0804.4201/ms.tex
1: %\documentclass{emulateapj}
2: 
3: \NeedsTeXFormat{LaTeX2e}[1996/06/01]
4: 
5: \documentclass[12pt,preprint]{aastex}
6: %\usepackage{epsfig}
7: \def\lsim{\raise0.3ex\hbox{$<$}\kern-0.75em{\lower0.65ex\hbox{$\sim$}}}
8: \def\gsim{\raise0.3ex\hbox{$>$}\kern-0.75em{\lower0.65ex\hbox{$\sim$}}}
9: 
10: \begin{document}
11: \title{Magnetically Regulated Star Formation in 3D: The Case of 
12: Taurus Molecular Cloud Complex} 
13:  
14: \author{Fumitaka Nakamura\altaffilmark{1} and Zhi-Yun Li\altaffilmark{2}} 
15: %
16: \altaffiltext{1}{Faculty of Education, Niigata University,
17: 8050 Ikarashi-2, Niigata 950-2181, Japan; fnakamur@ed.niigata-u.ac.jp}
18: \altaffiltext{2}{Department of Astronomy, University of Virginia,
19: P. O. Box 400325, Charlottesville, VA 22904; zl4h@virginia.edu}
20: 
21: \begin{abstract}
22: We carry out three-dimensional MHD simulations of star formation
23: in turbulent, magnetized clouds, including ambipolar diffusion
24: and feedback from protostellar outflows. The calculations focus
25: on relatively diffuse clouds threaded by a strong magnetic field 
26: capable of resisting severe tangling by turbulent motions and 
27: retarding global gravitational contraction in the cross-field 
28: direction. They are motivated by observations of the Taurus 
29: molecular cloud complex (and, to a lesser extent, Pipe Nebula), 
30: which shows an ordered large-scale magnetic field, as 
31: well as elongated condensations that are generally perpendicular 
32: to the large-scale field. We find that stars form in earnest in 
33: such clouds when enough material has 
34: settled gravitationally along the field lines that the 
35: mass-to-flux ratios of the condensations approach the critical 
36: value. Only a small fraction (of order $1\%$ or less) of the 
37: nearly magnetically-critical, 
38: condensed material is turned into 
39: stars per local free-fall time, however. The slow star formation 
40: takes place in condensations that are moderately supersonic; 
41: it is regulated primarily by magnetic fields, rather than turbulence. 
42: The quiescent condensations are surrounded by diffuse halos 
43: that are much more turbulent, as observed in the Taurus complex. 
44: Strong support for magnetic regulation of star formation in 
45: this complex comes from the extremely slow conversion of the 
46: already condensed, relatively quiescent C$^{18}$O gas into 
47: stars, at a rate two orders of magnitude below the maximum, 
48: free-fall value.
49: We analyze the properties of dense 
50: cores, including their mass spectrum, which resembles the 
51: stellar initial mass function. 
52: \end{abstract}
53: 
54: \keywords{ISM: clouds --- ISM: magnetic fields --- MHD --- stars: formation
55: --- turbulence}
56: 
57: 
58: \section{Introduction}
59: \label{intro}
60: 
61: Stars form in molecular clouds that are both turbulent and magnetized. 
62: The relative importance of the turbulence and magnetic field in 
63: controlling star formation is a matter of debate (McKee \& Ostriker 
64: 2007). Early quantitative studies have concentrated on the formation 
65: and evolution of individual (low-mass) cores out of quiescent, 
66: magnetically supported clouds (Nakano 1984; Shu et al. 1987; 
67: Mouschovias \& Ciolek 1999). Recent numerical simulations have 
68: focused more on the role of turbulence in the dynamics of larger 
69: clouds and core formation in them (as reviewed in, e.g., Mac Low 
70: \& Klessen 2004). Ultimately, the debate must be settled by 
71: observations, especially direct measurements of the magnetic 
72: field strength in the bulk of the molecular gas, which are 
73: generally difficult to do (e.g., Heiles \& Crutcher 2005). In the 
74: absence of such measurements, we have to rely on indirect evidence.
75: 
76: The best observed region of active star formation is arguably 
77: the nearby Taurus molecular cloud complex. Although it is not 
78: clear whether the complex is magnetically supported or not as 
79: a whole, the more diffuse regions are probably magnetically 
80: dominated. The strongest evidence comes from thin strands of 
81: $^{12}$CO emission that are aligned with the directions of 
82: the local magnetic field (Heyer et al. 2008), as traced by 
83: polarization of background 
84: star light. The magnetic field in the strands is apparently
85: strong enough to induce a measurable difference between the
86: turbulent velocities along and perpendicular to the field
87: direction. Heyer et al. concluded that the field strength 
88: is probably high enough to render the relatively diffuse 
89: striated region subcritical, with a magnetic flux-to-mass 
90: ratio greater than the critical value $2\pi G^{1/2}$ (Nakano 
91: \& Nakamura 1978). The CO striations are strikingly 
92: similar to those observed in the nearby Riegel-Crutcher HI 
93: cloud, mapped recently by McClure-Griffiths et al. (2006), 
94: using 21cm absorption against the strong continuum emission
95: towards the galactic center region. Its hair-like strands 
96: are also along the directions of the local magnetic field, 
97: again traced by the polarization vectors of the background 
98: star light. The authors estimated a field strength of $\sim 
99: 30$~$\mu$G inside the strands. Kazes \& Crutcher (1986) 
100: measured the line-of-sight field strength at a nearby 
101: location, and found $B_{los}\sim 18~\mu$G, which is 
102: consistent with the above estimate if a correction of $\sim 
103: 2$ is applied for projection effect. These are clear examples 
104: of magnetic domination in, respectively, cold HI
105: % (see also Orion Veil, Brogan et al.??) 
106: and relatively diffuse molecular clouds. 
107: 
108: It is unclear, however, how representative these two clouds are. 
109: In particular, the degree of magnetization is uncertain in 
110: giant molecular clouds (GMCs), where the majority of stars 
111: form. Elmegreen (2007) argued that the envelopes of GMCs, where 
112: most of the molecular gas resides, are probably magnetically 
113: critical or even subcritical, and can be long-lived if not for 
114: disruption by rapid star formation in GMC cores, which are 
115: probably supercritical, with internal dynamics dominated by 
116: turbulence; this supposition needs to be tested by magnetic 
117: field observations on GMC scales (e.g., Novak, et al. 2007). 
118: Nevertheless, the 
119: Taurus cloud complex and the Riegel-Crutcher cloud are probably  
120: best mapped cloud in CO and HI, respectively. How such relatively 
121: diffuse, apparently magnetically dominated clouds proceed to 
122: condense and collapse to form stars is the focus of our paper. 
123: 
124: Diffuse clouds are expected to condense more easily along than 
125: across the field lines as long as they are dominated by ordered 
126: large-scale magnetic fields. Such an anisotropic condensation 
127: can increase the density greatly with little corresponding 
128: increase in the magnetic field strength. This expectation is 
129: consistent with the available Zeeman measurements of field 
130: strength, which is more or less constant below a number density 
131: of order $10^3$~cm$^{-3}$ (Heiles \& Crutcher 2005). Above this 
132: density, the field strength tends to increase with density, indicating 
133: that contraction perpendicular to the field lines becomes 
134: significant. We interpret the break in the observed field 
135: strength-density relation as marking the point where the 
136: self-gravity of the cloud becomes dynamically important in the 
137: cross-field direction. Before this point is reached, one expects 
138: the self-gravity to first become significant along the field 
139: lines (where there is no magnetic support), pulling matter into 
140: condensations, which can be either a sheet, a filament, 
141: or even a knot, depending on the degree of anisotropy in the 
142: initial mass distribution. 
143: 
144: In the case of the Taurus molecular cloud complex, most of the 
145: dense molecular gas traced by $^{13}$CO and especially C$^{18}$O 
146: are distributed
147: in structures elongated more or less perpendicular to the 
148: large-scale magnetic field (e.g., Onishi et 
149: al. 1996, 2002; Goldsmith et al. 2008). It has long been suspected 
150: that they have condensed along the field lines (e.g., Heyer et 
151: al. 1987; Tamura et al. 1987). Cross-field contraction must have 
152: already taken place in the condensed structures, at least locally, 
153: since stars have been forming in these structures for at least
154: a few million years (Kenyon \& Hartmann 1995). Palla \& Stahler
155: (2002) examined the star formation pattern in the Taurus
156: clouds in both space and time, concluding that stars began 
157: to form at a low level at least 10 million years ago, in 
158: a spatially dispersed fashion. The majority of the stars are
159: formed, however, in the last 3 million years, near where the 
160: dense gas is observed today. This pattern of accelerating 
161: star formation is precisely what is expected of a magnetically
162: dominated cloud that condenses in two stages: first along the
163: field line, when little star formation is expected, except 
164: perhaps in some pockets that have an exceptionally weak local 
165: magnetic field to begin with or have their magnetic fluxes 
166: reduced by an exceptionally strong compression-enhanced ambipolar 
167: diffusion, and then across the field, when more active star
168: formation can take place throughout the dense sheets and 
169: filaments that have condensed along the field lines. A strong 
170: support for this supposition comes from the fact that star 
171: formation is slow even in dense gas: the rate of star 
172: formation for the last 3 million years is about ${\dot M}_*
173: \sim 5\times 10^{-5}~M_\odot$yr$^{-1}$ (Goldsmith et al. 2008), 
174: two orders of magnitude below the free-fall rate for the dense 
175: ($\sim 10^4$~cm$^{-3}$) C$^{18}$O gas (Onishi et al. 2002; see 
176: \S~\ref{taurus} for actual numbers); Krumholz \& Tan (2006) found 
177: that a similar result holds for objects of a wide range of density. 
178: Although the observed moderately supersonic motions (${\cal M}\sim 
179: 2$) can provide some support for the dense gas, we believe that 
180: the star formation is slow mainly because the bulk of the dense 
181: gas remains magnetically supported even after cross-field 
182: contraction has begun in localized regions in the condensed 
183: structures. 
184: This requirement can naturally be satisfied if such structures 
185: are marginally magnetically critical (Basu \& Ciolek 2000), 
186: where individual stars can form on a relatively short time 
187: scale, but the overall rate of star formation remains well 
188: below the free-fall value, as we have demonstrated using 
189: simulations in 2D sheet-like geometry (Li \& Nakamura 2004 
190: and Nakamura \& Li 2005; LN04 and NL05 hereafter; see also 
191: Kudoh et al. 2007 and Kudoh \& Basu 2008). The Taurus cloud 
192: complex is discussed further in \S~\ref{taurus}, along 
193: with the Pipe Nebula (\S~\ref{pipe}), which may represent an 
194: earlier phase of star formation in a magnetically dominated 
195: cloud (Lada et al. 2008; F. Alves et al., in preparation).   
196: 
197: The rest of the paper is organized as follows. Section~\ref{setup} 
198: describes the model formulation, including governing equations, 
199: initial and boundary conditions, and numerical code. It is followed
200: by three sections on numerical results. Section~\ref{result} concentrates 
201: on a standard model with a particular combination of initial turbulence 
202: and outflow feedback. It illustrates the essential features of the 
203: magnetically regulated star formation in three dimensions, including 
204: strong stratifications in density and turbulent speed, broad 
205: probability distribution functions (PDFs) of volume and column 
206: densities and, most importantly, low rate of star formation. This 
207: model is contrasted, in \S~\ref{other}, with others that have 
208: different levels of initial turbulence and outflow feedback. We 
209: devote \S~\ref{core} to a detailed analysis of dense cores, including
210: their shapes, mass spectrum, velocity dispersions, flux-to-mass
211: ratios, angular momenta, and whether they are bound or not. In 
212: \S~\ref{discussion}, we outline a general scenario of magnetically
213: regulated star formation in quiescent condensations of relatively
214: diffuse, turbulent clouds, and make connection to observations 
215: of the Taurus molecular cloud complex and Pipe Nebula. The main 
216: results of the paper are summarized in \S~\ref{conclude}. Readers 
217: interested only in the general scenario and their connection to 
218: observations can skip to \S~\ref{discussion}.
219: 
220: \section{Model Formulation}
221: \label{setup}
222: 
223: \subsection{Governing Equations}
224: 
225: The basic equations that govern the evolution of isothermal, weakly 
226: ionized, magnetized molecular gas are (see, e.g., Shu 1991)
227: \begin{equation}
228: {\partial \rho_n \over \partial t} + \nabla \cdot (\rho_n
229:  \mbox{\boldmath{$V$}}_n) = 0 \ ,
230: \end{equation}
231: \begin{equation}
232: \rho_n {\partial \mbox{\boldmath{$V$}}_n \over \partial t} +
233: \rho_n (\mbox{\boldmath{$V$}}_n \cdot \nabla \mbox{\boldmath{$V$}}_n)
234: = -\rho_n \nabla \Psi - c_s^2 \nabla \rho_n + {1 \over 4 \pi}
235: (\nabla \times \mbox{\boldmath{$B$}}) \times \mbox{\boldmath{$B$}} \ , 
236: \end{equation}
237: \begin{equation}
238: {\partial \mbox{\boldmath{$B$}} \over \partial t}
239: =\nabla \times (\mbox{\boldmath{$V$}}_i \times \mbox{\boldmath{$B$}}) \ ,
240: \end{equation}
241: \begin{equation}
242: \nabla ^2 \Psi = 4 \pi G \rho_n  \ , 
243: \end{equation}
244: where 
245: $\rho$, $\mbox{\boldmath{$V$}}$, $\mbox{\boldmath{$B$}}$, $c_s$,
246: and $\Psi$ are, respectively, 
247: density, velocity, magnetic field, isothermal sound speed, 
248: and gravitational potential, with the subscripts $n$ and $i$ 
249: denoting neutrals and ions. The drift velocity between ion 
250: and neutral is 
251: \begin{equation}
252: \mbox{\boldmath{$V$}}_i - \mbox{\boldmath{$V$}}_n = {t_c \over 4 \pi
253:  \rho_n} (\nabla \times \mbox{\boldmath{$B$}}) \times 
254: \mbox{\boldmath{$B$}}  \ , 
255: \end{equation}
256: where the coupling time between the magnetic field and neutral matter,
257: $t_c$, is given by 
258: \begin{equation}
259: t_c = {1.4 \over \gamma C \rho_n^{1/2}},
260: \end{equation}
261: in the simplest case where the coupling is provided by ions
262: that are well tied to the field lines and the ion density $\rho_i$
263: is related to the neutral density by the canonical expression
264: $\rho_i = C \rho_n^{1/2}$. We adopt an ion-neutral drag coefficient 
265: $\gamma = 3.5\times 10^{13}$~cm$^3$~g$^{-1}$~s$^{-1}$ and $C = 3 
266: \times 10^{-16}$~cm$^{-3/2}$~g$^{1/2}$ (Shu 1991). The factor 1.4 
267: in the above equation comes from the fact that the cross section 
268: for ion-helium collision is small compared to that of ion-hydrogen 
269: collision (Mouschovias \& Morton 1991). These governing equations 
270: are solved numerically subject to a set of initial and boundary 
271: conditions. 
272: 
273: \subsection{Initial and Boundary Conditions}
274: 
275: Our simulations are carried out in a cubic box (of size $L$), with
276: standard periodic conditions imposed at the boundaries. The cloud 
277: is assumed to have a uniform density $\rho_0$ initially. The 
278: corresponding Jeans length is 
279: \begin{equation} 
280: L_J= \left({\pi c_s^2/G \rho_0}\right)^{1/2}, 
281: \end{equation}
282: where the isothermal sound speed $c_s=1.88\times 10^4(T/10 
283: K)^{1/2}$~cm/s, with $T$ being the cloud temperature. The 
284: initial density $\rho_0 = 4.68\times 10^{-24} n_{H_2,0}$~g~cm$^{-3}$, 
285: where $n_{H_2,0}$ is the initial number density of molecular 
286: hydrogen, assuming 1 He for every 10 H atoms. 
287: 
288: We consider relatively diffuse molecular clouds, with a fiducial 
289: H$_2$ number density of $250$~cm$^{-3}$. It is the value adopted 
290: by Heyer et al. (2008) for the sub-region in Taurus that shows 
291: magnetically aligned $^{12}$CO striations. Scaling the density by 
292: this value, we have 
293: \begin{equation}
294: L_J=1.22 \left(\frac{T}{10 K}\right)^{1/2} 
295: \left(\frac{n_{H_2,0}}{250\ {\rm cm}^{-3}}\right)^{-1/2} {\rm pc}, 
296: \end{equation}
297: and a Jeans mass 
298: \begin{equation} 
299: M_J= \rho_0 L_J^3 = 31.6 M_\odot \left(\frac{T}{10 K}\right)^{3/2} 
300: \left(\frac{n_{H_2,0}}{250\ {\rm cm}^{-3}}\right)^{-1/2}.
301: \end{equation} 
302: The Jeans length is smaller than the dimensions of molecular cloud 
303: complexes such as the Taurus clouds, which are typically tens of 
304: parsecs. Although global simulations on the complex scale are
305: desirable, they are prohibitively expensive if individual star 
306: formation events are to be resolved at the same time. After some 
307: experimentation, we settled on a moderate size for the simulation 
308: box, with an initial Jeans number $n_J=L/L_J=2$. It represents 
309: a relatively small piece of a larger complex. Scaling the Jeans 
310: number by 2, we have a box size 
311: \begin{equation}
312: L=2.44 \left(\frac{n_J}{2}\right) \left(\frac{T}{10 K}\right)^{1/2} 
313: \left(\frac{n_{H_2,0}}{250\ {\rm cm}^{-3}}\right)^{-1/2} {\rm pc}, 
314: \end{equation}
315: and a total mass 
316: \begin{equation} 
317: M_{tot}= 253\ M_\odot \left(\frac{n_J}{2}\right)^3 \left(\frac{T}{10 
318: K}\right)^{3/2}\left(\frac{n_{H_2,0}}{250\ {\rm cm}^{-3}}\right)^{-1/2}
319: \end{equation} 
320: in the computation domain. Even though the initial Jeans number
321: adopted is relatively small, there is enough material inside 
322: the box to form many low-mass stars, especially in condensations 
323: where the Jeans mass is much reduced. 
324: 
325: We impose a uniform magnetic field along the $x$ axis at the beginning 
326: of simulation. The field strength $B_0$ is specified by the parameter 
327: $\alpha$, the ratio of magnetic to thermal pressure, through 
328: \begin{equation}
329: B_0=3.22~\alpha^{1/2} \left({T\over 10 K}\right)^{1/2}
330: \left(\frac{n_{H_2,0}}{250\ {\rm cm}^{-3}}\right)^{1/2} (\mu~G). 
331: \label{fieldstrength}
332: \end{equation}
333: In units of the critical value $2\pi G^{1/2}$ (Nakano \& Nakamura
334: 1978), the flux-to-mass ratio for the material inside the box is 
335: \begin{equation}
336: \Gamma_0 = 0.45 n_J^{-1} \alpha^{1/2}.
337: \label{centralgamma}
338: \end{equation}
339: Since we are interested in magnetic regulation of star formation,
340: we choose a relatively strong magnetic field with $\alpha=24$, 
341: corresponding to a mildly magnetically subcritical region with 
342: $\Gamma_0=1.1$ (or a dimensionless mass-to-flux ratio $\lambda
343: =0.91$) for $n_J=2$, which is close to the value of 1.2
344: used in the 2D simulations of LN04 and NL05. For the fiducial 
345: values for temperature and density, the initial field strength 
346: is $B_0=15.8\ \mu$G according to equation~(\ref{fieldstrength}); 
347: this value is not unreasonably high (see \S~\ref{taurus} for 
348: a discussion of the magnetic field in the Taurus region). We 
349: postpone an investigation of initially supercritical clouds to 
350: a future publication.  
351: 
352: Molecular clouds are highly turbulent, particularly for the 
353: relatively diffuse gas that we have assumed for the initial 
354: state. It is established that supersonic turbulence decays 
355: rapidly, with or without a strong magnetic field (e.g., Mac 
356: Low \& Klessen 2004). How the turbulence is maintained is 
357: uncertain (e.g., McKee \& Ostriker 2007). For the relatively 
358: small, parsec-scale region that is modeled here, a potential 
359: source is energy cascade from larger scales, which can 
360: in principle keep the turbulence in the sub-region for a 
361: time longer than the local crossing time. Ideally, this 
362: (and may other) form of driving should be included in a 
363: model, although it is unclear how this can be done in  
364: practice. We will take the limit that the turbulence decays 
365: freely, except for the feedback from protostellar outflows 
366: that are associated with forming stars. Other forms of 
367: driving will be considered elsewhere. Following the standard 
368: practice (e.g., Ostriker et al. 2001), at the beginning of 
369: the simulation ($t=0$) we stir the cloud with a turbulent 
370: velocity field of power spectrum $v_k^2 \propto k^{-3}$ in 
371: Fourier space.  
372: 
373: A useful time scale for cloud evolution is the global gravitational 
374: collapse time 
375: (Ostriker et al. 2001)
376: \begin{equation}
377: t_g= {L_J \over c_s} = 6.36 \times 10^6 \left(\frac{n_{H_2,0}} 
378: {250\ {\rm cm}^{-3}}\right)^{-1/2} ({\rm years}), 
379: \label{gravtime}
380: \end{equation}
381: which is longer than the free-fall time at the initial density, 
382: $t_{ff,0}=[3\pi/(32 G \rho_0)]^{1/2}$, by a factor of 3.27. In 
383: the presence of a strong magnetic field, as is the case in our
384: simulation, gravitational condensation is expected to occur 
385: preferentially along field lines, creating a flattened, sheet-like,
386: structure in which most star formation activities take place. 
387: Inside the sheet, the characteristic Jeans length is 
388: \begin{equation}
389: L_{s} \equiv {c_s^2 \over G \Sigma_0} = {1 \over \pi n_J} L_J 
390: =  0.195 \left(\frac{2}{n_J}\right) \left(\frac{T}{10 K}\right)^{1/2} 
391: \left(\frac{n_{H_2,0}}{250\ {\rm cm}^{-3}}\right)^{-1/2} {\rm pc}\  ,
392: \end{equation}
393: corresponding to a local gravitational collapse time 
394: \begin{equation}
395: t_{s} \equiv {L_{s} \over c_s} = {c_s \over G \Sigma_0} 
396: ={1 \over \pi n_J} t_g \ = 1.01 \times 10^6 \left(\frac{2}{n_J}\right) 
397: \left(\frac{n_{H_2,0}} {250\ {\rm cm}^{-3}}\right)^{-1/2} ({\rm years}) ,
398: \end{equation}
399: where the column density $\Sigma_0$ is measured along the initial
400: magnetic field lines, $\Sigma_0 = \rho_0 L = n_J \rho_0 L_J$.
401: The sheet has a Jeans number $n_{s}\equiv L/L_{s} = \pi n_J^2$, 
402: corresponding to $12.6$ for $n_J=2$, which is close to the value 
403: of 10 adopted in the 2D simulations of LN04 and NL05. Note that 
404: the gravitational collapse time $t_s$ is longer than the free-fall 
405: time of the condensed sheet 
406: \begin{equation}
407: t_{\rm ff,s}=\left({3\pi \over 32 G \rho_s}\right)^{1/2},
408: \label{freefalltime}
409: \end{equation}
410: by a factor of 2.31, where the characteristic sheet density is
411: given by  
412: \begin{equation}
413: \rho_s = \frac{\pi G \Sigma_0^2}{2 c_s^2} = \frac{\pi^2 n_J^2}{2} 
414: \rho_0 \ .
415: \end{equation}
416: 
417: \subsection{Numerical Code}
418: 
419: We solve the equations that govern the cloud evolution using a 3D MHD 
420: code based on an upwind TVD scheme, with the ${\nabla\cdot {\bf B}}=0$ 
421: condition enforced through divergence cleaning after each time step. 
422: The base code is the same as the one  
423: used in the cluster formation simulations of Li \& Nakamura (2006) 
424: and Nakamura \& Li (2007), except that we have now included ambipolar 
425: diffusion in the induction equation. The ambipolar diffusion term is 
426: treated explicitly, which puts a stringent requirement on the time 
427: step (since it is proportional to the square of the grid size, Mac 
428: Low et al. 1995), especially in the lowest density regions. To 
429: alleviate the problem, we set a density threshold, $\rho_{AD}=0.1 
430: \rho_0$, below which the rate of ambipolar diffusion is reduced to 
431: zero. A similar approach was taken by Kudoh et at. (2007), and can 
432: be justified to some extent by increased ionization of low (column) 
433: density gas from UV background (McKee 1989). Even with this measure, 
434: it still takes more than a month to run a standard $128^3$ simulation 
435: for two global gravitational collapse time $t_g$ (or $4\pi$ times the 
436: sheet collapse time $t_s$) on the vector machine available to us. 
437: Since our focus is on the evolution of, and star formation in, the 
438: condensed gas sheet, to speed up the simulations, we follow the 
439: initial phase of cloud evolution on a coarser grid of $64^3$ up to 
440: a maximum density of $\sim 40~\rho_0$, before switching to the 
441: $128^3$ grid. To gauge the numerical diffusion of magnetic field on 
442: such a relatively coarse grid, we have followed the evolution of a 
443: model cloud in the ideal MHD limit (with the ambipolar diffusion 
444: term turned off; Model I0 in Table \ref{tab:model parameter}) up 
445: to $2 t_g (=12.6t_s)$, and found that the mass-to-flux ratio is 
446: conserved to within $1.5\%$, 
447: which seems to be much better than the two-dimensional 
448: higher resolution runs with ZEUS ($512^3$, Krasnopolsky \& Gammie 2005). 
449: 
450: A difficulty with grid-based numerical codes such as ours in 
451: simulating star formation in turbulent clouds is that, once 
452: the collapse of a piece of the cloud runs away to form the 
453: first star, the calculation grinds to a halt without special 
454: treatment. Because the bulk of the stellar mass is believed 
455: to be assembled over a rather brief (Class 0) phase of $\sim
456: 3\times 10^4$~yrs (Andre et al. 1993) and the outflows are 
457: most active during this phase (much shorter than the fiducial 
458: characteristic time $t_s\sim 10^6$~yrs), we idealize the 
459: processes of individual star formation and outflow ejection 
460: as instantaneous, and treat them using the following simple 
461: prescriptions. When the density in a cell reaches a threshold 
462: value $\rho_{th}=10^3\rho_0$, we define around it a critical 
463: core using a version of CLUMPFIND algorithm 
464: (Williams, de Geus, \& Blitz 1994; 
465: see \S~\ref{core} for detail). The core identification 
466: routine is specified by a minimum threshold density $\rho_{\rm 
467: th, min}$ and the number of density levels $N$. They are set 
468: to $200~\rho_0$ and 5 respectively. The mean density of the 
469: core so identified is estimated at $\sim 10^5$~cm$^{-3}$, 
470: comparable to those of H$^{13}$CO$^+$ cores in Taurus studied 
471: by Onishi et al. (2002). We extract $30\%$ of the mass from 
472: the core (if its mass exceeds the local Jeans mass), and put 
473: it in a Lagrangian particle located at the cell center; its 
474: subsequent evolution is followed using the standard leap-frog 
475: method. The extracted mass fraction is 
476: motivated by the observations of J. Alves et al. (2007), 
477: who found a mass spectrum for dense cores that is similar to 
478: the stellar IMF in shape but more massive by a factor of $\sim 3$ 
479: (see also Ikeda, Sunada, \& Kitamura 2007). 
480: It is within the range of $25-75\%$ for the star formation
481: efficiency per dense core estimated theoretically by Matzner \& 
482: Mckee (2000). 
483: 
484: The material remaining in the core after mass extraction is 
485: assumed to be blown away in a two-component outflow: a bipolar 
486: jet of $30^\circ$ half-opening angle around the local magnetic 
487: field 
488: direction that carries a fraction $\eta$ of the total momentum, 
489: and a spherical component outside the jet that carries the 
490: remaining fraction. The total outflow momentum is assumed to 
491: be proportional to the stellar mass, with a proportionality 
492: coefficient of $P_*=35$~km/s. 
493: The prescriptions for star formation 
494: and outflow are similar to those used in Nakamura \& Li (2007). 
495: One difference is the use of CLUMPFIND here to define the critical 
496: core. Another is that, when an outflow reenters the computation 
497: box because of periodic boundary condition, we reduce its strength 
498: by a factor of ten if its speed exceeds $50~c_s$, to prevent 
499: it from strongly perturbing or even disrupting the condensed 
500: sheet. 
501: 
502: \section{Standard Model} 
503: \label{result}
504: 
505: We have run four models that include ambipolar diffusion, with 
506: parameters listed in Table \ref{tab:model parameter}. They
507: differ in random realization of the initial velocity field, 
508: relative strength of the jet and spherical outflow components,
509: and level of initial turbulence. An ideal MHD run was also 
510: performed for comparison. In this section, we focus on the 
511: model with initial turbulent Mach number ${\cal M}=3$ and 
512: jet momentum fraction $\eta=75\%$. It serves as the standard 
513: against which other models, to be discussed in the next 
514: section, are compared. 
515: 
516: The initial phase of cloud evolution is similar to those already 
517: presented in Krasnopolsky \& Gammie (2005), who carried out 
518: high-resolution, ideal MHD simulations of magnetically subcritical 
519: clouds in two dimensions, including self-gravity and an initially 
520: supersonic turbulence that decays freely, as in our simulation. 
521: The initial turbulence creates many shocklets that are oriented 
522: more or less perpendicular to the field lines; they appear as 
523: filaments of enhanced density and are the main sites of turbulence 
524: dissipation. As the turbulence decays, the fragments start to 
525: coalesce, eventually forming a single, approximate Spitzer sheet. 
526: In the ideal MHD limit, the sheet formation marks the end of 
527: cloud evolution. With ambipolar diffusion, it signals the beginning 
528: of a new phase---star formation\footnote{For a strong enough 
529: turbulence and/or initially inhomogeneous distribution of 
530: mass-to-flux ratio, stars can form before the formation of a 
531: well-defined sheet (see \S~\ref{discussion} for further
532: discussion).}. 
533: This latter phase is the focus of our discussion next.  
534: 
535: \subsection{Star Formation Efficiency}
536: \label{SFE}
537: 
538: We begin the discussion with a global measure of star formation in the
539: standard model: the efficiency of star formation, defined as the ratio 
540: of the mass of all stars
541: to the total mass of stars and gas. The star formation efficiency (SFE 
542: hereafter) is plotted in Fig.~\ref{fig:sfe} as a function of time in 
543: units of the gravitational collapse time of the sheet $t_s$. Sheet 
544: formation 
545: is finished by the end of the first global collapse time $t_g=2\pi\ 
546: t_s$. Soon after, the first star appears, around $t \approx 7~t_s$. 
547: By the end of simulation at $t=11.4~t_{s}$ (nearly $2~t_g$), 37 stars 
548: have formed, yielding an accumulative SFE of 6.5\%. The average mass 
549: of a star is close to $0.45 M_\odot \left(T/10 K\right)^{3/2}\left(n_{H_2,0}
550: /250 {\rm cm}^{-3} \right)^{-1/2}$, typical of low-mass stars. The 
551: rate of star formation is given by the slope of the SFE curve, which 
552: is estimated at $\sim 1.5\%$ per $t_s$. At this 
553: rate, it would take $\sim 65~t_s$ to deplete the gas due to star 
554: formation. The gas depletion time corresponds to $\sim 10~t_g$ or 
555: $\sim 33$ times the free-fall time $t_{{\rm ff},0}$ at the {\it 
556: initial} (uniform) density $\rho_0$. In other words, 
557: the star formation efficiency per {\it initial} free fall time 
558: (Krumholz \& McKee 2005) is SFE$_{{\rm ff},0}\sim 3\%$. Interestingly,
559: this rate is comparable to that of cluster formation in 
560: protostellar outflow-driven turbulence (Nakamura \& Li 2007). 
561: However, a large fraction of the gas is at a much higher density
562: than the initial density $\rho_0$ after the formation of the 
563: condensed sheet. It is more appropriate to measure the star 
564: formation rate using the free fall time at the characteristic 
565: sheet density $\rho_s=2 \pi^2\ \rho_0$, which is $t_{{\rm ff},s}= 
566: 0.43\ t_s$ (see equation~[\ref{freefalltime}]), yielding 
567: SFE$_{{\rm ff},s}\sim 0.7\%$. To help understand the rather slow 
568: and inefficient star formation, we next examine in some detail 
569: the cloud structure and dynamics at $t=11~t_s$, representative of 
570: the post-sheet formation phase of cloud evolution when stars 
571: are forming steadily, starting with overall mass distribution 
572: as revealed by column density. 
573: 
574: \subsection{Column Density Distribution}
575: 
576: The cloud material is expected to condense along the magnetic field
577: lines into a flattened structure. This expectation is borne out by
578: Figs.~\ref{colden_y} and \ref{colden_critical}, which show the 
579: distributions of column density along, respectively, the $y$- and 
580: $x$-axis (the direction of the initial magnetic field). Viewed
581: edge-on, the structure resembles a knotty, wiggly filament 
582: (Fig.~\ref{colden_y}). The dense filament is sandwiched by a 
583: low-density ``halo,'' which contains a number of spurs more or
584: less perpendicular to the filament. The spurs are roughly aligned 
585: with the global magnetic field. They resemble the diffuse, fuzzy 
586: $^{12}$CO streaks that are frequently observed to stream away 
587: from dense filaments (Goldsmith et al. 2008). Similar spurs are 
588: seen in the recent SPH-MHD simulations of Price \& Bate (2008). 
589: 
590: The face-on view of the condensed structure is quite different 
591: (Fig.~\ref{colden_critical}). It appears highly fragmented, with 
592: low-density voids as well as high-density peaks. Some 
593: over-dense regions cluster together to form a short filament; 
594: others are more scattered. The actual appearance of the cloud 
595: to outside observers will depend, of course, on the viewing 
596: angle. It can be filament-like (as in Fig.~\ref{colden_y}) or, 
597: more likely, patchy (as in Fig.~\ref{colden_critical}). Superposed 
598: on the face-on column density map is the distribution of the 
599: 33 stars that have formed up to the time plotted. With few 
600: exceptions, they follow the dense gas closely. Most of the stars
601: are found in small groups, which is broadly consistent with the 
602: distribution of young stars in Taurus (Gomez et al. 1993). Since 
603: the clouds is initially 
604: magnetically subcritical everywhere, in order for any localized 
605: region to collapse and form stars, it must become supercritical 
606: first. 
607: %
608: % plots recentered, and in units of Lg rather than Ls
609: %
610: 
611: A rough measure of magnetic criticality is provided by the 
612: average mass-to-flux ratio $\bar{\lambda}$ normalized to
613: the critical value,
614: \begin{equation}
615: \bar{\lambda}(y,z) = \int_{-L/2}^{+L/2} 2\pi G^{1/2}
616: \rho/B_x dx.  
617: \label{AverageLambda}
618: \end{equation}
619: Contours of $\bar{\lambda}=1$ are plotted in Fig.~\ref{colden_critical}, 
620: showing that most of the dense regions are indeed magnetically 
621: supercritical, and thus capable of forming stars in principle. 
622: The supercritical regions enclosed within the contours contain 
623: a significant fraction ($34\%$) of the total mass. 
624: The mass fraction of supercritical gas increases gradually with 
625: time initially, and remains more or less constant at later times 
626: ($t\gtrsim 6~t_s$), as shown in Fig. \ref{fig:supercritical}.
627: This value provides an upper limit to the efficiency of 
628: star formation.
629: If the supercritical regions collapse to form stars within a local 
630: free fall time, the star formation rate would be much higher 
631: than the actual value. An 
632: implication is that the bulk of the supercritical material 
633: remains magnetically supported to a large extent and is rather 
634: long lived. Only its densest parts (i.e., dense cores), 
635: which contain a small fraction of the supercritical mass, 
636: directly participate in star formation at any given time. 
637: The longevity is characteristic of the (mildly) supercritical 
638: regions that are produced through turbulence-accelerated 
639: ambipolar diffusion in nearly magnetically critical clouds. 
640: It is in contrast with the highly supercritical material 
641: in weakly magnetized clouds, which is expected to collapse 
642: promptly unless supported by turbulence. This same behavior 
643: was found in 2D calculations (NL05). 
644: 
645: 
646: To quantify the mass distribution further, we plot in Fig.~\ref{colden_PDF} 
647: the probability distribution function (PDF) of the column density 
648: along the $x-$axis, in the direction of initial magnetic field. The 
649: PDF peaks around $\Sigma \sim 1.7~\rho_0 L_J$ (where $\rho_0$ and 
650: $L_J$ are the density and Jeans length of the initial, uniform state), 
651: close to the average value of $2~\rho_0 L_J$. The distribution can be 
652: described reasonably well by a lognormal distribution, although there 
653: is significant deviation at both low and high density end. The 
654: considerable spread in column density may be surprising at the 
655: first sight, since one may expect little structure to develop in 
656: a subcritical sheet dominated by a strong, more or less uniform 
657: magnetic field, especially since the initial mass distribution is 
658: also chosen to be uniform. This expectation is consistent with the 
659: column density PDF of the ideal MHD counterpart of the standard 
660: model (plotted in the same figure for comparison), which is much 
661: narrower. The key difference is that ambipolar diffusion turns an 
662: increasingly large fraction of the initially subcritical cloud into 
663: supercritical material (until saturation sets in at late times, 
664: see Fig.~\ref{fig:supercritical}). Once produced, the supercritical 
665: material behaves drastically differently from the subcritical 
666: background. It is easier to compress by an external flow because 
667: its magnetic field is weaker. The compressed supercritical region
668: also rebounds less strongly, because the magnetic forces driving 
669: the rebound are cancelled to a larger extent by the self-gravity 
670: of the region. More importantly, the supercritical regions can 
671: further condense without external compression, via self-gravity. 
672: The accumulation of supercritical material and subsequent  
673: (magnetically-diluted) gravitational fragmentation in such material 
674: are responsible for the broad column density PDF. 
675: These key aspects of the problem are {\it not} captured by the 
676: ideal MHD simulation, but are well illustrated in 2D simulations 
677: that include ambipolar diffusion (e.g., NL05; Kudoh \& Basu 
678: 2006). Other aspects are better discussed in 3D, however. These 
679: include the dependence of various physical quantities on volume 
680: density, to which we turn our attention next. 
681: 
682: \subsection{Density Stratification} 
683: 
684: Our model cloud is strongly stratified in density. It contains dense 
685: cores embedded in a condensed sheet which, in turn, is surrounded 
686: by a more diffuse halo. The relative amount of mass distributed at 
687: different densities is displayed in Fig.~\ref{lognorm_density},
688: along with that for the ideal MHD case and for a Spitzer sheet. 
689: Whereas the mass distribution for the ideal MHD case resembles 
690: that of the Spitzer sheet closely, the same is not true for 
691: the standard model with ambipolar diffusion: its mass distribution 
692: is considerably broader, with more material at both the low and 
693: high density end. The former indicates that the diffuse material 
694: outside the sheet is dynamically active, as we show below. The 
695: excess at the high density end is produced, on the other hand, 
696: primarily by ambipolar diffusion, which enables pockets of 
697: the sheet material to condense across magnetic field lines 
698: gravitationally. The bulk ($62.5\%$) of the cloud material 
699: resides in the condensed sheet, which we define somewhat 
700: arbitrarily as the material in the density range between 
701: $5~\rho_0$ and $50~\rho_0$, corresponding to a range in 
702: H$_2$ number density between $1.25\times 10^3$~cm$^{-3}$ and 
703: $1.25\times 10^4$~cm$^{-3}$ (assuming $n_{H_2,0}=250$~cm$^{-3}$). 
704: The remaining cloud material is shared between the diffuse halo 
705: ($30.0\%$), defined 
706: loosely as the material with density below $5\ \rho_0$, and 
707: dense cores ($7.5\%$), defined as the material denser than
708: $50 \rho_0$ (see \S~\ref{core} below). 
709: 
710: Our model cloud is also strongly stratified in turbulent speed. 
711: Fig.~\ref{rms_vel} shows the mass weighted rms velocity as a 
712: function of density. At densities corresponding to the 
713: condensed sheet ($5-50~\rho_0$), the turbulent Mach number is 
714: about 2. The moderately supersonic turbulence persists in the 
715: condensed sheet for many local free fall times, despite the 
716: absence of external driving. It is driven by a combination 
717: of outflows (through either direct impact or MHD waves) and 
718: ambipolar diffusion-induced gravitational acceleration. 
719: The modest increase in rms velocity toward the high density 
720: end is because the denser gas is typically closer to star 
721: forming sites and is thus more strongly perturbed by outflows. 
722: The much higher level of turbulence in the diffuse halo is 
723: more intriguing. One may expect little turbulence to remain at 
724: the representative time $t=11~t_s$ 
725: shown in the figure, which is well beyond the time ($t\sim 6\ t_s$) 
726: of sheet formation --- a product of turbulence dissipation. The 
727: expectation is consistent with the ideal 
728: MHD simulation, which shows that, in the absence of external
729: driving, the turbulent motions die down quickly everywhere. 
730: The turbulence decay is changed by ambipolar diffusion 
731: which, accelerated by the initial strong turbulence, enables 
732: stars to form soon after sheet formation. The stars stir 
733: up the halo via protostellar outflows in at least two ways. 
734: First, the outflows are powerful enough to disrupt the 
735: star-forming cores, break out from the condensed sheet, and 
736: inject energy and momentum directly into the halo. Second, a 
737: fraction of the outflow energy and momentum is also 
738: deposited in the sheet, where gas motions can shake the 
739: field lines and send into the halo MHD waves whose amplitudes 
740: grow as the density decreases (Kudoh \& Basu 2003); the waves 
741: help maintain the turbulence in the diffuse region, particularly 
742: in the cross-field direction. 
743: 
744: 
745: Even though the turbulence speed is relatively high in the halo, 
746: it is still sub-Alfv\'enic. This is shown in 
747: Fig.~\ref{ave_MA}, where the average Alfv\'en Mach 
748: number $M_A$ is plotted as a function of density. The flow 
749: is moderately sub-Alfv\'enic at densities between $\sim 0.3$ 
750: and $\sim 10~\rho_0$, with $M_A$ within $20\%$ of $0.5$. The
751: Alfv\'en Mach number 
752: drops slowly towards lower densities, indicating that the 
753: lower-density gas becomes increasingly more magnetically
754: dominated in our model cloud. A caveat is that, in real 
755: clouds such as the Taurus complex, the dynamics of the 
756: low-density halo on the parsec-scale of our modeled region 
757: are expected to be tied to that of the larger complex. It 
758: is likely for the diffuse gas to receive energy cascaded 
759: from larger scales and become more turbulent than suggested 
760: by Fig.~\ref{ave_MA}. Some of this energy may even be fed 
761: through Alfv\'en waves into the condensed sheet and increase 
762: its velocity dispersion as well. The increase is not expected 
763: to be large, however, since the Alfv\'en speed becomes 
764: transonic at high densities (see Fig.~\ref{ave_MA}), where 
765: highly supersonic motions should damp out quickly through 
766: shock formation. In future refinements, it will be 
767: desirable to include external driving of the halo material, 
768: which may modify the way stars form, especially at early 
769: times (see discussion in \S~\ref{discussion}). Nevertheless, 
770: we believe that the decrease of turbulent speed with increasing 
771: density found in our simulation (and earlier in Kudoh \& 
772: Basu 2003) is a generic feature of the magnetically 
773: regulated star formation in three dimensions. It is broadly 
774: consistent with the observed trend (e.g., Myers 1995). 
775: 
776: Another generic feature is the distribution of magnetic field 
777: strength as a function of density. The distribution is shown 
778: in Fig.~\ref{ave_B}. There is a clear trend for the average 
779: field strength to remain more or less constant at relatively 
780: low densities, before increasing gradually towards the high 
781: density end. A similar trend is inferred 
782: in Zeeman measurements of field strength over a wide range of 
783: densities, and has been interpreted to mean that the initial 
784: stage of cloud condensation leading to star formation is 
785: primarily along the field lines, which increases the density 
786: without increasing the field strength (Heiles \& Crutcher 
787: 2005). This is precisely what happens in our model, 
788: where the initial accumulation of diffuse material into the 
789: condensed sheet is mostly along the field lines, and further 
790: condensation inside the sheet is primarily in the cross-field 
791: direction. The cross-field condensation leads to an increase 
792: in field strength with density, although by a factor that is 
793: less than that expected from flux-freezing because of ambipolar 
794: diffusion. 
795: %
796: 
797: \subsection{Sheet Fragmentation: Turbulent or Gravitational?} 
798: %
799: 
800: The star formation activities in our model are confined mostly to 
801: the condensed sheet. Its dynamics hold the key to understanding the 
802: magnetically regulated star formation. An overall impression of
803: the sheet can be gained from Figs.~\ref{colden_y} and 
804: \ref{colden_critical},
805: which showed an edge-on and face-on view of the sheet in column 
806: density. The sheet has a highly inhomogeneous 
807: mass distribution which is, of course, a pre-requisite for active
808: star formation. To illustrate the clumpy mass distribution further, 
809: we plot in Fig.~\ref{den_pot_vel} the distribution of the density  
810: in the ``mid-plane'' of the sheet, defined as the $yz$ plane 
811: (perpendicular to the initial magnetic field along the $x$ axis) 
812: that passes through the minimum of gravitational potential. The 
813: density distribution spans seven orders of magnitude, from 
814: $10^{-4}~\rho_0$ (the density floor adopted in the computation) to 
815: $10^3~\rho_0$ (the threshold for Lagrangian particle generation).
816: The question is: how is the density inhomogeneity created and
817: maintained? 
818: 
819: A widely discussed possibility is turbulent fragmentation, where 
820: regions of high densities are created by converging flows (e.g., 
821: Padoan \& Nordlund 2002; Mac Low \& Klessen 2004). There is indeed 
822: a random velocity field in the mid-plane, as shown by arrows in 
823: the figure. However, the flow speed is typically small, comparable 
824: to the sound speed in most places. 
825: It is unlikely for such a weak turbulence to create the high density 
826: contrast shown in Fig.~\ref{den_pot_vel} by itself. This conclusion 
827: is strengthened by the presence of a strong magnetic field, which 
828: cushions the flow compression. Indeed, the turbulent speeds in the 
829: plane of the sheet are significantly below the approximate
830: magnetosonic speed $v_{ms}=[B_x^2/(4\pi\rho)+c_s^2]^{1/2}$ in
831: most places, and are thus not expected to create large density 
832: fluctuations; they are essentially a collection of magnetosonic 
833: waves. The low level of turbulence points to gravity rather than 
834: flow motions as the primary driver of fragmentation in the sheet. 
835: %
836: % (1) potential loop holes... what if large compression in individual 
837: % regions involving a small fraction of mass??? i.e., tail effect???
838: % 
839: % Indeed the case at freshly shocked sites by active outflows
840: %
841:  
842: 
843: Gravity is expected to be dynamically important in the sheet, a structure 
844: formed by gravitational condensation along field lines in the 
845: first place. The fact that the sheet has a dimension much larger than
846: the thickness indicates that it contains many thermal Jeans masses. 
847: If the thermal pressure gradient was the main force that opposes 
848: the gravity, one would expect the bulk of the sheet to collapse 
849: in one gravitational time $t_s$. And yet, the vast majority of 
850: the sheet material remains uncollapsed after at least $\sim 5~t_s$. 
851: The longevity indicates that the 
852: gravitational fragmentation proceeds at a much reduced 
853: rate. The reduction is due, of course, to magnetic forces, 
854: which cancel out the gravitational forces in the plane of 
855: the sheet to a large extent (Shu \& Li 1997, Nakamura et al. 
856: 1999). The near cancellation 
857: can be inferred from Fig.~\ref{den_pot_vel}, where contours of
858: gravitational potential are plotted along with the density
859: and velocity vectors. If the gravitational forces were 
860: unopposed, one would expect the material to slide down the 
861: potential well, and picking up speed towards the bottom, 
862: with the infall speed reaching highly supersonic values. 
863: For example, for a potential drop of 8~$c_s^2$ (say between 
864: contours labeled  ``$-8$'' and ``$-16$''), an increase in speed 
865: by 4~$c_s$ is expected. These expectations are not met. The 
866: actual velocity field appears quite random, and generally 
867: transonic. The slow random motions in a rather deep 
868: gravitational potential well provide a clear indication 
869: for magnetic regulation of the sheet dynamics. Active 
870: outflows also play an important role in generating the density 
871: inhomogeneities observed in the condensed sheet. Their effects 
872: are discussed in \S~\ref{outflow} below. 
873: 
874: \subsection{Comparison to 2D Calculations}
875: \label{comparison}
876: 
877: %
878: There are a number of differences between the 3D simulation and the 
879: previous 2D calculations. The 2D treatment does not account for the 
880: mass stratification along the field lines, and thus cannot 
881: address the important issue of dynamical coupling between the 
882: condensed material and the diffuse halo. A related difference is
883: that, in 3D, there is an initial phase of mass settling to form the 
884: sheet, which is is absent in 2D. Furthermore, the mass of each star 
885: in the 2D models is held fixed (typically at $0.5~M_\odot$); in 3D, 
886: it is fixed at $30\%$ of the mass of the criticl core that forms 
887: the star. In addition, the prescription for outflow
888: \footnote{We note that the spherical component of the outflow 
889: in the standard 3D model carries a momentum close to  
890: (within $15\%$ of)  that of the 2D outflow in the plane of 
891: mass distribution in the standard model of NL05.}
892:  and the degree 
893: of initial cloud magnetization are also somewhat different. Despite 
894: these differences, the results of the 2D and 3D models are in 
895: remarkably good agreement. 
896: 
897: The most important qualitative agreement is that, once settled 
898: along the magnetic field lines, the condensed sheet material 
899: produces stars at a slow rate. In the standard 2D model of NL05, 
900: about $5.3\%$ of the total mass is converted into stars within 
901: a time interval of $3.35~t_s$ (see their Fig.~1), corresponding 
902: to a star formation rate of $\sim 1.6\%$ per sheet gravitational 
903: collapse time $t_s$. This rate is nearly identical to the value 
904: of $\sim 1.5\%$ estimated in \S~\ref{SFE} for the standard 3D 
905: model (which is also close to that of the 3D ${\cal M}=10$ model, 
906: see \S~\ref{other} below). The (face-on) mass distributions 
907: are also similar in 2D and 3D models. In both cases, dense, 
908: star-forming cores are embedded in moderately supercritical 
909: filaments which, in turn, are surrounded by more diffuse, 
910: magnetically subcritical material (compare 
911: Fig.~\ref{colden_critical} with the lower panels of 
912: Fig.~2 of NL05). More quantitatively, the mass fraction of 
913: supercritical material increases gradually initially and 
914: reaches a plateau of $\sim 30\%$ in both 2D and 3D models. The 
915: similarity extends to the turbulent speed of the condensed 
916: material, which is about twice the sound speed in both cases 
917: at late times (see Fig.~\ref{rms_vel} and Fig.~11 of NL05). 
918: 
919: The similarities indicate that 2D calculations capture the 
920: essence of the slow, magnetically regulated star formation 
921: in moderately subcritical clouds. The same conclusion was
922: drawn by Kudoh et al. (2007) based on 3D simulations of 
923: weakly perturbed magnetically subcritical sheets. The 
924: 2D calculations have the advantage of being much faster to 
925: perform. They can be done with higher spatial resolution 
926: ($512^2$ as compared to the $128^3$ used in this paper) 
927: and in larger numbers. The number of 3D simulations is 
928: necessarily more limited, especially because of the 
929: small time step required by our treatment of ambipolar 
930: diffusion using an explicit method. Nevertheless, we are 
931: able to perform several additional simulations, to check 
932: the robustness of the standard model, and to help elucidate 
933: the effects of two key ingredients: initial turbulence and 
934: outflow feedback.  
935: 
936: \section{Other Models} 
937: \label{other}
938: 
939: \subsection{Initial Turbulence}
940: \label{turbulence}
941: 
942: Even though the initial turbulence dissipates quickly, it seeds 
943: inhomogeneities in mass distribution and mass-to-flux ratio 
944: that are expected to affect star formation later on. To illustrate 
945: the effects of initial turbulence, we focus on a model identical 
946: to the standard model, except for a higher initial turbulent Mach 
947: number ${\cal M}=10$ (instead of $3$). The overall pattern of 
948: cloud evolution follows that seen 
949: in the standard model: dissipation of turbulence leads to 
950: gravitational settling of cloud material onto a condensed sheet, 
951: the densest parts of which collapse to form stars; the stars, 
952: in turn, stir up the sheet and its surrounding halo through 
953: protostellar outflows. One difference is that the stronger 
954: turbulence initiates star formation at an earlier time, as seen 
955: in Fig.~\ref{sfetwo}. The first star forms around $\sim 4.5~t_s$ 
956: in the ${\cal M}=10$ case, compared to $\sim 6.8~t_s$ for 
957: ${\cal M}=3$. Nevertheless, their rates of star formation are 
958: comparable, with values between $\sim 0.5\%$ and $\sim 1.0\%$ 
959: per free fall 
960: time $t_{ff,s}=0.43\ t_s$ in both cases. The earlier start, but 
961: still low rate, of star formation in the more turbulent case 
962: is consistent with the turbulence-accelerated, magnetically 
963: regulated scenario that we advocate based on previous 2D 
964: calculations (LN04, NL05; see also recent 3D simulations of 
965: Kudoh \& Basu 2008). 
966: 
967: The basic structure and kinematics of the initially more turbulent 
968: cloud are also similar to those of the standard model cloud. In 
969: both cases, the condensed sheet is highly fragmented but dynamically 
970: ``cold,'' with moderately supersonic random motions. It is surrounded 
971: by a halo moving at highly supersonic, but sub-Alfv\'enic, speeds. 
972: An interesting difference is that the sheet in the ${\cal M}=10$ 
973: case appears thicker, as illustrated in Fig.~\ref{colden_y_M10},
974: where an edge-on view of the column density distribution (along 
975: the $y$-axis, perpendicular to the initial magnetic field 
976: direction) is plotted for a representative time $t=8~t_s$. It is 
977: to be compared with Fig.~\ref{colden_y} for the standard model. 
978: The main reason for the larger apparent thickness is that the 
979: condensed sheet, while still thin intrinsically, is more warped. 
980: An example of warps is shown in Fig.~\ref{den_Bfield_M10}, where 
981: the density distribution in a plane perpendicular to the $y$-axis 
982: is plotted; superposition of such warps in different parts of 
983: the sheet along the line of sight gives rise to the more 
984: puffed-up appearance.
985: 
986: Associated with the sharp warp of condensed material in 
987: Fig.~\ref{den_Bfield_M10} is a strong 
988: distortion of magnetic field lines. Both are likely driven by 
989: protostellar outflows, which are the main source of energy 
990: that drives the cloud evolution at late times. In particular, 
991: outflows ejected at an angle to the large-scale magnetic field 
992: can excite large-amplitude Alfv\'en waves, which perturb the 
993: field lines in the halo that thread other parts of the condensed 
994: sheet. The large Alfv\'en speed in the diffuse halo allows 
995: different parts of the sheet to communicate with one another 
996: quickly, as emphasized recently by Shu et al. (2007) in the 
997: context of poloidally magnetized accretion disks. The 
998: communication raises the interesting possibility of global, 
999: magnetically mediated, oscillations for the condensed material 
1000: that may have observational implications. This issue warrants 
1001: a closer examination in the future.  
1002: 
1003: We conclude by stressing that condensations produced by gravitational
1004: settling along field lines do not have to appear highly elongated, 
1005: even when viewed edge-on. They can be further distorted by outflows 
1006: from stars formed in neighboring condensations or other external 
1007: perturbations that are not included in our current calculations. We 
1008: have experimented with simulations that allow active outflows to 
1009: re-enter the simulation box, and found that the condensed sheet 
1010: became more puffed up and, sometimes, even completely disrupted. 
1011: 
1012: \subsection{Outflow Feedback}
1013: \label{outflow}
1014: 
1015: Protostellar outflow plays an important role in our model. It 
1016: limits the amount of core mass that goes into a forming star 
1017: and injects energy and momentum back to the cloud. In our 
1018: standard model, we have assumed that $75\%$ of the outflow 
1019: momentum is carried away in a bipolar jet (of $30^\circ$ in
1020: half opening angle), and the 
1021: remaining $25\%$ in a spherical component. Here, we contrast 
1022: the standard model with one that has a stronger spherical 
1023: component (carrying $75\%$ of outflow momentum). The star
1024: formation efficiencies (SFEs) of both models are plotted 
1025: in Fig.~\ref{sfedependm3}. It is clear that the stronger 
1026: spherical component of outflow has reduced the rate (and
1027: efficiency) of star formation significantly, by a factor of 
1028: $\sim 2.3$. This trend is the opposite of what we found in 
1029: the case of cluster formation in dense clumps that are
1030: globally magnetically supercritical (Nakamura \& Li 2007):  
1031: for a given total outflow momentum, more spherical outflows 
1032: tend to drive turbulence on a smaller scale that dissipates 
1033: more quickly, leading to a higher rate of star formation. 
1034: 
1035: There are a couple of reasons for the seemingly contradictory 
1036: result. First, whereas the supercritical cluster-forming 
1037: clumps are supported mainly by turbulent motions, the bulk 
1038: of the star-forming condensations in subcritical or nearly 
1039: critical clouds are magnetically supported (in the cross-field 
1040: direction). As a result, faster dissipation of 
1041: turbulence leads to more rapid star formation in the former, 
1042: but not necessarily in the latter. Second, the jet component, 
1043: while crucial for supporting the more massive, rounder 
1044: cluster-forming clump, has a more limited effect on the flattened 
1045: condensation: it escapes easily from the sheet (and quickly 
1046: out of the simulation box) and its energy is mostly ``wasted'' 
1047: as far as driving turbulence in the star-forming, condensed 
1048: material is concerned. The spherical component, in contrast, 
1049: impacts the condensed material more directly, especially in 
1050: the vicinity of a newly formed star. The dispersal of the 
1051: residual core material after star 
1052: formation to a wider region appears to be the main reason 
1053: for the lower star formation rate in the case of stronger 
1054: spherical component: it takes longer for the more widely 
1055: dispersed material to recondense and form stars. The stars 
1056: in the stronger spherical component model are less clustered 
1057: than those in the standard model, consistent with the above 
1058: interpretation.  
1059: %
1060: 
1061: \section{Dense Cores}
1062: \label{core}
1063: 
1064: \subsection{Identification}
1065: 
1066: A variant of the CLUMPFIND algorithm of Williams et al. (1994) is
1067: ussed for core identification. We apply their ``friends-of-friends'' 
1068: algorithm to the density data cube in regions above a threshold 
1069: density, $\rho_{\rm th, min}$, and divide the density distribution 
1070: between $\rho_{\rm th, min}$ and $\rho_{\rm th, max}$ into 15 bins 
1071: equally spaced logarithmically. The maximum is set to $\rho_{\rm th, 
1072: max}=10^3\ \rho_0$, the density above which a star (Lagrangian 
1073: particle) is created if the core mass exceeds the local Jeans 
1074: mass. The minimum is set to $\rho_{\rm th, min}=50\ \rho_0$, 
1075: corresponding to $n_{H_2}=1.25 \times 10^4$~cm$^{-3}$ for the 
1076: fiducial value of initial density. We include only those 
1077: ``spatially-resolved'' cores containing more than 30 cells in 
1078: the analysis, to ensure that their properties are determined with 
1079: reasonable accuracy. The minimum core mass is thus $M_{c, \rm min}
1080: \equiv \rho_{\rm th, min} \times 30\Delta x\Delta y\Delta z=7.19 
1081: \times 10^{-2} M_{J,s}$, where $M_{J,s}\equiv \rho_s L_s^3 
1082: \simeq2.5M_\odot\left(T/10 K\right)^{3/2}\left(n_{H_2,0}/250 
1083: {\rm cm}^{-3} \right)^{-1/2}$ is the Jeans mass of the Spitzer 
1084: sheet. 
1085: 
1086: In our standard simulation, the cloud material settles into a large 
1087: scale gas sheet by the time $t\sim 6\ t_s$. Soon after, a dense core
1088: collapses in a run-away fashion to form the first star, signaling 
1089: the beginning of active star formation. We will limit our analysis 
1090: of dense cores to this star-forming phase, which complements the 
1091: analyses of, e.g., Gammie et al. (2003) and Tilley \& Pudritz (2007), 
1092: which focus on the phase prior to the runaway collapse. 
1093: 
1094: There are typically a handful of dense cores in our standard model at 
1095: any given time (see Fig.~\ref{colden_critical}), which is too small 
1096: for a statistical analysis. To enlarge the core sample, we rerun the 
1097: standard model with a different realization of the initial turbulent 
1098: velocity field, and include all ``resolved'' cores at six selected 
1099: times $t=6.0$, 7.0, 8.0, 9.0, 10.0, and 11.0~$t_s$ from both runs. 
1100: The total number of cores is 72. Since the majority of them do not 
1101: survive for more than 1~$t_s$ (they do last for several local
1102: free-fall times typically), there is relatively little overlap in 
1103: dense cores from one time to another. We treat all cores independently 
1104: regardless whether they appear at more than one selected time or 
1105: not. They have densities in the range 10$^4\sim 10^5$ cm$^{-3}$ 
1106: and radii of $0.04 \sim 0.12$ pc, corresponding roughly to 
1107: H$^{13}$CO$^+$ or N$_2$H$^+$ cores.
1108: 
1109: \subsection{Core Shape}
1110: 
1111: We compute the shape of a core using the eigenvalues of the 
1112: moment-of-inertia tensor:
1113: \begin{equation} 
1114: I_{ij} \equiv \int \rho x_i x_j dV  \ ,
1115: \end{equation}
1116: where the components of position vector, $x_i$ and $x_j$, 
1117: are measured relative to the center of mass.
1118: We compute the lengths of the three principal axes, $a, b,$ and $c$
1119: ($a\ge b \ge c$) by dividing the eigenvalues of the tensor, 
1120: $M_ca^2$, $M_cb^2$, and $M_c c^2$, by the core mass $M_c$.
1121: The core shape is then characterized by the axis ratios 
1122: $\xi\equiv b/a$ and $\eta \equiv c/a$.
1123: 
1124: Figure \ref{fig:coreshapenew} shows the distribution of axis ratios 
1125: of the dense cores in our sample. It is clear that the cores
1126: are generally triaxial; they show little clustering around either the 
1127: prolate ($b/a=c/a<1$) or oblate ($b/a=1, c<1$) line in the figure. 
1128: This result is perhaps to be expected, since the cores are distorted 
1129: by external turbulent flows and occasionally outflows, neither of 
1130: which are symmetric. 
1131: Following Gammie et al. (2003), we divide the $\xi$-$\eta$ plane 
1132: into three parts: a prolate group which lies above the line
1133: connecting $(\xi, \eta)=(1.0, 1.0)$ and (0.33, 0.0), an oblate group
1134: which lies below the line connecting $(\xi, \eta)=(1.0,1.0)$ and 
1135: (0.67, 0.0), and a triaxial group which is everything else.
1136: Of the 72 identified cores, 43\% (31) are prolate, 44\% (32) are
1137: triaxial, and 13\% (9) are oblate.  These fractions are in general 
1138: agreement with Gammie et al. (2003). One  
1139: difference is that the ratio $\eta=c/a$ tends to be smaller for 
1140: our cores [see Figure 10 of Gammie et al. (2003)]. It is likely 
1141: because dense cores in our model are embedded in a condensed sheet, 
1142: which has already settled 
1143: gravitationally along the field lines, although part of the 
1144: difference may also come from the different methods used in
1145: core identification. Note that more massive cores tend to be more 
1146: flattened along their shortest axis. This trend lends support to 
1147: the idea that gravitational squeezing plays a role in core 
1148: flattening. 
1149: 
1150: Another trend is that more massive cores tend to be more oblate. 
1151: For example, 9 out of 15 (60\%) cores having masses more than 1.5 
1152: times the average core mass 
1153: ($\left<M_c\right>=0.93~M_{J,s}$) 
1154: lie 
1155: below the line connecting $(\xi, \eta)=(1.0,1.0)$ and (0.5, 0.0)
1156: (the dotted line in Fig.~\ref{fig:coreshapenew}). Less massive cores
1157: tend to be more prolate. Their shapes may be more susceptible to
1158: external influences, such as ambient flows. 
1159: 
1160: \subsection{Virial Analysis}
1161: 
1162: 
1163: Virial theorem is useful for analyzing the dynamical states of dense 
1164: cores. We follow Tilley \& Pudritz (2007) and adopt the virial 
1165: equation in Eulerian coordinates (McKee \& Zweibel 1992): 
1166: \begin{equation}
1167: {1 \over 2} \ddot{I} + {1 \over 2} \int _S \rho r^2 
1168: \mbox{\boldmath $v$} \cdot d\mbox{\boldmath $S$} 
1169: = U+K+W+S+M+F \ ,
1170: % = {1 \over 2} \ddot{I'}
1171: \label{eq:virial}
1172: \end{equation}
1173: where
1174: $$
1175: %\begin{eqnarray}
1176: I=\int _V \rho r^2 dV\ , 
1177: U= 3 \int _V P dV \ ,
1178: K= \int _V \rho v^2 dV \ ,
1179: W= -\int _V \rho \mbox{\boldmath $r$} \cdot \nabla \Psi dV \ ,
1180: $$
1181: $$
1182: S= -\int _S \left[P\mbox{\boldmath $r$}+ 
1183: \mbox{\boldmath $r$}\cdot (\rho \mbox{\boldmath $v$} 
1184: \mbox{\boldmath $v$})\right] \cdot d\mbox{\boldmath $S$} \ ,
1185: M= {1 \over 4\pi}\int _V B^2 dV \ ,
1186: F= \int _S \mbox{\boldmath $r$} \cdot \mbox{\boldmath $T$}_M \cdot d
1187: \mbox{\boldmath $S$}  \ .
1188: $$
1189: %\end{eqnarray}
1190: Here, $\mbox{\boldmath $T$}_M$ is the Maxwell stress-energy tensor
1191: \begin{equation}
1192: \mbox{\boldmath $T$}_M={1 \over 4\pi}\left(
1193: \mbox{\boldmath $B$}\mbox{\boldmath $B$}-{1 \over 2} B^2 
1194: \mbox{\boldmath $I$}\right) \  ,
1195: \end{equation}
1196: where $\mbox{\boldmath $I$}$ is the unit tensor.
1197: The terms, $I$, $U$, $K$, $W$, $S$, $M$, and $F$ denote, respectively,  
1198: the moment of inertia, internal thermal energy, internal kinetic
1199: energy, gravitational energy, the sum of the thermal surface
1200: pressure and dynamical surface pressure, internal magnetic energy,
1201: and the magnetic surface pressure.
1202: The internal thermal, kinetic, and magnetic energies are always
1203: positive. The gravitational term is negative for all the 
1204: cores in our sample, although it can in principle be positive, 
1205: especially in crowded environments. Other terms can be either 
1206: positive or negative.
1207: The second term on the left hand side of equation (\ref{eq:virial})
1208: denotes the time derivative of the flux of moment of inertia
1209: through the boundary of the core. 
1210: As is the standard practice (e.g., Tilley \& Pudritz 2007), we 
1211: ignore the left hand side of equation (\ref{eq:virial}) in our 
1212: discussion, and consider a core to be in virial equilibrium if 
1213: the sum $U+K+W+S+M+F=0$. 
1214: 
1215: The equilibrium line is shown in Fig.~\ref{fig:virialparamnew}, where 
1216: the sum of the surface terms ($S+F$) is plotted against 
1217: the gravitational term ($W$), 
1218: both normalized to the sum of the internal terms ($U+K+M$). 
1219: The surface terms (dominated by the external kinetic term) 
1220: are generally important; they are more so for lower mass cores 
1221: than higher mass cores. 
1222: The majority of the cores more massive than 1.5 times
1223: the average core mass lie below the equilibrium line, and 
1224: are thus bound; they are expected to collapse and form stars. 
1225: The majority of lower mass cores lie, in contrast, above the line, 
1226: and may disperse away, if they do not gain more mass through
1227: accretion and/or merging with other cores, or further reduce 
1228: their turbulence support through dissipation and/or magnetic
1229: support through ambipolar diffusion. 
1230: 
1231: Tilley \& Pudritz (2007) found in their ideal MHD simulations 
1232: that as the mean magnetic field becomes stronger, the surface 
1233: terms contribute more to the virial equation.
1234: Indeed, for their most strongly magnetized cloud, a good
1235: fraction of cores have surface terms that are greater than 
1236: the sum of the internal terms ($U+K+M$). This is not the 
1237: case for our non-ideal MHD model, even though our clouds are 
1238: more strongly magnetized than theirs. The reason is that the 
1239: cloud dynamics is modified by ambipolar diffusion, especially 
1240: in dense cores, where the diffusion rate is the highest. A 
1241: related difference is that there are more bound cores in our 
1242: models than in their strong field models, where most of the 
1243: cores are magnetically subcritical (see also Dib et al. 2007).   
1244: Ambipolar diffusion enables these cores to become supercritical, 
1245: and thus more strongly bound. It is a crucial ingredient that 
1246: cannot be ignored for core formation, even for relatively 
1247: weakly magnetized clouds (of dimensionless mass-to-flux ratio 
1248: $\lambda$ of a few). For such clouds, the cores tend to be more 
1249: strongly magnetized (relative to their masses) than the cloud 
1250: as a whole, especially in driven turbulence (Dib et al/ 2007), 
1251: since only a fraction of the cloud mass along any given flux 
1252: tube is compressed into a core. 
1253: 
1254: The boundedness of a core is often estimated using the virial 
1255: parameter, the ratio of the virial mass to core mass, 
1256: particularly in observational studies. 
1257: In Figure \ref{fig:virialrationew}, we plot against core mass 
1258: the virial parameter, defined as 
1259: \begin{equation}
1260: \alpha_{\rm vir}={5 \Delta V_{1D}^2 R_c \over GM_c},
1261: \end{equation}
1262: where $\Delta V_{1D}$ is the one-dimensional velocity dispersion 
1263: including the contribution from thermal motion.
1264: There is a clear trend for more massive cores to have smaller 
1265: virial parameters. Indeed, nearly all cores 
1266: more massive than 1.5 times the average 
1267: ($M_c \gtrsim 1.5 \left<M_c\right>\simeq 1.4 M_{J,s}$)
1268: have virial parameters smaller than unity. 
1269: It is consistent with the fact that most of these 
1270: cores lie below the line of the virial equilibrium in 
1271: Fig.~\ref{fig:virialparamnew},  indicating that they are bound. 
1272: The reason for the agreement is that the surface terms are not 
1273: as significant as the gravitational term for massive cores, and
1274: that their magnetic support is greatly reduced by ambipolar 
1275: diffusion (consistent with the trend for cores with smaller 
1276: flux-to-mass ratios to have smaller virial parameters, see
1277: Fig.~\ref{fig:virialparamnew}).
1278: The reduction of surface terms makes the internal kinetic energy, 
1279: which is directly related to the virial parameter, more important.
1280: We conclude that the virial parameter can be used to gauge the 
1281: importance of self-gravity for dense cores formed in strongly 
1282: magnetized clouds, as long as ambipolar diffusion is treated 
1283: properly. The virial parameter can be fitted by a power law 
1284: $\alpha_{\rm vir} \propto M_c^{-2/3}$, in good agreement 
1285: with the scaling found by Bertoldi \& McKee (1992) for nearby 
1286: GMCs.
1287: 
1288: \subsection{Velocity Dispersion-Radius Relation}
1289: 
1290: The kinematics of molecular clouds and dense cores embedded in them 
1291: are probed with molecular lines. Since the pioneering work of Larson 
1292: (1981), many observational studies have shown that the linewidth of 
1293: a region correlates with the size of the region. The linewidth-size 
1294: relation is often approximated by a power law $\Delta V\propto 
1295: R^{0.5}$. However, this relation may not hold universally, especially
1296: on small scales. There are many cases where no clear correlation is 
1297: found between the two (e.g., Caselli \& Myers 1995; Lada et al. 1991; 
1298: Onishi et al. 2002).
1299: 
1300: In Fig.~\ref{fig:velocitysizenew}, we plot the 
1301: three-dimensional
1302: nonthermal velocity 
1303: dispersion against radius (equivalent to the linewidth-size relation) 
1304: for the cores in our sample. No correlation 
1305: is apparent between the two, consistent with the observations of 
1306: Onishi et al. (2002) for the dense cores in the Taurus molecular 
1307: clouds. There is a large scattering in the velocity dispersion-radius 
1308: plot, especially for cores with small masses. The reason may be that 
1309: some cores (especially the less massive ones) are not in virial 
1310: equilibrium and that their formation is strongly influenced by 
1311: external flows. 
1312: 
1313: \subsection{Angular Momentum Distribution}
1314: 
1315: Angular momentum is one of the crucial parameters that determine the 
1316: characteristics of formed stars or stellar systems. In Fig.~\ref{am1}, 
1317: we plot the specific angular momenta of the cores against their radii.
1318: There is no tight correlation between the specific angular momentum 
1319: and radius, although there is a slight tendency for more massive 
1320: cores to have somewhat larger specific angular momenta. For our 
1321: sample cores, the specific angular momentum ranges from 0.01 to 
1322: 0.4 $c_s L_s$ (where $c_s$ is the isothermal sound speed and $L_s$ the
1323: Jeans length of the Spitzer sheet). The distribution peaks around 
1324: 0.08 $c_s L$, corresponding 
1325: to $0.9 \times 10^{21}$ cm$^2$ s$^{-1}$ (for the fiducial cloud 
1326: parameters). For comparison, we 
1327: superpose in the figure the specific angular momenta of 7 starless 
1328: N$_2$H$^+$ cores deduced by Caselli et al. (2002). They are broadly 
1329: consistent with our results. The average for the N$_2$H$^+$ cores 
1330: is estimated at $\sim 1.6 \times 10^{21}$ cm$^2$ s$^{-1}$, somewhat 
1331: larger than the peak value of our distribution. 
1332: 
1333: An oft-used quantity for characterizing the rotation of a dense core 
1334: is the ratio of rotational energy to gravitational energy. For a 
1335: uniform sphere of density $\rho$ rotating rigidly at an angular 
1336: speed $\Omega$, it is given by $\beta=\Omega^2/(4\pi G\rho)$. 
1337: Even though our cores are not uniform, and do not rotate strictly 
1338: as a solid body, we have computed the value of $\beta$ using the 
1339: average angular speed and mean density. The distribution of  
1340: $\beta$ peaks around $\sim $ 0.02, in good agreement with 
1341: the average value of 0.018 for the starless N$_2$H$^+$ cores of 
1342: Caselli et al. (2002). It is also consistent with the well-known 
1343: result that rotation is generally not important as far as core 
1344: support against self-gravity is concerned (Goodman et al. 1993).  
1345: 
1346: \subsection{Flux-to-Mass Ratio}
1347: 
1348: Ambipolar diffusion enables dense cores to be less magnetized relative
1349: to their masses. In Fig.~\ref{fig:mass2fluxnew}, we plot the magnetic 
1350: flux-to-mass ratios of the cores against their column densities.
1351: The ratio is normalized to the critical value $2\pi G^{1/2}$, and the 
1352: column density to that of the initially uniform state, 
1353: $4\pi\rho_0 L_s$. 
1354: The magnetic flux-to-mass ratio is 
1355: approximated as $\Gamma_c\equiv \pi R_{c}^2 \left< B\right>/M_c$, where
1356: $\left<B\right>$ is the mean magnetic field strength inside a core
1357: (e.g., Dib et al. 2007) and the mean core radius $R_c$ is 
1358: defined as the radius of a sphere enclosing the same volume as 
1359: the core. The column density is calculated from the core mass 
1360: divided by the area of a circle whose radius is equal to $R_{c}$. 
1361: There is a clear trend for cores with higher column densities to have 
1362: lower flux-to-mass ratios. Indeed, all cores with masses greater 
1363: than $1.5\left<M_c\right>$ are magnetically supercritical, with 
1364: flux-to-mass ratios as low as $\sim 0.35$, indicating 
1365: that they have all experienced significant ambipolar diffusion. 
1366: The weakening of magnetic support may have enabled these cores to 
1367: condense to a high column density in the first place. 
1368: 
1369: \subsection{Core Mass Spectrum}
1370: 
1371: In Figure \ref{fig:coremf} we plot the mass spectrum for our sample 
1372: of dense cores. The core mass is normalized to the Jeans mass of 
1373: the Spitzer sheet, $M_{J,s}$. 
1374: The lowest mass at $\sim 0.1~M_{J,s}$ corresponds to the minimum 
1375: core mass determined from the criteria for core identification. There 
1376: is a prominent break around $1~M_{J,s}$ in the mass spectrum. Above 
1377: the break, the spectrum can be fitted roughly by a power law $dN/dM 
1378: \propto M^{-2.5}$, not far from the Salpeter stellar initial mass 
1379: function. We note that cores above the break contain more than 
1380: $\sim 100$ cells; their structures should be reasonably well resolved. 
1381: The core spectrum flattens below the break, which is also
1382: broadly consistent with observations. 
1383: Similar core spectra are also 
1384: obtained in the turbulent fragmentation model of core formation that
1385: involves much weaker magnetic field 
1386: and much stronger turbulence
1387: (e.g., Padoan \& Nordlund 2002).
1388: The characteristic mass below which the core mass spectrum flattens 
1389: apparently coincides with the Jeans mass of the condensed sheet, 
1390: which may result from magnetically regulated gravitational fragmentation.
1391: Caution must be exercised in trying to differentiate various 
1392: scenarios of core formation, such as the magnetically regulated 
1393: gravitational fragmentation and the turbulent fragmentation, based 
1394: on core mass spectrum and, perhaps by extension, stellar mass 
1395: spectrum. The stellar mass spectrum for the 52 stars formed in the 
1396: standard model and its close variant are plotted in Fig.~\ref{fig:coremf}
1397: for comparison. The stellar mass spectrum also resembles the 
1398: Salpeter IMF. 
1399: 
1400: \section{Discussion}
1401: \label{discussion}
1402: 
1403: \subsection{Magnetically-Regulated Star Formation in Quiescent 
1404: Condensations} 
1405: \label{scenario}
1406: 
1407: We have simulated star formation in relatively diffuse, mildly 
1408: magnetically subcritical clouds that are strongly turbulent 
1409: initially. The decay of initial turbulence leads to gravitational 
1410: condensation of the diffuse material along field lines into 
1411: a highly fragmented sheet (see Fig.~\ref{3D}), which produces 
1412: stars in small groups. The bulk of the condensed material 
1413: remains magnetically supported in the cross-field direction, 
1414: however, with typical motions that are moderately supersonic 
1415: and approximately magnetosonic. Only a small fraction of the 
1416: condensed material collapses gravitationally to form stars. 
1417: The calculations lead us to a picture of slow, inefficient, 
1418: star formation in a relatively quiescent, large-scale 
1419: condensation (which can be a filament if the cloud is flattened 
1420: in a direction perpendicular to the field lines to begin with) 
1421: that is embedded in a much more turbulent halo of diffuse gas. 
1422: 
1423: The above picture is an unavoidable consequence of the anisotropy 
1424: intrinsic to the large-scale magnetic support. Before the self-gravity 
1425: becomes strong enough to overwhelm the magnetic support in the 
1426: cross-field direction (to initiate widespread star formation), it  
1427: should be able to condense material along the field lines (where 
1428: there is no magnetic resistance) first. The condensed material 
1429: is expected to be relatively quiescent. Its field-aligned velocity 
1430: component is damped during the condensation process, whereas its 
1431: cross-field motions are constrained by the field lines. The 
1432: lateral magnetic support ensures that the bulk of the condensed 
1433: material is relatively long-lived, allowing more time for the 
1434: turbulence to dissipate. The dissipation is hastened by the high 
1435: density in the condensation, which makes it easier for shock 
1436: formation, because of low Alfv\'en speed. Indeed, when the condensed 
1437: layer accumulates enough material for its mass-to-flux 
1438: ratio to approach the critical value (a condition for widespread 
1439: star formation), its Alfv\'en speed drops automatically to a value 
1440: comparable to the sound speed (NL05, Kudoh et al. 2007). Rapid 
1441: dissipation of super-magnetosonic motions ensures that the bulk of 
1442: the condensed material moves at no more than moderately supersonic 
1443: speeds, as we find here and previously in 2D calculations (NL05). 
1444: It is in this relatively quiescent environment that most star 
1445: formation takes place.
1446: 
1447: The quiescent, star-forming, condensed material is surrounded by 
1448: a highly turbulent, fluffy halo. The stratification in turbulent 
1449: speed is a consequence of the stratification in density (Kudoh \& 
1450: Basu 2003): the amplitude of a wave increases as it propagates 
1451: from high density to low density, and faster motions can survive 
1452: for longer in the lower density region because of higher Alfv\'en 
1453: speed. We believe that this is a generic result independent of 
1454: the source of turbulence. In our particular simulations, the 
1455: saturated, highly supersonic turbulence in the halo at late times 
1456: is driven mainly by protostellar outflows internally. It is 
1457: conceivable, perhaps even likely, that most of the turbulence 
1458: in the diffuse regions of star forming clouds comes from energy 
1459: cascade from larger-scale molecular gas (perhaps even HI envelope). 
1460: Depending on the details of energy injection, the external driving 
1461: of turbulence can slow down the 
1462: gravitational settling along field lines and may even suppress 
1463: it altogether. Even when the large-scale gravitational settling 
1464: is suppressed, star formation can still occur, perhaps in a less 
1465: coherent fashion: stars can form in localized pockets of a 
1466: magnetically subcritical cloud directly from strong compression 
1467: through shock-enhanced ambipolar diffusion (NL05), especially
1468: when the shock is highly super-Alfv\'enic (Kudoh \& Basu 2008). 
1469: Vigorous driving may favor a mode of turbulence-accelerated star 
1470: formation at isolated locations, whereas decay of turbulence may 
1471: allow the gravity to pull the diffuse gas into large-scale 
1472: condensations, in which stars form in a more coherent fashion. 
1473: The first mode is now under investigation, and the results will 
1474: be reported elsewhere. The second corresponds to the picture 
1475: outlined above.  
1476: 
1477: The dichotomy is not unique to star formation in strongly magnetized 
1478: clouds. It is also present when the magnetic field is weak or 
1479: non-existent: whereas stars form rapidly in a clustered mode in a 
1480: decaying non-magnetic turbulence, they are produced more slowly, in 
1481: a more dispersed fashion, if the turbulence is driven constantly, 
1482: especially on small scales (e.g., Mac Low \& Klessen 2004). Different 
1483: modes of star formation may operate in different environments, and 
1484: there is a urgent need for sorting out which mode dominates in which 
1485: environment. In the next subsection, we will concentrate on the Taurus 
1486: molecular clouds, the prototype of distributed low-mass star forming 
1487: regions (see Kenyon et al. 2008 for a recent review), and the Pipe 
1488: nebula, which appears to be a younger version of the Taurus complex 
1489: (Lada et al. 2008). 
1490: 
1491: \subsection{Connection to Observations} 
1492: 
1493: \subsubsection{The Taurus Molecular Cloud Complex}
1494: \label{taurus}
1495: 
1496: The Taurus molecular cloud complex may represent the best case for 
1497: the picture of magnetically regulated star formation investigated 
1498: in this paper. One line of evidence supporting this assertion 
1499: comes from the studies of young stars by Palla \& Stahler (2002). 
1500: They inferred, using pre-main sequence tracks, that the star formation 
1501: in the complex has two phases. It began at least 10 Myrs ago, at 
1502: a relatively low level and in widely dispersed locations (Phase 
1503: I hereafter). In the last three million years or so, stars are 
1504: produced at a much higher rate (Phase II). The majority of the 
1505: stars produced in Phase II are confined within the dense, 
1506: striated, C$^{18}$O filaments (Onishi et al. 2002). Palla \& 
1507: Stahler interpret these results to mean that the Taurus clouds 
1508: evolved from an initial, relatively diffuse, state 
1509: quasi-statically. They envision that, during the early epoch 
1510: of quasi-static contraction, few places have contracted to 
1511: the point of forming dense cores and protostars, and hence only 
1512: scattered star formation activity is seen at isolated locations 
1513: (Phase I). 
1514: Eventually, further contraction of the cloud leads to the formation 
1515: of nearly contiguous, dense filaments where multiple cores can 
1516: form and collapse simultaneously, leading to elevated star formation 
1517: (Phase II). 
1518: % The authors did not speculate on the physical mechanisms 
1519: % that may be responsible for this behavior. 
1520: In view of the discussion 
1521: in the last subsection, it is natural to identify Phase I with the 
1522: shock-accelerated mode of star formation in a magnetically 
1523: subcritical or nearly critical cloud (or sub-region), when the 
1524: turbulence is still 
1525: strong enough to prevent large-scale gravitational condensation 
1526: along the field lines. In this picture, Phase II started when the
1527: turbulence had decayed to a low enough level that the gravity was  
1528: able to collect enough mass along the field lines into condensed 
1529: structures for a faster, more organized, mode of star formation 
1530: to take over. 
1531: 
1532: A salient feature of our picture is that the rate of star formation 
1533: in Phase II should remain well below the limiting free-fall rate 
1534: of the condensed material. There is strong evidence that the 
1535: conversion of dense material in the Taurus clouds into stars is 
1536: indeed slow. The dense, filamentary material is well traced by 
1537: C$^{18}$O, which Onishi et al. (1996) has studied in great detail. 
1538: These authors estimated a mass and average density for the C$^{18}$O 
1539: gas of $M_{{\rm C}^{18}{\rm O}}\sim 2900~M_\odot$ and $n_{H_2}\sim
1540: 4000$~cm$^{-3}$ (comparable to that in the condensed sheet in 
1541: our simulations). The latter corresponds a free fall time 
1542: $t_{{\rm C}^{18}{\rm O},ff}=4.86 \times 10^5$~yrs. The maximum, free-fall, 
1543: rate of turning C$^{18}$O material into stars is therefore 
1544: ${\dot M}_{*,{\rm C}^{18}{\rm O},ff}\sim 6\times
1545: 10^{-3}~M_\odot$~yr$^{-1}$. This 
1546: rate is to be compared with the actual rate of star formation in 
1547: the complex, ${\dot M}_*\sim 5\times 10^{-5}~M_\odot$yr$^{-1}$, 
1548: estimated by Goldsmith et al. (2008) based on an assumed average 
1549: mass of $0.6~M_\odot$ for the 230 stars observed in their
1550: mapped region (Kenyon et al. 2008) and a duration of 3~Myrs for 
1551: Phase II of star formation where most stars are produced (Palla 
1552: \& Stahler 2002). Therefore, the rate of converting the even 
1553: dense (relative to the diffuse $^{12}$CO gas) C$^{18}$O material 
1554: into stars is extremely slow, amounting to $\sim 1\%$ per local 
1555: free fall time. This rate is comparable to those obtained in our 
1556: simulations. 
1557: %
1558: % comparable to Krumholz's estimates for denser gas!
1559: % 
1560: 
1561: One may argue that the slow rate of star formation in C$^{18}$O
1562: gas is due to turbulence support. However, the C$^{18}$O gas 
1563: in the Taurus clouds is observed to be unusually quiescent. 
1564: Onishi et al. (1996) estimated a mean equivalent line-width 
1565: (defined as the integrated intensity divided by the peak 
1566: temperature of the line) of 0.65~km~s$^{-1}$, corresponding 
1567: to a 3D velocity dispersion of 0.45~km~s$^{-1}$ for a Gaussian
1568: profile. The velocity dispersion is $\sim 2.4$ times the 
1569: sound speed for a gas temperature of 10~K, indicating that 
1570: the moderately dense gas traced by C$^{18}$O is not 
1571: highly supersonic in general. Indeed, the velocity 
1572: dispersion is quite comparable to that of the condensed
1573: sheet material in our standard model at similar densities
1574: (see Fig.~\ref{rms_vel}). Nearly 1/3 ($937~M_\odot$ out 
1575: of $2900~M_\odot$) of the C$^{18}$O gas is classified 
1576: into 40 cores, which have a mean line-width (FWHM) of 
1577: 0.49~km~s$^{-1}$, comparable to that estimated above 
1578: for the C$^{18}$O gas as a whole. Onishi et al. concluded 
1579: that most of the cores are roughly gravitationally bound, 
1580: based on their masses, sizes and line-widths. These authors 
1581: argued that the general alignment of the minor axes of the 
1582: cores with the overall field direction indicates that the 
1583: magnetic field plays a role in their formation. We agree 
1584: with this assessment. 
1585: The relatively low velocity dispersion of the C$^{18}$O gas 
1586: and the morphology of dense cores provide indirect support 
1587: for the picture of magnetically regulated star formation. 
1588: 
1589: A more direct test of the picture would come from a 
1590: determination of the mass-to-flux ratio. 
1591: %
1592: % what is predicted? can moderately supercritical clouds 
1593: % work? 
1594: % 
1595: For a given field strength $B$, the critical column density 
1596: along the field line is
1597: \begin{equation}
1598: \Sigma_{\vert\vert,c}=58.7~B_{20}~M_\odot~{\rm pc}^{-2}, 
1599: \label{critical}
1600: \end{equation}
1601: corresponding to an $A_V \approx 2.5~B_{20}$, where $B_{20}$ 
1602: is the field strength in units of 20~$\mu$G. 
1603: Goldsmith et al. (2008) estimated a total mass of 
1604: $2.4\times 10^4 M_\odot$ in their 
1605: mapped area of $28$~pc$\times 21$~pc, yielding an average 
1606: column density of 40.8~M$_\odot$~pc$^{-2}$ along the line 
1607: of sight. Only $\sim 2/3$ of the surveyed area have detectable 
1608: $^{12}$CO emission in individual spectra, however (their 
1609: Masks 1 and 2). For these regions, the total mass and area 
1610: are $1.95\times 10^4 M_\odot$ and $\sim 392$~pc$^2$, yielding 
1611: a somewhat higher average column density of 49.7~M$_\odot$~pc$^{-2}$ 
1612: along the line of sight. If the column density along the field 
1613: lines is comparable to this value, the corresponding critical 
1614: field strength would be $16.9~\mu$G, according to 
1615: equation~(\ref{critical}). About half of the mass (9807~M$_\odot$) 
1616: resides in eight ``high-density'' regions (such as L1495 and 
1617: Heiles Cloud 2) that together occupy $\sim 31\%$ of the area, 
1618: with an average line-of-sight column density of 
1619: 79.5~M$_\odot$~pc$^{-2}$. This value would correspond to 
1620: a critical field strength of $27.1~\mu$G if the magnetic field 
1621: is exactly along the line of sight. There is, however, evidence 
1622: for a large-scale magnetic field in the plane of sky in the 
1623: Taurus cloud complex from the polarization of background 
1624: star light. Indeed, the velocity anisotropy measured by Heyer 
1625: et al. (2008) in a relatively diffuse sub-region points to 
1626: a magnetic field that is mainly in the plane of sky. If this 
1627: magnetic orientation is correct, and if matter condenses along the 
1628: field lines into flattened structures, then projection effects 
1629: may be considerable. Applying a moderate correction factor 
1630: of $\sqrt{2}$ (corresponding to a viewing angle of $45^\circ$) 
1631: would bring the average field-aligned column density of the 
1632: ``high-density'' regions down to 56.2~M$_\odot$~pc$^{-2}$, with 
1633: a corresponding  
1634: critical field strength of $19.2~\mu$~G. It is therefore likely 
1635: that a magnetic field of order 20~$\mu$G is strong enough to 
1636: regulate star formation in the Taurus cloud complex. 
1637: 
1638: The magnetic field strength has been measured for three dense 
1639: cores in Taurus through OH Zeeman measurements using Arecibo 
1640: telescope (Troland \& Crutcher 2008). They are B217-2, TMC-1 
1641: and L1544, with a line-of-sight field strength of $B_{\rm los} 
1642: \approx 13.5$, $9.1$, and $10.8$~$\mu$G respectively. Although 
1643: the true orientation of the magnetic field is unknown, a 
1644: statistically most probable correction factor of 2 would bring 
1645: the total field strength to $\sim 18.2 - 27.0~\mu$G. Given that 
1646: OH samples moderately dense gas approximately as C$^{18}$O 
1647: does, and that the field strength changes relatively little at 
1648: low to moderate densities (see Fig.~\ref{ave_B}), it is plausible 
1649: such a field permeates most of the space in the complex. A 
1650: worry is that such a strong field is not detected in several 
1651: other cores in the Taurus in the same Arecibo survey. Whether
1652: these non-detections can be reconciled with our picture of
1653: magnetically regulated star formation through, e.g., unfavorable 
1654: field orientation, remains to be seen. 
1655: 
1656: Indirect evidence for a relatively strong magnetic field 
1657: comes from the observation that the magnetic field as
1658: traced by the polarization vectors of the background star
1659: light is well ordered over many degrees in the sky, 
1660: indicating that the magnetic energy is larger than the 
1661: turbulent energy.
1662: If we use conservative estimates for H$_2$ number density 
1663: of $10^2$~cm$^{-3}$ (see Table 2 of Goldsmith et al. 2008), 
1664: and turbulent speed of 2~km~s$^{-1}$,
1665: we can put a lower limit to the field strength at $\sim 
1666: 15~\mu$G. The limit agrees with the value of $\sim 14~\mu$G 
1667: estimated by Heyer et al. (2008) based on comparing the 
1668: observed velocity anisotropy in a striated, relatively 
1669: diffuse region with MHD turbulence simulations; they point 
1670: out that 
1671: the value is to be increased if the magnetic field has 
1672: a substantial component along the line of sight. A
1673: strong line-of-sight magnetic field of $\sim 20~\mu$G 
1674: or more is deduced by Wolleben \& Reich (2004) at the 
1675: boundaries of molecular clouds a few degrees away from 
1676: the sub-region studied by Heyer et al. (2008), from 
1677: modeling the enhanced rotation measures inferred from 
1678: polarization observations at 21 and 18~cm. Subsequent 
1679: HI Zeeman observations did not confirm the deduced 
1680: field strength, however (C. Heiles, priv. comm.). 
1681: 
1682: Although none of the pieces of evidence discussed above is 
1683: conclusive individually, taken together, they make a 
1684: reasonably strong case for a global magnetic field of order 
1685: $\sim 20\mu$G in strength. The existence of such a field 
1686: may not be too surprising from an evolutionary point of view. 
1687: It is now established that the cold neutral medium (CNM) of 
1688: HI gas (likely precursor of Taurus-like molecular clouds)   
1689: is fairly strongly magnetized in general, with a median
1690: field strength of $\sim 6~\mu$G (Heiles \& Troland 2005). 
1691: For denser HI clouds, it is not difficult to imagine a 
1692: somewhat higher field strength. 
1693: A case in point is the nearby Riegel-Crutcher HI cloud 
1694: that we mentioned in the introduction. It was mapped 
1695: recently by McClure-Griffiths et al. (2006) using 21~cm 
1696: absorption against strong continuum emission towards the 
1697: Galactic center. They found long strands of cold neutral
1698: hydrogen that are remarkably similar to the CO striations 
1699: in Taurus studied by Heyer et al. (2008): both are aligned 
1700: with the local magnetic field. For the magnetic energy to
1701: dominate the turbulent energy inside the filaments, the
1702: field strength must be of order $30~\mu$G or larger. This
1703: value is consistent with $B_{los}\sim 18~\mu$G obtained 
1704: by Kazes \& Crutcher (1986) through Zeeman measurements 
1705: at a nearby location, if a correction of $\sim 2$ is
1706: applied for projection effects. The physical conditions 
1707: inside the HI filaments of the Riegel-Crutcher cloud are 
1708: not dissimilar to those of diffuse CO clouds: number 
1709: density $n_H\sim 460$~cm$^{-3}$, 
1710: velocity dispersion $\sigma \sim 3.5$~km~s$^{-1}$, and 
1711: temperature $T\sim 40$~K. The main difference is that the 
1712: column density is $\sim 10^{20}$~cm$^{-2}$, which is not 
1713: sufficient for shielding the CO from photodissociation 
1714: (van Dishoeck \& Black 1988). 
1715: 
1716: The above discussion leads us to the following scenario for 
1717: cloud evolution and star formation in the Taurus molecular
1718: clouds: we envision the cloud complex to have evolved from a 
1719: turbulent, strongly magnetized, non-self-gravitating HI cloud 
1720: similar to the Riegel-Crutcher cloud. Once enough material 
1721: has been accumulated, perhaps by converging flows 
1722: (Ballesteros-Paredes et al. 1999) or some other means
1723: (such as differential turbulence dissipation, which may 
1724: lead to a condensation through ``cooling flow'' along the 
1725: field lines, e.g., Myers \& Lazarian 1998), for CO 
1726: shielding, it becomes visible as a molecular cloud. 
1727: Depending on the degree of magnetization, level of 
1728: inhomogeneity, and strength of turbulence, star formation 
1729: can start quickly in some pockets (perhaps even during the 
1730: process of molecular cloud formation), despite of a strong 
1731: global magnetic field, either because they happen to be 
1732: less magnetized to begin with or because their magnetic 
1733: fluxes are force-reduced by shock-enhanced ambipolar 
1734: diffusion. This corresponds to the low-level, Phase I of 
1735: star formation deduced by Palla \& Stahler (2002). Once 
1736: enough material has settled along field lines for the  
1737: mass-to-flux ratio of the condensed material to approach 
1738: the critical value, the rate of star formation increases 
1739: drastically. We identify this phase of more organized star 
1740: formation in condensed structures as the more active, 
1741: Phase II of star formation deduced by Palla \& Stahler 
1742: (2002). Even during this relatively active phase, the star 
1743: formation rate is still well below the free-fall rate of 
1744: the condensed gas, because of magnetic regulation. 
1745: %
1746: % 330 solar masses in H13CO+, SFE rate 10 times slower than free 
1747: % fall rate?? 
1748: %
1749: 
1750: The above picture can be contrasted with that of Ballesteros-Paredes 
1751: et al. (1999) and Hartmann et al. (2001). These authors advocated 
1752: a picture of rapid cloud formation and star formation for the
1753: Taurus-Auriga complex, motivated mainly by a lack of ``post-T Tauri 
1754: stars'' older than $\sim 5$~Myrs in the region; indeed, the 
1755: majority of stars are younger than $\sim 3$~Myrs, as mentioned 
1756: earlier. They argue that it is difficult to understand how star 
1757: formation can occur simultaneously over a spatial extent of $\sim 
1758: 20$~pc (or more) over a timescale as short as a few million years, 
1759: given that the observed turbulent speed is $\sim 2$~km~s$^{-1}$, 
1760: corresponding to a crossing time of $\sim 10$~Myrs (or more), 
1761: unless the star formation is coordinated by some agent other than 
1762: the currently observed turbulence. They suggest that this agent is 
1763: external: a large-scale converging flow with speeds of 5-10~km~s$^{-1}$.  
1764: In our picture, it is the global gravity and a strong, large-scale, 
1765: internal magnetic field that coordinate the star formation: the 
1766: strong field suppresses vigorous star formation until enough 
1767: material has collected along field lines into condensed structures 
1768: for the gravity to start overwhelm magnetic support in the 
1769: cross-field directions. The degree to which the star formation is 
1770: actually synchronized will, of course, depend on how the mass is 
1771: initially distributed along the field lines and rate of turbulence 
1772: dissipation and replenishment, neither of which is, unfortunately, 
1773: well constrained at the present. Indeed, a hybrid scenario may be
1774: possible, with the mass accumulation along the field lines sped 
1775: up by a large-scale converging flow. Nevertheless, we believe 
1776: that, at least in the particular case of the Taurus molecular cloud 
1777: complex, for the large amount of dense, relatively quiescent, 
1778: C$^{18}$O gas to form stars at $\sim 1\%$ of the free-fall rate, 
1779: magnetic regulation is needed. A similar case can be made for the 
1780: Pipe Nebula.   
1781: 
1782: \subsubsection{Pipe Nebula}
1783: \label{pipe}
1784: 
1785: Although less well studied than the Taurus cloud complex, the Pipe 
1786: nebula is an important testing ground for theories of low-mass 
1787: star formation in relatively diffuse dark clouds. It has $\sim 
1788: 10^4~M_\odot$ of molecular material as traced by $^{12}$CO (Onishi 
1789: et al. 1999), not much different from the Taurus complex. The higher 
1790: density gas traced by $^{13}$CO and C$^{18}$O is concentrated in 
1791: a long filament that spans more than 10~pc, again similar to the
1792: Taurus case. The main difference is that there is no evidence
1793: for star formation activity except in the most massive core, B59. 
1794: The lower rate of star formation indicates that the Pipe nebula is 
1795: at an earlier stage of evolution compared to the Taurus complex, as 
1796: pointed out by Muench et al. (2007). The relative inactivity is 
1797: not due to a lack of dense, quiescent material that is a 
1798: pre-requisite of (low-mass) star formation. J. Alves et al. (2007) 
1799: identified 159 dense cores from extinction maps, many of which are 
1800: formally gravitationally unstable, with masses above the critical 
1801: Bonner-Ebert mass (Lada et al. 2008). In the absence of additional 
1802: support, such cores should collapse and form stars on the free-fall 
1803: time scale, which is $t_{ff}=3.7\times 10^5$~yrs for the inferred 
1804: median core density $n(H_2)=7\times 10^3$~cm$^{-3}$. This time 
1805: scale is much shorter than the turbulence crossing time $t_c=L/v 
1806: \sim 5\times 10^6$~yrs, over a region of $L \sim 15$~pc, and for 
1807: a typical turbulent speed of $v\sim 3$~km~s$^{-1}$. The mismatch 
1808: between the turbulent crossing time and star formation time is 
1809: therefore much more severe than that noted above for the Taurus 
1810: region. 
1811: 
1812: There are two possible resolutions to the above conundrum. If 
1813: the dense cores (at least those massive enough to be formally 
1814: gravitationally unstable) are indeed short-lived objects, with 
1815: a lifetime of order the average free-fall time ($\sim 3\times 
1816: 10^5$~yrs), their formation must be rapid and well coordinated. 
1817: It is tempting to 
1818: attribute the coordination to a large-scale converging flow, as 
1819: Ballesteros-Paredes et al. (1999) and Hartmann et al. (2001) 
1820: did for the Taurus case. However, to cover a distance of $15$~pc 
1821: (the size of the cloud, which is comparable to the length of 
1822: the chain of dense cores) in $\sim 3\times 10^5$ years, the 
1823: flow needs to converge at a speed of order $\sim 50$~km~$s^{-1}$, 
1824: which is uncomfortably high. A more likely alternative is that 
1825: the cores live much longer 
1826: than their free-fall times would indicate. While the longevity 
1827: is not necessarily a problem for those non-self-gravitating, 
1828: lower mass cores that can be confined by external pressure, 
1829: it is problematic for the more massive cores that should have 
1830: collapsed on the free-fall time scale, unless they are supported 
1831: by some (so-far invisible) agent in addition to the observed
1832: thermal and turbulent pressures. The most likely candidate is
1833: a magnetic field. There is some evidence for the existence of 
1834: an ordered magnetic field that is more or less perpendicular 
1835: to the long filament that contains the dense cores from stellar 
1836: polarization observations (F. Alves \& Franco 2007; F. Alves et
1837: al., in preparation), although much more work is needed, particularly 
1838: for the diffuse region outside the filament, to firm up or reject 
1839: this possibility. In any case, it is plausible that 
1840: the Pipe nebula represents an earlier stage of the magnetically 
1841: regulated cloud evolution and star formation than the Taurus 
1842: clouds: material has started to collect along the field lines 
1843: into dense, relatively quiescent, structures, but their 
1844: mass-to-flux ratios are yet to reach the critical value, 
1845: except perhaps in the most massive core, B59. In this case, 
1846: the ages of the cores are set by the rate of mass condensation 
1847: along the field lines, which may depend on the ill-understood 
1848: process of turbulence replenishment in the diffuse gas, and 
1849: may be accelerated by large-scale converging flows (Ballesteros 
1850: et al. 1999; Hartmann et al. 2001). 
1851: %
1852: % Southern coalsack??? earlier phase???   
1853: %
1854: 
1855: \subsection{Modes of Star Formation in Strongly Magnetized Clouds}
1856: \label{mode}
1857: 
1858: Strong magnetic fields tend to resist cloud condensation and 
1859: star formation. This resistance can be overcome by either 
1860: turbulence or gravity, leading to two distinct modes of star 
1861: formation. Which mode dominates depends on the relative
1862: importance of gravity and turbulence. When the turbulence is 
1863: strong enough to prevent gravitational settling along field 
1864: lines, star formation is still possible. Turbulence can 
1865: overcome magnetic resistance in localized regions that are 
1866: compressed by converging flows to high densities; such regions 
1867: are expected to have large magnetic field gradients and low 
1868: fractional ionization, both of which enhance the rate of 
1869: ambipolar diffusion. The 
1870: turbulence-accelerated ambipolar diffusion enables dense
1871: pockets to become self-gravitating 
1872: %\footnote{Dib et al. (2007) 
1873: %found that the majority of the dense cores formed in a rather 
1874: %weakly magnetized cloud of dimensionless mass-to-flux ratio 
1875: %of $\lambda=2.8$ with driven turbulence are nearly magnetically 
1876: %critical or even subcritical. Such cores are expected to be
1877: %strongly affected by enhanced ambipolar diffusion as well.} 
1878: and collapse into stars (NL05, Kudoh \& Basu 2008), even when 
1879: the background cloud 
1880: remains well supported. In this mode, the crucial step of 
1881: star formation---core formation---is driven by turbulent 
1882: converging flows but regulated by magnetic fields, with 
1883: self-gravity playing a secondary role. After a bound core 
1884: has formed, its further evolution is driven primarily by 
1885: the self-gravity, but remains regulated by the magnetic 
1886: field. 
1887: 
1888: The second mode corresponds to a turbulence that is not strong 
1889: enough to prevent gravitational settling, either because it 
1890: is weak to begin with or because it has decayed without being 
1891: adequately replenished. Condensation would occur along field 
1892: lines, which would lead to further increase in self-gravity 
1893: and more rapid dissipation of turbulence. In the condensed 
1894: structure, turbulence becomes less important dynamically, and 
1895: star formation is driven mainly by gravity in the cross-field 
1896: direction. The gravitational contraction is resisted primarily 
1897: by magnetic fields, although protostellar outflows play an 
1898: important role in dispersing away the dense gas that remains 
1899: after the formation of individual stars; the dispersal slows 
1900: down the global star formation. In this mode, the gravity and 
1901: magnetic fields are of primary importance, with turbulence 
1902: playing a secondary role. 
1903: 
1904: The above two modes of star formation may be interconnected: it 
1905: is natural to expect the second to follow the first, as a result 
1906: of turbulence dissipation. The transition from one to the 
1907: other should occur when the mass-to-flux ratios in the condensed 
1908: structures approach the critical value. It may be possible for 
1909: the mass-to-flux ratio to stay near the critical value due to 
1910: outflow feedback: if too much mass is added to the structure 
1911: so that its mass-to-flux ratio increases well above the critical 
1912: value, it would form stars at a higher rate, producing more 
1913: outflows that stop further mass accumulation, similar to the 
1914: outflow-regulated scenario of cluster formation envisioned 
1915: in Matzner \& McKee (2000). They have shown that regions of 
1916: outflow-regulated star formation tend to be overstable, 
1917: oscillating with increasingly large amplitudes. Indeed, the 
1918: outflows may unbind the condensed structure completely, 
1919: terminating its star formation quickly. An abrupt termination 
1920: is implied by the dearth of the well-observed star forming 
1921: regions in the declining phase of star formation (Palla \& 
1922: Stahler 2000). 
1923: 
1924: If, on the other hand, outflows fail to retard mass accumulation 
1925: along field lines, the star formation may change over to 
1926: yet another, more active, mode --- cluster formation. The 
1927: transition is expected to take place when the mass-to-flux ratios 
1928: of the condensations become substantially greater than the critical 
1929: value (say by a factor of two), making global (not just local) 
1930: contraction perpendicular to the field lines possible. The
1931: higher rate of star formation should lead to a stronger 
1932: outflow feedback, which could increase the turbulence 
1933: to a level that balances the global collapse (e.g., Nakamura 
1934: \& Li 2007; Matzner 2007). In this mode, protostellar 
1935: outflow-driven turbulence becomes the primary agent that 
1936: resists the global gravitational collapse, with the magnetic 
1937: field playing a secondary role. 
1938: 
1939: In summary, we envision three, perhaps sequential, modes of star 
1940: formation in strongly magnetized clouds, dominated (a) first by 
1941: turbulence and magnetic fields, (b) then by magnetic fields and 
1942: gravity, and (c) finally by gravity and turbulence (again). In 
1943: this sequence, stars form first directly out of turbulent diffuse 
1944: material in a dispersed fashion, then more coherently in 
1945: relatively quiescent, condensed structures, and finally as 
1946: clusters in dense turbulent clumps. The strong turbulence 
1947: is primordial initially. It is transformed into a protostellar 
1948: turbulence by outflows in regions of active cluster formation.
1949: %
1950: % B important in quiescent regions, emphasized by Cox; true for SF as well. 
1951: %
1952: % (b) is unique to strong field case
1953: 
1954: \section{Summary and Conclusions}
1955: \label{conclude}
1956: 
1957: We have performed 3D simulations of star formation in turbulent, 
1958: mildly magnetically subcritical clouds including protostellar 
1959: outflows and ambipolar diffusion. The simulations are motivated
1960: by observations of molecular gas and young stars in the Taurus
1961: cloud complex and the realization that cold neutral HI gas, a 
1962: likely progenitor of molecular gas, is magnetically subcritical
1963: in general. They are natural extensions of our previous 2D 
1964: calculations. The main results are:  
1965: 
1966: 1. The decay of turbulence leads to gravitational condensation 
1967: of material along the field lines. In the ideal MHD limit, the 
1968: resulting structure closely resembles a smooth Spitzer sheet, 
1969: as found previously (Krasnopolsky \& Gammie 2005). Ambipolar 
1970: diffusion changes the structure drastically. It creates 
1971: supercritical material that behaves differently from the 
1972: subcritical background in a fundamental way: it can be condensed 
1973: further by self-gravity. As a result, the distributions of mass 
1974: with respect to volume and column densities are greatly broadened. 
1975: Indeed, the probability distribution function (PDF) of the (more 
1976: easily observable) column density can be fitted approximately 
1977: by a broad, lognormal distribution (see Fig.~\ref{colden_PDF}), 
1978: as is the case for weakly magnetized clouds fragmented by strong 
1979: turbulence (e.g., Padoan \& Nordlund 2002; P. S. Li et al. 2004). 
1980: It may be difficult to distinguish the turbulent fragmentation 
1981: in weakly magnetized clouds and the ambipolar diffusion-enabled 
1982: gravitational fragmentation in strongly magnetized clouds based 
1983: on mass distribution. 
1984: 
1985: 2. The magnetized clouds evolve into a configuration strongly 
1986: stratified in density at late times, with dense cores embedded 
1987: in fragmentary condensations which, in turn, are surrounded 
1988: by a more diffuse halo (see Fig.~\ref{3D}). 
1989: The condensed material can be significantly warped and may appear 
1990: thicker than its intrinsic width, even when viewed edge-on, 
1991: because of superposition of warps along the line of sight. While 
1992: the diffuse halo is highly turbulent, the condensations are 
1993: more quiescent, with moderately supersonic velocity dispersions
1994: in general. One reason is that the bulk of the condensed material 
1995: is magnetically supported, which allows more time for the 
1996: turbulence to decay. Another is their relatively low Alfv\'en 
1997: speed (comparable to the sound speed), which forces highly 
1998: supersonic motions to dissipate quickly, through shock 
1999: formation. The moderately supersonic turbulence is maintained, 
2000: however, for many local free-fall times, through a combination 
2001: of outflow feedback, MHD waves, and gravitational motions 
2002: induced by ambipolar diffusion.
2003: %, even in the absence of any external driving. 
2004:   
2005: 3. Stars form at a low rate. Only one percent or less of the cloud 
2006: material is converted into stars in a free-fall time at the 
2007: characteristic density of the condensed material. The slow star 
2008: formation cannot be due to support by turbulence because the  
2009: star-forming, condensed material is rather quiescent. It is 
2010: regulated by magnetic fields. The low rate of star formation is 
2011: not sensitive to the level of initial turbulence, although more 
2012: turbulent clouds produce stars early, consistent with the picture 
2013: of turbulence-accelerated star formation in strongly magnetized 
2014: clouds. The rate of star formation is further reduced by 
2015: protostellar outflow, particularly the component in the plane 
2016: of mass condensation.   
2017: 
2018: 4. Dense cores formed in our simulations are typically triaxial.  
2019: More massive cores tend to have lower magnetic flux-to-mass
2020: ratios, be more tightly bound gravitationally, and be more 
2021: oblate. Virial analysis indicates that the surface term due to 
2022: external kinetic motions contributes significantly in general, 
2023: especially for less massive cores. We find a larger number 
2024: of bound cores than in previous ideal MHD simulations even 
2025: though our clouds are more strongly magnetized; they are made 
2026: possible by ambipolar diffusion. The cores have specific 
2027: angular momenta comparable to the observed values, and a mass 
2028: spectrum that resembles the Salpeter stellar IMF towards the 
2029: high mass end. The spectrum flattens near the characteristic 
2030: Jeans mass of the condensed material, which is a few solar 
2031: masses for typical parameters. The stellar mass spectrum also 
2032: resembles the Salpeter IMF. 
2033: 
2034: 5. The Taurus molecular cloud complex may represent the best case 
2035: for magnetically regulated star formation. The strongest evidence 
2036: comes from the existence of a large amount of dense, relatively 
2037: quiescent, C$^{18}$O gas, which produces stars at a rate two orders 
2038: of magnitude below the maximum free-fall rate. Most likely, the 
2039: star formation is regulated by a strong, ordered magnetic field. 
2040: Several lines of evidence suggest that a magnetic field of the 
2041: required strength (of order $\sim 20 \mu G$) is indeed present 
2042: in the region, although direct Zeeman measurements are available 
2043: for only three cores, and their interpretations are complicated 
2044: by projection effects. 
2045: 
2046: 6. We speculate that, depending on the (uncertain) rates of 
2047: turbulence replenishment and outflow feedback, there may be 
2048: three distinct modes of star formation in strongly magnetized 
2049: clouds: (a) inefficient star formation through 
2050: turbulence-accelerated ambipolar diffusion at dispersed 
2051: locations in relatively diffuse, magnetically subcritical 
2052: clouds, (b) more coherent, but still inefficient, 
2053: magnetically regulated star formation in relatively 
2054: quiescent, nearly magnetically critical condensations, 
2055: (c) more efficient cluster formation in magnetically 
2056: supercritical clumps that may be regulated by outflow-driven 
2057: protostellar turbulence. Much work remains to be done to 
2058: understand these modes and their possible interconnection. 
2059: 
2060: \acknowledgments 
2061: This work is supported in part by a Grant-in-Aid for Scientific Research 
2062: of Japan (18540234, 20540228),  
2063: and NSF and NASA grants (AST-0307368 and NAG5-12102). 
2064: Parts of this work were performed while the authors were in residence
2065: at the Kavli Institute for Theoretical Physics at UCSB.
2066: The numerical calculations were carried out mainly 
2067: on NEC SX8 at YITP in Kyoto University, 
2068: on Fujitsu VPP5000 at the Center for Computational Astrophysics, CfCA, 
2069: of the National Astronomical Observatory of Japan,
2070: and on Hitachi SR8000 at Center for Computational Science 
2071: in University of Tsukuba.
2072: 
2073: 
2074: 
2075: \begin{thebibliography}{99}
2076: \bibitem[]{}
2077: Alves, F. O. \& Franco, G. A. P. 2007, \aap, 470, 597
2078: \bibitem[]{}
2079: Alves, J., Lombardi, M. \& Lada, C. F. 2007, \aap, 462, 17
2080: \bibitem[]{}
2081: Andre, P., Ward-Thompson, D. \& Barsony, M. 1993, \apj, 406, 122
2082: \bibitem[]{}
2083: Ballesteros-Paredes, J. Hartmann, L., \& Vazquez-Semadeni, E. 1999,
2084:           \apj, 527, 285
2085: \bibitem[]{}
2086: Basu, S. \& Ciolek, G. E., 2004, \apj, 607, 39
2087: \bibitem[]{}
2088: Bertoldi, F. \& McKee, C. F. 1992, \apj, 395, 140
2089: \bibitem[]{}
2090: Caselli, P. \& Myers, P. C. 1995, \apj, 446, 665
2091: \bibitem[]{}
2092: Dib, S., Vazquez-Semadeni, E., Kim, J., Burkert, A., \& Shadmehri,
2093:           M. 2007, \apj, 661, 262 
2094: \bibitem[]{}
2095: Elmegreen, B. G. 2007, \apj, 668, 1064
2096: \bibitem[]{}
2097: Gammie, C. F., Lin, Y.-T., Stone, J. M., \& Ostriker, E. C. 2003, \apj,
2098:           592, 203
2099: \bibitem[]{}
2100: Goldsmith et al. 2008 (astro-ph/arXiv:0802.2206)
2101: \bibitem[]{}
2102: Gomez, M. Hartmann, L., Kenyon, S. J., \& Hewett, R. 1993, \apj, 105, 1927
2103: \bibitem[]{}
2104:   Hartmann, L., Ballesterios-Paredes, J. \& Bergin, E. A. 2001, \apj,
2105:           562, 852
2106: \bibitem[]{}
2107: Heiles, C. \& Crutcher, R. M. 2005, in Cosmic Magnetic Fields, Lecture
2108: Notes in Physics, vol. 664, pp137-182
2109: \bibitem[]{}
2110: Heiles, C. \& Troland, T. H. 2005, \apj, 624, 773
2111: \bibitem[]{}
2112:  Heyer, M., Gong, H., Ostriker, E. Brunt, C. 2008, (astro-ph/arXiv:0802.2084)
2113: \bibitem[]{}
2114: Heyer, M. H., Frederick, J., Snell, R. L., Schloerb, F. P., Strom,
2115:           S. E., Goldsmith, P. F., \& Strom, K. M. 1987, \apj, 321, 855 
2116: \bibitem[]{}
2117: Ikeda, N., Sunada, K., \& Kitamura, Y. 2007, \apj, 665, 1194
2118: \bibitem[]{}
2119: Kazes, I. \& Crutcher, R. M. 1986, \aap, 164, 328
2120: \bibitem[]{}
2121: Kenyon, S. J. \& Hartmann, L. 1995, ApJS, 101, 117
2122: \bibitem[]{}
2123: Kenyon et al. 2008, to appear in Handbook of Star Forming Regions, ASP 
2124: Conference Series, ed. B. Reipurth
2125: \bibitem[]{}
2126:   Krasnopolsky, R. \& Gammie, C. F. 2005, \apj, 635, 1126
2127: \bibitem[]{}
2128: Krumholz, M. R. \& Tan, J. C. 2007, \apj, 654, 304
2129: \bibitem[]{}
2130:         Krumhotz, M. \& McKee, C. F. 2005, ApJ, 630, 250
2131: \bibitem[]{}
2132:    Kudoh, T., Basu, S., Ogata, Y., \& Yabe, T. 2007, \mnras, 380, 499
2133: \bibitem[]{}
2134:    Kudoh, T., \& Basu, S. 2003, \apj, 595, 842
2135: \bibitem[]{}
2136:    Kudoh, T., \& Basu, S. 2006, \apj, 642, 270
2137: \bibitem[]{}
2138: Lada, E. A., Bally, J., \& Stark, A. A., 1991, \apj, 368, 432
2139: \bibitem[]{}
2140: Lada, C. J., Muench, A. A>, Rathborne, J., Alves, J.F., Lombardi,
2141:           M. 2008, \apj, 672, 410
2142: \bibitem[]{}
2143: Larson, R. B. 1981, \mnras, 194, 809 
2144: \bibitem[]{}
2145:   Li, P. S., Norman, M. L., Mac Low, M.-M. \& Heitsch, F. \apj, 605, 800
2146: \bibitem[]{}
2147:        Li, Z.-Y. \& Nakamura, F. 2004, \apj, 609, L83  
2148: \bibitem[]{}
2149:        Li, Z.-Y. \& Nakamura, F. 2006, \apj, 640, L187  
2150: \bibitem[]{}
2151:         Mac Low, M.-M., Klessen, R. S., Burkert, A., \& Smith, M. D. 
2152: 1998, \prl, 80, 2754
2153: \bibitem[]{}
2154:         Matzner, C. D. 2007, \apj, 659, 1394
2155: \bibitem[]{}
2156:         Matzner, C. D. \& McKee, C. F. 2000, ApJ, 545, 364
2157: \bibitem[]{412}
2158:    McClure-Griffiths, N. M., Dickey, J. M., Gaensler, B. M., Green,
2159:           A. J., \& Haverkorn, M. 2006, \apj, 652, 1339
2160: \bibitem[]{}
2161:         McKee, C. F. 1989, \apj, 345, 782 
2162: \bibitem[]{}
2163:   McKee, C. F. \& Ostriker, E. C. 2007, \araa, 45, 565
2164: \bibitem[]{}
2165: McKee, C. F. \& Zweibel, E. G. 1992, \apj, 399, 551
2166: \bibitem[]{}
2167:       Mouschovias, T.  \& Ciolek, G. 1999,
2168:    in The Origins of Stars and Planetary Systems, ed. C. Lada \&
2169:    N. Kylafis (Kluwer), p. 305
2170: \bibitem[]{}
2171: Muench, A. A. Lada, C. J., Rathborne, J. M. Alves, J. F., \& Lombardi,
2172:           M. 2007, \apj, 671, 1820
2173: \bibitem[]{}
2174: Myers, P. C. \& Lazarian, A. 1998, \apj, 507, 157
2175: \bibitem[]{}
2176:    Nakamura, F. \& Li, Z.-Y. 2005, \apj, 631, 411
2177: \bibitem[]{}
2178:    Nakamura, F. \& Li, Z.-Y. 2007, \apj, 662, 395
2179: \bibitem[]{}
2180: Nakamura, F., Matsumoto, T., Hanawa, T. \& Tomisaka, K. 1999, \apj,
2181:      510, 274 
2182: \bibitem[]{}
2183:    Nakano, T. 1984, Fundam. Cosmic. Phys., 9, 139
2184: \bibitem[]{}
2185:         Nakano, T. \& Nakamura, T. 1978, PASJ, 30, 681
2186: \bibitem[]{}
2187: Novak, G., Dotson, J. L. \& Li, H. 2007, \apj, in print (arXiv:0707.2818)
2188: \bibitem[]{}
2189:     Onishi, T., Mizuno, A., Kawamura, A., Ogawa, H. \& Fukui, Y. 
2190:        1996, \apj, 465, 815
2191: \bibitem[]{}
2192:     Onishi, T., Mizuno, A., Kawamura, A., Tachihara, K. \& Fukui, Y. 
2193:        2002, \apj, 575, 950
2194: \bibitem[]{}
2195:         Ostriker, E. C., Gammie, C. F. \& Stone, J. M. 2001, \apj, 546, 980
2196: \bibitem[]{}
2197:        Padoan, P. \& Nordlund, A. 2002, \apj, 576, 870
2198: \bibitem[]{}
2199: Palla, F. \& Stahler, S. W. 2000, \apj, 540, 255
2200: \bibitem[]{}
2201: Palla, F. \& Stahler, S. W. 2002, \apj, 581, 1194
2202: \bibitem[]{}
2203:        Price, D. \& Bate, M. R. 2008, \mnras, in press
2204: \bibitem[]{}
2205:         Shu, F. H. 1991, Physics of Astrophysics, Vol. 2: Gas Dynamics
2206:           (Mill Valley: University Science Books)
2207: \bibitem[]{}
2208:         Shu, F. H., Adams, F. C., \& Lizano, S. 1987, \araa, 25, 23
2209: \bibitem[]{}
2210:         Shu, F. H. \& Li, Z.-Y. 1997, \apj, 475, 251
2211: \bibitem[]{}
2212: Shu, F. H., Galli, D., Lizano, S., Glassgold, A. E. \& Diamond,
2213: P. H. 2007, \apj, 665, 535
2214: \bibitem[]{}
2215: Tamura, M., Nagata, T., Sato, S. \& Tanaka, M. 1987, \mnras, 224, 413
2216: \bibitem[]{}
2217:         Tilley, D. \& Pudritz, R. 2007, \mnras, 382, 73
2218: \bibitem[]{}
2219: Troland \& Crutcher 2008 (astro-ph/arXiv:0802.2253v1)
2220: \bibitem[]{}
2221: van Dishoeck, E. F. \& Black,J. H. 1988 , \apj, 334, 771
2222: \bibitem[]{}
2223: Williams, J. P., de Geus, E. J., \& Blitz, L. 1994, \apj, 428, 693
2224: \bibitem[]{}
2225: Wolleben, M. \& Reich, W. 2004, \aap, 427, 537
2226: \end{thebibliography} 
2227: 
2228: 
2229: 
2230: \begin{deluxetable}{lllll}
2231: %\tabletypesize{\scriptsize}
2232: %\rotate
2233: \tablecolumns{4}
2234: \tablecaption{Model Parameters}
2235: \tablewidth{\columnwidth}
2236: \tablehead{
2237:  \colhead{Model}     & \colhead{${\cal M}$} & \colhead{$\eta$} 
2238: & \colhead{AD}  & \colhead{Note} 
2239: }
2240: \startdata
2241: S0   & 3 & 0.75 & yes & standard model\\
2242: S1   & 3 & 0.75 & yes & different realization of initial velocity field\\
2243: S2   & 3 & 0.25 & yes & stronger spherical outflow component\\
2244: M0   & 10 & 0.75 & yes   & stronger initial turbulence \\
2245: I0   & 3 & N/A & no   & ideal MHD counterpart of standard model\\
2246: \enddata
2247: \tablecomments{
2248: All models have the same initial distributions of mass and magnetic 
2249: field. The field strength is given by the dimensionless ratio of 
2250: magnetic to thermal pressure $\alpha=24$, corresponding to a 
2251: mildly magnetically subcritical cloud of flux-to-mass ratio 
2252: $\Gamma_0=1.1$. The models differ in initial turbulent Mach number 
2253: ${\cal M}$, jet momentum fraction $\eta$, random realization 
2254: of the initial velocity field, and whether ambipolar diffusion 
2255: (AD) is included or not. 
2256: }
2257: \label{tab:model parameter}
2258: \end{deluxetable}
2259: 
2260: \clearpage
2261: \begin{figure}
2262: \plotone{f1.eps}
2263: \caption{Star formation efficiency (SFE) in the standard model as a 
2264: function of time in units of the collapse time of the condensed 
2265: sheet $t_s$. The slope of the curve indicates that less than 
2266: $1~\%$ of the cloud mass is converted into stars per local 
2267: free-fall time of the condensed sheet. }  
2268: \label{fig:sfe}
2269: \end{figure}
2270: 
2271: 
2272: \begin{figure}
2273: \plotone{f2.eps}
2274: \caption{An edge-on view of the condensed sheet at a representative
2275: time $t=11~t_s$, showing the column
2276: density distribution along the $y-$axis, which is perpendicular to
2277: the direction of initial magnetic field. The condensation is surrounded
2278: by a diffuse halo, with some spurs streaking away perpendicular to
2279: the sheet. The length is in units of $L_J$ (the Jeans length of
2280: the initial uniform state), and column density in the color-bar
2281: in units of $\rho_0 L_J$. The three lowest contours have values of 
2282: 0.05, 0.15, and 0.5, respectively, and are not shown in the
2283: color-bar. They are plotted to highlight the structure of the
2284: diffuse gas.} 
2285: \label{colden_y}
2286: \end{figure}
2287: 
2288: \begin{figure}
2289: \plotone{f3.eps}
2290: \caption{Face-on view of the same condensed sheet as in
2291:   Fig.~\ref{colden_y}, showing the column
2292: density distribution along the $x-$axis, the direction of initial 
2293: magnetic field. Superposed are stellar positions (crosses) and 
2294: contours of critical average mass-to-flux ratio (${\bar \lambda}
2295: =1$, see equation~[\ref{AverageLambda}]). The contours divide the 
2296: magnetically supercritical material capable of star formation from 
2297: the ambient subcritical material sterile to star formation.}  
2298: \label{colden_critical}
2299: \end{figure}
2300: 
2301: \begin{figure}
2302: \plotone{f4.eps}
2303: \caption{Mass fraction of magnetically supercritical material
2304: created in the initially subcritical cloud of the standard model
2305: through ambipolar diffusion. The mass fraction of supercritical 
2306: gas increases gradually initially, reaching a plateau at late 
2307: times. 
2308: }  
2309: \label{fig:supercritical}
2310: \end{figure}
2311: 
2312: 
2313: 
2314: \begin{figure}
2315: \plotone{f5.eps}
2316: \caption{The PDF of the face-on column density of the standard model 
2317: shown in Fig.~\ref{colden_critical} ({\it thick solid line}), with a 
2318: fitted lognormal distribution ({\it dashed-dotted line}) superposed.
2319: The distribution is normalized so that the peak value is close to 
2320: unity, and the column density is in units of $\rho_0 L_J$. Also
2321: plotted for comparison is the column density PDF for the ideal MHD
2322: counterpart of the standard model ({\it thin solid line}), which is 
2323: much narrower. The stark contrast between the two PDFs highlights the 
2324: key role of ambipolar diffusion in fragmenting strongly magnetized
2325: clouds.}  
2326: \label{colden_PDF}
2327: \end{figure}
2328: 
2329: \begin{figure}
2330: %\epsscale{0.6}
2331: \plotone{f6.eps}
2332: \caption{Distribution of mass as a function of density (in units 
2333: of initial density $\rho_0$, {\it thick solid line}), along with 
2334: that for the ideal MHD case ({\it thin solid line}) and 
2335: for a Spitzer sheet containing the same amount of mass ({\it 
2336: dashed line}). The mass distribution is much broader in the 
2337: presence of ambipolar diffusion, because of (magnetically-diluted) 
2338: gravitational fragmentation.}  
2339: \label{lognorm_density}
2340: \end{figure}
2341: 
2342: \begin{figure}
2343: %\plotone{rms_vel.eps}
2344: %\epsscale{0.6}
2345: \plotone{f7.eps}
2346: \caption{Distribution of mass-weighted rms velocity (in units of 
2347: sound speed $c_s$) as a function of density (in units of $\rho_0$),
2348: showing that the diffuse halo is much more turbulent than the 
2349: condensed sheet.}  
2350: \label{rms_vel}
2351: \end{figure}
2352: 
2353: \begin{figure}
2354: %\epsscale{0.6}
2355: \plotone{f8.eps}
2356: \caption{Distribution of the average Alfv\'en Mach number $M_A$ ({\it
2357:     solid line}) and Alfv\'en speed (in units of sound speed, {\it 
2358: dashed line}) as a function of density (in units of $\rho_0$), showing
2359: that the motions of the bulk condensed sheet materials are roughly
2360: Alfv\'enic whereas those of the halo materials are sub-Alfv\'enic.} 
2361: \label{ave_MA}
2362: \end{figure}
2363: 
2364: \begin{figure}
2365: %\epsscale{0.6}
2366: \plotone{f9.eps}
2367: \caption{Distribution of average magnetic field strength (in units 
2368: of the initial $B_0$, see eq.~[\ref{fieldstrength}]) as a function 
2369: of density (in units of $\rho_0$).} 
2370: \label{ave_B}
2371: \end{figure}
2372: 
2373: \begin{figure}
2374: %\epsscale{1.0}
2375: \plotone{f10.eps}
2376: \caption{Map of the logarithm of the density (in units of $\rho_0$) in 
2377: the ``midplane'' of the condensed sheet. Overlaid are contours of 
2378: gravitational potential (in units of sound speed squared) and velocity 
2379: vectors (with a unit vector of length equal to sound speed shown 
2380: above the panel for comparison). } 
2381: \label{den_pot_vel}
2382: \end{figure}
2383: 
2384: \begin{figure}
2385: \plotone{f11.eps}
2386: \caption{Star formation efficiencies as a function of time for two 
2387: models with initial turbulent Mach number ${\cal M}=10$ and $3$ 
2388: (standard model), respectively. Stars form earlier in the initially 
2389: more turbulent cloud, illustrating acceleration of star formation 
2390: by turbulence in strongly magnetized clouds.} 
2391: \label{sfetwo}
2392: \end{figure}
2393: 
2394: \begin{figure}
2395: \plotone{f12.eps}
2396: \caption{Same as Fig.~\ref{colden_y} but for an initial turbulent 
2397: Mach number ${\cal M}=10$, showing that the initially more 
2398: turbulent model has an apparently thicker condensed sheet than 
2399: the standard model.} 
2400: \label{colden_y_M10}
2401: \end{figure}
2402: 
2403: \begin{figure}
2404: \plotone{f13.eps}
2405: \caption{Map of the logarithm of the density (in units of $\rho_0$) 
2406: in an $xz$ plane that passes through the minimum of gravitational 
2407: potential. Superposed are unit vectors for magnetic field directions. 
2408: Note the prominent warp at $z\sim 0.3$. Superposition of such warps
2409: along the line of sight can make the condensed sheet appear thicker
2410: than its intrinsic thickness, as illustrated in Fig.~\ref{colden_y_M10}.  
2411: }
2412: \label{den_Bfield_M10}
2413: \end{figure}
2414: 
2415: \begin{figure}
2416: \plotone{f14.eps}
2417: \caption{Star formation efficiencies as a function of time for two 
2418: models with outflow momentum dominated by, respectively, the 
2419: bipolar jet ({\it upper curve}, standard model) and spherical component 
2420: ({\it lower curve}), showing that the spherical component of outflow
2421: is more efficient in slowing down star formation in flattened 
2422: condensations of strongly magnetized clouds. }
2423: \label{sfedependm3}
2424: \end{figure}
2425: 
2426: 
2427: \begin{figure}
2428: \plotone{f15.eps}
2429: \caption{Distribution of axis ratios of dense cores formed after the
2430:   first star was created. The abscissa and ordinate are the axis 
2431: ratios of $\xi=b/a$ and $\eta=c/a$, respectively. A large symbol 
2432: indicates a core with a mass greater than $1.5 \left<M_c\right>$, 
2433: while a small symbol indicates a core with a smaller mass. We divide 
2434: the $\xi$-$\eta$ plane into three parts: a prolate group which lies 
2435: above the line connecting $(\xi, \eta)=(1.0, 1.0)$ and (0.33, 0.0), 
2436: an oblate group which lies below the line connecting $(\xi,
2437: \eta)=(1.0, 1.0)$ and (0.67, 0.0), and a triaxial group which is 
2438: everything else. For comparison we draw the dotted line that connects 
2439: $(\xi, \eta)=(1.0,1.0)$ and (0.5, 0.0). 
2440: }
2441: \label{fig:coreshapenew}
2442: \end{figure}
2443: 
2444: \begin{figure}
2445: \plotone{f16.eps}
2446: \caption{The relationship between the sum of the two surface terms, 
2447: $S+F$, and the gravitational term, $W$, in the virial equation. 
2448: They are normalized to the sum of the internal terms, $U+K+M$.
2449: The solid line indicates the virial equilibrium, $U+K+W+S+M+F=0$.
2450: The large and small symbols are the same 
2451: as those of Figure \ref{fig:coreshapenew}.
2452: For the cores that lie below the line, the left hand side 
2453: of the virial equation (\ref{eq:virial}) is negative and 
2454: thus expected to be bound. 
2455: All others that lie above the line are unbound and expected to 
2456: disperse away, if they do not gain more mass through accretion
2457:  and/or merging with other cores, or reduce internal 
2458: support through turbulence dissipation and/or magnetic flux 
2459: reduction through ambipolar diffusion. 
2460: }  
2461: \label{fig:virialparamnew}
2462: \end{figure}
2463: 
2464: \clearpage
2465: \begin{figure}
2466: %\epsscale{0.6}
2467: \plotone{f17.eps}
2468: \caption{The virial parameter as a function of the core mass.
2469: Large, intermediate, and small symbols denote, respectively, 
2470: cores with $\Gamma_c \le 0.5$, $0.5 < \Gamma_c \le 0.7$, 
2471: and $\Gamma_c > 0.7$, where $\Gamma_c$ is the flux-to-mass ratio 
2472: of a core. The virial parameter can be fitted by a power law of
2473: $\alpha_{\rm vir}\propto M_c^{-2/3}$, indicated by the solid line.
2474: }  
2475: \label{fig:virialrationew}
2476: \end{figure}
2477: 
2478: \begin{figure}
2479: %\epsscale{0.6}
2480: \plotone{f18.eps}
2481: \caption{
2482: The nonthermal velocity dispersion as a function of core radius.
2483: The velocity dispersion is normalized to the sound speed $c_s$,
2484: and radius to the Jeans length of the Spitzer sheet $L_{s}$.
2485: The large and small symbols are the same as those of 
2486: Figure \ref{fig:coreshapenew}.
2487: There is apparently no correlation between the velocity 
2488: dispersion and the radius, in good agreement with 
2489: the observed linewidth-size relation of the dense cores in 
2490: the Taurus molecular clouds (Onishi et al. 2002). 
2491: } 
2492: \label{fig:velocitysizenew}
2493: \end{figure}
2494: 
2495: \begin{figure}
2496: %\epsscale{0.6}
2497: \plotone{f19.eps}
2498: \caption{Distribution of specific angular momentum as a function of core 
2499: radius. The specific angular momentum is normalized to $c_s L_s = 1.1
2500: \times 10^{22}$ cm$^{2}$ s$^{-1}$ (for our fiducial cloud parameters),
2501: and radius to the Jeans length of the Spitzer sheet $L_s$. 
2502: The large and small symbols are the same as those of Figure 
2503: \ref{fig:coreshapenew}. For comparison, the specific angular momenta 
2504: of 7 starless N$_2$H$^+$ cores are plotted as asterisks (data from Table 5 
2505: of Caselli et al. 2002). 
2506: }  
2507: \label{am1}
2508: \end{figure}
2509: 
2510: \begin{figure}
2511: %\epsscale{0.6}
2512: \plotone{f20.eps}
2513: \caption{The magnetic flux-to-mass ratio as a function of the column
2514:  density of the core.
2515: The large and small symbols are the same 
2516: as those of Figure \ref{fig:coreshapenew}.
2517: The cores below the solid line $\Gamma_c=1$ are 
2518: magnetically supercritical.
2519: The flux-to-mass ratio tends to be smaller for more massive cores.
2520: }  
2521: \label{fig:mass2fluxnew}
2522: \end{figure}
2523: 
2524: \clearpage
2525: \begin{figure}
2526: %\epsscale{0.4}
2527: \plotone{f21.eps}
2528: \caption{Core mass spectrum. The core mass is normalized to 
2529: the Jeans mass of the Spitzer sheet, $M_{J,s}$.
2530: The solid histogram indicates the mass spectrum of dense cores.
2531: For comparison, we show the stellar mass spectrum
2532: with the dotted histogram.
2533: The solid line denotes the power law $dN/dM \propto M^{-2.35}$, 
2534: the Salpeter IMF.
2535: There is a prominent break around $1 M_{J,s}$ in the core mass
2536:  spectrum. Above the break, the spectrum is in good agreement 
2537:  with the Salpeter IMF. 
2538: The stellar mass spectrum can be also fitted by the Salpeter IMF.
2539: }  
2540: \label{fig:coremf}
2541: \end{figure}
2542: 
2543: \begin{figure}
2544: %\epsscale{0.65}
2545: \plotone{f22.eps}
2546: \caption{3D view of density distribution and field lines of the  
2547: standard model at the same time as in
2548: Figs.~\ref{colden_y} and \ref{colden_critical}. The isodensity surfaces 
2549: have values $\rho=0.5$
2550: ({\it blue}), 15 ({\it green}) and 60~$\rho_0$ ({\it red}). They 
2551: represent three  distinct cloud components:
2552: turbulent diffuse halo,  fragmented condensed sheet, and dense cores.  
2553: The sheet is nearly magnetically
2554: critical,  whereas the halo and cores are significantly subcritical and 
2555: supercritical, respectively.}
2556: \label{3D}
2557: \end{figure}
2558: 
2559: 
2560: 
2561: \end{document}
2562: 
2563: 
2564: 
2565: 
2566: