim 3 / 4 a_0 \rho$. In our estimate of the dust mass (with our
1750: assumption that $dN/da \propto a^{-3.5}$), the emission is dominated by
1751: grains with $a \sim a_0$, since these are the most abundant grains with
1752: $\Qabs \sim 1$. Thus, $\kappa \sim 3 / 4 a_0 \rho$ for our estimate as
1753: well as for the Rayleigh limit, and these two different methods yield
1754: approximately the same mass in grains contributing significantly
1755: to the observed
1756: emission. We feel that our scenario, with a Dohnanyi size distribution,
1757: is more realistic than the standard scenario, in which all of the grains
1758: that contribute significantly to the emission are supposed to be in the
1759: Rayleigh limit.
1760:
1761: Combining the above estimate for the number of grains in the disk around
1762: $\beta$ Pic with the Spangler et al.~(2001) result, we estimate
1763: %
1764: \be %$
1765: \frac{dN}{da} \sim 4.7 \times 10^{25} \cm^{2.5} \left( \frac{12 \Myr}{t}
1766: \right)^2 a^{-3.5}~~~,~~~~~a \ltsim 1 \cm~~~,
1767: \ee %$
1768: %
1769: where $t$ is the age of the system.
1770:
1771: The Spangler et al.~(2001) result suggests that grains are lost at a rate
1772: $dN/dt \sim -2 N/t$. The following could be important sinks for the
1773: grains: 1. gravitational ejection from the system, 2. gravitational
1774: ejection into the star, 3. incorporation into planets or
1775: planetesimals, and 4. ejection of small shattering fragments by
1776: radiation pressure. We will not attempt to estimate the relative importance
1777: of these sinks in this paper. Rather, we will introduce the unknown
1778: factor $f_{\rm ej}$, equal to the fraction of the grains that are lost by
1779: gravitational ejection. Then, the specific dust luminosity is
1780: %
1781: \be
1782: L_{v, a}(t, \vej, a) = 7.8 \times 10^{18} \yr^{-1} \cm^{-1} (\cm/\s)^{-1}
1783: \left( \frac{12 \Myr}{t}\right)^3 \left( \frac{a}{\cm} \right)^{-3.5}
1784: f_{\rm ej} f_v(\vej)~~~,
1785: \ee
1786: %
1787: where we have taken time equal to the age of the star when the grains are
1788: ejected and
1789: $f_v(\vej) d\vej$ is the fraction of the grains ejected with speed
1790: between $\vej$ and $\vej + d\vej$.\footnote{We assume that $f_v$ is
1791: independent of $a$.}
1792:
1793: At very early ages, the primordial gas disk will not yet have dissipated,
1794: and the drag force on grains could dramatically reduce the dust
1795: luminosity. Also, we do not expect large grains to be ejected prior to
1796: planet formation. Thus, $L_{v, a} = 0$ for $t < t_{\rm cr}$, an unknown
1797: critical age. We will assume that $t_{\rm cr}$ lies in the range 3 to
1798: $10 \Myr$.
1799:
1800: \subsection{Flux at Earth}
1801:
1802: Assuming $f_{\rm survive} = 1$,
1803: the flux, at Earth, of grains with size $\ge a$, $F(a)$, is given by
1804: \be
1805: \label{eq:flux_at_earth}
1806: F(a) = 8.2 \times 10^{-5} \yr^{-1} \, {\rm km}^{-2} f_{\rm beam}
1807: f_{\rm ej} \left( \frac{d}{10\pc}
1808: \right)^{-2} \left( \frac{a}{10\micron} \right)^{-2.5}
1809: \int d\vej \, f_v(\vej) \frac{v_{d, \odot}}{\vej} Q(t_{\rm ej} = t -
1810: d/\vej)~~~,
1811: \ee
1812: where
1813: \be
1814: Q(t_{\rm ej}) = \cases{(12 \Myr/t_{\rm ej})^3 &,
1815: $t_{\rm ej} \ge t_{\rm cr}$\cr
1816: 0 &, $t_{\rm ej} < t_{\rm cr}$\cr}~~~.
1817: \ee
1818:
1819: If we assume $v_{d,\odot}/v_{ej}\approx10$, this yields a flux of
1820: $2.6\times10^{-21}(25\micron/a)^{2.5}(10\pc/d)^2\cm^{-2}\s^{-1}$,
1821: which may be compared with Fig. (\ref{fig:flux}); the AMOR point
1822: source supplies a flux at Earth of $10^{-17}\cm^{-2}\s^{-1}$. It would
1823: appear that the point source is not due to a debris disk.
1824:
1825: We will be interested in sources within the
1826: LB, so $f_{\rm survive}=1$ when $d < r_B$. With $\Phi \approx 0.94 \,
1827: {\rm V}$ and $\rho = 3.5 \g \cm^{-3}$ (appropriate for silicate grains
1828: in the LB), equation (\ref{eq:gyroradius}) yields a critical grain
1829: radius \be
1830: \label{eq:a_cr}
1831: a_{\rm cr} \equiv 10 \micron \left( \frac{d}{9.0 \pc} \right)^{1/2}
1832: \left( \frac{v}{10 \kms} \right)^{-1/2}~~~;
1833: \ee
1834: $f_{\rm survive} = 1$ when $a > a_{\rm cr}$. We will use
1835: $v_{\ast, \sun}$ as a good approximation to $v_{d, \sun}$ in equation
1836: (\ref{eq:a_cr}).
1837:
1838: Suppose the threshold of detectability requires 20 events per year, so
1839: that the threshold flux $F_{\rm th}=20 \yr^{-1} A_{col}^{-1}$. Then,
1840: we can estimate the threshold distance $d_{\rm th}$ out to which
1841: sources might be detected by making the optimistic assumptions that
1842: $t_{\rm ej} = t_{\rm cr}$ and $f_{\rm ej}=1$ (we will also assume that
1843: $f_{\rm beam}=1$). Inserting equation (\ref{eq:a_cr}) into equation
1844: (\ref{eq:flux_at_earth}) and assuming $v_{d, \odot}/\vej \sim 10$
1845: yields
1846: %
1847: \be
1848: d_{\rm th} \approx 0.42 \left( \frac{A_{col}}{1 \km^2}
1849: \right)^{4/13} \left( \frac{t_{\rm cr}}{12 \Myr} \right)^{-12/13}
1850: \pc~~~.
1851: \ee
1852: %
1853: In Table \ref{tab:d_th} we give $d_{\rm th}$ and $a_{\rm cr}$ for the
1854: cases that $T_{\rm cr} = 3$ or $10 \Myr$ and $A_{col} = 10^4 \km^2$ or
1855: $10^6\km^2$; recall that AMOR has $A_{col}\approx10\km^2$.
1856:
1857: Since the flux drops off as $t_{\rm ej}^{-3}$, it is highly advantageous
1858: to catch the star when $t_{\rm ej} \sim t_{\rm cr}$. In Figure
1859: \ref{fig:age_ranges}, we plot, as a function of $d$, the upper and lower
1860: ages for which $t_{\rm cr} \le t_{\rm ej} \le 2 t_{\rm cr}$, assuming a
1861: range of ejection speeds between $0.5 \kms$ and $2 \kms$. Since the
1862: total volume and favorable age range both increase with $d$, distant
1863: stars that are possibly too old to have observable IR excesses today
1864: may be a significant source population, despite the $d^{-2}$ decrease
1865: in flux. However, nearby identifiable Vega-like stars are more
1866: attractive sources, since we could then combine the dust flux information
1867: with other observations to learn more about a particular object.
1868:
1869: \subsection{Candidate Sources}
1870:
1871: \subsubsection{\label{sec:Gliese} Gliese Catalog Stars}
1872:
1873: Are there any nearby Vega-like stars for which we can expect a detectable
1874: dust flux at Earth? To answer this question, we consider stars from the
1875: Gliese Catalog with far-infrared excesses from IRAS (see Table X in
1876: Backman \& Paresce 1993).
1877:
1878: In Table \ref{tab:Gliese}, we give each star's distance $d$ and its space
1879: velocity $(U,V,W)$ with respect to the Sun. In calculating these quantities,
1880: we have taken coordinates, parallaxes, proper motions, and radial velocities
1881: from SIMBAD.\footnote{Note that we exclude one star from Table \ref{tab:Gliese} in
1882: Backman \& Paresce (1993), Gl 245, because SIMBAD does not give its
1883: radial velocity.} We estimate a star's age $T$ from its IR excess $\tau$,
1884: which is the fraction of the star's luminosity that is re-emitted in the
1885: IR by dust. Adopting the Spangler et al.~(2001) result that
1886: $\tau \propto t^{-2}$, we take
1887: \be
1888: \label{eq:T_tau}
1889: t = t(\beta \, {\rm Pic}) \left[ \frac{\tau(\beta \, {\rm Pic})}{\tau}
1890: \right]^{0.5}~~~.
1891: \ee
1892:
1893: The dust fluxes in Table \ref{tab:Gliese} were calculated using
1894: equation (\ref{eq:flux_at_earth}) with limiting grain size $a=\max(10
1895: \micron, a_{\rm cr})$ ($a_{\rm cr}$ is taken from eq.~\ref{eq:a_cr}
1896: with $v = v_{\ast, \sun}$), a flat distribution of ejection velocities
1897: between $0.5 \kms$ and $3 \kms$, $t_{\rm cr} = 3 \Myr$, and $f_{\rm
1898: beam} = f_{\rm ej} =1$. Thus, these estimates are optimistic. In \S
1899: \ref{sec:Discussion} below, we suggest how to build a radar detector
1900: with $A_{col}\approx3\times10^4\km^2$. Four of the stars in Table
1901: \ref{tab:Gliese} would yield 20 or more meteors per year (i.e., $F > 6
1902: \times 10^{-4} \yr^{-1} \km^{-2}$) for such a system. Since $t_{\rm
1903: ej}$ is substantially greater than $t_{\rm cr}$ for most of the stars,
1904: increasing $t_{\rm cr}$ to $10 \Myr$ yields reduced fluxes for only
1905: three stars: Gl 219, 297.1, and 673.1. The flux for Gl 219 ($\beta
1906: \,$Pic) vanishes, whereas the fluxes for Gl 297.1 and 673.1 remain
1907: greater than $2 \times 10^{-5} \yr^{-1} \km^{-2}$.
1908:
1909: The apparent location of a dust source on the sky is determined by the velocity
1910: of the dust relative to the Sun (eq.~\ref{eq:v_dust_Sun}). If the dust
1911: ejection speed $\vej$ were much greater than the star's speed
1912: $v_{\ast, \sun}$, then the dust would
1913: appear to come from the location of the star itself (since the speed of
1914: light is much greater than the star's speed). However, $\vej$ is
1915: typically an order of magnitude or more smaller than $v_{\ast, \sun}$.
1916: Thus, the location of the dust source on the sky is primarily determined by
1917: $\vec{v}_{\ast, \sun}$ and need not be anywhere near the actual location of
1918: the source.
1919:
1920: Since $\vec{v}_{\ast, \sun}$ is determined in part by the velocity
1921: $\vec{v}_{\sun, {\rm LSR}}$ of
1922: the Sun relative to the local standard of rest (LSR), we may expect some
1923: degree of clustering of the apparent directions around the solar apex
1924: (i.e., the direction of $\vec{v}_{\sun, {\rm LSR}}$).
1925: Dehnen \& Binney (1998) give $U=10.00 \pm 0.36$, $V=5.25 \pm 0.62$, and
1926: $W=7.17 \pm 0.38 \kms$ for $\vec{v}_{\sun, {\rm LSR}}$. They find that the
1927: velocity of the Sun with respect to nearby young stars is similar, except
1928: that $V \approx 12 \kms$ in this case. The coordinates of the solar
1929: apex with respect to the LSR are $(l, b) = (27.7\arcdeg, 32.4\arcdeg)$
1930: or $(\lambda, \beta) = (248.4\arcdeg, 32.2\arcdeg)$, while those with
1931: respect to nearby young stars are $(l, b) = (50.2\arcdeg, 24.7\arcdeg)$
1932: or $(\lambda, \beta) = (265.7\arcdeg, 48.9\arcdeg)$. In addition to
1933: Galactic coordinates $(l, b)$, we also make use of ecliptic coordinates:
1934: ecliptic longitude $\lambda$ and ecliptic latitude $\beta$. These
1935: coordinates are useful in radar studies of meteors, because most of the
1936: observed grains originate in the ecliptic plane.
1937: In Table \ref{tab:Gliese}, we give the apparent direction to the dust
1938: stream in ecliptic coordinates. Figure \ref{fig:Gliese} is a plot of
1939: these directions on the sky.
1940:
1941: \subsubsection{Young Clusters}
1942:
1943: Nearby young clusters could potentially yield strong dust fluxes. The
1944: cluster members may not have appeared in searches for the Vega Phenomenon,
1945: since the stellar luminosities may be too low. However,
1946: most nearby clusters have poor tuning between distance and age. There are
1947: some very young clusters that are too far away for the dust to have reached
1948: us yet and some nearer clusters that are much older (see Table 1 in
1949: Spangler et al.~2001).
1950:
1951: The recently discovered Tucana Association
1952: (Zuckerman, Song, \& Webb 2001) might be more suitable. Zuckerman \&
1953: Webb (2000) find a distance $\sim 45 \pc$, age $\sim 40 \Myr$ and space
1954: velocity $(U,V,W) \approx (-11, -21, 0) \kms$. Stelzer \& Neuh\"{a}user
1955: (2000) find a younger age of 10--30$\, \Myr$. Adopting numbers from
1956: Zuckerman \& Webb and making use of the convenient coincidence that
1957: $1 \kms = 1.02 \pc/\Myr$, we find that dust ejected at speed
1958: $\vej = 1.5 \kms$ when the stars were $15 \Myr$ old would be reaching us
1959: today. The position of the dust stream on the sky would be
1960: $(\lambda, \beta) = (126 \arcdeg, -46 \arcdeg)$.
1961:
1962: The Pleiades cluster is another potential source, with a distance of
1963: $118 \pc$ and an age of $\approx 120 \Myr$ (Spangler et al.~2001).
1964: Although the large distance would suppress the flux (because of both the
1965: $d^{-2}$ dependence in eq.~\ref{eq:flux_at_earth} and the need for
1966: larger grains in order to avoid magnetic deflection), this could be
1967: partially compensated by the large number of stars in the cluster.
1968: Robichon et al.~(1999) find $(U, V, W) = (-6.35, -24.37, -13.02) \kms$
1969: for the cluster motion. Ignoring the Galactic potential, this would yield
1970: $(\lambda, \beta) \approx (276\arcdeg, 73\arcdeg)$ for the position of the
1971: dust stream.
1972:
1973: \subsection{Has Dust from $\beta \,$Pic Been Detected?}
1974:
1975: Baggaley (2000) detected a ``discrete'' source of radar meteoroids with a
1976: central location $(\lambda, \beta) \approx (280 \arcdeg, -56 \arcdeg)$, an
1977: angular diameter of $\sim 30 \arcdeg$, and a dust speed relative to the
1978: Sun $v_{d, \sun} \approx 13 \kms$. He claimed that the central location
1979: and dust speed could be reproduced if $\beta \,$Pic were the source and
1980: the grains were ejected with speed $\vej \approx 29 \kms$.
1981:
1982: This result seems unlikely on theoretical grounds, since it is not
1983: clear how the grains can be ejected with such large speeds. In
1984: addition, we do not find that dust emitted from $\beta \,$Pic
1985: reproduces the location of the discrete source for any $\vej$.
1986: Baggaley did his calculations in the LSR frame, taking the Sun's
1987: motion with respect to the LSR from Binney \& Tremaine (1987), who
1988: give $(U_{\sun}, V_{\sun}, W_{\sun}) = (9, 12, 7) \kms$. Baggaley
1989: found the direction to the discrete source to be $(\lambda, \beta) =
1990: (49\arcdeg, -72\arcdeg)$, $(l, b) = (259\arcdeg, -28\arcdeg)$, whereas
1991: we find $(\lambda, \beta) = (58.6\arcdeg, -81.3\arcdeg)$, $(l, b) =
1992: (267.4\arcdeg, -34.1\arcdeg)$. Note that the two coordinate pairs
1993: given by Baggaley do not actually correspond to the same point on the
1994: sky.
1995:
1996: In Figure \ref{fig:baggaley}, we plot
1997: the sky position of a dust stream emitted by $\beta \,$Pic for various
1998: values of $\vej$ (triangles). The box indicates the central location of
1999: Baggaley's discrete source. For reasonable ejection speeds
2000: ($\vej \ltsim 3 \kms$), the location of the dust stream differs dramatically
2001: from that of the discrete source. For $\vej \approx 30 \kms$, the locations
2002: are much closer and $v_{d, \sun} \approx 13 \kms$. Although the dust
2003: stream does lie within the $\sim 30 \arcdeg$ extent of the source in this
2004: case, it is still $12 \arcdeg$ away from the center of the source.
2005:
2006: The particle flux coming from the discrete source is several orders of
2007: magnitude larger than we would expect from a debris disk at 20
2008: parsecs. From eqn. (\ref{eq:flux_at_earth}) we find that Beta Pic
2009: produces a flux of $a=10\micron$ particles of $10^{-3}$ per square
2010: kilometer per year at Earth. Given our estimate of the collecting area
2011: of the AMOR detector, somewhat less than $10\km^2$, AMOR is not
2012: capable of detecting such a low flux.
2013:
2014: Thus, it appears that Baggaley's discrete source is not related to
2015: $\beta \,$Pic.
2016:
2017: \subsection{Potential Distributed Sources}
2018:
2019: In addition to the discrete source, Figures 2b and c in Baggaley
2020: (2000) suggest the presence of a distributed, band-like feature.
2021: Here we discuss three possible sources for a distributed feature,
2022: namely the Galactic plane, Gould's Belt, and the spiraling of grains
2023: in the local magnetic field.
2024:
2025: \subsubsection{The Galactic Plane}
2026:
2027: The scale height of young stars above the Galactic plane is
2028: $\sim 90 \pc$ (Gilmore \& Reid 1983). Thus, we might expect to see a
2029: signature of the plane in the dust flux since we can detect grains from
2030: sources beyond $100 \pc$ if $a$ and $v$ are large enough (but $v$ must
2031: not be so large that the grains are destroyed). The effects
2032: of the Galactic gravitational potential cannot be ignored for such large
2033: distances, but we do not expect the Galactic potential to deflect grains
2034: out of the plane.
2035: As mentioned in \S \ref{sec:Gliese},
2036: the apparent direction of a dust stream with a low ejection speed is
2037: primarily determined by the velocity of the source with respect to the
2038: Sun. Since the Galactic plane is a feature in physical space rather than
2039: velocity space, it may seem that it should not actually appear as a band
2040: on the sky in a dust flux map. However, if the radial velocity of a
2041: source (with respect to the Sun) exceeds the dust ejection speed, then,
2042: unless the source passed near the current location of the Sun at some time
2043: in the past, the dust will never reach the Sun. Thus, though the velocity
2044: of the source with respect to the Sun determines the direction of the dust
2045: stream, the location of the source does restrict which velocities can
2046: give rise to an observable dust stream.
2047:
2048: From equations (\ref{eq:flux_at_earth}) and (\ref{eq:a_cr}),
2049: the dust flux $F \propto d^{-3.25}$, so it is not clear that distant
2050: stars in the plane can indeed produce a noticeable band on top of the
2051: relatively isotropic distribution produced by nearby stars. The
2052: following simple estimate shows that a band is expected, although at
2053: a low contrast level. For stars with distance $R < 90 \pc$, the volume
2054: element (in which young stars are found) is $dV = 4 \pi R^2 dR$, while
2055: $dV = (90 \pc) 2 \pi R dR$ for $R > 90 \pc$. Thus, the ratio of the
2056: flux $F_{\rm plane}$ due to the Galactic plane to the isotropic flux
2057: $F_{\rm iso}$ (due to nearby stars) is
2058: \be
2059: \frac{F_{\rm plane}}{F_{\rm iso}} \approx \frac{(180 \pc)\pi
2060: \int_{90 \pc}^{\infty} dR R^{-1.25}}{4 \pi \int_0^{90 \pc} dR
2061: R^{-0.25}} = 1.6~~~,
2062: \ee
2063: where we have multiplied each integrand by a factor $R$ to account for
2064: the increased probability of catching a star when $T_{\rm ej} \approx
2065: T_{\rm cr}$.
2066:
2067: Since the components of random stellar velocities perpendicular to the
2068: plane are smaller than those parallel to the plane,
2069: the Galactic plane should appear as a fuzzy band on
2070: the sky even considering only stars closer than $90 \pc$.
2071:
2072: Note that the observed dust flux band
2073: should be warped in such a way as to favor the direction of the solar
2074: apex.
2075:
2076: \subsubsection{Gould's Belt}
2077:
2078: Gould's Belt is a band on the sky, inclined $\approx 20\arcdeg$ with
2079: respect to the Galactic plane and with an ascending node $l_{\Omega} \approx
2080: 280\arcdeg$, along which young stars tend to lie
2081: (see P\"{o}ppel 1997 and de Zeeuw et al.~2001 for reviews).
2082: From a Hipparcos study of early-type stars, Torra, Fern\'{a}ndez, \&
2083: Figueras (2000) concluded that $\approx 60\%$ of the stars younger than
2084: $60 \Myr$ and within $600 \pc$ of the Sun are in Gould's Belt.
2085: OB associations, young star clusters, and molecular clouds have long been
2086: known to trace Gould's Belt. Recently, Guillout et al.~(1998) detected
2087: late-type stellar members of the Gould Belt population, by
2088: cross-correlating the ROSAT All-Sky Survey with the Tycho catalog.
2089: Thus, Gould's Belt could be responsible for a distributed grain flux
2090: feature.
2091:
2092: Guillout et al.~(1998) found that the three-dimensional structure of
2093: Gould's Belt is disk-like. The outer rim of the Gould Disk is
2094: ellipsoidal, with a semi-major axis of $\approx 500 \pc$ and a semi-minor
2095: axis of $\approx 340 \pc$. The center of the structure is $\approx 200 \pc$
2096: from the Sun, towards $l \approx 130\arcdeg$. On the near side
2097: ($l \approx 310\arcdeg$), the inner edge of the Gould Disk lies only
2098: $\approx 30 \pc$ from the Sun, although most of the young stars lie
2099: beyond $\approx 80 \pc$.
2100:
2101: Estimates of the age of Gould's Belt range from 20--90$\, \Myr$
2102: (see Torra et al.~2000 for a review); Torra et al.~favor an age between
2103: 30 and 60$\, \Myr$. Grains traveling with $\vej = 1.5 \kms$ reach a
2104: distance of $\approx 90 \pc$ from their star of origin in $60 \Myr$.
2105: Thus, only the near side of the Gould Disk should be visible in the
2106: dust flux.
2107:
2108: The kinematics of Gould's Belt is still poorly understood, but it is clear
2109: that the member stars are undergoing some sort of expansion. If the
2110: expansion dominates the random velocities, then the near side of Gould's
2111: Belt should appear on the opposite side of the sky in the dust flux,
2112: i.e., centered on $l \approx 130\arcdeg$. If the random velocities are
2113: more important, then we would expect to see a feature lying in the same
2114: direction as the near side of Gould's Belt, due to stars moving towards
2115: us.
2116:
2117: \subsubsection{Spiraling of Grains in the Local Magnetic Field}
2118:
2119: Thus far, we have only considered grains that are not dramatically
2120: deflected as they travel from their source to the Earth. Suppose the
2121: magnetic field were uniform throughout the Local Bubble. In this case,
2122: deflected grains would produce a wide band centered on the plane
2123: perpendicular to the field. The actual field presumably includes a
2124: significant random component, so the resulting distributed feature (if
2125: any) may be more complicated than a wide band.
2126:
2127: \section{\label{sec:AGB} DUST FROM AGB STARS}
2128:
2129: Stars with main sequence mass $M \ltsim 6 M_{\sun}$ spend time on the
2130: asymptotic giant branch (AGB). During this evolutionary phase, a wind is
2131: driven off the star, with mass loss rates as high as $\sim 10^{-5}
2132: M_{\sun} \yr^{-1}$. Grain formation in these outflows likely supplies a
2133: large fraction of the dust in the ISM. See Habing (1996) for a general
2134: review of AGB stars.
2135:
2136: Since the grains spend only a limited time in a high-density environment,
2137: the size distribution of the newly formed dust is thought to be dominated
2138: by grains with $a < 1 \micron$ (see, e.g., Kr\"{u}ger \& Sedlmayr 1997).
2139: Observations confirm the significant presence of submicron grains.
2140: Jura (1996) examined the extinction due to circumstellar dust around seven
2141: O-rich giants and found that it rises towards the ultraviolet.
2142: From observations of scattering, Groenewegen (1997) found a mean grain size
2143: $a \approx 0.16 \micron$ for the circumstellar envelope of the carbon star
2144: IRC +10 216.
2145:
2146: Even if submicron grains dominate the outflows from AGB stars, the
2147: luminosity of grains with $a \gtsim 10 \micron$ may be high enough to result
2148: in detectable fluxes at Earth, since the dust production rate is so high.
2149: Grains with isotopic ratios indicative of formation in AGB stars have
2150: been discovered in meteorites. Although most of these grains have
2151: $a \sim 0.5 \micron$, some are as large as $a \approx 10 \micron$
2152: (Zinner 1998).
2153:
2154: Suppose the grains all have radius $a$ and that a total dust mass $M_d$
2155: is emitted during the AGB lifetime $\tau$. The dust luminosity is then
2156: $L = 3 M_d / (4 \pi \rho a^3 \tau)$. We adopt $M_d \sim 0.01 M_{\sun}$
2157: and the canonical lifetime $\tau \sim 10^5 \yr$ (Habing 1996). The flux
2158: at Earth follows from equation (\ref{eq:specific_flux}). For simplicity,
2159: we take $f_{\rm beam}=1$ and $f_{\rm survive} =1$; also, since the typical
2160: grain ejection speed in an AGB wind is $\sim 10 \kms$ (Habing 1996),
2161: we take $v_{d, \sun}/v_{ej} = 1$. Thus, the flux as a function of AGB star
2162: distance $d$ is approximately
2163: \be
2164: F(d) \sim 114 \yr^{-1} \km^{-2} \left( \frac{M_d}{10^{-2} M_{\sun}} \right)
2165: \left( \frac{\rho}{3.5 \g \cm^{-3}} \right)^{-1} \left( \frac{a}{10 \micron}
2166: \right)^{-3} \left( \frac{\tau}{10^5 \yr} \right)^{-1} \left( \frac{d}
2167: {100 \pc} \right)^{-2}~~~.
2168: \ee
2169:
2170: In order for the grains to be traceable to their point of origin, we require
2171: that the gyroradius $r_B > d$. This yields a minimum acceptable grain size
2172: (eq.~\ref{eq:gyroradius}), so that
2173: \begin{eqnarray}
2174: \label{eq:flux_AGB}
2175: F(d) & \ltsim 8 \yr^{-1} \km^{-2} & \left( \frac{M_d}{10^{-2} M_{\sun}}
2176: \right)
2177: \left( \frac{\rho}{3.5 \g \cm^{-3}} \right)^{1/2} \left( \frac{\Phi}{0.5 \,
2178: {\rm V}} \right)^{-3/2} \left( \frac{B}{5 \mu {\rm G}} \right)^{-3/2} \times
2179: \nonumber \\
2180: & & \left( \frac{v}{10 \kms} \right)^{3/2}
2181: \left( \frac{\tau}{10^5 \yr}
2182: \right)^{-1} \left( \frac{d} {100 \pc} \right)^{-7/2}~~~.
2183: \end{eqnarray}
2184: The above flux is for the optimistic case that all of the dust mass is
2185: in the smallest grain size $a_t$ ($25\micron$ in this case) such that
2186: the grain will barely be undeflected. In cgs units, this flux is
2187: $2.5\times10^{-17}\cm^{-2}\s^{-1}$ tantalizingly close to the flux of
2188: the point source seen by Baggaley (2001).
2189:
2190: Jackson, Ivezi\'{c}, \& Knapp (2002) have studied the distribution of AGB
2191: stars with mass loss rates in the range $10^{-6}$--$10^{-5} M_{\sun}
2192: \yr^{-1}$. They find that the number density of AGB stars
2193: $n_{\rm AGB} \approx 4.4 \times 10^{-7} \pc^{-3}$ in the solar neighborhood.
2194: Thus, the number of sources within a distance $d$ from the Sun is
2195: $N(d) \approx 1.84 (d/100 \pc)^3$.
2196:
2197: In Table \ref{tab:AGB} we give the flux $F$ from equation
2198: (\ref{eq:flux_AGB}) and the number $N$ of sources for a few values of $d$.
2199: We also give the threshold grain size $a_t$ and the fraction $g$ of the
2200: emitted dust mass that would need to be in such large grains in order to
2201: yield an observed flux at Earth of $2 \times 10^{-3} \yr^{-1} \km^{-2}$
2202: (corresponding to 20 events per year with a collecting area of
2203: $10^4 \km^2$). We see from the table that AGB stars could be a significant
2204: source population for radar detection of extra-solar meteoroids if a small
2205: fraction of the grains in the outflow are large.
2206: Note that the source AGB star will not be visible as such, due to
2207: the large distances to AGB stars and their short lifetimes. Also, the
2208: distances are large enough that the grain trajectories will be affected
2209: by the Galactic potential.
2210:
2211: Even if all the grains in the AGB outflow are smaller than $10 \micron$,
2212: there is still a chance to see the AGB star as a source on the sky, since
2213: some AGB stars apparently have long-lived disks in which grain coagulation
2214: occurs. This phenomenon has been observed for binary stars; the presence
2215: of the companion apparently causes some of the outflowing grains to be
2216: deflected into a disk. A well-studied example is AC Her (Jura, Chen, \&
2217: Werner 2000). Jura et al.~find that the grains must have $a \gtsim 200
2218: \micron$ in order to remain gravitationally bound, due to the extreme
2219: luminosity of the evolved star. They estimate the mass of the dust
2220: disk (in grains with $a \ltsim 0.1 \cm$) to be $\approx 1.2 \times
2221: 10^{30} \g$. Their observations of infrared emission
2222: imply the presence of much smaller grains as well. They suggest that
2223: shattering during collisions between disk grains leads to a radiation
2224: pressure-driven outflow of smaller grains. The lifetime of such disks,
2225: and the frequency with which they occur, remain unknown. The large disk
2226: mass suggests the possibility of very large dust fluxes and the commonality
2227: of binaries suggests that the frequency could be high. Other examples of
2228: stars with circumbinary disks are HD 44179 (the central star of the Red
2229: Rectangle nebula; Jura, Turner, \& Balm 1997), IRAS 09425-6040
2230: (Molster et al.~2001), and SS Lep, 3 Pup, and BM Gem (Jura, Webb, \&
2231: Kahane 2001).
2232:
2233: If the AMOR point source is (or was) an AGB star, the star would
2234: currently be a young, hot, high proper motion white dwarf. The proper
2235: motion would point back to the vicinity of the AMOR source, after
2236: allowing for the uncertain aberration of the dust particles.
2237:
2238: \section{\label{sec:YSOs} DUST FROM YOUNG STELLAR OBJECTS}
2239:
2240: The accretion of material onto a forming star is accompanied by a bipolar
2241: outflow of material. These outflows appear as both highly collimated,
2242: supersonic ($v \sim 100$--$200 \kms$) jets and as large-angle molecular
2243: outflows (with $v \ltsim 25 \kms$). See Reipurth \& Bally (2001),
2244: K\"{o}nigl \& Pudritz (2000), Shu et al.~(2000), Eisl\"{o}ffel et
2245: al.~(2000), and Richer et al.~(2000) for reviews.
2246:
2247: The mechanism for launching the outflow is not yet well understood.
2248: If the material originates in the disk and if the density is high enough,
2249: then large grains can be entrained in the outflow, assuming there has been
2250: sufficient time for growth to large sizes. As the density decreases,
2251: the grains decouple from the gas and retain their launching speed, which
2252: could be $\sim 10$--$500 \kms$. Chugai (2001) has appealed to the ejection
2253: of large grains from young stellar objects (YSOs) to explain the early AMOR
2254: detections of interstellar meteors (Taylor, Baggaley, \& Steel
2255: 1996). He finds that the flux supplied from YSOs is a factor of 30
2256: below that claimed by the AMOR group.
2257:
2258: Shocks form when the jets run into the ambient ISM and grains could be
2259: destroyed in the shocked regions (known as Herbig-Haro objects).
2260: Observations have yielded contradictory results on this point.
2261: Beck-Winchatz, B\"{o}hm, \& Noriega-Crespo (1996) and B\"{o}hm \&
2262: Matt (2001) find that the gas-phase Fe abundance in Herbig-Haro objects is
2263: generally close to solar, indicating very efficient grain destruction.
2264: Mouri \& Taniguchi (2000), on the other hand, find the gas-phase Fe
2265: abundance to be $\approx 20\%$ solar, indicating only a modest amount of
2266: grain destruction. Grain destruction is not expected for the slower,
2267: wide-angle outflows. However, these outflows may consist of ambient
2268: material swept up by the jets (K\"{o}nigl \& Pudritz 2000), in which
2269: coagulated grains might not be present. Thus, it is not yet clear
2270: whether or not YSO outflows contain large grains.
2271:
2272: We can make an optimistic estimate of the dust luminosity by assuming
2273: that $\sim 1\%$ of the outflowing mass is in large grains, as we did for
2274: AGB stars. The mass loss rate can be as high as $\sim 10^{-6} M_{\sun}
2275: \yr^{-1}$ (Eisl\"{o}ffel et al.~2000), yielding a flux
2276: %
2277: \be %$
2278: F(d)\approx2.8\times 10^{-18}
2279: \left(100\pc\over d\right)^2{v_{d\odot}\over v_{ej}}
2280: \cm^{-2}\s^{-1},
2281: \ee %$
2282: %
2283: where we have assumed $a=25\micron$. For ejection in a jet, this is
2284: $\approx 100$ times smaller than the optimistic estimate for an AGB
2285: star at the same distance. However, the space density of YSO's is
2286: larger than that of AGB stars; TW Hydrae, a T Tauri star, is only
2287: $59\pc$ from the sun, and is associated with dozens of other young
2288: stars at similar distances. There is good evidence that TW Hydrae has
2289: a $\sim100\km\s^{-1}$ wind
2290: \citet{2002ApJ...572..310H,2000ApJ...534L.101W}. TW
2291: Hydrae may be associated with Gould's Belt \citet{2001A&A...368..866M}.
2292: While the (optimistic) flux
2293: estimate given here is lower by a factor of order $100$ than is
2294: realistically detectable by AMOR, TW Hydrae and stars associated with
2295: it might be seen by more sensitive future radar systems.
2296:
2297: The outflow may last $\sim 10^6 \yr$, 10 times longer than the AGB
2298: star lifetime. In the following stage of
2299: evolution, the YSO disk is dispersed, perhaps by photoevaporation. In
2300: this phase, the mass loss rate could be as high as $\sim 10^{-7}
2301: M_{\sun} \yr^{-1}$ (Hollenbach, Yorke, \& Johnstone 2000), yielding a
2302: maximum dust luminosity 10 times lower than for the bipolar outflows.
2303:
2304: Given the potentially large dust luminosities, YSOs could be likely
2305: candidate sources for extrasolar meteoroids. However, the young stars
2306: listed in Table \ref{tab:Gliese} are probably too old, given their
2307: proximity. For example, if disk dispersal is complete at an age of
2308: $3 \Myr$, then dust from the YSO phase of $\beta \,$Pic has been traveling
2309: for $\approx 9 \Myr$. At a speed of $10 \kms$, the dust would now be
2310: $\approx 90 \pc$ from $\beta \,$Pic, but the Sun is only $\approx 20 \pc$
2311: from $\beta \,$Pic. If, however, disk dispersal continues until an age
2312: of $10 \Myr$, then $\beta \,$Pic may appear as a YSO dust source today.
2313:
2314: Some nearby, young clusters could produce large total fluxes distributed
2315: over a wide area on the sky. Such clusters include TW Hydrae ($d \sim 55
2316: \pc$, $t \sim 10 \Myr$; Zuckerman \& Webb 2000), Upper Scorpius
2317: ($d \sim 145 \pc$, $t \sim \,$1--10$\Myr$; Spangler et al.~2001),
2318: Chamaeleon Ia and Ib ($d \sim \,$140--150$\pc$, $t \sim \,$1--40$\Myr$;
2319: Spangler et al.~2001), and Taurus ($d \sim 140 \pc$, $t \sim \,$10--40$\Myr$;
2320: Spangler et al.~2001).
2321:
2322: \section{DISCUSSION\label{sec:Discussion}}
2323: Our calculations suggest that there are a number of nearby sources of
2324: $10\micron$ or larger particles that yield sufficient fluxes to be
2325: detectable by ground based radar systems. Here we suggest ways to
2326: optimize radar systems for meteor detection.
2327:
2328: Meteoroids with $a\lesssim 10\micron$ cannot travel through the ISM for
2329: appreciable distances. Furthermore, it appears that the flux of meteoroids
2330: decreases with increasing meteoroid size. It follows that if one is interested
2331: in detecting particles from an identifiable source, large radar
2332: collecting area is more crucial than large radar power, once the power
2333: is sufficient to detect $a\approx10\micron$ particles.
2334:
2335: The returned radar power falls off as $1/R^3$ for radars that detect
2336: coherent emission from a substantial fraction of the meteor trail
2337: (such as AMOR), or as $1/R^4$ for meteor head detectors such as
2338: Arecibo. The minimum range to the meteor trail is given by the height
2339: of the trails (of order $100\km$), while the maximum range is of order
2340: $1000\km$ due to the curvature of the Earth. Increasing the mean
2341: range by a factor of 10 will increase the collecting area, and hence
2342: the flux, by a factor of 100, but requires an increase in radar power
2343: by a factor of $10^3$ to $10^4$, depending on the type of radar
2344: employed. While the cost of a high power radar transmitter increases
2345: rapidly with increasing power, it is clearly helpful to maximize the
2346: transmitted power.
2347:
2348: Another way to increase the radar range, and hence the collecting
2349: area, is to increase the gain of the antenna; the received power
2350: scales as the product $G_TG_R$. However, there are limits to the
2351: extent to which one can increase the antenna gain. If the radar is
2352: like AMOR, the width of the beam must exceed the length of a typical
2353: meteor trail, else the number of coherently emitting electrons will
2354: drop. The angular width of the trail is of order $H_p/R$, which ranges
2355: from $0.06$ to $0.006$ radians, or $3^\circ$ (in RA, if the beam looks
2356: due south or north) at the zenith to $0.3^\circ$ at the
2357: horizon. Optimally, the beam will cover the sky from the zenith to the
2358: horizon, so the antenna gain should be no larger than $\sim 1000$.
2359:
2360: Selecting a gain of this magnitude matches the beam size to the length
2361: of the meteor trail, but limits the geometric area $A_G$ of the
2362: antenna, partially defeating the purpose, which is to maximize the
2363: collecting area. This is the case with AMOR, where the choice of a
2364: narrow beam was dictated by the need for high precision measurement of
2365: the meteor position, but at the cost of collecting area. To get around
2366: this problem, we propose building an array of antennas, all powered by
2367: the same radio generator. The beams of the array would all have widths
2368: of order $3^\circ$, but the beams would be directed around the points
2369: of the compass. In principle one could have $\sim100$ such beams
2370: emanating from a single facility. The duration of a radar pulse is
2371: typically a microsecond, while the time between pulses is of order a
2372: millisecond. (The time to travel out and back $1000\km$ is about 6
2373: milliseconds; in that time the meteoroid will travel a distance
2374: $\sim240\m$.) One could send out pulses, directed at different points
2375: around the compass, separated by 10 milliseconds. The receiver would
2376: be turned off when the transmitter is on, about $10^{-4}$ of the
2377: time. Return pulses from trails at different compass points would
2378: overlap in time, but given sufficient receiver gains, cross talk
2379: between the receivers should be small enough to be acceptable.
2380:
2381: If one employs a sufficiently powerful radar, one can live with a low
2382: antenna gain and still achieve large collecting area. However, the need for
2383: high precision determinations of the meteor positions remains. This
2384: problem can be addressed by using separate transmitting and receiving
2385: antennae. The transmitter can have a broad beam, with a low gain but
2386: covering a large fraction of the sky. An antenna array can then be used as
2387: an interferometer to locate the meteor on the sky.
2388:
2389: Either a multi-beam radar or single broad beam radar with
2390: an interferometer could have a collecting area
2391: $A_{col}$ of order $800\km^2$, if the typical range were $R\sim
2392: 150\km$. Current distant early warning (DEW) radar may be interesting
2393: in this context. The existence of tens of such systems, with an
2394: aggregate collecting area of $\sim10,000\km^2$ only adds to their
2395: attractiveness as extrasolar meteor detectors.
2396:
2397: The collecting area grows dramatically, up to $35,000\km^2$,
2398: if the radar can detect $10\micron$ particles out to the
2399: horizon. This is $4000$ times the collecting area of the AMOR
2400: radar. AMOR sees about 10 extrasolar meteors every day; the proposed
2401: detector would see up to $40,000$ extrasolar meteoroids of size
2402: $35\micron$ a day, and about $10^6$ $10\micron$ extrasolar
2403: meteoroids per day. This suggests that one of the major constraints is
2404: handling the data, since there would be of order $10^9$ interplanetary
2405: dust particles detected per day.
2406:
2407: Such a radar would be capable of detecting meteoroids from a handful of
2408: nearby debris disks; table \ref{tab:Gliese} lists four nearby young
2409: Gliese stars that might have fluxes exceeding $6\times10^{-4}$
2410: particles per year per square kilometer. This large area radar system
2411: would also detect meteoroids from (former) AGB stars with fluxes as low as
2412: $2\times10^{-11}\cm^{-2}\s^{-1}$; in the optimistic case that a
2413: substantial fraction of emitted particles have $a\gtsim25\micron$,
2414: sources as distant as a kiloparsec could be seen. From table
2415: \ref{tab:AGB} this could be several hundred stars. The number of
2416: possible YSO sources could be somewhere between these two estimates.
2417:
2418: How feasible is such a radar? We have already mentioned the current DEW
2419: systems. However, we note that moderate enhancements to an
2420: AMOR type radar are all that are needed. The product of the
2421: transmitter and receiver gain could be enhanced by a factor of $5-10$,
2422: while the power could be enhanced by a factor of 20, to $2$ megawatts,
2423: for a total increase in received power of a factor of $\sim200$. This
2424: would allow for the detection of $a=10\micron$ size particles at
2425: ranges of $\sim500\km$. Further progress could be made by increasing
2426: the operating wavelength; a factor of two increase, to
2427: $\lambda=2000\cm$ would be sufficient to detect $10\micron$ size
2428: particles out to the horizon.
2429:
2430: The flux of interstellar meteoroids of different mass appears to follow a
2431: power law $mf_m\propto m^{-1.1}$, over five decades in mass. This may
2432: be compared to the Dohnanyi law, $f_m\propto m^{-\alpha}$, with
2433: $\alpha=11/6\approx 1.83$; in the case of interstellar meteoroids,
2434: $\alpha\approx2.1$. We note that the observed size distribution of
2435: asteroids has $\alpha\approx2$ for objects larger than $a=2.5\km$, and
2436: $\alpha\approx 1.4$ for $a$ between $2.5$ and $0.1\km$
2437: \cite{Ivezic}. However, the binding energy per gram can be expected to
2438: vary with size, a fact we appealed to in explaining the observation
2439: that meteoroids of very difference sizes ablate at the same height in the
2440: Earth's atmosphere. In our view, the rough agreement between the
2441: values of $\alpha$ might reflect a common origin in a collisional
2442: cascade, but is more likely a coincidence.
2443:
2444: Krivova \& Solanki \citep{2003A&A...402L...5K} recently suggested that
2445: interactions with a Jupiter mass planet could eject dust grains with
2446: velocities of order $10-70\km\s^{-1}$. They cited this result as
2447: support for the identification of the point source found by Baggaley
2448: (2000) with $\beta\,$ Pic. Our analytic calculations show that only a
2449: tiny fraction of ejected particles will have such high velocities, a
2450: result supported by direct numerical integrations.
2451:
2452:
2453: \section{CONCLUSIONS\label{sec:conclusions}}
2454:
2455: We have provided simple analytic estimates for the minimum size of
2456: radar detected meteoroids as a function of radar power and antenna
2457: gain. We give very rough estimates of the collecting area of radar
2458: detectors, and hence of fluxes of interstellar meteoroids, assuming that
2459: the claimed detection rates are correct. The fluxes of satellite and radar
2460: detected interstellar meteoroids appear to lie along a power law,
2461: $mf_m\propto m^{-1.1}$, corresponding to a $\alpha=2.1$, or a size
2462: scaling $\gamma=4.3$, in contrast to the Dohnanyi value $\gamma=3.5$.
2463:
2464: We examine three possible sources of large, $a>10\micron$ size
2465: extrasolar meteoroids. In descending order of dust luminosity, they are
2466: AGB stars, young stellar objects (with winds and/or jets), and debris
2467: disks. (Supernovae can supply similar mass fluxes of particles, but we
2468: expect the particles to be smaller.) We show that such large particles
2469: can travel in straight lines for tens of parsecs through the
2470: interstellar medium, which in principle will allow their sources to be
2471: traced.
2472:
2473: Since there are $\sim2$ AGB stars within 100pc of the sun at any time,
2474: we expect that there will be $\sim2$ sources of extrasolar meteoroids,
2475: with fluxes at Earth approaching $\sim8\yr^{-1}\km^{-2}$, on the sky
2476: at all times. Such a star might be responsible for the ``point
2477: source'' reported in Baggaley (2000). The AMOR radar, according to our
2478: estimates, is marginally capable of detecting such a small flux. The
2479: star itself would currently be a hot white dwarf with a high proper
2480: motion, within $\sim100\pc$ of the sun.
2481:
2482: There is one known T Tauri star within $100\pc$ of the sun, TW
2483: Hydrae. This star has a high velocity outflow ($\sim100\km\s^{-1}$)
2484: from a dusty disk. It is very likely expelling $10$ to $100\micron$
2485: size particles; unfortunately, the flux is clearly too small, by a
2486: factor of $\sim100$, to be detected by current radar systems.
2487:
2488: Similarly, currently known debris disk systems provide fluxes that are
2489: too small to be detected at present. We have shown that the system
2490: implicated by Baggaley (2000), $\beta$ Pic, is unlikely to provide a
2491: particle flux as large as that associated with his point source. We
2492: also show that the location of the apparent source on the sky is
2493: inconsistent with particles ejected from $\beta$ Pic by gravitational
2494: interaction with a Jupiter mass planet in the system; we showed that
2495: typical ejection velocities in that case are of order
2496: $1\km\s^{-1}$. Even if the ejection velocity is assumed to be
2497: $30\km\s^{-1}$, the most favorable value, the apparent position of the
2498: source does not match the observed position (see
2499: Fig. \ref{fig:baggaley}). Since the particle flux, ejection velocity,
2500: and position on the sky do not match the observations, we conclude
2501: that the point source is not associated with gravitational ejection
2502: from the debris disk of $\beta$ Pic.
2503:
2504: However, we would like to stress that all three types of sources, AGB
2505: stars, YSOs, and debris disks, should be detectable by future ground based
2506: radar systems. The pioneering AMOR system may have already detected
2507: meteoroids from an AGB star. Modest improvements in this type of ground
2508: based radar system, in particular better sky coverage, should allow
2509: for the detection of multiple examples of each type of source. An
2510: alternate possibility is to piggyback on already existing radar
2511: employed in distant early warning systems.
2512:
2513:
2514: \acknowledgements
2515: We are grateful to C.~D.~Matzner, D.~N.~Spergel, D.~D. Meisel, W.~J.~Baggaley,
2516: B.~T.~Draine, and R.~E.~Pudritz for helpful discussions. This research
2517: was supported by NSERC of Canada and has made use of the SIMBAD
2518: database, operated at CDS, Strasbourg, France, and of NASA's
2519: Astrophysics Data System.
2520:
2521: \appendix
2522: \section{\bf APPENDIX}
2523: In this appendix we rederive some of the relations used in section
2524: \ref{sec:detectable_fluxes}, employing the classical theory of
2525: meteors. In the meteoroid frame, the air flow has a kinetic energy flux
2526: %
2527: \be %$
2528: F={1\over2}\rho_av^3.
2529: \ee %$
2530: %
2531: Following the meteoritics literature, we define a shape factor $A$ to
2532: be the ratio of the area $A_m$ of the meteoroid divided by the two-thirds
2533: power of the volume $(m/\rho)$, or
2534: %
2535: \be %$
2536: A\equiv{A_m\over(m/\rho)^{2/3}}.
2537: \ee %$
2538: %
2539: For a sphere, $A=(9\pi/16)^{1/3}\approx 1.2$. The kinetic luminosity
2540: seen by the meteoroid is then $A_m F= {A\over2} (m/\rho)^{2/3}\rho_av^3$; assume
2541: that a fraction $\Lambda$ of this kinetic
2542: luminosity goes toward ablating the meteor. Then the time rate
2543: of change of the binding energy is
2544: %
2545: \be %$
2546: {dE\over dt}=-{\Lambda A\over2}\left({m\over\rho}\right)^{2/3}\cdot\rho_av^3.
2547: \ee %$
2548: %
2549: Typical estimates for $\Lambda$ are around $0.5$.
2550: It follows that the rate of ablation is
2551: %
2552: \be \label{ablation}%$
2553: {dm\over dt}=-{\Lambda A\over 2\zeta}
2554: \left({m\over \rho}\right)^{2/3}\rho_av^3,
2555: \ee %$
2556: %
2557: where, the reader will recall, the binding energy $E$ is related to
2558: the meteoritic mass $m$ by the heat of ablation $\zeta$, $E=\zeta m$.
2559: From equation (\ref{q}), the number of ions produced per
2560: centimeter along the path is
2561: %
2562: \be %$
2563: q=-{\beta\over v\mu}{dm\over dt},
2564: \ee %$
2565: %
2566: where $\mu$ is the mean molecular weight of the meteoroid (recall that
2567: $\beta$ is the number of ions produced by each meteoroid atom). Combining
2568: this with equation (\ref{ablation}), we find
2569: %
2570: \be %$
2571: q=\beta{\Lambda A\over 2\zeta\mu}
2572: \left({m\over \rho}\right)^{2/3}\rho_av^2.
2573: \ee %$
2574: %
2575: This is known as the ionization equation.
2576:
2577: The maximum line density along the track can be found by
2578: differentiating the ionization equation with respect to time:
2579: %
2580: \be \label{dqdt}%$
2581: {1\over q}{dq\over dt}={1\over\beta}{d\beta(v)\over
2582: dt}+{2\over3m}{dm\over dt}+{1\over\rho_a}{d\rho_a\over dt}+ {2\over
2583: v}{dv\over dt}.
2584: \ee %$
2585: %
2586: Note that $\beta(v)\sim v^n$, so that both the first and last terms on
2587: the right hand side are proportional to
2588: %
2589: \be %$
2590: {1\over v}{dv\over dt}
2591: \ee %$
2592: %
2593: We assume for the moment that this is smaller than either the
2594: mass or density variations. Then the maximum value of $q$ occurs when
2595: %
2596: \be %$
2597: {2\over 3m}{dm\over dt}=-{1\over \rho_a}{d\rho_a\over dt},
2598: \ee %$
2599: %
2600: or
2601: %
2602: \be %$
2603: {2\over3m}(-{q_{max}\mu\over \beta}v)=-{v\over H_p},
2604: \ee %$
2605: %
2606: where $H_p$ is the density scale height. Thus
2607: %
2608: \be %$
2609: q_{max}={3\over2}{\beta m\over\mu H_p}.
2610: \ee %$
2611: %
2612: In words, the maximum line density is given by spreading the meteoroid over
2613: approximately a scale height and accounting for the number $\beta$ of
2614: atoms that are ionized for each atom in the meteor.
2615:
2616:
2617: Next we justify the neglect of terms proportional to the derivative of
2618: the velocity. The momentum equation for the meteoroid is
2619: %
2620: \be %$
2621: m{dv\over dt}=mg-C_DA\left({m\over\rho}\right)^{2/3}\rho_av^2,
2622: \ee %$
2623: %
2624: where $C_D$ is the drag coefficient. The mean free path at meteor
2625: heights is of order $10\cm$, much larger than the typical radius of
2626: the meteoroids we are interested in. In that case $C_D=1$. The velocity
2627: derivative becomes
2628: %
2629: \be \label{dvdt}%$
2630: {1\over v}{dv\over dt}={g\over v}-A\left({m\over\rho}\right)^{2/3}\rho_av/m
2631: \ee %$
2632: %
2633: We note that $H_p\approx c^2/g$ where $c$ is the sound speed; then
2634: %
2635: \be %$
2636: {v\over H_p}\approx\left({v\over c}\right)^2{g\over v}>>{g\over v}
2637: \ee %$
2638: %
2639: so that the first term on the right hand side of (\ref{dvdt}) is
2640: negligible compared to $v/H_p$; recall that the meteoroids are highly
2641: supersonic. Next we compare the velocity gradient in (\ref{dqdt})
2642: (including only the second term on the rhs of
2643: (\ref{dvdt})),
2644: %
2645: \be %$
2646: {2+n\over v}{dv\over dt}\approx A\left({m\over\rho}\right)^{2/3}
2647: {\rho_av\over m}\sim a^2\rho_av/\rho
2648: a^3\sim{\rho_a\over \rho}{H_p\over a}{v\over H_p}
2649: \ee %$
2650: %
2651: to
2652: %
2653: \be %$
2654: {2\over 3m}{dm\over dt}=-{\Lambda A\over2\zeta}
2655: \left({m\over\rho}\right)^{2/3}
2656: {\rho_a v^3\over m}.
2657: \ee %$
2658: %
2659: The ratio is
2660: %
2661: \be %$
2662: {12\over \Lambda}{\zeta\over v^2}
2663: ={3\over4}\left({\zeta\over10^{11}\cm^2\s^{-2}}\right)
2664: \left({v\over 40\km\s^{-1}}\right)^2
2665: \left({10^{-1}\over\Lambda}\right),
2666: \ee %$
2667: %
2668: where we took $n=2$.
2669:
2670: McKinley gives estimates for $\Lambda$ that are of order unity. For
2671: these values, the velocity derivative is much smaller than the mass
2672: derivative. We conclude that the logarithmic mass derivative is
2673: comparable to or larger than the logarithmic velocity derivative.
2674:
2675:
2676:
2677: \begin{thebibliography}{}
2678:
2679: \bibitem[]{A84} Aumann, H. H. et al.~1984, \apj, 278, L23
2680: \bibitem[]{BP93} Backman, D. E. \& Paresce, F. 1993, in Protostars and
2681: Planets III, eds. E. H. Levy \& J. I. Lunine (Tucson: Univ.~of Arizona
2682: Press), 1253
2683: \bibitem[Baggaley(2000)]{Bag00} Baggaley, W. J. 2000, \jgr, 105, 10353
2684: \bibitem[]{BBST94} Baggaley, W. J., Bennett, R. G. T., Steel, D. I., \&
2685: Taylor, A. D. 1994, \qjras, 35, 293
2686: \bibitem[]{BBN96} Beck-Winchatz, B., B\"{o}hm, K.-H., \& Noriega-Crespo,
2687: A. 1996, \aj, 111, 346
2688: \bibitem[]{BT87} Binney, J. \& Tremaine, S. 1987, Galactic Dynamics
2689: (Princeton: Princeton U. Press)
2690: \bibitem[]{BM01} B\"{o}hm, K.-H. \& Matt, S. 2001, \pasp, 113, 158
2691: \bibitem[Bronshten 1983]{Bronshten} Bronshten, V. A. Physics of
2692: Meteoric Phenomena (Dordrecht: Ridel) 1983
2693: \bibitem[Campbell-Brown \& Jones(2003)]{2003MNRAS.343..775C}
2694: Campbell-Brown, M.~\& Jones, J.\ 2003, \mnras, 343, 775
2695: \bibitem[]{C01} Chugai, N. N. 2001, Solar System Research, 35, 307
2696: \bibitem[]{CVLFG97} Crifo, F., Vidal-Madjar, A., Lallement, R., Ferlet,
2697: R., \& Gerbaldi, M. 1997, \aap, 320, L29
2698: \bibitem[]{DB98} Dehnen, W. \& Binney, J. J. 1998, \mnras, 298, 387
2699: \bibitem[]{D69} Dohnanyi, J. S. 1969, \jgr, 74, 2531
2700: \bibitem[]{DS79} Draine, B. T. \& Salpeter, E. E. 1979, \apj, 231, 77
2701: \bibitem[]{dZ01} de Zeeuw, P. T., Hoogerwerf, R., de Bruijne, J. H. J.,
2702: Brown, A. G. A., \& Blaauw, A. 2001, Encyclopedia of Astronomy and
2703: Astrophysics (Bristol: Institute of Physics Publishing)
2704: \bibitem[]{EMRR00} Eisl\"{o}ffel, J., Mundt, R., Ray, T. P., \&
2705: Rodr\'{i}guez, L. F. 2000, in Protostars and Planets IV, eds.~V. Mannings,
2706: A. P. Boss, \& S. S. Russell (Tucson: Univ.~of Arizona Press), 815
2707: \bibitem[]{FN} Fowler, R.H. \& Nordheim, L.~1928,
2708: Proc. Roy. Soc. (London) A, 119, 173
2709: \bibitem[]{F99} Frisch, P. C. et al.~1999, \apj, 525, 492
2710: \bibitem[]{G86} Gillett, F. C. 1986, in Light on Dark Matter, ed.~F. P.
2711: Israel (Dordrecht: D. Reidel), 61
2712: \bibitem[Gilmore \& Reid(1983)]{1983MNRAS.202.1025G} Gilmore, G.~\& Reid,
2713: N.\ 1983, \mnras, 202, 1025
2714: \bibitem[]{GSSMN98} Guillout, P., Sterzik, M. F., Schmitt, J. H. M. M.,
2715: Motch, C., \& Neuh\"{a}user, R. 1998, \aap, 337, 113
2716: \bibitem[]{G97} Groenewegen, M. A. T. 1997, \aap, 317, 503
2717: %\bibitem[]{grun94} Grun, E., Gustafson, B., Man, I., Baguhl, M.,
2718: Morfill, G.E., Staubach, P., Taylor, A. \& Zook, H.A. 1994, \aap, 286, 915
2719: \bibitem[]{H96} Habing, H. J. 1996, A\&AR, 7, 97
2720: \bibitem[]{H98} Heiles, C. 1998, in The Local Bubble and Beyond, eds.
2721: D. Breitschwerdt, M. J. Freyberg, \& J. Tr\"{u}mper (Berlin: Springer-Verlag)
2722: \bibitem[Herczeg et al. (2002)]{2002ApJ...572..310H} Herczeg, G.~J., Linsky,
2723: J.~L., Valenti, J.~A., Johns-Krull, C.~M., \& Wood, B.~E.\ 2002, \apj, 572,
2724: 310
2725: \bibitem[]{Hetc98} Holland, W. S., et al.~1998, Nature, 392, 788
2726: \bibitem[]{HYJ00} Hollenbach, D. J., Yorke, H. W., \& Johnstone, D. 2000,
2727: in Protostars and Planets IV, eds.~V. Mannings, A. P. Boss, \& S. S. Russell
2728: (Tucson: Univ.~of Arizona Press), 401
2729: \bibitem[Ivezic et al. 2001]{Ivezic} Ivezi{\' c}, {\v Z}.~et al.\
2730: 2001, \aj, 122, 2749
2731: \bibitem[Jackson, Ivezi{\' c}, \& Knapp(2002)]{2002MNRAS.337..749J}
2732: Jackson, T., Ivezi{\' c}, {\v Z}., \& Knapp, G.~R.\ 2002, \mnras, 337, 749
2733: \bibitem[Janches et al. 2000]{JMMZ} Janches, D., Mathews, J. D.,
2734: Meisel, D. D., \& Zhou, Q.-H. 2000, Icarus, 145, 53
2735: \bibitem[Jones(1997)]{1997MNRAS.288..995J} Jones, W.\ 1997, \mnras, 288,
2736: 995
2737: \bibitem[Jones \& Halliday(2001)]{2001MNRAS.320..417J} Jones, W.~\&
2738: Halliday, I.\ 2001, \mnras, 320, 417
2739: \bibitem[]{JTH96} Jones, A. P., Tielens, A. G. G. M., \& Hollenbach,
2740: D. J. 1996, \apj, 469, 740
2741: \bibitem[]{J96} Jura, M. 1996, \apj, 472, 806
2742: \bibitem[]{JTB97} Jura, M., Turner, J., \& Balm, S. P. 1997, \apj, 474, 741
2743: \bibitem[]{JCW00} Jura, M., Chen, C., \& Werner, M. W. 2000, \apj, 541, 264
2744: \bibitem[]{JWK01} Jura, M., Webb, R. A., \& Kahane, C. 2001, \apj, 550, L71
2745: \bibitem[Kim \& Martin(1995)]{1995ApJ...444..293K} Kim, S.~\& Martin,
2746: P.~G.\ 1995, \apj, 444, 293
2747: \bibitem[]{KP00} K\"{o}nigl, A. \& Pudritz, R. E. 2000, in Protostars and
2748: Planets IV, eds.~V. Mannings, A. P. Boss, \& S. S. Russell (Tucson:
2749: Univ.~of Arizona Press), 759
2750: \bibitem[Krivova \& Solanki(2003)]{2003A&A...402L...5K} Krivova, N.~A.~\&
2751: Solanki, S.~K.\ 2003, \aap, 402, L5
2752: \bibitem[]{KS97} Kr\"{u}ger, D. \& Sedlmayr, E. 1997, \aap, 321, 557
2753: \bibitem[]{LBA00} Lagrange, A.-M., Backman, D. E., \& Artymowicz, P. 2000,
2754: in Protostars and Planets IV, eds. V. Mannings, A. P. Boss, \& S. S.
2755: Russell (Tucson: Univ.~of Arizona Press), 639
2756: \bibitem[Landgraf et al. 2000]{2000JGR...10510343L} Landgraf, M.,
2757: Baggaley, W.~J., Gr{\" u}n, E., Kr{\" u}ger, H., \& Linkert, G.\ 2000,
2758: \jgr, 105, 10343
2759: \bibitem[Levison \& Duncan(1994)]{1994Icar..108...18L} Levison, H.~F.~\&
2760: Duncan, M.~J.\ 1994, Icarus, 108, 18
2761: \bibitem[]{LD01} Li, A. \& Draine, B. T. 2001, \apj, 554, 778
2762: \bibitem[]{LG98} Li, A. \& Greenberg, J. M. 1998, \aap, 331, 291
2763: \bibitem[Makarov \& Fabricius(2001)]{2001A&A...368..866M} Makarov, V.~V.~\&
2764: Fabricius, C.\ 2001, \aap, 368, 866
2765: \bibitem[Manning 1958]{Manning} Manning, L. A. 1958, JGR, 63 181
2766: \bibitem[Mathews et al. 1997]{MMHGZ97} Mathews, J.D., Meisel, D.D.,
2767: Hunter, K.P., Getman, V.S. \& Zhou, Q. 1997, Icarus, 126, 157
2768: \bibitem[Meisel, Janches, \& Mathews(2002)]{2002ApJ...567..323M} Meisel,
2769: D.~D., Janches, D., \& Mathews, J.~D.\ 2002, \apj, 567, 323
2770: \bibitem[]{MMP83} Mathis, J. S., Mezger, P. G., \& Panagia, N. 1983, \aap,
2771: 128, 212
2772: \bibitem[Mathis, Rumpl, \& Nordsieck(1977)]{1977ApJ...217..425M} Mathis,
2773: J.~S., Rumpl, W., \& Nordsieck, K.~H.\ 1977, \apj, 217, 425
2774: \bibitem[McKinley(1961)]{McKinley61}McKinley, D. W. R. 1961, Meteor
2775: Science and Engineering, (New York: McGraw-Hill)
2776: \bibitem[]{MJM02} Meisel, D. D., Janches, D. \& Mathews, J. D. 2002,
2777: \apj, 579, 895
2778: \bibitem[]{MMP82} Mezger, P. G., Mathis, J. S., \& Panagia, N. 1982, \aap,
2779: 105, 372
2780: \bibitem[]{MYWNKJL01} Molster, F. J., Yamamura, I., Waters, L. B. F. M.,
2781: Nyman, L.-\AA., K\"{a}ufl, H.-U., de Jong, T., \& Loup, C. 2001, \aap,
2782: 366, 923
2783: \bibitem[]{MT00} Mouri, H. \& Taniguchi, Y. 2000, \apj, 534, L63
2784: \bibitem[Murray \& Dermott(1999)]{MD99}Murray, C.~D. \& Dermott,
2785: S.~F. 1999, Solar system dynamics (Cambridge: Cambridge University Press)
2786: \bibitem[]{P97} P\"{o}ppel, W. 1997, Fundamentals of Cosmic Physics, 18, 1
2787: \bibitem[Opik(1976)]{Opik76}\"Opik, E.~J. 1976, Interplanetary
2788: Encounters: Close-range Gravitational Interactions (New York: Elsvier)
2789: \bibitem[]{RB01} Reipurth, B. \& Bally, J. 2001, \araa, 39, 403
2790: \bibitem[]{RSCBC00} Richer, J. S., Shepherd, D. S., Cabrit, S., Bachiller,
2791: R., \& Churchwell, E. 2000, in Protostars and Planets IV, eds.~V. Mannings,
2792: A. P. Boss, \& S. S. Russell (Tucson: Univ.~of Arizona Press), 867
2793: \bibitem[]{RAMT99} Robichon, N., Arenou, F., Mermilliod, J.-C., \& Turon,
2794: C. 1999, \aap, 345, 471
2795: \bibitem[]{S99} Schneider, G., et al.~1999, \apj, 513, L127
2796: \bibitem[Scholl, Roques, \& Sicardy(1993)]{1993CeMDA..56..381S} Scholl, H.,
2797: Roques, F., \& Sicardy, B.\ 1993, Celestial Mechanics and Dynamical
2798: Astronomy, 56, 381
2799: \bibitem[]{SLCW99} Sfeir, D. M., Lallement, R., Crifo, F., \& Welsh, B. Y.
2800: 1999, \aap, 346, 785
2801: \bibitem[]{SNSL00} Shu, F. H., Najita, J. R., Shang, H., \& Li, Z.-Y.
2802: 2000, in Protostars and Planets IV, eds.~V. Mannings, A. P. Boss, \& S. S.
2803: Russell (Tucson: Univ.~of Arizona Press), 789
2804: \bibitem[]{ST84} Smith, B. A. \& Terrile, R. J. 1984, Science, 226, 1421
2805: \bibitem[]{SSSBZ01} Spangler, C., Sargent, A. I., Silverstone, M. D.,
2806: Becklin, E. E., \& Zuckerman, B. 2001, \apj, 555, 932
2807: \bibitem[]{SN00} Stelzer, B. \& Neuh\"{a}user, R. 2000, \aap, 361, 581
2808: \bibitem[Taylor, Baggaley, \& Steel(1996)]{TBS96} Taylor, A. D.,
2809: Baggaley, W. J., \& Steel, D. I. 1996, Nature, 380, 323
2810: \bibitem[]{TMSH94} Tielens, A. G. G. M., McKee, C. F., Seab, C. G., \&
2811: Hollenbach, D. J. 1994, \apj, 431, 321
2812: \bibitem[]{TFF00} Torra, J., Fern\'{a}ndez, D., \& Figueras, F. 2000,
2813: \aap, 359, 82
2814: \bibitem[]{WD01a} Weingartner, J. C. \& Draine, B. T. 2001a, \apj, 548, 296
2815: \bibitem[]{WD01b} Weingartner, J. C. \& Draine, B. T. 2001b, \apj, 553, 581
2816: \bibitem[]{WD01c} Weingartner, J. C. \& Draine, B. T. 2001c, \apjs, 134, 263
2817: \bibitem[]{W92} Whittet, D. C. B. 1992, Dust in the Galactic Environment
2818: (Bristol: Institute of Physics Publishing)
2819: \bibitem[Wilner, Ho, Kastner, \&
2820: Rodr{\'{\i}}guez(2000)]{2000ApJ...534L.101W} Wilner, D.~J., Ho, P.~T.~P.,
2821: Kastner, J.~H., \& Rodr{\'{\i}}guez, L.~F.\ 2000, \apjl, 534, L101
2822: \bibitem[]{WHKH02} Wilner, D.J., Holman, M.J., Kuchner, M.J. \& Ho,
2823: P.T.P. 2002, \apj, 569, L115
2824: \bibitem[Wisdom \& Holman(1991)]{1991AJ....102.1528W} Wisdom, J.~\& Holman,
2825: M.\ 1991, \aj, 102, 1528
2826: \bibitem[]{Zin98} Zinner, E. 1998, Ann Rev Earth Planet Sci, 26, 147
2827: \bibitem[]{Z01} Zuckerman, B. 2001, \araa, 39, 549
2828: \bibitem[]{ZB93} Zuckerman, B. \& Becklin, E. E. 1993, \apj, 414, 793
2829: \bibitem[]{ZSBW01} Zuckerman, B., Song, I., Bessell, M. S., \& Webb, R. A.
2830: 2001, \apj, 562, L87
2831: \bibitem[]{ZSW01} Zuckerman, B., Song, I., \& Webb, R. A. 2001, \apj, 559,
2832: 388
2833: \bibitem[]{ZW00} Zuckerman, B. \& Webb, R. A. 2000, \apj, 535, 959
2834:
2835: \end{thebibliography}
2836:
2837: \begin{figure}
2838: \epsscale{1.00}
2839: %\plotone{Fig_flux.eps}
2840: \plotone{figure1.ps}
2841: \caption{ Measured cumulative fluxes of extrasolar meteoroids. The data
2842: are from the Ulysses and Galileo satellites (crosses joined by a solid
2843: line), the Arecibo radar (solid squares) and the AMOR radar (solid
2844: triangle). The open triangle represent the ``point source'' seen by
2845: the AMOR radar. The dashed line represents the $mf_m\propto m^{-1.1}$
2846: scaling first noted by Landgraf et al. (2000) and given in equation
2847: (\ref{eq:observed flux}). The upper axis gives the particle radius $a$
2848: in microns, assuming that the meteoroids have a density of $3\g\cm^{-3}$.
2849: \label{fig:flux}
2850: }
2851: \end{figure}
2852:
2853: \begin{figure}
2854: \epsscale{1.00}
2855: %\plotone{Fig_mass_flux.eps}
2856: \plotone{figure2.ps}
2857: \caption{ Measured differential mass fluxes, $m\cdot mf_m$, with units
2858: of $\g\cm^{-3}\s^{-1}$, of extrasolar micrometeoroids. The
2859: long-dashed line is the Landgraf et al. (2000) scaling, while the
2860: dotted line is the limit found by assuming that half of all the
2861: metals in the local ISM (with $n=0.1\cm^{-3}$) are in the form of
2862: micrometeoroids. The solid squares are the fluxes reported by the
2863: Arecibo radar. The filled triangle is the flux estimated from the
2864: spatially distributed flux in Fig. 2 of Baggaley (2000). The open
2865: triangle is the flux in Baggaley's ``point source''. The horizontal
2866: dotted line is the upper limit on the flux given by assuming that
2867: half of all the heavy elements are in dust grains. The dashed line
2868: joining open circles is the flux corresponding to the ISM dust size
2869: distribution found by Kim and Martin (1995), normalized to the
2870: dotted line. This figure suggests that the bulk of the mass in dust
2871: consists of particles with masses of order $10^{-13}\g$, or sizes of
2872: order two tenths of a micron.
2873: \label{fig:mass flux}
2874: }
2875: \end{figure}
2876:
2877:
2878: \begin{figure}
2879: \epsscale{1.00}
2880: %\plotone{f_survive.eps}
2881: \plotone{figure3.ps}
2882: \caption{
2883: \label{fig:f_survive}
2884: The drag distance $D_{\rm drag}$ (short dash), gyroradius $r_B$ (solid), and
2885: grain destruction distance $D_{\rm dest}$ (long dash) versus grain speed $v$
2886: for four grain sizes, as indicated, and assuming supersonic grain speeds
2887: and $\rho = 3.5 \g \cm^{-3}$,
2888: $\nH = 1 \cm^{-3}$, $U=0.5 \, {\rm V}$, and $B = 5 \, \mu G$.
2889: }
2890: \end{figure}
2891:
2892: \begin{figure}
2893: \epsscale{1.00}
2894: %\plotone{age_ranges.eps}
2895: \plotone{figure4.ps}
2896: \caption{
2897: \label{fig:age_ranges}
2898: The upper and lower current ages $t$ of a Vega-like star for which we would
2899: observe grains emitted when the star's age was between $t_{\rm cr}$
2900: and $2 t_{\rm cr}$, versus the distance $d$ to the star.
2901: }
2902: \end{figure}
2903:
2904:
2905:
2906: \begin{figure}
2907: \epsscale{1.00}
2908: %\plotone{Fig_scatter.eps}
2909: \plotone{figure5.ps}
2910: \caption{ The geometry of hyperbolic motion in the two body
2911: problem. The scattering center (the planet) is at the origin; the
2912: test particle travels along the curved line, with an initial impact
2913: parameter given by $s$. Its closest approach to the planet occurs as
2914: it crosses the apsidal line (choosen here to be the x-axis), when it
2915: is a distance $q$ from the planet. The ingoing and outgoing
2916: asymptotes are denoted by dotted lines. The angle between the ingoing
2917: and outgoing asymptotes is $\gamma$.
2918: \label{fig:scatter}
2919: }
2920: \end{figure}
2921:
2922: \begin{figure}
2923: \epsscale{1.00}
2924: %\plotone{Fig_v_eject.eps}
2925: \plotone{figure6.ps}
2926: \caption{
2927: \label{fig:swift}
2928: Ejection velocities of small particles from a system with a solar mass
2929: star and a Jupiter mass planet at $5.2$AU.
2930: }
2931: \end{figure}
2932:
2933:
2934: \begin{figure}
2935: \epsscale{1.00}
2936: %\plotone{Gliese.eps}
2937: \plotone{figure7.ps}
2938: \caption{
2939: \label{fig:Gliese}
2940: Triangles: apparent locations of the dust flux from the Gliese stars in
2941: Table \ref{tab:Gliese}. Open box: Direction of the solar apex.
2942: Coordinates are ecliptic longitude ($\lambda$) and latitude ($\beta$).
2943: }
2944: \end{figure}
2945:
2946:
2947: \begin{figure}
2948: \epsscale{1.00}
2949: %\plotone{baggaley.eps}
2950: \plotone{figure8.ps}
2951: \caption{
2952: \label{fig:baggaley}
2953: The box marks the central location of the discrete source observed by
2954: Baggaley (2000) and the dashed line indicates its extent.
2955: Triangles mark the apparent location of a dust stream from $\beta \,$Pic for
2956: values of the ejection speed $\vej$ ranging from 2 to $40 \kms$ in steps of
2957: $2 \kms$. For a few cases, $\vej$ and the speed of the dust with respect
2958: to the Sun, $v_{d, \sun}$ are indicated. The ecliptic coordinates of
2959: $\beta \,$Pic itself are $(\lambda, \beta) = (82.5 \arcdeg, -74.4 \arcdeg)$.
2960: }
2961: \end{figure}
2962:
2963: %\include{radar_parameters}
2964: \include{t1}
2965:
2966: \begin{deluxetable}{cccc}
2967: \tablewidth{0pc}
2968: \tablecaption{Idealized Interstellar Environments
2969: \label{tab:phases}}
2970: \tablehead{
2971: \colhead{Phase}&
2972: \colhead{$T_{\rm gas}$}&
2973: \colhead{$\nH$}&
2974: \colhead{$x_e$}
2975: \\
2976: \colhead{}&
2977: \colhead{$\K$}&
2978: \colhead{$\cm^{-3}$}&
2979: \colhead{}
2980: }
2981: \startdata
2982: CNM &100 &30 &$1.5 \times 10^{-3}$\\
2983: WNM &6000 &0.3 &0.1\\
2984: LB &$10^6$ &$5\times 10^{-3}$ &1\\
2985: \enddata
2986: \end{deluxetable}
2987:
2988: \begin{deluxetable}{cccc}
2989: \tablewidth{0pc}
2990: \tablecaption{Threshold Distances for Observing Vega-Like Stars
2991: \label{tab:d_th}}
2992: \tablehead{
2993: \colhead{$A_{col}$}&
2994: \colhead{$t_{\rm cr}$}&
2995: \colhead{$d_{\rm th}$}&
2996: \colhead{$a_{\rm cr}$}
2997: \\
2998: \colhead{$\km^2$}&
2999: \colhead{$\Myr$}&
3000: \colhead{$\pc$}&
3001: \colhead{$\micron$}
3002: }
3003: \startdata
3004: $10^4$ &3 &26 &17\\
3005: $10^4$ &10 &8.6 &10\\
3006: $10^6$ &3 &108 &35\\
3007: $10^6$ &10 &36 &20\\
3008: %%$10^8$ &3 &447 &70\\
3009: %%$10^8$ &10 &147 &40\\
3010: \enddata
3011: \end{deluxetable}
3012:
3013: \begin{deluxetable}{cccccccccccc}
3014: \tablewidth{0pc}
3015: \tablecaption{Gliese Star Dust Fluxes
3016: \label{tab:Gliese}}
3017: \tablehead{
3018: \colhead{Gl \#}&
3019: \colhead{Name}&
3020: \colhead{$d$}&
3021: \colhead{$t$\tablenotemark{a}}&
3022: \colhead{$U$\tablenotemark{b}}&
3023: \colhead{$V$\tablenotemark{b}}&
3024: \colhead{$W$\tablenotemark{b}}&
3025: \colhead{$v_{\ast, \sun}$}&
3026: \colhead{$a_{\rm cr}$}&
3027: \colhead{$F$\tablenotemark{c}}&
3028: \colhead{$\lambda$\tablenotemark{d}}&
3029: \colhead{$\beta$\tablenotemark{d}}
3030: \\
3031: \colhead{}&
3032: \colhead{}&
3033: \colhead{pc}&
3034: \colhead{Myr}&
3035: \colhead{$\kms$}&
3036: \colhead{$\kms$}&
3037: \colhead{$\kms$}&
3038: \colhead{$\kms$}&
3039: \colhead{$\micron$}&
3040: \colhead{$\yr^{-1} \km^{-2}$}&
3041: \colhead{deg}&
3042: \colhead{deg}
3043: }
3044: \startdata
3045: 68.0 &DM+19 279 &7.5 &69.8 &34.5 &-24.7 &2.4
3046: &42.5 &4.4 &3.3E-5 &62 &33\\
3047:
3048: 71.0 &$\tau \,$Cet &3.6 &170 &18.6 &29.4 &12.7
3049: &37.0 &3.3 &6.3E-6 &94 &-54\\
3050:
3051: 111.0 &$\tau^1 \,$Eri &14.0 &39.6 &-26.2 &-16.9 &-13.0
3052: &33.8 &6.8 &1.5E-4 &262 &33\\
3053:
3054: 121.0 &$\tau^3 \,$Eri &26.4 &215 &19.4 &9.96 &-1.09
3055: &21.9 &11.6 &3.4E-8 &101 &-18\\
3056:
3057: 144.0 &$\epsilon \,$Eri &3.22 &59.7 &-3.01 &7.22 &-20.0
3058: &21.5 &4.1 &1.1E-4 &184 &8\\
3059:
3060: 167.1 &$\gamma \,$Dor &20.3 &107 &-22.9 &-18.1 &-13.8
3061: &32.3 &8.4 &1.2E-6 &261 &38\\
3062:
3063: 217.1 &$\zeta \,$Lep &21.5 &69.8 &-14.4 &-11.2 &-8.45
3064: &20.1 &10.9 &3.7E-6 &260 &38\\
3065:
3066: 219.0 &$\beta \,$Pic &19.3 &12.0 &-10.8 &-16.0 &-9.11
3067: &21.4 &10.0 &1.5E-3 &268 &51\\
3068:
3069: 248.0 &$\alpha \,$Pic &30.3 &194 &-35.4 &-19.3 &-10.1
3070: &41.6 &9.0 &1.0E-7 &268 &25\\
3071:
3072: 297.1 &B Car &21.4 &31.4 &19.8 &-15.1 &-34.0
3073: &42.2 &7.5 &1.1E-3 &140 &48\\
3074:
3075: 321.3 &$\delta \,$Vel &24.4 &351 &11.6 &-1.15 &-5.00
3076: &12.7 &14.6 &2.3E-9 &107 &17\\
3077:
3078: 364.0 &DM-23 8646 &14.9 &69.3 &-39.9 &-25.4 &6.27
3079: &47.7 &5.9 &1.4E-5 &290 &17\\
3080:
3081: 448.0 &$\beta \,$Leo &11.1 &135 &-20.1 &-16.1 &-7.80
3082: &26.9 &6.8 &1.2E-6 &268 &36\\
3083:
3084: 557.0 &$\sigma \,$ Boo &15.5 &123 &2.06 &15.9 &-5.17
3085: &16.9 &10.1 &5.7E-7 &165 &-38\\
3086:
3087: 673.1 &DM-24 13337 &25.7 &34.5 &-35.8 &-13.1 &-11.6
3088: &39.9 &8.5 &6.4E-4 &261 &20\\
3089:
3090: 691.0 &$\mu \,$Ara &15.3 &58.8 &-13.7 &-8.39 &-3.97
3091: &16.5 &10.1 &9.7E-6 &270 &26\\
3092:
3093: 721.0 &$\alpha \,$Lyr &7.76 &152 &-16.1 &-6.33 &-7.76
3094: &18.9 &6.7 &1.1E-6 &256 &26\\
3095:
3096: 820.0 &61 Cyg &3.48 &48.0 &-93.8 &-53.4 &-8.33
3097: &108 &1.9 &1.0E-3 &278 &23\\
3098:
3099: 822.0 &$\delta \,$Eql &18.5 &80.0 &5.72 &-28.6 &-10.3
3100: &30.9 &8.2 &4.3E-6 &18 &78\\
3101:
3102: 881.0 &$\alpha \,$PsA &7.69 &83.1 &-5.70 &-8.22 &-11.0
3103: &14.9 &7.6 &5.7E-6 &235 &54\\
3104:
3105: \enddata
3106: \tablenotetext{a}{Age of star from eq.~(\ref{eq:T_tau})}
3107: \tablenotetext{b}{Space velocity of the star relative to the Sun; $U$ is
3108: positive towards the Galactic center, $V$ is positive along the direction
3109: of the Galactic rotation, and $W$ is positive towards the North Galactic Pole.}
3110: \tablenotetext{c}{Dust flux calculated using equation
3111: (\ref{eq:flux_at_earth}) with a limiting grain size $a=10 \micron$, a
3112: flat distribution of ejection velocities between $0.5 \kms$ and $3 \kms$,
3113: $T_{\rm cr} = 3 \Myr$, and $f_{\rm survive} = f_{\rm beam} = f_{\rm ej}
3114: =1$.}
3115: \tablenotetext{d}{Apparent location of dust stream on the sky; $\lambda$ is
3116: ecliptic longitude and $\beta$ is ecliptic latitude. In most cases,
3117: $v_{\rm ej}=1 \kms$ is adopted; for Gl 219.0 ($\beta \,$Pic)
3118: $v_{\rm ej}=3 \kms$ is adopted, since the dust will not yet have reached us if
3119: $v_{\rm ej}=1 \kms$.}
3120: \end{deluxetable}
3121:
3122: \begin{deluxetable}{ccccc}
3123: \tablewidth{0pc}
3124: \tablecaption{AGB Star Dust Fluxes
3125: \label{tab:AGB}}
3126: \tablehead{
3127: \colhead{$d$\tablenotemark{a}}&
3128: \colhead{$F$\tablenotemark{b}}&
3129: \colhead{$N$\tablenotemark{c}}&
3130: \colhead{$a_t$\tablenotemark{d}}&
3131: \colhead{$g$\tablenotemark{e}}
3132: \\
3133: \colhead{\pc}&
3134: \colhead{$\yr^{-1} \km^{-2}$}&
3135: \colhead{}&
3136: \colhead{$\micron$}&
3137: \colhead{}
3138: }
3139: \startdata
3140: 100 &7.99 &1.84 &24.3 &2.50E-4\\
3141: 200 &0.706 &14.7 &34.3 &2.83E-3\\
3142: 500 &2.86E-2 &230 &54.2 &6.99E-2\\
3143: \enddata
3144: \tablenotetext{a}{distance to AGB star}
3145: \tablenotetext{b}{flux from eq.~(\ref{eq:flux_AGB})}
3146: \tablenotetext{c}{number of AGB stars within distance $d$ from Sun}
3147: \tablenotetext{d}{threshold grain size for traceability to source}
3148: \tablenotetext{e}{fraction of emitted grains that must have $a>a_t$ in
3149: order for flux at Earth to equal $2 \times 10^{-3} \yr^{-1} \km^{-2}$}
3150: \end{deluxetable}
3151:
3152: \end{document}
3153: \end
3154:
3155:
3156: