1: %\documentclass[]{aastex}
2: %\documentclass{emulateapj}
3: %\documentclass[onecolumn]{emulateapj}
4: \documentclass[useAMS,usenatbib,usegraphicx]{mn2e}
5: %\usepackage[dvips]{epsfig}
6: \voffset=-1.2cm
7: %\usepackage{epsfig}
8: %\usepackage{graphicx}
9: \usepackage{amsmath}
10: \def \bea {\begin{eqnarray}}
11: \def \ena {\end{eqnarray}}
12: \def \bee {\begin{equation}}
13: \def \ene {\end{equation}}
14: \def \bQ {{\bf Q}}
15: \def \be {{\bf e}}
16: \def \ba {{\bf a}}
17: \def \bb {{\bf b}}
18: \def \bphi {{\bf \Phi}}
19: \def \btheta {{\bf \Theta}}
20: \def \mc {\mbox{ cos }}
21: \def \ms {\mbox{ sin }}
22: \def \mcs {\mbox{ cos}^{2}}
23: \def \mss {\mbox{ sin}^{2}}
24:
25: \def \ma {\hat{\bf a}}
26: \def \me {\hat{\bf e}}
27: \def \mr {{\bf r}}
28: \def \mn {\hat{\bf N}}
29: \def \mm {\hat{\bf n}}
30: \def \mp {{\bf p}}
31: \title{Radiative Torques: Analytical Model and Basic Properties}
32:
33:
34: \author[A. Lazarian \& Thiem Hoang]{A. Lazarian
35: $^1$\thanks{E-mail:lazarian@astro.wisc.edu}, and Thiem Hoang
36: $^2$\thanks{E-mail: hoang@astro.wisc.edu} \\Department of Astronomy, University of Wisconsin, Madison,
37: WI 53706, USA}
38:
39: \begin{document}
40:
41: \date{Draft version \today}
42:
43: \maketitle
44:
45:
46: \begin{abstract}
47: We attempt to get a physical insight into grain alignment processes by studying basic
48: properties of radiative torques (RATs). For
49: this purpose we consider a simple toy model of a helical
50: grain that reproduces
51: well the basic features of RATs.
52: The model grain consists of a spheroidal body with a mirror attached at an
53: angle to it. Being very simple, the model
54: allows analytical description of RATs that act upon it.
55: We show a good correspondence of RATs obtained for this model and those
56: of irregular grains calculated by DDSCAT. Our analysis of the role of
57: different torque components for grain alignment reveals that one of the
58: three RAT components does not affect the alignment, but induces only
59: for grain precession. The other two components provide a generic alignment
60: with grain long axes perpendicular to the radiation direction, if the radiation
61: dominates the grain precession, and perpendicular to
62: magnetic field, otherwise.
63: The latter coincides with the famous predictions of the Davis-Greenstein
64: process, but our model does not invoke paramagnetic relaxation. In fact, we
65: identify a narrow range of angles between the radiation beam and
66: the magnetic field, for which the alignment is opposite to the
67: Davis-Greenstein
68: predictions. This range is likely to vanish, however,
69: in the presence of thermal wobbling of grains. In addition,
70: we find that a substantial part of
71: grains subjected to RATs gets aligned with low angular momentum,
72: which testifies, that most of the grains in diffuse interstellar medium
73: do not rotate fast, i.e. rotate with thermal or even sub-thermal velocities.
74: This tendency of RATs to decrease grain angular velocity
75: as a result of the RAT alignment decreases the degree of polarization, by decreasing the degree of internal
76: alignment, i.e. the alignment of angular momentum with the grain axes.
77: For the radiation-dominated environments, we find
78: that the alignment can take place on the time
79: scale much shorter than the time of gaseous damping of grain rotation. This
80: effect makes grains a more reliable tracer of magnetic fields.
81: In addition, we study a self-similar
82: scaling of RATs as a function of $\lambda/a_{eff}$. We show that the self-similarity is useful for studying grain alignment by a broad spectrum of radiation,
83: i.e. interstellar radiation field.
84: \end{abstract}
85:
86: \begin{keywords}
87: ISM- Magnetic fields- polarization, ISM: dust-extinction
88: \end{keywords}
89:
90: \section{Introduction}
91:
92: Magnetic fields
93: play a crucial role in many astrophysical processes, e.g. star formation,
94: accretion of matter, transport processes, including heat conduction and
95: propagation of cosmic rays. One of the easiest ways to study magnetic
96: field topology is via polarization of radiation arising from extinction
97: or/and emission by aligned dust grains. The new instruments like Scuba II (Bastien, Jenness \& Molnar 2005), Sharc II (Novak et
98: al. 2004) and an intended polarimeter for SOFIA open new horizons for tracing
99: of astrophysical magnetic fields with aligned grains.
100:
101: In this situation it is unacceptable that the processes of grain alignment
102: are not completely understood (see review by Lazarian 2003). The enigma that
103: surrounds grain alignment since its discovery in 1949 (Hall 1949; Hilner 1949)
104: makes the interpretation of the polarization in terms of magnetic fields
105: somewhat unreliable. The failure of grains to
106: align at high optical depths was discussed, for instance, in Goodman (1995).
107:
108: A recent progress in understanding of the grain alignment physics removed
109: many questions, but have not remedied the situation completely. Among the
110: milestones let us mention the recent revival of interest to radiative torques (henceforth RATs).
111: Introduced by Dolginov \& Mytrophanov (1976), those torques, that arise from
112: the interaction of irregular grains with a flow of photons,
113: were essentially forgotten till Draine \& Weingarter
114: (1996, 1997, henceforth DW96, DW97, respectively) provided quantitative numerical studies. While Dolginov \& Mytrophanov (1976) were somewhat vague on what
115: makes RATs important for a grain, DW96 demonstrated that their arbitrary
116: chosen irregular grains exhibit dynamically important RATs when subjected to
117: a typical interstellar radiation field (ISRF). Very importantly, Bruce Draine
118: incorporated RATs into the {\it publicly available} DDSCAT code (Draine \& Flatau
119: 1994), which stimulated a further progress in the field. First laboratory
120: studies of RATs were reported in Abbas et al. (2004).
121:
122: The renewed interest to RATs coincided with a crisis of the paramagnetic
123: alignment as it is described in textbooks (Purcell 1979; Spitzer \& McGlynn 1979; Mathis 1986).
124: Lazarian \& Draine (1999a) (hereafter LD99a) identified new elements of grain dynamics,
125: which they termed ``thermal flipping'' and ``thermal trapping''. Due to
126: thermal wobbling arising from the dissipative
127: coupling of grain vibrational and rotational degrees of freedom (Lazarian
128: 1994; Lazarian \& Roberge 1997) grains smaller than a critical radius
129: $a_c$ flip frequently and thus average out uncompensated torques. These
130: torques, that were first
131: discussed by Purcell (1979), were considered essential to make otherwise
132: inefficient paramagnetic alignment
133: (Davis \& Greenstein 1951; Jones \& Spitzer 1967) to account for
134: the polarimetric observations. A new dissipative coupling mechanism related
135: to nuclear spins of constituent atoms, that was described in Lazarian \&
136: Draine (1999b) (henceforth LD99b), resulted in $a_c$ larger than the typical cut-off
137: scale for grains in diffuse interstellar medium (ISM). As the other
138: mechanisms, e.g. mechanical alignment (Gold 1951, review by Lazarian
139: 2003 and references therein), have their limitations, this made the RAT alignment the only viable
140: mechanism to explain the ubiquity of interstellar polarization (DW97)
141: and possibly polarization arising from aligned dust in other astrophysical
142: environments (Lazarian 2007).
143:
144: The successes of RATs include a more recent work by
145: Cho \& Lazarian (2005),
146: where a substantial increase of the RAT efficiency with the grain
147: size was established. This work
148: explained the sub-millimeter polarization data for quiescent starless cores
149: (Wart-Thompson et al. 2000) by
150: appealing to the differential RAT alignment of large grains. For such
151: cores the analysis of all other mechanisms in Lazarian, Goodman \& Myers (1997)
152: predicted only marginal degrees of alignment. The studies elaborating the
153: approach in Cho \& Lazarian (2005), e.g. Pelkonen, Juvela \& Padoan (2007),
154: Bethell et al. (2007), provided theory-motivated predictions of the degree of
155: alignment for numerically simulated molecular clouds and cores.
156:
157: However, the above explanation as well as other explanations (see Lazarian 2003) are based on the plausibility of arguments, rather than on
158: the rigorous RAT alignment theory. Indeed, DW96 considered
159: RATs as a means of spin-up. This induced a naive
160: explanation of RAT alignment action that could be perceived in some
161: of the papers that followed the DW96 study.
162: There it was assumed that RATs were proxies of the
163: Purcell's torques (1979), that arise from the action of photons, rather
164: than from the action of H$_2$ formation over catalytic sites, as in the original mechanism. While the Purcell's torques
165: depend on the resurfacing and therefore short lived, RATs depend on
166: grain shape and can be long-lived. As a result, long-lived fast rotation of paramagnetic
167: grains should induce good paramagnetic alignment (Purcell 1979). This
168: understanding of RATs is not correct, as it is clear from a more careful
169: reading of DW96 and DW97.
170:
171: In fact, RATs can be subdivided into the parts that arise from isotropic
172: and anisotropic radiation fluxes. The part arising from anisotropic radiation,
173: for which we adopt a shorthand notation ``isotropic part'', is, indeed, similar
174: to the Purcell's torques. The ``anisotropic part'' is, however, both
175: usually stronger and has properties
176: different from the Purcell's torques. The
177: major difference arises from the fact that RATs
178: are defined in the laboratory, rather
179: than in the grain coordinate system. Thus, the presence of even a small
180: anisotropic component of radiation, which is a natural
181: condition for any realistic astrophysical system, is bound to change
182: the dynamics of grain. Note, that the alignment by the anisotropic
183: radiation was first discussed by Doginov \& Mitraphanov (1976). They, however,
184: concluded that prolate and oblate grains can be aligned differently. Lazarian
185: (1995) took into account internal relaxation and claimed that both prolate
186: and oblate grains should be aligned with longer axes perpendicular to magnetic
187: field. Nevertheless, the theory lacked a proper description of RATs.
188:
189: DW97 demonstrated numerically
190: that in the presence of anisotropic radiation
191: the grains can be aligned
192: by RATs in respect to magnetic field on the time scales much {\it shorter}
193: than the time scale for paramagnetic alignment\footnote{DW97 identifies this
194: time-scale with the gaseous damping time. In \S 5.5 we show that the
195: alignment could happen much faster in the presence of strong radiation sources}. In general, the magnetic field for the RAT alignment
196: acts through inducing fast Larmor precession; the alignment
197: potentially may happen both with long grain axes parallel and perpendicular
198: to magnetic field. Only the latter is consistent with polarimetric
199: observations, however (see Serkowski, Mathewson \& Ford 1975).
200:
201:
202: In the DW97 study, the alignment with longer grain axes perpendicular
203: to magnetic field (``right alignment'') happened more frequently
204: than the grain alignment with longer grain axes parallel to magnetic field
205: (``wrong alignment'').
206: This experimental evidence, based on a limited sampling, raised
207: worrisome questions.
208: Is this a general property of radiative torques
209: or just a coincidence? Do we expect to see more of ``wrong alignment'' if the
210: grain environment is different from the interstellar one? What are the
211: chances that we are fooled by the ``wrong alignment'' while interpreting
212: the polarimetry measurements in terms of the underlying magnetic fields?
213: It seems necessary to address these questions
214: if interpreting polarimetry data in terms
215: of underlying magnetic fields is sought.
216:
217:
218: Analytical calculations played an important role for formulating the models
219: of both paramagnetic and mechanical alignment (see Davis \& Greenstein 1951,
220: Jones \& Spitzer 1967, Purcell 1979, Spitzer \& McGlynn 1979).
221: Although such calculations
222: dealt with
223: intentionally idealized models of grains, they allowed deep insight into the relevant physics.
224: Dolginov \& Mytrophanov (1976) attempted an analytical modeling
225: for RATs. They used a model
226: grain containing two ellipsoids connected together at an angle.
227: The radiative torques were calculated for such a model grain by assuming that
228: the wavelength is much larger than the grain size, i.e., in the Rayleigh-Hans approximation (Dolginov \& Silantev 1976). However, adopting their shape,
229: we could not reproduce numerically their analytical predictions for RATs.
230: This induce us to seek analytical models that would correspond to the
231: DDSCAT calculations.
232:
233: Our approach in the present paper is to provide a physical insight into basic
234: RAT properties. In \S 2 we explain why we consider only RATs due to anisotropic radiation.
235: Then, we describe a simple grain model that is subject to RATs and allows
236: analytical descriptions (\S \ref{amo1}). In \S \ref{sec33} we present calculations of RATs for
237: a number irregular shapes and study the correspondence of their RATs with AMO. We also briefly consider possible generalizations of our model (\S \ref{amo2}). In \S \ref{kalign} we analyze the
238: alignment for both this model and irregular grains with respect to the radiation direction. In \S \ref{balign} we study the alignment of our model and irregular grains with respect to magnetic
239: field. Crossovers are studied in \S \ref{cross}, while we identify the conditions
240: for the magnetic field or the radiation direction
241: to act as the axis of alignment in \S \ref{sec9}.
242: As, even with modern computers, the calculations of RATs
243: for a variety of wavelengths is time consuming,
244: we address the question of the accuracy of presenting radiative torques
245: as the function of the ratio of grain size to the wavelength $\lambda/a_{eff}$
246: in \S \ref{self}.
247: The discussion of our results and the summary are provided in
248: \S \ref{discuss} and \S 12, respectively.
249:
250: \section{Isotropic and Anisotropic RATs}
251:
252: RATs can emerge even when the radiation field is isotropic. Devices similar to those used by Lebedev (1901) to measure radiation
253: pressure experience torques in the presence of the isotropic radiation (see also
254: the cartoon of a model with absorbing and reflecting strips in DW96).
255:
256: The dynamics of an irregular grain subjected to isotropic radiation is very similar
257: to a grain subjected to the Purcell's torques arising, for instance, from
258: H$_2$ formation. For instance, one would expect to have thermal trapping
259: of sufficiently small grains due to thermal fluctuations as described
260: in LD99a. Therefore RATs induced by isotropic radiation
261: (henceforth ``isotropic RATs'') only marginally alter
262: the problems that the paramagnetic alignment mechanism faces in explaining
263: observational data. In addition, as we mentioned above, the ``isotropic RATs'' are usually weaker
264: than those that arise when a grain is subjected to anisotropic radiation (see DW97).
265:
266: Due to the situation described above, for the rest of the paper we shall
267: associate RATs only with the part arising from anisotropic radiation, as, for instance, was done
268: in Cho \& Lazarian (2005). In other words, we treat the torques arising from isotropic radiation
269: as a particular realization of the Purcell's torques.
270:
271: Let us now introduce briefly some basic definitions. RAT ${\bf \Gamma}_{rad}$ is defined by
272: \bea
273: {\bf \Gamma}_{rad}=\frac{\gamma u_{\lambda}\lambda a_{eff}^{2}}{2}{\bf Q}_{\Gamma},\label{eq1}
274: \ena
275: where ${\bf Q}_{\Gamma}$ is the RAT efficiency, $\gamma$ is the anisotropy degree, and $u_{\lambda}$ is the energy density of
276: radiation field of the wavelength $\lambda$. Here $a_{eff}$ is the effective
277: size of the grain which is defined as the radius of a sphere of the same
278: volume with the irregular grain (similar to DW97). In general, ${\bf Q}_{\Gamma}$ is a
279: function of angles $\Theta, \beta, \Phi$ in which $\Theta$ is the angle
280: between the axis ${\bf a}_{1}$ corresponding to the maximal moment of inertia
281: (henceforth maximal inertia axis) with respect to the radiation
282: direction ${\bf k}$,
283: $\beta$ is the rotation angle of the grain around ${\bf a}_{1}$, and $\Phi$ is
284: the precession angle of ${\bf a}_{1}$ about ${\bf k}$ (see Fig. \ref{f1}). To help the
285: reader familiar with the earlier works on RATs, wherever possible, we use the
286: same the notations as in DW96 and DW97.
287:
288: \begin{figure}
289: \includegraphics[width=0.49\textwidth]{f1.eps}
290: \caption{The orientation of a grain, described by three principal axes
291: $\hat{a}_{1},\hat{a}_{2}, \hat{a}_{3}$, in the laboratory coordinate system (scattering reference
292: system)
293: $\hat{e}_{1},\hat{e}_{2}, \hat{e}_{3}$ is
294: defined by three angles $\Theta, \beta, \Phi$. The
295: direction of incident photon beam ${\bf k}$ is along $\hat{e}_{1}$.}
296: \label{f1}
297: \end{figure}
298:
299: The RAT efficiency can be decomposed into components in the scattering system via
300: \begin{align}
301: {\bf Q}_{\Gamma}(\Theta, \beta,\Phi)&=Q_{e1}(\Theta,\beta,0)\hat{e}_{1}\nonumber\\
302: &+Q_{e2}(\Theta, \beta, 0)(\hat{e}_{2}\mc\Phi+\hat{e}_{3}\ms\Phi)\nonumber\\
303: &+Q_{e3}(\Theta, \beta, 0)(\hat{e}_{3}\mc\Phi-\hat{e}_{2}\ms\Phi), \label{eq2}
304: \end{align}
305: where $\hat{e}_{1}$, $\hat{e}_{2}$, $\hat{e}_{3}$ are shown in Fig.
306: \ref{f2}. In addition, for the sake of simplicity, we have denoted
307: $Q_{e1}(\Theta,\beta,0)\equiv{\bf Q}_{\Gamma}(\Theta,\beta,0).\hat{e}_{1}$,
308: $Q_{e2}(\Theta,\beta,0)\equiv{\bf Q}_{\Gamma}(\Theta,\beta,0).\hat{e}_{2}$,
309: $Q_{e3}(\Theta,\beta,0)\equiv{\bf Q}_{\Gamma}(\Theta,\beta,0).\hat{e}_{3}$. In what
310: followings, we use $Q_{e1}, Q_{e2}, Q_{e3}$ for the RAT components and keep in mind that
311: they are functions of $\Theta, \beta$ at $\Phi=0$. However, in some
312: particular cases, these angles will be explicitly written.
313:
314:
315: \section{Introducing Analytical model (AMO) of a helical grain}\label{amo1}
316:
317: Let us consider an asymmetric grain shape consisting of a reflecting
318: spheroid and a square mirror with the side $l_{2}$ attached on a pole of
319: the length $l_{1}$. For the sake of simplicity, we assume that the mirror and
320: the pole are weightless. Also, both the mirror and the spheroid are assumed to
321: be perfectly reflecting. Moreover, we neglect
322: the shadowing of the mirror by the grain by assuming that $l_{1}\gg l_{2}$ (see
323: Fig. \ref{f2}).
324: \begin{figure}
325: \includegraphics[width=0.49\textwidth]{f2.eps}
326: \caption{Our grain model consists of a mirror connected to an oblate spheroid by a weightless
327: rod. The distance between the mirror and the spheroid is assumed to be much
328: larger than the mirror's size. Both the mirror and the spheroid are perfectly reflecting.}
329: \label{f2}
330: \end{figure}
331:
332: \subsection{RATs from a reflecting spheroidal body}
333: Consider first RATs acting on an oblate spheroidal body. As a consequence of
334: its symmetries, such a grain is not expected to
335: exhibit any spin-up arising from RATs, provided that the incident radiation is
336: not circularly polarized. The latter will be
337: our assumption for the rest of the paper.
338:
339: Consider a photon beam of wavelength $\lambda$, propagating in the ${\bf k}$-direction parallel
340: to the axis $\hat{e}_{1}$ of the lab coordinate system (see Figs \ref{f1} and \ref{f2}). The
341: momentum of a photon is deposited to the grain as it reflects from the grain
342: surface. As a result, the grain experiences a net torque.
343:
344: For simplicity, we assume that the grain rotates fast around the maximal
345: inertia axis, so averaging over such a rotation is suitable. Also, from
346: equation ({\ref{eq2}) it follows that we only need to find RATs for
347: $\Phi=0$. Therefore, the resulting RAT is
348: \begin{align}
349: {\bf \Gamma}_{rad}&=\frac{\gamma u_{rad}\lambda b^{2}}{2}(Q_{e1}\me_{1}+Q_{e2}\me_{2}+Q_{e3}\me_{3}),\label{eq3a}
350: \end{align}
351: where $b$ is the length of the major axis of the spheroid, RAT components are given by (see Appendix A for their derivation)
352: \begin{align}
353: Q_{e3}&=\frac{2ea}{\lambda}(s^{2}-1)K(\Theta, e)\ms2\Theta,\label{eq3}\\
354: Q_{e1}&=0,\label{eq4}\\
355: Q_{e2}&=0,\label{eq5}
356: \end{align}
357: where $s=a/b<1$ is the ratio of the minor to the major axes, $K(\Theta, e)$ is
358: a fitting function depending on $\Theta$ and the eccentricity of the oblate spheroid (see the lower panel in Fig. \ref{ap1}).
359:
360:
361: Following equations (\ref{eq3})-(\ref{eq5}) we see that a reflecting spheroidal grain does not produce
362: any $Q_{e1}, Q_{e2}$, but $Q_{e3}$ only.
363: Therefore, it is easy to see that
364: the only effect of RATs on the spheroidal grain is to cause the precession
365: in the plane perpendicular to the radiation direction ${\bf k}\|\me_{1}$.
366:
367:
368: \subsection{Torques from a reflecting mirror}
369: %\subsubsection{General expressions for RATs}
370: Consider now torques that act upon the perfectly reflecting mirror attached at an angle to the oblate spheroid
371: (see Fig.~\ref{f2}). The pole is considered too thin to interact with the
372: radiation.\footnote{In our model, the only purpose of the existence of the
373: pole is to minimize the effects of shadowing of the oblate grain core by the mirror.}
374: The normal unit vector $\mn$ which determines the orientation of the mirror in
375: the grain coordinate system is given by
376: \bea
377: \mn =n_{1}\ma_{1}+n_{2}\ma_{2},
378: \ena
379: where $\ma_{1}, \ma_{2},\ma_{3}$ are the principal axes of the grain
380: (i.e., the principal axes of the spheroid because the mirror and the pole are weightless). Here $n_{1}=\ms
381: \alpha, n_{2}=\mc\alpha$ with $\alpha$ is the angle between $\mn$ and $\ma_{2}$ (see Fig. \ref{f2}).
382:
383: Due to the rotation, the cross section of the mirror with the surface area
384: $A$, varies as (see Appendix B)
385: \begin{align}
386: A_{\perp}=A |\me_{1}. \mn|=A |n_{1} \mc\Theta-n_{2} \ms\Theta \mc\beta|,\label{eq7}
387: \end{align}
388:
389: Following the same above procedure (see Appendix B for detail), we get RAT
390: \begin{align}
391: {\bf \Gamma}_{rad}&=\frac{\gamma u_{rad}\lambda l_{2}^{2}}{2}(Q_{e1}\me_{1}+Q_{e2}\me_{2}+Q_{e3}\me_{3}),\label{eq7b}
392: \end{align}
393: where $l_{2}$ is the size
394: of the square mirror, $l_{1}$ is the length of the pole, and RAT components are given by
395: \begin{align}
396: Q_{e1}&=\frac{4l_{1}}{\lambda}|n_{1}\mc\Theta-n_{2}\ms\Theta\mc\beta|[n_{1}n_{2}\mcs\Theta\nonumber\\
397: &+\frac{n_{1}^{2}}{2}\mc\beta\ms2\Theta-\frac{n_{2}^{2}}{2}\mc\beta\ms2\Theta\nonumber\\
398: &-n_{1}n_{2}\mss\Theta\mcs\beta],\label{eq8}\\
399: Q_{e2}&=-\frac{4l_{1}}{\lambda}|n_{1}\mc\Theta-n_{2}\ms\Theta \mc\beta|[n_{1}^{2}\mc\beta\mcs\Theta\nonumber\\
400: &-\frac{n_{1}n_{2}}{2}\mcs\beta\ms2\Theta-\frac{n_{1}n_{2}}{2}\ms2\Theta+n_{2}^{2}\mc\beta\mss\Theta],\label{eq9}\\
401: Q_{e3}&=-\frac{4l_{1}}{\lambda}|n_{1} \mc\Theta-n_{2} \ms\Theta \mc\beta|
402: n_{1}\ms\beta[n_{1}\mc\Theta\nonumber\\
403: &-n_{2}\mc\beta\ms\Theta].\label{eq10}
404: \end{align}
405:
406: For $\Theta=0, \pi$, the RAT components for the mirror are
407: \begin{align}
408: Q_{e1}&=\frac{4 l_{1} n_{1}^{2}n_{2}}{\lambda},\label{eq11}\\
409: Q_{e2}&=-\frac{4 l_{1} n_{1}^{3}}{\lambda}\mc\beta,\label{eq12} \\
410: Q_{e3}&=-\frac{4 l_{1} n_{1}^{3}}{\lambda}\ms\beta,\label{eq13}
411: \end{align}
412: Equations (\ref{eq11}), (\ref{eq12}) and (\ref{eq13}) reveal that $Q_{e1}$ does not depend on $\beta$, but
413: $Q_{e2}$, $Q_{e3}$ are periodic functions of $\beta$. As a result,
414: when averaging over the rotation angle $\beta$ from $0$ to $2\pi$, $Q_{e2},
415: Q_{e3}$ vanish for $\Theta=0, \pi$ (see a proof based on
416: more general symmetry considerations
417: in \S \ref{sec42}).
418:
419: For arbitrary $\Theta$, assuming that the rotation of the grain around the shortest axis is very fast, we can
420: average RATs over $\beta$ from $0$ to $2\pi$. The resulting RAT components are given by
421: \begin{align}
422: Q_{e1}&=\frac{4\pi l_{1} n_{1}n_{2}}{\lambda} (3 \mbox{cos}^{2} \Theta - 1) f(\Theta, \alpha),\label{eq14}\\
423: Q_{e2}&=\frac{4\pi l_{1} n_{1}n_{2}}{\lambda} \ms 2\Theta g(\Theta, \alpha),\label{eq15}\\
424: Q_{e3}&=0,\label{eq16a}
425: \end{align}
426: where $f(\Theta, \alpha), g(\Theta, \alpha) $ are fitting functions depending on
427: $\Theta, \alpha$ which characterize the influence of variation of cross section on RATs. The analytical approximations for them are given in Appendix B. Note, that the dependence on $l_{1}$ arises in equations (\ref{eq14}) and
428: (\ref{eq15}) due to the assumption $\lambda \ll l_{1}$. In the opposite limit
429: we expect $\lambda$ to act as an arm for torques and therefore no dependences of the RAT
430: efficiencies $Q_{e1}, Q_{e2}$ on $l_{1}/\lambda$ to exist.
431:
432:
433: \subsection{AMO: RATs' Properties}\label{amo3}
434:
435: For the sake of simplicity, for the rest of
436: the paper, apart from the Appendices B2 and B3, we consider AMO for a single value
437: of angle $\alpha=\pi/4$.
438: Combining RATs produced by the reflecting oblate spheroid (see equations \ref{eq3}-
439: \ref{eq5}), and RATs induced by the reflecting mirror (see equations \ref{eq14}-
440: \ref{eq16a}), for $\alpha=\pi/4$ AMO has the following components
441: \begin{align}
442: Q_{e1}&= \frac{4\pi l_{1} n_{1}n_{2}}{\lambda} (3 \mbox{cos}^{2} \Theta - 1)f(\Theta, \pi/4),\label{eq18}\\
443: Q_{e2}&= \frac{4\pi l_{1} n_{1}n_{2}}{\lambda} \ms 2\Theta g(\Theta, \pi/4),\label{eq19}\\
444: Q_{e3}&=\frac{2e a(s^{2}-1)}{\lambda}K(\Theta, e) \ms 2\Theta,\label{eq20}
445: \end{align}
446: where the analytical and numerical fitting functions $f(\Theta, \pi/4)$ and
447: $g(\Theta, \pi/4)$ are shown in Figs B1 and B2, respectively.
448:
449: However, as we see further, $Q_{e3}$ does not affect the alignment apart from inducing the precession. To roughly estimate the latter, one does not need to know
450: the exact form of $K(\Theta, e)$ (see \S \ref{kalign}).
451:
452: Equation (\ref{eq18}) reveals clearly that $Q_{e1}$ is symmetric, while equation (\ref{eq19}) shows that
453: $Q_{e2}$ is asymmetric. In addition, $Q_{e2}$ is zero for $\Theta=0, \pi$ and
454: $\pi/2$, i.e., when the maximal inertia
455: axis ${\bf a}_{1}$
456: is parallel or perpendicular to the
457: light direction (see also Fig. \ref{f6}).
458:
459: In particular, two first components, $Q_{e1}, Q_{e2}$ depend on the product of projections of the normal vector,
460: i.e., $n_{1}n_{2}$, that suggests us to define the helicity of a grain. A grain of {\it right} helicity is defined so that, when
461: the maximal inertia axis ${\bf a}_{1}$ is parallel to the radiation beam,
462: it can rotate in clockwise sense around the radiation beam; a grain has {\it left} helicity if it rotates anti-clockwise around the radiation beam. With these definitions, we see from
463: equations (\ref{eq18}) and (\ref{eq19}) that
464: the grain with right helicity corresponds to the mirror being oriented so that
465: $n_{1}n_{2}>0$, and $n_{1}n_{2}<0$ for the grain with left helicity.
466: It is straightforward to change the grain helicity from right to left by a
467: rotation of the mirror over an angle of $\pi/2$. Thus, grains with
468: left and right helicity
469: are mirror symmetric, as expected. Naturally, the helicity of a spheroid is
470: zero.
471:
472: Fig. \ref{f6} shows components of RAT normalized over the maximum of $|Q_{e1}|$, so that they exhibit the functional dependence of RATs
473: on $\Theta$. Since $Q_{e3}$ arises from the reflecting spheroid, not subject to the mirror size, in order for AMO to be self-consistent, we normalize $Q_{e3}$ so that $Q_{e3}^{max}=Q_{e1}^{max}$.
474: It is evident from examining the upper and lower panels of Fig.~\ref{f6}
475: that the transition from left-handed to right-handed grains induces a simultaneous change of the both $Q_{e1}$ and $Q_{e2}$ components. This can be easily understood based on the fact that both $Q_{e1}$ and $Q_{e2}$ depend on the product $n_{1}n_{2}$, which defines the helicity of grain. Therefore, as the helicity changes, i.e., $n_{1}n_{2}$ reverses the sign, $Q_{e1}, Q_{e2}$ change synchronically. We shall see that this property
476: is also present for arbitrary-chosen irregular grains that we study numerically
477: in \S \ref{sec33}. Fig. \ref{f6} also shows clearly that the third component $Q_{e3}$ that arises from
478: the spheroidal body of the grain does not change as the
479: grain gets the opposite helicity. This exactly what we expect from a spheroid.
480:
481: \begin{figure}
482: \includegraphics[width=0.49\textwidth]{f6a.ps}
483: \includegraphics[width=0.49\textwidth]{f6b.ps}
484: \caption{The components of RAT normalized over the maximum of $|Q_{e1}|$ as a function of $\Theta$ for
485: right handed ({\it Upper Panel}) and left handed ({\it Lower Panel}) models of AMO.
486: The helicity change
487: is obtained by twisting the mirror
488: by $\pi/2$. The maximum of $Q_{e3}$ is also normalized to be equal $Q_{e1}^{max}$. Figs exhibit
489: zeros of $Q_{e2}$, and symmetries of RATs that we study for irregular grains
490: using DDSCAT in \S \ref{sec33}.}
491: \label{f6}
492: \end{figure}
493:
494:
495: \section{RATs: AMO versus Irregular Grains}\label{sec33}
496:
497: We justify AMO's utility
498: finding the correspondence of the functional form obtained
499: for the torques that our toy model experience with those exerted on
500: actual irregular grains. We start with finding generic properties
501: of RATs for irregular grains using general symmetry considerations and
502: follow further with numerical calculations.
503:
504:
505: \subsection{Symmetry considerations for RATs}
506:
507: Here we show
508: that some properties of RATs follow from general considerations based on the
509: analysis of symmetries.
510: For instance,
511: we have observed that for AMO $Q_{e2}$, $Q_{e2}$ become
512: zero after
513: $\beta$-averaging at points $\Theta=0$ and $\pi$, while $Q_{e1}$ does not
514: depend on $\beta$. This property is valid for arbitrarily shaped grains.
515: Indeed, when $\Theta$ is either 0 or $\pi$ the radiation direction presents the
516: axis of symmetry. It is obvious, therefore that changes of $\beta$
517: cannot change the RAT component along the radiation direction (i.e. does
518: not change $Q_{e1}$), while any perpendicular component of RAT (i.e. both
519: $Q_{e2}$ and $Q_{e3}$), should vanish as the result of $\beta$-averaging\footnote{In fact, these considerations prove not only the zero values of $Q_{e2}$ and $Q_{e3}$ but also their periodicity as a function of $\beta$.}
520:
521:
522: Further on, we discuss the properties of $\beta$-averaged RATs.
523: We can observe, that, similar to the case of AMO, the component of RATs
524: $Q_{e1}$ for irregular grains
525: is symmetric with respect to
526: $\Theta \to \pi-\Theta$ change (see Fig. \ref{f21}). This symmetry is not
527: exact, but it gets better for grains for which mutual shadowing of dipoles
528: gets less.
529: The symmetry of $Q_{e1}$
530: ensures that the torque along ${\bf k}$ has the same sign and similar magnitude when the grain flips over.
531: At the same time the RAT component $Q_{e2}$ is anti-symmetric (see Fig. \ref{f21}), it changes the
532: sign for a transformation $\pi
533: -\Theta$. This also corresponds to AMO. Similarly to $Q_{e3}$ the symmetry
534: of $Q_{e1}$ is only approximate.
535:
536:
537:
538: \subsection{Zero points of $Q_{e2}$ at $\Theta =\pi/2$}\label{sec42}
539:
540: For AMO, according to equation (\ref{eq20}), $Q_{e2}$ is equal to zero for $\Theta=\pi/2$. A
541: similar property also exists for irregular grains, as obviously seen in Figs
542: \ref{f21} and \ref{f23}. There it is shown that
543: when the maximal inertia axis is perpendicular to the radiation direction, the
544: magnitude of $Q_{e2}$ is very small. This can be explained in terms of
545: the interaction
546: of the electric dipoles
547: with the electric field vector of radiation as follows: the interaction between
548: electric field and electric dipoles induces their rotation around
549: ${\bf e}_{2}$ to emit circularly polarized photons. However since ${\bf a}_{1}$ is perpendicular to
550: ${\bf k}$,
551: the electric field is only able to induce the rotation of the electric dipoles in a plane
552: containing ${\bf e}_{2}$. As a result, the torque component vanishes.
553: Quantitatively, according to equation (C5) (see Appendix C), it follows that, when $\Theta=90^{0}$, we have
554: \bea
555: {\bf Q}_{abs}.{\bf e}_{2}\sim k \mbox{ cos } \beta
556: \mbox{Re}\sum_{j}[p_{j}.E_{inc}]e^{ik x_{j}}.
557: \ena
558:
559: Since $x_{j}=r\mbox{sin }\alpha \mbox{sin }\beta$, it follows
560: \bea
561: {\bf Q}_{abs}.{\bf e}_{2} \sim \sum_{j}\mbox{ cos }\beta \mbox{ cos }(A\mbox{
562: sin }\beta) [p_{j}.E_{inc}],
563: \ena
564: where the term $[p_{j}.E_{inc}]$ is a function that is independent of
565: $\beta$.
566: It is obvious that ${\bf Q}_{abs}.{\bf e}_{2}$ is a function of $\beta$ which
567: is zero when averaging is performed for $\beta$ over [$0$, $2 \pi$]. We calculated $Q_{abs}.e_{2}$ for
568: different $\lambda/a$ and grain shapes, and found that $Q_{e2}$ is indeed
569: close to zero at $\mbox{cos}\Theta=0$, which is consistent with our analytical
570: expectation.
571:
572: \subsection{DDSCAT Calculations}\label{sec43}
573:
574:
575: \begin{figure*}
576: \includegraphics[width=0.32\textwidth]{f8a.eps}
577: \includegraphics[width=0.32\textwidth]{f8b.eps}
578: \includegraphics[width=0.32\textwidth]{f8c.eps}
579: \includegraphics[width=0.32\textwidth]{f8d.eps}
580: \includegraphics[width=0.32\textwidth]{f8e.eps}
581: \includegraphics[width=0.32\textwidth]{f8f.eps}
582: \caption{Geometry of grains under study: shape 1, 2, 3 are similar to those of
583: DW97, shape 4, and 5 are created from 15 and 11 cubic blocks respectively,
584: and an ellipsoidal shape.}
585: \label{f18}
586: \end{figure*}
587:
588: Fig.~\ref{f18} presents the test grain shapes that we have calculated RATs for using DDSCAT. Parameters for calculations are given in Table \ref{tab2}. Shapes 1, 2 and 3 have been
589: used in DW97. We added to them shapes 4 and 5. In addition, we created a mirror
590: symmetric shape of shape 1, namely, shape 1* and provided the DDSCAT
591: calculations for a spheroidal grain (see more details in Table~\ref{tab2}). We
592: adopt dielectric functions for astronomical silicate in which a feature in
593: the ultraviolet is removed (see DW97; Weingartner \& Draine 2001; Cho \&
594: Lazarian 2005).
595:
596: \begin{table*}
597: %\renewcommand{\thetable}{2}
598: \caption{Grain shapes and parameters for calculation of RATs}
599: \begin{displaymath}
600: \begin{array}{rrrrr} \hline\hline\\
601: \multicolumn{1}{c}{\bf Grain~shapes} & \multicolumn{1}{c}{\bf Dipole~ \#} &{\bf Size ~(\mu m)} & \multicolumn{1}{c}{\bf Wavelength~ (\mu m)} &\multicolumn{1}{c}{\bf Helicity}\\[1mm]
602: \hline\\
603: {\rm Shape~ 1}&{\it 832000 }&{\it 0.05-0.2}&{\it ISRF}&{\it right}\\[1mm]
604: {\rm Shape~ 1^{*}}&{\it 53248}&{\it 0.2}&{\it 1.2}&{\it left}\\[1mm]
605: {\rm Shape~ 2}&{\it 45056}&{\it 0.2} &{\it ISRF}&{\it left}\\[1mm]
606: {\rm Shape~ 3}&{\it 102570} & {\it 0.2}&{\it ISRF} &{\it left} \\[1mm]
607: {\rm Shape~ 4}&{\it 15000}&{\it 0.2} &{\it 1.2}&{\it left}\\[1mm]
608: {\rm Shape~ 5}&{\it 11000}&{\it 0.2}&{\it 1.2} &{\it left} \\[1mm]
609: {\rm Hollow~1}&{\it 832000 }&{\it 1.0}&{\it 0.1}&{\it right}\\[1mm]
610: \hline\\[1mm]
611: \\[1mm]\hline
612:
613: \end{array}
614: \end{displaymath}
615: \label{tab2}
616: For all calculations here, we adopt the dielectric function for astronomical silicate.
617: \end{table*}
618:
619: \begin{figure}
620: \includegraphics[width=0.49\textwidth]{f9a.ps}
621: \caption{Figure shows the comparison of RATs normalized over the maximum of $|Q_{e1}|$ between AMO (the left handed grain) and
622: DDSCAT (for shapes 2 and 4 and monochromatic radiation of $\lambda=1.2\mu m$). Solid and dashed
623: lines show normalized RATs corresponding to $Q_{e1}^{max}/Q_{e2}^{max}=0.78$ and $1$ in which the functional forms are obtained from the analytical approximation given by equations (\ref{eq18})
624: and (\ref{eq19}) with tabulated functions $f, g$. Dot and dashed-dot
625: lines show normalized RATs for shape 2 and shape 4, respectively.}
626: \label{f61}
627: \end{figure}
628:
629:
630:
631: We discussed for AMO, that the sign of helicity can be changed
632: by taking the mirror image of the grain.
633: We performed a similar procedure
634: to the irregular grains and obtained results similar to the ones obtained
635: for AMO (see Figs. \ref{f23} and \ref{f23a}).
636:
637: Note, that we observe that $Q_{e1}$ and $Q_{e2}$ change synchronously when we
638: calculate torques for a mirror image of a grain. We see that the shape
639: 1 has one type of helicity, while shapes 4, 5 and mirror symmetric image
640: of shape 1, i.e. shape 1*, have another type of helicity.
641:
642: \begin{figure}
643: \includegraphics[width=0.49\textwidth]{f25b.ps}
644: \includegraphics[width=0.49\textwidth]{f25a.ps}
645: \caption{$Q_{e1}$ and $Q_{e2}$ for grains of left ({\it Upper Panel})
646: and right ({\it Lower Panel}) helicity. The symmetry of $Q_{e1}$ with the transformation $\pi -\Theta$ is clearly seen for all grain shapes. $Q_{e2}$ is antisymmetric
647: with respect to the
648: same transformation $\pi -\Theta$. A similarity with the torques
649: produced by AMO is evident (see also Fig \ref{f61}).}
650: \label{f21}
651: \end{figure}
652:
653: Fig.~\ref{f61} provides a comparison of normalized RATs between AMO and
654: DDSCAT calculations performed for two irregular grains induced by monochromatic radiation
655: field of $\lambda=0.2 \mu m$ (see more in \S \ref{para}). It can be seen that they possess the same symmetric properties as well as
656: zero points. Also, the functional form of normalized $Q_{e1}$ and $Q_{e2}$ calculated for
657: the irregular grains and AMO are remarkably similar, in particular for
658: the $Q_{e2}$ component. Typically, the RAT components for shape 2 are similar to those
659: of AMO
660: with $Q_{e1}^{max}/Q_{e2}^{max}=1$ ratio, while RATs of shape 4 are similar to those of AMO with $Q_{e1}^{max}/Q_{e2}^{max}=0.78$
661: . Hence, by changing the ratio of amplitudes of the RAT components for AMO, we
662: can obtain analytical expressions of RATs for a number of irregular
663: grains. To have RATs appropriate to irregular grains, it is necessary to use
664: DDSCAT to estimate the magnitude of RATs. Combining functional forms from AMO
665: and magnitude from DDSCAT, we can obtain analytical approximate expressions for
666: RATs components of
667: irregular grains. Note, that in Figs \ref{f6} and \ref{f61}, we normalized RATs over
668: $|Q_{e1}^{max}|=|Q_{e1}(\Theta=0)|$, that gives rise to $Q_{e1}^{max}$
669: remained the same for all realizations of
670: AMO and the irregular grains. It is easy to see that with this choice, AMO reproduces very
671: well $Q_{e2}$ for irregular grains, but gets $Q_{e1}$, which is a bit larger at $\mc\Theta=0$ ($\Theta=\pi/2$) than that for irregular grains. Potentially, this may mean that
672: more appropriate parametrization should include $Q_{e1}^{max}$, which is not
673: defined as $Q_{e1}$ at $|\mc\Theta|=1$ ($\Theta=0$ or $\pi$) as we do in this paper, but, for instance, the amplitude value of $Q_{e1}$, which is $|Q_{e1}(\Theta=0)|+|Q_{e1}(\Theta=\pi/2)|$. We feel, however, that the our present parametrization has the advantage of
674: simplicity and is sufficiently accurate.
675:
676:
677:
678:
679:
680:
681: \subsection{Parameter study} \label{para}
682:
683: Above we compared the properties of RATs in AMO with those
684: obtained numerically from DDSCAT for a few chosen grain shapes and radiation spectra.
685: To see how general our numerical results are,
686: we attempt a limited parameter study, namely, we study how the properties of
687: RATs vary with the spectrum of the incident radiation for different grain
688: shapes. One can view the AMO formulae as a {\it physically motivated} fit to RATs
689: acting on astrophysical grain with $a_{eff}/\lambda<1$. The parameter study
690: is intended to find out how good is this fit.
691:
692:
693: Fig. \ref{f23} shows $Q_{e1}$ and $Q_{e2}$ for the shape 1 produced by different
694: radiation fields. There the upper panel show that when monochromatic radiation
695: fields of $\lambda/a_{eff}$ increases, the symmetry of $Q_{e1}$ and zeros of
696: $Q_{e2}$ do not change. However, their amplitude decreases. In addition, the
697: symmetric property of $Q_{e1}$ and zeros of $Q_{e2}$ also remain unchanged
698: when being
699: averaged over different radiation spectrum (see the lower panel in
700: Fig. \ref{f23}).
701:
702: Now let consider RAT properties for different irregular grains. The upper panel in Fig. \ref{f23a} shows RATs for different
703: shapes: shape 1* which is a mirror symmetric copy of shape 1, and shapes 4 and 5 are
704: built from 15 and 11 cubic blocks, respectively.
705: We also see clearly that $Q_{e1}$ exhibits the symmetry, and $Q_{e2}$ exhibits
706: the asymmetry that we have already seen with AMO and other grain shapes.
707:
708:
709: From Figs \ref{f23} and \ref{f23a}, it follows that the form of $Q_{e1}, Q_{e2}$
710: for shape 1 is mirror-symmetric to the corresponding
711: RAT components applied to shape 1*. This mirror
712: symmetry is also evident when we compare $Q_{e1}, Q_{e2}$ with those
713: of shapes 4, 5 (see Fig.~\ref{f18}).
714: This implies, similar to AMO, irregular grains may be of right and left
715: helicities. A comparison between Figs~ \ref{f21},
716: \ref{f23}{\it upper}, and \ref{f23a}{\it upper} shows that shapes 1*, 2, 3, 4, 5 are of left helicity, while the shape 1 is
717: of right helicity. Note, that Fig. \ref{f23}{\it lower} clearly shows that the
718: helicity is independent of wavelength and is intrinsic attribute of
719: grains which is associated to their shape. Especially, we can obtain a
720: grain with the opposite helicity by performing a mirror symmetric
721: transformation,
722: which is illustrated by AMO in Fig.~\ref{f6}. We remind the reader, that the
723: correspondence between $Q_{e1}$ and $Q_{e2}$ for AMO and shape 2 and 4 is
724: illustrated by Fig.~\ref{f61}.
725: \begin{figure}
726: \includegraphics[width=0.49\textwidth]{f26a.ps}
727: \includegraphics[width=0.49\textwidth]{f26b.ps}
728: \caption{RATs for the shape 1 corresponding to various $\lambda/a_{eff}$ ({\it Upper
729: Panel}) and
730: RATs averaged over a range of wavelengths ({\it Lower Panel}).}
731: \label{f23}
732: \end{figure}
733:
734: \begin{figure}
735: \includegraphics[width=0.49\textwidth]{f27a.ps}
736: \includegraphics[width=0.49\textwidth]{f27c.ps}
737: \caption{{\it Upper panel:} RATs for different grain shapes. Shape 1* corresponds to mirror-symmetric image of
738: shape 1. Symmetric features of $Q_{e1}$ and zeros of $Q_{e2}$ are clearly
739: found. {\it Lower panel:} The third component of RATs $Q_{e3}$ for different shapes are
740: shown together with that of a spheroid with $e=0.5$ predicted by AMO. An analogy exists between the zeros of $Q_{e2}$ and $Q_{e3}$. Also, the shape of $Q_{e3}$ is not affected by the change of grain helicity.}
741: \label{f23a}
742: \end{figure}
743: For the third RAT component $Q_{e3}$, it exhibits analogous properties with
744: $Q_{e2}$ obviously seen in the lower panel in Fig. \ref{f23a}. Also, for
745: an axisymmetric shape, i.e., spheroid, $Q_{e3}$ is still significant. Note, that its functional form obtained by DDSCAT is very consistent with that predicted by AMO in see \S \ref{amo1} (see dot and dot-dashed line in Fig. \ref{f23a}).
746: Furthermore, we see
747: that it has similar forms for left (shape 2 and 3) and right (shape 1) helical
748: grains (see Fig.~\ref{f23a}{\it lower}). While the dependences of
749: $Q_{e1}$ and $Q_{e2}$ undergo a transformation when shape 1 is substituted
750: by the shape 1*, having the opposite helicity, a comparison of lower and
751: upper panels of Fig.~\ref{f23a} shows that the shape of $Q_{e3}$ component
752: stays the same.
753: This is expected, as
754: $Q_{e3}$ does not depend on the helicity of grains.
755:
756:
757: Also, the $Q_{e3}$ component has zeros
758: at $\Theta=0, \pi$, similar to $Q_{e2}$ (see the lower panel in Fig. \ref{f23a}). However, the anti-symmetry of $Q_{e3}$ is less prominent than for $Q_{e2}$.
759:
760: As we will see that the ratio $Q_{e1}^{max}/Q_{e2}^{max}$ is an important
761: parameter that determines the existence of high-$J$ attractor points. For
762: irregular grains, this ratio is a function of the ratio of wavelength to grain size, as shown in
763: Fig. \ref{f23*} for three irregular grains. The
764: peak of $Q_{e1}^{max}, Q_{e2}^{max}$ is different for different
765: shapes. The form of the curve for sufficiently large ratios of
766: $\lambda/a_{eff}$ can be approximated as $10 \frac{a_{eff}}{\lambda}$. This dependence can be used to
767: reduce the number of DDSCAT calculations necessary for determining the alignment
768: for arbitrary radiation fields.
769:
770:
771: \begin{figure}
772: \includegraphics[width=0.49\textwidth]{f28.ps}
773: \caption{Ratio of $Q_{e1}^{max}, Q_{e2}^{max}$ as function of
774: wavelength to grain size $\lambda/a_{eff}$ for different grain shapes and sizes. The shaded region corresponds to the parameter space in
775: which the high-$J$ attractor point is present for the alignment in respect
776: to the beam direction (see \S \ref{kalign}).}
777: \label{f23*}
778: \end{figure}
779:
780:
781: In general, we observe strong similarities between the plots of RAT components
782: obtained for very different grains. Thus we can expect that the RAT alignment
783: should be similar for such grains. We will discuss the alignment for AMO and irregular grains in \S~ \ref{kalign} and \S~\ref{balign}.
784:
785:
786: \subsection{RATs: Comparison with AMO}\label{sec94}
787:
788: The actual grains are not perfectly reflecting particles and the scattering that
789: they induce cannot be described by
790: geometric optics that we employ for AMO. Therefore the justification of the AMO utility
791: can be obtained via a comparison of the functional form of the torque components obtained for irregular grains with the AMO predictions. Naturally, one should not expect to see
792: the amplitudes of the torques to be the same. Therefore the comparison should be done
793: for the normalized torque components. However, we preserve the ratio of the components.
794:
795:
796: Naturally, our sample of RATs acting on grains studied with DDSCAT is
797: limited. It includes several grain sizes. For instance, for shape 1, we
798: studied for grain sizes of $0.05, 0.08, 0.1$ and $0.2 \mu m$. For other
799: shapes, the size $a_{eff}=0.2 \mu m$ is studied, except the hollow shape 1
800: with $a_{eff}=1 \mu m$. We calculated RATs for the entire spectrum of ISRF
801: corresponding to 21 wavelengths in the range $\lambda=0.1 \mu m$ to $100\mu m$, for
802: shape 1, 2 and 3, and the monochromatic radiation with $\lambda=1.2\mu m$ for
803: other shapes. This provides us with RATs calculated for 130 realizations of
804: grains and radiation fields \footnote{Each realization corresponds to a given grain
805: size and a given wavelength}. This makes it the most extensive sample of RATs
806: studied numerically. It is obvious, that in our paper we cannot present plots
807: of the RATs for all the realizations that we calculated (see e.g. Fig. \ref{f61}). A quantitative comparison based on the deviation testing for normalized RATs of all realizations and AMO
808: will be presented below.
809:
810: We show in Fig. \ref{f23*} that, for irregular grains, the relative
811: amplitude of $Q_{e1}$ versus $Q_{e2}$ changes both with the grain shape and
812: wavelengths. However, our studies in this paper shows that the
813: functional form of the RAT components for all the cases we studied is still
814: well represented by AMO (with different ratio of $Q_{e1}^{max}/Q_{e2}^{max}$, e.g., Fig. \ref{f61}). In other words, while DDSCAT studies of alignment for grains of a
815: few chosen shapes cannot reveal the generic properties of the RAT alignment,
816: revealing the correspondence of the functional dependences of the torques
817: between irregular grains and AMO provides a deep insight into the
818: alignment.
819:
820: Since we are only interested in the functional forms of RATs, let us introduce the mean deviation over $\Theta$ for the components $Q_{e1}$ and $Q_{e2}$ as followings
821: \begin{align}
822: \langle\Delta^{2}\rangle(Q_{ei})&=\frac{1}{\pi( Q_{ei}^{max})^{2}}\int_{0}^{\pi}(Q_{ei}^{DDSCAT}(\Theta)-Q_{ei}^{AMO}(\Theta))^{2} d\Theta,\label{eq82a}
823: \end{align}
824: where $Q_{ei}^{DDSCAT}(\Theta)$ denote $Q_{e1}(\Theta), Q_{e2}(\Theta)$ for irregular grains, $Q_{ei}^{AMO}(\Theta)$ denotes the torque components for AMO in which the relative magnitude are rescaled to have the same ratio $Q_{e1}^{max}/Q_{e2}^{max}$ with each realization of irregular grains\footnote{We note again that, throughout this paper, apart from Appendices B2 and B3, the functional forms of RAT components for AMO corresponds AMO with $\alpha=45^{0}$, and the ratio of their maximum is adjustable.}. In equation (\ref{eq82a}), $Q_{ei}^{max}$ is the maximum of $Q_{e1}$ and $Q_{e2}$, that is chosen the same for both AMO and irregular grains.
825:
826: We perform $\langle\Delta^{2}\rangle$ testing for our sample consisting of $130$ realizations of irregular grain shape, size and wavelength. To see the correspondence of AMO with different grain shape, size and wavelength, in Fig. \ref{chi_test0} we show $\langle\Delta^{2}\rangle$ as a function of $\lambda/a_{eff}$.
827:
828: \begin{figure}
829: \includegraphics[width=0.49\textwidth]{f29a.ps}
830: \includegraphics[width=0.49\textwidth]{f29b.ps}
831: \caption{Figs show $\langle \Delta^{2}\rangle$ testing as function of $\lambda/a_{eff}$ for normalized $Q_{e1}$ (upper panel) and $Q_{e2}$
832: (lower panel) between irregular grain shapes and sizes with AMO.}
833: \label{chi_test0}
834: \end{figure}
835: Fig. \ref{chi_test0}{\it upper} shows a good correspondence for the component $Q_{e1}$ between irregular grains and AMO. The value of $\langle\Delta^{2}\rangle$ ranges from as small as $10^{-3}$ to $2\times 10^{-1}$.
836: In addition, Fig. \ref{chi_test0}{\it lower} shows an extremely good correspondence for the component $Q_{e2}$ between all cases of irregular grains and AMO. The value of for different sizes of shape 1 are nearly the same, but it changes with grain shapes. For instance, shape 2 and 3 have the better correspondence with AMO than shape 1. The corresponding value of $\langle\Delta^{2}\rangle$ is about $2\times10^{-2}$ and $3\times 10^{-2}$ for shape 2, 3 and 1, respectively (see Fig \ref{chi_test0}{\it lower}). For both components and all grain shapes and sizes, $\langle\Delta^{2}\rangle$ do not change so much with respect to wavelength of radiation field for $\lambda/a_{eff}>3$. However, it increases when $\lambda/a_{eff}$ decreases. It is clear that the smaller $\langle\Delta^{2}\rangle$ is, the better correspondence is. Therefore, the observed worse correspondence for the range $\lambda/a_{eff}<3$ can been easily explained in terms of the RAT properties for irregular grains. In fact, following Fig. \ref{f23}{\it upper}, it can be seen that the components $Q_{e1}, Q_{e2}$ for $\lambda/a_{eff}=2.8$ are indeed less symmetric than for $\lambda/a_{eff}=8$ and $15$. In other words, the symmetry of RATs decreases toward small $\lambda/a_{eff}$.
837:
838: Note, that lower values of $\langle\Delta^{2}\rangle$ can be achieved for small $\lambda/a_{eff}$
839: ratios, if $Q_{e1}$ is considered separately for $\Theta$ within ranges
840: $[\pi, \pi/2]$ and $[\pi/2,0]$. This procedure can compensate for the
841: differences in the surface characteristics on the side of the grain towards the
842: light direction and opposite to it. As we discuss in \S \ref{amo2} such
843: differences are natural for less idealized models of helical grains. We do not persue this approach within this paper, prefering a simple model
844: with a reasonable fit to a more complex model that can provide a better fit.
845: Nevertheless, in future, dealing with particular cases, e.g.
846: graphite grains, the approach of considering
847: different ranges of $\Theta$ may prove advantageous.
848:
849:
850: Therefore, our quantitative comparison of the RATs based on $\langle\Delta^{2}\rangle$ for irregular grains with AMO implies that the functional form of RATs obtained
851: in the limit $\lambda\ll a_{eff}$ (AMO) is also valid for RATs in the limit
852: $\lambda\sim a_{eff}$ and $\lambda>a_{eff}$ (for irregular grains). This allows us to use AMO as a
853: representative model for describing RATs of realistic astrophysical grains,
854: e.g. study grain alignment using AMO.
855:
856: %Together with comparison for functional forms of RATs for AMO and irregular grains in \S \ref{sec33}, the $\chi^{2}$ testing gives us quantitatively the good correspondence between AMO and irregular grains.
857:
858: \section{AMO: Definition and Generalizations}\label{amo2}
859:
860: Our studies above show that the model of AMO corresponding to $\alpha=\pi/4$
861: corresponds well to the results of numerical calculations of RATs, provided that we treat $Q_{e1}^{max}/Q_{e2}^{max}$ as an adjustable parameter. Indeed, while we show in Appendix B that this ratio changes with $\alpha$, the range of the ratio variations is lower than for irregular grains (compare Fig. \ref{f5} and Fig \ref{f23*}).
862:
863: This adjustment is natural, as we do not really expect the scattering by
864: irregular dielectric grains to be entirely equivalent to scattering by mirrors
865: of our toy model of a grain.
866:
867: %The important provides us with a definition of ???
868:
869: Our analytical studies are based on AMO, which is provides a simple model
870: of RATs. If necessary,
871: AMO can be trivially generalized by adding additional mirrors to grain
872: surface and accounting for a partial shadowing of the mirrors by the
873: spheroidal grain body. In addition, one can consider not perfectly
874: reflective models, but models with refractive grain body and also refractive
875: plate instead of a mirror. This allows us to vary AMO's properties.
876:
877: An arbitrary attachment of mirrors may make $Q_{e1}$
878: less symmetric. For instance, a grain
879: shown in the upper panel of Fig. \ref{f17} has a mirror attached to one of its surfaces. Naturally,
880: turning this surface towards the beam of radiation produces torques that
881: are different from the case when the mirror is hidden by the elliptical body
882: of the grain.
883:
884: \begin{figure}
885: \includegraphics[width=0.49\textwidth]{f24a.ps}
886: \includegraphics[width=0.49\textwidth]{f24b.ps}
887: \caption{Models of a grain shape consisting of a mirror shadowed partially by
888: the oblate spheroid (upper panel) and a partially damped oscillator
889: attached to the oblate spheroid (lower panel).}
890: \label{f17}
891: \end{figure}
892:
893: In addition, the adopted AMO does not demonstrate account for
894: torques arising from absorption. If, however, one adds to the spheroidal
895: grain a one-dimensional damped oscillator with the axis at an angle to the
896: spheroid symmetry axis (see the lower panel in Fig. \ref{f17}), such a grain
897: would have a non-zero torque component $Q_{e2}$ arising from light absorption.
898: Note, that just adding absorbing component to one of the surfaces of the
899: reflecting mirror would distort the symmetry of torques $Q_{e1}$, which would
900: correspond to RATs calculated for irregular grains for small $\lambda/a_{eff}$
901: ratio.
902:
903:
904: While adding such effects could provide us with better fits of RATs obtained numerically for particular shapes, this would increase the
905: complexity of our model. We do not pursue this path here, therefore. However,
906: we can see that
907: in some cases, e.g. dealing with highly absorbing grains, e.g. graphite
908: grains, one may have to make the corresponding analytical model more
909: sophisticated.
910:
911:
912: Note, that some modifications do not require making the model
913: more complex.
914: Historically, spheroidal grains were used to demonstrate alignment. However,
915: for demonstrating the effects of H$_2$ torques Purcell (1979) considered a
916: ``brick-shaped'' grain. If we want to study the effects of H$_2$ torques
917: for AMO, it is natural to consider not a ``spheroid with a wing'', as
918: we have done in this paper, but a ``brick with a wing''. In terms of AMO this would not change the properties, apart from the value of
919: $Q_{e3}$ torque, which is not important for the alignment, anyhow.
920:
921:
922: Intentionally, our AMO is based on an intuitive macroscopic model.
923: Microscopic models based on
924: the analytical treatment of RATs are possible also. For instance,
925: light scattering by optically active (chiral)
926: sphere was analytically studied by Bohren (1974). The author derived analytical solutions for scattering matrix. Due to the optically activeness, such spherical grain can produce RATs. This model
927: would correspond to the Dolginov (1972) suggestion of quartz grains being spin up
928: and aligned by RATs.
929:
930: An important generalization of AMO is to consider that its inertial
931: properties are given not by a spheroid but by a triaxial ellipsoid. We
932: expect changes to result in difference in dynamics during periods of crossovers. The corresponding effects together with the modification of inertial properties of AMO are considered in Hoang \& Lazarian (2007).
933:
934: All in all, while it is rather easy to make AMO more sophisticated and provide
935: a better fit for RATs within different ranges of $\lambda$ and $a_{eff}$,
936: for the rest of the paper, we adopt a very simple model of AMO. This model
937: provides a reasonable fit to generic properties of
938: RATs acting on actual irregular grains. Thus studying grain alignment with AMO
939: should provide insight into the alignment processes acting upon
940: astrophysical grains.
941:
942:
943:
944:
945:
946: \section{Alignment with respect to light direction}\label{kalign}
947:
948: While the earlier studies dealt with the alignment in respect to magnetic
949: field,
950: in this section we show that RATs can align grains on their own, without any
951: influence of magnetic fields. In this case, the direction of radiation ${\bf k}$ is
952: the axis of alignment. As we further discuss in \S \ref{sec91} such alignment
953: happens in the presence of magnetic field when the rate of precession induced
954: by radiation is faster than the Larmor precession rate.
955: Dealing with this simple case also prepares
956: us for dealing with a more complicated cases of alignment in respect to magnetic field in \S \ref{balign}.
957:
958: \subsection{RATs: spin-up, alignment and precession}
959:
960:
961:
962: To understand the role of the RAT components, we calculate torques
963: that spin up, align, and induce grain precession. A RAT component that acts to spin up grains, $H$, is directed
964: along $\hat{{\bf a}}_{1}$, the component that aligns
965: grains, $F$, is directed along $\hat{\Theta}$, and RAT that causes precession, $G$,
966: is along $\hat{\Phi}$.
967: They are respectively given by (see DW97)
968: \begin{align}
969: H(\Theta, \Phi)&=\mbox{\bQ}_{\Gamma}.\mbox{\be}_{1}(\Theta,\Phi)\mbox{cos}\Theta+
970: \mbox{\bQ}_{\Gamma}.\mbox{\be}_{2}(\Theta,\Phi)\mbox{sin}\Theta\mbox{cos}\Phi\nonumber\\
971: &+\mbox{\bQ}_{\Gamma}.\mbox{\be}_{3}(\Theta,\Phi)\mbox{sin}\Theta\mbox{sin}\Phi\label{eq21},\\
972: F(\Theta, \Phi)&=-\mbox{\bQ}_{\Gamma}.\mbox{\be}_{1}(\Theta,\Phi)\mbox{sin}\Theta+
973: \mbox{\bQ}_{\Gamma}.\mbox{\be}_{2}(\Theta,\Phi)\mbox{cos}\Theta\mbox{cos}\Phi\nonumber\\
974: &+\mbox{\bQ}_{\Gamma}.\mbox{\be}_{3}(\Theta,\Phi)\mbox{cos}\Theta\mbox{sin}\Phi\label{eq22},\\
975: G(\Theta, \Phi)&=-\mbox{\bQ}_{\Gamma}.\mbox{\be}_{2}(\Theta,\Phi)\mbox{sin}\Phi+\mbox{\bQ}_{\Gamma}.\mbox{\be}_{3}(\Theta,\Phi)\mbox{cos}\Phi\label{eq23}.
976: \end{align}
977:
978: On the other hand, following equation (\ref{eq2}), RATs at a precession angle $\Phi$ are related to
979: which at $\Phi=0$ via
980: \begin{align}
981: \mbox{\bQ}_{\Gamma}.\mbox{\be}_{1}(\Theta,\Phi)&=Q_{e1}(\Theta,0),\label{eq24}\\
982: \mbox{\bQ}_{\Gamma}.\mbox{\be}_{2}(\Theta,\Phi)&=Q_{e2}(\Theta,0)\mbox{cos}\Phi-Q_{e3}(\Theta,0)\mbox{sin}\Phi,\label{eq25}
983: \\
984: \mbox{\bQ}_{\Gamma}.\mbox{\be}_{3}(\Theta,\Phi)&=Q_{e2}(\Theta,0)\mbox{sin}\Phi+Q_{e3}(\Theta,0)\mbox{cos}\Phi.\label{eq26}
985: \end{align}
986:
987: Plugging the above equations into (\ref{eq21}), (\ref{eq22}) and
988: (\ref{eq23}) we get
989: \begin{align}
990: G(\Theta, \Phi)&=Q_{e3}(\Theta,0)\label{eq27},\\
991: H(\Theta, \Phi)&=Q_{e1}(\Theta,0)
992: \mbox{ cos}\Theta + Q_{e2}(\Theta,0)\mbox{
993: sin}\Theta,\label{eq28} \\
994: F(\Theta, \Phi)&=-Q_{e1}(\Theta,0)\mbox{
995: sin}\Theta + Q_{e2}(\Theta,0)\mbox{ cos}\Theta.\label{eq29}
996: \end{align}
997:
998: If for $G(\Theta, \Phi)$ we are mostly interested in its amplitude, the
999: functional form of $F(\Theta, \Phi)$ and $H(\Theta, \Phi)$ is essential for
1000: grain alignment. The problem is that $F(\Theta, \Phi)$ and $H(\Theta, \Phi)$
1001: as well as their counterparts obtained in the presence of magnetic field (see
1002: equations \ref{eq61} and \ref{eq62}), vary substantially from one grain to another. For AMO
1003: different grains correspond to different ratio $Q_{e1}^{max}/Q_{e2}^{max}$. As we mentioned in \S \ref{sec43} and \S \ref{para}, irregular grains are different in terms of RATs for the radiation of different wavelengths and
1004: different grain sizes. However, for both AMO and irregular grains, the generic properties of the RAT components (i.e., symmetry of $Q_{e1}$, as well as the asymmetry and zeroes of $Q_{e2}, Q_{e3}$) is always unchanged. Therefore, unlike $Q_{e1}$ and $Q_{e2}$, the components
1005: $F(\Theta, \Phi)$ and $H(\Theta, \Phi)$ do not demonstrate a universal
1006: behavior and play an auxiliary role in our study.
1007:
1008:
1009: We see that the precessing torque depends only on the third component
1010: $Q_{e3}(\Theta,0)$, while the aligning and
1011: spinning torques are related to two first components, namely
1012: $Q_{e1}(\Theta,0), Q_{e2}(\Theta,0)$. We note that the functions $F, G, H$
1013: are the functions of only variable $\Theta$, in this case.
1014:
1015:
1016: For AMO with $\alpha=45^{0}$, substituting analytical expressions $f=f_{\pi/2}$ and $g$ given by equations (\ref{eq16})
1017: and (\ref{eq17b}) into equations (\ref{eq14}) and (\ref{eq15}), we get
1018:
1019: \begin{align}
1020: Q_{e1}(\Theta, 0)&=\frac{16 l_{1} n_{1}n_{2}|n_{2}|}{3\lambda}(5\mc^{2}\Theta-2),\label{eq30a}\\
1021: Q_{e2}(\Theta, 0)&=\frac{40 l_{1} n_{1}n_{2}|n_{2}|}{3\lambda}\ms2\Theta(1.191+0.1382\mcs\Theta).\label{eq31a}
1022: \end{align}
1023: Therefore, equations (\ref{eq27})-(\ref{eq29}) becomes
1024: \begin{align}
1025: G(\Theta)&=-\frac{2e a (1-s)}{\lambda} K(\Theta, e)\ms 2\Theta \label{eq30},\\
1026: H(\Theta)&=\frac{8 l_{1}n_{1}n_{2}|n_{2}|}{3\lambda}\mc\Theta[1+6.91\mss\Theta\nonumber\\
1027: &+\mcs\Theta(5+1.382\mss\Theta)],\label{eq31} \\
1028: F(\Theta)&=\frac{8 l_{1}
1029: n_{1}n_{2}|n_{2}|}{3\lambda}\ms\Theta[-1+6.91\mcs\Theta\nonumber\\
1030: &+1.382\mbox{cos}^4\Theta+5\mss\Theta].\label{eq32}
1031: \end{align}
1032:
1033: \subsection{Simplified Treatment of Crossovers} \label{cross}
1034: In general, the maximal inertia axis ${\bf a}_{1}$ of our
1035: model grain can precess about the vector of the angular momentum
1036: ${\bf J}$. In the present paper, however, for the sake of simplicity,
1037: we assume a perfect internal alignment, i.e. ${\bf J} \|
1038: {\bf a}_{1}$. This assumption coincides with that in DW97 and
1039: can be justified by the high efficiency of the internal relaxation within a
1040: wobbling grain.
1041: This relaxation stems from the Barnett relaxation discovered by Purcell (1979)
1042: and/or nuclear relaxation introduced in LD99b.
1043: However, these relaxation processes provide a good
1044: coupling only when $J \gg J_{th}\approx (kT_{d} I_{1})^{1/2}$ where $T_{d}$
1045: is the dust temperature, $I_{1}$ is the maximal moment of inertia of grain (Lazarian 1994),
1046: i.e. when a grain rotates with suprathermal velocities. This condition
1047: is not satisfied as a grain approaches crossover points,
1048: i.e. as $J\rightarrow J_{th}$.
1049:
1050: We adopt below a simplified treatment of crossovers, which is different, however, from the treatment of crossovers in DW97. There it was assumed that
1051: ${\bf J} \| {\bf a}_{1}$ up to the moment of the angular velocity getting zero.
1052: After that DW97 assumed that ${\bf J}$ changes its direction to the opposite,
1053: while the direction of ${\bf a}_{1}$ is preserved.
1054: Such a model of crossovers differs from the earlier work on crossovers, which
1055: suggests that, as the angular velocity goes to zero, the grain undergoes a flip,
1056: i.e., ${\bf J}$ preserves its direction, while the direction of ${\bf a}_{1}$
1057: changes to the opposite (Spitzer \& McGlynn 1979; Lazarian \& Draine 1997). Therefore, in what follows, we adopt a model in which ${\bf J} \| {\bf a}_{1}$
1058: up to a crossover; at the crossover ${\bf a}_{1}$ flips, while the direction of
1059: ${\bf J}$ is preserved. This makes the dynamics of grains very different from
1060: that in DW97\footnote{We may mention parenthetically, that the dynamics that
1061: we get has similarities to the dynamics of grains with thermal fluctuations
1062: accounted for, as in WD03. There, however, it was interpreted as a new
1063: effect related to thermal fluctuations.}. Although we accept that our
1064: model is not precise when $J \sim J_{th}$, we claim the our simplified model
1065: is ``roughly true''. The latter point is justified for regular crossovers
1066: (i.e., crossovers without thermal fluctuations) in \S \ref{cross}.
1067:
1068: \subsection{Stationary and singular points}
1069:
1070: \subsubsection{Equations of motion}
1071: To find out whether grains can be aligned with respect to ${\bf k}$, we need to
1072: find stationary states for grain motion exerted by RATs.
1073: We start with equations of motion
1074:
1075: \begin{align}
1076: \frac{dJ}{dt}&=M H(\Theta)-J,\label{eq33}\\
1077: \frac{d\Theta}{dt}&=M \frac{ F(\Theta)}{J},\label{eq34}
1078: \end{align}
1079: where time $t$ and angular momentum $J$ are scaled in the units of the gas
1080: damping time $t_{gas}$ (see Table \ref{tab1}) and the thermal angular
1081: momentum $I_{1}\omega_{T}$. Here $I_{1}$ is the maximal
1082: moment of inertia and $\omega_{T}=(\frac{kT_{gas}}{I_{1}})^{1/2}$ is
1083: the thermal angular velocity with $T_{gas}$ is the gas temperature. In
1084: equations (\ref{eq33}) and (\ref{eq34})
1085: \bea
1086: M=\frac{\gamma u_{rad}\lambda a_{eff}^{2}t_{gas}}{2I_{1}\omega_{T}}
1087: \ena
1088: contains important parameters of the physical problem. There $u_{rad}$ is the energy density of
1089: radiation field, $\lambda$ is the wavelength. In the present paper, for our estimates, we use the values
1090: provided in Table \ref{tab1} for
1091: interstellar medium. It will be explicitly stated in case other values of these parameters are used.
1092: \begin{table}
1093: \caption{Physical parameters for diffuse ISM }
1094: \begin{displaymath}
1095: \begin{array}{rrrr} \hline\hline\\
1096: \multicolumn{1}{c}{\it Definitions} & \multicolumn{1}{c}{\it Typical~ Values} \\[1mm]
1097: \hline\\
1098: {\rm Gas~ density}& {\rm n=30~ cm^{-3}}\\[1mm]
1099: {\rm Gas~temperature}& {\rm T_{gas}=100~K}\\[1mm]
1100: {\rm Gas~damping~time}& {\rm t_{gas}=4.6\times
1101: 10^{12}(\frac{\hat{\rho}}{\hat{n}\hat{T}_{gas}^{1/2}}) a_{-5}~ s}\\[1mm]
1102: {\rm Dust~temperature}& {\rm T_{d}=20~K}\\[1mm]
1103: {\rm Anisotropy~degree}&{\rm \gamma=0.1}\\[1mm]
1104: {\rm Grain~ size}&{\rm a_{eff}=0.2 \mu m}\\[1mm]
1105: {\rm Mean~wavelength}&{\rm \overline{\lambda}=1.2~ \mu m}\\[1mm]
1106: {\rm Mean~density~of~ISRF}&{\rm {u}_{rad}=8.64 \times 10^{-13}~ erg~ cm^{-3}}\\[1mm]
1107: \\[1mm]
1108: \hline\hline\\
1109: \end{array}
1110: \end{displaymath}
1111: The normalized values that we also use
1112: are $\hat{T}_{gas}=T_{gas}/100~ K,~\hat{n}=n/30~ g~ cm^{-3}$, and
1113: $a_{-5}=a_{eff}/10^{-5}~ cm $, and normalized grain density
1114: $\hat{\rho}=\rho/3~ g~ cm^{-3}$.
1115: \label{tab1}
1116: \end{table}
1117:
1118: \subsubsection{Stationary points}
1119:
1120: The stationary point is determined by setting the equations of
1121: motion (\ref{eq33}) and (\ref{eq34}) equal to zero. This point may be an attractor or a repellor point. We remind the
1122: reader that for a stationary point to be an attractor point, it requires (see Appendix C)
1123: \begin{align}
1124: \left. \frac{d F(\Theta)}{H(\Theta) d\Theta}\right|_{\Theta=\Theta_{s}} &<1,\label{eq44}\\
1125: \left. H(\Theta)
1126: \frac{dF(\Theta)}{d\Theta}\right|_{\Theta=\Theta_{s}}&<\left. F(\Theta)\frac{d H(\Theta)}{d\Theta}\right|_{\Theta=\Theta_{s}}.\label{eq45}
1127: \end{align}
1128:
1129:
1130: At the stationary points, $ F(\Theta_{s})=0$, so equations (\ref{eq44}) and (\ref{eq45}) reduce to
1131: \bea
1132: \left. \frac{d F(\Theta)}{d\Theta}\right|_{\Theta=\Theta_{s}}<0,\mbox{ for } H(\Theta_{s})>0,\label{eq46}
1133: \ena
1134: and
1135: \bea
1136: \left. \frac{d F(\Theta)}{d\Theta}\right|_{\Theta=\Theta_{s}}>0,\mbox{ for } H(\Theta_{s})<0.\label{eq47}
1137: \ena
1138:
1139: Consider that initially a grain has the maximal inertia axis ${\bf a}_{1}$ parallel to
1140: ${\bf k}$, i.e., $\Theta_{0}=0$, then $ F(\Theta_{0})=0, H(\Theta_{0})=Q_{e1}(\Theta_{0})$. These assumptions are
1141: usually used to estimate the grain angular velocity induced by RATs (see
1142: DW97; Cho \& Lazarian 2005). Therefore, the equations of motion become
1143: \begin{align}
1144: \frac{d\Theta}{dt}&=0,\\
1145: \frac{dJ}{dt}&=M Q_{e1}-J.
1146: \end{align}
1147: Solutions for the above equations are easily found,
1148: \begin{align}
1149: \Theta&=0,\label{eq38}\\
1150: J&=M Q_{e1}(\Theta)+(J_{0}-M Q_{e1}(\Theta_{0}))e^{-t},\label{eq39}
1151: \end{align}
1152: where $J_{0}$ is the initial angular momentum of grains.
1153: Equation (\ref{eq39}) shows that as $t \gg 1, J =M Q_{e1}(\Theta=0)$. In other words, grains
1154: initially parallel to ${\bf k}$ are spun up by RATs and aligned with ${\bf
1155: k}$ at angular momentum $J_{s}=M Q_{e1}(\Theta_{s}=0)$, regardless of its initial angular momentum $J_{0}$. However, the particular point of phase
1156: trajectory map may not provide stable orientation of the grain, i.e., may not be
1157: attractor points. \footnote{The shaded area with diagonal lines in Fig \ref{f13} corresponds to the existence of such points. We see that many wavelengths the condition is not satisfied.}
1158:
1159: For $\ms\Theta_{0} \ne 0$ , substituting $ F(\Theta), H(\Theta)$ from
1160: equations (\ref{eq31}) and (\ref{eq32}) into equations (\ref{eq33}) and (\ref{eq34}), we obtain
1161: \begin{align}
1162: \frac{dJ}{dt}&=\frac{8 l_{1} n_{1}n_{2}|n_{2}|}{3\lambda}M\mc\Theta[1+6.91\mss\Theta\nonumber\\
1163: &+\mcs\Theta(5+1.382\mss\Theta)]-J,\label{eq40}\\
1164: \frac{d\Theta}{dt}&=\frac{8 l_{1} n_{1}n_{2}|n_{2}|M}{3\lambda J}\ms\Theta[-1+6.91\mcs\Theta\nonumber\\
1165: &+1.382\mbox{cos}^4\Theta+5\mss\Theta].\label{eq41}
1166: \end{align}
1167: Hence, we can easily find a stationary point\footnote{Another stationary point $\Theta=\pi, J<0$ is forbidden since $J$ can not be negative. Note, that the analysis in this section is applied for the right helical grain, i.e., $n_{1}n_{2}>0$.}
1168: \begin{align}
1169: \Theta_{s}&= 0,\label{eq42}\\
1170: J_{s}&=\frac{48 l_{1} n_{1}n_{2}|n_{2}|M}{3\lambda}.\label{eq43}
1171: \end{align}
1172: From equations (\ref{eq31}) and (\ref{eq32}) we get
1173: \bea
1174: \left. \frac{d F(\Theta)}{d\Theta}\right|_{\Theta_{s}=0}=\frac{90.232 l_{1}
1175: n_{1}n_{2}|n_{2}|}{3\lambda},\label{eq48}
1176: \ena
1177: and $H(\Theta_{s}=0)>0$. As a result, the stationary point $\Theta_{s}=0$ is a repellor point.
1178:
1179: In a general case, $ F(\Theta), H(\Theta) $ are given by equations (\ref{eq28}) and (\ref{eq29}). Thus, substituting
1180: them into equations (\ref{eq46}) and (\ref{eq47}), we get a criteria for a stationary point
1181: $\Theta_{s}$ being an attractor point, namely,
1182: \bea
1183: \frac{-Q'_{e1}\ms\Theta_{s}+Q'_{e2}\mc\Theta_{s}}{Q_{e1}\mc\Theta_{s}+Q_{e2}\ms\Theta_{s}} <1,\label{eq50}
1184: \ena
1185: where $Q'_{e1}=\frac{dQ_{e1}}{d\Theta},Q'_{e2}=\frac{dQ_{e2}}{d\Theta}$. However, our
1186: approximate formulae above catches the phenomenon correctly and we still get
1187: repellor points for $\Theta_s=0$ with the exact expressions for AMO.
1188:
1189: As the range for ratios of $Q_{e1}/Q_{e_2}$ for an arbitrarily chosen irregular grain
1190: may be different from AMO, for some irregular grains we still may have attractor
1191: points for $\Theta_s=0$, as it is shown in Fig~\ref{f13} for shape~1
1192: subjected to interstellar radiation field. There the relevant data corresponds
1193: to $\psi=0$, as in this situation the direction of magnetic field and the light coincide
1194: and thus our considerations about the alignment in respect to ${\bf k}$ are applicable.
1195:
1196: \subsubsection{Singular points: crossovers}
1197:
1198: When $J=0$, we have a singular point (see equation \ref{eq34}).
1199: In terms of our simplified
1200: equations it corresponds to a crossover.
1201: As we discussed earlier, a crossover is a period in which a
1202: grain spins down to the point that the component of angular momentum
1203: parallel to ${\bf a}_{1}$ gets zero. In the presence of strong internal
1204: relaxation that tends to align ${\bf J}$ and ${\bf a}_{1}$, this means that
1205: $J$ should get small (Spitzer \& McGlynn 1979). Our equations above are
1206: derived in the assumption of ${\bf J}\| {\bf a}_{1}$ and therefore can not
1207: treat grain crossovers (cf. \S6~\ref{cross}). However, they can still trace the grain
1208: dynamics as the grain phase trajectory approaches the crossover and $J \to
1209: J_{th}$. Assuming that initial angular momentum of the grain $J_{0} \gg J_{th}$, we disregard the difference
1210: between $J=J_{th}$ and $J=0$. Thus, from equation~(\ref{eq34}) it follows that, in order to have a physical crossover,
1211: grains must have $J =0$, and the aligning
1212: torque $F(\Theta)$ must be zero. The latter condition is
1213: naturally satisfied since we found that $ F(\Theta)=0$ at stationary
1214: points. The former one is satisfied as RATs act to decelerate the
1215: grain rotation. Indeed, equation~(\ref{eq32}) shows that $ F(\Theta)>0$, i.e., it acts to increase $\Theta$, for every angles
1216: $\Theta<\pi$. Yet, equation (\ref{eq31}) shows that $ H(\Theta)<0$ for $\mbox{cos }\Theta <0$ and $ H(\Theta) >0$ for
1217: $\mbox{cos }\Theta >0$. Therefore, if initially the angular momentum of grains makes an
1218: angle $\mc \Theta <0$, then their angular momentum is
1219: decreased due to $ H(\Theta)< 0$; so grains approach directly to the state of $J =0$ (see
1220: the upper panel in Fig. \ref{f8} for right helical grain).
1221: On the other hand, for grains which initially have $\mc \Theta >0$,
1222: the aligning torque acts
1223: to increase $\Theta$, while the spinning torque increases their angular
1224: momentum. Eventually, grains attain the angle $\mc\Theta<0$ for which the
1225: spinning torque changes the sign, so they are decelerated to the state of $J=0$ (see
1226: Fig. \ref{f8} for phase maps of left and right helical grains).
1227:
1228: One can observe that
1229: when grains get to the crossover with very low $J$, their maximal inertia axis
1230: $\ma_{1}$ flips with respect to ${\bf J}$ to enter the opposite flipping
1231: state, i.e., grains flip from the upper to lower panel in Fig. \ref{f8}{\it upper}, for
1232: instance. Right after that, grains flip back to the initial state (upper
1233: panel). This back and forth flipping process takes place frequently. As a
1234: result, the crossover point, in our approach, can be treated as the attractor
1235: point at zero angular momentum, hereafter, called low-$J$ attractor
1236: point\footnote{In Hoang \& Lazarian (2007), we will show that zero angular
1237: momentum attractor points become attractor points at thermal angular momentum
1238: due to thermal wobbling}, to be distinguished from high-$J$ attractor points.
1239:
1240:
1241:
1242: \subsection{RATs: alignment by one component} \label{sec4p4}
1243:
1244: The grain alignment for the case at hand is uniquely related to two
1245: components $Q_{e1}$ and $Q_{e2}$. To understand which component causes the
1246: alignment, let us study the role of $Q_{e1}$ and $Q_{e2}$ separately.
1247:
1248: If $Q_{e2}=0$, then $\frac{dQ_{e2}}{d\Theta}(\Theta_{s}=0, \pi)=0$, equation
1249: (\ref{eq50}) for attractor points becomes
1250: \bea
1251: Q_{e1}(\Theta_{s}=0)>0,\label{eq51} \\
1252: Q_{e1}(\Theta_{s}=\pi)<0.\label{eq51*}
1253: \ena
1254: According to AMO, $Q_{e1}(\Theta_{s}=0, \pi)=\frac{16 l_{1} n_{1}n_{2}|n_{2}|}{3\lambda}(5\mbox{cos}^2\Theta_{s}-2)=1>0$. As a result, we
1255: expect that the stationary point $\Theta_{s}=0$ is an attractor point,
1256: i.e., grains are perfectly aligned with respect to ${\bf k}$. While, the stationary point $\Theta_{s}=\pi$ is a repellor point.\\
1257:
1258: For the case $Q_{e1}=0$, besides stationary points $\Theta_{s}=0, \pi$ as
1259: discussed above, there is another stationary point corresponding to
1260: $\Theta_{s}=\pi/2$ (see equation \ref{eq29}). For $\Theta_{s}=0, \pi$, equation (\ref{eq50}) reduces to
1261: \bea
1262: \frac{dQ_{e2}}{d\Theta}(\Theta_{s}=0)<0, \label{eq52} \\
1263: \frac{dQ_{e2}}{d\Theta}(\Theta_{s}=\pi)>0. \label{eq52*}
1264: \ena
1265: For AMO with $\alpha=45^{0}$, thus, $Q_{e2}=\frac{40 l_{1} n_{1}n_{2}|n_{2}|}{3\lambda}\ms 2\Theta (1.191+0.1382\mcs\Theta)$, we have $
1266: \frac{dQ_{e2}}{d\Theta}(\Theta_{s}=0, \pi)=\frac{106.2 l_{1} n_{1} n_{2}|n_{2}|}{3\lambda}\mc2\Theta_{s} >0$. It means
1267: that the condition (\ref{eq52}) is not satisfied, but equation (\ref{eq52*}) is fulfilled. In other words, the stationary point $\Theta_{s}=0$ is a repellor points, while the stationary point $\Theta_{s}=\pi$ is an attractor point.
1268:
1269: For a particular stationary point $\Theta_{s}=\pi/2$, equation (\ref{eq52}) in this case becomes
1270: \bea
1271: \frac{Q'_{e2}\mc\Theta_{s}}{Q_{e2}\ms\Theta_{s}}=0<1.\label{eq53}
1272: \ena
1273: Therefore, the stationary point $\Theta_{s}=\pi/2$ is, indeed, an
1274: attractor point. Note, equation (\ref{eq53}) which indicates that
1275: $\Theta_{s}=\pi/2$ is the attractor point is valid for an arbitrary form of $Q_{e2}$. These results show that the component $Q_{e2}$ acts to align
1276: grains in the direction perpendicular to the radiation beam for arbitrary grain shapes.
1277:
1278:
1279: \subsection{Phase trajectories}
1280:
1281: To show our predictions from the above analysis, we construct phase trajectory maps in which each RAT component acts separately. The difference of our
1282: phase maps from those in DW97 is that here we present the alignment in
1283: respect to ${\bf k}$ rather than to magnetic field. In addition, we
1284: treat crossovers differently from how they are treated in DW97. With the
1285: exception of Figs \ref{f12*} and \ref{f15}, to avoid over-crowding of our phase maps we do
1286: not draw arrows with $t_{gas}$ intervals.
1287:
1288:
1289:
1290: Throughout the present paper, the phase trajectory map represents the evolution of $J$ and angle between ${\bf J}$ and ${\bf k}$ or ${\bf B}$
1291: \footnote{We note that the angle between ${\bf J}$ and ${\bf k}$ or ${\bf B}$ is shown rather
1292: than the angle between ${\bf a}_{1}$ and ${\bf k}$ or ${\bf B}$ as in
1293: DW97.}. Each phase map
1294: has upper and lower panels, which correspond to ${\bf J}$ parallel and anti-parallel to
1295: ${\bf a}_{1}$, respectively. Also, a circle denotes an attractor point and a
1296: cross denotes a repellor point.
1297:
1298: Moreover, the title of the trajectory map presents information about the model (AMO) or grain shape, size for irregular grains. For trajectory maps labeled with AMO on their titles, it is a default that the RAT components are taken from the exact calculations of equations (\ref{eq8}-\ref{eq10}) for $\alpha=45^{0}$ which have $Q_{e1}^{max}/Q_{e2}^{max}=1.2$. For the maps of AMO in which the ratio $Q_{e1}^{max}/Q_{e2}^{max}$ is explicitly shown, it means that the functional forms of RATs are similar for the case $\alpha=45^{0}$, but the relative amplitudes are rescaled to the shown value.
1299:
1300: \begin{figure}
1301: \includegraphics[width=0.49\textwidth]{f10a.ps}
1302: \hfill
1303: \includegraphics[width=0.49\textwidth]{f10b.ps}
1304: \caption{Phase trajectory maps of grains in the absence of magnetic
1305: field corresponding to
1306: $Q_{e2}=0$ ({\it Upper Panel}) and
1307: $Q_{e1}=0$ ({\it Lower Panel}). The map in upper panel exhibits two attractor
1308: points A, B corresponding to a perfect alignment as pointed out by the
1309: analysis; E is a low-$J$ attractor point and two repellor points C, D.
1310: The lower panel shows two repellor
1311: points A, B and two attractor points C, D in which D corresponds to ``wrong''
1312: alignment.}
1313: \label{f7}
1314:
1315: \end{figure}
1316:
1317: The map for the case when only $Q_{e1}$ is present has three attractor points:
1318: A, B correspond to $\Theta_{s}=0$, and E corresponds to zero of $Q_{e1}$ at $\mbox{cos}\Theta=-0.6$ (see the upper panel in
1319: Fig. \ref{f7}). Therefore, $Q_{e1}$ acts to align grains with
1320: ${\bf J}$ parallel to ${\bf k}$, i.e, ${\bf a}_{1}$ parallel or anti
1321: parallel to ${\bf k}$ . In addition, there are two repellor points C,
1322: D. These features show clearly our predictions obtained with AMO.
1323:
1324: When $Q_{e1}=0$, Fig. \ref{f7} (lower panel) shows that the phase trajectory map has one attractor point D at
1325: $\Theta_{s}=\pi/2$, which corresponds to a perpendicular alignment, one
1326: low-$J$ attractor point C at $\Theta_{s}=\pi$, and two repellor points A, B as predicted by our above analysis. So
1327: $Q_{e2}$ can align some grains with the maximal inertia axis ${\bf a}_{1}$ perpendicular to the light direction. However, we
1328: discussed in \S 2 that $Q_{e2}$ is equal zero at the attractor point
1329: $\Theta_{s}=\pi/2$, while $Q_{e1}$ is different to zero at this point; thus the alignment with the long axes parallel to the direction of light does not
1330: occurs\footnote{We checked that even the addition of $Q_{e1}$ with an
1331: amplitude of $10^{-9}$ of the amplitude of $Q_{e2}$ can destroy such
1332: an alignment.} for the case of alignment with respect to ${\bf k}$ (see \S~\ref{sec5}).
1333:
1334: \subsection{RATs: alignment by joint action of torques, role of
1335: $Q_{e1}^{max}/Q_{e2}^{max}$ ratio}\label{qratio}
1336:
1337: So far we have dealt with the alignment by one component of RATs, and pointed out
1338: that $Q_{e1}$ acts to align grains with long axes $\perp$ to ${\bf k}$, while $Q_{e2}$
1339: can produce the alignment for some grains with long axes $\|$ to ${\bf k}$. In reality, grains are
1340: driven simultaneously by both components, so the alignment of grains depends
1341: on which component of RATs is predominant.
1342:
1343: As discussed in previous sections, for AMO, we have the stationary point
1344: which is determined by equations (\ref{eq42}) and (\ref{eq43}). Also, we have analytically shown that the
1345: stationary point $\Theta_{s}= 0$ ($\pi$ for left helical AMO) is the repellor point. To test our expectation, we produce phase trajectory maps of grains driven
1346: by RATs with the precessing, spinning-up and aligning components given by equations (\ref{eq30})-(\ref{eq32}). Fig. \ref{f8} shows the maps for right and left helical AMOs. It can be
1347: seen that the phase maps have two stationary points which positions are exactly given by analytical expressions
1348: (\ref{eq42}) and (\ref{eq43}). Among these stationary points, there are two repellor points
1349: A, B at $\Theta_{s}=0$ (or $\pi$ for lower panel), and one low-$J$ attractor point C at $\Theta_{s}=\pi$ (or $\Theta=0$ for lower panel).
1350: In addition, we see that for the right and left helical AMOs, the trajectory maps
1351: are related through a transformation: $\Theta \to \pi-\Theta$.
1352:
1353: \begin{figure}
1354: \includegraphics[width=0.49\textwidth]{f11a.ps}
1355: \hfill
1356: \includegraphics[width=0.49\textwidth]{f11b.ps}
1357: \caption{Upper and lower panels show the phase trajectory maps for grains with right and
1358: left helicity, respectively. Figs show that grains have a perfect alignment
1359: with respect to ${\bf
1360: k}$ with one low-$J$ attractor point C at $\Theta=\pi$ (upper panel) and
1361: $\Theta=0$ (lower panel). Also, two repellor points A,
1362: B are shown as predicted in our analysis. The maps for upper and lower panels are
1363: mirror-symmetric.}
1364: \label{f8}
1365: \end{figure}
1366:
1367: We have seen that , in both cases of grain alignment by one component and joint actions of
1368: components, the important features (e.g., attractor and repellor points) present in the trajectory maps that are constructed with exact RATs are consistent
1369: with the predictions based on the approximate formulae using the fitting
1370: function $f_{\pi/2}$ and $g$. It indicates that, though the fitting functions
1371: do not give the best fit in some particular angles $\Theta$, they can
1372: provide us, intuitively, the alignment property of grains by RATs.
1373:
1374:
1375: We found in the case of alignment by one component that if only $Q_{e1}$ is at
1376: work, it aligns grains with two attractor points with $J \gg J_{th}$, $\Theta=0$, i.e., high-$J$ attractor points. In contrast, when only
1377: $Q_{e2}$ acts, those points are repellor points; Instead $Q_{e2}$ produces an attractor point
1378: at $\Theta=\pi/2$. When both components act simultaneously, are there high-$J$
1379: attractor points? Obviously, we may conjecture that if $Q_{e1}$ is predominant
1380: over $Q_{e2}$, then high-$J$ attractor points should still appear. Otherwise,
1381: high-$J$ stationary points are repellor points.
1382:
1383: For our default AMO with $\alpha=\pi/4$, we have predicted that the stationary points $\Theta=0$ are
1384: always repellor points. This can also be understood in terms of the ratio of $Q_{e1}^{max}/Q_{e2}^{max} \sim 1$, for this case, i.e, the
1385: dominant criteria is not satisfied (see equations \ref{eq18} and \ref{eq19}).
1386:
1387: To study when we have high-$J$ attractor points, let us write RATs for AMO in
1388: a simplified form
1389: \begin{align}
1390: Q_{e1}&=\frac{Q_{e1}^{max}}{3}(5\mcs\Theta-2),\label{eq54}\\
1391: Q_{e2}&=Q_{e2}^{max}\ms2\Theta,\label{eq55}
1392: \end{align}
1393: where $Q_{e1}^{max}, Q_{e2}^{max}$ are maximal values of $Q_{e1}, Q_{e2}$.
1394:
1395: Clearly, stationary points for this model are $\Theta_{s}=0$ because aligning
1396: torque $F(\Theta_{s})=-Q_{e1}\ms\Theta_{s}+Q_{e2}\mc\Theta_{s}=0$ at
1397: $\Theta_{s}=0$.\\
1398: Using the criteria of an attractor point (i.e., equation \ref{eq50}), it follows that
1399: $\Theta_{s}=0$ is an attractor point if
1400: \bea
1401: \frac{1}{Q_{e1}}\frac{dQ_{e2}}{d\Theta}(\Theta_{s}=0)<1,\label{eq56}
1402: \ena
1403: Plugging equations (\ref{eq54}) and (\ref{eq55}) into equation (\ref{eq56}), we get
1404: \bea
1405: Q_{e1}^{max}> 2 Q_{e2}^{max}.\label{eq57}
1406: \ena
1407: From Fig. \ref{f5} and equation (\ref{eq57}), it follows
1408: that for the original AMO, i.e., AMO in which the relative magnitude of RAT component is not rescaled yet, the maximum of the ratio $max[Q_{e1}^{max}/Q_{e2}^{max}]$ is $1.3$, and
1409: therefore the
1410: stationary points $\Theta_{s}=0$ (corresponding to high $J$)
1411: are always repellor points. However, as we have discussed earlier, their relative magnitude is adjustable. Thus, AMO can produce the phase map with high attractor points for $\psi=0^{0}$, provided that it satisfies equation (\ref{eq57}). Note, that the criteria (\ref{eq57}) is only applicable for the case of alignment with respect to ${\bf k}$ or $\psi=0^{0}$ (i.e., ${\bf k}\| {\bf B}$). For an arbitrary angle $\psi$, the criteria is shown in the lower panel of Fig.~\ref{f13}. There it can be seen that for some irregular grains (i.e. shape~1, ISRF)
1412: $\Theta_s=0$ can correspond
1413: to high attractor points.
1414:
1415: \subsection{Alignment for irregular grains}
1416:
1417: Similar to the AMO case, we consider first the alignment of dust in
1418: respect to the radiation direction ${\bf k}$. In particular, to compare the
1419: action of RATs for AMO and an irregular grain, we consider the alignment that
1420: is induced by individual torque components.
1421:
1422: To calculate the phase map, we use again the parameters from
1423: Table~{\ref{tab1}. We exemplify the alignment for irregular grains using
1424: shape 1, which shows the maximum deviations from AMO predictions.
1425:
1426: Fig. \ref{f24} shows the phase trajectory maps of an irregular
1427: grain (shape 1) driven by RATs calculated by DDSCAT with either $Q_{e2}$ or $Q_{e1}$ taken to be
1428: zero.
1429:
1430: \begin{figure}
1431: \includegraphics[width=0.49\textwidth]{f30a.ps}
1432: \hfill
1433: \includegraphics[width=0.49\textwidth]{f30b.ps}
1434: \caption{Phase trajectory maps in the absence of magnetic field corresponding to
1435: $Q_{e2}=0$ ({\it Upper Panel}) and $Q_{e1}=0$ ({\it Lower Panel}).
1436: The map in the {\it Upper Panel} has
1437: three attractor points A, B, C in which B, C correspond to
1438: high angular momentum. The map in {\it Lower Panel} has one attractor
1439: point C corresponding to ``wrong'' alignment and two attractor points with $J=0$.}
1440: \label{f24}
1441: \end{figure}
1442:
1443: We see that these trajectory maps are similar to those constructed by RATs
1444: from AMO (see Fig. \ref{f7}). Indeed, for $Q_{e1}=0$ case, the map for AMO in Fig. \ref{f7} shows two attractor points at $\mc \Theta= 1$, which are
1445: found in Fig. \ref{f24}.
1446: However, in the situation when only $Q_{e2}$ acts, the upper panel in Fig.
1447: \ref{f24} shows an attractor point at $\Theta=\pi/2$ and $J\gg J_{th}$, which is somewhat
1448: different from
1449: Fig. \ref{f7}. This difference stems from the fact that $Q_{e2}$ for shape 1 is though small, but not equal to zero at $\Theta=\pi/2$. Therefore, RATs are still able to
1450: spin up grains and align them there.
1451: \begin{figure}
1452: \includegraphics[width=0.49\textwidth]{f31a.ps}
1453: \caption{Phase trajectory map in the absence of magnetic field when all torque components act together shows the alignment with one high-$J$ attractor point C , two low-$J$ attractor points A and B corresponding to perfect alignment with ${\bf k}$, and one repellor point D.}
1454: \label{f25}
1455: \end{figure}
1456:
1457: Fig. \ref{f25} shows the phase map when all torque components are at work corresponding to shape 1 and ISRF.
1458: It is shown that the trajectory map does not have the attractor point at $\Theta=\pi/2$. Thus, similarly to AMO, the attractor point corresponding to aligning grains
1459: with long axes parallel to the direction of radiation
1460: disappears when non-zero $Q_{e1}$ is accounted for.
1461:
1462: Moreover, from Fig.~\ref{f25} (shape 1) and the upper panel in Fig.
1463: \ref{f8} for AMO, it is seen that both maps have the same repellor point at
1464: $\mc \Theta=1$ in the upper panel. However, the point C in Fig. \ref{f25} is an attractor point, rather
1465: than a repellor point B as seen for AMO. This difference stems
1466: from the fact that for AMO, $Q_{e1}=\frac{4\pi l_{1}n_{1} n_{2}|n_{2}|}{\lambda}\frac{4}{3\pi}(5\mbox{cos}^{2}\Theta-2)$
1467: is completely symmetric, i.e., $Q_{e1}(\Theta=0)=Q_{e1}(\Theta=\pi)$, while
1468: for the Shape 1, Fig.~\ref{f21} (the solid-dot line) shows that
1469: $Q_{e1}(\Theta=0)<Q_{e1}(\Theta=\pi)$. Therefore, the stationary point
1470: $\Theta=0$ in the upper panel, does not satisfy the criteria for attractor points, i.e., it is a
1471: repellor point, while it is an attractor point in the lower panel.
1472:
1473:
1474:
1475:
1476: \section{Alignment with respect to ${\bf B}$}\label{balign}
1477:
1478: We showed in \S \ref{kalign} that grains can be aligned with respect to ${\bf k}$.
1479: Below we consider the case when magnetic field is essential in terms of grain
1480: precession (see Fig.~\ref{f3}). As earlier, we disregard the paramagnetic relaxation. We make an extensive use of the physical insight obtained with
1481: a more simple case of alignment in \S \ref{kalign}. Indeed, because of the precession about magnetic field the analytical treatment of the corresponding processes gets
1482: less transparent here compared to that in \S \ref{kalign}.
1483:
1484: \subsection{Equations of motion in presence of ${\bf B}$}
1485:
1486:
1487: \begin{figure}
1488: \includegraphics[width=0.49\textwidth]{f12.eps}
1489: \caption{Coordinate systems used to study grain alignment in which the
1490: external magnetic field {\bf B} defines the alignment axis.}
1491: \label{f3}
1492: \end{figure}
1493: In the presence of magnetic field, equations of motion in dimensionless units
1494: become
1495: \begin{align}
1496: \frac{d\phi}{dt}&= \frac{M}{\ms\xi}G(\xi, \psi,\phi)-\Omega_{B},\label{eq58}\\
1497: \frac{d\xi}{dt}&=M \frac{F(\xi, \psi,\phi)}{J},\label{eq59}\\
1498: \frac{dJ}{dt}&=M H(\xi, \psi,\phi)-J,\label{eq60}
1499: \end{align}
1500: where $\Omega_{B}$ is the Larmor precession rate of ${\bf J}$ around the magnetic
1501: field ${\bf B}$. Here $F, H, G$ are RAT components projected to three axes
1502: $\hat{\xi}, \hat{J}, \hat{\phi}$, which are given by (see DW97)
1503: \begin{align}
1504: F(\xi,\psi,\phi)&=Q_{e1}(\Theta,0)[-\ms\psi \mc\xi \mc\phi-\mc \psi \ms\xi]\nonumber\\
1505: &+Q_{e2}(\Theta, 0)[\mc \Phi(\mc \psi \mc\xi\mc\phi\nonumber\\
1506: &-\ms\psi \ms\xi)+\ms\Phi \mc\xi \ms\phi]\nonumber\\
1507: &+Q_{e3}(\Theta, 0)[\mc\Phi \mc\xi \ms\phi +\ms\Phi(\ms\psi \ms\xi\nonumber\\
1508: &-\mc\psi \mc\xi \mc\phi)],\label{eq61}\\
1509: H(\xi,\psi,\phi)&=Q_{e1}(\Theta, 0)[-\ms \psi \ms\xi\mc\phi+\mc\psi \mc\xi]\nonumber\\
1510: &+Q_{e2}(\Theta, 0)[\mc\Phi(\ms\psi \mc\xi\nonumber\\
1511: &+\mc\psi \ms\xi \mc\phi)+\ms\Phi \ms\xi \ms\phi]\label{eq62},\\
1512: G(\xi,\psi,\phi)&=Q_{e1}(\Theta, 0)[\ms \psi \ms\phi]\nonumber\\
1513: &+Q_{e2}[\ms\Phi\mc\phi-\mc\Phi\mc\psi \ms\phi]\nonumber\\
1514: &+Q_{e3}[\mc\Phi \mc\phi+\ms\Phi \mc\psi \ms\phi]\label{eq62b}.
1515: \end{align}
1516: Here $\Theta$ and $\Phi$ are related to $\xi$, $\psi$, $\phi$ via
1517: \begin{align}
1518: \mc \Theta&=\mc\xi \mc\psi -\ms\xi \ms\psi \mc\phi,\label{eq63}\\
1519: \Phi&=2\mbox{tan}^{-1}\frac{\ms\Theta-\ms\xi \ms\psi}{\ms\xi \ms\phi}.\label{eq64}
1520: \end{align}
1521: Equation (\ref{eq62}) reveals explicitely that the component $Q_{e3}(\Theta, 0)$ does not
1522: contribute to spinning up grains. On the other hand, we found numerically that
1523: the last term containing $Q_{e3}(\Theta, 0)$ in equation (\ref{eq61}) goes to
1524: zero after averaging over the precession angle $\phi$. Therefore, similar to
1525: the case of alignment with respect to ${\bf k}$, the only effect of
1526: $Q_{e3}(\Theta, 0)$ is to induce the grain precession.
1527:
1528: After averaging over the precession angle $\phi$, the equations of
1529: motion (\ref{eq58})-(\ref{eq60})
1530: are reduced to two equations for $\xi$ and $J$, whereas $F(\xi, \psi, \phi), H(\xi, \psi, \phi)$ are replaced by
1531: $\langle F(\xi, \psi)\rangle_{\phi}, \langle H(\xi, \psi)\rangle_{\phi}$.
1532:
1533: \subsection{Stationary points for arbitrary shaped grains}
1534: While in this section we deal with AMO, some results can be obtained in a
1535: general case of arbitrary shaped grains. More results of this nature are
1536: presented in \S 8.
1537:
1538: In the presence of magnetic field, aligning and spinning torques are complicated
1539: functions of RATs, involving $\psi, \xi, \phi$ variables. Therefore, it's not
1540: easy to derive general analytical expressions for stationary points. However, we can find some particular physically
1541: interesting situations that
1542: correspond to stationary points.
1543:
1544: For instance, the perfect alignment
1545: corresponds to the maximal inertia axis ${\bf a}_{1}$ parallel to the magnetic
1546: field, i.e., $\ms\xi_{s} =0$. For this angle, from equations (\ref{eq63}) and (\ref{eq64}), we have $\Theta=\psi$, and
1547: $\Phi=0$ or $\pi$. Hence, $Q_{e1}(\Theta,0)=Q_{e1}(\psi,0)$, $Q_{e2}(\Theta,0)=Q_{e2}(\psi,0)$. Equation~(\ref{eq61}) becomes
1548: \begin{align}
1549: F(\xi_{s},\psi,\phi)&=Q_{e1}(\psi,0)\ms\psi \mc\phi+Q_{e2}(\psi,0)\mc\Phi \mc\phi \nonumber\\
1550: &+Q_{e3}(\xi, 0)\mc\Phi\ms\phi. \label{eq65}
1551: \end{align}
1552:
1553: Obviously, $F(\xi_{s},\psi,\phi)$ is a function of the precession
1554: angle $\phi$ about the magnetic field ${\bf B}$. Thus, if the grain precesses rapidly around ${\bf B}$, then we can average $F(\xi_{s},\psi,\phi)$ over $\phi$ from $0$ to $2\pi$. As a result,
1555: \begin{align}
1556: \langle F(\xi_{s},\psi)\rangle_{\phi}&=Q_{e1}(\psi,0)\ms\psi \int_{0}^{2\pi}\mc\phi
1557: d\phi \nonumber\\
1558: &+Q_{e2}(\psi,0)\mc\Phi \int_{0}^{2\pi}\mc\phi
1559: d\phi\nonumber\\
1560: &+Q_{e3}\mc\Phi\int_{0}^{2\pi}\ms\phi d\phi=0,
1561: \label{eq66}
1562: \end{align}
1563: which implies that for {\it grains of an arbitrary shape}, and
1564: for {\it arbitrary direction of light with respect to the magnetic field},
1565: there are always two stationary points at $\xi_{s}=0$ and $\pi$. This very
1566: fact makes the alignment of the grains with long axes perpendicular to ${\bf B}$ in some sense
1567: the expected one, although it does not present a sufficient condition for
1568: such an alignment.
1569:
1570: If attractor points exist for $\xi_{s}$ different from $0$ or $\pi$, the
1571: alignment may get ``wrong'', i.e. with the maximal inertia axis of the grain
1572: tending to be parallel to the magnetic field. Here and below we adopt the convention
1573: that the alignment is ``right'' if it corresponds to the Davis-Greenstein predictions,
1574: which made the Davis-Greenstein mechanism so popular even in spite of its inefficiency.
1575: Needless to say, that
1576: RATs we seek ``right'' alignment, i.e. with long grain axes perpendicular
1577: to magnetic fields without appealing for paramagnetic relation.
1578:
1579: The ``wrong'' alignment may be expected, for instance,
1580: when the radiation beam is perpendicular to the
1581: magnetic field, i.e. $\psi=\pi/2$. For this $\psi$, consider
1582: the direction of ``wrong'' alignment corresponding to $\xi_{s}=\pi/2$.
1583: For $\psi=\pi/2$, and $\xi=\pi/2$, we have $\mc \Theta=-\mc \phi$, and
1584: $\Phi=\pi/2$ for $\phi <\pi$ and $-\pi/2$ for $\phi>\pi$. Therefore, equation
1585: (\ref{eq61}) becomes
1586: \begin{align}
1587: F(\xi_{s},\psi,\phi)&=Q_{e3}(\Theta,0)\ms \Phi.\label{eq67}
1588: \end{align}
1589: Thus
1590: \bea
1591: \langle F(\xi_{s},\psi)\rangle_{\phi}=\int_{0}^{2\pi}Q_{e3}(\phi,0)d\phi=0.\label{eq68}
1592: \ena
1593: Here we use the property $Q_{e3}(\phi, 0)\sim \ms2\phi$ in calculating the integral.
1594:
1595:
1596: \subsection{"Right" and "wrong" alignment for AMO}
1597:
1598: The introduction of fast precession arising from ${\bf B}$ makes the dynamics
1599: of grains more interesting. For instance, it allows for a parameter space
1600: for ``wrong'' alignment, high attractor points for AMO, shifts of the crossover points.
1601:
1602: \subsubsection{Torques considerations}
1603:
1604: Above, we found that in the presence of ${\bf B}$, two permanent
1605: stationary points are $\xi_{s}=0, \pi$. If being attractor points, they correspond to the ``right''
1606: alignment.
1607:
1608: Consider a case of suspected ``wrong'' alignment for AMO at $\xi_{s}\sim \pi/2$.
1609: Since as $\psi=\pi/2$, $\mc\Theta=-\mc\phi$, so RATs
1610: (see equations \ref{eq30a} and \ref{eq31a}) become
1611: \begin{align}
1612: Q_{e1}&=\frac{16 l_{1} n_{1}n_{2}|n_{2}|}{3\lambda}(5\mc^{2}
1613: \phi-2),\label{eq69}\\
1614: Q_{e2}&=\frac{40 l_{1} n_{1}n_{2}|n_{2}|}{3\lambda}\ms2\phi(1.191+0.1382\mcs\phi),\label{eq70}
1615: \end{align}
1616: Substituting equations (\ref{eq69}) and (\ref{eq70}) into equation
1617: (\ref{eq62}) and averaging over the precession angle $\phi$, we get
1618: \begin{align}
1619: \langle H(\xi, \psi)\rangle_{\phi}&=\frac{16 l_{1} n_{1}n_{2}|n_{2}|}{3\lambda}\int_{0}^{2\pi}(5\mc^{2}\phi-2)\mc\phi
1620: d\phi\nonumber\\
1621: &+\frac{40 l_{1} n_{1}n_{2}|n_{2}|}{3\lambda}[\int_{0}^{2\pi} \ms2\phi \ms
1622: \phi(1.191)d\phi\nonumber\\
1623: &+\int_{0}^{2\pi}\ms2\phi(0.1382\mcs\phi) d\phi]=0.\label{eq71}
1624: \end{align}
1625: The fact that the integral (\ref{eq71}) is equal to zero means
1626: that RATs do not spin up grains when they are perpendicular to the magnetic field.
1627:
1628: Now, let us study whether $\xi_{s}=\pi/2$ satisfies the condition of an attractor
1629: points. Fig.~\ref{f10}{\it upper} shows the spinning and aligning torques for
1630: $\psi=89.9^{0} \sim 90^{0}$ \footnote{Here we take $\psi=89.9^{0}$ to avoid
1631: the singularity that may appear due to zero of $\langle H\rangle$ as shown in
1632: equation (\ref{eq71}) when considering condition of attractor points} for
1633: the state ${\bf a}_{1}\| {\bf J}$. It shows that, there are stationary points at
1634: $\mc\xi=\pm 1$ and $\mc\xi_{s}=0.1, -0.1$ for the ${\bf J}$ parallel and anti-parallel to ${\bf a}_{1}$ corresponding to zeroes of $\langle F(\xi) \rangle_{\phi}$. For the stationary point C with $\mc\xi_{s}=0.1$, Fig. \ref{f10} shows $\left. \frac{d\langle
1635: F(\xi)\rangle_{\phi}}{d\xi}\right|_{\xi_{s}}>0$, and $\langle H(\xi_{s})\rangle_{\phi}>0$. Hence, this stationary point does not satisfy
1636: equation (\ref{eq46}), i.e, it is a repellor point. Meanwhile, the stationary point C' at $\mc\xi_{s}=-0.1$, we have $\left. \frac{d\langle
1637: F(\xi)\rangle_{\phi}}{d\xi}\right|_{\xi_{s}}<0$ and $\langle
1638: H(\xi_{s})\rangle_{\phi}>0$. As a result, the stationary point C' is an
1639: attractor point. Similarly, the point A at $\mc\xi_{s}=1$ is an attractor
1640: point since $\left. \frac{d\langle F(\xi)\rangle_{\phi}}{d\xi}\right|_{\xi_{s}}<0$ and $\langle
1641: H(\xi_{s})\rangle_{\phi}>0$; also, the point B is a low-$J$ attractor
1642: point. However, the points A' and B' are repellor points in the case ${\bf
1643: J}$ anti-parallel to ${\bf a}_{1}$.
1644:
1645: Therefore, the possibility of existence of ``wrong'' alignment is feasible, but it
1646: happens at a low angular momentum. So, RATs from anisotropic radiation field
1647: themselves can not maintain the ``wrong'' alignment with respect to magnetic field
1648: in the presence of thermal wobbling and the bombardment by the ambient gas (see Hoang \& Lazarian 2007).
1649:
1650: \subsubsection{Effect of isotropic torques on ``wrong'' alignment}
1651:
1652: If grains are sufficiently large not to experience frequent thermal flips (see LD99a),
1653: they
1654: can be subjected to regular isotropic torques, which include both Purcell's torques
1655: (Purcell 1979) and those from isotropic radiation flux (DW96). According to
1656: LD99b, the this corresponds to grains larger than $10^{-4}$~cm, which is much larger
1657: than the typical size of the grains in our Table~2. However, such large
1658: grains are relevant
1659: to many astrophysical environments, e.g. comets, dark clouds, accretion disks.
1660:
1661: The ``wrong'' alignment gets modified when we take into
1662: account RATs induced by isotropic radiation, or equivalently, Purcell's $H_{2}$ torques. Indeed, ``isotropic'' RATs and Purcell's spin-wheel are both parallel to the
1663: maximal inertia axis, i.e, ${\bf a}_{1}$. Therefore, the total spinning and
1664: aligning torques are $
1665: \langle H(\xi, \psi)\rangle_{\phi}+Q_{iso}, \langle F(\xi,
1666: \psi)\rangle_{\phi}$, respectively.
1667:
1668: Since the aligning torque $\langle F(\xi, \psi)\rangle_{\phi}$ depends uniquely on
1669: RATs induced by anisotropic radiation, the positions of ``wrong''
1670: attractor points do not change. Meanwhile, their angular momentum are added by
1671: a term resulted from ``isotropic'' RATs or Purcell pinwheel.
1672:
1673: Now, let introduce $H_{2}$ torques. $H_{2}$ torques, after averaging over the
1674: grain rotation around the maximal inertia axis, are given by
1675: \bea
1676: {\bf \Gamma}_{H_{2}}={\bf a}_{1}\frac{I_{1}\omega_{H_{2}}}{t_{gas}}p,
1677: \ena
1678: where
1679: \bea
1680: \omega_{H_{2}}\sim 5\times 10^{7} \left(\frac{f}{a_{-5}^{2}}\right)\left(\frac{l}{10
1681: A^{0}}\right) \left(\frac{E}{0.2 eV}\right)^{1/2}\left[\frac{n(H)}{n}\right] s^{-1}.
1682: \ena
1683: Here $l^{2}$ is the surface area of a catalytic site of the grain surface,
1684: $f$ is the efficiency of $H_{2}$ formation, $n(H)$ is the density of atomic hydrogen, $E$ is the kinetic energy of the
1685: escaping $H_{2}$ molecule, and $p$ is a random variable (see DW97).
1686: In addition to parameters given in Table \ref{tab1}, here we assume that $n(H)/n=1, f=\frac{1}{3}, l=10 A^{0}$, and $E=0.2 eV$; we consider then the role of $H_{2}$ torques for
1687: $p=1, -1$ corresponding to ${\bf \Gamma}_{H_{2}}$ parallel and anti-parallel
1688: to ${\bf a}_{1}$.
1689:
1690: \subsubsection{Trajectory maps}
1691:
1692: To illustrate the alignment with respect to the magnetic field for AMO,
1693: we construct phase trajectory maps using RATs obtained by averaging equations
1694: (\ref{eq8})-(\ref{eq10}), for different radiation directions
1695: $\psi$. To do this we adopt the parameters from Table \ref{tab1}.
1696: For $\psi=0^{0}$, i.e., ${\bf B}\| {\bf k}$, the phase map shown in the upper
1697: panel of Fig. \ref{f8} exhibits a perfect alignment of ${\bf J}$ with respect to ${\bf B}$.
1698:
1699: \begin{figure}
1700: \includegraphics[width=0.49\textwidth]{f13a.ps}
1701: \includegraphics[width=0.49\textwidth]{f13b.ps}
1702: \caption{Phase trajectory maps for AMO for $\psi=30^{0}$ ({\it
1703: Upper Panel})
1704: and $\psi=60^{0}$ ({\it Lower Panel}) both have zero attractor points C,
1705: and two repellor points A, B. }
1706: \label{f9}
1707: \end{figure}
1708:
1709: Fig.~\ref{f9} shows the phase trajectory maps for $\psi=30, 60^{0}$ with
1710: two repellors A and B, and one low-$J$ attractor point C.
1711:
1712: \begin{figure}
1713: \includegraphics[width=0.49\textwidth]{f14a.ps}
1714: \hfill
1715: \includegraphics[width=0.49\textwidth]{f14b.ps}
1716: \caption{Aligning and spinning torques for the particular direction of light
1717: $\psi\sim 90^{0}$ and the corresponding phase map for AMO. {\it Upper Panel}:
1718: Solid line shows the aligning torque $\langle F(\xi, \psi)\rangle_{\phi}$ with
1719: two zeroes corresponding to two stationary
1720: points at $\mc\xi=0.1, -0.1$ corresponding to the case ${\bf J}$ parallel and anti-parallel to ${\bf a}_{1}$, besides two zeroes $\mc\xi=\pm1$, while the
1721: dashed line
1722: shows the spinning up torque $\langle H(\xi, \psi)\rangle_{\phi}$. {\it Lower
1723: panel}: Phase map corresponding to RATs in the upper panel shows one
1724: ``wrong'' attractor point C' at very low $J/I_{1}\omega_{T}\sim 0$ and $\mc\xi=-0.1$.}
1725: \label{f10}
1726: \end{figure}
1727:
1728: \begin{figure}
1729: \includegraphics[width=0.49\textwidth]{f15a.ps}
1730: \space
1731: \includegraphics[width=0.49\textwidth]{f15b.ps}
1732: \caption{Diagram for the position ({\it Upper panel}) and angular momentum ({\it
1733: Lower panel}) of attractor points as function of $\psi$ for
1734: three cases,
1735: $Q_{iso}=-10^{-5}, p=0$ (no $H_{2}$ torques), $Q_{iso}=-10^{-5}, p=1$ ($\Gamma_{H2}$ is parallel to
1736: ${\bf a}_{1}$), and $Q_{iso}=-10^{-5}, p=-1$
1737: ($\Gamma_{H2}$ is anti-parallel to ${\bf a}_{1}$). Each symbol (diamond,
1738: triangle, and square) denotes an
1739: attractor point, and a symbol-line shows the extend of the ``wrong''
1740: alignment as a function of $\psi$. }
1741: \label{f11}
1742: \end{figure}
1743:
1744: For $\psi \sim 90^{0}$, the lower panel in Fig. \ref{f10} shows an attractor
1745: point C' at $\mc\xi=-0.1$ corresponding to
1746: $J \sim 0$ and two repellor points at $\mc\xi=1$,
1747: respectively. It is shown that the phase map is not symmetric between upper
1748: and lower panels compared to the maps for $\psi<90^{0}$. This stems from the
1749: fact that spinning and aligning torques are no longer symmetric. The existence of the attractor point C' indicates
1750: that there is, indeed, a ``wrong'' alignment
1751: situation as predicted by our analysis above. However, to what extend of $\psi$ the "wrong" alignment occurs?
1752:
1753:
1754: Fig. \ref{f11} shows the angle of attractor points (upper) and their corresponding angular momentum (lower) for a range $\psi=[89.5^{0},90^{0}]$. In each frame of Fig. \ref{f11}{\it lower}, we show angular momentum of attractor points present in both upper and lower panels of a trajectory map (upper and lower panel are labeled).
1755:
1756: For the case $Q_{iso}=10^{-5}, p=0$, the upper panel in Fig. \ref{f11} shows that there are always three attractor
1757: points in which one attractor point happens at $\mc\xi=-0.1, -0.05$ and $0$ for $\psi=89.5^{0}, 89.8^{0}$ and $90^{0}$, corresponding to "wrong" alignment, two other with $\mc\xi=\pm 1$. However, their angular momentum is very low, about $1.5 I_{1}\omega_{T}$ (top frame in Fig. \ref{f11}{\it lower}). When $H_{2}$ torques are taken into account, if $H_{2}$ torques are parallel to ${\bf a}_{1}$, i.e., $p=1$, then the attractor points $\mc\xi=\pm 1$ are lifted to $J/I_{1}\omega_{T}\sim 100$, while the angular momentum of the "wrong" attractor point is unchanged (see the middle frame in Fig. \ref{f11}{\it lower}). In contrast, if $p=-1$, the angular momentum of the
1758: "wrong" attractor point can be increased by $H_{2}$ torques to $J/I_{1}\omega_{T}=100$, but the attractor points $\mc=\pm 1$ are unchanged (see lower frame in Fig. \ref{f11}{\it lower}). Note, that the shift of "wrong" attractor point position toward $\mc\xi=0$ is also seen on the diagram when $\psi \to 90^0$ (Fig. \ref{f11}{\it upper}).
1759:
1760: In summary, the presence of "wrong" alignment with respect to magnetic fields can be expected, but it happens in a very narrow range of $\psi$ in the vicinity of $\psi=\pi/2$. In addition, their angular momentum is very low in the absence of isotropic torques such as $H_{2}$ torques and "isotropic" RATs.
1761: Naturally, the effects of isotropic torques are negligible if grains are flipping,
1762: and therefore are thermally trapped
1763: as it is discussed in LD99a.
1764:
1765: \subsection{Alignment for Irregular Grains}
1766:
1767: Similarly as in \S \ref{kalign} we present the phase trajectory maps for shape 1,
1768: which is the ``most irregular'' in terms of RATs.
1769: Consider first phase maps for shape 1 obtained for and $\psi=90^{0}$ first.
1770: (see Fig. \ref{f25a}). For this case, RATs for shape 1 create an attractor point at
1771: $\xi=90^{0}$, $J/I_{1}\omega_{T}=2$ in the phase map, which corresponds to
1772: {\it wrong alignment} (see Fig. \ref{f25a}).
1773: AMO also does produce such a "wrong" alignment but at much smaller angular
1774: velocity, $J/I_{1}\omega_{T} \sim 0$ (see Fig. \ref{f10}). This difference
1775: is resulted from the property of $Q_{e2}$ component. Following AMO, $Q_{e2}=0$ at
1776: $\Theta=\pi/2$, while for shape 1, this component is not zero at this angle.
1777:
1778: All in all, for various grains shapes studied, we found that there is a narrow range $\psi=85^{0}\to 90^{0}$
1779: in which there exist "wrong" alignment.
1780:
1781: \begin{figure}
1782: \includegraphics[width=0.49\textwidth]{f31b.ps}
1783: \caption{For $\psi=90^{0}$, the map shows three high-$J$ attractor point A, B and D and two low-$J$ attractor point C and E. Most grains in the lower frame of the map align on D with long axes parallel to ${\bf B}$, i.e., "wrong" alignment. }
1784: \label{f25a}
1785: \end{figure}
1786:
1787: In addition, we illustrate the alignment in respect to magnetic field with
1788: phase trajectories obtained for $\psi=30^{0}$. For shape 1, similar studies were
1789: performed in DW97. However, their treatment of crossovers was different from
1790: ours. Thus the phase trajectories that we observe are different. We do not see cyclic maps with
1791: grains emerging from a crossover to get accelerated by RATs. Instead, we see,
1792: similar to the case of AMO (see Fig. \ref{f9}{\it upper}), the low-$J$ attractor points (see Fig \ref{f25a}).
1793: \begin{figure}
1794: \includegraphics[width=0.49\textwidth]{f32.ps}
1795: \caption{Phase trajectory map for grain shape 1 and $\psi=30^{0}$ shows the alignment with two repellor points A and B an one low attractor point C. }
1796: \label{f25b}
1797: \end{figure}
1798:
1799: Our study provides important insight into the role of $Q_{e1}$ and $Q_{e2}$ components.
1800: For instance,
1801: in the presence of magnetic field $Q_{e1}$ tends to provide the
1802: alignment with long axes perpendicular to magnetic field, i.e. the
1803: ``right'' alignment. The possibility of ``wrong'' alignment within a limited
1804: range of $\xi$ is related to $Q_{e2}$. Similarly, the ratio of $Q_{e1}^{max}/Q_{e2}^{max}$
1805: determines the existence of the high-$J$ attractor points.
1806:
1807:
1808:
1809:
1810: \subsection{Ratio of $Q_{e1}^{max}/Q_{e2}^{max}$: existence of high-$J$ and shift of low-$J$ attractor points }\label{sec5}
1811:
1812: \begin{figure}
1813: \includegraphics[width=0.49\textwidth]{f16a.ps}
1814: \includegraphics[width=0.49\textwidth]{f16b.ps}
1815: \caption{Phase trajectory maps with $\psi=60^{0}$ for the case of AMO ({\it Upper Panel})
1816: and for the shape 4 (see Fig. \ref{f18}) corresponding to the same ratio
1817: $Q_{e1}^{max}/Q_{e2}^{max}=0.78$ ({\it Lower Panel}). The maps show that both AMO
1818: and DDSCAT give rise two high-$J$ attractor points A, B at $\mc \xi=-1$ and one low J
1819: attractor point C.}
1820: \label{f12}
1821: \end{figure}
1822: \begin{figure}
1823: \includegraphics[width=0.48\textwidth]{f17.ps}
1824: \caption{Similar to the upper panel in Fig. \ref{f12}, but grains start from
1825: angular momentum smaller than the value of high-$J$ attractor point. Filed circles show initial position of grains. Filled arrows show the time interval of $0.5 t_{gas}$. Some grains get aligned at low attractor point
1826: over a short time $t<0.5 t_{gas}$, some other get to the high J attractor points over $t \sim 3 t_{gas}$.}
1827: \label{f12*}
1828: \end{figure}
1829:
1830: The role of the ratio $Q_{e1}^{max}/Q_{e2}^{max}$ on the grain alignment has been studied in \S \ref{qratio}
1831: for the case of no magnetic field (see \S \ref{kalign} when this is applicable). Below
1832: we consider the role of this ratio when the magnetic field is present.
1833: We shall show that this ratio is important in determining whether
1834: grains have high-$J$ attractor points.
1835:
1836: In previous subsections we have derived the expressions for the stationary points,
1837: and presented the trajectory maps for the RAT alignment for a particular $\alpha$,
1838: namely,
1839: $\alpha=45^{0}$. It is clearly shown that stationary points $\mc\xi=\pm 1$ do
1840: not depend on the magnitude of $Q_{e1}, Q_{e2}$. However, their properties,
1841: i.e., whether they are attractor or repellor points do so.
1842:
1843:
1844: Let us first consider a realization of AMO in which the functional forms are
1845: established for $\alpha=45^{0}$ and their magnitudes are rescaled to have $Q_{e1}^{max}/Q_{e2}^{max}= 0.78$ (see Fig. \ref{f5}). Note, that this ratio is similar to the
1846: ratio of RATs obtained by DDSCAT for the irregular shape 4 with radiation
1847: field of wavelength $\lambda=1.2 \mu m$ (see Fig. \ref{f61}).
1848:
1849: Fig. \ref{f12} shows the obtained trajectory maps for $\psi=60^{0}$ for both
1850: AMO and shape 4. We see that both AMO and shape 4 produce the maps with two
1851: attractor points A, B at $\mc\xi=-1$ and one low-$J$ attractor point C, but
1852: one difference is that the percentage of grains on A and B for shape 4 is
1853: higher than for AMO. For AMO and $\psi=60^{0}$, the existence of high-$J$
1854: attractor points A and B with $Q_{e1}^{max}/Q_{e2}^{max}= 0.78$ is not found in the case $\alpha=45^{0}$ that has $Q_{e1}^{max}/Q_{e2}^{max}= 1.2>1$ in which these points are repellor points (see Fig. \ref{f9}). In other words, the existence of high-$J$ attractor points depends on
1855: the value of $Q_{e1}^{max}/Q_{e2}^{max}$ as predicted in \S~\ref{qratio}. Furthermore, the similarity in the
1856: phase maps between the AMO and shape 4, which have the similar ratio
1857: $Q_{e1}^{max}/Q_{e2}^{max}=0.78$, indicates a good
1858: correspondence of AMO with irregular grains in terms of grain dynamics.
1859:
1860:
1861: While in most cases, we show the phase trajectory maps that start from high
1862: values of $J$, in Fig. \ref{f12*} we also show the phase trajectories starting
1863: at $J=J_{th}$. The dynamics with this initial conditions are similar. However,
1864: the alignment can be achieved faster (see further discussion in \S \ref{fast}).
1865: The filled arrows in Fig. \ref{f12*} mark time scales of $0.5 t_{gas}$. We see that it takes approximately less than $0.5 t_{gas}$ to get
1866: to the lower-$J$ attractor point C for grains with initial orientation close to C, for the typical
1867: interstellar diffuse gas conditions. Also, it takes about $3 t_{gas}$ to get to the high-$J$ attractor points A and B (see Fig. \ref{f12*}). These time scales are usually much smaller than that for
1868: paramagnetic damping invoked in the Davis-Greenstein mechanism
1869:
1870:
1871:
1872: Now let us use the approximate functional form of RATs
1873: given by equations (\ref{eq54}) and
1874: (\ref{eq55}), and seek the range of $Q_{e1}^{max}/Q_{e2}^{max}$,
1875: in which the phase map has high J attractor points.
1876: Fig. \ref{f13} shows the ratio of
1877: $Q_{e1}^{max}/Q_{e2}^{max}$ for which there exist high-$J$ attractor
1878: points in the phase trajectory map of grains. It is shown that for $\psi <45^{0}$, $Q_{e1}$ is required to be
1879: dominant over $Q_{e2}$, at least $Q_{e1}^{max}=2Q_{e2}^{max}$, to have high-$J$ attractor points; their ratio is an
1880: increasing function of $\psi$. At $\psi=45^{0}$, one does not have high-$J$
1881: attractor points because $Q_{e1}, Q_{e2}$ are equally projected to ${\bf B}$
1882: and to the direction ${\bf \xi}$ perpendicular to ${\bf B}$.
1883:
1884: For $\psi>45^{0}$, high-$J$ attractor points occur when $Q_{e2}$
1885: becomes predominant, i.e., $Q_{e1}^{max}/Q_{e2}^{max}<1$ (see the lower panel
1886: in Fig. \ref{f13}).
1887: Also in Fig. \ref{f13}, the intermediate region with parallel lines correspond
1888: to the range in which the phase
1889: trajectory map
1890: has only low-$J$ attractor points, while high-$J$ stationary points are repellors.
1891:
1892: Fig. \ref{f13} shows also the existence of high J attractor points for
1893: irregular shapes 1 (both monochromatic radiation field and ISRF), 2 and 4 (only
1894: monochromatic radiation). For shape 1 and radiation of $\lambda=1.2\mu m$, its phase map has low attractor points
1895: for $\psi<60^{0}$ and high J attractor point for $\psi\ge 60^{0}$. But, the
1896: map for ISRF has J attractor point only in a small range
1897: $\psi<10^{0}$. Interestingly enough, for this wavelength, the ratio of
1898: $Q_{e1}^{max}/Q_{e2}^{max}$ exceeds those possible for the "original" AMO. As a result, as we mentioned
1899: earlier, an attractor point is possible for $\psi=0$, which also describe the
1900: situation of grain alignment in the absence of magnetic field. For such situations
1901: AMO has only repellor points (see \S\ref{kalign}).
1902:
1903: \begin{figure}
1904: \includegraphics[width=0.5\textwidth]{f18a.eps}
1905: \caption{Ratio of $Q_{e1}^{max}/Q_{e2}^{max}$ for which the phase map has
1906: high-J attractor points or low-$J$ attractor points as a function of $\psi$. Curved solid lines show
1907: predictions by AMO, solid lines (maps with low-$J$ attractor points only), dot lines
1908: (maps with high J attractor points) show the result for irregular grains.}
1909: \label{f13}
1910: \end{figure}
1911:
1912: \begin{figure}
1913: \includegraphics[width=0.49\textwidth]{f19a.eps}
1914: \includegraphics[width=0.49\textwidth]{f19b.eps}
1915: \caption{{\it Upper panel:} The shift of position of low-$J$
1916: attractor point as function of $Q_{e1}^{max}/Q_{e2}^{max}$ for different $\psi$. {\it Lower panel:} The shift of position of low-$J$
1917: attractor point with respect to $\psi$ for AMO with three different ratio $Q_{e1}^{max}/Q_{e2}^{max}$ and
1918: for irregular grains. The shaded area corresponding to the degree of
1919: alignment
1920: $R=1.5\mcs\xi-0.5\le0$ shows the range of ``wrong'' alignment.}
1921: \label{f13*}
1922: \end{figure}
1923:
1924:
1925: The differences in the range of $\psi$ in which
1926: the maps have high J attractor points predicted by AMO and irregular
1927: grains (i.e., dot lines extend outside the region surrounded by thick
1928: solid lines in Fig. \ref{f13}) exist, but it is moderate.
1929: In fact, the correspondence between the ranges of $\psi$ for the
1930: existence of the high-$J$
1931: attractor point revealed between our predictions and
1932: the actual irregular grains allows one to find out whether
1933: the high attractor point is expected if only $Q_{e1}^{max}/Q_{e2}^{max}$ ratio is
1934: known. Note, that the existence of high attractor points is important for evaluating
1935: the degree of grain alignment.
1936:
1937: Another new effect related to $Q_{e1}^{max}/Q_{e2}^{max}$ ratio is a regular
1938: shift of
1939: the position of the crossover point, which,
1940: as we discussed below, is also low-$J$ attractor point.
1941: Fig.~\ref{f13*}{\it upper} shows this shift as a function of the ratio $Q_{e1}^{max}/Q_{e2}^{max}$ for different $\psi$ predicted by AMO. It is shown that the low-$J$ attractor point tends to shift to $\mc\xi=0$ as $Q_{e2}^{max}$ increases. Particularly, for $\psi=89.9^{0}\sim 90^{0}$, the low-$J$ attractor point coincides to $\mc\xi=0$, i.e., grains are aligned with long axes perpendicular to magnetic field.
1942: The above observed tendency is consistent with our earlier discussion in \S~\ref{sec4p4} that $Q_{e2}$ act to align grains perpendicular to {\bf B}.
1943:
1944: Fig.~\ref{f13*}{\it lower} presents the shift as a function of $\psi$ for particular angles $\alpha$ of AMO and
1945: for irregular grains. It is clearly seen that, the low-$J$ attractor points for AMO and irregular grains are always at $\mc\xi=1$ for $\psi=0^{0}$, which means that no any "wrong" alignment possibility exists in this case. As $\psi$ increases, they shift to $\mc\xi=0$, and finally, fall on the perpendicular alignment angle $\mc\xi=0$ at $\psi\sim 90^{0}$. For shape~4, however, the low-$J$ attractor point can produce ``wrong''
1946: alignment earlier than for shape 2, i.e., when $\psi\sim 85^{0}$. This is because the shape 4 has the smaller ratio $Q_{e1}^{max}/Q_{e2}^{max}$ than the shape 2.
1947:
1948: We feel that $Q_{e1}^{max}/Q_{e2}^{max}$ provides a sufficiently good
1949: parametrization for torques of irregular grains. Therefore, in terms of
1950: practical calculations, establishing this ratio may be sufficient for
1951: describing the alignment of realistic irregular shapes. This would require
1952: much less computational efforts compared with obtaining the shapes for
1953: the entire range of $\Theta$.
1954:
1955:
1956:
1957:
1958:
1959: \section{Regular crossover}\label{cross}
1960:
1961: The dynamics of AMO that we observed above
1962: was very different from that of grains
1963: in DW97. As we can see in \S 8, the properties of the RAT components are similar
1964: for AMO and irregular grains, including those studied by DW97. Therefore
1965: the difference stems from the different treatment of crossovers in our and
1966: the DW97 models.
1967:
1968: The most striking difference between our and DW97 trajectory maps is that
1969: our crossover points correspond to $J=0$ and grains cannot get out of these
1970: points. Therefore our crossover points are also the attractor points. Note,
1971: that, unlike DW97, we do not observe cyclic maps. The latter are the artifact
1972: of their model of crossovers adopted there.
1973:
1974: In what follows, we first discuss crossovers in the most general terms and later consider
1975: a particular case of a crossover which is not affected by thermal wobbling
1976: (cf. WD03). Then we briefly discuss the possible effects of thermal wobbling.
1977:
1978: The crossover dynamics has two distinct regimes. The first one, described
1979: in Lazarian \& Draine (1997) takes place when the time of the crossover
1980: $t_{cros}\sim J_{\bot}/Q$ where $J_{\bot}$ is the angular momentum component
1981: perpendicular to ${\bf a}_{1}$ (Spitzer \& McGlynn 1979), i.e.
1982: the time during which the grain undergoes a regular flipping subjected
1983: to the torque $Q$, is shorter than the time of internal relaxation
1984: $t_{relax}$. For $H_2$ torques Lazarian \& Draine (1997) obtained
1985: that this regime is fulfilled for grains larger that $a_c\sim 10^{-5}$~cm.
1986: When LD99b introduced nuclear relaxation it became clear that
1987: the typical critical size for interstellar grains $\sim 10^{-4}$~cm. This size
1988: depends on the value of the torques rather weakly, thus we may also
1989: accept it as the critical size of the grains in diffuse ISM
1990: in the presence of radiative torques. However, in the vicinity of really strong
1991: radiation sources the critical size gets smaller. Note, that $a_c$ is larger
1992: than the ``typical'' $a_{eff}$ in Table~\ref{tab1}, but as we mentioned earlier, grains
1993: typical for diffuse interstellar gas are different from grains typical for other
1994: astrophysical environments.
1995:
1996: The randomization of grains during a crossover happens due to random
1997: processes associated with atomic bombardment and H$_2$ formation. This is
1998: a random walk process in which the squared deviation in angular momentum $(\delta J)^2$
1999: scales with the crossover time $t_{cros}$. Thus the deviation of the angular momentum
2000: in the process of a crossover scales as
2001: $\delta J/J_{th}\sim 1/J_{th}Q^{1/2}$.
2002: As $J_{th}\sim (kT_{d} I_1)^{1/2} \sim a_{eff}^{5/2}$ and assuming that $Q$ is proportional to the cross-section $\sim a_{eff}^2$ (see \S 10), we get the deviation $\delta J/J\sim a_{eff}^{-7/2}$, that decreases quickly with the grain size. Therefore the assumption of
2003: {\it no randomization during crossovers} should initially be accurate for grains much
2004: larger than $a_c$. Such grains according to Cho \& Lazarian (2005) are
2005: responsible for far infrared polarization emanating from dark starless cores.
2006:
2007: For grains less than the critical size $a_c$
2008: LD99a showed that
2009: the physics is different. As the angular momentum
2010: of a grain get comparable with $J_{th}$, such grain wobbles fast due to
2011: coupling of rotational and vibrational degrees of freedom by the internal
2012: relaxation (see Lazarian \& Roberge 1997). For H$_2$ torques this
2013: results in flipping that reverses the direction of torques and gets the
2014: grains ``thermally trapped'' (see LD99a).
2015:
2016: All in all, if
2017: grain dynamics is presented in the axes of $J/J_{th}$ and $\cos \gamma$ with $\gamma$ being the angle between ${\bf a}_{1}$ and ${\bf J}$,
2018: the assumption of ${\bf J}\|{\bf a}_{1}$ is accurate for $J/J_{th}\gg 1$. In
2019: the range of $J/J_{th}= [0,1]$ the actual crossover physics should be
2020: accounted for.
2021:
2022: In view of the discussion above, consider a regular crossover first, i.e. assume that
2023: $a_{eff}\gg a_c$. The assumption that ${\bf a}_{1}$ is always parallel to
2024: ${\bf J}$ breaks inevitably as $J_{\|}\rightarrow 0$. Indeed,
2025: whatever is the efficiency of the internal relaxation mechanism, there
2026: will be some residual $J_{\perp}$. Lazarian \& Draine (1997) showed
2027: that the mean value of $J_{\perp}$ during a regular crossover cannot be smaller than $J_{th}$.
2028: In addition, as $J$ gets small, the efficiency
2029: of relaxation drops and the gaseous bombardment increases $J_{\bot}$ further
2030: in accordance with
2031: the original Spitzer \& McGlynn (1979) theory of crossovers.
2032:
2033: To simplify our treatment we disregard all internal
2034: relaxation processes during the crossover.
2035: This is justifiable as for $a_{eff}\gg a_c$ the crossover happens on the time
2036: scale shorter than the internal (e.g. nuclear) relaxation time.
2037: Since the mirror is assumed weightless,
2038: for AMO the dynamics of free rotation coincides with that of a spheroid. As a result, for a given $J$, ${\bf a}_{1}$ precesses around ${\bf J}$ with a constant angle $\gamma$. The state of the grain is completely determined
2039: by describing ${\bf J}$ in the lab and body systems. Equations of motion for
2040: this case are
2041: \begin{align}
2042: \frac{d{\bf J}}{dt}&={\bf \Gamma}-\frac{{\bf J}}{t_{gas}},\label{cr1}\\
2043: \frac{J d\mc\gamma}{dt}&=-\frac{dJ}{dt}\mc\gamma+\frac{dJ_{\|}}{dt},\label{cr2}
2044: \end{align}
2045: where ${\bf \Gamma}$ is RAT, and $J_\|=J \mc\gamma$ is the
2046: component of angular momentum along the maximal inertia axis.
2047: In the coordinate system $J, \xi, \phi$, equation (\ref{cr1}) returns to the set of equations of motion (\ref{eq58})-(\ref{eq60}). Averaging over the angles, $\Phi$, and $\phi$ corresponding to the precession
2048: of ${\bf a}_{1}$ around ${\bf J}$ and ${\bf J}$ around ${\bf B}$,
2049: respectively, and solving three resulting equations for $J, \xi,\gamma$, we can obtain trajectory maps during the regime of low J and regular crossover.
2050:
2051: \begin{figure}
2052: \includegraphics[width=0.49\textwidth]{f20a.ps}
2053: \includegraphics[width=0.49\textwidth]{f20b.ps}
2054: \caption{Phase maps for AMO with $\psi=30^{0}$. {\it Upper panel} shows that all
2055: grains approach an attractor point C of $J=0$ corresponding to the same angle $\mc\xi=-1$. However regular
2056: crossovers corresponding to $J_{\|}=0$ occur at different angles as
2057: illustrated by {\it lower panel}.}
2058: \label{f15}
2059:
2060: \end{figure}
2061: In Fig. \ref{f15} we show the phase map for the case $\psi=30^{0}$.
2062: The most striking feature observed there is that the grain experiences not
2063: a single, but multiple crossovers. Unlike Purcell's torques,
2064: as the grain flips, RATs do not change their direction. As a result,
2065: they drive $J$ further down. In other words, the dynamics of grains
2066: roughly corresponds to what we observed in the earlier sections of the
2067: present paper (cf. DW97 where grains are spun up after passing $J=0$).
2068:
2069: Depending on magnetic field value, the crossover may happen on the
2070: time scale smaller than the Larmor precession period. In this case, different
2071: grains will undergo crossovers at different $\phi$ and therefore will
2072: experience different torques during crossovers. However, this would not change
2073: the qualitative picture, as RATs will drive $J$ to zero
2074: irrespectively of $\phi$. We shall discuss this issue in more details elsewhere.
2075:
2076: Let us consider qualitatively
2077: the issue of thermal fluctuations. For the Purcell's torques
2078: those can be ignored during the crossovers for grains with size $a_{eff}\gg a_c$. RATs, however, are different. They tend to decrease $J$ further rather than
2079: spin up the grains after a crossover. The time of relaxation, which is also
2080: the
2081: time over which thermal fluctuations of $\gamma$ take place, increases as
2082: $J$ decreases. Therefore in the absence of external collisions
2083: one can imagine two situations: according
2084: to one, RATs stop the grain completely, so no thermal wobbling of $\gamma$
2085: is possible, another is that $J$ settles at the value that is of the order
2086: of $J_{th}$. The latter seems more probable for irregular grains.
2087:
2088: An interesting consequence of the considerations above is that a grain tends
2089: to get to the $J=0$, but the irregularities in the motion
2090: of the grain axes in respect to the
2091: photon flow prevents the angular momentum from reaching zero. Thus, we can
2092: expect that for grains with moments of inertia closer to one of a spheroid the value of
2093: angular momentum at the low-$J$ attractor point is going to be smaller than for
2094: more irregular grains.
2095:
2096: To find the actual value of effective $J$, one should recall that grains are subject to substantial random excitations in the typical
2097: astrophysical conditions. Gaseous bombardment, interactions with ions,
2098: absorption of photons (see Draine \& Lazarian 1998 for a quantitative
2099: description of the processes), affects grain rotation at the low-J attractor
2100: point. We expect these
2101: influences to increase the mean value of $J$ above $J_{th}$.
2102:
2103: Another effect should also be present in the presence of random bombardment. If
2104: the phase maps have both high and low-$J$ attractor point, the bombardment allow
2105: more grains to get to a higher-$J$ attractor point. As low-$J$ attractor
2106: points are characterized by higher internal randomization, quite counter-intuitively, the random forcing can increase the degree of alignment.
2107:
2108: All in all, our considerations above justify the simplified model
2109: of crossovers that we adopt in the paper. We shall provide a more detailed
2110: treatment in Hoang \& Lazarian (2007).
2111:
2112:
2113: \section{Particular Cases}\label{sec9}
2114:
2115: Our earlier discussion covered two limiting cases, namely, (a) the
2116: light beam being the axis of alignment, which corresponds to the
2117: precession rate induced by RATs $t_k^{-1}$ much larger than the Larmor
2118: precession rate $t_B^{-1}$ and (b) when the magnetic field constitutes the
2119: axes of alignment, i.e. $t_B^{-1}\gg t_k^{-1}$. Can this case be ever astrophysically
2120: important?
2121:
2122: \subsection{Criterion for {\bf B} alignment} \label{sec91}
2123:
2124: Consider the precession of ${\bf a}_{1}$ or ${\bf J}$ around ${\bf k}$
2125: driven by the component $Q_{e3}$ which
2126: is perpendicular to the axis ${\bf a}_{1}$ and ${\bf k}$.
2127: The timescale for RAT precession is defined by
2128: \bea
2129: t_{k}=\frac{2\pi}{|d\phi/dt|}= {\frac{10^{11}}{\hat{Q}_{e3}}}(\hat{\rho}a_{-5}\hat{T})^{1/2}(\frac{1}{\hat{\lambda}\hat{u}_{rad}}) \mbox{s},\label{eq78}
2130: \ena
2131: where
2132: \bea
2133: \frac{d\phi}{dt}=\frac{\lambda a_{eff}^{2}\gamma}{I_{1}\omega} \mbox{{\bf
2134: Q}}_{\Gamma}.\hat{\Phi}=\frac{\lambda a_{eff}^{2}\gamma}{I_{1}\omega} Q_{e3}.\label{eq79}
2135: \ena
2136: Here $\omega$ is the angular velocity of a grain around the maximal inertia
2137: axis ${\bf a}_{1}$. $\hat{Q}_{e3}=Q_{e3}/10^{-2}$,
2138: $\hat{\lambda}=\lambda/1.2 \mu m$, $\hat{u}_{rad}=u_{rad}/8.64\times 10^{-13}
2139: ergs~ cm^{-3}$, and $\hat{\rho}=\rho/3 gcm^{-3}$, $a_{-5}=a_{eff}/10^{-5} cm$, and
2140: $\gamma=0.1$ for anisotropy of ISRF. In
2141: deriving equation (\ref{eq78}) the assumption $\omega=\omega_{T}$ is used.
2142: It is easy to see that for axisymmetric grain shapes,
2143: $Q_{e1}, Q_{e2}$ are
2144: equal to zero, while $Q_{e3}$ is non-zero (see equations \ref{eq3}-\ref{eq5}). Therefore, the third component produces fast precession of grains about ${\bf k}$.
2145:
2146: The precession rate above should be compared with
2147: the Larmor precession rate.
2148: A rotating grain acquires a magnetic moment by the Barnet effect
2149: which is shown to
2150: be much stronger than that arising from the rotation of its charged body
2151: (Dolginov \& Mytraphanov 1976). The
2152: interaction of the magnetic moment with the external magnetic field causes the gradual
2153: precession of the grain around the magnetic field direction.
2154: The rate of Larmor
2155: precession around magnetic field, $t_{B}$, is given by
2156: \bea
2157: t_{B}=3\times 10^{7} a_{-5}^{2} \hat{\rho}^{-1/2}\hat{\chi}^{-1}\hat{B}^{-1} ~\mbox{s},\label{eq80}
2158: \ena
2159: where $\hat{B}=B/5 \mu G$, $\hat{\chi}=\chi/3.3\times 10^{-3}$ are
2160: normalized magnetic field and magnetic susceptibility, respectively.
2161:
2162: The alignment of grains whether with the radiation or magnetic field depends on their
2163: precession rate around these axes. According to equations (\ref{eq78}) and (\ref{eq80}),
2164: the ratio of precession rate due to radiation
2165: and magnetic field is given by
2166: \begin{align}
2167: \frac{t_{k}}{t_{\mbox{B}}}&=\frac{3.33\times 10^{3}}{\hat{Q}_{e3}}\hat{\rho}\hat{\chi}\hat{T}^{0.5}a^{-1.5}_{-5}(\frac{\hat{B}}{\hat{\lambda}\hat{u}_{rad}}).\label{eq81}
2168: \end{align}
2169: It can be easily checked that for typical diffuse ISM,
2170: $t_{B}/t_{k} \sim 10^{-3}$, i.e. the Larmor precession is much faster than
2171: precession induced by radiation, therefore magnetic field is the
2172: alignment axis.
2173:
2174: For AMO, RATs can be
2175: very different as $Q_{e3}$ arises from a spheroidal shape, while
2176: the other two components arise from a mirror; the ratio of
2177: the sizes of the mirror and the spheroidal body may be arbitrary.
2178: For irregular grains, however, we can see all three RAT components to be roughly
2179: comparable (see Figs \ref{f21}-{\ref{f23a}). In this case, $t_{k}$ can be
2180: used as a proxy for the time of ``fast alignment'' (see \S \ref{fast}).
2181:
2182: \subsection{Astrophysical Implications}
2183:
2184: It is possible to estimate that for the magnetic field of
2185: 5 $\mu$G grains get aligned with the radiation beam when the density of
2186: energy in a beam is
2187: $\mbox{u}_{\mbox{rad}}>10^{3}\mbox{u}_{\mbox{ISRF}}$. The appropriate
2188: radiation fields is typical near stars and supernovae.
2189:
2190: Fig. \ref{f16} shows the variation of the ratio of precession timescales
2191: with distance
2192: for different stars in which we use $Q_{e3}=Q_{e3}^{max}$ (for shape 1 and ISRF) for equation (\ref{eq81}). It is apparent that near the stars, grains precess
2193: around ${\bf k}$ much faster than around magnetic field. However, the ratio increases
2194: with distance as the radiation field decreases, so the precession around magnetic field is faster than around light, and grains align
2195: with magnetic field at some distance.
2196:
2197: \begin{figure}
2198: \includegraphics[width=0.49\textwidth]{f21a.ps}
2199: \hfill
2200: \includegraphics[width=0.49\textwidth]{f21b.ps}
2201: \caption{The ratio of timescales for the precession about the light direction and
2202: about the magnetic field assuming that magnetic field is homogeneous with
2203: strength $B=0.5 \mu G$ through
2204: envelopes. {\it Upper Panel}: For three values of star temperature, the density dependence on
2205: the distance from the star is fixed; {\it Lower Panel}: For a given star with
2206: different functions of density.}
2207: \label{f16}
2208: \end{figure}
2209:
2210: Another case when the alignment can happen with respect to the direction of light
2211: is the case of cometary dust. This case may be somewhat more complex, as
2212: electric field can be present near the comet head. This can induce
2213: precession of grains with dipole electric moment and therefore provide
2214: yet another axis of alignment. We discuss this problem in Hoang \& Lazarian (2007).
2215:
2216: If $t_k^{-1}$ and $t_B^{-1}$ are comparable, the axis of alignment does not
2217: coincide with either beam or {\bf B}. This case can be relevant to some part
2218: of comet grains, but a discussion of it is beyond the scope of the present
2219: paper.
2220:
2221: Note that $t_k^{-1}$ is a function of $\Theta$. Therefore when its amplitude
2222: value is larger than $t_B^{-1}$ this does not guarantee that
2223: the effect of magnetic field is negligible. For AMO the alignment
2224: drives grains into the position corresponding to $\Theta=\pi/2$, for
2225: which the component $Q_{e3}$ gets zero and therefore the magnetic field
2226: again dominates. Thus the alignment is expected with respect to an intermediate
2227: axis. This may be important for explaining circular polarization arising
2228: from comets (Rosenbush et al. 2007). Naturally, not only precession arising
2229: from magnetic field, but also from electric field and mechanical torques (see
2230: \S 11.4) should be taken into account for that case.
2231:
2232: \subsection{Fast alignment}\label{fast}
2233: In this section we ignore the damping role of the
2234: ambient gas, and show that grains can be
2235: aligned by radiation on a timescale shorter than the gas damping time. Such
2236: alignment can be called ``fast alignment'' in analogy with the ``fast dynamo''
2237: process (see Vishniac, Lazarian \& Cho 2003 and ref. therein)
2238: that can amplify magnetic field on the timescales shorter than the
2239: magnetic diffusion time. In particular, supernovae flashes can align grains
2240: around them with respect to the direction of light (see \S \ref{kalign}).
2241:
2242: \begin{figure}
2243: \includegraphics[width=0.49\textwidth]{f22a.ps}
2244: \includegraphics[width=0.49\textwidth]{f22b.ps}
2245: \caption{Alignment in the absence of gas damping for two directions of
2246: radiation $\psi=0^{0}$ ({\it Upper panel}) and $ 70^{0}$ ({\it Lower
2247: panel}). Here, the gray circles represent
2248: the initial position of grains. In the upper and lower panel, the distance between
2249: two filled
2250: arrows corresponds to a time interval $\Delta t= 10 t_{k}$, and $\Delta t=50
2251: t_{k}$, respectively.
2252: }
2253: \label{f14}
2254: \end{figure}
2255:
2256: \begin{figure}
2257: \includegraphics[width=0.49\textwidth]{f23a.ps}
2258: \includegraphics[width=0.49\textwidth]{f23b.ps}
2259: \caption{Percentage of grains as function of the ratio of the alignment
2260: timescale due to photon, namely $t_{phot}$ to the precession time about the radiation direction, $t_{k}$, for AMO with
2261: $Q_{e1}^{max}/Q_{e2}^{max}=0.78$ ({\it upper panel}) and shape 4 ({\it lower panel}). Here $\epsilon=\frac{Q_{e1}^{max}}{Q_{e3}^{max}}$, that is chosen to be equal unity for AMO, and $\epsilon \sim 1$ for shape 4.}
2262: \label{f14*}
2263: \end{figure}
2264:
2265: In Fig.~\ref{f14}{\it upper} we show the map for $\psi=0^{0}$ in which the
2266: distance between two arrows represents a time interval equal 10 precession
2267: time $t_{k}$ that is defined in equation
2268: (\ref{eq78}). It can be seen that the grains that have initial orientations close to
2269: the attractor point A require alignment timescales $t_{phot}$\footnote{$t_{phot}$ is defined as the alignment timescale of grains due to photon or radiation} is approximately $40 t_{k}$ to get aligned (see Fig. \ref{f14}{\it upper}). In contrast, grains that have
2270: initial angles far away from the attractor point need more time (up to $\sim 160 t_{k}$) to get
2271: there (see Fig. \ref{f14*}).
2272:
2273: Fig. \ref{f14}{\it lower} shows a similar effect for the case $\psi=70^{0}$. However, the interval of two arrows represents a time interval $\Delta t=50t_{k}$.
2274: The map shows that grains bound to the low-$J$
2275: attractor point A reach it fast, about $70 t_{k}$, i.e., on the time scales much less than the rotational damping time. Some grains that otherwise would go to the
2276: high-$J$ attractor point stream to infinite $J$ in the absence of damping
2277: ({\it lower panel})\footnote{We may observe that phase trajectories in Fig. \ref{f14}{\it lower}
2278: directed to high attractor points also correspond to the aligned state
2279: of grains, although the stationary state requires $t_{gas}$ to be achieved. In this sense all grains get aligned fast.}. In practical terms the latter fact for this regime (i.e., alignment timescale
2280: $t_{phot} < t_{gas}$) is not so important, as most
2281: grains get aligned at the low-$J$ attractor points anyhow. We also observe the shift of low-$J$ attractor point A as in the upper panel of Fig. \ref{f13*}.,
2282:
2283: The corresponding time for the fast alignment is proportional to ratio of
2284: grain angular momentum and the component of the torque $\Gamma_{rad}$. In
2285: terms of RATs normalized components $Q_{\Gamma}$ that is related to ${\bf \Gamma}_{rad}$ via equation~ (\ref{eq1}) the relevant combination is given by
2286: equation (\ref{eq78}). The corresponding function $F$ depends only on $Q_{e1}$ and
2287: $Q_{e2}$. The amplitudes of those can be very different from the component
2288: $Q_{e3}$ that causes the grain precession.
2289: %However, one can see that for
2290: %arbitrary irregular grain, all the components $Q_{e1}, Q_{e3}$ have similar
2291: %amplitudes (see Fig. \ref{eq27}).
2292: Therefore for such grains, the ratio of
2293: $J/\Gamma_{rad} \propto J/F$ can be measured in units of ``the period of
2294: radiation induced precession'', namely, in $t_{k} \propto J/Q_{e3}$ (see
2295: equation \ref{eq78}). To make AMO more correspond to irregular grains in this respect,
2296: we make the choose the amplitude of the AMO components $Q_{e3}$ to be of the
2297: similar to the amplitude of $Q_{e1}$.
2298:
2299: Histograms showing the distribution of grains on the low-$J$ attractor point as a function of alignment timescale corresponding to AMO and an
2300: irregular grain shape 4 are shown in Fig. ~\ref{f14*}. It can be seen that, for AMO, about $45\%$ grains get aligned with respect to
2301: ${\bf k}$ over $t_{phot}\sim 35 t_{k}$ to $45 t_{k}$, and about $22\%$ of grains get there over $t_{phot}\sim 50 t_{k}$ to $65 t_{k}$. A few percent of grains require longer time to get
2302: aligned, up to $170t_{k}$ (see Fig. \ref{f14*}{\it upper}) . This relative inefficiency of alignment is a
2303: consequence of small amplitude of the function $F$ in the vicinity of the
2304: low-$J$ attractor points ($F=0$ at the stationary points). On the other hand, for the shape 4,
2305: Fig. \ref{f14*}{\it lower} shows that about $55\%$ and $21\%$ of grains get aligned with
2306: ${\bf k}$ over $35 t_{k}$ and $65 t_{k}$, respectively. Some others can get aligned over $170 t_{k}$. We see that the similar distribution of grains as functions of alignment time between shape 4 and AMO, though the slight difference in percentage of grains corresponding to each $t_{phot}/t_{k}$ present due to the fact that, their functional forms of their torques are not comletely the same.
2307:
2308: Fast alignment happens in {\it respect to magnetic field} provided that
2309: $t_{B}<t_k$ (see equation \ref{eq81}), but $t_{phot}<t_{gas}$. The ratio
2310: $t_{gas}/t_{B}=1.2\times 10^5 \frac{\hat{\rho}^{3/2}\hat{B}\hat{\chi}}{\hat{n}\hat{T}_{g}^{1/2}}a_{-5}^{-1} $ which provides a substantial parameter space if
2311: $t_{phot}$ is much larger
2312: that $t_k$, e.g. $t_{phot}\sim 10^2 t_k$. Comparing $t_{gas}$ in Table~2 and
2313: $t_k$ given by
2314: equation (\ref{eq78}) we may conclude that for typical grains in diffuse
2315: interstellar gas $t_{gas}$ is
2316: marginally smaller than $t_{phot}$ and therefore the grain phase
2317: trajectories are still determined by $t_{gas}$. However, closer to stars
2318: $t_{phot}$ provides the measure of the characteristic time of alignment\footnote{We see that the alignment times are more than $30 t_{k}$. This is due to the fact that the aligning torque gets weaker near the low-$J$ attractor points (see Fig. \ref{f10}{\it upper})}.
2319:
2320:
2321:
2322:
2323: \section{Fitting formulae for RATs}\label{self}
2324:
2325: Astrophysically motivated situations require calculations
2326: of RATs for grains of different sizes and at many wavelengths. This
2327: requires rather intensive numerical computations. Our encouraging
2328: results with AMO motivate us to consider whether we can predict the
2329: scalings of torques.
2330:
2331: Dolginov \& Mytrophanov (1976) associated RATs with the scattering
2332: of right and left handed photons by a grain. For this model one should
2333: conjecture that
2334: amplitude of RATs decreases rapidly with increasing of $\lambda/a_{eff}$
2335: as the grain-photon interactions get into the random walk regime. In other
2336: words, a sharp peak is expected for the torque efficiency for photons
2337: with $\lambda\sim a_{eff}$. AMO, on the contrary, suggests of a
2338: linear increase for $\lambda \ll a_{eff}$. Our computation, however, indicate that
2339: RAT gets constant for $\lambda < a_{eff}$. However, both computational and DDSCAT
2340: intrinsic limitations do not allow us to perform calculations
2341: for $\lambda/a_{eff}<0.1$. More studies with other techniques, e.g., ray-tracing
2342: one, are necessary.
2343:
2344: For grains with $\lambda\gg a_{eff}$, Lazarian (1995) suggested that the
2345: scaling of RATs efficiency should be $\sim (\lambda/a_{eff})^{-4}$.
2346: Such considerations disregard
2347: the variations of the optical constants. Therefore the testing is essential.
2348:
2349: We calculated RATs in function of $\lambda/a_{eff}$ for three
2350: grain shapes. Shape 1 and 2 are shown in Fig. \ref{f18}, and a hollow grain
2351: is produced from the shape 1 by removing the core of grain. The latter is done
2352: to reduce the amount of necessary DDSCAT computations while achieving
2353: smaller $\lambda/a_{eff}$ ratios.
2354:
2355: We use both the dielectric function for the smoothed astronomical
2356: silicate (DW97; Weingartner \& Draine 2001; Cho \& Lazarian 2005) and constant optical constant. Results are shown in
2357: Figs. \ref{f26} and \ref{f27}.
2358: \begin{figure}
2359: \includegraphics[width=0.49\textwidth]{f33a.ps}
2360: \includegraphics[width=0.49\textwidth]{f33b.ps}
2361: \caption{{\it Upper panel}: RATs for grain shape 1, as the function of
2362: wavelength to grain size for the
2363: direction $\psi=0^{0}$. We see that RATs decrease steeply as
2364: $\lambda/a_{eff}$ increases. Around $\lambda/a_{eff}\sim 1$ the dependence
2365: flattens. Several positions at which
2366: RATs decrease dramatically correspond to the peculiarities of
2367: dielectric susceptibility. If those peculiarities are
2368: ignored, RATs are self-similar. {\it Lower panel}:
2369: The same as the upper panel but for
2370: shape 2. RATs can be fitted by analytical functions with $\eta =3 \mbox{ or }
2371: 4$ (see equation \ref{eq86}).}
2372: \label{f26}
2373: \end{figure}
2374: \begin{figure}
2375: \includegraphics[width=0.49\textwidth]{f34a.ps}
2376: \includegraphics[width=0.49\textwidth]{f34b.ps}
2377: \caption{RATs for shape 1 and hollow grain with constant refractive index:
2378: self-similarity exhibits extremely well. They can be fitted by a step function
2379: and a power function with $\eta=3, 4$ ({\it Upper Panel}) and $\eta=4, 5$
2380: ({\it Lower Panel}) where $\eta$ is defined in equation (\ref{eq86}).}
2381: \label{f27}
2382: \end{figure}
2383:
2384: We see that the approximate
2385: self-similarity (i.e. the dependence on $\lambda/a_{eff}$)
2386: is an intrinsic property of radiative torques. When optical constant
2387: changes as a function of wavelength, RAT efficiencies for different grain
2388: sizes mostly
2389: differ at wavelengths corresponding to resonance absorption features.
2390:
2391: In addition, RATs have nearly constant magnitude as $\lambda \sim a_{eff}$, and decrease steeply with the ratio of wavelength to grain
2392: size. This is because the scattering of photon by irregular grains are strongest
2393: as $\lambda \sim a_{eff}$. We can fit our calculations for RATs (see Figs
2394: \ref{f26}, \ref{f27}) by a simple function
2395: given by
2396: \begin{align}
2397: Q_{\Gamma}&=0.4 \mbox{ for } \frac{\lambda}{a_{eff}}<1.8,\nonumber \\
2398: &=0.4(\frac{\lambda}{a_{eff}})^{-\eta} \mbox{ for } \frac{\lambda}{a}>1.8,\label{eq86}
2399: \end{align}
2400: where $\alpha$ is the spectral index that according to
2401: Figs \ref{f26} and \ref{f27} is between 3 and 4. The first case provides
2402: a good fit in the whole range of $\lambda/a_{eff}$. In contrast, the latter case
2403: gives better fit for the range of $\lambda/a_{eff} <20$. Cho \& Lazarian (2006)
2404: use the former
2405: fit formulae to calculate polarization degree for accretion disks because
2406: grains there are widely believed to be very large.
2407:
2408: For the case when ${\bf a}_{1}$ makes an angle $\Theta=45^{0}$ with respect to ${\bf
2409: k}$, we found that the self similarity is also valid. However, the curve of RATs is
2410: shallower, and can be fitted by a power index $\eta=-2$.
2411:
2412:
2413:
2414: To study the efficiency of the self similarity, we calculate rotation velocity of
2415: grains induced by RATs in which RATs are directly computed from DDSCAT and derived
2416: from the self similarity assumption. We use radiation intensity of a molecular cloud (see Mathis
2417: 1983) to calculate RATs for different optical depths $A_v$. Resulting rotation speed obtained with these two methods (see
2418: Fig. \ref{f28}) shows clearly that the self similarity provides
2419: a fair agreement between exact calculations and those based on
2420: the self-similarity arguments. This
2421: allows to reduce the DDSCAT computational efforts substantially.
2422:
2423: \begin{figure}
2424: \includegraphics[width=0.49\textwidth]{f35a.ps}
2425: \includegraphics[width=0.49\textwidth]{f35b.ps}
2426: \caption{Ratio of rotation speed to thermal rotation speed with respect to
2427: grain size obtained with exact RATs from DDSCAT ({\it Upper panel}) and using
2428: the self-similarity ({\it Lower panel}). Here, radiation direction is assumed to be parallel to the maximal inertia axis
2429: ${\bf a}_{1}$, that is constrained to be
2430: frozen to angular momentum ${\bf J}$. The figures show the rapid increase of the rotational
2431: speed with grain size, and even in the cloud core with high visual
2432: extinction $A_{V}=10$, large grain still have suprathermal rotation rate,
2433: ensuring that grain alignment is not susceptible to random collision by atomic gas.}
2434: \label{f28}
2435: \end{figure}
2436:
2437: \section{Discussion}\label{discuss}
2438:
2439: Our study above have approached an important problem of the RAT alignment mechanism by studying the fundamental properties of RATs. The goal of such studies
2440: was to change
2441: the status of the RAT alignment from an empirical fact to a theoretically
2442: understood process. Our work indicates that grain helicity is an essential
2443: property of realistic grains, which suggests that it should be accounted not only for RATs, but
2444: for mechanical alignment as well.
2445:
2446: \subsection{Evolution of ideas on RATs}
2447:
2448: For the first time grain spin-up due to differential scattering
2449: of left and right handed photons was considered in Dolginov (1972).
2450: The suggestion was limited to quartz grains, however. In Dolginov \&
2451: Mytrophanov (1976) it was realized that grains of made of more accepted
2452: astrophysical materials can be spun up, provided that they are somewhat
2453: twisted. This ground-breaking
2454: study qualitatively considered both the possibility
2455: of the RAT alignment with respect to magnetic field and the radiation beam.
2456: Regretfully, it was not appreciated in its time.
2457:
2458: A reason of the lukewarm response of the community to this work was
2459: probably related to both the absence of reliable estimates of RATs
2460: efficiencies and to the fact that some features of grain dynamics were
2461: unknown at that time. For instance, Dolginov \& Mytrophanov (1976) claimed
2462: that prolate grains should align along magnetic field, while oblate grains
2463: should align perpendicular to magnetic field. The former conclusion was based
2464: on the assumption that grain rotation along the axis of minimal inertia is
2465: stable. It is only later that Purcell (1979)
2466: discovered so-called Barnett relaxation and concluded that only rotation
2467: about the axis of maximal inertia is stable\footnote{The Barnett magnetization
2468: that induced Ed Purcell to think about the relaxation
2469: was described in the same Dolginov \& Mytrophanov (1976) paper.}. On the basis of this Lazarian
2470: (1995) claimed that the alignment of both prolate and oblate helical
2471: grains should happen with long axes perpendicular to magnetic field. However,
2472: it was not quite clear at that moment what makes the grain helical. In addition, Lazarian (1995) underestimated the importance of the RAT alignment.
2473:
2474: The quantitative stage of RAT studies was initiated by Draine \& Weingartner (DW96 and DW97).
2475: There RATs were shown to be generic for irregular
2476: (and therefore
2477: realistic) grains and the magnitude of RATs was reliably quantified.
2478: The alignment with respect to magnetic field was
2479: demonstrated in numerical simulations. On the basis of
2480: these simulations DW97 claimed that the RAT alignment
2481: was the dominant process responsible for polarization arising
2482: from dust in the diffuse ISM. However, the RAT alignment was an empirical
2483: fact in DW97 that was based on a limited sample. In this situation
2484: additional important
2485: complexities arising from including thermal wobbling of grains
2486: (WD03), gaseous bombardment etc. could not help to improving our
2487: physical understanding of the RAT alignment.
2488:
2489: What we have done in the present paper is we attempted to clarify the
2490: basics of the RAT alignment by seeking the generic properties of RATs that
2491: induce the alignment. We used analytical modeling which was tested with
2492: numerical DDSCAT simulations. As we discuss further, we hope that this work
2493: contributes to both more intuitive understanding of the alignment and to
2494: further elaborating of the mechanism in order to get precise
2495: predictions of the alignment degree in different astrophysical situations.
2496:
2497: It is worth mentioning that even we when know the analytical form of the torques,
2498: the dynamics of the system does not get completely trivial. It definitely
2499: exhibits interesting properties.
2500:
2501:
2502: \subsection{Our approach}
2503:
2504:
2505: While RATs were initially treated as a quantum effect arising
2506: from the difference of scattering of left and right handed photons (see
2507: Dolginov \& Mytrophanov 1976), above we presented an entirely
2508: classical model of a grain (see Fig.~\ref{f2}) that reproduces well their
2509: properties (see Fig~\ref{f61}).
2510:
2511: The gist of our approach above is to consider the basic generic
2512: properties of RATs and to relate of these properties with
2513: the RAT alignment. AMO plays a central role in our considerations.
2514: Our toy model of a helical grain allowed an analytical description, which
2515: enabled us to treat RATs analytically.
2516: In our study we concentrated on the properties of RAT components
2517: in the scattering system (see Fig.~\ref{f1}),
2518: i.e. on the properties of $Q_{e1}$, $Q_{e2}$ and
2519: $Q_{e3}$. These components show a remarkable similarity for grains of very
2520: different shapes (see Fig.~\ref{f23a}) and AMO (see Fig.~\ref{f61}). The chi
2521: squared test we present in \S \ref{sec94} returns the mean value of $\chi^{2}$ for
2522: both $ Q_{e1}$ and $Q_{e2}$ about $0.2$ (see Fig. \ref{chi_test0}). This provided
2523: us with the empirical justification of using AMO for studies of RAT alignment.
2524: AMO provides us with both useful intuitive model
2525: to think about the alignment and
2526: analytical formulae that allow straightforward quantitative calculations.
2527:
2528:
2529: We find that the basic properties of RATs obtained with AMO are very similar
2530: to the basic properties of RATs for arbitrary shaped grains. This, for
2531: instance, allows
2532: us to talk about helicity of grains being the most important attribute for the
2533: RAT alignment.
2534:
2535:
2536:
2537:
2538: Our major
2539: goal above was to get a better understanding of the physics of the
2540: RAT alignment. To do this we adopt a model, similarly to one in DW97, that disregards the wobbling
2541: of the grain axes with respect to angular momentum direction
2542: (cf. Lazarian 1994;
2543: Lazarian \& Roberge 1997; LD99ab), but treats
2544: crossovers differently, i.e. in the spirit of the Spitzer \& McGlynn (1979)
2545: model. This provides a substantial change in the dynamics of grains. For instance,
2546: we do not observe cyclic maps reported in DW97.
2547:
2548: While studying the properties of RATs
2549: we addressed the question of the necessary conditions for
2550: RAT alignment to happen and to fail.
2551: We also study the RAT alignment that takes place in the radiative-dominated
2552: environments, where the direction of radiation
2553: defines the axis of alignment (see Dolginov \& Mytrophanov 1979;
2554: Lazarian 2003). Such an alignment is both astrophysically important
2555: and provides a good insight into the physics of the RAT alignment.
2556:
2557: \subsection{Accomplishments and Limitations of the present study}
2558:
2559: We feel that our major accomplishment in the paper above was establishment of
2560: the analytical form of RATs and clarification of the role of different RAT
2561: components. We hope that
2562: AMO unveils the mystery that have surrounded the RAT alignment from the time
2563: of the mechanism introduction.
2564:
2565: Another important conclusion that follows from AMO is
2566: that the RAT alignment is not limited to
2567: grains with $\lambda \sim a_{eff}$, as it was believed
2568: before. As the similarities
2569: between the torques that we obtained for AMO in the $\lambda \ll a_{eff}$ limit
2570: and for irregular grains in the $\lambda \ge a_{eff}$ limit are striking,
2571: our work shows that the RAT alignment should take place also for large
2572: grains, that are present in accretion disks and dark cloud cores (see
2573: Cho \& Lazarian 2005, 2007).
2574:
2575: Obtaining generic properties of RATs makes the RAT alignment more a
2576: predictable theory and opens avenues for further theoretical advances,
2577: e.g. including thermal fluctuations, random bombardment, H$_2$ torques
2578: etc. Our
2579: establishing of a subdominant nature of one of the RAT components, namely
2580: $Q_{e3}$, simplifies the theoretical treatment of RATs.
2581: Insights into the generic properties of two other components allow
2582: to reduce the amount of numerical computations necessary to determine
2583: the degree of achievable alignment. For instance, in the
2584: current paper, we found that the existence of high-$J$ attractor points
2585: depends on the $Q_{e1}^{max}/Q_{e2}^{max}$ ratio. For practical applications
2586: it is important, that the criterion for this established with AMO works well with irregular grains.
2587: We note parathentically, that while $Q_{e1}$ and $Q_{e2}$ demonstrate universal
2588: behavior, the aligning and spinning torques (see equations (\ref{eq61}) and (\ref{eq62}))
2589: that present their combinations do not demonstrate this.
2590:
2591: The present study identifies a parameter space for which the
2592: RAT alignment may be suspected to be ``wrong'', i.e. to happen with long
2593: axes parallel to magnetic field. In addition, it provides simple scalings for
2594: RAT dependences on the ratio of the radiation wavelength to the grain
2595: size.
2596:
2597: Our study reveals new properties of the RAT alignment. First of all, the
2598: alignment may be fast, i.e. happen in a small fraction of gas damping time (see
2599: \S \ref{fast}). This has important consequences for the environments with fast
2600: changing radiation, i.e. circumstellar regions, interstellar medium in the
2601: vicinity of supernovae flashes etc. Moreover, we could see, that the alignment
2602: is different depending whether the initial angular momentum is small or large
2603: (see Figs \ref{f12} and \ref{f12*}).
2604:
2605: The approximate self-similarity of RATs (see \S 10) is another
2606: practically useful property of RATs. Combined with the established universality
2607: of the functional form of the components
2608: $Q_{e1}$ and $Q_{e2}$ and the established dependences of
2609: the properties of trajectory maps on the ratio of these components, this allows
2610: radically reduce DDSCAT calculations that may be necessary to find the
2611: expected
2612: degree of alignment for an ensemble of realistic grains subjected to a
2613: realistic
2614: radiation field. In fact, we find that RATs change the ratio of
2615: $Q_{e1}^{max}/Q_{e2}^{max}$ with $\lambda$ and this is the most important
2616: difference that the variation of the radiation wavelength entails. The
2617: functional forms of the torques do not change much and can be well approximated
2618: with those of AMO.
2619:
2620: One limitation of AMO is the upper limit of the ratio $Q_{e1}^{max}/Q_{e2}^{max}$, that
2621: makes AMO more appropriate to irregular grains with $\lambda >3 a_{eff}$ and $\lambda<a_{eff}$. It
2622: indicates that, though AMO is established based on the geometric optics, i.e.,
2623: $\lambda \ll a_{eff}$, it is also applicable for the opposite limit.
2624:
2625:
2626: In more general terms, our study proved that irregular
2627: grains can be characterized by helicity. Grain rotation
2628: provides the averaging that defines the helicity axis, while the
2629: irregularities define whether the helicity is left or right. The phase
2630: trajectories of grains that are the mirror-symmetric images of
2631: each other are mirror-symmetric (see
2632: Fig.~\ref{f8}). As expected, the torque component
2633: $Q_{e3}$, unlike the other two components, coincide
2634: for an irregular grain and its mirror-symmetric image (see the lower panel in Fig.~{\ref{f23a}). Indeed, this component is subdominant for most of the alignment
2635: processes and not related to grain helicity.
2636:
2637:
2638: In our study we do not directly address the grain alignment efficiency.
2639: Some statements can be made, however. For instance,
2640: when grains are aligned rotating suprathermally the direction of ${\bf J}$
2641: is immune to the randomization arising from the gaseous bombardment. In
2642: addition in this case,
2643: ${\bf J}\|{a}_1$ provides a good approximation. We find, however,
2644: that an appreciable subset of grains rotates with thermal velocities.
2645: For those the internal randomization of grain axes in respect to ${\bf J}$
2646: may be important. Does this signify a new crisis
2647: of the grain alignment theory? We do not believe so. Even in the absence
2648: of high-$J$ attractor points RATs will drive
2649: ${\bf J}$ back to low attractor points, which in most cases, as we discussed
2650: in the paper, correspond to the preferential alignment of grains with long
2651: axes perpendicular to magnetic field. As for the internal alignment,
2652: according to Roberge \& Lazarian (1997) for typical interstellar
2653: grains, this alignment is tangible even for $J\sim J_{th}$.
2654: A detailed study of the attainable degrees of alignment
2655: is provided elsewhere.
2656:
2657: We have not discussed RATs of the strongly absorbing materials, e.g. graphite. We expect the torque components to show more irregularity for such
2658: grains. As the grain alignment theory matures and extends to the
2659: environments different from molecular clouds and diffuse interstellar gas (see
2660: Lazarian 2007),
2661: the importance of the studies of wider range of materials will get
2662: more pressing.
2663:
2664:
2665: \subsection{Rates of Alignment and Rotation}
2666:
2667: As we mentioned above the RAT alignment can happen on time scales
2668: much shorter than the gaseous damping time. This finding corresponds to the
2669: notion in Dolginov \& Mytrophanov (1976), that the alignment happens on the
2670: time scale that is required for the radiative torques to deposit a grain
2671: with the angular momentum of the order of its initial angular momentum. Such a
2672: fast alignment makes grains good tracers of magnetic field when radiation
2673: direction changes quickly.
2674:
2675: The fast alignment takes place for low-$J$ attractor points of the grain
2676: phase trajectory map. These are the most probable attractor for the grains
2677: to end up with. Thus, most grains do not rotate suprathermally when subject
2678: to RATs. In this sense the RAT alignment tends to minimize grain angular
2679: momentum.
2680:
2681:
2682: While we expect that in the presence of thermal wobbling and
2683: gaseous bombardment most grains will rotate thermally,
2684: there is a radical difference between this effect and the effect
2685: of thermal trapping discussed in LD99ab. The effect
2686: of thermal trapping there is based on the compensation of the Purcell rocket
2687: torques, e.g. those related to H$_2$ formation, by thermal flipping of grains.
2688: The more efficient thermal flipping, the more efficient is the trapping and
2689: the less chance of a grain to get high angular momentum. On
2690: the contrary, we have seen in \S \ref{kalign} and \ref{balign} that without thermal wobbling, the significant fraction of grains
2691: ends up in the state of $J=0$. In other words, thermal fluctuations increase
2692: the value of $J$ to a {\it higher}, i.e. thermal value.
2693:
2694: In spite of the fact, that most of the grains tend to rotate with velocities
2695: much less than the maximal velocities, $\omega_{max}$,
2696: that RATs can spin the grain up, we
2697: believe that the parametrization of the alignment in terms of $\omega_{max}/\omega_T$, where $\omega_T$ is the thermal rotational velocity,
2698: may be a rough practical way of describing alignment. Indeed, the
2699: above ratio
2700: reflects the relative importance of RATs compared with those
2701: related to gas. When RATs force the grain into a low-$J$ attractor
2702: point, their ability to do this would also depend on this ratio. A further
2703: research should reveal more sophisticated and precise parametrization of
2704: the RAT alignment, however. This parametrization is necessary, for instance,
2705: to predict the expected alignment from the numerical simulations of magnetized
2706: molecular clouds (see Cho \& Lazarian (2005); Pelkonen et al. (2007); Bethell et al. (2007)).
2707:
2708:
2709: We have discussed in \S \ref{sec5} that
2710: for some phase trajectories high attractor points are available.
2711: The suprathermally rotating grains correspond to high-$J$ attractor
2712: points. It takes them about 3 damping times to reach such points (see
2713: also DW97). However, our analysis shows that grains get aligned even before
2714: they reach high-$J$ attractor points. Therefore the RAT alignment can happen over
2715: shorter time scales for all grains provided that the radiation is intensive enough.
2716:
2717: The predominance of low-$J$ attractor points has consequences that go beyond the
2718: problems of grain alignment. If grains rotate slowly, then loosely connected
2719: conglomerates constituting fractal grains can exist. Ever since the classical
2720: work by Purcell (1979), the suprathermal rotation had been thought to destroy such
2721: grains. When LD99ab showed that Purcell's torques may not be capable to spin-up
2722: grains less than $10^{-4}$~cm, it was still thought that radiative torques can do
2723: the job. Our work questions this (see also WD03).
2724:
2725:
2726:
2727: \subsection{Direction of Alignment}
2728:
2729: The alignment may happen with respect to radiation
2730: rather than to magnetic field if the precession induced by radiative torques
2731: is faster than the Larmor one (see \S \ref{sec9}).
2732: We found, that in
2733: the presence of magnetic field the alignment can still happen in
2734: respect to the direction of the beam or, equivalently, the direction
2735: of the anisotropy of radiation, provided that the rate of
2736: precession arising
2737: from the radiative torques is faster than rate of the Larmor precession.
2738: This is the case of comets sufficiently close to the Sun, ISM in the vicinity of
2739: supernovae and some circumstellar regions. Over vast expanses
2740: of diffuse interstellar medium and molecular clouds, however, the generic RAT
2741: alignment is with respect to magnetic field, thus enabling easy tracing of
2742: magnetic fields via polarimetry.
2743: Note, that our study shows
2744: that many features characteristic
2745: of the alignment in the absence of magnetic field carry over
2746: to the case when magnetic field is present.
2747:
2748:
2749: Our important finding is that while the generic alignment is
2750: ``right'', i.e. with the long grain axes perpendicular to magnetic
2751: field, for a range of angles $\psi$ between
2752: the magnetic field and the direction of the beam around $\psi=\pi/2$,
2753: the alignment may be ``wrong'', i.e. it happens with long grain axes parallel to magnetic field. However, the
2754: range of the angles is rather narrow.
2755: As a result, we do not expect the effect of ``wrong alignment'' to persist
2756: when grains undergo thermal wobbling. This wobbling is likely to
2757: vary the direction of the grain axes with respect to the light direction
2758: beyond the range angles in which the alignment is ``wrong''.
2759:
2760:
2761: \subsection{Magnetic Field and Gas Streaming}
2762:
2763: Unlike DW97, in the paper above we disregarded the effects of paramagnetic
2764: alignment altogether. When we consider dynamically important field, its
2765: only effect is to induce averaging due to Larmor precession. We believe
2766: that our approach is correct, as
2767: for paramagnetic grains the effects
2768: of paramagnetic relaxation are marginal on times over which the
2769: RAT alignment takes place.
2770:
2771: Gas streaming can induce its own alignment direction.
2772: Dolginov \& Mytrophanov (1976) assumed that
2773: magnetic field or a gaseous flow defines the axis of
2774: alignment depending on the ratio of Larmor precession time to
2775: that of mechanical
2776: alignment. On the basis of our study of $Q_{e3}$ with AMO,
2777: we believe that a more physically
2778: motivated distinction is related to the rate of Larmor precession
2779: versus the precession arising from the mechanical
2780: analogy of the $Q_{e3}$ torque. Such torque for an spheroidal grain
2781: can be obtained from formulae in Appendix~B by substituting the value of the
2782: gas atom momentum $mv$ instead of the photon momentum.
2783: The corresponding precession timescales ratio is (see equation \ref{eq79})
2784: \begin{align}
2785: \frac{t_{flow}}{t_{B}}= 3\times 10^{5}\frac{\hat{\rho}\hat{\chi}\hat{T}^{-0.5}\hat{a}_{-5}^{-0.5}}{\hat{\alpha}\hat{v}^{2}_{flow}\hat{n}_{H}}\frac{\hat{B}}{\hat{Q}_{e3, gas}},
2786: \label{t_fb}
2787: \end{align}
2788: where $\hat{Q}_{e3, gas}$ is the third component of torques induced by the
2789: gaseous flow, which is the analog of $Q_{e3}$ for RATs.
2790: In equation (\ref{t_fb}), $\alpha=\hat{\alpha}\times 0.1$ is the probability of elastic collision, $v_{flow}=\hat{v}_{flow}\times v_{thermal}$ is gas flow velocity,
2791: $n_{H}=\hat{n}_{H}\times 30$, $T=\hat{T}\times 100$ is gas density and temperature, respectively.
2792: equation (\ref{t_fb}) indicates that for sufficiently intensive gaseous flows, for
2793: instance, $\hat{v}_{flow}>10^{2} $, the alignment
2794: will indeed happen with respect to the flow direction. Note, that
2795: we predict that gaseous flows may define the direction of alignment for a
2796: wider parameter range compared to that in
2797: Dolginov \& Mytrophanov (1976). Moreover,
2798: we claim that mechanical flows can define the axis of alignment even for
2799: subsonic flow velocities, i.e. at those velocities for which the Gold alignment
2800: and its modifications (cf. below, however) are marginal.
2801:
2802:
2803: A conceivable situation is that the gaseous bombardment arising from
2804: grain streaming defines the axis of alignment, while RATs do the alignment
2805: job. This situation takes place when $Q_{e3, RATs}$ is less than $Q_{e3, gas}$,
2806: which defines
2807: \begin{equation}
2808: \frac{t_{flow}}{t_{k}}\sim \frac{u_{rad}}{\alpha m_{H}v_{flow}^{2}n_{H}}\sim 10^{1}\frac{\hat{u}_{rad}}{\hat{\alpha}\hat{v}_{flow}^{2}\hat{T}\hat{n}_{H}},\label{flrad}
2809: \end{equation}
2810: where $u_{rad}=\hat{u}_{rad}\times u_{ISRF}$.
2811:
2812: This can be the case of alignment in a part of comet
2813: atmosphere\footnote{Another case also relevant to the comet atmosphere
2814: is that the grains have electric
2815: dipole moments, while comet atmosphere has electric field. Then the comet
2816: electric field defines the axis of alignment.}. Naturally, combining
2817: equations~(\ref{t_fb}) and (\ref{flrad}) it is possible to establish when streaming
2818: defines the alignment axis in spite of the presence of magnetic field.
2819:
2820:
2821:
2822: \subsection{AMO and Mechanical Alignment of Helical Grains}
2823:
2824: Our present paper is devoted to RATs and the alignment that they entail.
2825: However, our consideration of a helical grain is quite general. In fact, the
2826: functional dependence of the torques that we obtain for our model grain
2827: is valid when atoms rather than photons are reflected from the mirror. Therefore
2828: we may predict that for elastic gas-grain collisions the helical
2829: grains\footnote{The mechanical alignment of helical grains was briefly
2830: discussed in Lazarian (1995) and
2831: Lazarian, Goodman \& Myers (1997), but was not elaborated there.} will
2832: align with long grain axes perpendicular to the flow
2833: in the absence of magnetic field and with long axes perpendicular to ${\bf B}$,
2834: when dynamically important magnetic field is present. If atoms attach to the
2835: grain surface and then are thermally ejected from it, this changes the values
2836: of torques by a factor of order unity. The only way that the uncompensated
2837: torques can vanish for a helical grain is if the correlation is lost between
2838: the place at which the atom strikes the grain surface and leaves it\footnote{Even
2839: in this case the local anisotropies of the surface at the place of atom
2840: impact can result in effective helicity similar to the case of damped
2841: oscillator in Fig. \ref{f10} (lower panel)}. The latter is rather improbable for sufficiently
2842: large grains. As the result, one has to conclude that helicity is a generic
2843: property of the interaction of irregular grains and atomic flows.
2844:
2845: This conclusion alters substantially the paradigm of grain motions in
2846: diffuse gas. Since the time when Gold (1951) proposed his first simple
2847: model of mechanical alignment, it was considered essential to have
2848: supersonic grain-gas velocities to achieve any tangible alignment.
2849: Indeed, both the Gold original process and the ones proposed for suprathermally
2850: rotating grains, namely cross-sectional and crossover alignments (Lazarian
2851: 1995) require the supersonic drift to ensure that the momentum deposited
2852: in a regular way exceeds one deposited due to thermal atomic motions.
2853: This is not a requirement for the alignment of helical grains! For those
2854: the regular momentum is deposited in proportion to the number of collisions,
2855: while the randomization adds up only as a random walk. In fact, the difference
2856: between the mechanical alignment of spheroidal and helical grains is similar
2857: to the difference between the stochastic
2858: Harwit (1970) alignment by stochastic absorption
2859: of photons and the RAT alignment. While the
2860: Harwit alignment requires very special
2861: conditions dominate, the regular RATs easily beat randomization.
2862:
2863: Similarly as in the case of RATs, where it is frequently
2864: possible to disregard the Harwit process, it should be possible to disregard
2865: the Gold-type processes\footnote{
2866: This alignment tends to minimize grain cross section,
2867: which
2868: means, for instance, that for grains streaming along magnetic fields the non-helical stochastic torques will tend to align grains with longer axes parallel to
2869: magnetic field, while helical torques will tend to align in the perpendicular
2870: direction.}
2871: for the alignment of helical grains. Note, that because of the property of
2872: helicity not to change sign during grain flipping, we do not expect to
2873: observe the thermal trapping effects described in Lazarian \& Draine (1999a)
2874: to be present for the mechanical spin-up of helical grains. The effects
2875: that decrease the efficiency of the mechanical alignment of helical grains
2876: are discussed elsewhere.
2877:
2878: In typical conditions of diffuse interstellar medium,
2879: the mechanical alignment of helical grains tends to act to
2880: align grains in the same direction as the RATs, i.e. with longer axes
2881: perpendicular to magnetic field. The relative role of the two mechanisms
2882: should be revealed by further research. The currently available data
2883: (see Lazarian 2007) agrees with the RAT mechanism being the primary source
2884: of alignment. However, the situations are possible, when mechanical
2885: alignment reveals magnetic field, when RATs fail to do so.
2886:
2887:
2888: \section{Summary}
2889: In this paper, we studied the properties of RATs, and how different RAT
2890: components affect the grain alignment. Briefly, our results are as follows:
2891:
2892: 1. We found that a simple model of a helical grain
2893: which consists of a reflecting spheroidal
2894: grain with an attached
2895: mirror reproduces well the functional
2896: dependences of RATs obtained for irregular
2897: grains using DDSCAT.
2898:
2899:
2900: 2. From the generic properties of RATs we predicted
2901: the preferential alignment of grains with long axes
2902: perpendicular to the direction towards the source of light, provided
2903: that magnetic field effect is subdominant.
2904:
2905: 3. The magnetic field is important and defines the
2906: axis of alignment when it induces the Larmor precession
2907: that is faster than the precession arising from the $Q_{e3}$-component of RATs,
2908: i.e. the component not related to grain helicity. This component is present for
2909: spheroidal grain, for instance.
2910:
2911:
2912: 4. When magnetic field is important, RATs tend to provide both the
2913: alignment of long grain axes perpendicular to magnetic field. With or
2914: without magnetic field, most of grains are driven to the low angular momentum
2915: attractor points, as RATs align grains. Grains can be driven to
2916: the low-$J$ attractor points on the time scales much less than the
2917: gaseous damping time. The very
2918: existence of the high angular
2919: momentum attractor points, and therefore grains rotating much faster than the
2920: thermal velocity, is not a default and
2921: depends on the ratio $Q_{e1}$ and $Q_{e2}$ components.
2922:
2923:
2924: 5. RATs can also induce ``wrong alignment'', i.e., the alignment
2925: with long grain
2926: axes parallel to magnetic field.
2927: The range of angles
2928: for ``wrong
2929: alignment'' is narrowly centered around the $\pi/2$ angle between
2930: the direction of light and magnetic field. This
2931: range is expected to be
2932: diminished when thermal fluctuations are accounted for. Thus the
2933: RAT alignment is capable to account for most of the observed polarization.
2934:
2935: 6. RATs exhibit approximate
2936: self-similarity that allows one to be expressed them as a function of
2937: the ratio of the wavelength to grain size. The dynamics of the grains
2938: can be reproduced relatively accurately when the
2939: self-similarity is used.
2940:
2941: 7. The RAT alignment is a particular case of the alignment of the
2942: helical grains. Therefore our results can be generalized to describe
2943: the mechanical
2944: alignment of irregular grains. Such an alignment is efficient for both
2945: supersonic and subsonic gaseous flows. The
2946: mechanical alignment may happen either with respect to the magnetic field or
2947: the direction of the flow depending on the rates of precession that are
2948: induce by the flow and magnetic field, respectively.
2949:
2950:
2951:
2952: \section*{Acknowledgments}
2953: We acknowledge the support by the NSF Center for Magnetic Self-Organization in Laboratory and Astrophysical
2954: Plasmas. AL acknowledge a partial support by the NSF grant AST 0507164.
2955: We thank the anonymous referee for his/her comments which improved the paper,
2956: especially for comments on the symmetries of the RAT components.
2957:
2958:
2959:
2960:
2961: \appendix
2962:
2963:
2964: \section{Reflecting Oblate Spheroid}
2965:
2966: The toy model of a grain (see Fig. \ref{f2}) allows us to derive analytical formulae for RATs.
2967: Let us consider RATs for an oblate reflecting spheroid characterized by semi-axes $a, b$
2968: where $s =a/b<1$ (see Fig. \ref{ap1}). ${\bf a}_{1},
2969: {\bf a}_{2}, {\bf a}_{3}$ are principal axes of the spheroid with moments of
2970: inertia $I_{1} > I_{2}=I_{3}$, respectively. Assuming that a photon beam of
2971: wavelength $\lambda$ is shined along
2972: ${\bf k} \| \hat{e}_{1}$, the photon reflections happen on the spheroidal surface at the location
2973: determined by a normal unit vector $\mn$ and a radius $\mr$. For the sake of
2974: simplicity, a perfect reflection is always assumed in this study.
2975: \begin{figure}
2976: \includegraphics[width=0.49\textwidth]{f36a.eps}
2977: \includegraphics[width=0.49\textwidth]{f36b.eps}
2978: \caption{{\it Upper panel} represents the coordinates and vectors for the
2979: spheroid. {\it Lower panel} shows the fitting function $K(\Theta,e)$ for $Q_{e3}$, depending on the eccentricity of the spheroid $e$ and its angle with light beam $\Theta$.}
2980: \label{ap1}
2981: \end{figure}
2982: The location of impact on the grain surface is specified by the radius
2983: ${\mr}$ and the normal vector ${\mn}$, which are, respectively, given by
2984: \begin{align}
2985: {\mr}&=a\ms\eta\ma_{1}+b\mc\eta\mc\xi\ma_{2}+b\mc\eta\ms\xi\ma_{3},\label{a1}\\
2986: {\mn}&=a_{1}\ms\eta\ma_{1}+b_{1} \mc\eta\mc\xi\ma_{2}+b_{1} \mc\eta\ms\xi\ma_{3},\label{a2}
2987: \end{align}
2988: where $a_{1}=[\ms^{2} \eta +(1-e^{2}\mc^{2}\eta]^{-1/2}, b_{1}=a_{1}(1-e^{2})^{1/2}$,
2989: and $e$ is the eccentricity of the spheroid, $\xi=[0,\pi], \eta=[-\pi/2,
2990: \pi/2]$ (see Fig. \ref{ap1}; see also Roberge et al. 1993).
2991:
2992: Due to the symmetry around $\ma_{1}$, we only need to find RATs for a single
2993: rotation angle, e.g. for $\beta=0$. Therefore, when the grain axis ${\bf
2994: a}_{1}$ makes an angle $\Theta$ with
2995: respect to the photon beam, we have
2996: \begin{align}
2997: \ma_{1}&=\mc\Theta\me_{1}+\ms\Theta\me_{2},\label{a3}\\
2998: \ma_{2}&=-\ms\Theta\me_{1}+\mc\Theta\me_{2},\label{a4} \\
2999: \ma_{3}&=\me_{3}.\label{a5}
3000: \end{align}
3001: Substituting equations (\ref{a3})-(\ref{a5}) into (\ref{a1}) and (\ref{a2}) we
3002: obtain
3003: \begin{align}
3004: {\mr}&=(a\mc\Theta\ms\eta-b\ms\Theta\mc\eta\mc\xi)\me_{1}\nonumber\\
3005: &+(a\ms\Theta\ms\eta+b\mc\Theta\mc\eta\mc\xi)\me_{2}\nonumber\\
3006: &+b\mc\eta\ms\xi\me_{3},\label{a6}\\
3007: {\mn}&=(a_{1}\mc\Theta\ms\eta-b_{1}\ms\Theta\mc\eta\mc\xi)\me_{1}\nonumber\\
3008: &+ (a_{1}\ms\Theta\ms\eta+b_{1}\mc\Theta\mc\eta\mc\xi)\me_{2}\nonumber\\
3009: &+b_{1} \mc\eta\ms\xi\me_{3}.\label{a7}
3010: \end{align}
3011: Hence, RAT produced by the photon beam is defined by
3012: \bea
3013: d{\bf \Gamma}_{rad}=\gamma \mr \times \Delta {\bf P}={-2p} \gamma F dA ({\bf k}.\mn)[\mr
3014: \times \mn], \label{a8}
3015: \ena
3016: where $\gamma$ is the anisotropy degree of radiation field, $p_{ph}$ is the
3017: momentum of each photon, $F$ is the flux of the incident light beam, and $dA$ is an area
3018: element on the grain surface given by
3019: \begin{align}
3020: dA=eb^{2} f(\eta)\mc\eta d\eta d\xi,\label{a9}
3021: \end{align}
3022: where $f(\eta)=\sqrt{\frac{1-e^{2}}{e^{2}}+\mss\eta}$.
3023:
3024: Therefore
3025: \begin{align}
3026: d{\bf \Gamma}_{rad}&=-2p_{ph} \gamma F (\me_{1}.\mn) dA
3027: [(r_{2}N_{3}-r_{3}N_{2})\me_{1}+(r_{3}N_{1}-r_{1}N_{3})\me_{2}\nonumber\\
3028: &+(r_{1}N_{2}-r_{2}N_{1})\me_{3}] .\label{a10}
3029: \end{align}
3030: Substituting $p_{ph}=\frac{h}{\lambda}, F=n_{ph}
3031: c=\frac{u_{rad}}{hc/\lambda} c=\frac{u_{rad}\lambda}{h}$, and
3032: $dA$ from equation (\ref{a9}) into the above equation, we obtain
3033: \begin{align}
3034: d{\bf \Gamma}_{rad}&=\frac{\gamma u_{rad}\lambda b^{2}}{2} [-\frac{4e}{\lambda}(\me_{1}.\mn)
3035: [(r_{2}N_{3}-r_{3}N_{2})\me_{1}\nonumber\\
3036: &+(r_{3}N_{1}-r_{1}N_{3})\me_{2}+(r_{1}N_{2}-r_{2}N_{1})\me_{3}]f(\eta)\mc\eta d\eta d\xi .\label{a11}
3037: \end{align}
3038: From equations (\ref{a7}) and (\ref{a8}), we get
3039: \begin{align}
3040: r_{2}N_{3}-r_{3}N_{2}&=\frac{ab_{1}-ba_{1}}{2}\ms2\eta\ms\xi\ms\Theta,\label{a12}\\
3041: r_{3}N_{1}-r_{1}N_{3}&=\frac{ab_{1}-ba_{1}}{2}\ms2\eta\ms\xi\mc\Theta,\label{a13}\\
3042: r_{1}N_{2}-r_{2}N_{1}&=\frac{ab_{1}-ba_{1}}{2}\ms2\eta\mc\xi.\label{a14}
3043: \end{align}
3044: Plugging equations (\ref{a12})-(\ref{a14}) into equation (\ref{a11}), and
3045: integrating over the surface, we obtain
3046: \begin{align}
3047: {\bf \Gamma}_{rad}&=\frac{\gamma u_{rad}\lambda b^{2}}{2}(Q_{e1}\me_{1}+Q_{e2}\me_{2}+Q_{e3}\me_{3}),\label{a15}
3048: \end{align}
3049: where the RAT components are given by
3050: \begin{align}
3051: Q_{e1}&=\int_{-\pi/2}^{\pi/2}\int_{0}^{\pi}\frac{-4e}{\lambda}\frac{ab_{1}-ba_{1}}{2}|a_{1}\ms\eta \mc\Theta\nonumber\\
3052: &-b_{1}\mc\eta\mc\xi\ms\Theta|
3053: \ms 2\eta \ms\xi \ms\Theta f(\eta) \mc\eta d\xi d\eta,\label{a16}\\
3054: Q_{e2}&=\int_{-\pi/2}^{\pi/2}\int_{0}^{\pi}\frac{-4e}{\lambda}\frac{ab_{1}-ba_{1}}{2}|a_{1}\ms\eta \mc\Theta\nonumber\\
3055: &-b_{1}\mc\eta\mc\xi\ms\Theta|
3056: \ms 2\eta \ms\xi\mc\Theta f(\eta) \mc\eta d\xi d\eta,\label{a17}\\
3057: Q_{e3}&=\int_{-\pi/2}^{\pi/2}\int_{0}^{\pi}\frac{-4e}{\lambda}\frac{ab_{1}-ba_{1}}{2}|a_{1}\ms\eta \mc\Theta\nonumber\\
3058: &-b_{1}\mc\eta\mc\xi\ms\Theta|
3059: \ms 2\eta \mc\xi f(\eta) \mc\eta d\xi d\eta.\label{a18}
3060: \end{align}
3061: Note, that we use the absolute value for ${\me}_{1}.{\mn}$ because
3062: in the integral over $\eta$, we always use the range $[-\pi/2, \pi/2]$.
3063:
3064: Due to the presence of the term $\ms\xi$ in equations (\ref{a16}) and
3065: (\ref{a17}), their integrals over the range $\xi=[0, \pi]$ vanish. The
3066: integral given by equation (\ref{a18}) provides us a function of $\Theta$ which can be fitted
3067: by a function $\ms 2\Theta$ and a fitting factor
3068: \bea
3069: Q_{e3}(\Theta)= \frac{2e a}{\lambda}(s^{2}-1) K(\Theta, e)\ms2\Theta,\label{ap19}
3070: \ena
3071: where $ K(\Theta, e)$ is a function of $e$ and $\Theta$ (see the lower panel
3072: in Fig.~\ref{ap1}).
3073:
3074: Thus, unpolarized
3075: radiation produces only the third component of RATs, i.e. $Q_{e3}$ for the
3076: spheroid, while two first components $Q_{e1}, Q_{e2}$ vanish.
3077:
3078: \section{Mirror on a pole model}
3079: \subsection{RATs calculations}\label{mirror}
3080: Consider now a grain consisting of a square reflective mirror of side $l_{2}$,
3081: attached to the spheroid by a pole of the length $l_{1}$ (see
3082: Fig. \ref{f2}). Here we calculate RATs acting on the mirror. Its orientation in the grain coordinate system, ${\ma}_{1},
3083: {\ma}_{2}, {\ma}_{3}$, is characterized by a normal unit vector $\mn$, given by
3084: \bea
3085: \mn=n_{1}\ma_{1}+n_{2}\ma_{2},\label{b1}
3086: \ena
3087: where $n_{1}=\ms \alpha, n_{2}=\mc\alpha$ with $\alpha$ is the
3088: angle between $\mn$ and $\ma_{2}$. (see Fig. \ref{f2}). The use of the grain
3089: coordinate system is appropriate as the spheroidal body determines the grain
3090: inertia.
3091:
3092: Since $l_{1} \gg l_{2}$, the radius vector determining the position of
3093: each reflecting event on the mirror surface $\mr$ is nearly parallel to ${\bf
3094: \ma}_{3}$, thus
3095: \bea
3096: \mr =l_{1}\ma_{3}.
3097: \label{b2}
3098: \ena
3099: The orientation of the grain in the lab coordinate system is determined by
3100: $\Theta$ and $\beta$ as follows,
3101: \begin{align}
3102: \ma_{1}&=\mc \Theta \me_{1}+\ms \Theta \me_{2},\label{b3}\\
3103: \ma_{2}&=\mc\beta [-\ms \Theta \me_{1}+\mc \Theta \me_{2}]+\ms\beta \me_{3},\label{b4}\\
3104: \ma_{3}&=-\ms\beta [-\ms \Theta \me_{1}+\mc \Theta \me_{2}]+\mc\beta \me_{3}.\label{b5}
3105: \end{align}
3106: Plugging equations (\ref{b3}-\ref{b5}) into equations (\ref{b1}) and (\ref{b2}) we get
3107: \begin{align}
3108: \mn&=\me_{1}(n_{1}\mc\Theta-n_{2}\mc\beta\ms\Theta)\nonumber\\
3109: &+\me_{2}(n_{1}\ms\Theta+n_{2}\mc\beta\mc\Theta)\nonumber\\
3110: &+\me_{3}n_{2}\ms\beta,\label{b6}\\
3111: \mr&=l_{1}(\me_{1}\ms\beta\ms\Theta-\me_{2}\ms\beta\mc\Theta+\me_{3}\mc\beta).\label{b7}
3112: \end{align}
3113: Hence, RAT produced by the photon beam acting on the mirror is
3114: \bea
3115: d{\bf \Gamma}_{rad}=\gamma \mr \times \Delta {\bf P}=(-2p_{ph}) \gamma F dA ({\bf k}.\mn)[\mr\times \mn]. \label{b8}
3116: \ena
3117: Integrating over the full mirror, RAT becomes
3118: \bea
3119: {\bf \Gamma}_{rad}={-2p} \gamma F A_{\perp} ({\bf k}.\mn)[\mr
3120: \times \mn],\label{b9}
3121: \ena
3122: where $A_{\perp}=A|\me_{1}.\mn|$ is the cross section of the mirror with
3123: respect to the photon flux.
3124:
3125: Substituting $p_{ph}$, $F$, ${\bf k}\equiv\me_{1}$, and
3126: $A_{\perp}=l_{2}^{2}|\me_{1}.\mn|$ into equation (\ref{b9}), we get
3127: \begin{align}
3128: {\bf \Gamma}_{rad}&=\frac{\gamma u_{rad}\lambda l_{2}^{2}}{2} (-\frac{4}{\lambda})|\me_{1}.\mn|(\me_{1}.\mn)
3129: [(r_{2}N_{3}-r_{3}N_{2})\me_{1}\nonumber\\
3130: &+(r_{3}N_{1}-r_{1}N_{3})\me_{2}+(r_{1}N_{2}-r_{2}N_{1})\me_{3}].\label{b10}
3131: \end{align}
3132: From equations (\ref{b6}) and (\ref{b7}), we obtain
3133: \begin{align}
3134: \frac{r_{2}N_{3}-r_{3}N_{2}}{l_{1}}&=-n_{2}\ms\beta \mc\Theta \ms\beta\nonumber\\
3135: &-\mc\beta(n_{1}\ms\Theta+n_{2}\mc\beta \mc\Theta ) \nonumber\\
3136: &=(-n_{1} \mc \beta\ms\Theta -n_{2}\mc\Theta),\label{b11}\\
3137: \frac{r_{3}N_{1}-r_{1}N_{3}}{l_{1}}&=
3138: \mc\beta(n_{1}\mc\Theta-n_{2}\mc\beta\ms\Theta)\nonumber\\
3139: &-\ms\beta \ms\Theta n_{2}\ms\beta \nonumber\\
3140: &=n_{1}\mc\beta\mc\Theta-n_{2}\ms\Theta\label{b12},\\
3141: \frac{r_{1}N_{2}-r_{2}N_{1}}{l_{1}}&=n_{1}\ms\beta.
3142: \end{align}
3143: Therefore, we can write
3144: \begin{align}
3145: {\bf \Gamma}_{rad}&=\frac{\gamma u_{rad}\lambda l_{2}^{2}}{2}(Q_{e1}\me_{1}+Q_{e2}\me_{2}+Q_{e3}\me_{3}),\label{b13}
3146: \end{align}
3147: where the RAT components are given by
3148: \begin{align}
3149: Q_{e1}&=-\frac{4l_{1}}{\lambda}|\me_{1}.\mn|(n_{1}\mc\beta\ms\Theta+n_{2}\mc\Theta)[-n_{1}\mc\Theta\nonumber\\
3150: &+(n_{2}\mc\beta)\ms\Theta],\label{b14}\\
3151: Q_{e2}&=-\frac{4l_{1}}{\lambda}|\me_{1}.\mn|(n_{1}\mc\beta\mc\Theta-n_{2}\ms\Theta)[n_{1}\mc\Theta\nonumber\\
3152: &-n_{2}\mc\beta \ms\Theta],\label{b15}\\
3153: Q_{e3}&=-\frac{4l_{1}}{\lambda}|\me_{1}.\mn|n_{1}\ms\beta[n_{1}\mc\Theta-n_{2}\mc\beta\ms\Theta].\label{b16}
3154: \end{align}
3155: Simplifying the above equations, we get
3156: \begin{align}
3157: Q_{e1}&=\frac{4l_{1}}{\lambda}|n_{1} \mc\Theta-n_{2}
3158: \ms\Theta\mc\beta|[n_{1}n_{2}\mcs\Theta\nonumber\\
3159: &+\frac{n_{1}^{2}}{2}\mc\beta\ms 2\Theta-\frac{n_{2}^{2}}{2}\mc\beta\ms 2\Theta-n_{1}n_{2}\mss\Theta \mcs\beta],\label{b17}\\
3160: Q_{e2}&=-\frac{4l_{1}}{\lambda}|n_{1}
3161: \mc\Theta-n_{2}\ms\Theta\mc\beta|[n_{1}^{2}\mc\beta\mcs\Theta\nonumber\\
3162: &-\frac{n_{1}n_{2}}{2}\mcs\beta\ms 2\Theta-\frac{n_{1}n_{2}}{2}\ms 2\Theta+n_{2}^{2}\mc\beta\mss\Theta],\label{b18}\\
3163: Q_{e3}&=-\frac{4l_{1}}{\lambda}|n_{1} \mc\Theta-n_{2} \ms\Theta \mc\beta| n_{1}\ms\beta[n_{1}\mc\Theta\nonumber\\
3164: &-n_{2}\mc\beta\ms\Theta]. \label{b19}
3165: \end{align}
3166:
3167: Averaging over the rotational angle $\beta$ in the range $[0, 2\pi]$, we get
3168: \begin{align}
3169: Q_{e1}&=\frac{4\pi l_{1}n_{1}n_{2}}{\lambda} (3 \mbox{cos}^{2} \Theta - 1) f(\Theta, \alpha),\label{b20}\\
3170: Q_{e2}&= \frac{4\pi l_{1}n_{1}n_{2}}{\lambda} \ms 2\Theta g(\Theta, \alpha),\label{b21} \\
3171: Q_{e3}&=0,\label{b22}
3172: \end{align}
3173: where $f(\Theta, \alpha), g(\Theta, \alpha) $ are fitting functions depending
3174: on $\alpha$ and $\Theta$. The dependence on $\Theta$ characterizes the
3175: influence of variation of the mirror cross section on RATs. We will find these fitting functions in the following section.
3176:
3177:
3178: \subsection{Fitting functions}
3179: As we have seen above, RATs can be decomposed into analytical terms and fitting functions which are
3180: functions of both $\Theta$ and $\alpha$ (see equations \ref{b17} and
3181: \ref{b18}). Below we discuss an analytical approximation to
3182: the fitting functions $f(\Theta, \alpha), g(\Theta, \alpha)$.
3183:
3184: In the vicinity of $\Theta=0, \pi$, $|n_{1}\mc\Theta-n_{2}\ms\Theta\mc\beta| \sim n_{1}$, so
3185: this factor does not make $\beta$ -averaging more involved, however.
3186:
3187: As $\Theta \sim \pi/2$, equations (\ref{b14}) and (\ref{b15}) can be written
3188: as
3189: \begin{align}
3190: Q_{e1}&=\frac{4l_{1}}{\lambda}|-n_{2}\mc\beta|[n_{1}n_{2}\mcs\Theta+\frac{n_{1}^{2}}{2}\mc\beta\ms2\Theta\nonumber\\
3191: &-\frac{n_{2}^{2}}{2}\mc\beta\ms2\Theta-n_{1}n_{2}\mss\Theta\mcs\beta],\label{eq8a}\\
3192: Q_{e2}&=-\frac{4l_{1}}{\lambda}|-n_{2}\mc\beta|[n_{1}^{2}\mc\beta\mcs\Theta
3193: -\frac{n_{1}n_{2}}{2}\mcs\beta\ms2\Theta\nonumber\\
3194: &-\frac{n_{1}n_{2}}{2}\ms2\Theta+n_{2}^{2}\mc\beta\mss\Theta],\label{eq9a}
3195: \end{align}
3196: Integrating equations (\ref{eq8a}) and (\ref{eq9a}) over $\beta $ in a range
3197: $[0, 2\pi]$, we get
3198: \begin{align}
3199: Q_{e1}&=\frac{4l_{1}}{\lambda}\frac{4n_{1}n_{2}|n_{2}|}{3}(5\mcs\Theta-2),\label{eq8b}\\
3200: Q_{e2}&=\frac{4l_{1}}{\lambda}\frac{10 n_{1}n_{2}|n_{2}|}{3}\ms2\Theta.\label{eq9b}
3201: \end{align}
3202: Comparing equations (\ref{b17}) and (\ref{b18}) with (\ref{eq8b}) and
3203: (\ref{eq9b}), we have
3204: \begin{align}
3205: f_{\pi/2}(\Theta,\alpha)&=\frac{|\mbox{cos}\alpha| (5 \mcs\Theta-2)}{3\pi(3\mcs\Theta-1)},\label{eq16}\\
3206: g_{\pi/2}(\Theta,\alpha)&=\frac{10|\mbox{cos}\alpha|}{3\pi}.\label{eq17}
3207: \end{align}
3208:
3209: The fitting functions $f_{\pi/2}(\Theta, \alpha)$ and the tabulated function $f(\Theta, \alpha)$ for $\alpha=45^{0}$ are shown in Fig. (\ref{f4}).
3210: \begin{figure}
3211: \includegraphics[width=0.49\textwidth]{f3.ps}
3212: \caption{Tabulated and fitting functions $f$ for $\alpha = 45^{0}$. Shaded lines show the vicinity of the singular line $\mc\Theta=1/\sqrt{3}$}
3213: \label{f4}
3214: \end{figure}
3215:
3216: \begin{figure}
3217: \includegraphics[width=0.49\textwidth]{f4.ps}
3218: \caption{Tabulated and fitting functions $g$ for $\alpha =45^{0}$.}
3219: \label{f4*}
3220: \end{figure}
3221:
3222: \begin{figure}
3223: \includegraphics[width=0.49\textwidth]{f37a.eps}
3224: \includegraphics[width=0.49\textwidth]{f37b.eps}
3225: \caption{Fitting function $f$ for two tilted angles of the mirror
3226: in the body system: upper and lower panel correspond to $\alpha =30^{0}, 60^{0}$. Shaded lines show the vicinity of the singular line $|\mc\Theta|=1/\sqrt{3}$}
3227: \label{fap3}
3228: \end{figure}
3229:
3230: \begin{figure}
3231: \includegraphics[width=0.49\textwidth]{f38a.eps}
3232: \includegraphics[width=0.49\textwidth]{f38b.eps}
3233: \caption{Fitting and tabulated functions $g$ for two tilted angles of the mirror
3234: in the body system: upper and lower panel correspond to $\alpha =30^{0}, 60^{0}$.}
3235: \label{fap4}
3236: \end{figure}
3237:
3238:
3239: Interestingly enough, the approximation of the fitting
3240: functions obtained in the vicinity of $\Theta \sim \pi/2$ presents
3241: a reasonable approximation for the fitting function $f$ over the
3242: entire range of $\Theta$.
3243: Indeed, Fig.~\ref{f4} shows the tabulated function $f(\Theta, \alpha)$
3244: obtained by substituting $Q_{e1}, Q_{e2}$ resulted from numerically averaging
3245: equations (\ref{b14}) and (\ref{b15}) into equations (\ref{b17}) and (\ref{b18})(solid lines) and the fitting function $ f_{\pi/2}(\Theta,\alpha)$ given by equation (\ref{eq16}) (dashed line) for a particular case of $\alpha=45^{0}$. It is shown that the fitting function $f_{\pi/2}(\Theta,\alpha)$ has a good agreement with $f(\Theta, \alpha)$. Similar results are also found for other $\alpha$ angles (see Fig.~\ref{fap3}).
3246:
3247: However, the fitting function $g_{\pi/2}$ given by equation~(\ref{eq17}) is independent
3248: of $\Theta$ and therefore does not provide a
3249: good fit with the tabulated function $g$. In this case, we perform a quadratic
3250: fitting for tabulated functions for different $\alpha$ angles. The results for
3251: $\alpha=30^{0}$, $45^{0}$ and $60^{0}$ are given by
3252: \begin{align}
3253: g(\Theta, \alpha=30^{0})&=1.4675-0.5\mcs\Theta,\label{eq17a}\\
3254: g(\Theta, \alpha=45^{0})&=1.191+0.1382\mcs\Theta,\label{eq17b}\\
3255: g(\Theta, \alpha=60^{0})&=0.864+0.869\mcs\Theta.\label{eq17c}
3256: \end{align}
3257: Tabulated and fitting functions g are shown in Fig.\ref{f4*} for $\alpha=45^{0}$ (see Fig.~\ref{fap4} for $g$ corresponding to other angles $\alpha$).
3258:
3259: Figs (\ref{fap3}) and (\ref{fap4}) show $f, g$ for $\alpha=30^{0}$ and $60^{0}$.
3260: Fig. \ref{fap3} shows that there is a change in sign of the tabulated function $f$ as $|\mc\Theta| \rightarrow 1/\sqrt{3}$ between $\alpha=30^{0}$ and $60^{0}$. This stems from the fact that, for $\alpha <45^{0}$, the influence of $A_{\perp}$ shifts the zero of $Q_{e1}$ from $|\mc\Theta|=1/\sqrt{3}$ to $|\mc\Theta|<1/\sqrt{3}$. As a result, when the term $3 \mbox{cos}^{2} \Theta - 1$ approaches the zero from the left of $\mc\Theta=-1/\sqrt{3}$, for instance, then $Q_{e1}$ still positive, which causes $f$ rises sharply to the positive direction as in Fig. \ref{fap3}{\it upper}.
3261:
3262: As an example, in the paper we mostly discuss the case in which the mirror is tilted by an angle
3263: of $\alpha=\pi/4$ for which there is the best fit between the fitting function
3264: $f_{\pi/2}$ and the tabulated function $f$ except a narrow range in the
3265: vicinity of $|\mc \Theta|=1/\sqrt{3}$ (see Fig. \ref{f4}). Similarly,
3266: $g$ given by equation (\ref{eq17b}) also provides a good fit for $Q_{e2}$
3267: (see Fig. \ref{f4*}).
3268: Therefore, RATs acting upon the mirror can be roughly approximated by equations (\ref{b17})-(\ref{b18}) with
3269: $f, g$ are given in equations (\ref{eq16}) and (\ref{eq17b}).
3270:
3271: Needless to say, that such an approximation entails errors in calculating
3272: the torques $Q_{e1}$ and $Q_{e2}$. Fig.~\ref{f6*} shows the torques with $f,
3273: g$ being the tabulated and fitting functions. It can be seen that there is a very good fit for the component $Q_{e2}$, but there exhibit some deviation for the component $Q_{e1}$ toward the range $\mc\Theta \sim \pm 1$. These errors could be
3274: decreased by using more sophisticated analytical fits, e.g. piecewise
3275: analytical fit. However, this would decrease the heuristic value of the
3276: formulae. Therefore while using the approximate fits (\ref{eq16}) and
3277: (\ref{eq17b})
3278: for some analytical calculations below, for most of the quantitative estimates,
3279: including those related to evolving the phase trajectories, we tabulate
3280: the fitting functions $f$ and $g$. Naturally, the latter
3281: is equivalent to the direct use of averaged
3282: equations (\ref{b14})-(\ref{b15}) for RAT components.
3283: \begin{figure}
3284: \includegraphics[width=0.49\textwidth]{f5.ps}
3285: \caption{Fig shows the comparison of RAT components. Dashed lines show RATs obtained from the analytical approximation given by equations
3286: (\ref{b17})-(\ref{b18}) and solid
3287: lines show RATs obtained by numerically averaging equations (\ref{b14})-(\ref{b15}) over $\beta$. }
3288: \label{f6*}
3289: \end{figure}
3290:
3291:
3292: What is the purpose of having the analytical form of the fitting functions, if
3293: they do not provide exact fits for the torque components (see Fig. \ref{f6*})? We
3294: may see the actual RATs acting on irregular grains (e.g. Fig. \ref{f61}) also differ
3295: somewhat from the corresponding AMO RATs. These differences, however, do not
3296: change substantially, as we show in the rest of the paper, the alignment of
3297: irregular grains in comparison with the alignment predicted by AMO. Thus we
3298: believe that the slightly distorted AMO with the appropriate analytically
3299: fitting functions should also reflect correctly the generic properties of the
3300: RAT alignment.
3301: \subsection{Dependences on $\alpha$}
3302: \begin{figure}
3303: \includegraphics[width=0.49\textwidth]{f7.eps}
3304: \caption{Variation of $Q_{e1}^{max}/Q_{e2}^{max}$ as a function of $\alpha$.
3305: This ratio determines the existence of high attractor points and the shift
3306: of a crossover point as we describe in \S \ref{qratio}. Shaded areas show the range of $\alpha$ in which the symmetry of RATs are significantly affected by the cross section.}
3307: \label{f5}
3308: \end{figure}
3309: Fig. \ref{f5} shows that the variation of the mirror orientation can change the ratio of their maximum $Q_{e1}^{max}/Q_{e2}^{max}$. Here we denote $Q_{e1}^{max}$ the maximum of
3310: $|Q_{e1}|$ which is exactly its amplitude at $\Theta=0$ or $\pi$, and
3311: $Q_{e2}^{max}$ the maximum of $Q_{e2}$. In Fig. \ref{f5} it is obviously shown that
3312: $Q_{e1}^{max}/Q_{e2}^{max}$ increases with $\alpha$ increasing due to the increase of the cross-section $A_{\perp}$.
3313: However, the ratio $Q_{e1}^{max}/Q_{e2}$ is limited to the upper limit of $1.3$. We will see later
3314: that the ratio plays an important role on the alignment that is discussed in \S~ \ref{sec5}.
3315:
3316: Moreover, the variation of $\alpha$ does not only give rise to the change of the ratio $Q_{e1}^{max}/Q_{e2}^{max}$, but it can also affect the symmetric functional form of RATs. Our calculations show that for $\alpha$ smaller than $20^{0}$ or larger than $80^{0}$, the RATs have some changes in the
3317: functional form of the $Q_{e1}$ and $Q_{e2}$ compared to what is shown in Fig. \ref{f6}.
3318:
3319: Let us explain why for $\alpha$ small, the
3320: symmetry of RATs for AMO is affected. Consider equation (\ref{b17}) with $n_{1}=\ms\alpha, n_{2}=\mc\alpha$
3321: \begin{align}
3322: Q_{e1}&=\frac{4l_{1}}{\lambda}|\ms\alpha \mc\Theta-\mc\alpha
3323: \ms\Theta\mc\beta|[\ms\alpha \mc\alpha\mcs\Theta\nonumber\\
3324: &+\frac{\mss\alpha}{2}\mc\beta\ms 2\Theta-\frac{\mcs\alpha}{2}\mc\beta\ms 2\Theta\nonumber\\
3325: &-\ms\alpha\mc\alpha\mss\Theta \mcs\beta].\label{apc1}
3326: \end{align}
3327: The first term is the cross-section $A_\perp$, and the second term is $Q_{e1}$ without
3328: cross-section.
3329: The second term when averaged gives rise to $Q_{e1} \sim 5\mcs\Theta-2$. It indicates that if the cross-section is constant, $Q_{e1}$ is fully symmetric. Consider $\alpha$ small, $cos\alpha$ is significant, and equation (\ref{apc1}) reduces to
3330: \begin{align}
3331: Qe1&=\frac{4l_{1}}{\lambda}|-\mc\alpha \ms\Theta \mc\beta|[\ms\alpha \mc\alpha\mcs\Theta\nonumber\\
3332: &+\frac{\mss\alpha}{2}\mc\beta\ms 2\Theta-\frac{\mcs\alpha}{2}\mc\beta\ms 2\Theta\nonumber\\
3333: &-\ms\alpha\mc\alpha\mss\Theta \mcs\beta].\label{apc2}
3334: \end{align}
3335: Obviously, when averaging over $\beta$, the absolute term containing $\ms\Theta$ contributes to modify substantially the symmetry of the resulting torques. The same problem also occurs for $Q_{e2}$. We found that for both $\alpha \le 20^{0}$ or $\alpha \ge 80^{0}$, the functional forms of RATs are very influenced. However, for $\alpha$ within $[20^{0},80^{0}]$, their functional forms are not much different with what shown in Fig. \ref{f6} for $\alpha=45^{0}$.
3336:
3337:
3338:
3339:
3340: \section{RATs with DDSCAT}
3341: In the discrete electric dipole approximation (Draine \& Flateau 1994), a grain is presented as an
3342: ensemble of electric dipoles. The interaction between the electric field of
3343: incident light and the
3344: dipoles produces radiative forces and torques.
3345: RATs produced by radiation on a grain consisting of $N$ electric
3346: dipoles are
3347: \bea
3348: {\bf \Gamma}_{rad}=\sum_{j=1}^{N}{\bf r}_{j}\times {\bf F}_{j}+\sum_{j=1}^{N} {\bf p}_{j}\times {\bf E}_{j},\label{rad0}
3349: \ena
3350: where ${\bf r}{j}$ is radius of $j^{th}$ dipole, $ {\bf E}_{j}, {\bf F}_{j}$
3351: are electric field at the location of, and radiative force
3352: which acts on the $j^{th}$ dipole. Radiative force ${\bf F}$ present in equation (\ref{rad0}) is produced by
3353: the gradient of electric field in the grain and the Lorentz force due to vibration of electric
3354: dipole in magnetic field.
3355:
3356: Radiation field inside the
3357: grain consists of that of incident and scattered light. Each
3358: dipole receives the incident light which induces its vibration and scattered
3359: light produced by all electric dipoles except the dipole under study. Hence,
3360: the total RAT can be written in a different form
3361: \bea
3362: {\bf \Gamma}_{rad}={\bf \Gamma}_{inc}+{\bf \Gamma}_{sca}.\label{rad1}
3363: \ena
3364: RAT efficiency ${\bf Q}_{\Gamma}$ is defined as followings,
3365: \bea
3366: {\bf \Gamma}_{rad}=\frac{u_{\lambda}a_{eff}^{2}\lambda}{2}\gamma{\bf Q}_{\Gamma},\label{rad11}
3367: \ena
3368: where $\gamma$ is the anisotropy, and $u_{\lambda}$ is the energy density of radiation field of wavelength
3369: $\lambda$.
3370: Hence, equation (\ref{rad1}) can be rewritten
3371: \bea
3372: {\bf Q}_{\Gamma}={\bf Q}_{inc}+{\bf Q}_{sca}.\label{rad2}
3373: \ena
3374: Here ${\bf Q}_{inc} \equiv {\bf Q}_{abs} $ and ${\bf Q}_{sca}$ are given by (DW96)
3375: \bea
3376: {\bf Q}_{inc}=\frac{4k}{a_{eff}^{2}|E_{inc,0}|^{2}} {\bf Re}\sum_{j=1}^{N}{\bf p}_{j}(0)\times {\bf E}_{inc,0}e^{i{\bf k}.{\bf r}_{j}} \nonumber \\
3377: -i{\bf k}\times \sum_{j=1}^{N}{\bf r}_{j}[{\bf p}_{j}.{\bf E}_{inc,0}]e^{i{\bf k}.{\bf r}_{j}},\label{rad3}\\
3378: {\bf Q}_{sca}=\frac{-k^{5}}{\pi a_{eff}^{2}|{\bf E}_{inc,0}|^{2}}\int d\Omega {\bf Re}(S_{E}^{*}V_{B}+S_{B}^{*}V_{E}).\label{rad4}
3379: \ena
3380: In above equations, ${\bf E}_{inc}$ denotes electric field of incident light,
3381: $S_{E}, S_{B}$, $V_{E}, V_{B}$ are given by
3382: \begin{align}
3383: S_{E}&=\sum_{j=1}^{N}[\mr_{j}-(\mm.\mr_{j})\mm-\frac{2i}{k}\mm].\mp_{j}(0)exp(-ik\mm.\mr_{j}),\\
3384: S_{B}&=\mr.\sum_{j=1}^{N}\mp_{j}(0)\times \mr_{j}exp(-ik\mm.\mr_{j}),\\
3385: V_{E}&=\sum_{j=1}^{N}{\mp_{j}(0)-\mm[\mm.\mp_{j}(0)}exp(-ik\mm.\mr_{j})\\
3386: V_{B}&=-\mn\times V_{E},
3387: \end{align}
3388: where $\mp_{j}$ is $j^{th}$ electric dipole moment, k is wave number, $\mr$ is the radius vector and $\mm$ is normal unit vector.
3389:
3390: We use DDSCAT code to compute separately the components arising from
3391: absorption, scattering
3392: and total RAT for a grain in which its direction with respect to ${\bf k}$ is
3393: determined by the angles $\Theta,\beta,\Phi$ (see Fig. \ref{f1}). Here $\Theta$ is the angle between ${\bf a}_{1}$
3394: and ${\bf k}$; $\beta$ is the rotational angle of the grain around ${\bf a}_{1}$; and
3395: $\Phi$ is the precession angle of ${\bf a}_{1}$ about ${\bf k}$.
3396: For our study (see Table \ref{tab2}), we compute RATs for the
3397: spectrum of the interstellar radiation field, over 21 directions of $\Theta$ from $0$ to $\pi$ and 21 values of $\beta$ from
3398: $0$ to $2\pi$, at $\Phi=0$.
3399:
3400: RAT can be decomposed into components in the scattering system via
3401: \bea
3402: {\bf Q}_{\Gamma}=Q_{e1}\hat{e}_{1}+Q_{e2}\hat{e}_{2}+Q_{e3}\hat{e}_{3}, \label{q4}
3403: \ena
3404: where $\hat{e}_{1}$, $\hat{e}_{2}$, $\hat{e}_{3}$ are shown in Fig.
3405: \ref{f1}.
3406:
3407: Mean radiative torque over wavelengths, $\overline{\bQ(\Theta,\beta,\Phi)}$ is defined by
3408: \bea
3409: \overline{\bQ}=\frac{\int \bQ_{\lambda}u_{\lambda}d\lambda}{\int
3410: u_{\lambda}d\lambda} \label{ra1}.
3411: \ena
3412:
3413: Since $\beta$ varies very fast due to the swift rotation of the
3414: grain around the axis
3415: of major inertia ${\bf a}_{1}$, we can average RATs over $\beta$ from $0$ to
3416: $2\pi$.
3417:
3418: \section{Attractor and Repellor points}
3419: Here we derive the condition for which a stationary point becomes an attractor
3420: and a repellor point following the approach in DW97.
3421: Assuming that the stationary point has the angle $\xi_{s}$ with respect to
3422: magnetic field and angular momentum $J_{s}$, one can expand the right hand sides
3423: of equations of motion around this point. As a result, equations (\ref{eq59})
3424: and (\ref{eq60}) give
3425: \bea
3426: \frac{d\xi}{dt}=\frac{\langle
3427: F(\xi_{s})\rangle_{\phi}}{J_{s}}+\frac{d\langle
3428: F\rangle_{\phi}}{J d\xi}(\xi_{s},J_{s})-\frac{\langle F(\xi_{s})\rangle_{\phi}}{J^{2}_{s}}(J-J_{s}),\label{ap16}\\
3429: \frac{dJ}{dt}=\langle
3430: H(\xi_{s})\rangle_{\phi}-J_{s}+\frac{d\langle H\rangle_{\phi}}{d\xi}(\xi-\xi_{s})-(J-J_{s})
3431: ,\label{ap17}
3432: \ena
3433: where $\langle H(\xi)\rangle_{\phi},\langle F(\xi)\rangle_{\phi}$ are spinning
3434: and aligning torques already averaged over the precession angle $\phi$.
3435: Since for stationary points $\xi_{s}, J_{s}$, we have $\langle
3436: F(\xi_{s})\rangle_{\phi}=0, \langle H(\xi_{s})\rangle_{\phi}-J_{s}=0$, equations
3437: (\ref{ap16}) and (\ref{ap17}) become
3438:
3439: \bea
3440: \frac{d\xi}{dt}=A(\xi-\xi_{s})+B(J-J_{s}),\label{ap18}\\
3441: \frac{d\omega}{dt}=C(\xi-\xi_{s})+D(J-J_{s}),\label{ap19}
3442: \ena
3443: where
3444: \bea
3445: A=\frac{d\langle F\rangle_{\phi}}{Jd\xi}(\xi_{s}, J_{s}),\\
3446: B=-\frac{\langle F(\xi_{s})\rangle_{\phi}}{J^{2}_{s}},\\
3447: C=\frac{d\langle H\rangle_{\phi}}{d\xi}(\xi_{s}, J_{s}),\\
3448: D=-1.
3449: \ena
3450: To have an attractor point, one requires
3451: \bea
3452: A+D<0,\label{ap20}\\
3453: BC-AD<0.\label{ap21}
3454: \ena
3455: In other words,
3456: \bea
3457: \frac{d\langle F\rangle_{\phi}}{\langle H\rangle_{\phi}d\xi}(\xi_{s},J_{s})<1,\label{ap22}\\
3458: H\frac{d\langle F\rangle_{\phi}}{d\xi}(\xi_{s},J_{s})<\langle
3459: F\rangle_{\phi}\frac{d\langle H\rangle_{\phi}}{d\xi}(\xi_{s},J_{s}),\label{ap23}
3460: \ena
3461: where we have substituted $J_{s}=H(\xi_{s})$.
3462:
3463: \section{Effective grain size for AMO}
3464:
3465: For phase maps, we numerically average equations (\ref{eq8})-(\ref{eq10}) to obtain
3466: exact RATs for AMO, rather than using approximate formulae as in the analysis. However, for this case, the absolute magnitude of torques
3467: matters. Therefore, we normalize AMO
3468: in the following way. Assuming that the size $l_{2}$ of the mirror and $l_{1}$ are chosen so that the RAT for AMO has the magnitude equal to that of irregular grain of an effective size $a_{eff}$. Thus, following equations (\ref{eq1}) and (\ref{eq7b}) we have
3469: \begin{align}
3470: \frac{\lambda u_{rad}l_{2}^{2}}{2}Q_{AMO}&=\frac{\lambda u_{rad}a_{eff}^{2}}{2}Q_{DDSCAT},\label{eq66*}
3471: \end{align}
3472: where $Q_{AMO}$ and $Q_{DDSCAT}$ are the magnitudes of RATs for AMO and an irregular grain, respectively.
3473: We can simplify further by normalizing RAT components over the maximum of $Q_{e1}$, and let $Q_{AMO}, Q_{DDSCAT}$ equal the maximum of $Q_{e1}$ for AMO and irregular grain. As a result, for AMO, we have $Q_{AMO}=|Q_{e1}|^{max}=\frac{16\pi n_{1}n_{2}^{2}l_{1}}{\lambda}$, and substituting into equation (\ref{eq66*}), we get
3474: \begin{align}
3475: l_{2}^{2}l_{1}&=\frac{\lambda}{16\pi n_{1}n_{2}^{2}}a_{eff}^{2} Q_{DDSCAT}.\label{eq67*}
3476: \end{align}
3477: We can define the effective size of AMO subject to the mirror size and the rod length as
3478: \begin{align}
3479: a_{AMO}^{2}&=l_{2}^{2}(\frac{l_{1}}{\lambda}).\label{eq67_1}
3480: \end{align}
3481: Equations (\ref{eq67*}) and (\ref{eq67_1}) enables us to find the effective size of AMO that produce the same the RAT magnitude with the irregular grain of size $a_{eff}$, given by
3482: \begin{align}
3483: a_{AMO}^{2}&=\frac{1}{16\pi n_{1}n_{2}^{2}}a_{eff}^{2} Q_{DDSCAT}.\label{eq68*}
3484: \end{align}
3485:
3486:
3487: As we discuss in \S \ref{self} that the magnitude of RATs for irregular grains can be crudely approximated as
3488: \begin{align}
3489: Q_{DDSCAT}&=0.4(\frac{\lambda}{a_{eff}})^{\eta},\label{eq69*}
3490: \end{align}
3491: where $\eta=0$ for $\lambda< 1.8~a_{eff}$ and $\eta=3$ or $4$ for $\lambda>1.8~ a_{eff}$ (see \S \ref{self}).
3492:
3493: Equations (\ref{eq69*}) and (\ref{eq68*}) allow us to roughly estimate the the effective size of AMO, $a_{AMO}$ as a function of the effective size of irregular grain, $a_{eff}$, provided that the wavelength of radiation field is known.
3494:
3495: \begin{thebibliography}{8.}
3496: \bibitem{} Abbas M.M., Craven P. D., Spann J. F. et al. 2004, ApJ, 614, 781
3497: \bibitem{} Bastien P., Jenness T., Molnar J. 2005, ASPC, 343, 69
3498: \bibitem{} Bethell T., Cherpunov A., Lazarian A., Kim J. 2006, ApJ, in press
3499: \bibitem{} Bohren, C Craig 1974, Chemical Physics Letters, 29, 458
3500: \bibitem{} Cho J., Lazarian A. 2005, ApJ, 631, 361
3501: \bibitem{} Cho J., Lazarian A. 2007, ApJ, submitted
3502:
3503: \bibitem{} Davis L., Greenstein J.L. 1951, ApJ, 114, 206
3504: \bibitem{} Dolginov A.Z. 1972 Ap\&SS, 16, 337
3505: \bibitem{} Dolginov A.Z. Mytrophanov I.G. 1976, Ap\&SS, 43, 291
3506: \bibitem{} Dolginov A.Z. Silantev N.A. 1976, Ap\&SS, 43, 337
3507: \bibitem{} Draine B., Flatau P. 1994, J. Opt. Soc. Am. A., 11, 1491
3508: \bibitem{} Draine B., Lee H. 1984, ApJ, 285, 89
3509: \bibitem{} Draine B. 1985, ApJS, 57, 587
3510: \bibitem{} Draine B., Lazarian, A. 1998, ApJ, 508, 157
3511: \bibitem{} Draine B., Weingartner J. 1996, ApJ, 470, 551 (DW96)
3512: \bibitem{} Draine B., Weingartner J. 1997, ApJ, 480, 633 (DW97)
3513: \bibitem{} Gold T. 1951, Nature, 169, 322
3514:
3515: \bibitem{} Goodman A., Jones T., Lada E., Myers P. 1995, ApJ, 448, 748
3516:
3517: \bibitem{} Hall J. 1949, Science, 109, 166
3518:
3519:
3520: \bibitem{} Harwit M. 1970, Nature, 226, 61-63
3521:
3522: \bibitem{} Hildebrand R., Davidson J. A., Dotson J.L, Wovell C.D., \\
3523: Novak G., Vaillancourt J.E. 2000, PASP, 112, 1215
3524: \bibitem{} Hildebrand R. 2002, in {\it Astrophysical Spectropolarimetry},\\
3525: ed. by J. Trujillo-Bueno,
3526: F. Moreno-Insertis, \\\& F. Sanchez (Cambridge, UK: Cambridge\\
3527: Univ. Press), p. 265
3528: \bibitem{} Hiltner W. 1949, Science, 109, 165
3529: \bibitem{} Hoang T., Lazarian A., 2007, MNRAS, submitted
3530: \bibitem{} Hoang T., Lazarian A., in preparation
3531:
3532: \bibitem{} Jones M., Spitzer L., 1967, ApJ, 147, 943
3533: \bibitem{} Lazarian A. 1994, MNRAS, 268, 713
3534: \bibitem{} Lazarian A. 1995, ApJ, 453, 229
3535: \bibitem{} Lazarian A. 1997a, MNRAS, 288, 609
3536: \bibitem{} Lazarian A. 1997b, ApJ, 483, 296
3537: \bibitem{} Lazarian A. 2003, Journal of \\
3538: Quantitative Spectroscopy and Radiative Transfer, 79, 881
3539: \bibitem{} Lazarian A. 2007, Journal of \\
3540: Quantitative Spectroscopy and Radiative Transfer, 106,225
3541: \bibitem{} Lazarian A., Goodman A.A., Myers P.C. 1997, ApJ, 490, 273
3542: \bibitem{} Lazarian A., Roberge W. 1997, ApJ, 484, 230
3543: \bibitem{} Lazarian A., Draine B. 1999a, ApJ, 516, L37 (LD99a)
3544: \bibitem{} Lazarian A., Draine B. 1999b, ApJ, 520, L67 (LD99b)
3545: \bibitem{} Lebedev P. 1901 Ann. der Physik, 6, 433
3546: \bibitem{} Lee H., Draine B. 1985, ApJ, 290, 211
3547: %\bibitem{} Mathis J., Whitney B., Wood K. 2002, ApJ, 574, 812
3548: \bibitem{} Mathis J., 1986, ApJ, 308, 281
3549: \bibitem{} Mathis J., Mezger P., Panagia N. 1983, A\&A, 128, 212
3550: %\bibitem{} Mathis J., Rumpl W., Nordsieck, K.
3551: 1977, ApJ, 217, 425 (MRN)
3552: \bibitem{} Novak G., et al. Millimeter and \\
3553: Submillimeter Detectors for Astronomy II. In Antebi J, D. Lemke, editors.\\
3554: Proceedings of the SPIE, vol. 5498; 2004 p. 278
3555:
3556: \bibitem{} Pelkonen V., Juvela M., Padoan P. 2007, A\&A, 461, 551
3557: \bibitem{} Purcell E. 1979, ApJ, 231, 404
3558: \bibitem{} Purcell E., Spitzer L. 1971, ApJ, 167, 31
3559: \bibitem{} Roberge W., Hanany S. 1990, B.A.A.S., 22, 862
3560: \bibitem{} Roberge W., Hanany S., Messinger D. 1995, 453, 238
3561: \bibitem{} Roberge W.G., Lazarian A. 1999, MNRAS, 305, 615
3562: \bibitem{} Rosenbush V., Kolokolova L.,Lazarian A., Shakhovskoy N., \&
3563: Kiselev N. 2007, Icarus, 186, 317
3564: \bibitem{} Serkowski K, Mathewson DS, Ford VL 1975, ApJ, 196,261 ≈
3565: \bibitem{} Spitzer L., McGlynn TA. 1979, ApJ, 231, 417
3566: \bibitem{} Vishniac E., Lazarian A., \& Cho, J. 2003, in Turbulence and
3567: Magnetic Fields in Astrophysics, Eds. E. Falgarone, and T. Passot, LNP, 614,
3568: 376
3569: \bibitem{} Weingartner J., Draine B. 2001, ApJ, 548, 296
3570: \bibitem{} Weingartner J., Draine B. 2003, ApJ, 589, 289 (WD03)
3571: \bibitem{} Ward-Thompson D., Kirk J.M., Crutcher R.M., Greaves J.S., Holland W.S., Andre P. 2000, ApJ, 537, L135
3572: \end{thebibliography}
3573:
3574: \newpage
3575: \clearpage
3576:
3577:
3578: \end{document}
3579: