1: \documentclass[12pt,eqsecnum,preprint,flushrt]{aastex}
2: %\documentclass[10pt,eqsecnum,manuscript,flushrt]{aastex}
3: \slugcomment{Submitted to ApJ}
4: \usepackage{graphicx}
5: %\usepackage{amsmath}
6: \usepackage{vmargin}
7: \usepackage{amssymb}
8:
9: \def\pd#1#2{\frac{\partial #1}{\partial #2}}
10: \def\od#1#2{\frac{d #1}{d #2}}
11: \def\pdd#1#2{\frac{\partial^2 #1}{\partial #2^2}}
12: \def\odd#1#2{\frac{d^2 #1}{d #2^2}}
13: \newcommand\Partial{\boldsymbol{\partial}}
14: \newcommand{\Del}{\mathbf{\nabla}}
15: \newcommand{\Cross}{\boldsymbol{\times}}
16:
17: \def\ahalf{{\frac{1}{2}}}
18: \def\oneover#1{{\frac{1}{#1}}}
19: \def\prn#1{{\left(#1\right)}}
20: \def\brk#1{{\left[#1\right]}}
21: \def\brc#1{{\left\{#1\right\}}}
22: \def\abs#1{{\left|#1\right|}}
23: \newcommand{\sech}{\text{sech}}
24: \newcommand{\xhat}{\mathbf{\hat x}}
25: \newcommand{\yhat}{\mathbf{\hat y}}
26: \newcommand{\zhat}{\mathbf{\hat z}}
27: \newcommand{\rhat}{\mathbf{\hat r}}
28: \newcommand{\thetahat}{\mathbf{\hat \theta}}
29: \newcommand{\varphihat}{\mathbf{\hat \varphi}}
30:
31: \newcommand{\text}[1]{\rm{#1}}
32: \newcommand{\eqref}[1]{(\ref{#1})}
33: \newcommand{\iint}{\int\!\!\!\int}
34:
35: \title{X-winds in Action}
36: \author{Mike J. Cai$^1$, Hsien Shang$^1$, Hsiao-Hsuan Lin$^{2}$, Frank H. Shu$^3$
37: } \affil{$^1$Academia Sinica, Institute of Astronomy
38: and Astrophysics, Taiwan\\
39: $^2$Department of Physics, University of Southern California\\
40: $^3$Department of
41: Physics, University of California, San Diego}
42: \email{mike@asiaa.sinica.edu.tw}
43:
44: \begin{abstract}
45: The interaction of accretion disks with the magnetospheres of young
46: stars can produce X-winds and funnel flows. With the assumption of
47: axial symmetry and steady state flow, the problem can be formulated
48: in terms of quantities that are conserved along streamlines, such as
49: the Bernoulli integral (BI), plus a partial differential equation
50: (PDE), called the Grad-Shafranov equation (GSE), that governs the
51: distribution of streamlines in the meridional plane. The GSE plus
52: BI yields a PDE of mixed type, elliptic before critical surfaces
53: where the flow speed equals certain characteristic wave speeds are
54: crossed and hyperbolic afterward. The computational difficulties
55: are exacerbated by the locations of the critical surfaces not being
56: known in advance. To overcome these obstacles, we consider a
57: variational principle by which the GSE can be attacked by
58: extremizing an action integral, with all other conserved quantities
59: of the problem explicitly included as part of the overall
60: formulation. To simplify actual applications we adopt the cold
61: limit of a negligibly small ratio of the sound speed to the speed of
62: Keplerian rotation in the disk where the X-wind is launched. We
63: also ignore the obstructing effects of any magnetic fields that
64: might thread a disk approximated to be infinitesimally thin. We then
65: introduce trial functions with adjustable coefficients to minimize
66: the variations that give the GSE. We tabulate the resulting
67: coefficients so that other workers can have analytic forms to
68: reconstruct X-wind solutions for various astronomical,
69: cosmochemical, and meteoritical applications.
70: \end{abstract}
71:
72: \keywords{stars: pre-main-sequence; winds; ISM: accretion disks;
73: jets and outflows; MHD}
74: \begin{document}
75: \maketitle
76: \section{Introduction}
77:
78: Accretion, disks, and jets are ubiquitous in astrophysics (see,
79: e.g., Blandford \& Rees 1992). A consensus has been reached that an
80: extra needed ingredient to obtain outflow from inflow is the
81: presence of strong magnetic fields that thread a disk conventionally
82: assumed to be rotating at Keplerian speeds about a central
83: gravitating object, taken in this paper to be a newly born star.
84: Differences come in ascribing the origin of the magnetic fields in
85: the disk itself or in the central star (K\"onigl \& Pudritz 2000,
86: Shu et al. 2000).
87:
88: Disk winds have been extensively studied, both analytically via the
89: assumptions of self-similarity in 2-D space for axisymmetric,
90: time-independent flows (e.g., Blandford \& Payne 1982; Contopoulos
91: \& Lovelace 1994) or by taking advantage of arbitrary variations of
92: the gas pressure (e.g., Tsinganos \& Trussoni 1991) or by studying
93: the asymptotic properties of the collimation (Heyvaerts \& Norman
94: 1997); and numerically by finite-element methods attacking the
95: axisymmetric, time-independent, Grad-Shafranov equation (e.g., in
96: the relativistic regime by Camenzind 1987) or by finite-difference
97: treatments of the time-dependent equations of ideal MHD in 2- and
98: 3-D (e.g., Uchida \& Shibata 1986, Pudritz et al. 2006). For a
99: review of these types of calculations, see Ferreira (2004).
100:
101: The most highly developed semi-analytic theory for the second
102: viewpoint is called X-wind theory, in which fast jets arising in
103: young stellar objects (YSOs) owe their existence to the interaction
104: of the accretion disk with the magnetosphere of the central star.
105: The interaction of accretion disks with strongly magnetized central
106: stars has also been studied numerically (e.g., Goodson, Bohm, \&
107: Winglee 1999; Long et al. 2005, Ustyugova et al. 2006). Although
108: both funnel flows and X-like winds have been found, they have yet to
109: appear simultaneously in numerical simulations, probably because the
110: numerical calculations have not yet proceeded to steady state where
111: the condition of disk-locking applies (Shu et al. 1994). Pure
112: X-wind theory assumes for simplicity that the disk itself is
113: unmagnetized, in fact, all that is needed for the theory to work is
114: for open field lines to be concentrated in a narrow annulus near the
115: inner edge of an accretion disk.
116:
117: Recently, Bacciotti et al. (2002) and Coffey et al. (2004)
118: identified jet rotation in four T Tau systems, DG Tau, TH 28, RW
119: Aur, and LkH$\alpha$ 321, of an amount too large to be compatible
120: with X-winds, but consistent with launching from disks at radii of
121: 0.5-2 AU. Later, Cabrit et al. (2006) showed from mm-wave radio
122: measurements that the disk rotation in RW Aur is actually in the
123: opposite sense to that deduced for the jet from optical lines.
124: Moreover, Pety et al. (2006) find that HH 30, which is observed
125: nearly edge-on and therefore should have had the clearest signature
126: for jet rotation, showed no evidence for outflow rotation at mm
127: wavelengths, a conclusion reinforced by optical and ultraviolet
128: observations of the HH 30 jet by Coffey et al. (2007). While the
129: positive results remain for the three other systems, the case of HH
130: 30, where longitudinal velocities occur in the direction transverse
131: to the line of sight, suggests that the slight line asymmetries in
132: the other cases may be more associated with unequal jumps in the
133: velocity of shocked, high-speed jets, than to the rotation of
134: collimated outflows.
135:
136: In contrast, no one has proposed any explanation other than X-winds
137: for the correlated inflow-outflow signatures seen in SU Aur by
138: Giampapa et al. (1993) and Johns \& Basri (1995). Apart from SZ 68
139: (Johns-Krull \& Hatzes 1997), we are unaware of any other T Tau star
140: that shows a tilted-dipole magnetic-field geometry, and it could be
141: that the dipole component is small on the surface of most T Tau
142: stars (Johns-Krull 2007). Fortunately, Mohanty \& Shu (2007) show
143: that while funnel flows are sensitive to the detailed assumptions
144: made concerning multi-pole structure on the surfaces of the central
145: stars, the properties of the X-wind depend mostly only on the amount
146: of trapped flux in the X-region (see also the observational evidence
147: relating to this point collected by Johns-Krull \& Gafford 2002).
148:
149: Recent calculations show that YSOs are unlikely to lose enough
150: magnetic flux in the process of gravitational collapse to make the
151: level of magnetization ignorable in the resultant circumstellar
152: accretion disks (Galli et al. 2006; Shu et al. 2006, 2007). Indeed,
153: the disks are sufficiently magnetized in many cases that, in
154: quasi-steady state, they rotate at sub-Keplerian rates until the the
155: inner disk-edge is reached. Thus, there are open questions of how
156: much of the trapped flux near the inner edge is to be attributed to
157: the central star versus the disk, and how such disks reacquire
158: near-Keplerian rates of rotation at their inner edges. We ignore
159: these complications in the present study of the X-wind phenomenon,
160: but we note that the methods introduced here are easily modified to
161: attack the more complex problem when the accretion disk interacting
162: with a stellar magnetosphere is itself strongly magnetized.
163:
164: The original X-wind model supposed the outflow to occur from the
165: equator of a magnetized star spun to breakup by a presence of an
166: accretion disk that abutted its surface (Shu et al. 1988). Later, in
167: order to accommodate the slow rotators, such as the classical T
168: Tauri stars which are only rotating at one tenth of breakup (Vogel
169: \& Kuhi 1981, Bouvier et al. 1991, Edwards et al. 1993), Shu et al.
170: (1994a) generalized the X-wind picture to include the case of
171: relatively low accretion when the magnetosphere of the star would
172: truncate the accretion disk at an inner edge before the disk reached
173: the stellar surface (typically a circle of radius 0.2 on the scale
174: of Fig.1, where the disk's inner edge is taken to be at $\varpi =
175: 1$). In a quasi-steady-state where most of the mass of the central
176: star is built up by disk accretion, the magnetic coupling between
177: the star and the disk regulates the star to corotate at the
178: Keplerian frequency at the truncation radius. For a protostar with
179: magnetic dipole moment $\mu_*$, mass $M_*$, mass accretion rate
180: $\dot M_D$, Ostriker \& Shu (1995) estimate this radius to be
181: \begin{equation}
182: R_X = \Phi_{\text{dx}}^{-4/7} \prn{\frac{\mu_*^4}{GM_*\dot
183: M_D^2}}^{1/7},
184: \end{equation}
185: where $\Phi_{\text{dx}}$ is an order unity dimensionless number that
186: parameterizes the amount of stellar magnetic flux that is trapped in
187: the disk. Inside this radius, matter is channeled to the star via a
188: funnel flow. The excess angular momentum of the
189: accreting material is deposited in the magnetic field in the form of
190: Maxwell torque, and then transported back to the disk. The gain of
191: angular momentum and approximate field freezing would try to move
192: the footpoint of the funnel-flow field lines outward.
193:
194: Exterior to the truncation radius $R_X$, the equatorial inward drift
195: in the accretion disk creates an angle between the stellar
196: magnetic-field lines and the disk normal. If approximate field
197: freezing holds as the accretion proceeds, a fraction of the field
198: lines will develop an angle larger than $30^\circ$, when matter
199: frozen to this flux tube becomes unstable to magnetocentrifugal
200: fling (Blandford \& Payne 1982). These field lines are thus
201: responsible for driving a magnetohydrodynamic (MHD) wind from the
202: disk. Since the wind removes angular momentum from the disk, the
203: footpoints of those field lines in the disk will try to migrate
204: inward. The radially inward press of the footpoints of the wind
205: field lines footpoints and the radially outward press of the
206: footpoints of the funnel-flow field lines create a magnetic
207: X-configuration that distinguishes the model from similar variants
208: in the literature (the historical choice of the name from the
209: X-point of the equivalent gravitational potential in the co-rotating
210: frame is common to many models). In quasi-steady state where radial
211: advection into the X is balanced by the resistive diffusion of field
212: lines out of the X, Shu et al. (1994b) estimated that enough stellar
213: flux could be trapped in a small X-region near the inner disk edge
214: to have large dynamical effects, namely, the truncation of the disk
215: by a funnel flow out of the disk plane accompanied by an X-wind that
216: carries away most of the excess angular momentum transported into
217: the X-region.
218:
219: Apart from the original numerical estimates, there are reasons to
220: suppose that if turbulent resistivity is the source of the diffusion
221: across magnetic field lines (Shu et al. 2007), then the fractional
222: size of the X-region in units of $R_X$ is given by the ratio of
223: sound speed at the surface of the disk where the X-wind is launched
224: to the local Keplerian speed at $R_X$. For the inner disk of a
225: classical T Tauri star, the thermal sound speed is $a\sim 5$ km/s
226: while the Keplerian speed at $R_X$ is $v_K \sim 100$ km/s (Najita et
227: al. 2007); thus, the ratio $\epsilon$ is a small number $\sim 0.05$.
228: In an asymptotic analysis where $\epsilon$ is taken to $\rightarrow
229: 0$, the X-wind tied to the trapped field lines in the X-region would
230: emerge from virtually a single point in the meridional plane with a
231: fan-like geometry. Seen by an observer rotating at the Keplertian
232: angular frequency of $R_X$, gas flows along streamlines that
233: coincide with field lines if field freezing is assumed, and both
234: patterns of streamlines and field lines remain stationary in the
235: corotating frame.
236:
237: \begin{figure}[ht]
238: \begin{center}
239: \includegraphics[width = 4in, angle = 0]{f1.ps}
240: \end{center}
241: \caption{Funnel flow (red curves), X-wind streamlines (blue curves),
242: and field lines dead to magnetocentrifugal fling (black curves)
243: according to Ostriker \& Shu (1995) and Shang et al. (1998). The
244: magnetic field near the origin (center of YSO) is modeled as a
245: magnetic dipole, and all of the field lines contained in the X-wind
246: have their counterparts (with reversed directions) in opened stellar
247: field lines that lie inside a hollow cone dead to flow surrounding
248: the $z$-axis. Exact pressure balance across the sheet current that
249: divides the X-wind and dead field lines holds near and far from the
250: Y-point at $\varpi \approx 1.3$, $z\approx 0.7$, but this balance is
251: only approximate at intermediate distances, which accounts for why
252: the tilted upside-down Y of the separatrix does not have the equal
253: angles of 120$^\circ$ that would characterize an exact
254: Y-configuration appropriate for coronal conditions. In fact, the
255: field lines in red and black are computed as if they were vacuum
256: magnetic fields; only the portion depected in blue are attacked via
257: the solution of the Grad-Shafranov equation or its
258: variational-principle analog discussed in the present paper. The
259: neutral line separating the black and blue field lines is replaced
260: by a separatrix of prescribed locus satisfying the approximate
261: pressure balance as described above. (see \S 4).} The implied
262: poloidal and toroidal current flows are discussed in Ostriker \& Shu
263: (1995) and Shu et al. (1995); in particular, there is no current
264: flow along the $z$-axis which is taken to be a region of vacuum
265: longitudinal field. \label{funnel-wind}
266: \end{figure}
267:
268: Viewed in this fashion, the overall problem can be broken into
269: smaller pieces and tackled separately. Using a formulation with a
270: precedent in the work of Lovelace et al. (1986), Shu et al. (1988,
271: 1994b) wrote down the mathematical equations that describe a steady
272: state axisymmetric flow in the corotating frame (Grad \& Rubin 1958,
273: Shafranov 1966). By the method of matched asymptotic expansions,
274: Shu et al. demonstrated the existence of an inner solution in the
275: X-region where the flow makes a sonic transition, and an outer
276: solution where the sound speed is formally taken to zero and the
277: X-region shrinks down to a point. Najita \& Shu (1994) computed
278: numerically the portion of the X-wind in which the fluid velocity is
279: sub-Alfv\'enic, and the governing equation is elliptic. Ostriker \&
280: Shu (1995) solved the problem of the funnel flow and the field
281: configuration in the dead zone (in which the field lines do not
282: depart sufficiently from disk normal to load any matter) in an
283: approximation that treated the accretion flow onto the star as
284: highly sub-Alfv\'enic. Shu et al. (1995) constructed asymptotic
285: solutions that describe the logarithmically slow, far-wind
286: collimation into jets. The free boundaries between various parts of
287: the problem (funnel flow, dead zone, and X-wind) were determined by
288: pressure balance on either side.
289:
290: In this paper, we wish to address the X-wind part of the overall
291: problem. In order for the X-wind to accelerate from rest to
292: supersonic speeds, it must smoothly pass three surfaces on which the
293: flow velocity is equal to the slow MHD, Alfv\'en, and fast MHD
294: velocity, respectively (Heinemann \& Olbert 1978, Sakurai 1985).
295: These critical surfaces manifest themselves as singularities in the
296: governing equation (see Weber \& Davis 1967), and thus need to be
297: handled analytically. In an axisymmetric problem, if the shapes of
298: the streamlines are known in the meridional plane, the conserved
299: quantities of the problem (mass to flux loading, angular momentum
300: flux including that carried in the Maxwell stress, and Bernoulli's
301: integral along a streamline) suffice to give a completely analytic
302: solution, including the locations and conditions required to cross
303: the critical surfaces smoothly. Unfortunately, the streamline
304: distribution in the meridional plane is not known a priori but must
305: be obtained, in principle, from a solution of the Grad-Shafranov
306: equation (GSE). The spatial location of the critical surfaces are
307: part of the overall solution of the GSE; indeed, they characterize
308: the regions where this PDE is elliptic or hyperbolic. The mixed
309: character of the GSE makes a direct numerical attack extremely
310: difficult when self-similarity does not apply, perhaps the hardest
311: problem in the mathematical theory of nonlinear PDEs of second-order
312: (see Garabedian 1986). The current work sidesteps the mathematical
313: solution of the GSE as a nonlinear PDE of second order, and
314: approaches it instead as a much more amenable problem in variational
315: calculus.
316:
317: The rest of this paper is organized as following. In \S 2 we review
318: the basic formulation in terms of a stream function and an Alfv\'en
319: discriminant that yields the partial-differential equations -- the
320: so-called Grad-Shafranov and Bernoulli equations -- that govern a
321: steady, axisymmetric, X-wind flow. In \S 3 we write down an action
322: whose variations with respect the stream function $\psi(\varpi,z)$
323: and the Alfv\'en discriminant ${\cal A}(\varpi,z)$ yield,
324: respectively, the Grad-Shafranov equation and the Bernoulli equation
325: when the action reaches an extremum. We also perform a
326: transformation where we replace the vertical coordinate $z$ in
327: cylindrical coordinates $(\varpi, \varphi, z)$ by $\psi$. In \S 4,
328: we show how to incorporate boundary conditions into the problem, as
329: well as how to take advantage of the fact that analytic forms are
330: known for the solutions in the near-neighborhood of the X-point and
331: in the asymptotic regime far from the X-point (Shu et al. 1994b,
332: 1995). In \S 5, we outline a practical implementation of the
333: principle of extremal action, making use of only variations of
334: $\psi$ -- or, more precisely, of $z(\varpi,\psi)$ in our actual
335: working space -- as the substitute to attacking the Grad-Shafranov
336: equation, while we solve Bernoulli's equation directly for reasons
337: that are expounded upon in this section. In \S 6, we present
338: numerical results for three specific cases of mass-loading onto wind
339: flux-tubes, finding good agreement with previous approximate
340: solutions obtained by Shang (1998) that have been used for many
341: different astrophysical applications (e.g., Shang et al. 1998, 2002,
342: 2004). In \S 7, we summarize the recipes needed to convert the
343: numerical solutions of \S 6 into practical dimensional models. We
344: then offer our conclusions and suggestions for needed future
345: research.
346:
347: \section{Basic Equations} \label{Basic Equations}
348: \newcommand{\Veff}{V_{\text{eff}}}
349: \newcommand{\A}{\mathcal{A}}
350: From the fundamental parameters of the problem, we may construct
351: units of length, time, and density as $R_X$, $\Omega_X^{-1}$, and
352: $\dot M_w/4\pi R_X^3\Omega_X$, respectively. By assuming
353: axisymmetry and stationarity in a frame that is rotating with
354: angular velocity $\Omega_X$, we may write down the dimensionless
355: governing equations in the above units.
356: \begin{mathletters}
357: \begin{eqnarray}
358: &&\Del \cdot (\rho \mathbf{u}) = 0, \label{continuity_eqn}\\
359: &&\Del \prn{\ahalf \abs{\mathbf{u}}^2} + (2 \mathbf{e}_z + \Del
360: \times \mathbf{u}) \times \mathbf{u} = -\frac{\epsilon}{\rho} \Del \rho
361: -\Del \Veff + \oneover{\rho} (\Del \times \mathbf{B}) \times \mathbf{B},
362: \label{Euler_eqn}\\
363: &&\mathbf{B} \times \mathbf{u} = 0,\label{field_freezing}\\
364: &&\Del \cdot \mathbf{B} = 0, \label{monopole}
365: \end{eqnarray}
366: \end{mathletters}
367: where $\epsilon \equiv a/R_X \Omega_X$ is the sound speed measured
368: in units of Keplerian velocity at the X-point, and is assumed to be
369: a small parameter of the problem. The effective potential in the
370: corotating frame is defined as
371: \begin{equation}
372: \Veff = -\frac{1}{\sqrt{\varpi^2 + z^2}} -\ahalf \varpi^2 +
373: \frac{3}{2}.
374: \label{Veff}
375: \end{equation}
376: Here we have added a constant term to the effective potential so
377: that its numerical value is zero at the X-point.
378:
379: \subsection{Constants of Motion}
380: \newcommand{\oneoverrho}{\beta^2 - \varpi^2\A}
381: The continuity equation \eqref{continuity_eqn} is satisfied
382: identically if we define the poloidal velocity through a stream
383: function (Shu et al. 1988, 1994a):
384: \begin{equation}
385: \rho u_\varpi \equiv \oneover{\varpi} \pd{\psi}{z}, \quad \rho
386: u_z \equiv - \oneover{\varpi} \pd{\psi}{\varpi}.
387: \label{psidef}
388: \end{equation}
389: For steady state axisymmetric flow in the corotation frame, the
390: field freezing condition \eqref{field_freezing} demands that the
391: magnetic field and the velocity are related by (see, e.g., Mestel 1968)
392: \begin{equation}
393: \mathbf{B} = \beta \rho \mathbf{u}.
394: \label{betadef}
395: \end{equation}
396: With this identification, the continuity equation
397: \eqref{continuity_eqn} and the absence of magnetic monopoles
398: \eqref{monopole} imply $\mathbf{u} \cdot \Del \beta = 0$. In terms
399: of the stream function, this means $\beta$ is constant along each
400: streamline, or $\beta = \beta(\psi)$.
401:
402: The Euler equation describes momentum and energy balance in three spatial
403: dimensions. If we take the component along the fluid velocity by
404: taking the inner product of \eqref{Euler_eqn} with $\mathbf{u}$, we
405: obtain the Bernoulli's equation (BE) along streamlines
406: \begin{equation}
407: \mathbf{u} \cdot \Del H = 0 \; \text{where} \; H \equiv \ahalf
408: \abs{\mathbf{u}}^2 + \epsilon^2 \ln \rho + \Veff.
409: \end{equation}
410: In other words, $H = H(\psi)$, and the energy per unit mass of an
411: isothermal gas, including its specific enthalpy, is conserved along
412: a streamline in the corotating frame where the flow occurs parallel
413: to $\bf B$. Similarly, if we take the toroidal component of the
414: Euler equation \eqref{Euler_eqn}, we obtain a third conserved
415: quantity along stream lines, the angular momentum of the gas
416: allowing for that part carried away by the Maxwell torque of the
417: field:
418: \begin{equation}
419: J \equiv \varpi^2 + \varpi(1-\beta^2 \rho) u_\varphi = J(\psi).
420: \label{Jdef}
421: \end{equation}
422: %The existence of three conserved quantities is no accident. It is
423: %well known from Noether's theorem that symmetries imply conserved
424: %quantities along particle's trajectory. In the context of X-wind,
425: %we have identified three symmetries: axisymmetry, stationarity, and
426: %field freezing. The corresponding integrals of motion are $J$
427: %(which is composed of angular momentum from the gas and Maxwell
428: %torque from the field), $H$ (which consists of the Jacobi's constant
429: %and specific enthalpy), and $\beta$ (which describes mass loading
430: %onto field lines).
431: As we shall see, the determination of the conserved quantities,
432: $H(\psi)$and $J(\psi)$, is achieved by demanding that the X-wind
433: crosses the slow MHD and fast MHD surfaces smoothly. The loading of
434: mass onto flux, which is governed by $\beta(\psi)$ is freely
435: specifiable within certain limits to be detailed below.
436:
437: The last component of the Euler equation describes momentum balance
438: in the direction perpendicular to the poloidal field lines. It is
439: the famous Grad-Shafranov equation (Heinemann \& Olbert 1978, Sakurai 1985):
440: \begin{equation}
441: \Del \cdot (\A \Del \psi) + \oneover{\A}\prn{\frac{J}{\varpi^2}
442: -1}\frac{J'}{\varpi^2}
443: + \frac{{\beta^2}'\brc{\Veff + \epsilon^2 \ln \brk{\epsilon^2
444: h/(\oneoverrho)}}}{(\oneoverrho)^2}
445: - \frac{\epsilon^2 h'/h}{\oneoverrho} =0,
446: \label{GSE-general}
447: \end{equation}
448: where we have rescaled Bernoulli's function as
449: \begin{equation}
450: H \equiv -\epsilon^2\ln (\epsilon^2h),
451: \end{equation}
452: so that $h$ remains an order unity quantity in our calculation. Here
453: $\A$ is the Alfv\'en discriminant defined by
454: \begin{equation}
455: \A \equiv \frac{M_A^{-2} -1}{\varpi^2 \rho},
456: \label{calAdef}
457: \end{equation}
458: where
459: \begin{equation}
460: M_A^2 \equiv \frac{\rho u^2}{B^2} = \oneover{\beta^2 \rho}
461: \end{equation}
462: is the Alfv\'en Mach number. Hence $\A$ is positive when the total
463: velocity is less than the Alfv\'en speed, and negative when the
464: total velocity is larger than the Alfv\'en speed. From the form of
465: the GSE \eqref{GSE-general}, we see that the conserved angular
466: momentum flux $J$ is not freely specifiable. It is determined by the
467: condition of smooth Alfv\'en transition. In order for the solution
468: to remain continuous and differentiable, one must impose
469: \begin{equation}
470: J = \varpi^2 \; \text{ whenever } \; \A = 0.\label{J-condition}
471: \end{equation}
472: The elimination of $\rho$ in the equations in favor of $\A$ is based
473: on numerical considerations, since $\rho$ will in general vary by
474: many orders of magnitude, while $\A$ only varies moderately. As we
475: argued in the previous section, the sound speed $\epsilon$ is likely
476: to be small. In terms of these variables, the BE takes the
477: form
478: \begin{equation}
479: \abs{\Del \psi}^2 + \oneover{\A^2}\prn{\frac{J}{\varpi^2} -1 }^2 +
480: \frac{2\varpi^2\brc{\Veff +
481: \epsilon^2 \ln \brk{\epsilon^2 h/(\oneoverrho)}}}{(\oneoverrho)^2} =
482: 0.\label{BE-general}
483: \end{equation}
484:
485: \subsection{The Cold Limit}
486: With $\A$ implicitly defined in the BE \eqref{BE-general}, the GSE
487: is a PDE of mixed type, which demands different numerical methods in
488: different regions (see Heinemann \& Olbert 1978 and Appendix
489: \ref{Character}). There are three relevant signal speeds (which we
490: term sonic, slow, and fast in the appendix) involved in an MHD flow
491: (see Jackson 1975 or Shu 1992). The loci where the poloidal fluid
492: speed equals those signal speeds separate the flow into four
493: regions. As the poloidal velocity exceeds the sonic speed, the
494: governing GSE changes from elliptic to hyperbolic. A wise strategy
495: might start with the search of appropriate boundary conditions in
496: the disk where $u_p^2 = 0$, and at the sonic surface (whose location
497: is still undetermined), followed by a standard scheme (e.g.,
498: relaxation) to obtain the interior solution. Beyond the sonic
499: surface, the GSE becomes hyperbolic. The boundary condition on the
500: sonic surface now serves as the initial condition, which we use to
501: integrate forward along characteristics toward the slow surface. We
502: then follow similar procedures to obtain solutions from the slow
503: surface to the fast surface, and beyond.
504:
505: A significant simplification can be achieved when the sound speed is
506: negligible, as in the outer problem of the X-wind (see
507: \S 4 of Shu et al. 1994b). The governing equations are treated as
508: power series expansion in $\epsilon$. The leading term in the GSE
509: \eqref{GSE-general} and the BE \eqref{BE-general} are
510: \begin{mathletters}
511: \begin{eqnarray}
512: &&\Del \cdot (\A \Del \psi) + \oneover{\A}\prn{\frac{J}{\varpi^2}
513: -1}\frac{J'}{\varpi^2} + \frac{{\beta^2}'\Veff}{(\beta^2 - \varpi^2 \A)^2}
514: =0, \label{GSE}\\
515: &&\abs{\Del \psi}^2 + \oneover{\A^2}\prn{\frac{J}{\varpi^2} -1 }^2 +
516: \frac{2\varpi^2 \Veff}{(\beta^2 - \varpi^2 \A)^2} =
517: 0.\label{BE}
518: \end{eqnarray}
519: \end{mathletters}
520: Notice the lowest order term in $H$ vanishes independent of the form
521: of $h$. In this limit, both the sonic speed and the slow speed
522: reduce to zero, and the first elliptic and hyperbolic parts of the
523: flow shrink down to the X-point. We are thus spared the vicissitudes
524: of this portion of the problem. Once the fluid leaves the X-region
525: (with poloidal velocity greater than the slow speed), it proceeds to
526: the fast surface, where the governing equation becomes hyperbolic.
527: %By setting $\epsilon= 0$, one can easily check that the poloidal
528: %component of the fast MHD wave travels at the Alfv\'en velocity, and
529: %the fast mode transition is made when the poloidal component of the
530: %fluid velocity exceeds the Alfv\'en velocity (see Appendix A).
531: %We offer the following interpretation
532: %for this mathematical result. For a general MHD flow, when the sound
533: %speed is taken to be zero, the slow speed vanishes and the
534: %\textit{total} fast speed reduces to the Alfv\'en speed (see,
535: %e.g., Jackson 1975, Shu1992). In other words, there is only one
536: %velocity at which MHD waves can propagate. If there is no symmetry
537: %in the problem, one would indeed expect the fast surface to coincide
538: %with the Alfv\'en surface. However, in an axisymmetric
539: %configuration, waves are forbidden to travel in the toroidal
540: %direction, and the only relevant part of the fluid velocity is its
541: %poloidal component.\footnote{Waves are local disturbances; they are
542: %unaware of the global symmetry. Hence even though the fast wave can
543: %only travel in the poloidal direction, its phase velocity is still
544: %given by the total Alfv\'en velocity.} Therefore the axisymmetry
545: %of the problem breaks the degeneracy of Alfv\'en and fast surface.
546:
547: Najita \& Shu (1994) solved the GSE in the sub-Alfv\'enic region. By
548: introducing a generalized coordinate system, they were able to map
549: the location of the Alfv\'en surface to a known location, and
550: determine the functional form of $\beta(\psi)$ based on the position
551: and shape of the Alfv\'en surface. Their numerical scheme to find
552: $\beta (\psi)$ by iteration encountered a systematic ``drift
553: problem", however, and an artificial ``Alfv\'en seam'' was invented
554: to cope with this difficulty.
555:
556: In a later treatment by Shang (1998), $\beta(\psi)$ was
557: specified in advance, limited in its functional form by
558: considerations of how the gas exits the X-region, an analysis that
559: we repeat in \S 4.2 (see also \S 5). The GSE was not solved as a
560: PDE, but rather as an error estimator in a Weber-Davis type of
561: analysis, where $\psi$ as a trial function of spatial location is
562: obtained by interpolating between the known analytic forms in the
563: X-point neighborhood (see \S 4.1) and at asymptotic infinity (see \S
564: 4.3). The interpolation formula has a number of degrees of freedom,
565: which are adjusted to give ``least error'' in some sense when the
566: trial solution for $\psi$ is substituted back into the GSE. The
567: rest of the problem, including the constraints of the conserved
568: quantities and smooth passage through the Alfv\'en and fast
569: surfaces, are performed exactly. She verified the result derived by
570: Goldreich and Julian (1970) that passage through the Alfv\'en surface
571: is automatic in such a scheme if one has guaranteed it through the
572: fast surface. In fact, \S 5.2 demonstrates the
573: falsity of the frequent claim made otherwise in the literature that
574: $J(\psi)$ is set at the Alfv\'en surface; the claim holds only if one
575: already has a solution such that the wind passes smoothly through
576: the fast surface.
577:
578: %The fact that the discriminant $\Delta$ changes sign through a
579: %vanishing denominator suggests that the hyperbolic part of the flow
580: %possesses its own demon. From any standard textbook on PDE
581: %\citep[e.g., ][]{Garabedian:1986}, $\Delta$ is proportional to the
582: %Jacobian in the transformation from original coordinates to the
583: %characteristics coordinates. A diverging Jacobian implies a
584: %singular coordinate transformation. In our particular example, the
585: %characteristics of the GSE are degenerate on the fast surface, and
586: %can not foliate the entire two dimensional physical space.
587:
588: \section{Variational Principle}
589: Based on the above arguments, the X-wind is a fierce mathematical
590: beast, and a direct numerical attack is unlikely to subdue it fully.
591: To construct a global solution of the X-wind that accelerates
592: elements of plasma from the disk to super-magnetosonic speeds, we
593: must resort to a different approach. Consider the following action
594: written down by inspection:\footnote{In their pioneering development
595: of magnetized stellar winds in a Grad-Shafranov formalism, Heinemann
596: \& Olbert (1978) noted in passing that the resultant equations could
597: be derived from a principle of least action, which differs in
598: detailed form from that used in this paper, but is in the same
599: spirit. However, they, and subsequent workers who have made similar
600: observations, did not exploit the principle to obtain actual wind
601: solutions.}
602: \begin{equation}
603: S = \int \brc{\ahalf \mathcal{A} |\Del \psi|^2
604: -\oneover{2\mathcal{A}} \prn{\frac{J}{\varpi^2} -1}^2 +
605: \frac{V_{\text{eff}} + \epsilon^2 \ln [\epsilon^2 h/(\oneoverrho)]
606: - \epsilon^2}{\beta^2 - \varpi^2 \mathcal{A}}} d^3x,
607: \label{action}
608: \end{equation}
609: It is straight forward to demonstrate that variation against $\psi$
610: yields the GSE \eqref{GSE-general}, while variation against $\A$
611: gives the BE \eqref{BE-general}. The challenges of constructing
612: solutions to a nonlinear PDE of mixed type is now transformed to
613: tuning trial functions of $\psi$ and $\mathcal{A}$ until a
614: local extremum of the action \eqref{action} is reached.
615:
616: To formulate a scheme that is easy to implement numerically, we
617: consider a change of independent variables from the usual
618: cylindrical coordinates
619: \begin{displaymath}
620: (\varpi, z, \varphi) \rightarrow (\varpi, \psi, \varphi).
621: \end{displaymath}
622: For a given value of $\psi$, the functional form of $z(\varpi)$
623: determines the shape of the given streamline, and
624: $\mathcal{A}(\varpi)$ offers information on the velocity distribution
625: along that streamline. Written in these new coordinates, and taking
626: the cold limit as $\epsilon \rightarrow 0$, the action reads
627: \begin{equation}
628: S = 2\pi\iint \brc{\frac{1}{2} \A \prn{\pd{z}{\psi}}^{-1}
629: \brk{1 + \prn{\pd{z}{\varpi}}^2}
630: -\frac{1}{2\A}\pd{z}{\psi} \prn{\frac{J}{\varpi^2} -1}^2 +
631: \pd{z}{\psi}\frac{V_{\text{eff}}}{\beta^2 - \varpi^2 \A}} \varpi
632: d\psi d\varpi \label{new_action}
633: \end{equation}
634: Since $\A$ only enters the action as a constraint rather than a
635: dynamic variable (i.e., its derivative is absent in the action),
636: variation with respect $\A$ yields the BE as before, but now written
637: in a different set of coordinates,
638: \begin{equation}
639: \prn{\pd{z}{\psi}}^{-2} \brk{1 + \prn{\pd{z}{\varpi}}^2}
640: + \oneover{\A^2} \prn{\frac{J}{\varpi^2} -1}^2 +
641: \frac{2\varpi^2\Veff}{(\beta^2 - \varpi^2 \A)^2} = 0
642: \end{equation}
643: Variation with respect to $z$ gives
644: \begin{eqnarray*}
645: &&\delta S_z = 2\pi\iint \Bigg \{-\frac{1}{2} \A \prn{\pd{z}
646: {\psi}}^{-2}
647: \pd{\delta z}{\psi}
648: \brk{1 + \prn{\pd{z}{\varpi}}^2}+\A \prn{\pd{z}{\psi}}^{-1}
649: \prn{\pd{z}{\varpi}} \pd{\delta z}{\varpi}\\
650: &&\qquad -\frac{1}{2\A}\pd{\delta z}{\psi} \prn{\frac{J}{\varpi^2}
651: -1}^2 +
652: \pd{\delta z}{\psi}\frac{V_{\text{eff}}}{\beta^2 - \varpi^2 \A}
653: +\pd{z}{\psi}\frac{V_{\text{eff},z}}{\beta^2 - \varpi^2 \A}
654: \delta z \Bigg \} \varpi
655: d\psi d\varpi.
656: \end{eqnarray*}
657: Integrating by parts, we have
658: \begin{eqnarray*}
659: &&\delta S_z = 2\pi\int \brc{-\frac{1}{2} \A \prn{\pd{z}{\psi}}^{-2}
660: \brk{1 + \prn{\pd{z}{\varpi}}^2}
661: -\frac{1}{2\A}\prn{\frac{J}{\varpi^2}
662: -1}^2 + \frac{V_{\text{eff}}}{\beta^2 - \varpi^2 \A}}
663: \varpi \delta z \Bigg \vert_{\psi = 0}^{\psi =1}d\varpi\\
664: &&\qquad +2\pi \int\A \prn{\pd{z}{\psi}}^{-1}
665: \prn{\pd{z}{\varpi}} \delta z \varpi \Bigg \vert_{\varpi =
666: 1}^{\varpi = \infty}
667: d\psi\\
668: &&\qquad +2\pi\iint \pd{}{\psi}\Bigg \{\frac{1}{2} \A
669: \prn{\pd{z}{\psi}}^{-2}
670: \brk{1 + \prn{\pd{z}{\varpi}}^2}+\frac{1}{2\A}\prn{\frac{J}
671: {\varpi^2}
672: -1}^2 - \frac{V_{\text{eff}}}{\beta^2 - \varpi^2 \A}
673: \Bigg \} \varpi \delta z d\psi d \varpi\\
674: &&\qquad +2\pi \iint \Bigg\{
675: \pd{z}{\psi}\frac{\varpi V_{\text{eff},z}}{\beta^2 - \varpi^2 \A}
676: - \pd{}{\varpi} \brk{\A \prn{\pd{z}{\psi}}^{-1}
677: \prn{\pd{z}{\varpi}}\varpi }\Bigg \} \delta z
678: d\psi d\varpi.
679: \end{eqnarray*}
680: Since we specify the boundary condition $z = 0$ on $\psi = 0$, and
681: $z = Z(\varpi)$ on $\psi = 1$, where $Z(\varpi)$ is a known
682: function, we see that $\delta z$ vanishes on these two boundaries.
683: As we shall see later \citep[see ][]{paper2, paper5}, the
684: solution near the X-point and asymptotically can be constructed
685: analytically. Thus $\delta z$ also vanishes when $\varpi =1$ and
686: $\varpi \rightarrow \infty$ in our variational scheme, and both
687: surface terms vanish in the above expression. In order for the
688: action to be stationary against any choice of $\delta z$, the
689: solution must satisfy the Euler Lagrange equation,
690: \begin{eqnarray}
691: &&\frac{\delta S}{\delta z} = \pd{}{\psi}\brc{\frac{1}{2} \A
692: \prn{\pd{z}{\psi}}^{-2}
693: \brk{1 + \prn{\pd{z}{\varpi}}^2}+\frac{1}{2\A}\prn{\frac{J}
694: {\varpi^2}
695: -1}^2 - \frac{V_{\text{eff}}}{\beta^2 - \varpi^2 \A}
696: } \varpi \nonumber \\
697: &&\qquad + \pd{z}{\psi}\frac{\varpi}{\beta^2 - \varpi^2 \A}
698: \pd{V_{\text{eff}}}{z}
699: - \pd{}{\varpi} \brk{\A \prn{\pd{z}{\psi}}^{-1}
700: \prn{\pd{z}{\varpi}}\varpi } = 0.
701: \label{dSdz}
702: \end{eqnarray}
703: Dividing both sides by $\varpi$, we can simplify the Euler-Lagrange equation
704: \eqref{dSdz} to obtain
705: \begin{eqnarray}
706: &&-\ahalf \pd{\A}{\psi}\brc{
707: \prn{\pd{z}{\psi}}^{-2}
708: \brk{1 + \prn{\pd{z}{\varpi}}^2}+ \frac{1}{\A^2}
709: \prn{\frac{J}{\varpi^2}
710: -1}^2 + \frac{2\varpi^2 \Veff}{(\beta^2 - \varpi^2 \A)^2}} \nonumber \\
711: &&\qquad + \frac{1}{\A}\prn{\frac{J}{\varpi^2}
712: -1} \frac{J'}{\varpi^2} + \frac{{\beta^2}' \Veff}{(\beta^2
713: - \varpi^2 \A)^2} - \oneover{\varpi}\prn{\pd{z}{\psi}}^{-1}
714: \pd{}{\varpi} \brk{\A
715: \prn{\pd{z}{\varpi}}\varpi } \nonumber\\
716: &&+ \prn{\pd{z}{\psi}}^{-1}\pd{}{\psi}\brc{\A
717: \prn{\pd{z}{\psi}}^{-1}
718: \brk{1 + \prn{\pd{z}{\varpi}}^2}}= 0.\label{GSE_halfdo}
719: \end{eqnarray}
720: We notice that the coefficient of $\partial \A/\partial \psi$ is
721: simply the BE, which vanishes at a local extremum of the action. One
722: may easily check that the other terms yield the conventional GSE,
723: but written in our new coordinates.
724:
725: \section{Boundary Conditions}
726: \subsection{X-point}\label{X-point}
727: With the new coordinates, the computational domain is bounded by
728: $\psi \in [0, 1]$ and $\varpi \in [1, \infty)$. The X-point in
729: these coordinates is a singularity given by $\varpi = 1$ for all
730: values of $\psi$. Fortunately, we have analytic solutions there.
731: From this point onward, we shall work with a scaled Alfv\'en
732: discriminant
733: \begin{equation}
734: \chi \equiv \frac{\A}{\beta^2}.
735: \end{equation}
736: This function has the advantage of remaining finite even when
737: $\beta$ diverges. For a given functional form of $\beta$ (which
738: tells us how matter is loaded onto the field lines), the Alfv\'en
739: discriminant has the series expansion in $\varpi-1$.
740: \begin{equation}
741: \chi_X = 1 + \chi_1(\psi) (\varpi-1) + \chi_2(\psi) (\varpi-1)^2 + ...,
742: \end{equation}
743: Here a subscript $X$ reminds us that this series solution is valid
744: near the X-point. In order to match asymptotically onto the outer
745: limit of the inner problem (Shu et al. 1994b), the density $\rho \equiv
746: \beta^{-2}(1- \varpi^2 \chi)^{-1}$ must diverge as $(\varpi-1)^{-2}$
747: near the X-point. This requirement translates to $\chi_1 = -2$ for
748: all values of $\psi$. Similarly, for the coordinate $z$ (now a
749: dependent variable), we expand it as,
750: \begin{equation}
751: z_X = z_1(\psi) (\varpi-1) + z_2(\psi)(\varpi-1)^2 + ....
752: \end{equation}
753: Notice that the Jacobian near the X-point is
754: \begin{displaymath}
755: \sqrt{g} = z_1'(\varpi-1) \; \text{ as } \; \varpi \rightarrow 1.
756: \end{displaymath}
757: which is expected since the entire line of $\psi \in [0, 1]$ is
758: mapped into a single point, and the Jacobian must vanish in this
759: situation. Substituting the series expansion into the GSE
760: \eqref{GSE_halfdo} in the transformed coordinates, the lowest order
761: is $(\varpi-1)^{-2}$.
762: \begin{displaymath}
763: \pd{}{\psi}\brk{\frac{\beta}{z_1'}
764: \prn{1+z_1^2}} = 0,
765: \end{displaymath}
766: which has the solution
767: \begin{equation}
768: z_1 = \tan \vartheta, \qquad \vartheta =
769: \frac{1}{K} \int_0^\psi \beta d\psi.\label{-2nd_order}
770: \end{equation}
771: If we assume that the upper boundary of the X-wind ($\psi = 1$) near
772: the X-point forms an angle of $\theta_X$ with the $x$ axis, we have
773: \begin{displaymath}
774: \tan \theta_X \approx \frac{z}{\varpi-1} \Big \vert_{\chi
775: \rightarrow 1} = z_1(\psi = 1) = \tan \frac{1}{K} \int_0^1 \beta
776: d\psi.
777: \end{displaymath}
778: Thus, the integration constant is given by
779: \begin{equation}
780: K = \frac{\bar \beta}{\theta_X}, \qquad \bar\beta = \int_0^1
781: \beta d\psi.
782: \end{equation}
783: Substituting back into the Bernoulli's equation allows us to solve
784: for $\chi_2$,
785: \begin{equation}
786: \chi_2 = 3- \frac{\sqrt{4\cos^2 \vartheta -1}}{K \beta \cos^2
787: \vartheta} \label{chi_2}
788: \end{equation}
789: For better numerical accuracy, we carry out the computation to next
790: order. The next term in the series expansion of the GSE is $O[(\varpi-1)^{-1}]$:
791: \begin{displaymath}
792: \pdd{\mathcal{Q}}{\vartheta} + \mathcal{Q} = \sin
793: \vartheta,
794: \end{displaymath}
795: where $\mathcal{Q} = z_2\cos^3\vartheta$. Given the boundary
796: condition $z=0$ at $\psi=0$ for all values of $\chi$, the above
797: equation may be solved to give
798: \begin{equation}
799: z_2 = \ahalf \sec^2\vartheta \prn{q \tan \vartheta - \vartheta
800: }.\label{z_2}
801: \end{equation}
802: The integration constant $q$ can be determined by expanding the
803: upper boundary near the X-point. Substituting into the BE, we can determine the last
804: term without the knowledge of $J$ as
805: \begin{equation}
806: \chi_3 = \frac{\cos^2\vartheta -2 + \tan \vartheta(q \tan
807: \vartheta - \vartheta)}{2K \beta \cos^2 \vartheta \sqrt{4
808: \cos^2 \vartheta -1}} + \frac{\sqrt{4 \cos^2 \vartheta -1}
809: (4-q \sec^2\vartheta)}{2K \beta \cos^2 \vartheta} - 4. \label{varpi_3}
810: \end{equation}
811:
812: \subsection{Specifying Mass Loading and Difficulties with the
813: Boundary Layer} \label{Boundary_Layer} Since the X-wind is driven
814: magnetocentrifugally, one would naively expect that it is bounded
815: away from the polar axis (at least in the immediate vicinity of the
816: X-point) by some curve which intersects the disk. In the outer
817: limit of the inner problem (see eq. 3.10d of Shu et al. 1994b), the gas
818: pressure $p = \epsilon^2 \rho$ takes the form
819: \begin{displaymath}
820: p \rightarrow \brk{\frac{\bar \beta}{\vartheta_x(0) \beta}}
821: \sigma^{-2} (4\cos^2\vartheta -1)^{-1/2}, \; \text{ as } \; \sigma
822: \rightarrow \infty,
823: \end{displaymath}
824: where $\tan\vartheta = z/(\varpi-1)$. As $\vartheta_x \rightarrow
825: \pi/3$ (which is the critical angle for the last matter carrying
826: streamline), the pressure diverges unless
827: \begin{displaymath}
828: \beta \propto (4 \cos^2\vartheta -1)^{-1/2}, \; \text{ as } \; \psi
829: \rightarrow 1.
830: \end{displaymath}
831: Substitute this functional form into the lowest order equation
832: \eqref{-2nd_order}, finite magnetic field and pressure on the last
833: streamline demands
834: \begin{equation}
835: \beta \propto (1-\psi)^{-1/3}, \;\text{ as }\; \psi \rightarrow 1.
836: \end{equation}
837:
838: The divergence of $\beta$ should not come as a surprise. Recall
839: that the last streamline is defined to be the boundary between the
840: X-wind and the dead zone. To ensure analyticity across this
841: boundary, we must have $\rho \rightarrow 0$ as $\psi \rightarrow 1$.
842: Now since neither the magnetic field nor the velocity become
843: singular, we must take $\beta \rightarrow \infty$ on that last
844: streamline, so that the product $\beta^2 \rho^2 = B^2/u^2$ remains
845: finite. With this limit in place, we see that the rescaled Alfv\'en
846: discriminant
847: \begin{displaymath}
848: \chi = \frac{1 - 1/\beta^2 \rho}{\varpi^2} \rightarrow
849: \oneover{\varpi^2}
850: \end{displaymath}
851: remains positive for all points along the last streamline. In other
852: words, the flow on the last streamline is always sub-Alfv\'enic,
853: since the Alfv\'en speed is infinite there. This behavior of the
854: last streamline requires a double limiting procedure if we were to
855: accurately construct the asymptotic solution on that interface. We
856: speculate that this difficulty is an indication that the last
857: streamline needs to be treated as a boundary layer. This
858: speculation is reinforced by the fact that the last X-wind
859: streamline is the outer bounding surface to a sheet of axisymmetric
860: current defined by opened stellar field lines that reverse poloidal
861: directions as the current sheet is crossed and we find ourselves in
862: the dead zone of the overall X-wind/funnel-flow configuration (see
863: Fig. 1). Until we actually construct such a
864: boundary-layer/current-sheet theory, we adopt a simple modification
865: to deal with the problem: we truncate the formal wind solution at
866: some $\psi_1<1$, below which $J$ and $\beta$ remain finite. We then
867: add the part between $\psi_1$ and $1$ to the dead zone fields of the
868: problem, i.e., treat the last few streamlines as opened vacuum
869: fields and impose the pressure balance condition at $\psi_1$.
870:
871: \subsection{Asymptotic Solution}\label{Asymptotic Solution}
872: The asymptotic solution at large distances from the X-point was
873: constructed by Shu et al. (1995). In particular, for a wind
874: reaching more or less constant terminal velocity, its density scales
875: roughly as $\rho \propto \varpi^{-2}$, and the Alfv\'en discriminant
876: $\chi \rightarrow -1/\beta^2 \rho \varpi^2$ is a slowly varying
877: function of $r$. By ignoring all radial derivatives compared to
878: angular derivatives, the GSE and the BE admit solutions of the form
879: \begin{equation}
880: \chi = - 1/\beta C, \qquad \sin \theta = \sech [C^{-1}I(C, \psi)],
881: \qquad I(C, \psi) = \int_0^\psi \frac{\beta
882: d\psi}{\sqrt{2J - 3- 2C \beta}},\label{sol_asymp}
883: \end{equation}
884: where $\theta$ is the usual polar angle in spherical coordinates,
885: and $C$ is a ``constant of integration'' that vary slowly in $r$. In
886: an inertial frame, the wind reaches a terminal velocity given by
887: \begin{equation}
888: v_w = (2J-3-2C \beta)^{1/2}.
889: \end{equation}
890: To determine the constant $C$, we impose pressure balance between
891: the X-wind and the dead zone. Since the dead zone field lines carry
892: no inertia, they do not develop a toroidal component, and the
893: poloidal field satisfies the vacuum equation (Ostriker \& Shu1994).
894: Asymptotically, we do not expect the field lines to pinch toward the
895: rotational axis since the hoop stress is vanishingly small (Shu et
896: al. 1995). For simplicity, we assume the boundary layer deviates
897: only slightly from a cylindrical surface at the asymptotic infinity
898: (an assumption which shall be checked a posteriori for consistency).
899: For any given (large) value of $r$, we can approximate the boundary
900: locally by $\varpi = $ const. Then a particular solution is
901: $\mathbf{B} = B_{\rm hc} \zhat$. For the hollow-cone region to trap
902: the same amount of {\it net} flux as the wind part and have a
903: cross-sectional area of $\pi \varpi_{\rm hc}^2$, we have
904: \begin{displaymath}
905: B_{\rm hc} = \frac{2\phi_{\rm hc} \bar \beta}{\varpi_{\rm hc}^2},
906: \end{displaymath}
907: where $\phi_{\rm hc}$ is a number ranging from $1$ to $3$ depending
908: on the fraction of closed field lines in the dead zone (compare Fig.
909: 1 of this paper with Fig. 1 of Shu et al. 2001). The maximal case
910: $\phi_{\rm hc} = 3$ has three times as many field lines as the
911: minimal case $\phi_{\rm hc} = 1$, but the extra field lines cancel
912: in oppositely directed pairs and contribute no net flux. The
913: overall solution does not depend sensitively on the number
914: $\phi_{\rm hc}$, and we take $\phi_{\rm hc}$ henceforth to be unity
915: as illustrated in Figure 1 of the current paper.
916:
917: In contrast, the wind region is dominated by the toroidal field
918: \begin{displaymath}
919: B_{w, \varphi} = \beta \rho u_\varphi = -\frac{J-\varpi^2}
920: {\varpi^3 \chi \beta} \rightarrow -\frac{C}
921: {\varpi}.
922: \end{displaymath}
923: The poloidal field $B_{w,p} = \beta \rho v_w \propto 1/\varpi^2$ is
924: much weaker in this limit. By equating the magnetic pressures on
925: both sides of the boundary, $B_{\rm hc}^2 = B_{w, \varphi}^2$, we obtain
926: \begin{equation}
927: C = \frac{2\bar \beta}{\varpi_{\rm hc}} = \frac{2\bar \beta}{r}
928: \cosh [C^{-1} I(C, 1)],
929: \end{equation}
930: which implicitly defines $C(r)$. Since $I$ only depends very weakly
931: on $C$, this expression shows that $C \rightarrow 0$ logarithmically
932: as $r \rightarrow \infty$. Notice that this limiting behavior of
933: $C$ ensures that $\varpi_{\text{hc}}$ deviates from a constant only
934: logarithmically slowly, which validates our assumption on the
935: geometry of the boundary layer. Written in our coordinates, the
936: asymptotic geometry of each streamline is given by
937: \begin{equation}
938: z = \varpi \sinh[C^{-1} I(C, \psi)].
939: \end{equation}
940: In other words, each streamline is approximately radial, with a
941: logarithmic collimation toward the axis.
942:
943: With $\rho \rightarrow C/\beta \varpi^2$ and $v_w \rightarrow $
944: constant, the poloidal and toroidal Alfv\'en speeds are given by
945: \begin{equation}
946: v_{A,p}^2 = B_p^2/\rho \rightarrow C \beta v_w^2/\varpi^2, \qquad
947: v_{A,\varphi}^2 = B_{\varphi}^2/\rho \rightarrow C\beta.
948: \end{equation}
949: Thus the Alfv\'en speed is dominated by the toroidal component,
950: which decreases logarithmically. This means that the (poloidal)
951: terminal velocity is super-fast in the asymptotic regime, and the
952: wind has to make a fast mode transition along each streamline. The
953: above analysis simply reiterates the claims made in Appendix
954: \ref{Character} that the asymptotic behavior of the flow is governed
955: by a hyperbolic differential equation.
956:
957: %\begin{displaymath}
958: % \oneover{r^2\sin\theta} \prn{\sin\theta \chi
959: % \beta^2 \psi_{,\theta}}_{,\theta} + \oneover{\chi \beta^2}
960: % \prn{\frac{J}{\varpi^2} -1} \frac{J'}{\varpi^2} +
961: % \frac{2\Veff}{\beta^4(1-\chi \varpi^2)^2}
962: % \beta \beta' = 0.
963: %\end{displaymath}
964: %Upon multiplication by $2 \chi \beta^2 r^2 \sin^2\theta
965: %\psi_{,\theta}$, and take the leading term in $\Veff$, we have
966: %\begin{displaymath}
967: % \pd{}{\theta}\brk{\prn{\sin\theta \chi
968: % \beta^2 \psi_{,\theta}}^2} + 2
969: % \prn{\frac{J}{\varpi^2} -1} \pd{J}{\theta} -
970: % \frac{\chi \varpi^4}{(1-\chi \varpi^2)^2} \pd{\log \beta^2}{\theta}
971: % = 0.
972: %\end{displaymath}
973: %In the same limit, the Bernoulli's equation \eqref{BE} can be
974: %written as
975: %\begin{displaymath}
976: % (\sin \theta \psi_{,\theta})^2 =
977: % \frac{(\varpi^2 - 3)\varpi^4}{\beta^4(1-\chi\varpi^2)^2} -
978: % \frac{\varpi^2}{\chi^2 \beta^4}
979: % \prn{\frac{J}{\varpi^2} -1}^2.
980: %\end{displaymath}
981: %We have to keep two leading terms in $\Veff$ here because for
982: %streamlines with $\beta \sim O(1)$, $\alpha^2 \chi^2 \beta^4
983: %\rightarrow 1$, which causes the $\varpi^2$ term to cancel.
984: %Substitute Bernoulli's equation back into the Grad-Shafranov
985: %equation, we have
986: %\begin{displaymath}
987: % \pd{}{\theta}\brk{\frac{\chi^2\varpi^6}{(1-\chi\varpi^2)^2}} +
988: % \pd{\varpi^2}{\theta}\brk{\prn{\frac{J^2}{\varpi^4} -1}} -
989: % \frac{\chi \varpi^4}{(1-\chi \varpi^2)^2} \pd{\log \beta^2}{\theta}
990: % = 0.
991: %\end{displaymath}
992: %It is advantageous now to choose the asymptotic surface to be
993: %$\varpi = \varpi_\infty =$ const. so that the Grad-Shafranov
994: %equation reads
995: %\begin{equation}
996: % \chi \varpi_\infty^2 \pd{}{\theta}\brk{\log \frac{\chi^2
997: % \varpi_\infty^6}{(1-\chi\varpi_\infty^2)^2}} -
998: % \pd{\log \beta^2}{\theta}
999: % = 0.\label{GSE_asymptotic}
1000: %\end{equation}
1001: %Equation \eqref{GSE_asymptotic} can be integrated to give
1002: %\begin{equation}
1003: % \chi = \frac{1}{\varpi_\infty^2}-\frac{1}{C
1004: % \beta},\label{chi_asymp}
1005: %\end{equation}
1006: %where $C$ is a (slow varying) function of $r$. The other solution
1007: %\begin{displaymath}
1008: % \chi = \frac{C \beta - \varpi_\infty^2}{(C \beta - 2
1009: % \varpi_\infty^2) \varpi_\infty^2}
1010: %\end{displaymath}
1011: %is extraneous because it gives a diverging $\chi$ for $C \beta \sim
1012: %2\varpi_\infty^2$, and it does not have the correct limit when
1013: %$\varpi_\infty \rightarrow \infty$ with finite $\beta$. Adopting the
1014: %finite solution \eqref{chi_asymp}, the Bernoulli's equation becomes
1015: %\begin{equation}
1016: % \sin\theta\pd{\psi}{\theta} =-\frac{C}{\beta}\brk{
1017: % \varpi_\infty^2 - 3
1018: % - \frac{\varpi_\infty^6}{(C \beta - \varpi_\infty^2)^2}
1019: % \prn{\frac{J}{\varpi_\infty^2} -1}^2}^{1/2}.\label{BE_asymptotic}
1020: %\end{equation}
1021: %The sign is chosen so that $\psi$ decreases with increasing
1022: %$\theta$. In the regions where $\beta$ is finite (and hence $J$ is
1023: %finite), the asymptotic solution requires $\varpi_\infty^2 \gg J$.
1024: %Then the asymptotic boundary is located at $\chi = -1/\beta C$, as
1025: %discovered by \citet{paper5}. In the same region, the Bernoulli's
1026: %equation can be approximated as
1027: %\begin{displaymath}
1028: % \sin \theta\pd{\psi}{\theta} =-\frac{C}{\beta}
1029: % \brk{- 3 -
1030: % 2\beta C +2J}^{1/2},
1031: %\end{displaymath}
1032: %which has solution
1033:
1034: \section{Global Solutions}
1035: As a particular example, let us suppose that diffusive mass loading
1036: onto field lines in the X-region produces a $\beta$ function which
1037: has the form
1038: \begin{equation}
1039: \beta = \frac{2}{3} \bar \beta (1-\psi)^{-1/3}.\label{beta}
1040: \end{equation}
1041: It is easy to verify that $\int_0^1 \beta d\psi = \bar \beta$. To be
1042: definite, let us also assume that the upper boundary of the
1043: X-wind near the X-point forms the maximum angle $\vartheta_X =
1044: \pi/3$ with the equatorial plane in order for magnetocentrifugal
1045: acceleration to operate. The $O[(\varpi-1)^{-2}]$ solution
1046: \eqref{-2nd_order} takes the form
1047: \begin{equation}
1048: z_1 = \tan \vartheta = \tan \frac{\pi}{3} \brk{1-(1-\psi)^{2/3}}.
1049: \label{-2nd_order_solution}
1050: \end{equation}
1051: In fact we have chosen a very special value for the opening angle.
1052: Recall that the formal boundary between the X-wind and the dead zone
1053: is characterized by vanishing $\rho$ with finite magnetic field.
1054: That results in $\beta \rightarrow \infty$ and $\chi = \varpi^{-2}$.
1055: If $\vartheta_X$ were smaller than $\pi/3$, one may check that the
1056: boundary condition $\chi = \varpi^{-2}$ agrees with the series
1057: solution of \S \ref{X-point} to the second order for all values of
1058: $q$. This integration constant is computed by expanding the shape of
1059: the last streamline near the X-point. However, when $\vartheta_X =
1060: \pi/3$, the series solution agrees with the boundary condition only
1061: if the quantity $q$ in equation (\ref{z_2}) satisfies
1062: \begin{displaymath}
1063: q = \oneover{3} \prn{\frac{7}{4} + \frac{\pi}{\sqrt{3}}}.
1064: \end{displaymath}
1065: If $\vartheta_X > \pi/3$, the solution becomes discontinuous. This
1066: behavior is consistent with our physical intuition. When the flow
1067: is cold, the upper boundary of the X-wind is imposed by pressure
1068: balance. As one eases up the external pressure, $\vartheta_X$
1069: increases. However, even when the external pressure drops to zero,
1070: the matter carrying streamlines are confined to $\vartheta \le
1071: \pi/3$, since it is the boundary where centrifugal effects can
1072: overcome gravity. At least near the X-point, there is no freedom to
1073: choose the shape of the last streamline. Any excursion across this
1074: boundary requires additional pressure support from the X-wind, which
1075: calls for a warm rather than cold outflow.
1076:
1077: In the particular example we are studying here, the second order
1078: coefficient for $z$ becomes
1079: \begin{equation}
1080: z_2 = \ahalf \sec^2 \vartheta \brk{\prn{\frac{7}{12} +
1081: \frac{\pi}{3\sqrt{3}}}
1082: \tan \vartheta - \vartheta}.
1083: \end{equation}
1084: For numerical tractability, we place the boundary layer at $\psi =
1085: 0.99$. Given the choice of mass loading in equation \eqref{beta},
1086: the $\beta$ function is not much larger than unity there.
1087:
1088: \subsection{Fixing the Free Function $J(\psi)$}
1089: \label{find_J}
1090: The BE \eqref{BE} is actually
1091: a quartic algebraic equation for the Alfv\'en discriminant once the
1092: shape of the streamlines are known. The Alfv\'en surface here is not
1093: a real singularity of the equation; it simply ensures that $\chi =
1094: 0$ is a solution when $J = \varpi^2$. In other words smooth crossing
1095: of the Alfv\'en surface does not uniquely determine the value of
1096: $J$. To see this, let us define
1097: \begin{displaymath}
1098: \mathcal{L} = \varpi \rho u_\varphi = - \oneover{\beta^2 \chi}
1099: \prn{\frac{J}{\varpi^2} -1}.
1100: \end{displaymath}
1101: It is always negative and asymptotes to zero for the wind since the
1102: magnetic field lines form a trailing spiral. The BE can be written
1103: as
1104: \begin{equation}
1105: (\abs{\Del \psi}^2 + \mathcal{L}^2)(\beta^2 \mathcal{L}
1106: + J - \varpi^2)^2 +
1107: 2\varpi^2 \Veff \mathcal{L}^2 = 0.\label{mod_BE}
1108: \end{equation}
1109: As long as $J$ is larger than some critical value, there are always
1110: real and finite solutions to this equation, which means the Alfv\'en
1111: surface is automatically crossed. On the other hand, the fast point
1112: is a real critical point for the BE. A smooth fast mode transition
1113: demands the BE to have a double root at the critical point (see Fig
1114: \ref{fast_crossing}). That means not only does the left hand side of
1115: \eqref{mod_BE} need to vanish, its derivative with respect to
1116: $\mathcal{L}$ must vanish as well.
1117:
1118: After some algebra, these
1119: requirements can be written as
1120: \begin{equation}
1121: \mathcal{L} = \abs{\Del \psi}^{2/3}
1122: (J - \varpi^2)^{1/3} \beta^{-2/3}.\label{fast}
1123: \end{equation}
1124: This expression is simply a statement that at the fast point, the
1125: poloidal fluid velocity is equal to the magnetosonic speed, which
1126: for $\epsilon = 0$ is equal to the total Alfv\'en speed. Notice that
1127: both equations \eqref{mod_BE} and \eqref{fast} are automatically
1128: satisfied by $\mathcal{L} = (J-\varpi^2) = 0$. This solution,
1129: however, is unphysical since it has a discontinuity on the Alfv\'en
1130: surface when $\chi = 0$. Substituting equation \eqref{fast} back in
1131: to the BE \eqref{mod_BE}, we have
1132: \begin{equation}
1133: \brk{\abs{\Del \psi}^{2/3}
1134: \beta^{4/3} + (J - \varpi^2)^{2/3}}^3 +
1135: 2\varpi^2 \Veff= 0.\label{degen_root}
1136: \end{equation}
1137: For a given value of $J$, the solutions to this equation give the
1138: locations where the BE has degenerate roots. If $J<J_c$, equation
1139: \eqref{degen_root} has no roots in the super-Alfv\'en region
1140: ($\varpi^2 > J$). If $J>J_c$, then equation \eqref{degen_root} has
1141: two roots in the super-Alfv\'en part of the flow. The desired
1142: solution is obtained when $J=J_c$, and there is only one double root
1143: occurring at the fast mode transition point (see Fig. 2).
1144: \begin{figure}[ht]
1145: \begin{center}
1146: \includegraphics[width = 4.5in, angle = 0]{f2.ps}
1147: \end{center}
1148: \caption{Determination of the critical value of $J$ that allows a
1149: fast mode transition}
1150: \label{fast_crossing}
1151: \end{figure}
1152:
1153: \subsection{Interpolation Schemes and Numerical Strategy}
1154: Our strategy is then to find interpolations between the X-point
1155: solution in \S \ref{X-point} and the asymptotic solution of \S
1156: \ref{Asymptotic Solution} so that the action \eqref{new_action} is
1157: extremized. Since the action involves an integral extending to $r
1158: \rightarrow \infty$, and the streamlines are approximately radial,
1159: in general the action integral is infinite. In the X-wind problem,
1160: however, the assumption of stationarity is an approximation that
1161: must fail physically at very large distances from the X-point. If
1162: the flow extends all the way to spatial infinity, then steady state
1163: cannot be established in finite time. To make the practical aspect
1164: of this problem manageable, we opt to truncate the action integral
1165: at some finite spatial surface, and assume that the solution is
1166: identical to the asymptotic solution beyond that point. Then the
1167: interpolation requires the intermediate solution to join smoothly on
1168: to the asymptotic solution at the boundary. Since the parameter $C$
1169: that appears in the asymptotic solution is purely a function of $r$,
1170: it is natural to choose the boundary surface at $r = r_0 <\infty$.
1171: Thus, along a given streamline labeled by $\psi$, the action
1172: involves an integral over the range $\varpi \in [1, \varpi_\infty]$,
1173: where
1174: \begin{equation}
1175: \varpi_\infty = \frac{2 \bar \beta}{C} \frac{\cosh[C^{-1}I(C,
1176: 1)]}{\cosh[C^{-1} I(C, \psi)]} \label{varpi_infty}
1177: \end{equation}
1178: Here $I$ is the integral defined in equation \eqref{sol_asymp}, and
1179: the asymptotic value of $z$ is given by
1180: \begin{equation}
1181: z_\infty(\varpi, \psi) = \varpi_\infty \sinh \brk{C^{-1} I(C, \psi)},
1182: \end{equation}
1183:
1184: Since the asymptotic behavior of the streamlines are predominantly
1185: radial with a logarithmic collimation toward the pole, we may
1186: approximate them by linear functions. There is a large class of
1187: basis functions in which $z(\psi, \varpi)$ can be expanded. To avoid
1188: unphysical oscillations introduced by higher order polynomial
1189: interpolations, we approximate $z$ by a cubic spline such that the
1190: second derivative $z_{\varpi\varpi}$ is a continuous piecewise
1191: linear function.
1192: \begin{equation}
1193: z_{\varpi \varpi} = f_i + (\varpi-\varpi_i)\frac{f_{i+1} - f_i}
1194: {\varpi_{i+1} - \varpi_i} , \; \text{ for } \;
1195: \varpi_i \le \varpi < \varpi_{i+1},
1196: \end{equation}
1197: where $i = 0...N-1$, with $\varpi_0 = 1$ and $\varpi_N =
1198: \varpi_\infty$. The boundary conditions on $z_{\varpi\varpi}$ read
1199: \begin{equation}
1200: f_0 = 2z_2, \qquad f_N = 0.
1201: \end{equation}
1202: Direct integration yields \citep{Press:1992}
1203: \begin{eqnarray}
1204: &&z = a y_i + b y_{i+1} + c f_i + d f_{i+1},\label{y-spline}\\
1205: &&z_\varpi = \frac{y_{i+1}-y_i}{\varpi_{i+1}-\varpi_i} -
1206: \frac{3a^2-1}{6}(\varpi_{i+1} -
1207: \varpi_i)f_i + \frac{3b^2 -1}{6} (\varpi_{i+1} - \varpi_i)
1208: f_{i+1},\label{yp-spline}
1209: \end{eqnarray}
1210: where
1211: \begin{eqnarray*}
1212: &&a \equiv \frac{\varpi_{i+1} -\varpi}{\varpi_{i+1} - \varpi_i},
1213: \qquad b \equiv 1-a = \frac{\varpi-\varpi_i}{\varpi_{i+1}-\varpi_i},\\
1214: &&c \equiv \frac{1}{6}(a^3-a)(\varpi_{i+1}-\varpi_i)^2, \qquad d \equiv
1215: \frac{1}{6} (b^3-b)(\varpi_{i+1}-\varpi_i)^2,
1216: \end{eqnarray*}
1217: and $y_i \equiv z(\varpi_i)$ for that particular streamline. The
1218: $y_i$ are determined by demanding $z_\varpi$ is continuous
1219: throughout the domain. Explicitly,
1220: \begin{equation}
1221: \frac{\varpi_i-\varpi_{i-1}}{6} f_{i-1} + \frac{\varpi_{i+1} - \varpi_{i-1}}{3} f_i +
1222: \frac{\varpi_{i+1} - \varpi_i}{6} f_{i+1} = \frac{y_{i+1} - y_i}{\varpi_{i+1} -
1223: \varpi_i} - \frac{y_i - y_{i-1}}{\varpi_i -
1224: \varpi_{i-1}},\label{smooth_condition}
1225: \end{equation}
1226: which is a set of $N-2$ linear equations for the $N$ $y_i$. The
1227: boundary conditions $y_0 = 0$ and $y_N = z_\infty$ close the
1228: equations, and allows unique determination of $y_i$ once $f_i$ are
1229: given. Since we have information on the slope of the solution on
1230: both boundaries, they impose two further constraints
1231: \begin{eqnarray}
1232: &&z_1= \frac{y_{1}-y_0}{\varpi_{1}-\varpi_0} - \frac{1}{3}(\varpi_{1} -
1233: \varpi_0)f_0 - \frac{1}{6} (\varpi_{1} - \varpi_0)
1234: f_{1},\label{ypXconstraint}\\
1235: &&z_{\infty,\varpi}= \frac{y_{N}-y_{N-1}}{\varpi_{N}-\varpi_{N-1}}
1236: + \frac{1}{6}(\varpi_{N} -
1237: \varpi_{N-1})f_{N-1} + \frac{1}{3} (\varpi_{N} - \varpi_{N-1})
1238: f_{N}.\label{ypinftyconstraint}
1239: \end{eqnarray}
1240: To demonstrate the principles, we choose $N=3$, so that all the
1241: $f_i$ are constrained. For a given set of $\varpi_i$, the equations
1242: \eqref{smooth_condition}, \eqref{ypXconstraint}, and
1243: \eqref{ypinftyconstraint} form a set of four linear equations, which
1244: can be solved by standard means. We also define $\varpi_2$ by
1245: \begin{equation}
1246: \frac{\varpi_2-\varpi_1}{\varpi_1 - 1} = \frac{\varpi_\infty -
1247: \varpi_2}{\varpi_2 - \varpi_1},\label{varpi_2}
1248: \end{equation}
1249: i.e., we demand that the interval between interpolation points to
1250: increase exponentially. Thus the shape of each streamline is
1251: parameterized by a single variable, $\varpi_1$.
1252: %\begin{align*}
1253: % &f_0 \frac{\varpi_1-\varpi_{0}}{6} - 2\prn{\varpi_{2} - \varpi_{0}}\brk{
1254: % \frac{y_0}{(\varpi_{1}-\varpi_0)^2} +\frac{f_0}{3} +\frac{z_1}{\varpi_1
1255: % -\varpi_0}} \\
1256: % &\qquad -
1257: % \prn{\varpi_{2} - \varpi_1} \brk{\frac{y_{3}}{(\varpi_{3}-\varpi_{2})^2}
1258: % +\frac{f_{3}}{3} - \frac{z_{\infty,\varpi}}{\varpi_{3} -
1259: % \varpi_{2}}} - \frac{y_{0}}{\varpi_1 -
1260: % \varpi_{0}}\\
1261: % &\qquad = - y_1 \brk{\frac{1}{\varpi_{2} -
1262: % \varpi_1} + \frac{1}{\varpi_1 - \varpi_{0}} + \frac{2(\varpi_2
1263: % -\varpi_0)}{(\varpi_1 - \varpi_0)^2}} + y_2\brk{\frac{1}{\varpi_{2} -
1264: % \varpi_1} -\frac{\varpi_2-\varpi_1}{(\varpi_3 - \varpi_2)^2}} \\
1265: % &-\prn{\varpi_2-\varpi_{1}}\brk{\frac{y_0}{(\varpi_{1}-\varpi_0)^2}
1266: % + \frac{f_0}{3} + \frac{z_1}{\varpi_1
1267: % -\varpi_0}} -\frac{y_{3}}{\varpi_{3} -
1268: % \varpi_2} \\
1269: % &\qquad -2 \prn{\varpi_{3} - \varpi_{1}}\brk{\frac{y_{3}}{(\varpi_{3}-\varpi_{2})^2}
1270: % +\frac{1}{3}f_{3} - \frac{z_{\infty,\varpi}}{\varpi_{3} -
1271: % \varpi_{2}}} +
1272: % \frac{\varpi_{3} - \varpi_2}{6} f_{3} \\
1273: % &\qquad = y_1 \brk{\frac{1}{\varpi_2- \varpi_1} - \frac{\varpi_2 - \varpi_1}
1274: % {(\varpi_1-\varpi_0)^2}} - y_2\brk{\frac{1}{\varpi_{3} -
1275: % \varpi_2} + \frac{1}{\varpi_2 -
1276: % \varpi_{1}} + \frac{2(\varpi_3-\varpi_1)}{(\varpi_3 - \varpi_2)^2}}\\
1277: % &f_{1} = \frac{6y_{1}}{(\varpi_{1}-\varpi_0)^2} -6\frac{y_0}
1278: % {(\varpi_{1}-\varpi_0)^2}- 2f_0 -6\frac{z_1}{\varpi_1
1279: % -\varpi_0},\\
1280: % &f_{2}= \frac{6y_{2}}{(\varpi_{3}-\varpi_{2})^2} -6\frac{y_{3}}{(\varpi_{3}-\varpi_{2})^2}
1281: % -2f_{3} + 6\frac{z_{\infty,\varpi}}{\varpi_{3} -
1282: % \varpi_{2}}.
1283: %\end{align*}
1284:
1285: %\begin{equation}
1286: %\begin{split}
1287: % &\chi = \chi_\infty
1288: % + \oneover{2\bar \beta \beta} \prn{1-\frac{\varpi-1}{\varpi_0-1}}^2
1289: % P_\chi, \\
1290: % &\chi_\infty = \oneover{\varpi^2}
1291: % - \frac{\varpi}{2\bar \beta \beta}, \qquad P_\chi =
1292: % \frac{\sum_{i=0}^{M} a_i(\psi)
1293: % (\varpi-1)^i}{1+\sum_{i=0}^N b_i(\psi) (\varpi-1)^i}.
1294: %\end{split}
1295: %\end{equation}
1296:
1297: The action integral and the asymptotic solution can be treated as
1298: solutions to a set of simultaneous ``ordinary'' differential
1299: equations
1300: \begin{equation}
1301: \od{S}{\psi} = \int_1^{\varpi_\infty(\psi)} L \varpi
1302: d\varpi,\qquad
1303: \od{I}{\psi} = \frac{\beta}{\sqrt{2J(\psi) - 3 - 2 C
1304: \beta(\psi)}},\label{num_int}
1305: \end{equation}
1306: subject to the boundary conditions
1307: \begin{displaymath}
1308: S(0) = 0, \qquad I(0) = 0.
1309: \end{displaymath}
1310: Here $L$ represents the Lagrangian appearing in the action
1311: \eqref{new_action}. For each value of $\psi$, to compute the right
1312: hand side of equation \eqref{num_int}, one needs the values of
1313: $\varpi_1(\psi)$, $I(\psi)$ and $I(\psi_1)$, where $\psi_1 = 0.99$
1314: is the label of the boundary layer discussed in \S
1315: \ref{Boundary_Layer}. Ideally, one would like to specify the shape
1316: of the last streamline by fixing the values of $\varpi_1(\psi_1)$
1317: and $I(\psi_1)$ as boundary conditions, and vary the function
1318: $\varpi_1(\psi)$ in a constrained manner to achieve a local extremum
1319: of the action. In practice, we find it more convenient to implement
1320: a scheme where only $\varpi_1(\psi_1)$ is given, and $I(\psi_1)$ is
1321: determined as an eigenvalue. This approach allows more freedom in
1322: the parameter search for the desired $\varpi_1(\psi)$. With each
1323: streamline fully parameterized, one can proceed to determine the
1324: necessary value of $J(\psi)$ that allows a smooth fast mode
1325: transition according to the procedure outlined in \S \ref{find_J}.
1326:
1327: Once $J(\psi)$ and $I(\psi)$ are both known, we can easily solve the
1328: BE \eqref{BE} as an algebraic equation along each streamline for
1329: $\mathcal{L}$ using standard techniques such as Laguerre's method
1330: (see Press et al. 1992). In particular, note that we do {\it not}
1331: actually use the extremal property of the action principle with
1332: respect to $\cal A$ to attack the Bernoulli equation, but effect
1333: direct solutions of it instead. Increased numerical accuracy
1334: constitutes only one reason for a mixed procedure, where we do find
1335: the extremal action through variations of $\psi$, or equivalently,
1336: through variations of $z(\varpi, \psi)$, as a substitute for solving
1337: the Grad-Shafranov equation. There is a yet more practical reason.
1338: It turns out the the correct solution sits on a saddle, where the
1339: extremal action is minimized by variations of $\psi$ but maximized
1340: by variations of $\cal A$. This combination makes a numerical
1341: search for the extremal action extremely difficult to execute in
1342: practice, perhaps even impossible, if the search is carried out in
1343: the double-function space of allowable $\psi$ and $\cal A$.
1344:
1345: One further obstacle to
1346: overcome is that the action integral \eqref{new_action} is
1347: logarithmically divergent at the X-point. Recall that the Jacobian
1348: of the coordinate transformation vanishes at the X-point since it
1349: maps the entire axis of $\varpi = 1$ onto a single point. A series
1350: expansion of the Lagrangian using the series solution of section
1351: \S\ref{X-point} shows that it diverges as
1352: \begin{equation}
1353: L_* = \frac{K\beta}{2(\varpi-1)}.
1354: \end{equation}
1355: Fortunately, this term does not enter into the variation scheme, and
1356: we may safely remove it as a counter term from the Lagrangian, as is
1357: the standard practice in quantum-field theory.
1358:
1359: Finally, the function $\varpi_1(\psi)$ is modeled by a Hyman
1360: filtered spline \citep{Hyman:1983} interpolating over evenly spaced
1361: control points $\psi_i \in [0, \psi_1]$. The values of $\varpi_1$
1362: at these control points, $\varpi_1(\psi_i)$, are the parameters we
1363: can adjust in our variation scheme. We restrict the parameter space
1364: to that satisfies the condition that the streamlines do not cross
1365: and that each streamline is monotonic. We then adopt a genetic
1366: algorithm to search for a set of $\varpi_1(\psi_i)$ that gives a
1367: local extremum of the action \eqref{new_action}.
1368:
1369: \section{Numerical Results}
1370: We compute the streamlines for three cases of average mass loading
1371: corresponding to $\bar \beta = 1, 2, 3$. In each case, we place the
1372: outer boundary of the computational domain at a constant radius so
1373: that it intersects the last streamline at $\varpi_\infty(\psi_1) =
1374: 20$ (which yields $C = 0.1 \bar \beta$). After a multidimensional
1375: search, we locate the desired set of control points that extremize
1376: the action. They are tabulated in Table \ref{control_pts}, and the
1377: function $\varpi_1(\psi)$ is interpolated between these points as
1378: described in the previous section.
1379: \begin{table}[ht]
1380: \begin{center}
1381: \begin{tabular}{|c|c|c|c|c|c|c|}
1382: \hline $\bar \beta$ & $\varpi_1(0.0)$ & $\varpi_1(0.2)$
1383: &$\varpi_1(0.4)$
1384: &$\varpi_1(0.6)$ &$\varpi_1(0.8)$ &$\varpi_1(0.99)$\\
1385: \hline
1386: 1.0 &29.687 & 29.817 & 23.281 & 18.809 & 8.310 &6.000 \\
1387: 2.0 &28.281 & 29.165 & 24.644 & 17.774 & 11.150 &6.000\\
1388: 3.0 &28.384 & 28.985 & 23.990 & 19.965 & 10.139 &6.000\\
1389: \hline
1390: \end{tabular}
1391: \caption{Values of control points $\varpi_1(\psi_i)$ that yield a
1392: local extremum of the action. The last value $\varpi_1(0.99)$ is
1393: fixed as a boundary condition.} \label{control_pts}
1394: \end{center}
1395: \end{table}
1396:
1397: For each converged solution, we can numerically integrate the
1398: asymptotic equation to evaluate $I(\psi)$. For practical purposes,
1399: we present here an interpolation formula that is a seventh degree
1400: polynomial in $\beta^{-1}$, and the coefficients are tabulated in Table
1401: \ref{I_inter}. Once $I(\psi)$ is known, one may determine the outer
1402: boundary of the computational domain in accordance with the
1403: asymptotic condition \eqref{sol_asymp}. With the combination of
1404: $\varpi_1(\psi)$ and $I(\psi)$, we are able to reconstruct the
1405: streamlines with the spline interpolation scheme, and they are
1406: depicted in Figure \ref{crit}. The location of the Alfv\'en surface
1407: determines the value of $J$ as a function of $\psi$, which
1408: ultimately allows us to compute the angular momentum being
1409: transported as well as the terminal velocity along each streamline.
1410: For convenience, we also present an interpolation formula for $J(\psi)$ as a
1411: polynomial in $\beta$, with the coefficients tabulated in Table
1412: \ref{J_inter}.
1413: \begin{table}[ht]
1414: \begin{center}
1415: \begin{tabular}{|c|c|c|c|c|c|c|c|c|}
1416: \hline
1417: $\bar\beta$&$I_0$&$I_1$&$I_2$&$I_3$&$I_4$&$I_5$&$I_6$&$I_7$\\
1418: \hline 1&0.732&-0.446&0.886&-0.511&-1.781&2.648&-1.397&0.263\\
1419: 2&0.842&0.413&-2.504&2.699&-11.768&24.849&-22.870&7.602\\
1420: 3&1.164&-2.915&30.674&-194.996&585.524&-998.495&$930.236$&$-347.764$ \\
1421: \hline
1422: \end{tabular}
1423: \caption{Interpolation formula for $I_{\text{int}}(\psi) =
1424: \sum_{i=0}^7 I_i \beta^{-i}$. The interpolated function agrees with
1425: the numerical values to within $0.5\%$.} \label{I_inter}
1426: \end{center}
1427: \end{table}
1428:
1429: \begin{figure}[ht]
1430: \begin{center}
1431: \includegraphics[width = 2.9in, angle = 90]{f3a.ps}
1432: \includegraphics[width = 2.9in, angle = 90]{f3b.ps}
1433: \includegraphics[width = 2.9in, angle = 90]{f3c.ps}
1434: \includegraphics[width = 2.9in, angle = 90]{f3d.ps}
1435: \includegraphics[width = 2.9in, angle = 90]{f3e.ps}
1436: \includegraphics[width = 2.9in, angle = 90]{f3f.ps}
1437: \end{center}
1438: \caption{Solutions for $\bar \beta = 1, 2, 3$. The dotted curves
1439: represent the streamlines labeled by constant $\psi$, and the solid curves
1440: are isodensity contours separated by logarithmic intervals. The dashed
1441: curves are the location of the Alfv\'en surface in each case, and the
1442: dashed-dotted curves mark the fast surface where the GSE becomes
1443: hyperbolic.}
1444: \label{crit}
1445: \end{figure}
1446:
1447: \begin{table}[ht]
1448: \begin{center}
1449: \begin{tabular}{|c|c|c|c|c|c|c|c|c||c|}
1450: \hline
1451: $\bar\beta$&$J_0$&$J_1$&$J_2$&$J_3$&$J_4$&$J_5$&$J_6$&$J_7$&$\bar J$\\
1452: \hline
1453: 1&-2.791&17.214&-13.640&-10.294 & 22.483 &-13.656 &3.628 &-0.362 & 2.638\\
1454: 2&-14.944&45.162&-43.065&21.660&-6.265&1.057&-0.0969&0.00373 & 4.356\\
1455: 3&-20.285&40.676&-25.088&8.006&-1.440&0.148&-0.00825&0.000191& 6.202\\
1456: \hline
1457: \end{tabular}
1458: \end{center}\caption{Interpolation formula for $J_{\text{int}}(\psi) =
1459: \sum_{i=0}^7 J_i \beta^i$. The last column gives the value $\bar J$
1460: of $J(\psi)$ averaged over $\psi$ from 0 to 0.99. The interpolated
1461: function agrees with the numerical values to within $1\%$. }
1462: \label{J_inter}
1463: \end{table}
1464:
1465: The solid lines in Figure 3 show the logarithmically spaced contours
1466: of constant density. It is evident that even though the dotted
1467: streamlines become asymptotically radial and only collimate
1468: logarithmically slowly, the density becomes cylindrically stratified
1469: very quickly, giving the X-wind the illusion of a jet-like
1470: appearance (Shang et al. 1998, 2002).
1471:
1472: Detailed comparisons of the results obtained here with those given
1473: by Shang (1998) show some differences, but the main impression is
1474: how remarkably well the solutions obtained by the two very different
1475: methods for the same mass-to-flux loading $\beta (\psi)$ agree with
1476: one another. Shang (1998) had a similar experience in comparing her
1477: approximate, but analytic, solutions for the sub-Alfv\'enic region
1478: to the exact, but numerical, solutions obtained by Najita \& Shu
1479: (1994).
1480:
1481: We attribute the fortunate circumstance to the following causes. If
1482: one is given somehow the geometric shape of the streamlines (or,
1483: equivalently, the field lines in the meridional plane), then the
1484: Weber-Davis procedure used by Shang, which includes an exact
1485: solution of Bernoulli's equation, would give an exact solution of
1486: the two-dimensional flow problem, provided one takes care to cross
1487: each of the critical points properly. In realistic circumstances,
1488: the geometric shape of streamlines in the meridional plane is not
1489: given a priori, but is to be found from the Grad-Shafranov equation
1490: (or, equivalently, from minimizing the action by variations of the
1491: stream function $\psi$). However, if one has analytic solutions to
1492: the Grad-Shafranov equation (from the work of Shu et al. 1994b and
1493: 1995) near and far from the X-point, then there are only so many
1494: ways that one can adjust the function $z(\varpi, \psi)$ for values
1495: of $\psi$ from 0 to 1 and of $\varpi$ close to 1 (or dimensionally,
1496: $R_X$) to $\varpi \gg 1$ (or $R_X$) that will connect the shape of
1497: the streamlines near the X-point (a fan) smoothly to those
1498: appropriate at asymptotic infinity (radial outflow). The procedures
1499: used by Shang (1998) and those used here to make such adjustments
1500: differ, but the global solution is relatively insensitive to these
1501: details as long as one gets the conserved quantities: mass-to-flux
1502: loading $\beta(\psi)$, angular momentum distribution $J(\psi)$, and
1503: Bernoulli's constant $H(\psi) = 0$ correctly.
1504:
1505: \section{Discussion and Conclusions}
1506:
1507: \subsection{Recipe for Use of Results}
1508:
1509: For the convenience of the reader, we summarize the recipes needed
1510: to convert the results of the previous section into numerical X-wind
1511: models for astronomical and meteoritical applications. Begin with
1512: the equation that describes the dimensionless locus of a streamline
1513: for given $\psi$ with numerical value between 0 and 1:
1514: \begin{equation}
1515: z = z(\varpi,\psi),
1516: \end{equation}
1517: where the functional form of $z(\varpi, \psi)$ is computed
1518: numerically by the technique described in \S 5.2.
1519:
1520: The reconstruction of streamline shapes, i.e., the function
1521: $z(\varpi,\psi)$, is performed over three radial intervals whose end
1522: points are $\varpi_0 \equiv 1$, $\varpi_1(\psi) > 1$,
1523: $\varpi_2(\psi) > \varpi_1(\psi)$, and $\varpi_3(\psi) \equiv
1524: \varpi_\infty(\psi) > \varpi_2(\psi)$ that give a geometrically
1525: increasing separation:
1526: \begin{equation}
1527: \frac{\varpi_2-\varpi_1}{\varpi_1 - 1} = \frac{\varpi_\infty -
1528: \varpi_2}{\varpi_2 - \varpi_1} ,
1529: \label{separation-interval}
1530: \end{equation}
1531: where $\varpi_\infty (\psi)$ is given by equation
1532: \eqref{varpi_infty}:
1533: \begin{equation}
1534: \varpi_\infty = \frac{2 \bar \beta}{C} \frac{\cosh[C^{-1}I(C,
1535: 1)]}{\cosh[C^{-1} I(C, \psi)]} .
1536: \end{equation}
1537: For practical computations, we choose $C = 0.1\bar \beta$ so that
1538: $\varpi_\infty = 20$ on the $\psi =1$ streamline. The asymptotic
1539: integral $I(C, \psi)$ in equation \eqref{sol_asymp} can be
1540: approximated by a seventh degree polynomial in $\beta^{-1}$:
1541: \begin{equation}
1542: I(\psi) = I_0 + I_1 \beta^{-1}(\psi) + \dots + I_7 \beta^{-7}(\psi),
1543: \end{equation}
1544: where the coefficients $I_0$, $I_1$, $\dots$, $I_7$ are given in
1545: Table 2 for the three values of $\bar \beta$ = 1, 2, 3. The function
1546: $\varpi_1(\psi)$ represents the first nontrivial abscissa of the
1547: spline beyond the X-point for each value of $\psi$ and is tabulated
1548: in Table \ref{control_pts} for $\psi_i = 0.0, 0.2, 0.4, 0.6, 0.8,
1549: 0.99$. For intermediate values, we interpolate $\varpi_1$ by a
1550: piecewise cubic polynomial:
1551: \begin{equation}
1552: \varpi_1 = h_0(\psi_i)+h_1(\psi)(\psi-\psi_i) +h_2(\psi_i)(\psi-\psi_i)^2
1553: +h_3(\psi_i)(\psi-\psi_i)^3 \text{ for } \psi_i
1554: \le \psi < \psi_{i+1}.
1555: \end{equation}
1556: In Table \ref{Hyman_coef}, we list the values of $h_j(\psi_i)$ for
1557: each case of $\bar \beta$. To get $\varpi_2(\psi)$ for any value of
1558: $\psi$, one should use equation \eqref{separation-interval} after
1559: first computing $\varpi_1(\psi)$ and $\varpi_\infty(\psi)$ at the
1560: desired value of $\psi$.
1561:
1562: \begin{table}[ht]
1563: \begin{center}
1564: \begin{tabular}{|c|c|c|c|c|c|c|}
1565: \hline & $\psi$ & $0.0$ & $0.2$ & $0.4$ & $0.6$ &
1566: $0.8$\\
1567: \hline$\bar \beta = 1$& $h_0$&$29.687$&$29.817$ & $23.281$ & $18.809$ & $8.310$\\
1568: & $h_1$&$17.315$&$-1.950$ & $-27.520$ & $-37.428$ & $-31.809$\\
1569: & $h_2$&$-153.650$&$-333.100$ & $126.938$ &$-254.103$ &$103.428$ \\
1570: & $h_3$&$351.625$ &$897.250$ &$-505.688$ &$893.830$ & $0.000$\\
1571: \hline $\bar \beta = 2$ &$h_0$ & $28.281$ &$29.165$ & $24.644$ &
1572: $17.774$ & $11.150$\\
1573: &$h_1$ &$17.933$ & $-9.093$ & $-28.478$ & $-33.735$ & $-30.035$\\
1574: &$h_2$ &$-67.563$ &$-105.763$ & $-61.800$ & $-9.272$ & $15.422$\\
1575: &$h_3$ &$0.000$& $191.000$&$162.188$ & $61.737$ & $0.000$\\
1576: \hline $\bar \beta = 1$&$h_0$ & $28.384$ & $28.985$ & $23.990$ &
1577: $19.965$ & $10.139$\\
1578: &$h_1$ & $16.995$ &$-9.015$ & $-22.550$ & $-34.628$ & $-35.107$\\
1579: &$h_2$ & $-79.800$ & $-171.725$ & $96.763$ & $-215.142$ & $70.117$\\
1580: &$h_3$ & $49.250$ &$459.625$ & $-423.188$ & $713.150$ & $0.000$\\
1581: \hline
1582: \end{tabular}
1583: \caption{Interpolation coefficients for $\varpi_1(\psi)$}
1584: \label{Hyman_coef}
1585: \end{center}
1586: \end{table}
1587:
1588: The shape of each streamline given by $\psi$ = const in the three
1589: radial intervals whose end points are $\varpi_0(\psi)=1$,
1590: $\varpi_1(\psi)$, $\varpi_2(\psi)$, and
1591: $\varpi_3(\psi)=\varpi_\infty(\psi)$ is then described by a
1592: piecewise cubic polynomial, whose form, suppressing the implicit
1593: dependence on $\psi$, is given by equation \eqref{y-spline}:
1594: \begin{equation}
1595: z(\varpi) = y_1a + y_2b+ {(\varpi_{i+1}-\varpi_i)^2\over 6}\left[
1596: f_1(a^3-a)+f_2(b^3-b)\right],
1597: \end{equation}
1598: where
1599: \begin{equation}
1600: a\equiv {\varpi_{i+1}-\varpi\over \varpi_{i+1}-\varpi_i}, \qquad b
1601: \equiv {\varpi-\varpi_i\over \varpi_{i+1}-\varpi_i}.
1602: \end{equation}
1603: The coefficients $y_1$, $y_2$, $f_1$, and $f_2$ are listed in Table
1604: \ref{spline_coef} for discrete values of $\psi$ = 0.0, 0.2., 0.4,
1605: 0.6, 0.8, 0.99 in the three cases $\bar \beta = 1$, 2, 3. A Hyman
1606: limited spline may be used to compute the streamlines for other
1607: values of $\psi$.
1608: \begin{table}[ht]
1609: \begin{center}
1610: \begin{tabular}{|c|c|c|c|c|c|c|c|}
1611: \hline &$\psi$ &$0.0$ &$0.2$ &$0.4$ &$0.6$ &$0.8$ &
1612: $0.99$\\
1613: \hline $\bar \beta = 1$ &$y_1$ &$0.0$ & $12.764$ & $21.872$ & $41.255$ & $51.367$
1614: &$525.487$\\
1615: & $y_2$ & $0.0$ & $366.579$ & $627.01$ &$1044.15$ & $1402.6$ &$3648.96$\\
1616: & $f_1$ & $0.0$ & $3.834 \times 10^{-3}$ &$3.083 \times 10^{-2}$ & $0.256$ & $4.120$
1617: & $118.485$\\
1618: & $f_2$ & $0.0$ & $-2.901 \times 10^{-5}$ & $-4.685 \times 10^{-4}$ & $-1.033\times 10^{-2}$
1619: & $-0.301$ & $-48.051$\\
1620: \hline $\bar \beta = 2$ &$y_1$ &$0.0$ &$12.091$ &$22.694$ &$29.659$ & $37.935$ & $73.230$\\
1621: &$y_2$ &$0.0$ & $86.922$ & $156.801$ & $208.943$ & $267.766$ & $383.996$\\
1622: & $f_1$ &$0.0$ & $2.359 \times 10^{-3}$ & $1.489\times 10^{-2}$ &
1623: $9.705 \times 10^{-2}$ & $0.750$ & $9.943$\\
1624: & $f_2$ &$0.0$ & $-1.991 \times 10^{-4}$ & $-1.518 \times 10^{-3}$ &
1625: $ -1.333 \times 10^{-2}$& $-0.158$ & $-4.246$\\
1626: \hline $\bar \beta = 3$ &$y_1$ & $0.0$ & $14.138$ & $22.415$ & $33.841$ & $26.286$ & $44.094$\\
1627: &$y_2$ & $0.0$ & $82.541$ & $104.01$ & $133.299$ & $131.469$ & $173.659$\\
1628: &$f_1$ & $0.0 $ & $1.903 \times 10^{-2}$ & $2.353 \times 10^{-2}$ & $5.056 \times 10^{-2}$
1629: & $0.376$ & $2.950$\\
1630: &$f_2$ & $0.0$ & $-3.651\times 10^{-3}$ & $-5.610 \times 10^{-3}$ &
1631: $-1.722 \times 10^{-2}$ & $-0.102$ & $-1.424$\\
1632: \hline
1633: \end{tabular}
1634: \caption{Spline coefficients for the
1635: streamlines.}\label{spline_coef}
1636: \end{center}
1637: \end{table}
1638:
1639: The partial derivatives of $\psi$ with $\varpi$ or $z$ are now given
1640: by the usual rules of multivariate calculus:
1641: \begin{equation}
1642: \left( {\partial \psi \over \partial \varpi}\right)_z = -{(\partial
1643: z/\partial \varpi)_\psi\over (\partial z/ \partial \psi)_\varpi};
1644: \qquad \left( {\partial \psi\over \partial z}\right)_\varpi =
1645: {1\over (\partial z/\partial \psi)_\varpi}.
1646: \end{equation}
1647: Table 3 gives $J(\psi)$ as a seventh order polynomial in $\beta(\psi)$:
1648: \begin{equation}
1649: J(\psi) = J_0 + J_1\beta(\psi) + \dots +J_7\beta^7(\psi),
1650: \end{equation}
1651: where $\beta(\psi)$ is itself given by
1652: \begin{equation}
1653: \beta(\psi) = {2\over 3}\bar \beta (1-\psi)^{-1/3},
1654: \end{equation}
1655: with $\bar \beta$ = 1, 2, 3 in the three chosen model cases. The
1656: coefficients tabulated in Table 3 give a $J(\psi)$ that guarantees
1657: that equation \eqref{BE},
1658: \begin{equation}
1659: \abs{\Del \psi}^2 + \oneover{\A^2}\prn{\frac{J}{\varpi^2} -1 }^2 +
1660: \frac{2\varpi^2 \Veff}{(\beta^2 - \varpi^2 \A)^2} =
1661: 0,
1662: \label{Bernoulli}
1663: \end{equation}
1664: has one real root for $\cal A$ in the computational domain when
1665: $V_{\rm eff}$ is given by equation \eqref{Veff}:
1666: \begin{equation}
1667: \Veff = -\frac{1}{\sqrt{\varpi^2 + z^2}} -\ahalf \varpi^2 +
1668: \frac{3}{2}.
1669: \end{equation}
1670: By solving equation (\ref{Bernoulli}) as a fourth-order polynomial,
1671: we may obtain the relevant value for the Alfv\'en discriminant $\cal
1672: A$. Then the density can be computed through equation
1673: \eqref{calAdef}
1674: \begin{equation}
1675: \rho = (\beta^2 - \varpi^2
1676: \A)^{-1}.
1677: \label{density}
1678: \end{equation}
1679: Note that this equation produces $\rho = \beta^{-2}$ at the
1680: Alfv\'enic transition ${\cal A} = 0$.
1681:
1682: With the density in place, we may obtain the two components of dimensionless
1683: poloidal velocity from the definition \eqref{psidef} of $\psi$:
1684: \begin{equation}
1685: u_\varpi \equiv \oneover{\varpi\rho} \pd{\psi}{z}, \quad
1686: u_z \equiv - \oneover{\varpi\rho} \pd{\psi}{\varpi}.
1687: \end{equation}
1688: The toroidal velocity in the corotating frame is given by equation \eqref{Jdef}
1689: \begin{equation}
1690: u_\varphi = {J(\psi)-\varpi^2\over \varpi (1-\beta^2\rho)}.
1691: \label{toroidalvel}
1692: \end{equation}
1693: Note that $J(\psi) = \varpi^2$ where $\beta^2\rho = 1$ keeps
1694: the toroidal velocity $u_\varphi$ well-behaved across the Alfv\'en surface, which is
1695: not one of the critical surfaces of the overall problem.
1696:
1697: The vector magnetic field may now be obtained from equation \eqref{betadef}:
1698: \begin{equation}
1699: {\bf B} = \beta \rho {\bf u},
1700: \label{vectorB}
1701: \end{equation}
1702: whereas
1703: the azimuthal velocity in the inertial frame is given by
1704: \begin{equation}
1705: v_\varphi = u_\varphi + \varpi ,
1706: \label{inertialvphi}
1707: \end{equation}
1708: with the term $\varpi$ from the frame rotation being cancelled at
1709: large $\varpi$ where $u_\varphi \rightarrow -\varpi$ because $\rho$
1710: vanishes as $1/\varpi^2$ at large distances from the rotation axis.
1711: Finally, to convert the computed quantities to their dimensional
1712: counterparts, we must multiply velocities, densities, and magnetic
1713: fields by $R_X\Omega_X$, $\dot M_w/4\pi R_X^3\Omega_X$, and
1714: $(\Omega_X\dot M_w/R_X)^{1/2}$, respectively.
1715:
1716: For interpolations or extrapolations in $\bar \beta$, we recommend
1717: computation first of the dimensionless density, velocity, and
1718: magnetic fields for the three cases $\bar \beta =$ 1, 2, 3, and then
1719: direct interpolations or extrapolations of those fields. Other
1720: techniques starting farther back in the process run the danger of
1721: obtaining complex roots of $\cal A$ (i.e., complex values of $\rho$)
1722: from the solution of the quartic equation (\ref{Bernoulli}) because
1723: of slight inaccuracies in computing the numerical coefficients.
1724:
1725: \subsection{Summary}
1726:
1727: In this paper, we have presented a technique by which solutions to
1728: the so-called Grad-Shafranov equation for X-wind flow can be solved,
1729: not by attacking the partial differential equation directly, but by
1730: choosing trial functions that minimize an appropriate action
1731: integral. While this method has been applied before in problems of
1732: plasma confinement in the fusion community, we believe that the
1733: example given here is its first application in astrophysics for the
1734: notorious case when magnetohydrodynamical flows cross critical
1735: surfaces that change the character of the underlying PDE.
1736:
1737: Many empirical arguments suggest that funnel flows and X-winds do
1738: underlie the accretion hot-spots, jets, and winds of YSOs, although
1739: a dipolar field geometry near the star (see Fig. 1) may be an
1740: over-simplification (Ardila et al. 2002, Unruh et al. 2004,
1741: Johns-Krull 2007). Fortunately, although the fractional areal
1742: coverage of hot spots depends on the detailed multipole structure of
1743: the surfaces of actual young stars, the general validity of X-wind
1744: theory depends only on the level of trapped flux in the X-region and
1745: is insensitive to the magnetic geometry on the star as long as the
1746: fields are strong (Mohanty \& Shu 2007). The trapped flux in the
1747: X-wind models of this paper are computed as
1748: \begin{equation}
1749: 2\pi \bar\beta \left(GM_*\dot M_w\over \Omega_X\right)^{1/2},
1750: \end{equation}
1751: and should be compared with the magnetic flux (area times mean
1752: field) in hot-spots on one hemisphere's surface of the star impacted
1753: by the corresponding funnel flow. (Both fluxes are 1/3 of the total
1754: trapped flux in the X-region and equal the net flux of the dead
1755: region.) For T Tauri stars, the comparison is pretty good (see,
1756: e.g., Johns-Krull \& Gafford 2002).
1757:
1758: Apart from relative simplicity, the semi-analytical solutions
1759: summarized in \S 7.1 have many other advantages. For example, the
1760: solutions hold over a formally infinite dynamic range, showing the
1761: asymptotic, logarithmically slow, collimation into jets missing in
1762: many numerical simulations. These properties make the models of
1763: this paper especially suitable for a wide variety of astronomical
1764: and meteoritical applications, such as detailed comparisons with
1765: observations, trajectories of solids entrained in the wind, and
1766: interactions with neighboring circumstellar or interstellar matter.
1767: A needed generalization for future research is the inclusion of the
1768: effects of the intrinsic magnetization of the surrounding accretion
1769: disk.
1770:
1771: \bigskip\bigskip
1772: We thank the Physics Department and the Center for Astrophysics and
1773: Space Sciences of UCSD for support. The Academia Sinica and the
1774: National Science Council (NSC) of Taiwan also provided funding
1775: through their grants to the Theoretical Institute for Advanced
1776: Research in Astrophysics (TIARA). The research of MJC is supported
1777: in part by the NSC grant 95-2112-M-001-44.
1778:
1779: \appendix \section{Character of Governing Equation}\label{Character}
1780:
1781: The GSE \eqref{GSE-general} resembles the steady state heat
1782: diffusion equation with a variable diffusion coefficient $\A$. This
1783: analogy is actually misleading since we do not know its overall
1784: character until we substitute in the implicit dependence of $\A$ on
1785: $\psi$ by solving the (algebraic) BE and examine the characteristics
1786: of the GSE. To do so, let us first differentiate the BE with
1787: respect to $\varpi$ and $z$.
1788: \begin{eqnarray*}
1789: &&2(\psi_{,\varpi} \psi_{,\varpi \varpi} + \psi_{,z}\psi_{,z \varpi}) -
1790: \frac{2\A_{,\varpi}}{\A^3}\prn{\frac{J}{\varpi^2} -1}^2 \\
1791: &&\qquad +
1792: \frac{2\varpi^4 \A_{,\varpi}}{(\oneoverrho)^3}\brk{2\Veff +
1793: 2\epsilon^2 \ln \prn{\frac{\epsilon^2 h}{\oneoverrho}}+ \epsilon^2}
1794: + ... = 0,\\
1795: &&2(\psi_{,\varpi} \psi_{,\varpi z} + \psi_{,z}\psi_{,z z}) -
1796: \frac{2\A_{,z}}{\A^3}\prn{\frac{J}{\varpi^2} -1}^2 \\
1797: &&\qquad +
1798: \frac{2\varpi^4 \A_{,z}}{(\oneoverrho)^3}\brk{2\Veff +
1799: 2\epsilon^2 \ln \prn{\frac{\epsilon^2 h}{\oneoverrho}}+ \epsilon^2}
1800: + ... = 0,
1801: \end{eqnarray*}
1802: where in the above equations, a subscript denotes partial derivative
1803: and the ellipsis symbols include terms that are irrelevant in
1804: determining the character of the GSE. These equations may be solved
1805: for $\A_{,\varpi}$ and $\A_{,z}$ to give
1806: \begin{displaymath}
1807: \A_{,\varpi} = \oneover{\mathcal{P}}\prn{\psi_{,\varpi} \psi_{,\varpi\varpi} +
1808: \psi_{,z} \psi_{,z\varpi}} + ..., \qquad \A_{,z} = \oneover{\mathcal{P}}
1809: \prn{\psi_{,\varpi} \psi_{,\varpi z} + \psi_{,z}\psi_{,zz}} + ...,
1810: \end{displaymath}
1811: where
1812: \begin{displaymath}
1813: \mathcal{P} =
1814: \frac{\varpi^2}{\oneoverrho}
1815: \abs{\Del \psi}^2 + \oneover{\A^3}\prn{\frac{J}{\varpi^2} -1 }^2 \frac{\beta^2}{\oneoverrho}
1816: - \frac{\epsilon^2\varpi^4}{(\oneoverrho)^3},
1817: \end{displaymath}
1818: after we eliminate $\Veff$ in the expression by using the BE
1819: \eqref{BE}. The second derivative terms in the GSE \eqref{GSE} can
1820: now be written in the form
1821: \begin{displaymath}
1822: a \psi_{,\varpi\varpi} + 2b\psi_{,\varpi z} + c \psi_{,zz} + ... =
1823: 0,
1824: \end{displaymath}
1825: where
1826: \begin{displaymath}
1827: a = \A + \frac{\psi_{,\varpi}^2}{\mathcal{P}}, \qquad b =
1828: \frac{\psi_{,\varpi} \psi_{,z}}{\mathcal{P}}, \qquad c = \A +
1829: \frac{\psi_{,z}^2}{\mathcal{P}}.
1830: \end{displaymath}
1831: The character of the GSE is determined by the quantity $\Delta = b^2
1832: - ac$ (Garabedian 1986): it is elliptic, parabolic, or
1833: hyperbolic if $\Delta$ is negative, zero, or positive. We may
1834: compute $\Delta$ for our GSE explicitly.
1835: \begin{equation}
1836: \Delta = - \A^2\left\{
1837: {|\Del \psi|^2 + (J\varpi^{-2} -1 )^2\A^{-2}
1838: - \epsilon^2\A \varpi^4[\beta (\beta^2-\varpi^2\A)]^{-2}\over
1839: \varpi^2 \A\beta^{-2} |\Del \psi|^2 + (J\varpi^{-2} -1)^2\A^{-2}
1840: - \epsilon^2 \A \varpi^4[\beta (\beta^2-\varpi^2\A)]^{-2}}\right\}.
1841: \end{equation}
1842: The interpretation of this expression becomes transparent if we
1843: transform back into the physical quantities. After some algebra, we
1844: have
1845: \begin{equation}
1846: \Delta = - \A^2\brk{
1847: \frac{u^2
1848: - \epsilon^2(1-M_A^2)}{
1849: (1-M_A^2) (u_p^2-\epsilon^2) + u_\varphi^2}}
1850: % = \A^2\brk{
1851: % \frac{(v_A^2 + \epsilon^2) u_p^2
1852: % - \epsilon^2v_{Ap}^2}{u_p^4 -
1853: % u_p^2 (v_A^2 + \epsilon^2)+\epsilon^2v_{Ap}^2}}
1854: =\A^2\brk{
1855: \frac{(v_A^2 + \epsilon^2) (u_p^2
1856: - v_s^2)}{(u_p^2 - v_{-p}^2)(u_p^2 - v_{+p}^2)
1857: }}.\label{PDE_character}
1858: \end{equation}
1859: where $v_{Ap} \equiv \sqrt{B_p^2/\rho}$ is the poloidal component of
1860: the Alfv\'en velocity $v_A \equiv \sqrt{B^2/\rho}$ with $B^2 =
1861: B_p^2+B_\varphi^2$, $u_p$ is the poloidal fluid velocity, and $v_s$
1862: is defined by
1863: \begin{displaymath}
1864: v_s^2 \equiv \frac{\epsilon^2v_{Ap}^2}{v_A^2 + \epsilon^2}.
1865: \end{displaymath}
1866: In the limit where $v_A^2 \gg \epsilon^2$, it reduces to the thermal
1867: sound speed. In addition, $v_{\pm p}$ denote the poloidal component
1868: of the fast and slow MHD wave speeds, respective, and are given by
1869: \begin{equation}
1870: v_{\pm p}^2 = \ahalf (v_A^2 + \epsilon^2)\brk{1 \pm \sqrt{1- \frac{4v_s^2}
1871: {v_A^2 + \epsilon^2}}}.
1872: \end{equation}
1873: A moment of thought reveals that $v_s < v_{-p} < v_{+p}$. The
1874: significance of the equation \eqref{PDE_character} is now clear. The
1875: governing GSE is elliptic when $u_p^2 < v_s^2$ or $v_{-p}^2 < u_p^2
1876: < v_{+p}^2$, and it is hyperbolic when $v_s^2 < u_p^2 < v_{-p}^2$ or
1877: $u_p^2 > v_{+p}^2$ \citep{Heinemann:1978,Sakurai:1985}.
1878:
1879: To be definite, we shall refer to the loci where the poloidal
1880: velocity squared equals $v_s^2$, $v_{-p}^2$ and $v_{+p}^2$ as
1881: sonic, slow, and fast surfaces, respectively. Despite
1882: the deceiving appearance of the GSE \eqref{GSE-general}, note that it does not
1883: change character on the Alfv\'en surface when $\A = 0$ (or
1884: equivalently when $M_A = 1$ and the total fluid velocity in the corotating frame
1885: equals the total Alfv\'en speed); it remains elliptic until the fast
1886: surface. The fact that the asymptotic flow is described by a
1887: hyperbolic PDE is consistent with our physical intuition. When the
1888: fluid speed is super-magnetosonic, no information can be sent
1889: upstream into the flow. Thus the asymptotic behavior of the X-wind is determined
1890: by the ``initial condition'' at the place when the fluid velocity
1891: first becomes equal to the fastest signal propagation speed -- a
1892: defining feature of hyperbolic problems.
1893:
1894: In the cold limit the discriminant $\Delta$ has the simplification,
1895: \begin{equation}
1896: \Delta = -\A^2\left\{
1897: {|\Del \psi|^2 + (J\varpi^{-2} -1 )^2\A^{-2} \over
1898: \varpi^2 \A\beta^{-2} |\Del \psi|^2 + (J\varpi^{-2} -1)^2\A^{-2}}\right\}
1899: = -\A^2\left(\frac{1
1900: }{1 -u_p^2/v_A^2}\right).
1901: \end{equation}
1902: This equation explicitly states that the transition to the
1903: hyperbolic portion of the solution is done through the fast surface,
1904: where the poloidal fluid velocity is equal to the magnetosonic
1905: speed, which is the total Alfv\'en speed $v_A = B/\sqrt{\rho}$ when
1906: $\epsilon$ is set to zero. The axial symmetry of the assumed
1907: problem guarantees that any compressions or rarefactions occur only
1908: in the meridional plane, so the relevant speed of signal propagation
1909: in the limit $\epsilon \rightarrow 0$ is the magnetosonic speed
1910: relative to the poloidal motion of the fluid.
1911: %Furthermore, since
1912: %this transition is made through a diverging $\Delta$, the
1913: %characteristics of the flow are degenerate on the fast surface, and
1914: %one would need an infinite numerical accuracy and resolution to
1915: %handle this portion of the calculation via a conventional
1916: %computational scheme.
1917:
1918: \newpage
1919: \begin{thebibliography}{}
1920: \bibitem[Ardila et al.(2002)]{Ardila:2003}
1921: Ardila, D. R., Basri, G., Walter, F. M., Valenti, J. A.,
1922: Johns-Krull, C. M. 2002, ApJ, 567, 1013
1923:
1924: \bibitem[Askey(1982)]{Askey:1982}
1925: Askey, R. 1982 Lett. in Math. Phys, 6, 299
1926:
1927: \bibitem[Bacciotti et al(2002)]{Bacciotti:2002}
1928: Bacciotti, F., Ray, T. P., Eisl\"offel, J., Solf, J. 2002, ApJ, 576, 222
1929:
1930: \bibitem[Blandford \& Rees(1992)]{Blandford:1992}
1931: Blandford, R. D., Rees, M. J. 1992, in Testing the AGN Paradigm, AIP
1932: Conf. Proc., 254, 3
1933:
1934: \bibitem[Blandford \& Payne(1982)]{Blandford:1982}
1935: Blandford, R. D., Payne, D. G. 1982 MNRAS, 199, 883
1936:
1937: \bibitem[Bouvier(1990)]{Bouvier:1990}
1938: Bouvier, J. 1990, AJ, 99, 946
1939:
1940: \bibitem[Cabrit et al(2006)]{Cabrit:2006}
1941: Cabrit, S., Pety. J., Presenti, N., Dougados, C. 2006, A\&A, 452, 897
1942:
1943: \bibitem[Camenzind(1987)]{Camenzind:1987}
1944: Camenzind, M. 1987, A\&A, 184, 341
1945:
1946: \bibitem[Coffey et al(2004)]{Coffey:2004}
1947: Coffey, D., Bacciotti, F., Woitas, J., Ray, T. P., Eisl\"offel, J.
1948: 2004, ApJ, 604, 708
1949:
1950: \bibitem[Coffey et al(2007)]{Coffey:2007}
1951: Coffey, D., Bacciotti, F., Ray, T. P., Esil\"offel, J., Woitas, J.
1952: 2007, ApJ, 663, 350
1953:
1954: \bibitem[Contopoulos \& Lovelace(1994)]{Contopoulos:1994}
1955: Contopoulos, J., Lovelace, R.V.E. 1994, ApJ, 429, 139
1956:
1957: \bibitem[Edwards et al.(1993)]{Edwards:1993}
1958: Edwards, S., Ray, T., Mundt, R. 1993, in Protostars \& Planets III,
1959: ed. E. H. Levy \& J. Lunine (Tucson: Univ. of Arizona Press), 567
1960:
1961: \bibitem[Ferreira(2004)]{Ferreira:2004}
1962: Ferreira, J. 2004, Ap\&SS, 293, 83
1963:
1964: \bibitem[Garabedian(1986)]{Garabedian:1986}
1965: Garabedian, P. R. 1986, Partial Differential Equations, (Providence:
1966: Americal Mathematical Society)
1967:
1968: \bibitem[Giampapa et al(1993)]{Giampapa:1993}
1969: Giampapa, M.S., Basri, G., Johns, C.M., Imhoff, C.L. 1993, ApJS, 89, 321
1970:
1971: \bibitem[Goldreich \& Julian(1970)]{Goldreich:1970}
1972: Goldreich, P., Julian, W. H. 1970, ApJ, 160, 971
1973:
1974: \bibitem[Goodson et al(1999)]{Goodson:1999}
1975: Goodson, A. P., Bohm, K. H., Winglee, R. M. 1999, ApJ, 524, 142
1976:
1977: \bibitem[Grad \& Rubin(1958)]{Grad:1958}
1978: Grad, H., Rubin, H. 1958, in Proc. Conf. Internat. Atomic Energy
1979: Agency 31, (Geneva: Internat. Atomic Energy Agency)
1980:
1981: \bibitem[Heinemann \& Olbert(1978)]{Heinemann:1978}
1982: Heinemann, M., Olbert, S. 1978, JGR, 83, 2457
1983:
1984: \bibitem[Heyvaerts \& Normal(1997)]{Heyvaerts:1997}
1985: Heyvaerts, J., Norman, C.A. 1997, in Herbig--Haro Flows and the
1986: Birth of Stars, IAU Symp. 182, ed. B. Reipurth \& C. Bertout
1987: (Kluwer), 275
1988:
1989: \bibitem[Hyman(1983)]{Hyman:1983}
1990: Hyman, J. 1983, SIAM, J. Sci. Stat. Comput., 4, 645
1991:
1992: \bibitem[Jackson(1975)]{Jackson:1975}
1993: Jackson, J. D. 1975, Classical Electrodynamics, (New York: John
1994: Wiley \& Sons)
1995:
1996: \bibitem[Johns \& Basri(1995)]{Johns:1995}
1997: Johns, C., Basri, G. 1995, ApJ, 449. 341
1998:
1999: \bibitem[Johns-Krull(2007)]{Johns-Krull:2007}
2000: Johns-Krull, C. M. 2007, arXiv0704.2923v1 [astro-ph]
2001:
2002: \bibitem[Johns-Krull \& Gafford(2002)]{Johns-Krull:2002}
2003: Johns-Krull, C. M., Gafford, A. D. 2002, ApJ, 573, 685
2004:
2005: \bibitem[Johna-Krull \& Hatzes(1997)]{Johns-Krull:1997}
2006: Johns-Krull, C.M., Hatzes, A.P. 1997, ApJ, 487, 896
2007:
2008: \bibitem[K\"onigl \& Pudritz(2000)]{Konigl:2000}
2009: K\"onigl, A., Pudritz, R. E. 2000, in Protostars \& Planets IV, ed.
2010: V. Mannings, A. P. Boss, \& S. S. Russell (Tucson: Univ. of Arizona
2011: Press), 759
2012:
2013: \bibitem[Long et al(2005)]{lrl07}
2014: Long, M., Romanova, M. M., Lovelace, R. V. E. 2005, ApJ, 634, 1214
2015:
2016: \bibitem[Lovelace et al(1986)]{Lovelace:1986}
2017: Lovelace, R.V.E., Mehanian, C., Mobarry, C.M., Sulkanen, M.E. 1986,
2018: ApJS, 62, 1
2019:
2020: \bibitem[Mestel(1968)]{Mestel:1968}
2021: Mestel, L. 1968, MNRAS, 138, 359
2022:
2023: \bibitem[Mohanty \& Shu(2007]{Mohanty:2007}
2024: Mohanty, S., Shu, F.H. 2007, ApJ, in preparation
2025:
2026: \bibitem[Najita et al.(2007)]{Najita:2007}
2027: Najita, J. R., Carr, J. S., Glassgold, A. E., Valenti. J. A. 2007,
2028: Protostars \& Planets V, ed. B. Reipurth, D. Jewitt, \& K. Keil
2029: (Tucson: Univ. of Arizona Press), 507
2030:
2031: \bibitem[Najita \& Shu(1994)]{paper3}
2032: Najita, J. R., Shu, F. H. 1994, ApJ, 429, 808
2033:
2034: \bibitem[Ostriker \& Shu(1995)]{paper4}
2035: Ostriker, E. C., Shu, F. H. 1995, ApJ, 447, 813
2036:
2037: \bibitem[Pety et al(2006)]{Pety:2006}
2038: Pety, J., Gueth, F., Guilloteau, S., Dutrey, A. 2006, A\&A, 458, 841
2039:
2040: \bibitem[Press(1992)]{Press:1992}
2041: Press, W. H., Teukolsky, S. A., Vetterling, W. T., \& Flannery, B.
2042: P. 1992, Numerical Recipes in C++ (Cambridge University Press)
2043:
2044: \bibitem[]{}
2045: Pudritz, R.E., Rogers, C.S., Ouyed, R. 2006, MNRAS, 365, 113
2046:
2047: \bibitem[Ramsay(1988)]{Ramsay:1988}
2048: Ramsay, J. O. 1988, Statist. Sci., 3, 425
2049:
2050: \bibitem[Ramsay(1998)]{Ramsay:1998}
2051: Ramsay, J. O. 1998, J. R. Statist. Soc. B, 60, Part 2, 365
2052:
2053: \bibitem[Sakurai(1985)]{Sakurai:1985}
2054: Sakurai, T. 1985, A \& A, 152, 121
2055:
2056: \bibitem[Shafranov(1966)]{Shafranov:1966}
2057: Shafranov, V. D. 1966, Rev. Plasma Phys, 2, 103
2058:
2059: \bibitem[Shang(1998)]{Shang:1998a}
2060: Shang, H. 1998, PhD Thesis, Univ. of California, Berkeley
2061:
2062: \bibitem[Shang et al.(1998)]{Shang:1998b}
2063: Shang, H., Shu, F. H., Glassgold, A. E. 1998, ApJ, 493, 91
2064:
2065: \bibitem[Shang et al.(2002)]{Shang:2002}
2066: Shang, H., Glassgold, A. E., Shu, F. H., Lizano, S. 2002, ApJ, 564,
2067: 853
2068:
2069: \bibitem[Shang et al.(2004)]{Shang:2004}
2070: Shang, H., Lizano, S., Glassgold, A. E., Shu, F. H. 2004, ApJ, 612,
2071: L69
2072:
2073: \bibitem[Shu et al. (1988)]{paper0}
2074: Shu, F. H., Lizano, S., Ruden, S. P., Najita, J. 1988, ApJ, 328, L19
2075:
2076: \bibitem[Shu(1992)]{Shu:1992}
2077: Shu, F. H. 1992, The Physics of Astrophysics, Vol II: Gas Dynamics,
2078: (Mill Valley: University Science Books)
2079:
2080: \bibitem[Shu et al.(1994a)]{paper1}
2081: Shu, F. H., Najita, J., Ostriker, E., Wilkin, F., Ruden, S.,
2082: Lizano, S. 1994 ApJ, 429, 781
2083:
2084: \bibitem[Shu et al.(1994b)]{paper2}
2085: Shu, F. H., Najita, J., Ruden, S. P., Lizano, S. 1994b, ApJ, 429,
2086: 797
2087:
2088: \bibitem[Shu et al.(1995)]{paper5}
2089: Shu, F. H., Najita, J., Ostriker, E. C., Shang, H. 1995 ApJ, 455,
2090: L155
2091:
2092: \bibitem[Shu et al.(2000)]{Shu:2000}
2093: Shu, F. H., Najita, J. R., Shang, H., Li, Z.-Y., 2000, in
2094: Protostars \& Planets IV, ed. V. Mannings, A. P. Boss, \& S. S.
2095: Russell (Tucson: Univ. of Arizona Press), 789
2096:
2097: \bibitem[Shu et al.(2001)]{Shu:2001}
2098: Shu, F. H., Shang, H., Gounelle, M., Glassgold, A. E., Lee, T. 2001,
2099: ApJ, 548, 1029
2100:
2101: \bibitem[Tsinganos \& Trussoni(1991)]{Tsinganos:1991}
2102: Tsinganos, K., Trussoni, E. 1991, A\&A, 249, 156
2103:
2104: \bibitem[Uchida \& Shibata(1985)]{Uchida:1985}
2105: Uchida, Y., Shibata, K. 1985, PASJ, 37, 515
2106:
2107: \bibitem[Unruh et al.(2004)]{Unruh:2004}
2108: Unruh, Y. C. et al. 2004, MNRAS, 348, 1301
2109:
2110: \bibitem[Ustyugova et al(2006)]{Ustyugova:2006}
2111: Ustyugova, G.V., Koldoba, A., Romanova, M.M., Lovelace, R.V.E. 2006, 646, 304
2112:
2113: \bibitem[Van Dyke(1964)]{van Dyke:1964}
2114: Van Dyke, M. 1964 Perturbation Methods in Fluid Mechanics, (New
2115: York: Academic Press)
2116:
2117: \bibitem[Vogel \& Kuhi(1981)]{Vogel:1981}
2118: Vogel. S., Kuhi, L. 1981, ApJ, 245, 960
2119:
2120: \bibitem[Weber \& Davis(1967)]{Weber:1967}
2121: Weber, E. J., Davis, L. 1967, ApJ, 148, 217
2122: \end{thebibliography}
2123:
2124: \end{document}
2125: