0807.1373/ms.tex
1: \documentclass[12pt,preprint]{aastex}%
2: \usepackage{amsfonts}
3: \usepackage{amssymb}
4: \usepackage{graphicx}
5: %TCIDATA{OutputFilter=latex2.dll}
6: %TCIDATA{Version=5.00.0.2570}
7: %TCIDATA{CSTFile=aastex.cst}
8: %TCIDATA{LastRevised=Sunday, June 15, 2008 15:32:46}
9: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
10: %TCIDATA{<META NAME="SaveForMode" CONTENT="1">}
11: %TCIDATA{Language=American English}
12: \slugcomment{Preprint}
13: \shorttitle{Gravity-powered Jet Dynamo}
14: \shortauthors{Bellan}
15: \newtheorem{theorem}{Theorem}
16: \newtheorem{acknowledgement}[theorem]{Acknowledgement}
17: \newtheorem{algorithm}[theorem]{Algorithm}
18: \newtheorem{axiom}[theorem]{Axiom}
19: \newtheorem{claim}[theorem]{Claim}
20: \newtheorem{conclusion}[theorem]{Conclusion}
21: \newtheorem{condition}[theorem]{Condition}
22: \newtheorem{conjecture}[theorem]{Conjecture}
23: \newtheorem{corollary}[theorem]{Corollary}
24: \newtheorem{criterion}[theorem]{Criterion}
25: \newtheorem{definition}[theorem]{Definition}
26: \newtheorem{example}[theorem]{Example}
27: \newtheorem{exercise}[theorem]{Exercise}
28: \newtheorem{lemma}[theorem]{Lemma}
29: \newtheorem{notation}[theorem]{Notation}
30: \newtheorem{problem}[theorem]{Problem}
31: \newtheorem{proposition}[theorem]{Proposition}
32: \newtheorem{remark}[theorem]{Remark}
33: \newtheorem{solution}[theorem]{Solution}
34: \newtheorem{summary}[theorem]{Summary}
35: \newenvironment{proof}[1][Proof]{\noindent\textbf{#1.} }{\ \rule{0.5em}{0.5em}}
36: \begin{document}
37: %
38: %TCIMACRO{\TeXButton{Defining underset}{\makeatletter\def\stackunder
39: %#1#2{\mathrel{\mathop{#2}\limits_{#1}}}
40: %\makeatother}}%
41: %BeginExpansion
42: \makeatletter\def\stackunder#1#2{\mathrel{\mathop{#2}\limits_{#1}}}
43: \makeatother
44: %EndExpansion
45: %
46: 
47: %TCIMACRO{\TeXButton{Title}{\title{Dust-driven Dynamos in Accretion Disks}}}%
48: %BeginExpansion
49: \title{Dust-driven Dynamos in Accretion Disks}%
50: %EndExpansion
51: %
52: 
53: %TCIMACRO{\TeXButton{Author}{\author{P. M. Bellan}}}%
54: %BeginExpansion
55: \author{P. M. Bellan}%
56: %EndExpansion
57: %
58: 
59: %TCIMACRO{\TeXButton{Affiliation}{\affil
60: %{Applied Physics, Caltech, Pasadena CA 91125, USA}}}%
61: %BeginExpansion
62: \affil{Applied Physics, Caltech, Pasadena CA 91125, USA}%
63: %EndExpansion
64: %
65: 
66: %TCIMACRO{\TeXButton{Email}{\email{pbellan@caltech.edu}}}%
67: %BeginExpansion
68: \email{pbellan@caltech.edu}%
69: %EndExpansion
70: %
71: 
72: %TCIMACRO{\TeXButton{Begin abstract}{\begin{abstract}}}%
73: %BeginExpansion
74: \begin{abstract}%
75: %EndExpansion
76: Magnetically driven astrophysical jets are related to accretion and involve
77: toroidal magnetic field pressure inflating poloidal magnetic field flux
78: surfaces. Examination of particle motion in combined gravitational and
79: magnetic fields shows that these astrophysical jet toroidal and poloidal
80: magnetic fields can be powered by the gravitational energy liberated by
81: accreting dust grains that have become positively charged \ by emitting
82: photo-electrons. Because a dust grain experiences magnetic forces after
83: becoming charged, but not before, charging can cause irreversible trapping of
84: the grain so dust accretion is a consequence of charging. Furthermore,
85: charging causes canonical angular momentum to replace mechanical angular
86: momentum as the relevant constant of the motion. The resulting effective
87: potential has three distinct classes of accreting particles distinguished by
88: canonical angular momentum, namely (i) \textquotedblleft
89: cyclotron-orbit\textquotedblright, (ii) \textquotedblleft
90: Speiser-orbit\textquotedblright, and (iii) \textquotedblleft zero canonical
91: angular momentum\textquotedblright\ particles. Electrons and ions are of class
92: (i) but depending on mass and initial orbit inclination, dust grains can be of
93: any class. Light-weight dust grains develop class (i) orbits such that the
94: grains are confined to nested poloidal flux surfaces, whereas grains with a
95: critical weight such that they experience comparable gravitational and
96: magnetic forces can develop class (ii) or class (iii)\ orbits, respectively
97: producing poloidal and toroidal field dynamos.
98: %TCIMACRO{\TeXButton{End abstract}{\end{abstract}} }%
99: %BeginExpansion
100: \end{abstract}
101: %EndExpansion
102: \ \ \ \ \ \ %
103: 
104: %TCIMACRO{\TeXButton{Keywords}{\keywords
105: %{Accretion, jet, MHD, canonical angular momentum, dynamo, kinetic theory, Störmer potential, collisions, Hamiltonian dynamics, orbit, Speiser orbit, dusty plasma, photo-emission }%
106: %} }%
107: %BeginExpansion
108: \keywords
109: {Accretion, jet, MHD, canonical angular momentum, dynamo, kinetic theory, Störmer potential, collisions, Hamiltonian dynamics, orbit, Speiser orbit, dusty plasma, photo-emission }
110: %EndExpansion
111: \ 
112: 
113: \pagebreak
114: 
115: \section{Introduction}
116: 
117: Magnetohydrodynamically driven plasma jets having topology and dynamics
118: analogous to astrophysical jets have been produced in laboratory experiments
119: by
120: %TCIMACRO{\TeXButton{\citet{Hsu2002,Bellan2005}}{\citet{Hsu2002,Bellan2005}} }%
121: %BeginExpansion
122: \citet{Hsu2002,Bellan2005}
123: %EndExpansion
124: and by
125: %TCIMACRO{\TeXButton{\citet{Lebedev2005}}{\citet{Lebedev2005}}}%
126: %BeginExpansion
127: \citet{Lebedev2005}%
128: %EndExpansion
129: ; see discussion by
130: %TCIMACRO{\TeXButton{\citet{Blackman2007}}{\citet{Blackman2007}}}%
131: %BeginExpansion
132: \citet{Blackman2007}%
133: %EndExpansion
134: . The feature of azimuthal symmetry, common to both the lab experiments and to
135: real astrophysical jets, has important implications for the structure of the
136: magnetic field. This is because an azimuthally symmetric magnetic field can be
137: expressed using a cylindrical coordinate system $\{r,\phi,z\}$ as$\ $
138: \begin{equation}
139: \mathbf{B=}\frac{1}{2\pi}\left(  \nabla\psi\times\nabla\phi+\mu_{0}I\nabla
140: \phi\right)  \ \label{B general2}%
141: \end{equation}
142: where the poloidal flux $\psi(r,z,t)$ is defined by
143: \begin{equation}
144: \psi(r,z,t)=\int_{0}^{r}2\pi r^{\prime}dr^{\prime}B_{z}(r^{\prime},z,t)
145: \label{psi def2}%
146: \end{equation}
147: and the poloidal electric current $I(r,z)$ is defined by
148: \begin{equation}
149: I(r,z,t)=\int_{0}^{r}2\pi r^{\prime}dr^{\prime}J_{z}(r^{\prime},z,t).
150: \label{Idef2}%
151: \end{equation}
152: The definition of $I(r,z)$ is consistent with Ampere's law for the toroidal
153: field, since using $\nabla\phi=\hat{\phi}/r\ $in Eq.\ref{B general2} gives the
154: toroidal magnetic field to be
155: \begin{equation}
156: B_{\phi}=\frac{\mu_{0}I}{2\pi r}. \label{Bphi}%
157: \end{equation}
158: 
159: 
160: Equations \ref{B general2}-\ref{Idef2} describe the magnetic field and
161: electric current distribution of \textit{any} axisymmetric magnetic field.
162: Because astrophysical jets are azimuthally symmetric, their magnetic field
163: must be of the form prescribed by Eqs.\ref{B general2}- \ref{Idef2} and,
164: indeed, it is generally believed that astrophysical jets involve large-scale
165: poloidal magnetic fields threading an accretion disk (e.g., see
166: %TCIMACRO{\TeXButton{\citet{Livio2002,Ferreira2004}}{\citet
167: %{Livio2002,Ferreira2004}}}%
168: %BeginExpansion
169: \citet{Livio2002,Ferreira2004}%
170: %EndExpansion
171: ) and in addition, toroidal magnetic fields. Application of Ampere's law to
172: Eq.\ref{B general2} shows that the poloidal and toroidal currents are
173: respectively given by
174: \begin{equation}
175: \mathbf{J}_{pol}=\frac{1}{2\pi}\nabla I\times\nabla\phi\label{Jpol2}%
176: \end{equation}
177: and
178: \begin{equation}
179: \mathbf{J}_{tor}=\ -\frac{r^{2}\nabla\cdot\left(  r^{-2}\nabla\psi\right)
180: }{2\pi\mu_{0}}\ \nabla\phi\label{Jtor}%
181: \end{equation}
182: showing that poloidal magnetic fields are produced by a toroidal electric
183: current and toroidal magnetic fields are produced by a poloidal current;
184: toroidal vectors are those vectors in the $\phi$ direction and poloidal
185: vectors are those in any combination of the $r$ and $z$ directions.
186: \ Knowledge of the two stream-function quantities $I(r,z,t)$ and $\psi(r,z,t)$
187: is thus necessary and sufficient to determine the complete vector magnetic
188: field and the complete vector current density.
189: 
190: The term `magnetic axis' has traditionally been assigned different meanings in
191: the respective contexts of astrophysics and laboratory toroidal magnetic
192: confinement devices (e.g., tokamaks, reversed field pinches, or spheromaks).
193: Specifically, a local maximum in $r$-$z$ space of $\psi(r,z)$ is called a
194: magnetic axis in the context of toroidal confinement devices whereas the $z$
195: symmetry axis of the magnetic field is called the magnetic axis in \ the
196: context of astrophysics. To avoid confusion, we will call the location of a
197: maximum of $\psi(r,z)$ the poloidal flux magnetic axis. Poloidal magnetic
198: field lines follow level contours of $\psi$ and so one can envision the
199: projection of the magnetic field in the $r$-$z$ plane as being like a set of
200: roads, each at a different altitude, encircling a mountain peak at a specific
201: $r$-$z$ location which is the poloidal flux magnetic axis (also called an
202: O-point). \ Since a toroidal current at infinity is not physical, and since
203: the net magnetic flux enclosed by a circle with infinite radius must vanish as
204: field lines cannot go to infinity, $\psi$ must vanish at infinity.
205: Furthermore, mathematical regularity of physical quantities requires $\psi$ to
206: vanish on the $z$ axis
207: %TCIMACRO{\TeXButton{\citep{Lewis1990}}{\citep{Lewis1990}}}%
208: %BeginExpansion
209: \citep{Lewis1990}%
210: %EndExpansion
211: . Thus, a non-trivial $\psi$ can only be finite in the region $0<r<\infty$,
212: $-\infty<z<\infty.$ The simplest situation of physical interest is therefore
213: where $\psi$ has a single maximum in the $r$-$z$ plane. We will consider this
214: situation, namely a single poloidal field magnetic axis with $\psi(r,z)$
215: symmetric with respect to $z.$ This situation has been previously considered
216: by
217: %TCIMACRO{\TeXButton{citet{Lovelace2002}}{\citet{Lovelace2002}} }%
218: %BeginExpansion
219: \citet{Lovelace2002}
220: %EndExpansion
221: and implies via Eq.\ref{Jtor} that a toroidal current circulates in an
222: accretion disk to produce the poloidal magnetic field
223: \begin{equation}
224: \mathbf{B}_{pol}=\mathbf{\ }\frac{1}{2\pi}\nabla\psi\times\nabla\phi.
225: \label{Bpol def}%
226: \end{equation}
227: An inescapable feature of this topology is that because $\nabla\psi=0$ at the
228: maximum of $\psi$, i.e., at the poloidal field magnetic axis, the poloidal
229: magnetic field has a null on the poloidal field magnetic axis. In the $z=0$
230: plane, the poloidal flux $\psi$ thus starts from zero at $r=0,$ increases to a
231: maximum at the poloidal field magnetic axis, and then decays to zero as
232: $r\rightarrow\infty.$ \ 
233: 
234: We define $a\ $to be the radius of the poloidal field magnetic axis. In
235: addition, we define $\left\langle B_{z}\right\rangle $ to be the
236: \textit{spatially-averaged} axial magnetic field linked by the poloidal field
237: magnetic axis and $\psi_{0}$ to be the value of the poloidal magnetic flux at
238: the poloidal field magnetic axis, so
239: \begin{equation}
240: \left\langle B_{z}\right\rangle =\frac{\int_{0}^{a}dr2\pi rB_{z}(r,0)}%
241: {\int_{0}^{a}dr2\pi r}=\frac{\psi_{0}}{\pi a^{2}}. \label{Bbar}%
242: \end{equation}
243: \qquad\qquad
244: 
245: The axial field $B_{z}=(2\pi r)^{-1}\partial\psi/\partial r$ reverses sign at
246: $r=a$ and the radial field $B_{r}=-(2\pi r)^{-1}\partial\psi/\partial z$
247: reverses sign at $z=0.$ $\ $An analytic representation for a physically
248: realizable generic flux function satisfying all these properties is derived in
249: Appendix \ref{Generic Flux Function}. This generic flux function is
250: \begin{equation}
251: \psi(r,z)=\ \ \frac{27\left(  r/a\right)  ^{2}\psi_{0}}{8\left(  \left(
252: \frac{r}{a}+\frac{1}{2}\right)  ^{2}+\left(  \frac{z}{a}\right)  ^{2}\right)
253: ^{3/2}} \label{generic}%
254: \end{equation}
255: and has the properties that (i)\ $\psi(r,z)$ has a maximum value of $\psi_{0}$
256: at $r=a,$ $z=0,$ (ii) $\psi\sim r^{2}$ for $r\ll a$ and $z=0,$ (iii) $\psi\sim
257: r^{-1}$ for $r\gg a,z$ and (iv) for $r\ll a/2$ or $r\gg a\ $ and for $z\gg a$
258: the contours of $\psi$ are identical to the contours of the poloidal flux
259: produced by a current loop located at $r=a/2,$ $z=0.$ This flux function thus
260: encompasses simpler models which assume a uniform axial magnetic field $B_{z}%
261: $; these simpler models would correspond to the $r,z\ll a$ region here since
262: in this region $\psi\sim r^{2}$ which corresponds to having a uniform axial
263: magnetic field $B_{z}.$ This flux function could also\ be used to describe the
264: far-field of a dipole by assuming that $r,z\gg a.$ Since any real axial
265: magnetic field must always be generated by a toroidal current located at some
266: finite radius, any real situation will have a poloidal flux function
267: qualitatively similar to Eq.\ref{generic}. The flux function prescribed in
268: Eq.\ref{generic} is similar in essence to the flux function used in Fig.1 of
269: %TCIMACRO{\TeXButton{citet{Lovelace2002}}{\citet{Lovelace2002}}}%
270: %BeginExpansion
271: \citet{Lovelace2002}%
272: %EndExpansion
273: .
274: 
275: Figure \ref{GenericFluxFunction-new} plots $\psi(r,z)$ as prescribed by
276: Eq.\ref{generic} and shows that $\psi(r,z)$ has its maximum at the poloidal
277: field magnetic axis $r\ =a,$ $z=0$. This flux function corresponds to a
278: smoothly varying toroidal current density prescribed by Eq.\ref{Jtor}
279: concentrated in the vicinity of $r=a,$ $z=0.$ Since for $z=0$ and small $r$,
280: this function has the asymptotic dependence $\psi\simeq27\psi_{0}\left(
281: r/a\right)  ^{2}$, it corresponds to an approximately uniform axial magnetic
282: field $B_{z}\simeq27\psi_{0}/\pi a^{2}$ for $r,z\ll a.$ The $r\ll a$
283: inner-region $B_{z}$ is thus 27 times stronger than the average $B_{z}$ field
284: between $0$ and $a.$ The total toroidal current $\mathcal{I}_{\phi}$
285: associated with the generic flux function given by Eq.\ref{generic} is
286: calculated in Appendix \ref{current associated with flux function} using the
287: integral form of Ampere's law and found to be
288: \begin{equation}
289: \mathcal{I}_{\phi}=\frac{27\psi_{0}}{\ \pi a\mu_{0}}\ \ . \label{Iphi}%
290: \end{equation}
291: 
292: 
293: \begin{figure}[ptb]
294: \caption{Plot of the normalized generic flux function $\psi(r,z)/\psi_{0}\ $in
295: coordinates normalized to the radius of the poloidal field magnetic axis, that
296: is to the radial position of the maximum of $\psi(r,z)\,.$ Contours of
297: iso-surfaces shown on top; these correspond to projection of poloidal magnetic
298: field onto $r$-$z$ plane.}%
299: \plotone{f1.eps}\label{GenericFluxFunction-new}%
300: \end{figure}
301: 
302: The laboratory jets involve the mutual interaction between poloidal and
303: toroidal magnetic fields powered by laboratory capacitor banks. The jet
304: acceleration mechanism results from the pressure of the toroidal magnetic
305: field inflating flux surfaces associated with the poloidal magnetic field. The
306: question arises as to what powers the toroidal and poloidal magnetic fields in
307: an actual astrophysical situation. Existing models of astrophysical jets are
308: based on the magnetohydrodynamic (MHD) approximation of plasma behavior and
309: typically assume (i)\ the poloidal field is pre-existing and (ii)\ the
310: toroidal field results from a rotating accretion disk twisting up this assumed
311: primordial poloidal field. The purpose of this paper is to present an
312: alternate model wherein it is postulated that the toroidal and poloidal field
313: result instead from a non-MHD dusty plasma dynamo mechanism that converts the
314: gravitational energy of infalling dust grains into an electrical power source
315: that drives poloidal and toroidal electric currents creating the respective
316: toroidal and poloidal fields. A brief outline of how infalling charged dust
317: can drive poloidal currents has been presented in
318: %TCIMACRO{\TeXButton{\citet{Bellan2007}}{\citet{Bellan2007}}}%
319: %BeginExpansion
320: \citet{Bellan2007}%
321: %EndExpansion
322: .
323: 
324: This model obviously requires existence of sufficient infalling dust to
325: provide the jet power. Since the dust-to-gas mass ratio in the Interstellar
326: Medium (ISM) is 1\%, one might be tempted to argue that any jet driven by the
327: proposed dust infall mechanism would be limited to having\ less than 1\% of
328: the power available from infalling gas, a constraint that would contradict
329: observations. However, in
330: %TCIMACRO{\TeXButton{\citet{Bellan2008}}{\citet{Bellan2008}} }%
331: %BeginExpansion
332: \citet{Bellan2008}
333: %EndExpansion
334: (to be referred to as Paper I), we showed that the dust-to-gas mass ratio in a
335: molecular cloud can be substantially enriched compared to the ISM\ value
336: (e.g., the dust to gas mass ratio in a molecular cloud could be enriched
337: 20-fold compared to the 1\% ISM value). This enrichment occurs because
338: accreting dust slows down much more in proportion to its initial velocity than
339: does accreting gas so that the density amplification resulting from dust
340: slowing down is much greater than the corresponding density amplification of gas.
341: 
342: The condition for the toroidal magnetic field to inflate the poloidal magnetic
343: field and create a jet can be expressed as
344: \begin{equation}
345: \frac{\mu_{0}I}{\psi}>\lambda\label{lambda threshold}%
346: \end{equation}
347: where $\lambda$ is a parameter of the order of the inverse characteristic
348: linear dimension in the radial direction. The ratio $I/\psi$ can be thought of
349: as the ratio of the electric current flowing along a flux tube to the magnetic
350: flux content of the flux tube and is proportional to the twist of the magnetic
351: field. Equation \ref{lambda threshold}, well-established in spheromak
352: formation physics
353: %TCIMACRO{\TeXButton{\citep
354: %{Barnes1990,Jarboe1994,Geddes1998,Bellan2000,Hsu2005}}{\citep
355: %{Barnes1990,Jarboe1994,Geddes1998,Bellan2000,Hsu2005}}}%
356: %BeginExpansion
357: \citep{Barnes1990,Jarboe1994,Geddes1998,Bellan2000,Hsu2005}%
358: %EndExpansion
359: , is essentially a statement that jet expansion (i.e., poloidal field
360: inflation)\ occurs when the toroidal magnetic field pressure force $\sim
361: B_{\phi}^{2}A_{1}$ acting on area $A_{1}$ exceeds the restraining force
362: $B_{z}^{2}A_{2}$ of the poloidal magnetic field `tension' acting on area
363: $A_{2}$. Here $A_{1}$ and $A_{2}$ are not exactly the same because the
364: toroidal and poloidal fields do not act over the same areas. \ The equivalence
365: between Eq. \ref{lambda threshold} and the condition $B_{\phi}^{2}>B_{z}%
366: ^{2}A_{2}/A_{1}$ is seen by substituting $\mu_{0}I=2\pi aB_{\phi}$ from
367: Ampere's law and $\psi\sim B_{z}\pi a^{2}$ in Eq.\ref{lambda threshold}.
368: 
369: Paper I divided the regions of interest into successively smaller concentric
370: regions and considered dust and gas behavior in the outermost regions.
371: Simultaneous gas and dust accretion were considered and it was shown that the
372: dust could be considered as a perturbation on the gas, so that the gas
373: accretion problem could be solved first without considering dust and then the
374: solution of this gas accretion problem could be used as an input for the dust
375: accretion problem. Below is a listing showing which regions were considered in
376: Paper I, which are considered in this paper, and which will be considered in a
377: future paper; the nominal radii scales and star mass are from Table 3 in Paper I:
378: 
379: \begin{description}
380: \item \textit{ISM scale (considered in Paper I):} The outermost scale is that
381: of the Interstellar Medium (ISM). The ISM\ has a gas density $\sim$10$^{7}$
382: m$^{-3}$, a dust-to-gas mass ratio of 1 percent, a gas temperature
383: $T_{g}^{ISM}\sim100$ K$,$ and is optically thin. The ISM is assumed to be
384: spatially uniform and to bound a molecular cloud having radius $r_{edge}$
385: $\sim10^{5}$ a.u.
386: 
387: \item \textit{Molecular cloud scale (considered in Paper I):}\ The molecular
388: cloud scale has much higher density than the ISM and is characterized by force
389: balance between gas self-gravity\ and gas pressure. The molecular cloud scale
390: is sub-divided into a large, radially non-uniform low-density outer region and
391: a small, approximately uniform, high-density inner core region. Clouds have a
392: characteristic scale given by the Jeans length $r_{J}\sim1.4\times10^{4}$ a.u.
393: The radial dependence of gas density is provided by the Bonnor-Ebert sphere
394: solution which acts as the outer boundary of the Bondi accretion scale.
395: 
396: \item \textit{Bondi accretion scale (considered in Paper I):} The Bondi
397: accretion scale is $\sim r_{B}\sim4.3\times10^{3}$ a.u. which is sufficiently
398: small that gas self-gravity no longer matters so equilibrium is instead
399: obtained by force balance between gas pressure and the gravity of a central
400: object assumed to be a star having mass $M\sim0.4M_{\odot}$ The Bondi scale is
401: sub-divided into three concentric radial regions: an outermost region where
402: the gas flow is subsonic, a critical transition radius at exactly $r_{B}$
403: where the flow is sonic, and an innermost region where the gas flow is
404: free-falling and supersonic.
405: 
406: \item \textit{Collisionless dusty plasma scale (considered in this paper): }
407: Free-falling dust grains collide with each other in one of the above scales
408: and coagulate to form large-radius grains which are collisionless and
409: optically thin. The optically thin dust absorbs UV photons from the star,
410: photo-emits electrons and becomes electrically charged. The charged dust
411: grains are subject to electromagnetic forces in addition to gravity. Motions
412: of charged dust grains relative to electrons result in electric currents with
413: associated poloidal and toroidal magnetic fields [see preliminary discussion
414: in
415: %TCIMACRO{\TeXButton{\citet{Bellan2007}}{\citet{Bellan2007}}}%
416: %BeginExpansion
417: \citet{Bellan2007}%
418: %EndExpansion
419: ]. \ This region is assumed to have a scale of $10-10^{3}$ a.u.\ \ and
420: corresponds to the scale of $a,$ the radius of the poloidal magnetic field
421: axis. It is assumed that a distributed toroidal current peaked at a nominal
422: radius $a$ is responsible for producing a poloidal field having the generic
423: profile given in Fig.\ref{GenericFluxFunction-new}.
424: 
425: \item \textit{Jet scale (to be considered in a future publication): }The
426: electric currents interact with the magnetic fields to produce
427: magnetohydrodynamic forces that drive astrophysical jets in a manner
428: consistent with Eq.\ref{lambda threshold} and analogous to that reported in \
429: %TCIMACRO{\TeXButton{\citet{Hsu2002,Hsu2005}}{\citet{Hsu2002,Hsu2005}} }%
430: %BeginExpansion
431: \citet{Hsu2002,Hsu2005}
432: %EndExpansion
433: and
434: %TCIMACRO{\TeXButton{\citet{Bellan2005}}{\citet{Bellan2005}}}%
435: %BeginExpansion
436: \citet{Bellan2005}%
437: %EndExpansion
438: . This region is assumed to have a scale $\ll10^{3}$ a.u., possibly as small
439: as a few a.u. and will involve a deformation of the generic poloidal field
440: profile given in Fig.\ref{GenericFluxFunction-new} because of the pressure of
441: toroidal magnetic field inflating the poloidal flux surfaces.
442: \end{description}
443: 
444: \section{Outline of model}
445: 
446: We will show how infalling collisionless dust grains can develop special three
447: dimensional orbits suitable for sustaining both toroidal and poloidal dynamos.
448: This result is obtained by considering Hamiltonian particle dynamics in the
449: combination of the gravitational field of a star with mass $M$ and a
450: three-dimensional axisymmetric magnetic field topology consistent with
451: previous models of magnetically driven astrophysical jets [e.g.,
452: %TCIMACRO{\TeXButton{citet{Lovelace1976}}{\citet{Lovelace1976}}}%
453: %BeginExpansion
454: \citet{Lovelace1976}%
455: %EndExpansion
456: ,
457: %TCIMACRO{\TeXButton{citet{Li2001}}{\citet{Li2001}}}%
458: %BeginExpansion
459: \citet{Li2001}%
460: %EndExpansion
461: ,
462: %TCIMACRO{\TeXButton{citet{Lovelace2002}}{\citet{Lovelace2002}}}%
463: %BeginExpansion
464: \citet{Lovelace2002}%
465: %EndExpansion
466: , and
467: %TCIMACRO{\TeXButton{citet{Lynden-Bell2003}}{\citet{Lynden-Bell2003}}}%
468: %BeginExpansion
469: \citet{Lynden-Bell2003}%
470: %EndExpansion
471: ]. The reason why dust grains develop these special orbits will be shown to be
472: due to charging of dust grains via photo-emission of electrons. The analysis
473: involves using Hamiltonian mechanics to generalize the centrifugal potential
474: so as to include magnetic force, i.e., the St\"{o}rmer effective potential is
475: used. St\"{o}rmer potentials have been previously used for investigating
476: auroral particles
477: %TCIMACRO{\TeXButton{citep{Stormer1955}}{\citep{Stormer1955}}}%
478: %BeginExpansion
479: \citep{Stormer1955}%
480: %EndExpansion
481: , electron and ion motion in the magnetosphere
482: %TCIMACRO{\TeXButton{citep{Shebalin2004,Lemaire2003}}{\citep
483: %{Shebalin2004,Lemaire2003}} }%
484: %BeginExpansion
485: \citep{Shebalin2004,Lemaire2003}
486: %EndExpansion
487: and most recently, charged dust grain motion in the magneto-gravitational
488: fields of Saturn and Jupiter
489: %TCIMACRO{\TeXButton{citep{Dullin2002,Mitchell2003}}{\citep
490: %{Dullin2002,Mitchell2003}}}%
491: %BeginExpansion
492: \citep{Dullin2002,Mitchell2003}%
493: %EndExpansion
494: . St\"{o}rmer potentials are also commonly used to characterize particle
495: orbits in tokamaks
496: %TCIMACRO{\TeXButton{citep{Rome1979}}{\citep{Rome1979}} }%
497: %BeginExpansion
498: \citep{Rome1979}
499: %EndExpansion
500: and St\"{o}rmer potentials were found to be important in the MHD-driven jet
501: experiment reported by
502: %TCIMACRO{\TeXButton{\citet{Tripathi2007}}{\citet{Tripathi2007}}}%
503: %BeginExpansion
504: \citet{Tripathi2007}%
505: %EndExpansion
506: . We will restrict the analysis to showing how toroidal and poloidal field
507: dynamos can be sustained in steady state by these special Hamiltonian particle
508: orbits; the much more complicated problem of how a dynamo grows from a seed
509: magnetic field will not be addressed here. These special orbits are quite
510: different from conventional cyclotron orbits. As reviewed in Appendix
511: \ref{Orbit review} a dynamo cannot be sustained by particles executing
512: cyclotron orbits because cyclotron orbits and associated drifts are
513: diamagnetic, i.e., create magnetic fields that oppose the field in which the
514: particle is orbiting.
515: 
516: The importance of a Hamiltonian analysis can be appreciated by considering the
517: gedanken experiment where the charge to mass ratio of a particle in \ a
518: combined gravitational-magnetic field is assumed to be increased from zero
519: (neutral particle) to that of an electron or ion. The particle will thus make
520: a transition from Kepler to cyclotron orbital motion. The details of how this
521: transition occurs have been examined by
522: %TCIMACRO{\TeXButton{\citet{Bellan2007}}{\citet{Bellan2007}} }%
523: %BeginExpansion
524: \citet{Bellan2007}
525: %EndExpansion
526: in the context of uniform-magnetic-field orbits restricted to a plane. The
527: present paper will address this issue in the more general context of three
528: dimensional particle orbits in a spatially non-uniform three dimensional
529: magnetic field having dipole-like topology appropriate for an accretion disk;
530: similar dipole topology has been previously invoked for accretion disks by
531: %TCIMACRO{\TeXButton{citet{Lovelace2002}}{\citet{Lovelace2002}}}%
532: %BeginExpansion
533: \citet{Lovelace2002}%
534: %EndExpansion
535: . Our analysis identifies five distinct classes of orbits and shows that the
536: class to which a given charged particle belongs depends both on its\ charge to
537: mass ratio and on the circumstances under which the charged particle was
538: created from an initially neutral particle. The interaction between the
539: distinct symmetries of the magnetic and gravitational fields removes the
540: isotropy of the incident neutral particles existent prior to charging so that
541: the newly formed charged particles separate into groups having qualitatively
542: different types of orbits. Some orbits correspond to a simple accretion, some
543: involve accretion and production of a dynamo driving toroidal current, and
544: some involve accretion and a dynamo driving poloidal current. The type of
545: orbit a charged particle develops depends on both the angular momentum and the
546: angle of incidence of the parent neutral particle.
547: 
548: The paper is organized with the goal of being concise while also realizing
549: that some readers may not be familiar with the concepts of adiabatic versus
550: non-adiabatic orbits, Speiser orbits, St\"{o}rmer effective potentials, and
551: how conservation of canonical angular momentum results in confinement of an
552: adiabatic particle to the vicinity of a poloidal flux surface. Rather than
553: reviewing these concepts in an introductory section , they are instead
554: discussed in appendices.
555: 
556: \section{Reduction of collisionality due to dust agglomeration}
557: 
558: Paper I showed that dust grains are collisionally decoupled from gas in the
559: ISM and then become collisionally coupled to gas in the Bonner-Ebert and Bondi
560: regions of a molecular cloud. Because of the spherical focusing of the dust
561: and gas inflows, the dust density increases to a level such that dust-dust
562: collisions become important. When dust grains collide with each other they may
563: agglomerate to form larger dust grains.
564: %TCIMACRO{\TeXButton{\citet{Przygodda2003}}{\citet{Przygodda2003}} }%
565: %BeginExpansion
566: \citet{Przygodda2003}
567: %EndExpansion
568: and
569: %TCIMACRO{\TeXButton{\citet{vanBoekel2003}}{\citet{vanBoekel2003}} }%
570: %BeginExpansion
571: \citet{vanBoekel2003}
572: %EndExpansion
573: have reported direct observational evidence of grain growth in circumstellar
574: disks while, in addition,
575: %TCIMACRO{\TeXButton{\citet{Jura1980}}{\citet{Jura1980}}}%
576: %BeginExpansion
577: \citet{Jura1980}%
578: %EndExpansion
579: ,
580: %TCIMACRO{\TeXButton{\citet{Miyake1993}}{\citet{Miyake1993}}}%
581: %BeginExpansion
582: \citet{Miyake1993}%
583: %EndExpansion
584: ,
585: %TCIMACRO{\TeXButton{\citet{Pollack1994}}{\citet{Pollack1994}}}%
586: %BeginExpansion
587: \citet{Pollack1994}%
588: %EndExpansion
589: ,
590: %TCIMACRO{\TeXButton{\citet{D'Alessio2001}}{\citet{D'Alessio2001}}}%
591: %BeginExpansion
592: \citet{D'Alessio2001}%
593: %EndExpansion
594: , and
595: %TCIMACRO{\TeXButton{\citet{Dullemond2005}}{\citet{Dullemond2005}} }%
596: %BeginExpansion
597: \citet{Dullemond2005}
598: %EndExpansion
599: provided detailed calculations showing a strong tendency for dust grain growth
600: when dust grains collide with each other. This agglomeration will increase the
601: dust grain radius $r_{d}$ while keeping the dust mass density $\rho_{d}$
602: constant. We will consider first how this agglomeration affects dust-gas
603: collisions and then how it affects dust-dust collisions.
604: 
605: Since the mean free path is much larger than the grain radius, the drag force
606: on a dust grain due to collisions with gas molecules is of the Epstein-type
607: and given by
608: %TCIMACRO{\TeXButton{\citep{Lamers1999}}{\citep{Lamers1999}} }%
609: %BeginExpansion
610: \citep{Lamers1999}
611: %EndExpansion
612: \begin{equation}
613: F_{drag}=-(u_{d}-u_{g})\rho_{g}\sigma_{d}\sqrt{c_{g}^{2}+(u_{d}-u_{g})^{2}}
614: \label{collision freq}%
615: \end{equation}
616: where $c_{g}$ is the gas thermal velocity, $u_{d}$ is the dust grain velocity,
617: $\sigma_{d}$ is the dust grain cross-sectional area, and $u_{g}$ is the mean
618: velocity of the gas (i.e., the fluid velocity). In the innermost Bondi region
619: where flow is supersonic, we may approximate $c_{g}\simeq0$ and work in a
620: frame moving with\ $u_{g}$ by defining $\Delta u_{d}=u_{d}-u_{g}.$ The dust
621: equation of motion in this frame is thus%
622: 
623: \begin{equation}
624: m_{d}\frac{d\Delta u_{d}}{dt}=-\left(  \Delta u_{d}\right)  ^{2}\rho_{g}%
625: \sigma_{d}.\ \label{collision}%
626: \end{equation}
627: Defining $\xi$ to be distance in the direction of dust motion so $\Delta
628: u_{d}=d\xi/dt,$ Eq.\ref{collision} can be recast as%
629: \begin{equation}
630: \frac{d\Delta u_{d}}{d\xi}\Delta u_{d}=-\left(  \Delta u_{d}\right)
631: ^{2}\ \sigma_{d}\frac{\rho_{g}}{m_{d}}\ .
632: \label{collision with integrating factor}%
633: \end{equation}
634: Integration gives%
635: \begin{equation}
636: \Delta u_{d}(\xi)=\Delta u_{d}(0)\exp\left(  -\xi/l_{dg}\right)
637: \label{slowing down}%
638: \end{equation}
639: where the dust-gas collision mean free path is
640: \begin{equation}
641: l_{dg}=\frac{m_{d}}{\rho_{g}\sigma_{d}}. \label{mean free path}%
642: \end{equation}
643: 
644: 
645: Since the dust cross-section and mass are given respectively by
646: \begin{equation}
647: \sigma_{d}=\pi r_{d}^{2} \label{sigma_d}%
648: \end{equation}
649: and%
650: \begin{equation}
651: m_{d}=\frac{4\pi r_{d}^{3}\rho_{d}^{int}}{3}\ \label{md}%
652: \end{equation}
653: where $\ $ $\rho_{d}^{int}$ is the intrinsic density of a dust grain, the
654: dust-gas collision mean free path can be expressed as%
655: \begin{equation}
656: l_{dg}=\frac{4\rho_{d}^{int}}{3\rho_{g}}r_{d}\ \label{ldg rd}%
657: \end{equation}
658: which shows that dust agglomeration increases the dust-gas mean free path and
659: so will tend to make dust collisionless with respect to gas.
660: 
661: Let us now consider how agglomeration affects dust-dust collisions. We first
662: note that the condition for dust-dust collisions to be significant is closely
663: related to the condition for the dust to be optically thick: if $l$ is the
664: characteristic length of a configuration, the condition for collisions to be
665: significant is $\rho_{d}\sigma_{d}l/m_{d}>1$ whereas the condition for the
666: dust to be optically thick is $Q_{eff}\rho_{d}\sigma_{d}l/m_{d}>1$ where $\ $
667: $Q_{eff}$ is an extinction efficiency parameter that depends on the ratio of
668: the dust radius to the light wavelength. The dust-dust collision mean free
669: path is thus%
670: \begin{equation}
671: l_{dd}=\frac{m_{d}}{\rho_{d}\sigma_{d}}=\frac{4\rho_{d}^{int}}{3\rho_{d}}%
672: r_{d}\ \label{ldd}%
673: \end{equation}
674: so if, as argued in Paper I, \ the dust mass density $\rho_{d}$ has been
675: enriched to be a significant fraction of the gas mass density $\rho_{g},$ the
676: dust-dust collision mean free path $l_{dd}$ will be the same order of
677: magnitude as the dust-gas mean free path $l_{dg}.$ Agglomeration will thus
678: tend to increase both the dust-dust and dust-gas collision mean free paths,
679: and furthermore will cause the dust to become optically thin. We will assume
680: that dust grains agglomerate when the dust number density $\ n_{d}=\rho
681: _{d}/m_{d}$ becomes sufficiently large for dust-dust collisions to occur and
682: that this agglomeration results in an increase in $r_{d}$ until the dust
683: grains become collisionless and optically thin again. We will not attempt to
684: follow the dynamics of the agglomeration process, relying instead on the
685: analysis in the papers cited above. Our starting point then will be assuming
686: the existence of collisionless dust grains exposed to star light, having
687: radius $r_{d}$ larger than in the ISM, and as discussed in Paper I, having a
688: dust to gas mass density ratio substantially enriched compared to the 1\%
689: value in the ISM.
690: 
691: \section{Review: Neutral particle motion in a gravitational
692: field\label{Kepler}}
693: 
694: For reference and in order to define terms to be used later in a more complex
695: context, we first review the elementary problem of the motion of a neutral
696: particle of mass $m$ in the gravitational field of a star of mass $M$. The
697: particle we have in mind could be a a dust grain with radius $r_{d}$
698: sufficiently large to be collisionless over the distance from its starting
699: point to the star.
700: 
701: The equations governing the motion of this neutral particle are spherically
702: symmetric whereas the motions of a charged particle in an azimuthally
703: symmetric electromagnetic field are cylindrically symmetric. An axisymmetric
704: magnetic field is assumed to exist in the lab frame and the $z$ axis is
705: defined by the direction of this magnetic field at the origin. Although the
706: neutral particle trajectory is unaffected by this magnetic field, we
707: nevertheless use the magnetic field coordinate system to define the lab frame.
708: Depending on what is being emphasized, the lab frame will be characterized by
709: either a cylindrical coordinate system $\{r,\phi,z\}\ $or by a Cartesian
710: coordinate system \thinspace$\{x,y,z\}$ so that $x=r\cos\phi,$ and
711: $y=r\sin\phi.$ Because the force is central, the neutral particle angular
712: momentum vector $\mathbf{L}=m\mathbf{r\times\dot{r}}$ is invariant and so the
713: neutral particle moves in an orbital plane normal to $\mathbf{L}$. The lab and
714: orbital planes are sketched in Fig.\ref{Orbital-Plane-coordinate-system}.\ The
715: $x$ axis of the lab frame is defined to be in the direction of the unit vector
716: $\hat{x}=$ $\hat{z}\times\mathbf{L}/L$ and the $y$ axis of the lab frame is
717: defined to be in the direction of the unit vector $\hat{y}=\hat{z}%
718: \times\left(  \hat{z}\times\mathbf{L}/L\right)  $. The orbital plane is tilted
719: with respect to the lab frame by an angle $\theta$ about the $x$ axis. The
720: $x^{\prime}$ axis of the orbital frame is defined to be coincident with the
721: $x$ axis of the lab frame and the $y^{\prime}$ axis of the orbital plane is an
722: uptilted version of the $y$ axis of the lab frame.
723: 
724: \begin{figure}[ptb]
725: \caption{Lab frame has Cartesian coordinates $x,y,z$ and the magnetic field is
726: axisymmetric with respect to the lab frame $z$ axis. The orbital plane of a
727: neutral particle is normal to the neutral particle angular momentum vector
728: $\mathbf{L}$ which is tilted by an angle $\theta$ with respect to the $z$
729: axis. The orbital plane Cartesian coordinates are $x^{\prime},y^{\prime}$
730: where the $x^{\prime}$ axis is coincident with the $x$ axis. \ The neutral
731: particle makes a circular Kepler, elliptical Kepler, or cometary orbit in its
732: orbital plane (cometary orbit shown).}%
733: \label{Orbital-Plane-coordinate-system}%
734: \plotone{f2.eps}\end{figure}
735: 
736: $\theta=0$ corresponds to prograde motion in the lab frame (i.e., the neutral
737: particle moves in the same sense as the toroidal current that produces the
738: magnetic field $B_{z}$ on the $z$ axis), $\theta=\pi$ corresponds to
739: retrograde motion in the lab frame, and $\theta=\pi/2$ corresponds to a polar
740: orbit. For purposes of following the trajectory in the orbital plane it is
741: convenient to use cylindrical coordinates $\rho,\eta$ defined in the
742: orbital\ plane such that $x^{\prime}=\rho\cos\eta$ and $y^{\prime}=\rho
743: \sin\eta.$ The Hamiltonian for a neutral particle moving in its orbital plane
744: can then be written as
745: \begin{equation}
746: H=\frac{1}{2}mv_{\rho}^{2}+\frac{L^{2}}{2m\rho^{2}}-\frac{mMG}{\rho}
747: \label{vt2}%
748: \end{equation}
749: where
750: \begin{equation}
751: L=m\rho v_{\eta}, \label{p}%
752: \end{equation}
753: the magnitude of the mechanical angular momentum vector, is an invariant
754: positive scalar. The Kepler angular frequency at a reference radius $a\ $\ is
755: defined as
756: \begin{equation}
757: \Omega_{0}=\sqrt{MG/a^{3}}. \label{Kepler0}%
758: \end{equation}
759: The value of $a$ is chosen to be the radius of the poloidal magnetic field
760: axis. Normalized quantities are defined as%
761: \begin{equation}%
762: \begin{array}
763: [c]{c}%
764: \bar{\rho}=\rho/a,\tau=\Omega_{0}t,\bar{v}_{\rho}=\frac{v_{\rho}}{\Omega_{0}%
765: a}\\
766: \bar{L}=\ \frac{L}{m\Omega_{0}a^{2}},\quad\bar{H}=\frac{H}{m\Omega_{0}%
767: ^{2}a^{2}}.
768: \end{array}
769: \label{norm}%
770: \end{equation}
771: Equation \ref{vt2} can then be expressed in dimensionless form as
772: \begin{equation}
773: \bar{H}=\frac{\bar{v}_{\rho}^{2}}{2}+\frac{\bar{L}^{2}}{2\bar{\rho}^{2}}%
774: -\frac{1\ }{\bar{\rho}}. \label{nondim}%
775: \end{equation}
776: 
777: 
778: The last two terms depend on $\bar{\rho}$ and so constitute an effective
779: potential
780: \begin{equation}
781: \bar{\chi}(\bar{\rho})=\frac{\bar{L}^{2}}{2\bar{\rho}^{2}}-\frac{1}{\bar{\rho
782: }}. \label{chi netural}%
783: \end{equation}
784: This effective potential depends parametrically on $\bar{L}$ which is a
785: property of the particle and not the environment. Two different particles at
786: the same position but having different values of $\bar{L}$ will have different
787: effective potentials and so march to a \textquotedblleft different
788: drummer\textquotedblright. This \textquotedblleft different
789: drummer\textquotedblright\ concept will re-appear later in a more elaborate
790: fashion when the motion of charged particles is considered.
791: 
792: $\bar{\chi}(\bar{r})$ attains its minimum value $\chi_{\min}=-1/2\bar{L}^{2}$
793: at the normalized radius $\bar{\rho}=\bar{L}^{2}$. \ A particle with energy
794: equal to this minimum has $\bar{v}_{\rho}=0$ and therefore has a circular
795: orbit with angular frequency $d\eta/d\tau=\ L/\Omega_{0}ma^{2}=\ \bar{L}%
796: /\bar{\rho}^{2}.$ Hence, if $\bar{L}=\ 1$ the minimum-energy particle traces
797: out a circular Kepler orbit with $d\eta/d\tau=1$ and has an energy $\bar
798: {H}=-1/2.$ A particle with energy $-1/2<\bar{H}<0$ cannot escape to infinity
799: and so has a bounded elliptical Kepler orbit. The effective potential
800: prescribed by Eq.\ref{chi netural} for a particle with $\bar{L}=\ 1$ is shown
801: in Fig.\ref{Effective-Potential-Neutral-Particle}(a).
802: 
803: \begin{figure}[ptb]
804: \caption{(a)\ Effective potential for a neutral particle having $\bar{L}=1$;
805: (b)\ effective potential for a charged particle with appropriate values of
806: canonical angular momentum and poloidal flux function.}%
807: \label{Effective-Potential-Neutral-Particle}%
808: \plotone{f3.eps}\end{figure}
809: 
810: Reflection (pericenter) of a particle \ occurs when $\bar{v}_{\rho}=0$ in
811: which case Eq.\ref{nondim} gives%
812: \begin{equation}
813: \bar{\rho}_{pericenter}=\frac{\bar{L}^{2}}{1+\sqrt{1+2\bar{L}^{2}\bar{H}}}.
814: \label{perigee}%
815: \end{equation}
816: Reflection at the pericenter \ can be considered to be the consequence of a
817: potential barrier preventing the particle from accessing the region $\bar
818: {\rho}<\bar{\rho}_{pericenter};$ the effective potential in the inaccessible
819: region exceeds the total available energy.$\ $Thus an unbounded particle with
820: $\bar{L}=1$ also has the effective potential shown Fig.
821: \ref{Effective-Potential-Neutral-Particle}(a), but unlike the bounded $\bar
822: {H}=-1/2$ Kepler particle, the unbounded particle reflects from the pericenter
823: potential barrier and so has a cometary orbit.
824: 
825: In order for an incoming unbound particle to access a given $\bar{\rho}$
826: without being reflected at some larger radius, the condition that $\bar
827: {v}_{\rho}^{2}~$cannot be negative gives the constraint on angular momentum
828: that
829: \begin{equation}
830: \bar{L}^{2}<2\bar{\rho}^{2}\bar{H}+2\bar{\rho}. \label{bound on p}%
831: \end{equation}
832: Since a particle with zero angular momentum will simply fall into the central
833: object, in order for a particle to be both unbounded and \ able to access the
834: radius $\bar{\rho}$ its angular momentum is constrained to lie in the range%
835: \begin{equation}
836: 0<\bar{L}^{2}<2\bar{\rho}^{2}\bar{H}+2\bar{\rho}. \label{p range}%
837: \end{equation}
838: 
839: 
840: Solution of the equation of motion
841: %TCIMACRO{\TeXButton{\citep{Goldstein1950}}{\citep{Goldstein1950}} }%
842: %BeginExpansion
843: \citep{Goldstein1950}
844: %EndExpansion
845: shows that the orbit can be expressed as
846: \begin{equation}
847: \frac{1}{\ \bar{\rho}}=\frac{1-\sqrt{1+2\bar{L}^{2}\ \bar{H}}\cos\left(
848: \eta-\alpha\right)  }{\bar{L}^{2}}\ \ \label{solution}%
849: \end{equation}
850: where $\alpha,$ which we call the clock angle in the orbital plane, is the
851: angle between the symmetry line of the orbit (the line passing through the
852: central object and the pericenter position) and the lab frame $x$ axis (which
853: is also the $x^{\prime}$ axis of the orbital plane).
854: 
855: The Cartesian orbit coordinates $\bar{x}^{\prime}=\bar{\rho}\cos\eta\ $and
856: $\bar{y}^{\prime}=\bar{\rho}\sin\eta$ in the orbital plane (denoted by a prime
857: to distinguish this plane from the lab frame)\ are%
858: \begin{equation}%
859: \begin{array}
860: [c]{c}%
861: \bar{x}^{\prime}=\frac{\bar{L}^{2}\cos\eta}{1-\sqrt{1+2\bar{L}^{2}\ \bar{H}%
862: }\cos\left(  \eta-\alpha\right)  }\\
863: \bar{y}^{\prime}=\frac{\bar{L}^{2}\sin\eta}{1-\sqrt{1+2\bar{L}^{2}\ \bar{H}%
864: }\cos\left(  \eta-\alpha\right)  }%
865: \end{array}
866: \label{unbounded Cartesian}%
867: \end{equation}
868: If the effective potential had a different shape, say the shape shown in
869: Fig.\ref{Effective-Potential-Neutral-Particle}(b) with $\bar{\chi}%
870: \rightarrow0$ at large $\bar{\rho},$ then a particle with $\bar{H}\geq0$ could
871: be trapped in one of the two minima of this effective potential. However, a
872: particle coming from infinity would still be unbounded and would just reflect
873: from some potential barrier$.$ The inability of a static Hamiltonian system to
874: trap a particle coming from infinity is independent of the shape of the
875: Hamiltonian and results from the intrinsic time reversibility of Hamiltonian dynamics.
876: 
877: \section{Comparison of gravitational/magnetic forces to Poynting-Robertson
878: force and to radiation pressure}
879: 
880: \qquad The analysis in this paper is based on the assumption that the
881: trajectory of charged dust grains results primarily from a competition between
882: gravitational and magnetic forces with the possibility that in certain
883: situations electrostatic forces and collisional drag can\ also be important.
884: Two other types of forces, namely those due to the Poynting-Robertson effect
885: and due to radiation pressure, also exist and so it is important to check to
886: see if these additional forces need to be taken into account. This will be
887: done by making a comparison with the nominal magnetic force on a charged dust
888: grain. The magnetic force depends on the strength of the magnetic field, a
889: quantity which has been estimated in self-consistent fashion in Paper III to
890: be in the range $10^{-8}$ to $10^{-6}$ T (i.e., 0.1 to 10 mG) for a nominal
891: YSO jet-disk system where the dust grains have coagulated to a nominal radius
892: $r_{d}=$3 $\mu$m. This estimate of the magnetic field is in rough order of
893: magnitude agreement with measurements reported by
894: %TCIMACRO{\TeXButton{\citet{Chrysostomou1994}}{\citet{Chrysostomou1994}}}%
895: %BeginExpansion
896: \citet{Chrysostomou1994}%
897: %EndExpansion
898: , by
899: %TCIMACRO{\TeXButton{\citet{Roberts1997}}{\citet{Roberts1997}} }%
900: %BeginExpansion
901: \citet{Roberts1997}
902: %EndExpansion
903: and by
904: %TCIMACRO{\TeXButton{\citet{Itoh1999}}{\citet{Itoh1999}}}%
905: %BeginExpansion
906: \citet{Itoh1999}%
907: %EndExpansion
908: , and also is in agreement with the expectation that the magnetic fields in a
909: disk jet system should be much stronger than the nominal $10^{-10}$ T (i.e., 1
910: $\mu$G) magnetic fields of the ISM.
911: 
912: The radiation pressure acting on a dust grain at a distance $r$ from a star
913: with luminosity $L$ is
914: \begin{equation}
915: P_{rad}=\frac{L}{4\pi r^{2}c}Q_{rad}(r_{d}) \label{Prad}%
916: \end{equation}
917: where $Q_{rad}(r_{d})$ is the efficiency with which the photons are
918: absorbed/reflected by the dust grain. This pressure results in a radial
919: outwards force $F_{rad}=P_{rad}\sigma_{d}$. If the dust grain radius is much
920: larger than $\lambda_{rad}$ the wavelength of the radiation, then
921: $Q_{rad}\simeq1$ whereas if the dust grain radius is much smaller than the
922: wavelength of the radiation then $Q_{rad}\sim(\lambda_{rad}/r_{d})^{4}\ll1.$
923: The nominal $r_{d}\sim3$ $\mu$m dust grains assumed here are much larger than
924: the nominal light wavelength and so $Q_{rad}\sim1.$
925: 
926: Since the gravitational force $F_{g}=mMG/r^{2}$ is also in the radial
927: direction, the force due to radiation pressure and gravity compete; the ratio
928: of radiation pressure force to gravitational force on a dust grain is
929: \begin{equation}
930: \alpha=\frac{LQ_{rad}\sigma_{d}}{4\pi cm_{d}MG}=\frac{3}{16\pi}\frac{L}%
931: {MG\rho^{int}c}\frac{Q_{rad}(r_{d})}{r_{d}} \label{radiation pressure ratio}%
932: \end{equation}
933: where Eqs.\ref{sigma_d} and \ref{md} have been used. Assuming $r_{d}=3$ $\mu
934: $m, nominal luminosity $L=L_{\odot}=$ $4\times10^{26}$ watts, $M=M_{\odot},$
935: intrinsic dust density $\rho^{int}=2\times10^{3}$ kg m$^{-3},$ and
936: $Q_{rad}(r_{d})=1$ gives $\alpha=10^{-1}$ so radiation pressure can be ignored
937: compared to gravitational force.
938: 
939: The force on a dust grain due to the Poynting-Robertson effect is smaller by a
940: factor $v/c$ compared to the radiation pressure force, is in the toroidal
941: direction, opposes the Keplerian orbital motion $v_{K}=\sqrt{MG/r}$, and so
942: constitutes a drag force
943: \begin{equation}
944: F_{PR}=\ F_{G}\alpha\frac{v_{K}}{c}=\frac{\ Q_{rad}r_{d}^{2}}{4c^{2}\ \ }%
945: \sqrt{\frac{MGL^{2}}{r^{5}\ }}. \label{Poyntin-Roberston force}%
946: \end{equation}
947: The toroidal component of the magnetic force acting on a charged particle has
948: magnitude%
949: \begin{equation}
950: F_{mag}=Zev_{r}B \label{magnetic force}%
951: \end{equation}
952: where $Z$ is the charge. The grains typically have non-circular trajectories
953: with $v_{r}$ being of the order of the Kepler velocity $v_{K}$ so $F_{mag}\sim
954: Zev_{K}B.$ The ratio of Poynting-Robertson force to magnetic force is thus%
955: \begin{equation}
956: \frac{F_{PR}}{F_{mag}}=\ \frac{F_{G}}{F_{mag}}\alpha\frac{v_{K}}{c}.
957: \label{FPR to FG ratio}%
958: \end{equation}
959: Since the dust grains are assumed to be in a regime where they are acted on by
960: magnetic forces which are at least comparable to gravitational forces, i.e.,
961: $\ F_{mag}\gtrsim F_{G}$ and since $\alpha\ll1$ and $v_{K}/c\ll1$ it is seen
962: that the\ force due to Poynting-Robertson effect is negligible compared to
963: magnetic forces and so the Poynting-Robertson effect, like radiation pressure,
964: may be neglected.
965: 
966: \section{Electromagnetic particle Hamiltonian with gravity}
967: 
968: Hamilton-Lagrange methods are mathematically equivalent to the particle
969: equation of motion and so describe all physically allowed orbits (e.g.,
970: cyclotron, drift, Speiser, etc.). Furthermore, because the Hamilton-Lagrange
971: approach clarifies effects of spatial symmetries, deeper insight into orbital
972: dynamics is obtained than provided by direct integration of the equation of
973: motion. Direct integration nevertheless provides insight as well by providing
974: an independent verification of the predictions of Hamilton-Lagrange methods.
975: This two-pronged approach (Hamilton-Lagrange and direct orbit
976: integration)\ provides a powerful method for examining particle motion in
977: non-adiabatic situations.
978: 
979: The Lagrangian of a particle with mass $m_{\sigma}$ and charge $q_{\sigma}$ in
980: the combination of an axisymmetric electromagnetic field and the spherically
981: symmetric gravitational potential of a\ mass $M$ central object is
982: \begin{equation}%
983: \begin{array}
984: [c]{ccl}%
985: \mathcal{L} & = & \frac{m_{\sigma}}{2}\left(  v_{r}^{2}+r^{2}\dot{\phi}%
986: ^{2}+v_{z}^{2}\right) \\
987: &  & +q_{\sigma}\left(  r\dot{\phi}A_{\phi}(r,z,t)+v_{z}A_{z}(r,z,t)\right) \\
988: &  & -q_{\sigma}V(r,z,t)+\frac{m_{\sigma}MG}{\left(  r^{2}+z^{2}\right)
989: ^{1/2}}%
990: \end{array}
991: \label{Lagrangian}%
992: \end{equation}
993: where $V(r,z,t)$ is the electrostatic potential and a gauge with $A_{r}=0$ is
994: assumed. The canonical angular momentum is \
995: \begin{equation}
996: P_{\phi}\equiv\frac{\partial\mathcal{L}}{\partial\dot{\phi}}=m_{\sigma}%
997: r^{2}\dot{\phi}+q_{\sigma}rA_{\phi} \label{Pphi}%
998: \end{equation}
999: and, since $\ \mathbf{B}_{pol}=\nabla\times\left[  (2\pi r)^{-1}\psi\hat{\phi
1000: }\right]  =\nabla\times\left(  A_{\phi}\hat{\phi}\right)  $ implies $\psi=2\pi
1001: rA_{\phi}$, the canonical angular momentum can be expressed in terms of the
1002: poloidal magnetic flux as
1003: \begin{equation}
1004: P_{\phi}=m_{\sigma}r^{2}\dot{\phi}+\frac{q_{\sigma}}{2\pi}\psi(r,z,t).
1005: \label{Pphi flux}%
1006: \end{equation}
1007: Lagrange's equation $\dot{P}_{\phi}=\partial\mathcal{L}/\partial\phi
1008: \ $provides the important result that $\ $
1009: \begin{equation}
1010: P_{\phi}=const., \label{Pphi const}%
1011: \end{equation}
1012: i.e., $P_{\phi}$ is a constant of the motion because the system is
1013: axisymmetric. In the limit of a strong magnetic field, the second term in
1014: Eq.\ref{Pphi flux} dominates the first and leads to the constraint that a
1015: particle orbit must stay very nearly on a surface of constant $\psi;$ this is
1016: the basis for particle confinement in axisymmetric toroidal fusion devices
1017: (tokamaks, reversed field pinches, and spheromaks). Any deviation of a
1018: particle from a constant $\psi$ surface is a consequence of finite $m_{\sigma
1019: }$. When finite $m_{\sigma}$ is taken into account, it is seen that the
1020: particle must stay within a poloidal Larmor radius of a constant $\psi$
1021: surface, where poloidal Larmor radius means the cyclotron radius evaluated
1022: using the local poloidal field magnitude. Equation \ref{Pphi flux} may be
1023: solved for $\dot{\phi}$ to give
1024: \begin{equation}
1025: \dot{\phi}=\frac{P_{\phi}-\frac{q_{\sigma}}{2\pi}\psi(r,z,t)}{m_{\sigma}r^{2}%
1026: }. \label{solve phidot}%
1027: \end{equation}
1028: 
1029: 
1030: The corresponding Hamiltonian is
1031: \begin{equation}
1032: H=\frac{m_{\sigma}}{2}\left(  v_{r}^{2}+r^{2}\dot{\phi}^{2}+v_{z}^{2}\right)
1033: +q_{\sigma}V(r,z,t)-\frac{m_{\sigma}MG}{\left(  r^{2}+z^{2}\right)  ^{1/2}}.
1034: \label{Hamiltonian0}%
1035: \end{equation}
1036: By using Eq.\ref{solve phidot} to substitute for $\dot{\phi}\,\ $in
1037: Eq.\ref{Hamiltonian0}, the Hamiltonian can be expressed as%
1038: \begin{equation}%
1039: \begin{array}
1040: [c]{ccl}%
1041: H & = & \frac{m_{\sigma}}{2}\left(  v_{r}^{2}+v_{z}^{2}\right) \\
1042: &  & +\frac{\left(  P_{\phi}-\frac{q_{\sigma}}{2\pi}\psi(r,z,t)\right)  ^{2}%
1043: }{2m_{\sigma}r^{2}}\\
1044: &  & +q_{\sigma}V(r,z,t)-\frac{m_{\sigma}MG}{\left(  r^{2}+z^{2}\right)
1045: ^{1/2}}.
1046: \end{array}
1047: \label{Hamiltonian}%
1048: \end{equation}
1049: 
1050: 
1051: We now consider situations where $\psi$ is time-independent and $V=0\ $so the
1052: Hamiltonian reduces to
1053: \begin{equation}
1054: H=\frac{m_{\sigma}}{2}\left(  v_{r}^{2}+v_{z}^{2}\right)  +\frac{\left(
1055: P_{\phi}-\frac{q_{\sigma}}{2\pi}\psi(r,z)\right)  ^{2}}{2m_{\sigma}r^{2}%
1056: }-\frac{m_{\sigma}MG}{\sqrt{r^{2}+z^{2}}}. \label{H}%
1057: \end{equation}
1058: Since the Lagrangian does not explicitly depend on time, $H=const.$ and the
1059: particle energy is conserved. In the $q_{\sigma}\psi=0$ limit, $P_{\phi}$
1060: reduces to the mechanical angular momentum $p_{\phi}=mr^{2}\dot{\phi
1061: }=m\mathbf{r\times\dot{r}\cdot}\hat{z}=L\cos\theta$ in which case the dynamics
1062: reduces to the neutral particle orbital mechanics reviewed in Sec.\ref{Kepler}%
1063: . Thus, if $q_{\sigma}\psi=0$ bounded orbits correspond to $H<0$ and
1064: \ circular Kepler\ orbits correspond to $H$ having the minimum value of the
1065: effective potential well. Unbounded $q_{\sigma}\psi=0$ orbits correspond to
1066: $H\geq0$. Note that $p_{\phi}=L\cos\theta$ is a signed quantity, unlike $L.$
1067: 
1068: If $q_{\sigma}\psi$ is finite then $\left(  P_{\phi}-q_{\sigma}\psi
1069: (r,z)/2\pi\right)  ^{2}/2m_{\sigma}r^{2}$ is the appropriate term which
1070: contributes to the effective potential. This term, called the St\"{o}rmer
1071: potential, manifests \ a variety of qualitatively different spatial profiles
1072: depending on the relationship between $P_{\phi}$ and $q_{\sigma}\psi
1073: (r,z)/2\pi.$ These profiles are shown in
1074: Fig.\ref{effective-potential-scan.eps} for a sequence of decreasing values of
1075: $P_{\phi}.$ Very large positive $P_{\phi}$ gives prograde orbits similar to
1076: unmagnetized prograde cometary orbits and very large negative $P_{\phi}$ gives
1077: retrograde orbits similar to unmagnetized retrograde cometary orbits; in both
1078: these cases the strong centrifugal repulsion at small $r$ causes the particle
1079: to have an unbounded cometary orbit.
1080: 
1081: \_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_
1082: 
1083: Caption for Fig. \ref{effective-potential-scan.eps}
1084: 
1085: Left: Plot of $\psi(r,z)/\psi_{0}$ v. $r/a\,\ $for $z=0$ with sequence of
1086: values of $2\pi P_{\phi}/q\psi_{0}$ shown as dotted line. Right: Corresponding
1087: dependence of effective potential term $\left(  P_{\phi}-q\psi(r,z)/2\pi
1088: \right)  ^{2}/r^{2}$ showing that potential wells develop at locations where
1089: $2\pi P_{\phi}/q\psi_{0}$ intersects $\psi(r,z)/\psi_{0}.$ These wells
1090: correspond to cyclotron motion if the intersection is away from the maximum of
1091: $\psi$ and to Speiser orbits if the intersection is at or near the maximum of
1092: $\psi.$ A potential well at $r=0$ develops if $P_{\phi}=0$ as seen in the
1093: sixth set of plots from top; this results in drain-hole orbits. Dotted line in
1094: right-hand second plot from top has the vertical scale multiplied by 100 to
1095: enable visualization of the outer minimum and the dotted line on the third
1096: plot from the top has the vertical scale multiplied by 2000 times. Note
1097: changes of scale in right-hand plots.
1098: 
1099: \_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_\_
1100: 
1101: \begin{figure}[ptb]
1102: \caption{ caption on previous page}%
1103: \label{effective-potential-scan.eps}%
1104: \epsscale{0.48} \plotone{f4.eps}\end{figure}
1105: 
1106: \bigskip\pagebreak
1107: 
1108: On the other hand, if $P_{\phi}$ and $q_{\sigma}\psi(r,z)/2\pi$ have
1109: comparable magnitude, complex effective potential structures can result. For
1110: example if at some location $P_{\phi}$ $\ $equals $q_{\sigma}\psi(r,z)/2\pi$
1111: then $\left(  P_{\phi}-q_{\sigma}\psi(r,z)/2\pi\right)  ^{2}/2m_{\sigma}r^{2}$
1112: \ vanishes at this location, giving a \textit{localized minimum} in the
1113: overall effective potential.
1114: 
1115: If two separated positions exist where $P_{\phi}$ equals $q_{\sigma}%
1116: \psi(r,0)/2\pi$ then two distinct minima exist, but if only one position
1117: exists where $P_{\phi}$ equals $q_{\sigma}\psi(r,0)/2\pi$ then only one
1118: minimum exists. The former situation occurs when $P_{\phi}$ lies somewhere
1119: between $0$ and the maximum of $\psi$ and leads to cyclotron orbits with
1120: associated grad-$B$ $\ $and curvature drifts; in this case
1121: Eq.\ref{solve phidot} shows that the sign of $\dot{\phi}$ oscillates as the
1122: particle oscillates back and forth across the minimum of the effective
1123: potential, and the orbit is a cyclotron orbit. \ The situation of only one
1124: minimum occurs when the value of $P_{\phi}$ is approximately the maximum of
1125: $q_{\sigma}\psi/2\pi$ and gives Speiser orbits ($\dot{\phi}$ has fixed sign
1126: and the orbit is paramagnetic). For a review of the distinction between the
1127: diamagnetism of cyclotron orbits (and associated grad-$B$ $\ $and curvature
1128: drifts) and the paramagnetism of Speiser orbits see Appendix
1129: \ref{Orbit review}.
1130: 
1131: Yet another situation is where $P_{\phi}=0$. \ Because $\psi\sim r^{2}$ at
1132: small $r$, this special case removes the singularity of $\left(  P_{\phi
1133: }-q_{\sigma}\psi(r,0)/2\pi\right)  ^{2}/2m_{\sigma}r^{2}$ at $r=0$ and so
1134: eliminates centrifugal force repulsion altogether. The $P_{\phi}=0$ case gives
1135: trajectories which spiral down towards the central object\ while crossing
1136: magnetic field lines; part of the magnetic force cancels the centrifugal force
1137: so all that is left is gravity and a residual inward magnetic force. This
1138: situation is completely different from either cyclotron orbits or Speiser
1139: orbits and has only been previously discussed in the more limited context of
1140: two dimensional situations
1141: %TCIMACRO{\TeXButton{\citep{Bellan2007}}{\citep{Bellan2007}}}%
1142: %BeginExpansion
1143: \citep{Bellan2007}%
1144: %EndExpansion
1145: . Finally, there is also the special situation discussed by
1146: %TCIMACRO{\TeXButton{\citet{Schmidt1979}}{\citet{Schmidt1979}} }%
1147: %BeginExpansion
1148: \citet{Schmidt1979}
1149: %EndExpansion
1150: where $P_{\phi}=-r^{2}q_{\sigma}(2\pi)^{-1}\partial\left(  \psi/r\right)
1151: /\partial r$ in which case the charged particle executes an axis-encircling
1152: cyclotron orbit. While possible in principle, axis-encircling cyclotron orbits
1153: will be not be considered here because they would correspond to particles
1154: having extreme energies (e.g., a cyclotron radius of many a.u.).
1155: 
1156: Thus, there are five qualitatively distinct types of feasible trajectories
1157: depending on the relationship between $P_{\phi}$ and $q_{\sigma}\psi/2\pi.$ As
1158: labeled in Fig.\ref{effective-potential-scan.eps} and in order of descending
1159: signed value of the invariant $P_{\phi}\ $as shown by dashed horizontal line
1160: in left column of this figure, these are:
1161: 
1162: \begin{enumerate}
1163: \item prograde centrifugally dominated orbits [$P_{\phi}$ much larger than the
1164: peak of $q_{\sigma}\psi(r,z)/2\pi)$]
1165: 
1166: \item Speiser orbits [$P_{\phi}$ just grazes the peak of $q_{\sigma}%
1167: \psi(r,z)/2\pi$]
1168: 
1169: \item cyclotron orbits [$P_{\phi}$ well below the peak of $q_{\sigma}%
1170: \psi(r,z)/2\pi$ but much greater than zero]
1171: 
1172: \item $P_{\phi}=0$ orbits [which we will call \textquotedblleft drain-hole"
1173: orbits for reasons to be discussed later], and
1174: 
1175: \item retrograde centrifugally dominated orbits [$P_{\phi}$ negative and much
1176: less than zero)].
1177: \end{enumerate}
1178: 
1179: In the above list, we have removed the constraint that $z=0$ so, for example,
1180: in case \#3 (cyclotron orbits), the locations in the $r$-$z$ plane where
1181: $P_{\phi}$ and $q_{\sigma}\psi(r,z)/2\pi$ are equal corresponds to a specific
1182: closed curve in the $r$-$z$ plane; i.e., a specific $\psi$ iso-surface as
1183: shown in the projection of $\psi(r,z)$ at the top of Fig.
1184: \ref{GenericFluxFunction-new}.
1185: 
1186: The various possible values of $P_{\phi}$ can be considered as the different
1187: \textquotedblleft drummers\textquotedblright\ that dictate the effective
1188: potentials governing the motion of different particles located at the same
1189: position. A related example of this different \textquotedblleft
1190: drummers\textquotedblright\ situation \ has been reported by
1191: %TCIMACRO{\TeXButton{\citet{Tripathi2007}}{\citet{Tripathi2007}} }%
1192: %BeginExpansion
1193: \citet{Tripathi2007}
1194: %EndExpansion
1195: and involves two particles at the same location having velocities with equal
1196: magnitudes but opposite directions; the two particles have such extremely
1197: different effective potentials that one particle is expelled from a magnetic
1198: flux tube (hill-shaped effective potential) whereas the other remains in the
1199: flux tube (valley-shaped effective potential).
1200: 
1201: \subsection{Mechanism for accretion of collisionless particles}
1202: 
1203: Accretion is the process of converting unbounded orbits (i.e., cometary
1204: orbits)\ into bounded orbits. Accretion of a collisionless neutral particle is
1205: clearly impossible if such a particle is governed by dynamics of a
1206: time-independent Lagrangian because converting an unbounded orbit into a
1207: bounded orbit would require changing the particle energy $H$ and such a change
1208: is forbidden for a particle having a Lagrangian that does not explicitly
1209: depend on time.
1210: 
1211: We now postulate an accretion mechanism as follows:\ \textit{photo-emission
1212: acts as an effective switch which alters the form of the Hamiltonian equation
1213: governing particle dynamics}. The particle energy $H$ and mechanical angular
1214: momentum $mr^{2}\dot{\phi}$ \textit{do not change} during the switching, but
1215: after photo-emission has occurred, $H$ and $mr^{2}\dot{\phi}$ become
1216: parameters in a \textit{different Hamiltonian system} which has a different
1217: topography of potential barriers. For example, photo-emission can transform
1218: the neutral particle effective potential shown in
1219: Fig.\ref{Effective-Potential-Neutral-Particle}(a) into the\ charged particle
1220: effective potential shown in Fig.\ref{Effective-Potential-Neutral-Particle}(b).
1221: 
1222: The switching is postulated to occur when an incident neutral dust grain
1223: absorbs sufficient energetic photons from the star. The photon absorption
1224: causes the dust grain to photo-emit electrons and therefore become positively
1225: charged
1226: %TCIMACRO{\TeXButton{citep{Lee1996,Sickafoose2000}}{\citep
1227: %{Lee1996,Sickafoose2000}}}%
1228: %BeginExpansion
1229: \citep{Lee1996,Sickafoose2000}%
1230: %EndExpansion
1231: . The photo-emitted electrons become free electrons equal in number to the
1232: dust grain charge $Z$. The photo-emission process has effectively caused the
1233: initial neutral dust grain to disintegrate into a single heavy positively
1234: charged fragment (the charged dust grain) and $Z$ light negatively charged
1235: fragments (the photoelectrons). The motion of each fragment is governed by the
1236: Hamiltonian for a charged particle and this Hamiltonian is considerably
1237: different in form from the Hamiltonian that governed the neutral particle
1238: motion. \ 
1239: 
1240: The charge $q_{d}$ of a dust grain charged by photo-emission is given by
1241: \begin{equation}
1242: \frac{q_{d}}{4\pi\varepsilon_{0}r_{d}}\approx\ W_{photon}-W_{wf} \label{Z}%
1243: \end{equation}
1244: where $W_{photon}$ is the energy in eV of an incident photon that causes
1245: photo-emission of a primary photo-electron and $W_{wf}$ is the work-function
1246: in eV of the material
1247: %TCIMACRO{\TeXButton{\citep{Shukla2002}}{\citep{Shukla2002}}}%
1248: %BeginExpansion
1249: \citep{Shukla2002}%
1250: %EndExpansion
1251: .
1252: %TCIMACRO{\TeXButton{\citet{Lee1996}}{\citet{Lee1996}} }%
1253: %BeginExpansion
1254: \citet{Lee1996}
1255: %EndExpansion
1256: has shown that the effective photon energy is $W_{photon}\simeq8$ eV for
1257: nominal solar parameters and the effective work function of typical dust is
1258: $W_{wf}\simeq6$ eV so that the energy of emitted photo-electrons is $\sim2$ eV.
1259: 
1260: Combination of Eqs. \ref{md} and \ref{Z} show that the dust charge to mass
1261: ratio will be
1262: \begin{equation}
1263: \frac{q_{d}}{m_{d}}=\frac{\ 3\varepsilon_{0}\left(  W_{photon}-W_{wf}\right)
1264: }{\ \rho_{d}^{int}r_{d}^{2}}\ \label{charge to mass ratio}%
1265: \end{equation}
1266: which will be many of orders of magnitude smaller than the charge to mass
1267: ratios of electrons or ions. The number $Z$ of charges on a dust grain will
1268: be
1269: \begin{equation}
1270: Z=\frac{4\pi\varepsilon_{0}r_{d}\left(  W_{photon}-W_{wf}\right)  }{e}.
1271: \label{Znumber}%
1272: \end{equation}
1273: 
1274: 
1275: Charging a dust grain to $q_{d}$ takes a finite time interval, but for
1276: simplicity we will assume that this charging occurs at a single time defined
1277: as $t=0.$ Charging will not change either the instantaneous position or
1278: velocity of a particle.
1279: 
1280: Photo-emission at $t=0$ therefore decomposes an incident neutral dust grain
1281: into positive and negative product particles each inheriting the same position
1282: and velocity at $t=0_{+}$ that the neutral dust grain had at $t=0_{-}.$
1283: Position and velocity can consequently be considered to be continuous
1284: functions at $t=0$ so the canonical angular momentum with which a newly formed
1285: charged particle is endowed is%
1286: \begin{equation}%
1287: \begin{array}
1288: [c]{ccc}%
1289: P_{\phi} & = & m_{\sigma}r_{\ast}^{2}\dot{\phi}_{\ast}+q_{\sigma}\psi(r_{\ast
1290: },z_{\ast})/2\pi\\
1291: & = & L_{\sigma}\cos\theta\ +q_{\sigma}\psi(r_{\ast},z_{\ast})/2\pi
1292: \end{array}
1293: \label{Pphiborne}%
1294: \end{equation}
1295: where the subscript $\ast$ denotes the value of a coordinate at the instant of
1296: charging, i.e., at $t=0$. For simplicity we assume that photo-emission occurs
1297: when the distance between the incident neutral and the central object is at
1298: some critical spherical radius $R_{\ast}$ so that charging and the setting of
1299: $t=0$ occurs when the particle crosses the surface of the fictitious $R_{\ast
1300: }$ sphere, i.e., when $r=r_{\ast}$ and $z=z_{\ast}$ are such that
1301: \begin{equation}
1302: r_{\ast}^{2}+z_{\ast}^{2}=R_{\ast}^{2}. \label{ionization sphere}%
1303: \end{equation}
1304: 
1305: 
1306: Depending on the value of $L_{\sigma},$ the angle of inclination $\theta$, the
1307: magnitude of $q_{\sigma},$ and the value of $\psi(r_{\ast},z_{\ast}),$ all
1308: possible finite values of $P_{\phi}$ can occur, including positive, negative,
1309: and zero. The $r$-$z$ plane topography of the effective potential can change
1310: completely from what it was at $t=0_{-}$ because the centrifugal force
1311: potential $p_{\phi}^{2}/2mr^{2}$ responsible for the neutral particle
1312: potential barrier at small $r$ is replaced by the St\"{o}rmer term $(P_{\phi
1313: }-q_{\sigma}\psi(r,z))^{2}/2\pi mr^{2}$. $\ $Figure
1314: \ref{effective-potential-scan.eps}\ demonstrates that variation of particle
1315: mass and variation of the incoming orbit plane inclination angle $\theta$ $\ $
1316: results in a range of $p_{\phi}$, $q_{d}$ values and hence a range of
1317: $P_{\phi}$ values corresponding to cyclotron, \textquotedblleft
1318: drain-hole\textquotedblright, Speiser, or cometary orbits. If the new orbit is
1319: cyclotron, drain-hole, or Speiser, then photo-emission has prevented the
1320: particle from returning to infinity, i.e., the particle has accreted.
1321: Photo-emission changes the \textquotedblleft rules of the
1322: game\textquotedblright\ \ by effectively erecting a new potential
1323: barrier\ which traps a previously unbound particle. The \textquotedblleft old
1324: game\textquotedblright\ (i.e., neutral particle Keplerian motion as reviewed
1325: in Sec.\ref{Kepler}) did not depend on particle mass or $\theta,$ but the
1326: \textquotedblleft new game\textquotedblright\ does and leads to a mass- and
1327: $\theta$-dependent sorting of incoming charged grains and their associated
1328: photo-emitted electrons into qualitatively different classes of orbits.
1329: 
1330: This process whereby neutral particles enter a magnetic field from outside,
1331: become charged, and then become subject to magnetic forces is called `neutral
1332: beam injection' in the context of tokamak physics and `pickup' in the context
1333: of solar physics. Neutral beam injection is used routinely for tokamak heating
1334: and current drive
1335: %TCIMACRO{\TeXButton{citep{Simonen1988,Akers2002}}{\citep
1336: %{Simonen1988,Akers2002}}}%
1337: %BeginExpansion
1338: \citep{Simonen1988,Akers2002}%
1339: %EndExpansion
1340: . Pickup is important in the solar wind
1341: %TCIMACRO{\TeXButton{citep{Gloeckler2001}}{\citep{Gloeckler2001}}}%
1342: %BeginExpansion
1343: \citep{Gloeckler2001}%
1344: %EndExpansion
1345: , in planetary atmospheres
1346: %TCIMACRO{\TeXButton{citep{Hartle2006}}{\citep{Hartle2006}}}%
1347: %BeginExpansion
1348: \citep{Hartle2006}%
1349: %EndExpansion
1350: , and in producing source particles for comic rays
1351: %TCIMACRO{\TeXButton{citep{Ellison1998}}{\citep{Ellison1998}}}%
1352: %BeginExpansion
1353: \citep{Ellison1998}%
1354: %EndExpansion
1355: . However, to the best of the author's knowledge, charging of incoming neutral
1356: particles has not been previously proposed as a means for accreting matter
1357: around a star and instead accretion of matter around a star has always been
1358: argued to be the result of the \ viscosity of neutral particles, i.e.,
1359: collisions of neutral particles with each other, as discussed for example in
1360: %TCIMACRO{\TeXButton{citet{Lynden-Bell1974}}{\citet{Lynden-Bell1974}}}%
1361: %BeginExpansion
1362: \citet{Lynden-Bell1974}%
1363: %EndExpansion
1364: ,
1365: %TCIMACRO{\TeXButton{citet{Shakura1976}}{\citet{Shakura1976}}}%
1366: %BeginExpansion
1367: \citet{Shakura1976}%
1368: %EndExpansion
1369: , and
1370: %TCIMACRO{\TeXButton{citet{Pringle1981}}{\citet{Pringle1981}}}%
1371: %BeginExpansion
1372: \citet{Pringle1981}%
1373: %EndExpansion
1374: . The viscosity-based neutral particle accretion models suffer from not
1375: knowing what to do with the angular momentum of incident particles; this issue
1376: has motivated substantial work on developing the rather complicated non-linear
1377: turbulence-based magneto-rotational instability model as a means for
1378: transporting excess angular momentum outwards. In contrast, the model proposed
1379: here inherently accounts for angular momentum and\ so does not need any
1380: \textquotedblleft add-on turbulence\textquotedblright\ to transport angular
1381: momentum outwards.
1382: 
1383: Trapping via photo-emission has the remarkable feature that the special class
1384: of charged particles created with zero canonical angular momentum will spiral
1385: all the way down to the central object. \ These $P_{\phi}=0$ (drain-hole)
1386: particles falling towards $r=0$ are in what is effectively a loss cone in
1387: canonical angular momentum space. The drain-hole particle motion constitutes a
1388: gravity-driven dynamo
1389: %TCIMACRO{\TeXButton{\citep{Bellan2007}}{\citep{Bellan2007}} }%
1390: %BeginExpansion
1391: \citep{Bellan2007}
1392: %EndExpansion
1393: because the accumulation of these particles near $r=0$ produces a radially
1394: outward electric field while their flow produces a radially inward electric
1395: current (a dynamo is characterized by having opposed internal electric field
1396: and electric current). Since $J_{r}=-\left(  2\pi r\right)  ^{-1}\partial
1397: I/\partial z$ (see Eq.\ref{Jpol2}), creation of this radially inward current
1398: which is symmetric with respect to $z$ implies creation of an anti-symmetric
1399: function $I(r,z)$ which in turn implies creation of an anti-symmetric toroidal
1400: field $B_{\phi}\ $(see Eq.\ref{Bphi}). Equally remarkable, particles for which
1401: $P_{\phi}$ is near the maximum of $q_{\sigma}\psi/2\pi$ develop Speiser-type
1402: \textit{paramagnetic} orbits in the vicinity of $r=a$, $z=0$ and so can
1403: constitute the toroidal current that produces the poloidal flux (see
1404: discussion of Speiser orbit paramagnetism in Appendix \ref{Orbit review}). The
1405: creation of Speiser-orbit particles is \ conceptually similar to toroidal
1406: current drive in a tokamak via tangential neutral beam injection
1407: %TCIMACRO{\TeXButton{\citep{Simonen1988}}{\citep{Simonen1988}}}%
1408: %BeginExpansion
1409: \citep{Simonen1988}%
1410: %EndExpansion
1411: . Because the drain-hole and Speiser dynamos are both axisymmetric, both
1412: violate the essential claim of Cowling's anti-dynamo theorem
1413: %TCIMACRO{\TeXButton{\citep{Cowling1934}}{\citep{Cowling1934}} }%
1414: %BeginExpansion
1415: \citep{Cowling1934}
1416: %EndExpansion
1417: that axisymmetric dynamos cannot exist. This violation is not a problem
1418: because Cowling's theorem is based on MHD\ and so does not take into account
1419: drain-hole or Speiser orbits.
1420: 
1421: The Hamiltonian for an incoming neutral dust grain of mass $m_{n}$ can be
1422: written as
1423: \begin{equation}
1424: H\ =\frac{m_{n}v_{r}^{2}}{2}\ +\frac{m_{n}v_{\phi}^{2}}{2}+\frac{m_{n}%
1425: v_{z}^{2}}{2}-\frac{m_{n}MG}{\sqrt{r^{2}+z^{2}}}. \label{neutral}%
1426: \end{equation}
1427: This neutral dust grain absorbs energetic photons at $t=0,$ photo-emits $Z$
1428: free electrons, and consequently becomes positively charged with a charge of
1429: $Z.$ The mass of the neutral is related to the mass $m_{+}$ of the positively
1430: charged dust grain by $m_{n}=m_{+}+Zm_{e}$ where $m_{e}$ is the electron mass.
1431: Prior to this charging process, Eq.\ref{neutral} can be written as
1432: \begin{equation}%
1433: \begin{array}
1434: [c]{ccc}%
1435: H & = & \frac{\left(  m_{+}+Zm_{e}\right)  v_{r}^{2}}{2}\ +\frac{\left(
1436: m_{+}+Zm_{e}\right)  v_{\phi}^{2}}{2}\\
1437: &  & +\frac{\left(  m_{+}+Zm_{e}\right)  v_{z}^{2}}{2}-\frac{\left(
1438: m_{+}+Zm_{e}\right)  MG}{\sqrt{r^{2}+z^{2}}}%
1439: \end{array}
1440: \label{H before ionization}%
1441: \end{equation}
1442: so the positively charged dust grain with mass $m_{+}$ and the \thinspace$Z$
1443: electrons can each be thought of as executing identical neutral-type orbits
1444: before photo-emission occurs.
1445: 
1446: \bigskip\bigskip
1447: 
1448: At the instant before charging, the neutral particle mechanical angular
1449: momentum is
1450: \begin{equation}
1451: \ p_{\phi}=m_{n}r_{\ast}^{2}\dot{\phi}_{\ast}. \label{pmech}%
1452: \end{equation}
1453: At the instant after charging the newly created positively charged dust grain
1454: and its associated photo-emitted electrons all have the same values of
1455: $r_{\ast}$ and $\dot{\phi}_{\ast}.$ The canonical angular momentum of the
1456: positively charged dust grain will therefore be%
1457: \begin{equation}
1458: P_{\phi}^{+}=m_{+}r_{\ast}^{2}\dot{\phi}_{\ast}+Ze\psi(r_{\ast},z_{\ast}%
1459: )/2\pi\label{Pphi positive}%
1460: \end{equation}
1461: and the canonical momentum of each associated electron will be%
1462: \begin{equation}
1463: P_{\phi}^{e}=m_{e}r_{\ast}^{2}\dot{\phi}_{\ast}-e\psi(r_{\ast},z_{\ast})/2\pi.
1464: \label{Phi electron}%
1465: \end{equation}
1466: The initial neutral dust grain will be called the \textquotedblleft
1467: parent\textquotedblright\ particle while the positively charged dust grain
1468: resulting from photo-emission and its associated $Z~$photo-emitted electrons
1469: will be called \textquotedblleft child particles\textquotedblright\ that are
1470: \textquotedblleft siblings\textquotedblright\ of each other. The set of child
1471: particles resulting from the charging of a specific neutral dust grain will be
1472: called a \textquotedblleft family\textquotedblright. \ The canonical momenta
1473: $P_{\phi}^{+}$ and $P_{\phi}^{e}$ are now the appropriate orbit invariants for
1474: $t>0$ whereas \ the mechanical angular momenta $m_{\sigma}r^{2}\dot{\phi}$ of
1475: the individual siblings will \textit{not} be invariant for $t>0.$ Although the
1476: mechanical angular momentum of an individual sibling is not conserved, the
1477: total mechanical angular momentum of the family is conserved since summing
1478: Eq.\ref{Pphi positive} and $Z$ times Eq.\ref{Phi electron} gives $\left[
1479: p_{\phi}\right]  _{family}=P_{\phi}^{+}+ZP_{\phi}^{e}\ $where $\left[
1480: p_{\phi}\right]  _{family}$ $=m_{n}r_{\ast}^{2}\dot{\phi}$ is the sum of the
1481: mechanical angular momentum of the charged dust grain and all its sibling
1482: electrons. Because the siblings can physically separate from each other, the
1483: mechanical angular momentum $\left[  p_{\phi}\right]  _{family}$ is not a
1484: locally defined quantity and so is not a constant of the motion of either a
1485: single particle or, as in ideal hydrodynamics, of a fluid element. However,
1486: \textit{the mechanical angular momentum of the entire system is conserved}
1487: because the angular momentum of each family is globally conserved.
1488: 
1489: The kinetic energy of each sibling at the instant before photo-emission \ is
1490: the same as the value at the instant after photo-emission. If $H$ is
1491: decomposed into the contributions from the various \ siblings, it is seen that
1492: each sibling's $H_{\sigma}$ is the same before and after photo-emission.
1493: Assuming zero electrostatic potential $\ $for now, but allowing the child
1494: particle to be at arbitrary $z$, the Hamiltonian of each sibling is
1495: \begin{equation}%
1496: \begin{array}
1497: [c]{ccl}%
1498: H_{\sigma}\  & = & \frac{m_{\sigma}v_{r}^{2}}{2}\ +\frac{m_{\sigma}v_{z}^{2}%
1499: }{2}\\
1500: &  & +\frac{\left(  m_{\sigma}r_{\ast}^{2}\dot{\phi}_{\ast}+\frac{q_{\sigma}%
1501: }{2\pi}\left[  \psi(r_{\ast},z_{\ast})-\psi(r,z)\right]  \right)  ^{2}%
1502: }{2m_{\sigma}r^{2}}\\
1503: &  & -\frac{m_{\sigma}MG}{\sqrt{r^{2}+z^{2}}}.
1504: \end{array}
1505: \label{H-positive particle}%
1506: \end{equation}
1507: 
1508: 
1509: Using Eq.\ref{Bbar} we now define
1510: \begin{equation}
1511: \left\langle \omega_{c\sigma}\right\rangle =\frac{q_{\sigma}\left\langle
1512: B_{z}\right\rangle }{m_{\sigma}} \label{avg omegac}%
1513: \end{equation}
1514: as the spatially-averaged cyclotron frequency over the area bounded by the
1515: poloidal field magnetic axis. We now normalize all quantities to appropriate
1516: combinations of $a$ and $\Omega_{0},$ the Kepler angular frequency at $a$
1517: prescribed by Eq.\ref{Kepler0}. The normalized time, cylindrical coordinates,
1518: velocities, and magnetic flux are thus
1519: \begin{equation}%
1520: \begin{array}
1521: [c]{ccl}%
1522: \tau & = & \Omega_{0}t\\
1523: \bar{r} & = & r/a\\
1524: \bar{z} & = & z/a\\
1525: \bar{v}_{r} & = & v_{r}/a\Omega_{0}\\
1526: \bar{v}_{z} & = & v_{z}/a\Omega_{0}\\
1527: \bar{L} & = & L_{\sigma}/m_{\sigma}a^{2}\Omega_{0}\\
1528: \bar{p}_{\phi} & = & \bar{r}^{2}d\phi/d\tau=\bar{L}\cos\theta\\
1529: \bar{\psi}(r,z) & = & \frac{\psi(r,z)}{\psi(a,0)}\ =\frac{\psi(r,z)}%
1530: {\left\langle B_{z}\right\rangle \pi a^{2}}\\
1531: \bar{H} & = & \frac{H}{m\Omega_{0}^{2}a^{2}}%
1532: \end{array}
1533: \label{norm quantities}%
1534: \end{equation}
1535: in which case Eq.\ref{H-positive particle} becomes%
1536: \begin{equation}%
1537: \begin{array}
1538: [c]{ccl}%
1539: \bar{H} & = & \frac{\ \bar{v}_{r}^{2}}{\ 2}+\frac{\ \bar{v}_{z}^{2}}{\ 2}\\
1540: &  & +\frac{\left(  \bar{L}\cos\theta\ \ +\frac{\left\langle \omega_{c\sigma
1541: }\right\rangle }{2\Omega_{0}}\ \left[  \bar{\psi}(\bar{r}_{\ast},\bar{z}%
1542: _{\ast})-\bar{\psi}(\bar{r},\bar{z})\right]  \right)  ^{2}}{2\ \bar{r}^{2}}\\
1543: &  & -\ \frac{1}{\sqrt{\bar{r}^{2}+\bar{z}^{2}}}.
1544: \end{array}
1545: \label{norm H}%
1546: \end{equation}
1547: The effective potential is now
1548: \begin{equation}%
1549: \begin{array}
1550: [c]{cl}%
1551: \chi(\bar{r},\bar{z})= & \frac{\left(  \stackrel{\textnormal{mechanical}}%
1552: {\overbrace{\bar{L}\cos\theta}}\ \ +\stackrel{\textnormal{magnetic}}{\overbrace
1553: {\frac{\left\langle \omega_{c\sigma}\right\rangle }{2\Omega_{0}}\ \left[
1554: \bar{\psi}(\bar{r}_{\ast},\bar{z}_{\ast})-\bar{\psi}(\bar{r},\bar{z})\right]
1555: }}\right)  ^{2}}{2\ \bar{r}^{2}}\\
1556: & -\stackunder{\textnormal{gravitational}}{\underbrace{\frac{1}{\sqrt{\bar{r}%
1557: ^{2}+\bar{z}^{2}}}}}%
1558: \end{array}
1559: \label{effective potential magnetized}%
1560: \end{equation}
1561: where the mechanical, magnetic, and gravitational contributions have been
1562: labeled. Before photo-emission, the mechanical angular momentum is invariant
1563: so $\bar{p}_{\phi}(\bar{r}_{\ast},\bar{z}_{\ast})=\bar{L}\cos\theta$ is just
1564: the normalized mechanical angular momentum that the parent had when it was at
1565: infinity. Invoking Eq.\ref{p range}, it is seen that
1566: \begin{equation}
1567: 0\leq\bar{L}<\sqrt{2\bar{R}_{\ast}^{2}\bar{H}+2\bar{R}_{\ast}}%
1568: \ \label{pstar range}%
1569: \end{equation}
1570: is required since incident neutral dust grains with mechanical angular
1571: momentum outside this range would have reflected at larger radii than $\bar
1572: {R}_{\ast}$ and so would not have been able to access the radius $\bar
1573: {R}_{\ast}.$
1574: 
1575: The normalized canonical angular momentum with which a typical sibling charged
1576: particle is endowed is
1577: \begin{equation}
1578: \bar{P}_{\phi}=\bar{L}\cos\theta+\frac{\left\langle \omega_{c\sigma
1579: }\right\rangle }{2\Omega_{0}}\ \bar{\psi}(\bar{r}_{\ast},\bar{z}_{\ast}).
1580: \label{norm can momentum}%
1581: \end{equation}
1582: The parameters underlying Fig.\ref{Effective-Potential-Neutral-Particle}(b)
1583: can now be understood. This figure is a plot of $\chi(\bar{r},\bar{z})$ v.
1584: $\bar{r}$ for $\bar{z}=0\ $where $\bar{P}_{\phi}$ is calculated for the
1585: situation where $\left\langle \omega_{c\sigma}\right\rangle /\Omega_{0}=40,$
1586: $\bar{r}_{\ast}=0.5,$ $\bar{z}_{\ast}=0$, $\theta=0$, and $\bar{L}=1.$
1587: Charging of an $\ $ $\bar{L}=1$ dust grain via photo-emission causes the
1588: effective potential governing the dust grain to change from the form given in
1589: Fig.\ref{Effective-Potential-Neutral-Particle}(a) \ to the form given in
1590: Fig.\ref{Effective-Potential-Neutral-Particle}(b).
1591: 
1592: When the magnetic term in Eq.\ref{effective potential magnetized} becomes
1593: comparable to the mechanical term or much larger, orbital dynamics for the
1594: siblings become very different from the orbital dynamics of the neutral parent
1595: that existed before photo-emission. Various orbits can occur for the siblings.
1596: Because of the complexity of these three dimensional orbits, we will first
1597: consider orbits confined to the $\bar{z}=0$ plane and then generalize to fully
1598: three dimensional orbits ranging over finite $\bar{z}.$
1599: 
1600: \subsection{Distribution of Cometary, Speiser, Cyclotron, and Drain-Hole
1601: Orbits}
1602: 
1603: As reviewed in Sec.\ref{Kepler}, neutral particles orbits are degenerate with
1604: respect to their orbital plane inclination angle $\theta$ (see
1605: Fig.\ref{Orbital-Plane-coordinate-system}). However, once particles become
1606: charged, they are no longer restricted to an orbital plane, and furthermore,
1607: as seen from Eq.\ref{effective potential magnetized}, the effective potential
1608: of a charged particle has a strong dependence on the value of $\theta$ that
1609: its parent particle had. This dependence was manifested in the discussion of
1610: Fig. \ref{effective-potential-scan.eps} where it was noted that particles with
1611: $\left\vert 2\pi P_{\phi}/q_{\sigma}\psi_{0}\right\vert \gg1$ are essentially
1612: unmagnetized and have Keplerian cometary orbits, particles with $2\pi P_{\phi
1613: }\theta/q_{\sigma}\psi_{0}\simeq1$ have Speiser orbits, particles with
1614: $0\ll2\pi P_{\phi}/q_{\sigma}\psi_{0}\ll1$ have cyclotron orbits, and
1615: particles with $2\pi P_{\phi}/q_{\sigma}\psi_{0}\simeq0$ have drain-hole
1616: orbits. This discussion can be made more quantitative by defining
1617: $\Lambda\equiv2\pi P_{\phi}/q_{\sigma}\psi_{0};$ note that $\Lambda$
1618: corresponds to the horizontal dashed lines in the left hand column of
1619: Fig.\ref{effective-potential-scan.eps}. Equation \ref{norm can momentum} can
1620: then be recast as
1621: \begin{equation}
1622: \Lambda=\frac{2\Omega_{0}}{\left\langle \omega_{c\sigma}\right\rangle }\bar
1623: {L}\cos\theta\ +\ \ \bar{\psi}(\bar{r}_{\ast},\bar{z}_{\ast}). \label{lambda}%
1624: \end{equation}
1625: Thus particles with $\left\vert \Lambda\right\vert \gg1$ have Keplerian
1626: cometary orbits, particles with $\Lambda\simeq1$ have Speiser orbits,
1627: particles with $0\ll\Lambda\ll1$ have cyclotron orbits, and particles with
1628: $\Lambda\simeq0$ have drain-hole orbits.
1629: 
1630: Assuming $\bar{H}\ll1$ and $\bar{R}_{\ast}\sim1,$ Eq.\ref{pstar range} implies
1631: that only particles with $0<\bar{L}<\sqrt{2}$ can access a given location.
1632: Because there will be a distribution of all possible $\bar{L}$'s within this
1633: allowed range, we consider a particle with the mean of these allowed values as
1634: being representative and so assume that $\bar{L}=\sqrt{2}/2$ is the normalized
1635: angular momentum of this representative nominal particle.
1636: 
1637: Since $\left\langle \omega_{c\sigma}\right\rangle =q_{\sigma}\left\langle
1638: B_{z}\right\rangle /m_{\sigma}$, $\left\langle B_{z}\right\rangle =\psi
1639: _{0}/\pi a^{2}$, and $\Omega_{0}=\sqrt{MG/a^{3}}$ this nominal particle will
1640: have
1641: \begin{equation}
1642: \Lambda=K\cos\theta\ +\ \ \bar{\psi}(\bar{r}_{\ast},\bar{z}_{\ast})
1643: \label{def lambda}%
1644: \end{equation}
1645: where
1646: \begin{equation}
1647: K=\frac{m_{\sigma}}{q_{\sigma}}\frac{\ \pi\ \sqrt{2aMG\ }}{\psi_{0}%
1648: }\ \label{lambda rd}%
1649: \end{equation}
1650: parameterizes the competition between gravitational and magnetic forces. Using
1651: Eq. \ref{charge to mass ratio} to give the charge to mass ratio it is seen
1652: that
1653: \begin{equation}
1654: K=\ \frac{\ \pi\rho_{d}^{int}\sqrt{2aMG\ }}{3\varepsilon_{0}\left(
1655: W_{photon}-W_{wf}\right)  \psi_{0}}r_{d}^{2}; \label{K rd}%
1656: \end{equation}
1657: thus $K$ increases when $r_{d}$ increases as a result of dust grain coagulation.
1658: 
1659: Speiser and drain hole particles occur when gravitational and magnetic forces
1660: are comparable in magnitude, i.e., when $K$ is of order unity. For a given
1661: star mass $M,$ poloidal flux magnetic axis radius $a,$ and magnetic flux
1662: $\psi_{0},$ this means that Speiser and drain hole particles will occur when
1663: coagulation has caused the dust grains to have a certain critical radius which
1664: is of order%
1665: \begin{equation}
1666: r_{d}^{crit}\sim\frac{1}{\left(  2aMG\right)  ^{1/4}}\sqrt{\frac
1667: {3\varepsilon_{0}\left(  W_{photon}-W_{wf}\right)  \psi_{0}}{\pi\rho_{d}%
1668: ^{int}\ }}. \label{rdnom}%
1669: \end{equation}
1670: $\ $If $r_{d}\gg r_{d}^{crit}$ then gravity will dominate and the dust grains
1671: will behave like neutral particles whereas if $r_{d}\ll r_{d}^{crit}$ then
1672: magnetic forces will dominate and charged dust grains will mainly have
1673: cyclotron orbits. Since coagulation causes $r_{d}$ to increase monotonically,
1674: there should always be some time when $r_{d}$ $\sim r_{d}^{crit}$\ and $K$ is
1675: of order unity. This argument indicates that the dust-driven dynamo mechanism
1676: should take place as a well-defined temporal stage in the accretion process;
1677: before this stage $r_{d}$ is too small and after this stage $r_{d}$ is too large.
1678: 
1679: Since $\bar{\psi}(\bar{r}_{\ast},\bar{z}_{\ast})$ ranges between $0$ and $1$,
1680: let us consider the nominal situation where $\bar{\psi}(\bar{r}_{\ast},\bar
1681: {z}_{\ast})=1/2$ in which case%
1682: \begin{equation}
1683: \Lambda_{nom}=K\cos\theta\ +\frac{1}{2}. \label{lambda nom}%
1684: \end{equation}
1685: If $K\simeq1/2,$ Speiser particles result for $\cos\theta=1$ (i.e., neutral
1686: parent was prograde) and drain-hole particles result for $\cos\theta
1687: =-1\ $(i.e., neutral parent was retrograde ). If $K\ll1/2,$ the orbits will be
1688: cyclotron. Finally if $K\gg1/2,$ the orbits will be cometary if $\cos\theta$
1689: is not close to zero. The categorization implied by Eq.\ref{lambda nom} is
1690: shown schematically in Fig. \ref{orbit-distributionAI-PS-DCS2.0-Ascii5-25.eps}.
1691: 
1692: \begin{figure}[ptb]
1693: \caption{Distribution of orbits as function of $K,\theta.$ Radius $K$ is
1694: proportional to $r_{d}^{2}.$ Prograde orbits have $\theta=0$, retrograde
1695: orbits have $\theta=\pi$, and polar orbits have $\left\vert \theta\right\vert
1696: =\pi/2$.}%
1697: \label{orbit-distributionAI-PS-DCS2.0-Ascii5-25.eps}%
1698: \epsscale{1.0} \plotone{f5.eps}\end{figure}
1699: 
1700: \subsection{Light-weight particles ($K\ll1$) \label{light-weight trapping}}
1701: 
1702: \subsubsection{Generic accretion mechanism}
1703: 
1704: \begin{figure}[ptb]
1705: \caption{An incident neutral particle ($\bar{H}=0,$ $\bar{\rho}_{pericenter}%
1706: =0.1,$ $\theta=30,\alpha=0^{0}$) becomes charged due to photoemission of
1707: electrons at $R_{\ast}=2.$ The child particle mass is such that $\left\langle
1708: \omega_{c\sigma}\right\rangle /\Omega_{0}=200$ and the child particle becomes
1709: magnetically trapped, staying within a poloidal Larmor orbit of a constant
1710: $\psi$ surface. In this example, the child particle is mirror trapped and so
1711: cannot enter the strong magnetic field region at small $\bar{r}.$ (a) $\bar
1712: {x}-\bar{y}$ plane, charging occurs where orbit abruptly changes, poloidal
1713: field magnetic axis shown as dashed circle (b)\ $\bar{r}-\bar{z}$ plane
1714: showing magnetic mirroring of child particle at large magnetic field (poloidal
1715: flux contours $\bar{\psi}(\bar{r},\bar{z})$ shown as dashed lines). The orbit
1716: the parent neutral particle would have continued to have if it had not become
1717: charged is shown by dotted line (both projections).}%
1718: \label{mirror..pericenter=0.1_theta=30_alpha=0_rho0=4_rstar=2_omegaratio=200.eps}%
1719: \epsscale{0.5} \plotone{f6.eps}\end{figure}
1720: 
1721: Potential barriers occur at locations where $\chi(\bar{r},\bar{z})>$ $\bar
1722: {H}.$ We first consider motion of a sibling particle constrained to stay in
1723: the $\bar{z}=0$ plane (i.e., the particle starts with $\bar{v}_{z}=0$ and no
1724: forces exist that push it off the $\bar{z}=0$ plane). In this case the
1725: effective potential is a function of $\bar{r}$ only and is%
1726: \begin{equation}
1727: \chi(\bar{r},0)=\frac{\left(  \stackrel{\textnormal{mechanical}}{\overbrace{\bar
1728: {L}\cos\theta}}\ \ +\stackrel{\textnormal{magnetic}}{\overbrace{\frac{\left\langle
1729: \omega_{c\sigma}\right\rangle }{2\Omega_{0}}\ \left[  \bar{\psi}(\bar{r}%
1730: _{\ast},0)-\bar{\psi}(\bar{r},0)\right]  }}\right)  ^{2}}{2\ \bar{r}^{2}%
1731: }-\stackrel{\textnormal{gravitational}}{\overbrace{\frac{1}{\bar{r}}}}.
1732: \label{chi z plane}%
1733: \end{equation}
1734: If $\chi(\bar{r},0)$ exceeds $\bar{H}$ at some radius $\bar{r}>$ $\bar
1735: {r}_{\ast}$, the particle becomes trapped within a finite extent region as
1736: indicated in Fig.\ref{Effective-Potential-Neutral-Particle}(b) or equivalently
1737: by the third and fourth rows, right hand column of Fig.
1738: \ref{effective-potential-scan.eps}. \ If $\bar{r}_{\ast}$ is small or large
1739: compared to unity so $r_{\ast}$ is not near the peak of $\bar{\psi},$ then
1740: $\bar{\psi}(\bar{r}_{\ast},0)\ll1.$ Because the particle is assumed to be
1741: light-weight (i.e., \thinspace$r_{d}\ll r_{d}^{crit}$), its average-field
1742: cyclotron frequency $\left\langle \omega_{c\sigma}\right\rangle $ will be much
1743: larger than the Kepler frequency $\Omega_{0}.$ Since $\bar{L}$ is of order
1744: unity, the light-weight particle will have $\left\vert \left\langle
1745: \omega_{c\sigma}\right\rangle /2\Omega_{0}\right\vert \gg\bar{L}$ in which
1746: case%
1747: \[
1748: \max\left\{  \frac{\ \left(  \bar{L}\cos\theta+\frac{\left\langle
1749: \omega_{c\sigma}\right\rangle \ }{2\Omega_{0}}\left[  \bar{\psi}(\bar{r}%
1750: _{\ast},0)-\bar{\psi}(\bar{r},0)\right]  \right)  ^{2}}{2\bar{r}^{2}%
1751: \ }\right\}  \simeq\max\left\{  \frac{1}{2\bar{r}^{2}\ }\left(  \frac
1752: {\left\langle \omega_{c\sigma}\right\rangle \ }{2\Omega_{0}}\bar{\psi}(\bar
1753: {r},0)\right)  ^{2}\right\}
1754: \]
1755: so the peak of $\chi(\bar{r},0)$ will occur where $\bar{\psi}(\bar{r},0)$
1756: takes on its maximum value, namely unity. The maximum of $\chi(\bar{r},0)$ for
1757: a light-weight particle is thus
1758: \begin{equation}
1759: \chi(1,0)\simeq\frac{1}{2}\left(  \frac{\left\langle \omega_{c\sigma
1760: }\right\rangle }{2\Omega_{0}}\right)  ^{2}-1 \label{generic barrier}%
1761: \end{equation}
1762: where \ the $-1$ term comes from the gravitational potential at $\bar
1763: {r}=1,z=0.$ This gives the necessary condition for trapping light-weight
1764: charged dust grains to be
1765: \begin{equation}
1766: \frac{\left\vert \left\langle \omega_{c\sigma}\right\rangle \right\vert
1767: }{\ \Omega_{0}}>2\sqrt{2}. \label{trapping}%
1768: \end{equation}
1769: Because a typical light-weight grain has $\left\vert \left\langle
1770: \omega_{c\sigma}\right\rangle \right\vert \gg2\sqrt{2}\Omega_{0},$ a
1771: light-weight grain confined to the $z=0$ plane will become trapped in a
1772: finite-sized region upon being charged, i.e., it will have accreted. The same
1773: will be true for the associated sibling photo-electrons since they also have
1774: $\left\vert \left\langle \omega_{ce}\right\rangle \right\vert \gg2\sqrt
1775: {2}\Omega_{0}.$ Figure
1776: \ref{mirror..pericenter=0.1_theta=30_alpha=0_rho0=4_rstar=2_omegaratio=200.eps}%
1777: \ shows a direct numerical integration of the equation of motion demonstrating
1778: this basic accretion mechanism in three dimensions: a light-weight neutral
1779: dust grain disintegrates at a certain location into an $\left\vert
1780: \left\langle \omega_{c\sigma}\right\rangle \right\vert \gg2\sqrt{2}\Omega_{0}$
1781: positively charged dust grain \ (there would also be $Z$ associated
1782: photo-electrons which for clarity are not shown in the figure but would also
1783: have cyclotron-type orbits). An actual dust grain would start with an
1784: infinitesimal energy $0<\bar{H}\ll1;$ the calculation here uses $\bar
1785: {H}=0\,\ $as representative of this infinitesimal $\bar{H}$ since the
1786: difference between an orbit with $\bar{H}=0$ and an orbit with infinitesimal
1787: $\bar{H}$ is insignificant at any finite distance. The newly created charged
1788: particles are trapped to the vicinity of a $\bar{\psi}(\bar{r},\bar
1789: {z})=const.$ poloidal flux surface (poloidal flux surfaces are shown by dashed
1790: lines in
1791: Fig.\ref{mirror..pericenter=0.1_theta=30_alpha=0_rho0=4_rstar=2_omegaratio=200.eps}%
1792: (b)). Because of $\mu$ conservation, the charged particles can also be
1793: mirror-trapped, so while on the the constant $\bar{\psi}$ surface, they
1794: reflect from regions of this surface where the magnetic field is strong.
1795: Figure
1796: \ref{mirror..pericenter=0.1_theta=30_alpha=0_rho0=4_rstar=2_omegaratio=200momenta.eps}%
1797: (a) plots the time dependence of $\bar{p}_{\phi}$ for the particle shown in
1798: Fig.\ref{mirror..pericenter=0.1_theta=30_alpha=0_rho0=4_rstar=2_omegaratio=200.eps}%
1799: .
1800: 
1801: Figure
1802: \ref{mirror..pericenter=0.1_theta=30_alpha=0_rho0=4_rstar=2_omegaratio=200momenta.eps}%
1803: (b) plots the canonical angular momentum $\bar{P}_{\phi}$ (solid line) and the
1804: kinetic/potential energies (dashed lines labeled `KE' and `PE' in figure). It
1805: is seen that $\bar{p}_{\phi}$ is conserved before charging whereas $\bar
1806: {P}_{\phi}$ is the conserved quantity after charging. Also, the total energy
1807: (kinetic + potential, dashed line labeled `Tot' in figure) remains zero.
1808: Strictly speaking, this plot should be considered as referring to the neutral
1809: dust grain until charging, and then to the charged dust grain after charging
1810: so the jump in $\bar{P}_{\phi}$ at the charging time seen in the figure does
1811: not violate the requirement that $\bar{P}_{\phi}$ is a constant of the motion
1812: for\ a specific particle.
1813: 
1814: Figure
1815: \ref{notmirror..pericenter=0.1_theta=30_alpha=0_rho0=4_rstar=0.8_omegaratio=200.eps}
1816: shows the three dimensional orbit of a light-weight charged dust grain with
1817: slightly different parameters so that it is not mirror trapped. The derivation
1818: of the non-dimensional equation of motion used here is given in Appendix
1819: \ref{Equation of motion}. \begin{figure}[ptb]
1820: \caption{(a)\ Mechanical angular momentum $\bar{p}_{\phi}$ v. time $\tau$ and
1821: (b)\ kinetic energy (KE), potential energy (PE) and canonical angular momentum
1822: $\bar{P}_{\phi}$ v. time for the calculation shown in Fig.
1823: \ref{mirror..pericenter=0.1_theta=30_alpha=0_rho0=4_rstar=2_omegaratio=200.eps}%
1824: . Mechanical angular momentum $\bar{p}_{\phi}\ \ $ is conserved before
1825: charging, but oscillates after charging; canonical angular momentum $\bar
1826: {P}_{\phi}$ is much larger than mechanical angular momentum because of strong
1827: magnetic field and is conserved after charging. }%
1828: \label{mirror..pericenter=0.1_theta=30_alpha=0_rho0=4_rstar=2_omegaratio=200momenta.eps}%
1829: \epsscale{0.5} \plotone{f7.eps}\end{figure}\begin{figure}[ptbptb]
1830: \caption{Same parameters as Fig.
1831: \ref{mirror..pericenter=0.1_theta=30_alpha=0_rho0=4_rstar=2_omegaratio=200.eps}%
1832: , except $R_{\ast}=0.8.$ Charged particle is now not mirror-trapped.}%
1833: \label{notmirror..pericenter=0.1_theta=30_alpha=0_rho0=4_rstar=0.8_omegaratio=200.eps}%
1834: \epsscale{0.5} \plotone{f8.eps}\end{figure}
1835: 
1836: \subsubsection{Width of light-weight particle trapping well and relation to
1837: cyclotron orbits \label{trapping well width}}
1838: 
1839: If $\left\vert \left\langle \omega_{c\sigma}\right\rangle \right\vert
1840: /2\Omega_{0}\gg\ \left\vert \bar{p}_{\phi}(\bar{r}_{\ast},\bar{z}_{\ast
1841: })\right\vert $, the magnetic term in Eq.\ref{effective potential magnetized}
1842: dominates the mechanical term as soon as $\bar{r}$ deviates slightly from
1843: $r_{\ast}.$ This implies existence of a narrow trench-like potential
1844: well\ with minimum very close to $r_{\ast}.$ The effective potential shown
1845: Fig.\ref{Effective-Potential-Neutral-Particle}(b) has such a trench; this
1846: situation involves a particle confined to the $\bar{z}=0$ plane and the trench
1847: is at $\bar{r}=\bar{\rho}=0.55.$ This situation is also evident in the third
1848: and fourth rows, right hand column of Fig. \ref{effective-potential-scan.eps}.
1849: If $\left\vert \left\langle \omega_{c\sigma}\right\rangle \right\vert
1850: \gg\Omega_{0}$ the gravitational term is completely overwhelmed by the
1851: magnetic term so the trench bottom in the $z=0$ plane is where
1852: \begin{equation}
1853: \ \bar{\psi}(\bar{r})=\frac{\Omega_{0}}{\left\langle \omega_{c\sigma
1854: }\right\rangle }\bar{p}_{\phi}(\bar{r}_{\ast},\bar{z}_{\ast})\ +\bar{\psi
1855: }(r_{\ast}).\ \label{well bottom}%
1856: \end{equation}
1857: Taylor expansion of $\ \bar{\psi}(\bar{r})$ near $r_{\ast}$ gives
1858: \begin{equation}%
1859: \begin{array}
1860: [c]{cc}%
1861: \ \bar{\psi}(\bar{r})= & \bar{\psi}(\bar{r}_{\ast})+\ \left(  \bar{r}-\bar
1862: {r}_{\ast}\right)  \left(  \frac{\partial\bar{\psi}}{\partial\bar{r}}\right)
1863: _{\bar{r}=\bar{r}_{\ast}}\\
1864: & +\frac{1}{2}\left(  \bar{r}-\bar{r}_{\ast}\right)  ^{2}\left(
1865: \frac{\partial^{2}\bar{\psi}}{\partial\bar{r}^{2}}\right)  _{\bar{r}=\bar
1866: {r}_{\ast}}+...
1867: \end{array}
1868: \label{Taylor}%
1869: \end{equation}
1870: If $\bar{r}_{\ast}$ is not close to unity, then $\bar{\psi}(\bar{r}_{\ast})$
1871: is not close to its maximum value so the leading term in the Taylor expansion
1872: is the one involving $\partial\bar{\psi}/\partial\bar{r}.$ Using
1873: Eq.\ref{Taylor} to substitute for $\ \bar{\psi}(\bar{r})$ in
1874: Eq.\ref{well bottom} gives the trench bottom to be at%
1875: \begin{equation}
1876: \ \ \bar{r}\ \ =\bar{r}_{\ast}+\frac{2}{\left(  \partial\bar{\psi}%
1877: /\partial\bar{r}\right)  _{\bar{r}=\bar{r}_{\ast}}}\frac{\Omega_{0}%
1878: }{\left\langle \omega_{c\sigma}\right\rangle }\bar{p}_{\phi}(\bar{r}_{\ast
1879: },\bar{z}_{\ast})\ \label{more well bottom}%
1880: \end{equation}
1881: so the trench bottom is close to $\bar{r}_{\ast}$ because $\left\vert
1882: \left\langle \omega_{c\sigma}\right\rangle \right\vert \gg\Omega_{0}$ is being
1883: assumed. If $\bar{r}_{\ast}<1$ then $\left(  \partial\bar{\psi}/\partial
1884: \bar{r}\right)  _{\bar{r}=\bar{r}_{\ast}}$ is positive and vice versa since
1885: $\bar{\psi}$ has its maximum value at $\bar{r}=1.$ The trench bottom will thus
1886: be outside of $\bar{r}_{\ast}$ if $\bar{r}_{\ast}<1$ so $\left\langle
1887: \omega_{c\sigma}\right\rangle \left(  \partial\bar{\psi}/\partial\bar
1888: {r}\right)  _{\bar{r}=\bar{r}_{\ast}}$is positive and vice versa if $\bar
1889: {r}_{\ast}>1$. The sign of $\ \bar{\psi}(\bar{r})-\bar{\psi}(\bar{r}_{\ast})$
1890: oscillates as the particle bounces back and forth in the trench. Using
1891: Eq.\ref{solve phidot} expressed in normalized variables, \ and noting
1892: that\ $\bar{P}_{\phi}=\bar{p}_{\phi}(\bar{r}_{\ast},\bar{z}_{\ast})+\bar{\psi
1893: }(\bar{r}_{\ast},\bar{z}_{\ast})\left\langle \omega_{c\sigma}\right\rangle
1894: /2\Omega_{0}\ \ $it is seen that the azimuthal velocity
1895: \begin{equation}%
1896: \begin{array}
1897: [c]{ccl}%
1898: \frac{d\phi}{d\tau} & = & \frac{\bar{P}_{\phi}\ -\frac{\left\langle
1899: \omega_{c\sigma}\right\rangle }{2\Omega_{0}}\bar{\psi}(\bar{r})}{\bar{r}^{2}%
1900: }\\
1901: & = & \frac{\bar{p}_{\phi}(\bar{r}_{\ast},\bar{z}_{\ast})\ -\frac{\left\langle
1902: \omega_{c\sigma}\right\rangle }{2\Omega_{0}}\left[  \bar{\psi}(\bar{r}%
1903: )-\bar{\psi}(\bar{r}_{\ast})\right]  }{\bar{r}^{2}}%
1904: \end{array}
1905: \label{phidot}%
1906: \end{equation}
1907: has an oscillating polarity. The combined oscillation of $\bar{r}$ and
1908: \ $d\phi/d\tau$ corresponds to the particle tracing out Larmor orbits with
1909: gyro-center at the trench bottom
1910: %TCIMACRO{\TeXButton{citep{Schmidt1979}}{\citep{Schmidt1979}}}%
1911: %BeginExpansion
1912: \citep{Schmidt1979}%
1913: %EndExpansion
1914: .
1915: 
1916: To summarize: In the $\left\vert \left\langle \omega_{c\sigma}\right\rangle
1917: \right\vert /\Omega_{0}\gg1$ situation (i.e., light-weight particles) Eq.
1918: \ref{effective potential magnetized} provides an effective potential whereby
1919: the magnetized charged particle is confined to the vicinity of a constant
1920: $\psi$ surface, just like a charged particle in a tokamak
1921: %TCIMACRO{\TeXButton{citep{Rome1979}}{\citep{Rome1979}}}%
1922: %BeginExpansion
1923: \citep{Rome1979}%
1924: %EndExpansion
1925: . The particle motion over the constant $\psi$ surface can be understood as a
1926: sum of parallel to $\mathbf{B}$ motion, cyclotron motion, and particle drifts
1927: (curvature, grad $B,$ etc.) as given by Eq.\ref{particle drifts}. Furthermore,
1928: regions where $\mu B$ \ is large can constitute an additional potential
1929: barrier (i.e., magnetic mirror) that prevents those subsets of particles
1930: having inadequate velocity parallel to $\mathbf{B}$\textbf{ }from accessing
1931: the entire constant $\psi$ surface. Because of magnetic mirroring by
1932: $\mu\nabla B$ forces, particles in these subsets are confined to the weaker
1933: magnetic field regions of a constant $\psi$ surface as seen in
1934: Fig.\ref{mirror..pericenter=0.1_theta=30_alpha=0_rho0=4_rstar=2_omegaratio=200.eps}%
1935: (b).
1936: 
1937: \subsection{Speiser orbit particles ($K\cos\theta\simeq1/2$)}
1938: 
1939: When $\bar{r}_{\ast}\simeq1$ and $\bar{z}_{\ast}\simeq0$, photoemission occurs
1940: near the peak of $\psi(\bar{r},\bar{z})$, i.e., where $\nabla\psi$ $\simeq0$;
1941: see Fig. \ref{effective-potential-scan.eps} second row from top where
1942: $P_{\phi}$ is just grazing the peak of $\psi.$ The linear term in the Taylor
1943: expansion in Eq.\ref{Taylor} is therefore negligible. Since $\partial^{2}%
1944: \bar{\psi}/\partial\bar{r}^{2}$ is negative near the maximum of $\bar{\psi},$
1945: Eq.\ref{Taylor} becomes
1946: \begin{equation}
1947: \bar{\psi}(\bar{r})\simeq\bar{\psi}(\bar{r}_{\ast})-\ \frac{1}{2}\left(
1948: \bar{r}-\bar{r}_{\ast}\right)  ^{2}\left\vert \left(  \frac{\partial^{2}%
1949: \bar{\psi}}{\partial\bar{r}^{2}}\right)  _{\bar{r}=\bar{r}_{\ast}}\right\vert
1950: . \label{chi Speiser}%
1951: \end{equation}
1952: Equation \ref{phidot} then reduces to%
1953: \begin{equation}
1954: \frac{d\phi}{d\tau}\ =\frac{\bar{p}_{\phi}(\bar{r}_{\ast},0)\ +\frac
1955: {\left\langle \omega_{c\sigma}\right\rangle }{4\Omega_{0}}\ \left(  \bar
1956: {r}-\bar{r}_{\ast}\right)  ^{2}\left\vert \left(  \frac{\partial^{2}\bar{\psi
1957: }}{\partial\bar{r}^{2}}\right)  _{\bar{r}=\bar{r}_{\ast}}\right\vert }{\bar
1958: {r}^{2}} \label{preSpeiser}%
1959: \end{equation}
1960: and for $\bar{p}_{\phi}(\bar{r}_{\ast},0)\ $being positive (i.e., parent
1961: particle was prograde), $d\phi/d\tau$ is always positive. This corresponds to
1962: Speiser-type orbits because when the particles bounce back and forth across
1963: the peak of $\psi$, they are bouncing back and forth between regions where the
1964: poloidal magnetic field $\sim\partial\psi/\partial r$ changes sign. As
1965: discussed in Section \ref{Speiser} of Appendix \ref{Orbit review}, this
1966: results in paramagnetism, i.e., positively charged particles moving in the
1967: positive $\phi$ direction and so \textit{producing} rather than opposing a
1968: $B_{z}$ field. Creation of Speiser-orbiting particles sustains the poloidal
1969: magnetic field against losses and will amplify an initial seed poloidal field;
1970: creation of Speiser particles therefore constitutes a dynamo for driving
1971: toroidal current.
1972: 
1973: Figure
1974: \ref{speiser_pericenter=0.95_theta=18_alpha=0_rho0=4_rstar=1.2_omegaratio=10.eps}
1975: shows the creation of a Speiser orbit by photo-emission charging of a neutral
1976: particle near the poloidal field magnetic axis. Figure
1977: \ref{speiser_pericenter=0.95_theta=18_alpha=0_rho0=4_rstar=1.2_omegaratio=10.eps}%
1978: (a) shows that the orbit is paramagnetic (i.e., particle moves in positive
1979: $\phi$ direction) while Fig.
1980: \ref{speiser_pericenter=0.95_theta=18_alpha=0_rho0=4_rstar=1.2_omegaratio=10.eps}%
1981: (b) shows that the orbit involves repeated reflection from the interior of a
1982: poloidal flux surface in the manner discussed in Section \ref{Speiser} of
1983: Appendix \ref{Orbit review}.\begin{figure}[ptb]
1984: \caption{Speiser orbit resulting from parent with $\bar{H}=0$, $\bar{\rho
1985: }_{pericenter}=0.95,$ $\theta=18^{0},$ $\alpha=0^{0}.$ Charging occurs at
1986: $\bar{R}_{\ast}=1.2$ and the child particle is a positive particle with
1987: $\omega_{c\sigma}/\Omega_{0}=10;\ $ (a) the $\bar{x}$-$\bar{y}$ plane orbit is
1988: counter-clockwise corresponding to paramagnetic motion; (b) the $\bar{r}%
1989: $-$\bar{z}$ plane orbit involves the particle continuously reflecting from the
1990: interior of a toroidal flux tube.}%
1991: \label{speiser_pericenter=0.95_theta=18_alpha=0_rho0=4_rstar=1.2_omegaratio=10.eps}%
1992: \epsscale{0.5} \plotone{f9.eps}\end{figure}
1993: 
1994: \subsection{Drain-hole particles ($K\cos\theta\simeq-1/2$)}
1995: 
1996: Charged particles born with $\bar{P}_{\phi}=0\ $are called `drain-hole'
1997: particles because they behave as if they are going down a drain. The
1998: properties of drain-hole particles restricted to the $z=0$ plane were briefly
1999: examined in
2000: %TCIMACRO{\TeXButton{\citet{Bellan2007}}{\citet{Bellan2007}}}%
2001: %BeginExpansion
2002: \citet{Bellan2007}%
2003: %EndExpansion
2004: ; here the more general 3D situation will be considered. Using
2005: Eq.\ref{norm can momentum} and $\bar{p}_{\phi}(\bar{r}_{\ast},\bar{z}_{\ast
2006: })=\bar{L}\cos\theta$ the $\bar{P}_{\phi}=0$ condition corresponds to
2007: \begin{equation}
2008: \bar{L}\cos\theta=-\frac{\left\langle \omega_{c\sigma}\right\rangle }%
2009: {2\Omega_{0}}\ \bar{\psi}(\bar{r}_{\ast},\bar{z}_{\ast}%
2010: )\ \label{drain-perigee}%
2011: \end{equation}
2012: implying that $\cos\theta$ is negative in which case the parent particle must
2013: have been retrograde. The effective potential (see
2014: Eq.\ref{effective potential magnetized}) for the $\bar{P}_{\phi}=0$ class of
2015: particles reduces to
2016: \begin{equation}
2017: \chi(\bar{r},\bar{z})\simeq\frac{\left\langle \omega_{c\sigma}\right\rangle
2018: ^{2}}{8\Omega_{0}^{2}}\frac{\left(  \bar{\psi}(\bar{r},\bar{z})\right)  ^{2}%
2019: }{\ \bar{r}^{2}}\ -\ \frac{1}{\sqrt{\bar{r}^{2}+\bar{z}^{2}}}
2020: \label{chi free fall}%
2021: \end{equation}
2022: which has a funnel (i.e., drain-like) profile near $\bar{r}=0,$ $\bar{z}=0$
2023: due to the second (gravitational) term and a hill on the funnel side wall with
2024: peak near $\bar{r}=\bar{a},$ $\bar{z}=0$ due to the first (St\"{o}rmer) term.
2025: A particle initially on the hill (i.e., near the poloidal field magnetic axis)
2026: will fall down the hill into the drain-like funnel; see right hand column,
2027: second row from bottom in Fig. \ref{effective-potential-scan.eps} for plot of
2028: first term in Eq.\ref{chi free fall}. Thus, no matter where a $\bar{P}_{\phi
2029: }=0$ particle starts in $\bar{r},\bar{z}$ space, it eventually follows a
2030: spiral path down to\ $\bar{r}=0,\bar{z}=0;$ no centrifugal force will ever
2031: push it back outwards because the first term in Eq.\ref{chi free fall} has no
2032: singularity at $\bar{r}=0$ (recall that $\bar{\psi}\sim\bar{r}^{2}$ for small
2033: $\bar{r},\bar{z}$). The sense of this downward spiraling trajectory will be in
2034: the $-\left\langle \omega_{c\sigma}\right\rangle $ direction as shown by
2035: Eq.\ref{phidot}. Since $\bar{\psi}\sim\bar{r}^{2}$ for small $\bar{r},\bar{z}$
2036: it is seen from Eq.\ref{phidot} that drain-hole particles have a limiting
2037: angular velocity
2038: \begin{equation}
2039: \lim_{\bar{r},\bar{z}\rightarrow0}\frac{\ d\phi}{d\tau}\ =-\ \frac
2040: {\left\langle \omega_{c\sigma}\right\rangle }{2\Omega_{0}\ }\lim_{\bar{r}%
2041: ,\bar{z}\rightarrow0}\left(  \frac{\bar{\psi}(\bar{r},\bar{z})}{\bar{r}^{2}%
2042: }\right)  =const. \label{drain-hole ang vel}%
2043: \end{equation}
2044: 
2045: 
2046: Combination of Eqs. \ref{p range} and \ref{drain-perigee} show that drain-hole
2047: particles can only be created if the accessibility condition%
2048: \begin{equation}
2049: \frac{\left\langle \omega_{c\sigma}\right\rangle ^{2}}{4\Omega_{0}^{2}%
2050: }\ \left(  \bar{\psi}(\bar{r}_{\ast},\bar{z}_{\ast})\right)  ^{2}<\left(
2051: 2\bar{\rho}^{2}\bar{H}+2\bar{\rho}\right)  \cos^{2}\theta
2052: \ \label{accessibility}%
2053: \end{equation}
2054: is satisfied, a condition that $\left\langle \omega_{c\sigma}\right\rangle
2055: /\Omega_{0}$ not be too large. Since $\bar{H}\simeq0$ is assumed, $\cos
2056: \theta\simeq-1$ for drain-hole particles, and since $\bar{\rho}=\sqrt{\bar
2057: {r}^{2}+\ \bar{z}^{2}}$ is required to be larger than the pericenter, this
2058: condition becomes
2059: \begin{equation}
2060: \left\vert \bar{\psi}(\bar{r}_{\ast},\bar{z}_{\ast})\right\vert <\ \left\vert
2061: \frac{2\Omega_{0}}{\left\langle \omega_{c\sigma}\right\rangle }\sqrt
2062: {2\bar{\rho}_{pericenter}}\right\vert . \label{drain hole creation criterion}%
2063: \end{equation}
2064: Because Eq.\ref{drain hole creation criterion} requires $\left\langle
2065: \omega_{c\sigma}\right\rangle /\Omega_{0}$ to be small, drain-hole particles,
2066: like Speiser particles, result from dust grains with grain radii consistent in
2067: order of magnitude with Eq.\ref{rdnom}.
2068: 
2069: Particles with $\bar{P}_{\phi}$ exactly zero (\textquotedblleft
2070: perfect\textquotedblright\ drain-hole particles) fall down the gravitational
2071: potential all the way to the central object at the origin $\bar{r}=0,$
2072: $\bar{z}=0.$ Particles that are not quite perfect drain-hole particles will
2073: have small, but finite $\bar{P}_{\phi}$ and so will reflect when close to the
2074: central object.
2075: 
2076: \begin{figure}[ptb]
2077: \caption{Drain-hole particle ($\bar{H}=0,$ $\bar{\rho}_{pericenter}=0.2,$
2078: $\theta=170^{0},\alpha=0,$ $\bar{R}_{\ast}=0.8$) falls across magnetic field
2079: all the way to the central object. This is a heavy particle (dust grain)\ and
2080: has $\omega_{c\sigma}/\Omega_{0}=1.6;$ (a)\ shows orbit projection in $\bar
2081: {x}$-$\bar{y}$ plane, (b) shows projection in $\bar{r}$-$\bar{z}$ plane.
2082: Dotted line shows trajectory parent would have continued to have if it had not
2083: become charged.}%
2084: \label{drain-hole_pericenter=0.3_theta=170_alpha=0_rho0=4_rstar=0.8_omegaratio=1.6.eps}%
2085: \epsscale{0.5} \plotone{f10.eps}\end{figure}
2086: 
2087: \begin{figure}[ptb]
2088: \caption{(a)\ Mechanical angular momentum $p_{\phi}$ for drain-hole particle
2089: is not conserved when particle becomes charged, (b)\ canonical angular
2090: momentum $P_{\phi},$ kinetic energy (KE) and potential energy (PE). The
2091: magnitudes of the kinetic and potential energy increase without bound as the
2092: particle falls towards the central object. The canonical angular momentum is
2093: conserved and is near zero upon charging.}%
2094: \label{drain-hole_pericenter=0.3_theta=170_alpha=0_rho0=4_rstar=0.8_omegaratio=1.6momenta-adobe.eps}%
2095: \epsscale{0.5} \plotone{f11.eps}\end{figure}
2096: 
2097: Figure
2098: \ref{drain-hole_pericenter=0.3_theta=170_alpha=0_rho0=4_rstar=0.8_omegaratio=1.6.eps}
2099: shows a numerical calculation of a drain-hole particle orbit with $\bar{H}=0.$
2100: The solid line in Fig.
2101: \ref{drain-hole_pericenter=0.3_theta=170_alpha=0_rho0=4_rstar=0.8_omegaratio=1.6.eps}%
2102: (a) shows the projection of the drain-hole orbit in the $\bar{x}$-$\bar{y}$
2103: plane. The orbit the neutral particle would have had if it had not encountered
2104: any photons and so remained neutral is shown as a dotted line. Figure
2105: \ref{drain-hole_pericenter=0.3_theta=170_alpha=0_rho0=4_rstar=0.8_omegaratio=1.6.eps}%
2106: (b)\ shows the projection in the $\bar{r}$-$\bar{z}$ plane with surfaces of
2107: constant $\psi$ indicated (the dotted line again shows the orbit the neutral
2108: particle would have had if it had not encountered any photons). The drain-hole
2109: particle has a retrograde orbit (clockwise sense resulting from its angle of
2110: inclination $\theta>90^{0}$). Figure
2111: \ref{drain-hole_pericenter=0.3_theta=170_alpha=0_rho0=4_rstar=0.8_omegaratio=1.6momenta-adobe.eps}%
2112: (a) shows that the mechanical angular momentum is not constant after charging
2113: while Fig.
2114: \ref{drain-hole_pericenter=0.3_theta=170_alpha=0_rho0=4_rstar=0.8_omegaratio=1.6momenta-adobe.eps}%
2115: (b) shows that the canonical angular momentum remains constant at zero after
2116: charging. Figure
2117: \ref{drain-hole_pericenter=0.3_theta=170_alpha=0_rho0=4_rstar=0.8_omegaratio=1.6momenta-adobe.eps}%
2118: (b) also shows how the magnitudes of the potential and kinetic energies
2119: increase without bound as the particle descends towards the central object
2120: while the total energy stays zero (kinetic, potential and total energies shown
2121: as dashed lines).
2122: 
2123: When drain-hole particles approach the central object, the gravitational term
2124: in Eq.\ref{chi free fall} dominates (recall that $\bar{\psi}\sim\bar{r}^{2}$
2125: at small $\bar{r}$ and near $\bar{z}=0$). It therefore makes sense to use
2126: spherical coordinates in this region in which case the Hamiltonian is
2127: approximately
2128: \begin{equation}
2129: 0\simeq\frac{1}{2}\left(  \frac{d\bar{R}}{d\tau}\right)  ^{2}-\frac{1}{\bar
2130: {R}} \label{free-fall}%
2131: \end{equation}
2132: where $\bar{R}$ is the spherical radius and $\bar{H}\simeq0$ has been assumed.
2133: Equation \ref{free-fall} shows that the free-fall velocity scales as
2134: \begin{equation}
2135: \ \left\vert \frac{d\bar{R}}{d\tau}\right\vert \ =\sqrt{\frac{2}{\bar{R}\ }}
2136: \label{free-fall velocity}%
2137: \end{equation}
2138: and particle flux conservation over a spherical surface $4\pi\bar{R}^{2}$
2139: shows that $4\pi\bar{R}^{2}n_{dh}(\bar{R})d\bar{R}/d\tau=const.$ where
2140: $n_{dh}(\bar{R})$ is the density of drain-hole particles. Thus, if the
2141: incoming drain-hole particles do not accumulate, spherical focusing combined
2142: with the accelerating free-fall velocity shows that the density of drain-hole
2143: particles scales as%
2144: \begin{equation}
2145: n_{dh}(\bar{R})\sim\frac{1}{\bar{R}^{2}d\bar{R}/d\tau}\sim\frac{1}{\bar
2146: {R}^{3/2}}. \label{drain-hole density}%
2147: \end{equation}
2148: 
2149: 
2150: Accumulation of the drain-hole particles in the vicinity of the central object
2151: will also increase the density of drain-hole particles with time. There is
2152: thus both a temporal increase and a geometrically-induced increase of the
2153: drain-hole particle density as $\bar{R}$ decreases. Since the sibling
2154: \ electrons were left stranded at large $\bar{r},$ what results is the
2155: establishment of a large positive charge density near the central object and
2156: an equal-magnitude negative charge density at large $\bar{r}$. The flow
2157: pattern of the drain hole particles and the location of the stranded electrons
2158: is sketched in Fig.\ref{f12.eps}. Eventually the positive space charge near
2159: the central object becomes so large that it produces a repulsive electrostatic
2160: electric field that balances the gravitational force acting on any additional
2161: drain hole particles. The large positive potential near $\bar{r}=0$ will tend
2162: to drive\ axial electric currents flowing away from the $\bar{z}=0$ plane
2163: resulting in the loss or neutralization of some of the drain-hole particles.
2164: The axial electric current could result from attraction of electrons near
2165: $\bar{r}=0$ towards the $\bar{z}=0$ plane or from expulsion of positive
2166: particles away from the $\bar{z}=0$ plane. Either electron attraction or
2167: positive particle repulsion will deplete the positive space charge density
2168: near $\bar{r}=0.$ There will then have to be a replenishing flow of additional
2169: drain hole particles into the $\bar{r}=0$ region to compensate for this
2170: depletion of positive space charge.\ Being very low mass, the stranded
2171: electrons have very small Larmor orbit radius and so are constrained to stay
2172: essentially right on the poloidal flux surface on which they were
2173: photo-emitted (see Secs.\ref{light-weight trapping} and
2174: \ref{trapping well width}). The electron flow is thus at a much larger
2175: $|\bar{z}|$ than the drain-hole particle radially inward flow which is
2176: concentrated near the $\bar{z}=0$ plane. The vertical separation between the
2177: respective radially inward flows of positive and negative particles means that
2178: bipolar toroidal magnetic fields will be generated in the interstitial regions
2179: between the electron flow and the drain-hole flow (see positive and negative
2180: $B_{\phi}$ regions in Fig.\ref{f12.eps}).
2181: 
2182: \begin{figure}[ptb]
2183: \caption{Drain hole dust grains fall across poloidal field lines towards
2184: central object leaving behind stranded electrons which are confined to
2185: poloidal flux surface on which they are born. Drain hole particles accumulate
2186: near central object creating large positive charge there. This repels positive
2187: particles (drain hole particles, ions) to flow axially away from $z=0$ plane
2188: and also attracts stranded electrons which can flow on poloidal flux surface.
2189: The result is a clockwise poloidal current flow pattern in upper-half $r$-$z$
2190: plane, giving a positive $B_{\phi}$ in region linked by poloidal current and a
2191: negative $B_{\phi}$ in lower-half $r$-$z$ plane where poloidal current flow is
2192: counter-clockwise.}%
2193: \label{f12.eps}%
2194: \plotone{f12.eps}\end{figure}
2195: 
2196: \begin{figure}[ptb]
2197: \caption{Flow of conventional electric current for drain hole particles and
2198: their associated stranded electrons. The electric field on the $z=0$ plane is
2199: radially outwards while the current flow is radially inwards so
2200: $\mathbf{J\cdot E}$ is negative, indicating that the infall of the drain-hole
2201: particles constitutes a dynamo. The $\mathbf{J\times B}$ force (which is
2202: essentially due to the gradient of $B_{\phi}^{2}$ and which is strongest at
2203: small $r$) drives a bipolar axial jet.}%
2204: \label{f13.eps}%
2205: \plotone{f13.eps}\end{figure}
2206: 
2207: If no electric current is allowed to flow, the situation is like a
2208: free-standing battery not connected to any load, i.e., a situation where there
2209: is a voltage differential across the battery terminals, but no current flows.
2210: However, if bipolar axial currents are allowed to flow, then the situation is
2211: like a battery connected to a load and the resulting radially inward
2212: drain-hole particle current in the $\bar{z}=0$ plane is like the internal
2213: current in a battery. The overall current flow pattern sketched in
2214: Fig.\ref{f12.eps} results from a combination of drain hole particle and
2215: electron motion. This pattern is sketched in Fig.\ref{f13.eps} as a
2216: conventional electric current. The geometry of the current flow pattern and
2217: electromotive force driving this current is identical to the geometry and flow
2218: patterns in the laboratory configuration simulating astrophysical jets
2219: described in
2220: %TCIMACRO{\TeXButton{citet{Hsu2002,Hsu2005}}{\citet{Hsu2002,Hsu2005}} }%
2221: %BeginExpansion
2222: \citet{Hsu2002,Hsu2005}
2223: %EndExpansion
2224: and
2225: %TCIMACRO{\TeXButton{citet{Bellan-a2005}}{\citet{Bellan2005}}}%
2226: %BeginExpansion
2227: \citet{Bellan2005}%
2228: %EndExpansion
2229: . The electric field due to the drain hole particles corresponds to the
2230: electric field produced by the capacitor bank used in the laboratory
2231: experiment. This geometry\ and symmetry is also identical to that proposed by
2232: %TCIMACRO{\TeXButton{citet{Lovelace1976}}{\citet{Lovelace1976}}}%
2233: %BeginExpansion
2234: \citet{Lovelace1976}%
2235: %EndExpansion
2236: , the only difference being the means by which the radial electric field is
2237: produced. The magnetic fields in the lab and astrophysical plasmas have the
2238: same toroidal/poloidal topology.
2239: 
2240: The drain-hole current is\ thus powered by gravity and has $J_{r}$ radially
2241: inward with $E_{r}$ radially outward so that $\mathbf{J\cdot E}$ is negative,
2242: consistent with the condition for a dynamo. The drain-hole particles have
2243: retrograde motion so their mechanical angular momentum is negative. This
2244: negative mechanical angular momentum is removed by the braking torque
2245: $\mathbf{r\times F=}$ $\left(  r\hat{r}\right)  \times\left(  J_{r}\hat
2246: {r}\times B_{z}\hat{z}\right)  =-rJ_{r}B_{z}\hat{z}$ which is positive since
2247: $J_{r}$ is negative and $B_{z}$ is positive.
2248: 
2249: The radially inward current is symmetric with respect to $z.$ This property
2250: provides enough information to determine the symmetry properties of $I(r,z).$
2251: The radially inward drain-hole current means that $J_{r}<0$ and $J_{z}=0$ in
2252: the $z=0$ plane. Since Eq.\ref{Jpol2} shows that $J_{r}=-(2\pi r)^{-1}\partial
2253: I/\partial z$ and $J_{z}=$ $(2\pi r)^{-1}\partial I/\partial r$ the condition
2254: $J_{z}=0$ means $I(r,z)$ must vanish in the $z=0$ plane$.$ Furthermore $I$
2255: must be an odd function of $z$ in order for $J_{r}=-(2\pi r)^{-1}\partial
2256: I/\partial z$ to be finite in the $z=0$ plane. Finally $\partial I/\partial z$
2257: should be positive in order to have $J_{r}<0.$ Thus, $I(r,z)$ should be
2258: positive for $z>0$ and negative for $z<0$ so that, as sketched in
2259: Fig.\ref{f13.eps}, there will be a bipolar axial current flowing along the $z$
2260: axis outwards from the $z=0$ plane. The accumulation of drain-hole particles
2261: constitutes the engine that drives the poloidal electric current that drives
2262: the astrophysical jet. The $z$-symmetry of $\psi(r,z)$ and $z$-antisymmetry of
2263: $I(r,z)$ has been noted previously by
2264: %TCIMACRO{\TeXButton{\citet{Ferreira1995}}{\citet{Ferreira1995}}}%
2265: %BeginExpansion
2266: \citet{Ferreira1995}%
2267: %EndExpansion
2268: .
2269: 
2270: Electromagnetic power flow from this dynamo can also be interpreted in terms
2271: of the Poynting flux $\mathbf{S=E\times B}/\mu_{0}.$ Azimuthal symmetry
2272: applied to Faraday's law shows that $E_{\phi}$ is zero for a steady-state
2273: situation in which case the $z$-component of the Poynting flux reduces to
2274: $S_{z}=E_{r}B_{\phi}/\mu_{0}.$ Because $E_{r}$ and $B_{\phi}$ are both
2275: positive for $z>0$ whereas $E_{r}$ is positive while $B_{\phi}$ is negative
2276: for $z<0$, it is seen that $S_{z}$ is positive for $z>0$ and negative for
2277: $z<0.$ Thus, the Poynting flux associated with this dust-driven dynamo injects
2278: energy into bipolar astrophysical jets flowing normally outward from the $z=0$ plane.
2279: 
2280: One can ask just how close to exactly zero $\bar{P}_{\phi}$ has to be in order
2281: for a particle to behave as a drain-hole particle. Exact $\bar{P}_{\phi}=0$
2282: would enable a particle to spiral down all the way to the center of the
2283: central object, an obviously unrealistic situation because the particle would
2284: vaporize as it approached the stellar surface. A more realistic question then
2285: is how small does $\bar{P}_{\phi}$ have to be in order for a drain-hole
2286: particle to fall to some specified normalized radius $\bar{R}_{small}\ \ $that
2287: is much less than unity. $\bar{R}_{small}$ would presumably be of the order of
2288: the radius at which the astrophysical jet starts and so would be of the order
2289: of the thickness of the accretion disk or somewhat smaller. Since the
2290: dimensionless form of Eq.\ref{H} is
2291: \begin{equation}
2292: \bar{H}=\ \frac{1}{2}(\bar{v}_{r}^{2}+\bar{v}_{z}^{2})\ +\chi(\bar{r},\bar{z})
2293: \label{H TP}%
2294: \end{equation}
2295: where the effective potential is
2296: \begin{equation}
2297: \chi(\bar{r},\bar{z})=\frac{1}{2\bar{r}^{2}}\left(  \bar{P}_{\phi}%
2298: \ \ -\ \frac{\left\langle \omega_{c\sigma}\right\rangle }{2\Omega_{0}}%
2299: \bar{\psi}(\bar{r},\bar{z})\right)  ^{2}-\frac{1}{\sqrt{\bar{r}^{2}+\bar
2300: {z}^{2}}} \label{chi TP}%
2301: \end{equation}
2302: and since $\bar{H}\simeq0$, the turning point for a drain-hole particle is
2303: where $\chi(\bar{r},\bar{z})\simeq0.$ Because $\bar{\psi}(\bar{r},\bar
2304: {z})\rightarrow0$ at small $\bar{r},$ the inner turning point will therefore
2305: be where $\bar{P}_{\phi}^{2}=2\bar{r}^{2}/\sqrt{\bar{r}^{2}+\bar{z}^{2}}.$
2306: Assuming that\ the inner turning point is at $\bar{r}\ \sim\bar{R}_{small}$
2307: and $\tilde{z}\simeq0,$ the inner turning point is where $\bar{P}_{\phi}%
2308: ^{2}=2\bar{R}_{small}.$ A sufficient condition for assuming $\bar{P}_{\phi
2309: }\simeq0$ is thus $\bar{P}_{\phi}^{2}<2\bar{R}_{small}$, i.e.,
2310: \begin{equation}
2311: -\sqrt{2\bar{R}_{small}}<\bar{L}\cos\theta+\frac{\left\langle \omega_{c\sigma
2312: }\right\rangle }{2\Omega_{0}}\ \bar{\psi}(\bar{r}_{\ast},\bar{z}_{\ast}%
2313: )<\sqrt{2\bar{R}_{small}} \label{Pphi crit}%
2314: \end{equation}
2315: and particles satisfying this condition will fall to a normalized radius
2316: $\bar{R}<$ $\bar{R}_{small}.$ For given $\bar{L},$ $\left\langle
2317: \omega_{c\sigma}\right\rangle /2\Omega_{0},$ and $\ \bar{\psi}(\bar{r}_{\ast
2318: },\bar{z}_{\ast})$ this corresponds to a narrow range in $\theta$ centered
2319: about the angle at which $\bar{P}_{\phi}$ equals zero $\ $exactly. Equation
2320: \ref{Pphi crit} can be expressed as $\ \cos\left(  \theta+\Delta
2321: \theta/2\right)  <\cos\theta<\cos\left(  \theta-\Delta\theta/2\right)  $ where%
2322: \begin{equation}%
2323: \begin{array}
2324: [c]{ccc}%
2325: \bar{L}\cos\left(  \theta+\Delta\theta/2\right)  & = & -\sqrt{2\bar{R}%
2326: _{small}}-\frac{\left\langle \omega_{c\sigma}\right\rangle }{2\Omega_{0}%
2327: }\ \bar{\psi}(\bar{r}_{\ast},\bar{z}_{\ast})\\
2328: \bar{L}\cos\left(  \theta-\Delta\theta/2\right)  & = & \sqrt{2\bar{R}_{small}%
2329: }-\frac{\left\langle \omega_{c\sigma}\right\rangle }{2\Omega_{0}}\ \bar{\psi
2330: }(\bar{r}_{\ast},\bar{z}_{\ast}).
2331: \end{array}
2332: \label{def theta minmax}%
2333: \end{equation}
2334: Subtracting these two equations from each other shows that the range
2335: $\Delta\theta$ for drain-hole particles to reach $\bar{R}_{small}$ is
2336: \begin{equation}
2337: \Delta\theta\simeq\frac{2\sqrt{2\bar{R}_{small}}}{\bar{L}\sin\theta}.
2338: \label{delta theta}%
2339: \end{equation}
2340: The solid angle of incident particles lying between $\theta$ and
2341: $\theta+\Delta\theta$ is $2\pi\sin\theta\Delta\theta$ and so the fraction
2342: $f_{dh}$ of all incident particles with angular momentum $\bar{L}$ that become
2343: drain-hole particles and fall to $\bar{R}<$ $\bar{R}_{small}$ is $\ $ $\ $
2344: \begin{equation}
2345: f_{dh}=\ \frac{2\pi\sin\theta\Delta\theta}{4\pi}\ =\frac{\ \sqrt{2\bar
2346: {R}_{small}}}{\bar{L}\ }. \label{drain hole fraction}%
2347: \end{equation}
2348: 
2349: 
2350: \subsection{Drain hole dynamo power}
2351: 
2352: The strength of the equilibrium radial electric field produced by drain-hole
2353: particles can be estimated as follows: Before any drain-hole particles
2354: accumulate at small $\bar{r},$ there is no radial electric field, but as the
2355: drain-hole particles accumulate, the radial outward electric field will
2356: develop. The force due to the radial outward electric field will oppose the
2357: gravitational and magnetic forces causing the inward motion of the drain-hole
2358: particles. The balance between these opposing forces is quantified by the
2359: radial equation of motion. In cylindrical un-normalized coordinates the radial
2360: equation of motion governing drain-hole particles with $v_{z}=0$ in the $z=0$
2361: plane (where $B_{\phi}=0$ due to $z$-antisymmetry of $I(r,z)$) is
2362: \begin{equation}
2363: m_{d}\left(  \ddot{r}-r\dot{\phi}^{2}\right)  =q_{d}\left(  E_{r}+r\dot{\phi
2364: }B_{z}\right)  -\frac{m_{d}MG}{r^{2}}.\label{drain-hole motion}%
2365: \end{equation}
2366: $B_{z}$ is approximately uniform at small $r$ so $\psi\simeq\pi r^{2}B_{z}%
2367: \ $at small $r$ in which case the drain-hole particle condition $P_{\phi
2368: }=m_{d}r^{2}\dot{\phi}+q_{d}\psi/2\pi=0$ implies $\ \dot{\phi}=-q_{d}%
2369: B_{z}/2m_{d}.\ $On eliminating $\dot{\phi}\ $in Eq.\ref{drain-hole motion},
2370: the radial equation of motion governing drain-hole particles is%
2371: \begin{equation}%
2372: \begin{array}
2373: [c]{ccl}%
2374: \ \ddot{r} & = & \frac{q_{d}}{m_{d}}E_{r}-r\frac{\ q_{d}^{2}B_{z}^{2}}%
2375: {4m_{d}^{2}\ }\ -\frac{MG}{r^{2}}\\
2376: & \simeq & \frac{q_{d}}{m_{d}}E_{r}\ \ -\frac{MG}{r^{2}}%
2377: \end{array}
2378: \label{drain-hole motion2}%
2379: \end{equation}
2380: where the second line is for small $r.$ When $E_{r}=0$, the drain-hole
2381: particles fall inwards with gravitational acceleration, but as $E_{r}$ builds
2382: up because of accumulation at small $r$ of fallen-in positively charged
2383: drain-hole particles, Eq.\ref{drain-hole motion2} shows that this electric
2384: field will oppose the gravitational force and retard the infall. The
2385: drain-hole particles will continue to fall in and accumulate, thereby
2386: increasing $E_{r}$ until the radially outward repulsive electrostatic force
2387: due to $E_{r}$ becomes so strong as to balance gravity and cause $\ddot{r}$ to
2388: vanish. Thus, gravitational force is balanced by the radially outward force
2389: from the space charge electric field of the accumulated positively charged
2390: drain-hole particles. The saturation electric field is
2391: \begin{equation}
2392: \ E_{r}\ \ =\frac{m_{d}MG}{q_{d}r^{2}}.\label{drain hole Er}%
2393: \end{equation}
2394: 
2395: 
2396: Using $E_{r}=-\partial V/\partial r,$ integration of Eq.\ref{drain hole Er}
2397: from large $r$ to the jet radius $r_{jet}$ gives the voltage at the jet to be%
2398: \begin{equation}
2399: V_{jet}=\frac{m_{d}MG}{q_{d}r_{jet}}. \label{Vjet}%
2400: \end{equation}
2401: 
2402: 
2403: The jet electric current corresponds to the charge per second carried inward
2404: by the drain-hole particles. The number of drain-hole particles accreting per
2405: second is $\dot{M}_{dh}/m_{d}$ where $\dot{M}_{dh}$ is the mass accretion rate
2406: per second of drain-hole particles and $m_{d}$ is the mass of an individual
2407: drain-hole particle. Thus, the poloidal electric current is
2408: \begin{equation}
2409: I_{jet}=q_{d}\dot{M}_{dh}/m_{d}. \label{Ijet}%
2410: \end{equation}
2411: The jet electric power is
2412: \begin{equation}
2413: P_{jet}=I_{jet}V_{jet}=\frac{\ \dot{M}_{dh}MG}{\ r_{jet}}\ \ \ \label{Pjet}%
2414: \end{equation}
2415: which is just the rate at which gravitational potential energy is released by
2416: drain-hole particles falling from large $r$ to the jet radius. The drain-hole
2417: dynamo converts the gravitational energy released from accretion into
2418: electrical power suitable for driving bipolar jets that are moving away from
2419: the $z=0$ plane. The jet power is proportional to both the central object mass
2420: $M$ and the drain-hole mass accretion rate $\dot{M}_{dh}$. Paper I showed that
2421: the dust mass accretion rate can be a substantial fraction of the total mass
2422: accretion so $P_{jet}$ can be a substantial fraction of the power of all
2423: accreting material. The jet power accelerates the jet material to escape
2424: velocity and so is equal to the power available from accreting drain-hole dust
2425: grains. Thus, assuming that the axial starting point for jet particles is of
2426: the order of $r_{jet},$ the power required to drive the jet particles to
2427: escape velocity is $P_{jet}=\dot{M}_{jet}MG/r_{jet}$ and so the jet mass
2428: outflow would be approximately equivalent to the drain hole particle accretion
2429: rate, i.e., $\dot{M}_{jet}\simeq\dot{M}_{dh}.$ The particles in the jet would
2430: not, in general, be the drain hole particles, but instead would be plasma
2431: magnetohydrodynamically accelerated using the drain-hole accretion as the
2432: power source. Assuming $\bar{L}\simeq1\ $ and $\bar{R}_{small}=r_{jet}%
2433: /a\sim0.1$ in Eq.\ref{drain hole fraction}, the fraction of retrograde
2434: particles that are drain hole and able to reach $r_{jet}$ would be
2435: $f_{dh}=\ \allowbreak0.4;$ the fraction of combined retrograde and prograde
2436: dust grains would thus be 0.2. The example in paper I\ showed that because of
2437: differences in proportional slowing down, the dust accretion rate would be
2438: enriched to be 20\% of the gas accretion rate. This gives $\dot{M}_{dh}%
2439: /\dot{M}_{g}\sim0.2\times0.2=0.04$ and so predicts a jet power that would be
2440: about 1/25 of the power associated with all accreting dust and gas. This ratio
2441: of outflow power to accretion power is consistent with the estimate given by
2442: %TCIMACRO{\TeXButton{\citet{Bacciotti2004}}{\citet{Bacciotti2004}} }%
2443: %BeginExpansion
2444: \citet{Bacciotti2004}
2445: %EndExpansion
2446: using HST observations of T Tauri jets.
2447: 
2448: \section{Torque and angular momentum}
2449: 
2450: An important question repeatedly asked about accretion disks and jets is the
2451: role played by jets in satisfying conservation of \ mechanical angular
2452: momentum of the accreting material. It will now be shown that mechanical
2453: angular momentum is exactly conserved in our model.
2454: 
2455: Because of axisymmetry, the canonical angular momentum of the $j^{th}$
2456: $\ $charged dust grain
2457: \begin{equation}
2458: P_{\phi d,j}^{+}=m_{+}rv_{\phi}^{j}+Ze\psi(r,z)/2\pi\ \label{canmomentum}%
2459: \end{equation}
2460: and the canonical angular momentum of the $k^{th}$ electron
2461: \begin{equation}
2462: P_{\phi,k}^{e}=m_{e}rv_{\phi}^{k}-e\psi(r,z)/2\pi\label{e can momentum}%
2463: \end{equation}
2464: at any position $r,z$ are both invariants, i.e., $P_{\phi d}^{+}=const.$ and
2465: $P_{\phi}^{e}=const.$ In general, the dust grains and the electrons at any
2466: position $r$,$z$ will have quite different values of $v_{\phi}\ $but, in order
2467: for the plasma to be macroscopically quasi-neutral, there must be
2468: approximately $Z$ electrons adjacent to each dust grain$.$ Because electron
2469: and dust grain trajectories differ, these neighboring electrons will typically
2470: not be the original sibling electrons photo-emitted when the dust grain became charged.
2471: 
2472: The radial and axial velocities of a specific dust grain or electron can be
2473: written as $v_{r}^{\sigma,j}\ =dr^{\sigma,j}/dt$ and $\ v_{z}^{\sigma
2474: ,j}\ =dz^{\sigma,j}/dt$ where $r^{\sigma,j}(t)\ $ and $z^{\sigma,j}(t)$ are
2475: the position of the $j^{th}$ particle of species $\sigma$. Since $P_{\phi}$ is
2476: conserved for each individual particle, the time derivatives of the $P_{\phi}%
2477: $'s of a dust grain at a location $r,z$ and its neighboring neutralizing $Z$
2478: electrons respectively give $dP_{\phi d,j}^{+}/dt=0$ and $dP_{\phi,k}%
2479: ^{e}/dt=0.$ Using $d\psi/dt=v_{r}\partial\psi/\partial r+v_{z}\partial
2480: \psi/\partial z$ for the time derivative of $\psi$ measured in the particle
2481: frame, respective time derivatives of Eqs.\ref{canmomentum} and
2482: \ref{e can momentum} give
2483: \begin{equation}
2484: \frac{d}{dt}\left(  m_{+}rv_{\phi}^{+,j}\right)  =\ -\frac{Ze}{2\pi}\left(
2485: \frac{\partial\psi}{\partial r}v_{r}^{+,j}+\frac{\partial\psi}{\partial
2486: z}v_{z}^{,j}\right)  \label{ang momentum dust deriv}%
2487: \end{equation}
2488: and
2489: \begin{equation}
2490: \frac{d}{dt}\left(  Zm_{e}rv_{\phi}^{e,k}\right)  \ =\frac{Ze}{2\pi}\left(
2491: \frac{\partial\psi}{\partial r}v_{r}^{e,k}+\frac{\partial\psi}{\partial
2492: z}v_{z}^{e,k}\right)  . \label{ang momentum electron deriv}%
2493: \end{equation}
2494: 
2495: 
2496: Using $B_{r}=-(2\pi r)^{-1}\partial\psi/\partial z$ and $B_{z}=(2\pi
2497: r)^{-1}\partial\psi/\partial r$ from Eq.\ref{Bpol def} and summing
2498: Eqs.\ref{ang momentum dust deriv} and \ref{ang momentum electron deriv} over
2499: the dust grains and their associated $Z$ neutralizing electrons at location
2500: $r,z$ gives
2501: \begin{equation}
2502: \frac{dL_{\phi}}{dt}=r\left(  J_{z}B_{r}-J_{r}B_{z}\right)  =\ \ r\hat{\phi
2503: }\cdot\mathbf{J}_{pol}\mathbf{\times B}_{pol} \label{MHD ang momentum}%
2504: \end{equation}
2505: where $J_{r}$, $J_{z}$ are the respective radial and axial current densities
2506: and $L_{\phi}$ is the total mechanical angular momentum density taking into
2507: account both dust grains and electrons. Thus, from the macroscopic point of
2508: view there is a torque about the $z$ axis, namely $\hat{z}\cdot\mathbf{r\times
2509: F}=\hat{z}\times\left(  r\hat{r}+z\hat{z}\right)  \mathbf{\cdot}\left(
2510: \mathbf{J}_{pol}\times\mathbf{B}_{pol}\right)  =r\hat{\phi}\cdot
2511: \mathbf{J}_{pol}\mathbf{\times B}_{pol}$ acting to change the local mechanical
2512: angular momentum density.
2513: 
2514: On the other hand, using Eqs.\ref{Bpol def} and \ref{Jpol2} it is seen that
2515: when this torque is integrated over the entire volume to infinity,
2516: \begin{equation}%
2517: \begin{array}
2518: [c]{ccl}%
2519: \int d^{3}r\,r\hat{\phi}\cdot\mathbf{J}_{pol}\times\mathbf{B}_{pol}\  & = &
2520: \int\ d^{3}r\,r^{2}\nabla\phi\cdot\left(  \frac{1}{2\pi}\nabla I\times
2521: \nabla\phi\right)  \times\left(  \frac{1}{2\pi}\nabla\psi\times\nabla
2522: \phi\right) \\
2523: & = & \frac{1}{4\pi^{2}}\int\ d^{3}r\,r^{2}\nabla\phi\times\left(  \ \nabla
2524: I\times\nabla\phi\right)  \cdot\left(  \ \nabla\psi\times\nabla\phi\right) \\
2525: & = & \frac{1}{4\pi^{2}}\int\ d^{3}r\,\nabla I\cdot\left(  \nabla\psi
2526: \times\nabla\phi\right) \\
2527: & = & \frac{1}{4\pi^{2}}\int\ d^{3}r\,\nabla\cdot\left(  I\left(  \nabla
2528: \psi\times\nabla\phi\right)  \right) \\
2529: & = & 0
2530: \end{array}
2531: \label{integrate torque}%
2532: \end{equation}
2533: since both $I$ and $\nabla\psi$ vanish at $\infty.$ Thus, the total mechanical
2534: angular momentum of the system is exactly \ conserved because there is no net
2535: torque applied to the whole system.
2536: 
2537: $I$ is an odd function of $z$ and $\psi$ is an even function of $z,$ and in
2538: the jet $\mathbf{J}_{pol}$ is nearly parallel to $\mathbf{B}_{pol}.$ This
2539: suggests the following generic form for the current
2540: \begin{equation}
2541: \mu_{0}I(r,z)\simeq\lambda\psi(r,z)\tanh\left(  \frac{z}{h(r)}\right)
2542: \label{generic current}%
2543: \end{equation}
2544: where $h(r)$ represents the height of the accretion disk at radius $r\ $and
2545: $\lambda,$ the current per flux, has units of inverse length. Thus far from
2546: the $z=0$ plane, Eq.\ref{generic current} has the form $\mu_{0}I/\psi=\lambda
2547: $sign$(z)$ so that the jet above the $z=0$ plane has the opposite handedness
2548: of the jet below the $z=0$ plane. The parameter $\lambda$ is closely related
2549: to the current per flux in a force-free system (i.e., a system satisfying
2550: $\ \nabla\times\mathbf{B=}\lambda\mathbf{B}$), but differs slightly because
2551: here $\lambda$ refers to just the ratio of the poloidal current to the
2552: poloidal flux. Using Eq.\ref{MHD ang momentum} it is seen that the density of
2553: MHD\ torque about the $z$ axis is of the generic form%
2554: \begin{equation}%
2555: \begin{array}
2556: [c]{ccl}%
2557: \frac{dL_{\phi}}{dt} & = & \frac{1}{4\pi^{2}r\ }\left(  -\frac{\partial
2558: I}{\partial r}\frac{\partial\psi}{\partial z}+\frac{\partial I}{\partial
2559: z}\frac{\partial\psi}{\partial r}\right) \\
2560: & = & \frac{\lambda}{4\mu_{0}\pi^{2}r}\left\{  -\frac{\partial\ }{\partial
2561: r}\left[  \psi(r,z)\tanh\left(  \frac{z}{h(r)}\right)  \right]  \frac
2562: {\partial\psi}{\partial z}+\frac{\partial\ }{\partial z}\left[  \psi
2563: (r,z)\tanh\left(  \frac{z}{h(r)}\right)  \right]  \frac{\partial\psi}{\partial
2564: r}\right\}  .
2565: \end{array}
2566: \label{dLphitdt}%
2567: \end{equation}
2568: We assume that $\partial h/\partial r\ll1$ so the radial scale length at which
2569: $h$ changes is much larger than $h.$ Also, from symmetry $\partial
2570: \psi/\partial z=0$ on the $z=0$ midplane. Together, these conditions imply
2571: that near the midplane the last term in Eq.\ref{dLphitdt} dominates so near
2572: the midplane
2573: \begin{equation}
2574: \frac{dL_{\phi}}{dt}\simeq\frac{\lambda}{4\mu_{0}\pi^{2}hr}\frac{\psi
2575: (r,z)}{\cosh^{2}(z/h)}\frac{\partial\psi}{\partial r}=\frac{\lambda\psi
2576: (r,z)}{2\pi\mu_{0}h}\frac{B_{z}(r,z)}{\cosh^{2}(z/h)}. \label{approx dLphidt}%
2577: \end{equation}
2578: As seen from Eq.\ref{integrate torque} the torque density is proportional to
2579: $\ -\nabla I\times\nabla\phi\cdot\nabla\psi\sim-r\mathbf{J}_{pol}\cdot
2580: \nabla\psi$ and so is positive for poloidal current flow away from the
2581: poloidal field magnetic axis and negative for poloidal current flow towards
2582: the poloidal field magnetic axis$.$The \ torque density vanishes as
2583: $r\rightarrow0$ and as $r\rightarrow\infty$ and at the poloidal field magnetic
2584: axis because $r\mathbf{\ }\nabla\psi\rightarrow0$ at $r=0,$ $\mathbf{J}%
2585: _{pol}\rightarrow0$ at $r=\infty$, and $\nabla\psi=0$ at the poloidal field
2586: magnetic axis. The direction of poloidal current flow is shown in
2587: Fig.\ref{f13.eps}. This \ torque acts on the drain hole particles and their
2588: associated electrons since these particles are the carriers of the poloidal
2589: current as sketched in Fig.\ref{f12.eps} and Fig.\ref{f13.eps}. Unlike the
2590: drain-hole particles, no \ torque $r\hat{\phi}\cdot\mathbf{J\times B}$ about
2591: the $z$ axis acts on the Speiser particles because the current associated with
2592: the Speiser particles is in the $\phi$ direction.
2593: 
2594: \section{Conclusions}
2595: 
2596: We have shown that charging of collisionless dust grains incident upon a star
2597: causes the dust grain orbital dynamics to change from a relatively simple
2598: Kepler form to more complicated motion involving competition between magnetic
2599: and gravitational forces. This competition gives rise to five qualitatively
2600: different types of orbits. Two of these, the retrograde and prograde cometary
2601: orbits are just perturbations of Kepler cometary orbits. The orbit of a
2602: particle where magnetic forces overwhelm gravitational forces is just a Larmor
2603: (cyclotron) orbit and in this case the particle is constrained to remain
2604: within a poloidal Larmor radius of a poloidal flux surface in a manner similar
2605: to tokamak confinement. Particles where magnetic and gravitational forces are
2606: comparable can have two very different types of orbit depending on whether the
2607: incident particle is prograde or retrograde. Prograde particles of this latter
2608: type develop Speiser orbits; these orbits are paramagnetic with respect to the
2609: poloidal magnetic field and so can be the source of the poloidal magnetic
2610: field. Retrograde particles having comparable magnetic and gravitational
2611: forces can have a peculiar behavior whereby centrifugal force is eliminated
2612: with the result that the charged particle falls in towards the star along a
2613: spiral orbit. The accumulation of these \textquotedblleft
2614: drain-hole\textquotedblright\ particles near the star provides a radial
2615: electric field oriented so as to drive the poloidal currents and toroidal
2616: magnetic fields of an astrophysical jet.
2617: 
2618: This paper showed the existence of these different types of orbits, how their
2619: orientation is suitable for generating the poloidal and toroidal magnetic
2620: fields associated with an accretion disk and astrophysical jet, and how
2621: questions of angular momentum conservation are inherently resolved. A future
2622: paper will investigate the quantitative values of dust grain parameters
2623: required to produce toroidal and poloidal fields in the accretion disk of a
2624: young stellar object.
2625: 
2626: Finally, we offer some remarks regarding the effect of deviations from axisymmetry. The model presented here assumed perfect magnetic field axisymmetry 
2627: field whereas actual accretion disks are observed to have varying amounts of non-axisymmetry. This situation is  analogous
2628: to toroidal magnetic fusion devices such as tokamaks, reversed field pinches, and spheromaks all of which are modeled to first
2629: approximation as being axisymmetric, but in reality have  deviations from axisymmetry due to waves, turbulence, instability, and errors in machine construction.
2630: It is known from these devices that a modest breaking of symmetry  does not invalidate the results of the axisymmetric model, but rather weakens the
2631: conclusions, e.g., instead of cyclotron-orbiting particles being perfectly confined to the vicinity of a poloidal flux surface, when there is deviation from
2632: axisymmetry  cyclotron-orbiting particles can slowly wander away from  the  poloidal flux surface they started on.
2633: One would expect that deviations from axisymmetry in accretion disks  would cause a similar transport of cyclotron particles across poloidal flux surfaces.
2634: Because symmetry breaking  causes the canonical angular momentum of particles to change, it could be 
2635: considered as being somewhat like a collision that changes the canonical angular momentum of each of two  particles involved in a collision
2636: while conserving the total canonical angular momenta. Hence,  deviations from axisymmetry
2637: would  cause a   jiggling   of the canonical angular momenta of individual particles so that particles  on
2638: the borderline between being drain-hole and cyclotron or on the borderline between being Speiser and cyclotron might spend part of the  time (i.e., between jiggles) 
2639: being one type and part of the time  being the neighboring type. Similarly cyclotron particles that are on the borderline between being
2640: mirror-trapped and not mirror-trapped would, as they get kicked into and out of the mirror loss-cone,  spend part of the  time being mirror-trapped and 
2641: part of the  time not being mirror-trapped. However, at any given time there would
2642: be a certain fraction of particles of each type, i.e., a certain fraction would be cyclotron, a certain fraction would be drain-hole,   
2643: a certain fraction would be
2644: Speiser, and a certain fraction would be  cometary.
2645:   
2646: \pagebreak
2647: 
2648: \appendix
2649: 
2650: 
2651: \begin{center}
2652: \textbf{APPENDICES}
2653: \end{center}
2654: 
2655: \section{Derivation of generic poloidal flux
2656: function\label{Generic Flux Function}}
2657: 
2658: If all the toroidal current $\mathcal{I}_{\phi}$ is concentrated at the
2659: poloidal location $r=R_{0}$ and $z=0$, then the toroidal current density is
2660: $\mathbf{J}_{tor}=\ \hat{\phi}\mathcal{I}_{\phi}\delta(z)\delta(r-R_{0}).$ On defining%
2661: 
2662: \begin{equation}
2663: k^{2}=\frac{4R_{0}r}{\left(  R_{0}+r\right)  ^{2}+z^{2}} \label{k2 def}%
2664: \end{equation}
2665: analytic solution of Eq.\ref{Jtor} using $\mathbf{J}_{tor}=\ \hat{\phi
2666: }\mathcal{I}_{\phi}\delta(z)\delta(r-R_{0})$ gives
2667: %TCIMACRO{\TeXButton{\citep{Jackson1999}}{\citep{Jackson1999}} }%
2668: %BeginExpansion
2669: \citep{Jackson1999}
2670: %EndExpansion
2671: \begin{equation}
2672: \psi(r,z)=\ \ \frac{\mu_{0}\mathcal{I}_{\phi}}{k\ }\sqrt{R_{0}r}\left[
2673: \left(  2-k^{2}\ \right)  K(k)-2E(k)\right]  \ \label{elliptic}%
2674: \end{equation}
2675: where $E$ and $K$ are complete elliptic integrals. Equation \ref{elliptic}
2676: describes the situation where all the current density is concentrated at
2677: $r=R_{0},$ $z=0$, i.e., the current flows in a wire of zero cross-section
2678: located at $r=R_{0},$ $z=0.$ This equation can also be used to (i)\ describe
2679: the field observed at locations far from the poloidal field magnetic axis of a
2680: distributed current localized in the vicinity of the poloidal field magnetic
2681: axis and (ii)\ as the Green's function for a distributed toroidal current.
2682: This is because for an observer who is far from $r=R_{0},$ $z=0,$ the field of
2683: a distributed toroidal current localized near $r=R_{0},$ $z=0$ is
2684: indistinguishable from the field of a zero cross-section wire carrying the
2685: same total current. Equation \ref{elliptic} has a logarithmic singularity at
2686: the wire location because the wire has infinitesimal diameter.
2687: 
2688: Two analytic limits are of interest for Eq.\ref{elliptic}. The first is where
2689: $r\ll R_{0}$ so
2690: \begin{equation}
2691: k^{2}\simeq\frac{4R_{0}r}{R_{0}^{2}+z^{2}} \label{k2 approx 1}%
2692: \end{equation}
2693: and the second is where $r\gg R_{0}$ so
2694: \begin{equation}
2695: k^{2}\simeq\frac{4R_{0}r}{r^{2}+z^{2}}. \label{k2 approx 2}%
2696: \end{equation}
2697: The former gives the field near the loop axis and the latter gives the field
2698: at locations far from the current loop. In both cases $k^{2}$ is small
2699: compared to unity and so the small argument asymptotic expansions of the
2700: complete elliptic integrals can be used, namely,%
2701: \begin{equation}
2702: E(k)=\frac{\pi}{2}\left(  1-\frac{k^{2}}{4}\ -\frac{3}{64}k^{4}-...\right)
2703: ,\qquad K(k)=\frac{\pi}{2}\left(  1+\frac{k^{2}}{4}+\frac{9}{64}%
2704: k^{4}+...\right)  . \label{elliptic asymptotic}%
2705: \end{equation}
2706: Thus for small $k,$ it is seen that $\left(  2-k^{2}\right)  K(k)-2E(k)\simeq
2707: \pi k^{4}/16$ in which case$\ $%
2708: \begin{equation}
2709: \psi(r,z)=\ \frac{\pi\mu_{0}\mathcal{I}_{\phi}}{2\ \ }\frac{R_{0}^{2}r^{2}%
2710: }{\left(  \left(  R_{0}+r\right)  ^{2}+z^{2}\right)  ^{3/2}}%
2711: \ \label{psi approx}%
2712: \end{equation}
2713: so for $r\ll R_{0}$
2714: \begin{equation}
2715: \lim_{r\ll a}\psi(r,z)\simeq\frac{\pi\mu_{0}\mathcal{I}_{\phi}}{2\ \ }%
2716: \frac{R_{0}^{2}r^{2}}{\left(  R_{0}^{2}+z^{2}\right)  ^{3/2}}
2717: \label{small r psi}%
2718: \end{equation}
2719: and for $r\gg R_{0}$%
2720: \begin{equation}
2721: \lim_{r\gg a}\psi(r,z)\simeq\frac{\pi\mu_{0}\mathcal{I}_{\phi}}{2\ \ }%
2722: \frac{R_{0}^{2}r^{2}}{\left(  r^{2}+z^{2}\right)  ^{3/2}}. \label{large r psi}%
2723: \end{equation}
2724: For purposes of discussion and also numerical computation, it is convenient to
2725: choose Eq.\ref{psi approx} to represent the poloidal flux of a generic
2726: toroidal current \textit{everywhere}. Making this choice for the poloidal flux
2727: function (instead of the prescription given by Eq.\ref{elliptic}) means that
2728: $\psi(r,z)$ has a smooth hill-top at $r=2R_{0}$ rather than a logarithmic
2729: singularity at $r=R_{0}$ and has the same behavior far from $r=R_{0},$ $z=0$
2730: as does Eq.\ref{elliptic}.
2731: 
2732: Thus, a useful analytic representation for a nonsingular, physically
2733: realizable flux function is obtained by recasting Eq.\ref{psi approx} \ in the
2734: form
2735: \begin{equation}
2736: \psi(r,z)=\ \ \frac{27\left(  r/a\right)  ^{2}}{8\left(  \left(  \frac{r}%
2737: {a}+\frac{1}{2}\right)  ^{2}+\left(  \frac{z}{a}\right)  ^{2}\right)  ^{3/2}%
2738: }\psi_{0}. \label{realistic psi}%
2739: \end{equation}
2740: This has a maximum of $\psi_{0}$ at $r=a,$ scales as $r^{2}$ for small $r,$
2741: and scales as $r^{-1}$ for large $r.$ Equation \ref{Jtor} can be used to
2742: calculate the associated toroidal current density which will be sharply peaked
2743: near $r=a$ and $z=0.$ The $\psi(r,z)$ prescribed by Eq.\ref{realistic psi} has
2744: the features that it provides a dipole-like field far from the $z$ axis and a
2745: nearly uniform axial field near the $z$ axis, corresponds to a realistic
2746: distributed toroidal current, has a well-defined poloidal field magnetic axis,
2747: is analytically tractable, and is convenient for numerical computation of
2748: representative particle orbits in a physically relevant magneto-gravitational field.
2749: 
2750: \section{Current associated with flux function
2751: \label{current associated with flux function}}
2752: 
2753: Using Ampere's law to relate the toroidal current and the poloidal magnetic
2754: field it is seen that
2755: \begin{equation}
2756: \mathcal{I}_{\phi}=\frac{1}{\mu_{0}}\oint_{C}\mathbf{B}_{pol}\mathbf{\cdot
2757: }d\mathbf{l} \label{Amperes law}%
2758: \end{equation}
2759: where the contour $C$ links the total toroidal current $\mathcal{I}_{\phi}$.
2760: By letting the line integral go to infinity in the radial and $z$ directions
2761: it is seen that only the portion of the line integral along the $z$ axis makes
2762: a finite contribution so%
2763: \begin{equation}%
2764: \begin{array}
2765: [c]{ccl}%
2766: \mathcal{I}_{\phi} & = & \ \frac{1}{\mu_{0}}\int_{-\infty}^{\infty}%
2767: B_{z}(0,z)dz\\
2768: & = & \frac{27\psi_{0}}{16\pi\mu_{0}}\lim_{r\rightarrow0}\frac{1}{r}%
2769: \frac{\partial\ }{\partial r}\left(  \frac{r^{2}}{a^{2}}\int_{-\infty}%
2770: ^{\infty}\ \frac{dz\ }{\ \left(  \left(  \frac{r}{a}+\frac{1}{2}\right)
2771: ^{2}+\left(  \frac{z}{a}\right)  ^{2}\right)  ^{3/2}}\right)  \ .
2772: \end{array}
2773: \label{Iphi 1}%
2774: \end{equation}
2775: 
2776: 
2777: Defining $b=r/a+1/2$ and $z/a=b\sinh\vartheta$ the $z$ integral can be
2778: expressed as%
2779: \begin{equation}%
2780: \begin{array}
2781: [c]{ccl}%
2782: \int_{-\infty}^{\infty}\ \frac{dz\ }{\ \left(  \left(  \frac{r}{a}+\frac{1}%
2783: {2}\right)  ^{2}+\left(  \frac{z}{a}\right)  ^{2}\right)  ^{3/2}} & = &
2784: a\int_{-\infty}^{\infty}\ \frac{b\cosh\vartheta\,d\vartheta}{\ \left(
2785: b^{2}+b^{2}\sinh^{2}\vartheta\right)  ^{3/2}}\\
2786: & = & \frac{a}{b^{2}}\left[  \tanh\vartheta\right]  _{-\infty}^{\infty}\\
2787: & = & \frac{2a}{\left(  r/a+1/2\right)  ^{2}}.
2788: \end{array}
2789: \label{integral details}%
2790: \end{equation}
2791: Since
2792: \begin{equation}
2793: \lim_{r\rightarrow0}\frac{1}{r}\frac{\partial\ }{\partial r}\left[
2794: \frac{r^{2}}{a^{2}}\frac{2a}{\left(  r/a+1/2\right)  ^{2}}\right]  =\frac
2795: {16}{a} \label{lim integral}%
2796: \end{equation}
2797: the total toroidal current is
2798: \begin{equation}
2799: \mathcal{I}_{\phi}=\frac{27\psi_{0}}{\ \pi a\mu_{0}}.
2800: \label{solve Iphi integral}%
2801: \end{equation}
2802: 
2803: 
2804: \section{Review:\ Distinction between diamagnetic (adiabatic) orbits and
2805: paramagnetic (Speiser) orbits\label{Orbit review}}
2806: 
2807: \subsection{Diamagnetism of cyclotron (Larmor) orbits}
2808: 
2809: We first review charged particle motion in a uniform magnetic field
2810: $\mathbf{B}=B_{z}\hat{z}$ (so $\psi=B_{z}\pi r^{2}$) and no electric field;
2811: orbital motion in more complex fields will be discussed later. The particle
2812: motion is prescribed by the Lorentz equation
2813: \begin{equation}
2814: m_{\sigma}\frac{d\mathbf{v}}{dt}=q_{\sigma}\mathbf{v}\times B_{z}\hat{z}\ .
2815: \label{Lorentz}%
2816: \end{equation}
2817: If the particle is restricted to the $z=0$ plane, the respective radial and
2818: azimuthal components of Eq.\ref{Lorentz} are%
2819: \begin{equation}
2820: m_{\sigma}\left(  \ddot{r}-r\dot{\phi}^{2}\right)  =q_{\sigma}r\dot{\phi}B_{z}
2821: \label{Lorentz r}%
2822: \end{equation}%
2823: \begin{equation}
2824: \frac{m_{\sigma}}{r}\frac{d}{dt}\left(  r^{2}\dot{\phi}\right)  =-q_{\sigma
2825: }\dot{r}B_{z}. \label{Lorentz phi}%
2826: \end{equation}
2827: We consider circular motion (i.e., cyclotron or Larmor orbits) so $\ r=const.$
2828: in which case Eq.\ref{Lorentz phi} gives $\dot{\phi}=const.$ and
2829: Eq.\ref{Lorentz r} then becomes%
2830: \begin{equation}
2831: \dot{\phi}=-\omega_{c\sigma} \label{phidot diamag}%
2832: \end{equation}
2833: where $\omega_{c\sigma}=q_{\sigma}B_{z}/m_{\sigma}$ is the signed cyclotron
2834: frequency. The minus sign in Eq.\ref{phidot diamag} indicates that cyclotron
2835: motion is \textit{diamagnetic}. Thus if a gyrating charged particle is
2836: considered as a $\phi$-directed current, the polarity of this current is such
2837: as to create a magnetic field which opposes the initial field $B_{z}$, i.e.,
2838: cyclotron orbits tend to depress the value of $\psi.$ The diamagnetism of
2839: cyclotron orbits means that cyclotron orbits cannot be the source for the
2840: assumed poloidal magnetic field $\psi(r,z)$ nor the means by which this field
2841: is sustained against dissipation.
2842: 
2843: \subsection{Adiabatic orbits}
2844: 
2845: When the magnetic field is non-uniform or there are electric fields, and if
2846: these additional features are sufficiently weak that to lowest order the
2847: Larmor orbit (cyclotron orbit) description is approximately correct, then
2848: additional charged particle motions occur which are superimposed on the Larmor
2849: orbits $\mathbf{v}_{L}(t);$ these additional motions are adiabatic in the
2850: sense of classical mechanics. Defining\ $v_{\parallel}$ as the velocity
2851: component parallel to the magnetic field and $\mathbf{v}_{\perp}$ as the
2852: component perpendicular to the magnetic field these motions are the standard
2853: drifts
2854: %TCIMACRO{\TeXButton{citep{Longmire1967,Chen1984}}{\citep
2855: %{Longmire1967,Chen1984}}}%
2856: %BeginExpansion
2857: \citep{Longmire1967,Chen1984}%
2858: %EndExpansion
2859: , namely the $E\times B$ drift $\mathbf{v}_{E}=\mathbf{E\times B}/B^{2},$ the
2860: polarization drift $\mathbf{v}_{p}=m_{\sigma}q_{\sigma}^{-1}B^{-2}%
2861: d\mathbf{E}_{\perp}/dt$, the curvature drift $\mathbf{v}_{c}=-m_{\sigma
2862: }v_{\parallel}^{2}\hat{B}\cdot\nabla\hat{B}\times\mathbf{B/}q_{\sigma}B^{2}$,
2863: and the grad $B$ drift $\mathbf{v}_{\nabla B}=-\mu\nabla B\times
2864: \mathbf{B/}q_{\sigma}B^{2}$ where $\mu=m_{\sigma}v_{\perp}^{2}/2B$ is the
2865: magnetic moment, an adiabatic invariant. There is also a `force' drift
2866: $\mathbf{v}_{F}$ $=\mathbf{F\times B}/q_{\sigma}B^{2}$ where $\mathbf{F}$ is a
2867: generic non-electromagnetic force, which here is gravity, so $\mathbf{F}%
2868: =mMG\nabla\left(  r^{2}+z^{2}\right)  ^{-1/2}.$ Taking into account all these
2869: drifts, the velocity of an adiabatic-orbit charged particle becomes%
2870: \begin{equation}%
2871: \begin{array}
2872: [c]{cc}%
2873: \mathbf{v}= & v_{\parallel}\hat{B}+\mathbf{v}_{L\sigma}(t)\ +\frac
2874: {\mathbf{E\times B}}{B^{2}}+\frac{m_{\sigma}}{q_{\sigma}B^{2}}\frac
2875: {d\mathbf{E}_{\perp}}{dt}\ -\frac{m_{\sigma}v_{\parallel}^{2}\hat{B}%
2876: \cdot\nabla\hat{B}\times\mathbf{B}}{q_{\sigma}B^{2}}\\
2877: & -\frac{\mu\nabla B\times\mathbf{B}}{q_{\sigma}B^{2}}+\mathbf{\ }%
2878: \ \frac{m_{\sigma}MG}{q_{\sigma}B^{2}}\nabla\left(  \frac{1}{\sqrt{r^{2}%
2879: +z^{2}}}\right)  \times\mathbf{B.}%
2880: \end{array}
2881: \label{particle drifts}%
2882: \end{equation}
2883: The last four drifts in Eq.\ref{particle drifts} explicitly involve
2884: $q_{\sigma}$ and thus produce macroscopic currents. When these currents are
2885: summed and, in addition, diamagnetic current is taken into account, \ the
2886: result is equivalent to the MHD\ equation of motion where the polarization
2887: drift plays the role of the inertial term
2888: %TCIMACRO{\TeXButton{\citep{Goldston1995,Bellan2006}}{\citep
2889: %{Goldston1995,Bellan2006}}}%
2890: %BeginExpansion
2891: \citep{Goldston1995,Bellan2006}%
2892: %EndExpansion
2893: . The ideal MHD\ concept of frozen-in flux is directly equivalent to $\mu$
2894: conservation because $\mu$ conservation corresponds to conservation of the
2895: magnetic flux linked by a cyclotron orbit. Thus, the ideal MHD\ concept of
2896: frozen-in flux is based on the adiabatic invariance of cyclotron orbits.
2897: 
2898: The poloidal flux function specified by Eq.\ref{generic} corresponds to a
2899: magnetic field generated by a toroidal current flowing in the positive $\phi$
2900: direction (counterclockwise direction); the $B_{z}$ component of this field is
2901: positive for $r<a$ and negative for $r>a$ where $a$ is the location of the
2902: poloidal field magnetic axis. The poloidal magnetic field has both curvature
2903: and gradients so that away from field nulls, particles should have parallel
2904: motion and cyclotron orbits together with superimposed curvature and grad $B$
2905: drifts. Figure \ref{Orbit-Inside-Poloidal-Field-Magnetic-Axis} shows the
2906: numerically calculated orbit of a particle located in the $z=0$ plane in a
2907: magnetic field prescribed by Eq.\ref{generic} and located inside the poloidal
2908: field magnetic axis (indicated by dashed circle). It is seen that the particle
2909: makes cyclotron orbits with a superimposed drift due to curvature and $\nabla
2910: B$. The cyclotron orbit is clockwise consistent with the assertion that
2911: cyclotron motion is diamagnetic. Figure
2912: \ref{Orbit-Outside-Poloidal-Field-Magnetic-Axis} shows the situation for a
2913: particle located at a radius outside the poloidal field magnetic axis. The
2914: sense of the cyclotron orbit is now reversed as is the polarity of $B_{z}$ so
2915: the cyclotron orbit is again diamagnetic. For both inside and outside
2916: particles the drift motion is clockwise and so opposes the original toroidal
2917: current creating the poloidal flux and so the curvature and $\nabla B$ drift
2918: motion can also be considered diamagnetic.
2919: 
2920: \begin{figure}[ptb]
2921: \caption{Orbit of a positive particle in the $z=0$ plane located inside the
2922: poloidal field magnetic axis (indicated by dashed circle), coordinates are
2923: normalized to the poloidal field magnetic axis radius. $B_{z}$ is positive
2924: inside the circle and negative outside.}%
2925: \label{Orbit-Inside-Poloidal-Field-Magnetic-Axis}%
2926: \epsscale{1.0} \plotone{f14.eps}\end{figure}
2927: 
2928: \begin{figure}[ptb]
2929: \caption{Orbit for a positively charged particle located in the $z$ plane
2930: outside the poloidal field magnetic axis (dashed circle).}%
2931: \label{Orbit-Outside-Poloidal-Field-Magnetic-Axis}%
2932: \epsscale{1.0} \plotone{f15.eps}\end{figure}
2933: 
2934: The current associated with the gravitational force drift is%
2935: \begin{equation}
2936: \mathbf{J}_{g}=\sum_{\sigma}n_{\sigma}q_{\sigma}\mathbf{v}_{F}=\ \frac{\varrho
2937: MG}{\ B^{2}}\nabla\left(  \frac{1}{\sqrt{r^{2}+z^{2}}}\right)  \times
2938: \mathbf{B} \label{Jg}%
2939: \end{equation}
2940: where $\varrho=\sum m_{\sigma}n_{\sigma}$ is the mass density.
2941: 
2942: From a macroscopic (i.e., MHD) point of view, the force associated with the
2943: gravitational drift current exactly balances the gravitational force component
2944: perpendicular to the magnetic field since%
2945: 
2946: \begin{equation}%
2947: \begin{array}
2948: [c]{ccl}%
2949: \mathbf{J}_{g}\times\mathbf{B} & = & \ \frac{\varrho MG}{\ B^{2}}\left(
2950: \nabla\left(  \frac{1}{\sqrt{r^{2}+z^{2}}}\right)  \times\mathbf{B}\right)
2951: \times\mathbf{B}\\
2952: & = & -\varrho MG\nabla_{\perp}\left(  \frac{1}{\sqrt{r^{2}+z^{2}}}\right)
2953: \end{array}
2954: \label{grav force}%
2955: \end{equation}
2956: If $I=0$ on the $z=0$ plane (as is consistent with astrophysical jet symmetry
2957: used by
2958: %TCIMACRO{\TeXButton{citet{Lovelace1976}}{\citet{Lovelace1976}}}%
2959: %BeginExpansion
2960: \citet{Lovelace1976}%
2961: %EndExpansion
2962: ), the gravitational drift is not defined on the poloidal field magnetic axis
2963: because $\mathbf{B}$ vanishes on the poloidal field magnetic axis and the
2964: theory of particle drifts fails. In other words, going from the first to the
2965: second line in Eq.\ref{grav force} at the poloidal field magnetic axis would
2966: involve dividing zero by zero (since $B=0$ on the poloidal field magnetic axis).
2967: 
2968: When summed over species, the curvature and grad $B$ drifts correspond to
2969: currents which balance macroscopic pressure gradients (when diamagnetic
2970: current is included) and the polarization current corresponds to the inertial
2971: term in the MHD\ equation of motion. This analysis shows, as discussed in
2972: %TCIMACRO{\TeXButton{\citet{Bellan2007}}{\citet{Bellan2007}}}%
2973: %BeginExpansion
2974: \citet{Bellan2007}%
2975: %EndExpansion
2976: , that plasma particles undergoing cyclotron motion and drifts do not have
2977: Keplerian orbits. It also shows that the poloidal field magnetic axis is a
2978: special place where conventional particle drift theory fails.
2979: 
2980: \subsection{Non-adiabatic motion: the Speiser orbit\label{Speiser}}
2981: 
2982: An extreme form of magnetic non-uniformity occurs where the magnetic field
2983: reverses direction. In this case an orbit quite distinct from the cyclotron
2984: orbit and its associated adiabatic drifts occurs. This non-adiabatic orbit,
2985: called a meandering or Speiser orbit
2986: %TCIMACRO{\TeXButton{\citep{Speiser1965}}{\citep{Speiser1965}}}%
2987: %BeginExpansion
2988: \citep{Speiser1965}%
2989: %EndExpansion
2990: , consists of semi-circles of counterclockwise motion interspersed with
2991: semi-circles of clockwise motion.
2992: 
2993: A numerically calculated Speiser orbit for a positively charged particle in
2994: the $z=0$ plane is shown in Fig.\ref{SpeiserOrbit}. The particle oscillates
2995: across the poloidal field magnetic axis between the inside region where
2996: $B_{z}>0$ and the outside region where $B_{z}<0.$ The result is a net
2997: counterclockwise motion so, in contrast to cyclotron orbits, Speiser orbits
2998: are \textit{paramagnetic.} The paramagnetism of Speiser orbits has been
2999: considered an important aspect of current sheets in Earth's magnetotail,
3000: [e.g., see
3001: %TCIMACRO{\TeXButton{\citet{Zelenyi2000}}{\citet{Zelenyi2000}}}%
3002: %BeginExpansion
3003: \citet{Zelenyi2000}%
3004: %EndExpansion
3005: ], but to the author's knowledge this paramagnetism has not been previously
3006: considered in the axisymmetric three-dimensional geometry discussed here which
3007: is relevant to accretion disks and astrophysical jets. In particular, we will
3008: show that the poloidal flux function can be considered as a consequence of
3009: Speiser orbits such as shown in Fig.\ref{SpeiserOrbit}. \ Speiser orbits are
3010: not consistent with the drift approximation (i.e., $E\times B$ drift, grad $B$
3011: drift, curvature drift, etc.) \ because the drift approximation is based on
3012: the assumption that, to lowest order, the particle is undergoing cyclotron
3013: motion. The inconsistency between Speiser orbits and the drift approximation
3014: is obvious when one considers that the drift approximation fails where $B$
3015: reverses polarity whereas Speiser orbits depend on this reversal.
3016: 
3017: If motion in the $z$ direction is also allowed, then because $B_{r}$ also
3018: reverses at the poloidal field magnetic axis, the particle can also oscillate
3019: vertically across the poloidal field magnetic axis to make vertical Speiser
3020: orbits. The combined $r$ and $z$ Speiser motion means that particles moving at
3021: an arbitrary angle across the poloidal field magnetic axis will reflect from
3022: interior surfaces of the nested poloidal flux surfaces concentric with the
3023: poloidal field magnetic axis. These nested poloidal flux surfaces can thus be
3024: imagined as the walls of a toroidal tunnel and the Speiser orbit particles can
3025: be considered as reflecting from the interior walls of this toroidal tunnel
3026: while moving in the counterclockwise direction to trace out paramagnetic
3027: orbits and create poloidal flux.
3028: 
3029: \begin{figure}[ptb]
3030: \caption{Speiser orbit. The charged particle bounces back and forth across the
3031: field null at the poloidal field magnetic axis resulting in a counterclockwise
3032: (i.e., paramagnetic) orbit.}%
3033: \label{SpeiserOrbit}%
3034: \epsscale{1.0} \plotone{f16.eps}\end{figure}
3035: 
3036: \section{Equation of motion and its solutions\label{Equation of motion}}
3037: 
3038: The Hamiltonian orbit analysis presented here shows that photo-emission
3039: creates new effective potential barriers. The topography of these barriers
3040: depends on a combination of environmental factors, particle properties, and
3041: the location of the charging. Representative orbits obtained by numerically
3042: integrating the equation of motion$\ $have been presented and are consistent
3043: with the predictions of the Hamiltonian theory. We outline here the derivation
3044: of the dimensionless equation of motion; this derivation gives insights into
3045: several fundamental issues regarding the dynamics, especially the influence of
3046: initial conditions.
3047: 
3048: The equation of motion for a charged particle in a combined electromagnetic
3049: and gravitational field is
3050: \begin{equation}
3051: m_{\sigma}\frac{d^{2}\mathbf{x}}{dt^{2}}=q_{\sigma}\left(  \mathbf{E+v\times
3052: B}\right)  +m_{\sigma}MG\nabla\frac{1}{\sqrt{r^{2}+z^{2}}}.
3053: \label{motion em gravity}%
3054: \end{equation}
3055: 
3056: 
3057: Using Eq.\ref{B general2} for the magnetic field, the equation of motion can
3058: thus be written as%
3059: \begin{equation}%
3060: \begin{array}
3061: [c]{cl}%
3062: m_{\sigma}\frac{d^{2}\mathbf{x}}{dt^{2}}= & q_{\sigma}\mathbf{E}\\
3063: & \mathbf{+}\frac{q_{\sigma}}{2\pi}\frac{d\mathbf{x}}{dt}\times\left(
3064: \frac{\partial\psi}{\partial\mathbf{x}}\times\frac{\partial\phi}%
3065: {\partial\mathbf{x}}+\mu_{0}I\frac{\partial\phi}{\partial\mathbf{x}}\right) \\
3066: & +m_{\sigma}MG\nabla\left(  \frac{1}{\sqrt{r^{2}+z^{2}}}\right)  .
3067: \end{array}
3068: \label{eq motion}%
3069: \end{equation}
3070: Then, using the definitions given in Eq.\ref{norm quantities}, the equation of
3071: motion can be expressed in dimensionless form as%
3072: \begin{equation}
3073: \frac{d^{2}\mathbf{\bar{x}}}{d\tau^{2}}=\ \mathbf{\bar{E}}+\frac{\left\langle
3074: \omega_{c\sigma}\right\rangle }{2\Omega_{0}}\frac{d\mathbf{\bar{x}}}{d\tau
3075: }\times\left(  \ \frac{\partial\bar{\psi}}{\ \partial\mathbf{\bar{x}}}%
3076: \times\frac{\partial\phi}{\ \partial\mathbf{\bar{x}}}+\frac{\mu_{0}I}%
3077: {a\pi\left\langle B_{z}\right\rangle }\frac{\partial\phi}{\ \partial
3078: \mathbf{\bar{x}}}\right)  -\frac{\mathbf{\bar{x}}}{\left\vert \mathbf{\bar{x}%
3079: }\right\vert ^{3}} \label{norm eq motion}%
3080: \end{equation}
3081: where
3082: \begin{equation}
3083: \mathbf{\bar{E}}=\frac{q_{\sigma}}{am_{\sigma}\Omega_{0}^{2}}\mathbf{E}%
3084: =-\frac{q_{\sigma}}{a^{2}m_{\sigma}\Omega_{0}^{2}}\frac{\partial V}%
3085: {\partial\mathbf{\bar{x}}}=-\frac{\partial\bar{V}}{\partial\mathbf{\bar{x}}}
3086: \label{Enorm}%
3087: \end{equation}
3088: is the dimensionless electric field and
3089: \begin{equation}
3090: \bar{V}=\frac{aq_{\sigma}V\ }{\ m_{\sigma}MG}\ \label{Vnorm}%
3091: \end{equation}
3092: is the dimensionless electrostatic potential.
3093: 
3094: Equation \ref{norm eq motion} clearly shows that the dynamics change from
3095: being gravitationally dominated to being magnetically dominated according to
3096: the ratio $\left\langle \omega_{c\sigma}\right\rangle /\Omega_{0}$. The
3097: possibility of complex interactions between gravitational and magnetic forces
3098: when $\left\langle \omega_{c\sigma}\right\rangle /\Omega_{0}$ is of order
3099: unity is also evident. The coefficient $\mu_{0}I/a\pi\left\langle
3100: B_{z}\right\rangle $ is related to the pitch of a twisted field. The
3101: Hamilton-Lagrange formalism shows that $I$ plays a subservient role for
3102: particle orbits compared to $\psi$ because canonical angular momentum depends
3103: on $\psi$, not $I$. However, large $I$ increases $\left\vert B\right\vert $
3104: and so contributes to the effective potential $\mu|B|$ thereby providing
3105: additional possibilities for localization. Thus, if $\mu\ \ $is large, a
3106: particle is not only constrained to stay on a constant $\psi$ surface, but
3107: \ is additionally constrained to stay out of regions on this surface where
3108: $\mu|B|\ $is large. If poloidal currents flow, then the associated
3109: $\mathbf{J}_{pol}\mathbf{\times B}_{tor}$ forces drive jets which inflate and
3110: distend the $\psi$ surfaces. Thus, the orbits will depend indirectly on $I$
3111: when the jet dynamics alter the shape of the constant $\psi$ surfaces.
3112: 
3113: Using the relations
3114: \begin{equation}%
3115: \begin{array}
3116: [c]{cl}%
3117: \frac{\partial\phi}{\ \partial\mathbf{\bar{x}}}= & \frac{\hat{\phi}}{\bar{r}%
3118: }=\frac{-\hat{x}\bar{y}+\hat{y}\bar{x}}{\bar{x}^{2}+\bar{y}^{2}}\ \\
3119: \frac{\partial\bar{\psi}}{\ \partial\mathbf{\bar{x}}}= & \frac{\partial
3120: \bar{\psi}}{\ \partial\bar{r}}\hat{r}+\frac{\partial\bar{\psi}}{\ \partial
3121: \bar{z}}\hat{z}=\frac{\partial\bar{\psi}}{\ \partial\bar{r}}\left(  \frac
3122: {\hat{x}\bar{x}+\hat{y}\bar{y}}{\sqrt{\bar{x}^{2}+\bar{y}^{2}}}\right)
3123: +\frac{\partial\bar{\psi}}{\ \partial\bar{z}}\hat{z}\ ,
3124: \end{array}
3125: \label{grad phi psi}%
3126: \end{equation}
3127: the normalized equation of motion can be expressed in Cartesian coordinates as%
3128: \begin{equation}%
3129: \begin{array}
3130: [c]{cl}%
3131: \frac{d^{2}\mathbf{\bar{x}}}{d\tau^{2}}= & -\frac{\partial\bar{V}}%
3132: {\partial\mathbf{\bar{x}}}\\
3133: & +\frac{\left\langle \omega_{c\sigma}\right\rangle }{2\bar{r}\Omega_{0}}%
3134: \frac{d\mathbf{\bar{x}}}{d\tau}\times\left(
3135: \begin{array}
3136: [c]{c}%
3137: \ \frac{\partial\bar{\psi}}{\ \partial\bar{r}}\hat{z}-\frac{\partial\bar{\psi
3138: }}{\ \partial\bar{z}}\frac{\mathbf{\bar{r}}}{\bar{r}}\smallskip\\
3139: +\frac{\mu_{0}I}{a\pi\left\langle B_{z}\right\rangle }\frac{\left(  -\hat
3140: {x}\bar{y}+\hat{y}\bar{x}\right)  }{\bar{r}}%
3141: \end{array}
3142: \right) \\
3143: & -\ \ \frac{\left(  \mathbf{\bar{r}}+\bar{z}\hat{z}\right)  }{\left\vert
3144: \bar{r}^{2}+\bar{z}^{2}\right\vert ^{3/2}}%
3145: \end{array}
3146: \label{motion Cartesian}%
3147: \end{equation}
3148: where $\mathbf{\bar{r}}=\bar{x}\hat{x}+\bar{y}\hat{y}\ $ and $\bar{r}%
3149: =\sqrt{\bar{x}^{2}+\bar{y}^{2}}$. Equation \ref{motion Cartesian} is in a form
3150: suitable for numerical computation and has been used to provide the orbital
3151: plots shown earlier.
3152: 
3153: At this point it is convenient to use the generic poloidal flux function given
3154: by Eq.\ref{generic} so the unity-maximum, dipole-like, normalized flux
3155: function will be
3156: \begin{equation}
3157: \bar{\psi}(\bar{r},\bar{z})=\ \ \frac{27\bar{r}^{2}}{8\left(  \left(  \bar
3158: {r}+\frac{1}{2}\right)  ^{2}+\bar{z}^{2}\right)  ^{3/2}}\ \label{norm generic}%
3159: \end{equation}
3160: with%
3161: \begin{equation}
3162: \frac{\partial\bar{\psi}}{\ \partial\bar{r}}=\ \frac{27\bar{r}\left(
3163: \ \bar{r}+1+4\bar{z}^{2}-\ 2\bar{r}^{2}\right)  }{16\left(  \left(  \bar
3164: {r}+\frac{1}{2}\right)  ^{2}+\bar{z}^{2}\right)  ^{5/2}} \label{dpsibydr}%
3165: \end{equation}
3166: and
3167: \begin{equation}
3168: \frac{\partial\bar{\psi}}{\ \partial\bar{z}}=-\ \ \frac{81\bar{r}^{2}\bar{z}%
3169: }{8\left(  \left(  \bar{r}+\frac{1}{2}\right)  ^{2}+\bar{z}^{2}\right)
3170: ^{5/2}}. \label{dpsibydz}%
3171: \end{equation}
3172: Thus, the normalized poloidal magnetic field components are
3173: \begin{equation}
3174: \bar{B}_{r}=\frac{81\bar{r}\bar{z}}{16\pi\left(  \left(  \bar{r}+\frac{1}%
3175: {2}\right)  ^{2}+\bar{z}^{2}\right)  ^{5/2}} \label{Br norm}%
3176: \end{equation}
3177: and
3178: \begin{equation}
3179: \bar{B}_{z}=\frac{27\left(  \ \bar{r}+1+4\bar{z}^{2}-\ 2\bar{r}^{2}\right)
3180: }{32\pi\left(  \left(  \bar{r}+\frac{1}{2}\right)  ^{2}+\bar{z}^{2}\right)
3181: ^{5/2}}. \label{Bz norm}%
3182: \end{equation}
3183: 
3184: 
3185: We now consider the problem of establishing appropriate initial conditions for
3186: an incoming neutral particle.\ For purposes of starting a computation we
3187: assume the particle is located at some initial radial position $\bar{\rho}%
3188: _{0}$ in the orbital plane such that $\bar{\rho}_{0}>\bar{\rho}_{pericenter}$
3189: where $\bar{\rho}_{pericenter}~\ $is given by Eq.\ref{perigee}. Solving
3190: Eq.\ref{nondim} for the initial inward radial velocity gives%
3191: \begin{equation}
3192: \bar{v}_{\rho0}=-\sqrt{2\bar{H}-\frac{\bar{L}^{2}}{\bar{\rho}_{0}^{2}}%
3193: +\frac{2\ }{\bar{\rho}_{0}}}. \label{initial vrhro}%
3194: \end{equation}
3195: \bigskip and the corresponding initial orbital frame azimuthal velocity is
3196: \begin{equation}
3197: \bar{v}_{\eta0}=\frac{\bar{L}}{\bar{\rho}_{0}}.
3198: \label{initial orbital azimuthal}%
3199: \end{equation}
3200: 
3201: 
3202: Equation \ref{solution} can be solved for the initial polar angle in the
3203: orbital frame as
3204: \begin{equation}
3205: \eta=\alpha+\cos^{-1}\left(  \frac{1-\bar{L}^{2}/\bar{\rho}_{0}}{\sqrt
3206: {1+2\bar{L}^{2}\ \bar{H}}}\right)  . \label{initial eta}%
3207: \end{equation}
3208: Using Eq.\ref{unbounded Cartesian} the initial orbital frame Cartesian
3209: coordinates are thus%
3210: \begin{equation}%
3211: \begin{array}
3212: [c]{ccl}%
3213: \bar{x}^{\prime} & = & \frac{\bar{L}^{2}\cos\eta_{0}}{1-\sqrt{1+2\bar{L}%
3214: ^{2}\ \bar{H}}\cos\left(  \eta_{0}-\alpha\right)  }\\
3215: \bar{y}^{\prime} & = & \frac{\bar{L}^{2}\sin\eta_{0}}{1-\sqrt{1+2\bar{L}%
3216: ^{2}\ \bar{H}}\cos\left(  \eta_{0}-\alpha\right)  }\\
3217: \bar{z}^{\prime} & = & 0.
3218: \end{array}
3219: \label{initial orbital Cartesian}%
3220: \end{equation}
3221: The orbital frame Cartesian velocity components are related to the orbital
3222: frame cylindrical velocity components by%
3223: \begin{equation}%
3224: \begin{array}
3225: [c]{cl}%
3226: \bar{v}_{x^{\prime}0}= & \bar{v}_{\rho0}\cos\eta_{0}-\bar{v}_{\eta0}\sin
3227: \eta_{0}\\
3228: \bar{v}_{y^{\prime}0}= & \bar{v}_{\rho0}\sin\eta_{0}+\bar{v}_{\eta0}\cos
3229: \eta_{0}\\
3230: \bar{v}_{z^{\prime}0}= & 0.
3231: \end{array}
3232: \label{initial orbital vel}%
3233: \end{equation}
3234: We now take into account that the orbital frame Cartesian coordinate system is
3235: rotated by the angle of inclination $\theta$ about the $x$ axis with respect
3236: to the lab frame coordinate system. The $x$ and $x^{\prime}$ components of
3237: both position and velocity are the same in the two frames but the $y$ and $z$
3238: components are related by \
3239: \begin{equation}%
3240: \begin{array}
3241: [c]{cl}%
3242: \bar{y} & =-\bar{z}^{\prime}\sin\theta+\bar{y}^{\prime}\cos\theta\\
3243: \bar{z} & =\bar{z}^{\prime}\cos\theta+\bar{y}^{\prime}\sin\theta.
3244: \end{array}
3245: \label{oribtal to lab transformation}%
3246: \end{equation}
3247: Since $\bar{z}^{\prime}$ is by definition zero in the orbital frame, the
3248: initial lab frame Cartesian coordinates are%
3249: \begin{equation}%
3250: \begin{array}
3251: [c]{cl}%
3252: \bar{x}_{0}= & \bar{x}_{0}^{\prime}\\
3253: \bar{y}_{0}= & \bar{y}_{0}^{\prime}\cos\theta\\
3254: \bar{z}_{0}= & \bar{y}_{0}^{\prime}\sin\theta.
3255: \end{array}
3256: \label{initial lab Cartesian coord}%
3257: \end{equation}
3258: Since $v_{z}^{\prime}\ $is similarly zero in the orbital frame, in analogy to
3259: Eq.\ref{initial lab Cartesian coord}, the initial lab frame Cartesian
3260: velocities are%
3261: \begin{equation}%
3262: \begin{array}
3263: [c]{cl}%
3264: \bar{v}_{x0}= & \bar{v}_{x^{\prime}0}\\
3265: \bar{v}_{y0}= & \bar{v}_{y^{\prime}0}\cos\theta\\
3266: \bar{v}_{z0}= & \bar{v}_{y^{\prime}0}\sin\theta.
3267: \end{array}
3268: \label{initial lab Cartesian vel}%
3269: \end{equation}
3270: 
3271: 
3272: Thus, if one wishes to start the numerical computation at the radius
3273: $\bar{\rho}_{0}\,$on the trajectory of an incoming particle with orbit
3274: parameters $\{\bar{H}$,$\bar{L},\theta,\alpha\},$ Eqs.\ref{initial vrhro},
3275: \ref{initial orbital azimuthal}, \ref{initial eta},
3276: \ref{initial lab Cartesian coord} and \ref{initial lab Cartesian vel} give the
3277: appropriate initial position and velocity lab frame Cartesian components.
3278: Before charging, the orbits are degenerate with respect to choice of $\theta$
3279: or $\alpha,$ but after charging there is a strong dependence on these two
3280: angles. In particular, if $0\leq\theta<90^{0}$ the orbit is prograde and
3281: Speiser type orbits are possible if the charging occurs near the poloidal
3282: field magnetic axis. On the other hand if $90^{0}<\theta\leq180^{0}$ the orbit
3283: is retrograde and drain-hole orbits are possible. Thus, a subclass of prograde
3284: incident neutral particles transform upon charging into the
3285: toroidal-current/poloidal-field dynamo while a subclass of retrograde neutral
3286: particles transform upon charging into the poloidal-current/toroidal-field
3287: dynamo that drives a bipolar astrophysical jet. Because $\theta$ and $\alpha$
3288: also affect the angle between the velocity vector and the magnetic field at
3289: charging, $\theta$ and $\alpha$ affect the value of $\mu$ and hence the extent
3290: to which accreted particles with cyclotron orbits will be mirror trapped to
3291: subregions of constant $\psi$ surfaces. For example, if $\alpha=0$ then
3292: variation of the angle of inclination $\theta$ for a given $\ $ $\bar{\rho
3293: }_{pericenter},$ and charging radius $\bar{R}_{\ast}$ will determine whether
3294: the charged particles created upon disintegration of an incoming neutral
3295: particle will be normal trapped particles, untrapped particles, drain-hole
3296: particles, or Speiser particles.
3297: 
3298: \pagebreak
3299: 
3300: \bibliographystyle{authordate1}
3301: \bibliography{ApJ-gravitydynamoJuly2006}
3302: \ 
3303: \end{document}