0806.2482/ms.tex
1: %\batchmode
2: %%\documentclass[12pt,preprint]{aastex}
3: %\documentclass[11pt,preprint]{aastex}
4: \documentclass[narrow]{elsart1p}
5: %%\documentclass[manuscript]{aastex}
6: %%\documentclass[10pt,preprint2]{aastex}
7: %%\documentstyle[emulateapj]{article}
8: 
9: %TO DO COMMENTS
10: %calculate fractional mass above twice initial value
11: %quote typical value of total mass in system?
12: %check masses in pext=10 run
13: %Oct 31: changed 55.9 to 31.4 based on hand measurement of 6 new 512^2 runs of model 7
14: %discuss pressure driven mode eventually leading to grav driven instability
15: 
16: \usepackage{natbib}
17: \usepackage{epsfig}
18: \usepackage{amssymb}
19: \journal{New Astronomy}
20: 
21: \newcommand{\ga}{\gtrsim}
22: \newcommand{\la}{\lesssim}
23: \newcommand{\bl}[1]{\mbox{\boldmath$ #1 $}}
24: %\newcommand{\vnp}{\bl{v}_{n,p}}
25: %\newcommand{\vip}{\bl{v}_{i,p}}
26: \newcommand{\vnp}{\bl{v}_{\rm n}}
27: \newcommand{\vip}{\bl{v}_{\rm i}}
28: \newcommand{\cs}{c_{\rm s}}
29: \newcommand{\csq}{c_{\rm s}^2}
30: \newcommand{\ceffsq}{C^2_{\rm eff}}
31: \newcommand{\zhat}{\mbox{\boldmath$ \hat{z}$}}
32: \newcommand{\rhat}{\mbox{\boldmath$ \hat{r}$}}
33: \newcommand{\phihat}{\mbox{\boldmath$ \hat{\phi}$}}
34: \newcommand{\beq}{\begin{equation}}
35: \newcommand{\eeq}{\end{equation}}
36: \newcommand{\barray}{\begin{eqnarray}}
37: \newcommand{\earray}{\end{eqnarray}}
38: \newcommand{\Mdot}{\dot{M}}
39: \newcommand{\Msun}{M_{\odot}}
40: \newcommand{\ul}{\underline{\hspace{20pt}}}
41: \newcommand{\bfig}{\begin{figure}}
42: \newcommand{\efig}{\end{figure}}
43: \newcommand{\FT}{{\cal F}}
44: \newcommand{\FTinv}{{\cal F}^{-1}}
45: 
46: \newcommand{\few}{\rm few~}
47: \newcommand{\cmc}{~{\rm cm}^{-3}}
48: \newcommand{\cms}{~{\rm cm}^{-2}}
49: \newcommand{\Alf}{Alfv\'en\ }
50: \newcommand{\tff}{t_{\rm ff}}
51: \newcommand{\cm}{~{\rm cm}}
52: \newcommand{\kms}{~{\rm km~s}^{-1}}
53: \newcommand{\s}{~{\rm s}}
54: \newcommand{\pc}{~{\rm pc}}
55: \newcommand{\erg}{~{\rm erg}}
56: \newcommand{\K}{~{\rm K}}
57: \newcommand{\muG}{~\mu{\rm G}}
58: \newcommand{\yr}{~{\rm yr}}
59: \newcommand{\Myr}{~{\rm Myr}}
60: \newcommand{\AU}{~{\rm AU}}
61: \newcommand{\mH}{m_{\rm H}}
62: \newcommand{\Htwo}{{\rm H}_{2}}
63: \newcommand{\mn}{m_{\rm n}}
64: \newcommand{\mi}{m_{\rm i}}
65: \newcommand{\Pext}{P_{\rm ext}}
66: \newcommand{\Pexttil}{\tilde{P}_{\rm ext}}
67: \newcommand{\mui}{\mu_0}
68: \newcommand{\muc}{\mu_{\rm core}}
69: \newcommand{\muenv}{\mu_{\rm env}}
70: \newcommand{\mucrit}{\mu_{\rm crit}}
71: \newcommand{\tni}{\tau_{\rm ni}}
72: \newcommand{\tnii}{\tau_{\rm ni,0}}
73: \newcommand{\tniitil}{\tilde{\tau}_{\rm ni,0}}
74: \newcommand{\rhon}{\rho_{\rm n}}
75: \newcommand{\rhoni}{\rho_{\rm n,0}}
76: \newcommand{\nn}{n_{\rm n}}
77: \newcommand{\nnc}{n_{\rm n,c}}
78: \newcommand{\nni}{n_{\rm n,0}}
79: \newcommand{\nion}{n_{\rm i}}
80: \newcommand{\nii}{n_{\rm i,0}}
81: \newcommand{\sign}{\sigma_{\rm n}}
82: \newcommand{\signc}{\sigma_{\rm n,c}}
83: \newcommand{\signi}{\sigma_{\rm n,0}}
84: \newcommand{\signmax}{\sigma_{\rm n,max}}
85: \newcommand{\Nni}{N_{\rm n,0}}
86: \newcommand{\Beq}{B_{z,\rm eq}}
87: \newcommand{\Beqc}{B_{z,\rm eq,c}}
88: \newcommand{\Beqi}{B_{z,\rm eq,0}}
89: \newcommand{\Bref}{B_{\rm ref}}
90: \newcommand{\Breftil}{\tilde{B}_{\rm ref}}
91: \newcommand{\vn}{v_{\rm n}}
92: \newcommand{\vnx}{v_{{\rm n},x}}
93: \newcommand{\vnmax}{|v_{\rm n}|_{\rm max}}
94: \newcommand{\vimax}{|v_{\rm i}|_{\rm max}}
95: \newcommand{\lamcrit}{\lambda_{\rm T,cr}}
96: \newcommand{\lammax}{\lambda_{\rm T,m}}
97: \newcommand{\lamgm}{\lambda_{\rm g,m}}
98: \newcommand{\mgm}{M_{\rm g,m}}
99: \newcommand{\taugm}{\tau_{\rm g,m}}
100: \newcommand{\kion}{k_{\rm i}}
101: \newcommand{\xion}{x_{\rm i}}
102: \newcommand{\lamavg}{\langle \lambda \rangle}
103: \newcommand{\Mavg}{\langle M \rangle}
104: \newcommand{\savg}{\langle s \rangle}
105: \newcommand{\axisravg}{\langle b/a \rangle}
106: \newcommand{\Zavg}{\langle Z \rangle}
107: \newcommand{\sigin}{\langle \sigma w \rangle_{\rm{i\Htwo}}}
108: \newcommand{\Psim}{\Psi_{\rm M}}
109: 
110: %\newcommand{\cbyaavg}{\langle c \rangle/\langle }
111: 
112: %\slugcomment{to be submitted to NewAst}
113: %\shorttitle{Nonlinear Fragmentation}
114: %\shortauthors{Basu, Ciolek, \& Wurster}
115: 
116: 
117: \begin{document}
118: 
119: \begin{frontmatter}
120: 
121: \title{Nonlinear Evolution of Gravitational Fragmentation Regulated by
122: Magnetic Fields and Ambipolar Diffusion}
123: 
124: 
125: \author[label1]{Shantanu Basu\corauthref{cor}}, 
126: \corauth[cor]{Corresponding author.}
127: \ead{basu@astro.uwo.ca}
128: \author[label2]{Glenn E. Ciolek},
129: \ead{cioleg@rpi.edu}
130: and \author[label1]{James Wurster}
131: 
132: \address[label1]{Department of Physics and Astronomy, University of
133: Western Ontario, London, Ontario N6A~3K7, Canada}
134: \address[label2]{Department of Physics, Applied Physics, and Astronomy,
135: Rensselaer Polytechnic Institute, 110 W. 8th Street, Troy, NY 12180,
136: USA}
137: 
138: \begin{abstract}
139: We present results from an extensive set of simulations of gravitational
140: fragmentation in the presence of magnetic fields and ambipolar diffusion.
141: The thin-sheet approximation is employed, with an ambient magnetic field
142: that is oriented perpendicular to the plane of the sheet.
143: Nonlinear development of fragmentation instability leads to substantial
144: irregular structure and distributions of fragment spacings, fragment masses,
145: shapes, and velocity patterns in model clouds. We study the effect of 
146: dimensionless free parameters that characterize the 
147: initial mass-to-flux ratio, neutral-ion coupling, and external pressure 
148: associated with the sheet.
149: The average fragmentation spacing in the nonlinear phase of evolution is in
150: excellent agreement with the prediction of linear perturbation theory.
151: Both significantly subcritical {\it and} highly supercritical clouds 
152: have average fragmentation scales $\lamavg \approx 2 \pi Z_0$, where   
153: $Z_0$ is the initial half-thickness of the sheet. In contrast, the
154: qualitatively unique transcritical modes can have $\lamavg$ that 
155: is at least several times larger. Conversely, fragmentation dominated by
156: external pressure can yield dense cluster formation with much smaller
157: values of $\lamavg$. The time scale for nonlinear
158: growth and runaway collapse of the first core is $\approx 10$ times 
159: the calculated growth time $\taugm$ of the eigenmode with minimum growth time,
160: when starting from a uniform background state with small-amplitude
161: white-noise perturbations. Subcritical and transcritical models 
162: typically evolve on a significantly longer time scale than the supercritical
163: models.
164: Infall motions in the nonlinear fully-developed contracting cores
165: are subsonic on the core scale in subcritical and transcritical clouds,
166: but are somewhat supersonic in supercritical clouds. 
167: Core mass distributions are sharply peaked with a steep decline to large
168: masses, consistent with the existence of a preferred mass scale for each
169: unique set of dimensionless free parameters. 
170: However, a sum total of results for various 
171: initial mass-to-flux ratios yields a broad distribution reminiscent of observed
172: core mass distributions.
173: Core shapes are mostly near-circular in the plane of the sheet for subcritical
174: clouds, but become progressively more elongated for clouds with increasing
175: initial mass-to-flux ratio. Field lines above the cloud midplane
176: remain closest to vertical in the ambipolar-drift driven core formation
177: in subcritical clouds, and there is increasing amount of magnetic field 
178: curvature for clouds of increasing mass-to-flux ratio. 
179: Based on our results, we conclude that fragmentation spacings, 
180: magnitude of infall motions, core
181: shapes, and, especially, the curvature of magnetic field 
182: morphology, may serve as  
183: indirect observational means of determining a cloud's ambient
184: mass-to-flux ratio.
185: \end{abstract}
186: 
187: \begin{keyword}
188: {ISM: clouds \sep ISM: magnetic fields \sep MHD \sep stars: formation}
189: \end{keyword}
190: 
191: \end{frontmatter}
192: 
193: \section{Introduction}
194: \label{s:intro}
195: \subsection{Molecular Cloud Cores}
196: 
197: Star formation occurs in dense cores within
198: interstellar molecular clouds.
199: The existence and properties of dense cores
200: are well established by studies of molecular spectral line emission 
201: \citep{Myer83b,Bens,Jiji}, submillimeter dust emission 
202: \citep{Ward94,Andr,Kirk}, and infrared absorption
203: \citep{Bacm,Teix,Lada07}. 
204: The core formation process 
205: has been the subject of intense theoretical study for the past few 
206: decades. Ideas range from 
207: uninhibited gravitational fragmentation instability 
208: \citep{Jean,Lars85,Lars03},
209: to ambipolar diffusion in a magnetically supported cloud 
210: \citep{Mest,Mous78,Shu87},
211: to a very rapid fragmentation due to pre-existing turbulent flows
212: \citep{Pado,Kless,Gamm}.
213: The last two ideas are related to scenarios for the nature of relatively
214: low-density molecular cloud envelopes, which contain most of the mass 
215: of molecular clouds \citep[see][]{Gold}. 
216: In one scenario, they are magnetically-dominated
217: and evolve due to ambipolar diffusion, a gravitationally-driven redistribution of mass 
218: amongst magnetic flux tubes \citep{Mous78}. The relatively low
219: global star formation rate and efficiency in Galactic molecular clouds
220: \citep[see][]{McKe} is used to argue
221: that molecular cloud envelopes have a subcritical mass-to-flux ratio 
222: \citep[see e.g.][]{Shu99,Elme07}. 
223: Indirect empirical evidence based on the Chandrasekhar-Fermi (CF) 
224: method and velocity anisotropy measurements
225: does imply that the low density ($n \sim 100 \cmc$) regions are 
226: subcritical \citep{Cort,Heye}. The contrasting view is that
227: the mass-to-flux ratio is supercritical on large scales, and that the
228: star formation rate and efficiency are governed by supersonic turbulence
229: \citep{MacL}. This requires either a continual replenishment of rapidly dissipating
230: turbulence, or a relatively rapid dispersal of the cloud envelopes.
231: The focus of this paper 
232: is the fragmentation of embedded and relatively quiescent 
233: dense regions, and not on the nature of the larger envelopes.
234: %We consider a variety of initial mass-to-flux ratios, ionization
235: %fractions, and external pressures acting on these regions.
236: We study the effect of small-amplitude perturbations
237: on a variety of cloud models with different mass-to-flux ratios, 
238: degrees of magnetic coupling, and external pressures.
239:  
240: %XXX- may have to remove the refs below
241: %\citep{Mous77,Mous78,Mous80,Tomi90,BM94}.
242: %XXX - have to add refs ot McKe and MacL
243: %XXX- do we add refs to Goldsmith, Blitz, others above?
244: 
245: For the purpose of this paper, we refer to any individual unit of star
246: formation, which leads typically to a single or small multiple star system, 
247: as simply a ``core''. However, astronomers often subdivide this concept
248: into two observational categories, that of  
249: ``prestellar core'' and ``prestellar condensation''. 
250: The former term is often used to describe 
251: the extended (mean density $n \sim 10^5 \cmc$, size $s \sim 0.1$ pc,
252: spacing $\lambda \sim 0.25$ pc) objects in regions of distributed
253: star formation like the Taurus molecular cloud, while the latter
254: term typically describes the more compact ($n \ga 10^6-10^7 \cmc$,
255: $s \sim 0.02-0.03$ pc, $\lambda \sim 0.03$ pc) objects that are located 
256: within cluster-forming cores in, for example, Ophiuchus, Serpens,
257: Perseus, and Orion \citep[see discussion in][]{Ward07}. 
258: Both the prestellar cores in e.g., Taurus, and the
259: cluster-forming cores (which harbor multiple condensations) 
260: in the other clouds occupy only a very small fraction of the total 
261: volume and mass of their larger, more turbulent, molecular cloud complex
262: \citep[e.g.,][]{John,Gold}.
263: A compendium of current data from
264: spectral line emission, dust emission, and infrared absorption tends to
265: show that cores exhibit central density concentration
266: \citep{Ward94,Andr}
267: subsonic inward motions \citep{Tafa,Will,Lee,Case} that often extend beyond the 
268: nominal core boundary, near-uniform gas temperatures \citep{Bens}, subsonic 
269: internal turbulence \citep{Myer83a,Full,Good98}, 
270: and near-critical magnetic field strengths, when detected by the Zeeman
271: effect \citep{Crut99,Bour} or inferred by the CF method 
272: \citep{Crut04}.
273: 
274: The measured subsonic infall motions constitute indirect evidence for a force
275: that mediates gravity.
276: The mediative force may be
277: due to the magnetic field, whose strength in dense regions is very close to 
278: the critical value for collapse \citep{Crut99}.
279: %If the molecular cloud envelopes have a subcritical mass-to-flux 
280: If the fragmenting region has a subcritical mass-to-flux 
281: ratio, then the formation of cores will be regulated by
282: ambipolar diffusion, i.e. neutral molecules diffusing past ions that
283: are tied to magnetic fields, and will occur on a time scale that may significantly
284: exceed the local dynamical time for typical ionization fractions. 
285: %Such a gravitationally-driven redistribution of mass amongst magnetic
286: %flux tubes \citep{Mous78} provides an explanation for the inefficiency
287: %of star formation, and certainly allows time for magnetic braking to 
288: %resolve much of the angular momentum problem of star formation 
289: %\citep{Mous77,Mous78,Mous80,Tomi90,BM94}.
290: %In general, theoretical arguments based on the global star formation rate and
291: %efficiency in Galactic molecular clouds have been used to justify 
292: %the assumption of subcritical envelopes \citep[see e.g.][]{Shu99,Elme07}. 
293: %Indirect evidence based on the CF method and velocity anisotropy measurements
294: %does imply that the low density ($n \sim 100 \cmc$) regions are subcritical \citep{Cort,Heye}.
295: Conversely, if the dense gas is supercritical, then the fragmentation takes 
296: place on a dynamical time scale. \citet[hereafter BC04]{BC04} studied the nonlinear
297: development of fragmentation instability in clouds that were initially 
298: either critical or decidedly supercritical. They found that supercritical
299: fragmentation is characterized by somewhat supersonic motions on the 
300: core scale ($\sim 0.1$ pc) while critical fragmentation is characterized by
301: subsonic inward motions on those scales. The nonlinear fragmentation of 
302: decidedly subcritical clouds
303: is also presented in this paper as part of a comprehensive 
304: parameter study of
305: the effects of mass-to-flux ratio, initial ionization fraction, and external pressure.
306: We believe that the three broad categories of subcritical, transcritical, 
307: and supercritical fragmentation should all occur in the interstellar medium,
308: and even within separate regions of a single molecular cloud complex. Our
309: results can help to distinguish which mode of fragmentation is occurring in 
310: a given observed region.
311: 
312: The linear theory of fragmentation of a partially ionized, magnetic thin sheet
313: has been presented by \citet[hereafter CB06]{CB06}. 
314: An important quantity is the dimensionless critical mass-to-flux ratio, 
315: $\mu = 2 \pi G^{1/2} \sigma/B$, where $\sigma$ is the column density of a 
316: sheet and $B$ is the strength of the magnetic field that is oriented 
317: perpendicular to the plane of the sheet. In the limit of flux-freezing,
318: fragmentation can occur only if $\mu > 1$, i.e. the sheet is supercritical. 
319: Conversely, gravitationally-driven fragmentation instability cannot occur at all 
320: if $\mu < 1$, i.e. the sheet is subcritical.
321: CB06 show that the inclusion of ambipolar drift in such a model means that 
322: fragmentation can occur for {\it all mass-to-flux ratios}, but on widely 
323: varying time scales and length scales. 
324: A very important result of CB06
325: is their Fig. 2 (see also Fig. 1 in this paper), which demonstrates that transcritical ($\mu \approx 1$) fragmentation
326: has a preferred scale that can be many times larger
327: than $\lammax$, the wavelength of maximum growth rate in the thermal
328: (i.e. nonmagnetic) limit. The latter
329: is actually the preferred scale for both highly supercritical {\it and} 
330: highly subcritical clouds. This is because gravitational
331: instability in subcritical clouds develops by ambipolar drift of neutrals 
332: past near-stationary
333: magnetic field lines, and occurs on the ambipolar diffusion
334: time scale rather than the dynamical time scale \citep[see also][]{Lang78,Mous78}.
335: The ``resonant'' transcritical modes with preferred scale $\lamgm \gg \lammax$ 
336: occur due to a combination of ambipolar drift and field-line dragging;
337: the latter leads to magnetic restoring forces that can stabilize the
338: perturbations unless they are of the required large size.
339: The time scale for ambipolar-diffusion mediated gravitational instability
340: is also presented extensively in CB06. They showed that highly
341: supercritical clouds undergo gravitational instability on the 
342: %The transcritical modes have
343: %a time scale that is intermediate to the 
344: dynamical time $t_{\rm d} 
345: \approx Z/\cs$, where $Z$ is the half-thickness (effectively the scale height)
346: of the sheet and $\cs$ is the isothermal sound speed.
347: On the other hand,
348: highly subcritical clouds undergo instability on the
349: quasistatic ambipolar diffusion
350: time $\tau_{\rm AD} \approx 10\,Z/\cs$ for typical ionization
351: fraction. Transcritical clouds undergo a hybrid instability on an
352: intermediate time scale -- see also \citet{Zwei} for a 
353: similar result. Also, we note that the inclusion of nonlinear fluctuations 
354: can reduce the ambipolar diffusion time under certain circumstances
355: \citep{Fatu,Zwei02}; we do not model such effects in this paper. 
356:  
357: %This allows us to make detailed comparisons with the results of linear
358: %theory (CB06). 
359: Our nonlinear simulations represent a significant extension of the
360: parameter space of models from that presented by BC04. It similarly 
361: extends the parameter space studied by \cite{Inde}, who presented
362: nonaxisymmetric evolution of an infinitesimally thin subcritical sheet, 
363: including the effects of magnetic tension but ignoring magnetic pressure.
364: Recent fully three-dimensional simulations by \citet{Kudo07}, including 
365: magnetic fields and ambipolar diffusion, have confirmed the basic results
366: of BC04. However, that paper presents three representative models, and an extensive
367: parameter study remains computationally out of reach.
368: Here, in addition to a parameter study, we carry out large numbers of
369: simulations for {\it each} unique set of parameters. This is done 
370: in order to compile statistics on core spacings, mass distributions, and shapes.
371: All simulations end at the time of the runaway central collapse of the first 
372: core, hence we are compiling information about the early phases of star 
373: formation in a molecular cloud. The subsequent history of the molecular
374: cloud, after it has been stirred up by the initial star formation, remains 
375: to be determined.
376: This paper represents the product of a total of over
377: 700 separate simulations for models with 14 distinct sets of dimensionless
378: parameters.
379: 
380: An alternate approach to modeling fragmentation is  
381: to input turbulence directly into the dense thin-sheet
382: model \citep[e.g.,][]{Li04,Naka05}, rather than relegating its presence
383: to an unmodeled envelope. This results in highly supersonic 
384: motions within the dense sheet itself, and more rapid formation of
385: cores than in our models. We do not pursue that approach
386: in the present study but leave it open to future investigation.
387: 
388: \subsection{Relation to Global Cloud Structure}
389: Our simulations represent an intermediate approach between attempting a
390: global model of large-scale molecular
391: cloud structure and of modeling the interior collapse of individual
392: cores. Large-scale models need to account for the overall
393: structure and self-gravity of molecular clouds and cannot
394: be realistically studied using a periodic-box model. 
395: Truly global three-dimensional models are computationally expensive
396: and still rarely attempted. Smoothed particle hydrodynamics (SPH)
397: techniques have been used successfully to model entire
398: cluster-forming regions \citep{Bate,Bonn}, and can explain some
399: important features of star formation like mass segregation, binary
400: fraction, and the initial mass function. These models are not fully global
401: in that they do not include the effect of the molecular cloud envelopes.
402: They also do not include the effect of magnetic
403: fields or feedback from outflows, so they cannot address the observed 
404: low global star formation efficiency (SFE), $1-5$\% \citep{Lada03}. 
405: However, the latest version of such models \citep{Pric} does include 
406: flux-frozen magnetic fields and yields somewhat lower SFE's than
407: the non-magnetic models.
408: Another three-dimensional approach which is fully global and based on
409: finite-difference or SPH techniques is modeling the formation 
410: of a molecular cloud from large-scale supersonic gas flows and the early evolution
411: of the cloud \citep{Vazq06,Vazq07,Heit,Henn}.
412: These models yield an 
413: initially flattened cloud whose subsequent evolution is affected by
414: several instabilities (thermal instability, thin-shell instability,
415: and Kelvin-Helmholtz instability). The overall structure evolves away from
416: a sheet-like configuration, but individual segments may be treated as such.
417: Of the above mentioned studies, only the work of \citet{Henn} includes the 
418: effect of magnetic fields, which may work to suppress some of the
419: instabilities. It is not clear whether this is an important factor in 
420: that model. However, the smaller-scale simulation of
421: \citet{Naka08} does show the formation of a sheet-like structure due to
422: the presence of a dynamically important magnetic field.  
423: %These simulations also do not contain a dynamically important
424: %large-scale magnetic field, which may work to suppress some of the
425: %instabilities.
426: Amongst our models, the cases with high bounding pressure may be
427: most appropriate for the scenario of cloud formation due to colliding flows.
428: Our low external pressure models are more appropriate to the 
429: alternate scenario where the flows that form a molecular cloud 
430: are driven by gravity and/or channeled by a large scale magnetic
431: field, e.g. by the magneto-Jeans instability \citep{Kim} or the 
432: Parker instability \citep{Park}.
433: 
434: Despite the various theoretical emphases on flattened structure, we note
435: that observations of molecular clouds reveal complex morphologies with
436: projected shapes that may not look like globally flattened structures. 
437: Molecular clouds have been described variously 
438: in the literature as stratified objects supported by
439: internally generated turbulence \citep{McKe99}, or as 
440: fractal objects \citep{Elme96} in which the internal
441: pressure is not as relevant. Our local sheet 
442: model may be applicable to the dense subregions of the clouds (where
443: stars actually form in weak or rich clusters) in either
444: scenario. As an example from the first scenario described above, 
445: one-dimensional  
446: %If there is stratification in a cloud due to the interplay of 
447: %gravity and non-thermal or turbulent support, it so happens that 
448: %our model can be reasonably applied in this situation. 
449: global models \citep{Kudo03,Kudo06,Foli} of molecular clouds 
450: reveal that internally-driven turbulence 
451: yields large-amplitude motions in lower-density envelopes, 
452: while retaining transonic motions in embedded dense regions; the
453: fragmentation of the latter may be described by our models. 
454: From an observational point of view, we may apply our models to  
455: dense star forming regions such as L1495 and HCl 2 in the Taurus
456: molecular cloud \citep[see][]{Gold}, the L1688 cluster-forming core
457: in Ophiuchus \citep{Mott}, or better yet to the Pipe Nebula
458: \citep{Muen}, which represents an even earlier stage of evolution,
459: with a large number of relatively
460: quiescent prestellar cores that are found
461: in a dense elongated region (the ``stem'' of the Pipe). 
462: This region may represent the best available laboratory
463: for the study of the early stage of star formation, before 
464: feedback from star formation has significantly modified a
465: cloud's internal sructure and motions. There is
466: also evidence that the stem of the Pipe Nebula is flattened along the direction
467: of the mean magnetic field \citep{Alve}. This property is similar
468: to the better established result for the elongated structures in
469: Taurus \citep{Good90,Gold}.
470: 
471: %Our view is that the supersonic motions are confined to lower density
472: %envelope.
473: %On the smaller scale, 
474: %collapse with B and ion neutral friction includes
475: %many axisymmetric models and New thin disk but nonaxisymmetric models
476: %(VB; ideal MHD only so far). Refer to 3D collapse?
477: %Can't follow much of accretion phase.
478: %
479: % refer to VB papers later, in results section when discussing shapes
480: 
481: %We carry out full parameter study and model
482: %many runs. Our model is the nonaxisymmetric extension of earlier models
483: %but have a more uniform background (not initially peaked) and the ability to study
484: %multiple core formation.  We avoid question of how initial compression got there
485: %in the first place.
486: 
487: %We stick to linear
488: %analysis and stop calculation upon formation of 1st runaway core.
489: %Turbulent fragmentation best dealt with in fully 3D model.
490: %Will deal with nonlinear perturbations in a future paper.
491: %Need to fully analyze the linear perturbation case and compare to
492: %linear analysis before moving on to more complex nonlinear IC's.
493: 
494: %We carry results to nonlinear regime and analyze stochasticity of 
495: %nonlinear runs. Determine the intrinsic width of mass distribution
496: %due to stochasticity of fragmentation process alone.
497: 
498: \section{Physical Model}
499: \label{s:pm}
500: 
501: We consider the evolution of weakly ionized, magnetic
502: interstellar molecular clouds. The clouds are isothermal, 
503: having a temperature $T$. As presented in CB06, 
504: we model a cloud as a planar sheet or layer of infinite extent 
505: in the $x$- and $y$- directions of a Cartesian coordinate
506: system ($x$, $y$, $z$). At each time $t$ the sheet has a local 
507: vertical half-thickness $Z(x,y,t)$. We take our model clouds 
508: to be {\em thin}: by this we mean that for any physical quantity 
509: $f(x,y,z,t)$ the condition $f/\nabla_p f \gg Z$ is always satisfied, 
510: where $\nabla_p \equiv \hat{\bl{x}}\partial/\partial x
511: + \hat{\bl{y}}\partial/\partial y$ is the planar gradient 
512: operator. The magnetic field that threads a cloud has
513: the form
514: \beq
515: \bl{B}(x,y,z,t)= 
516: \left\{ 
517: \begin{array}{l}
518: \Beq(x,y,t)\zhat \hspace{9.3em} \mbox{for $|z| \leq Z(x,y,t)$}, \\
519: %\hspace{15.7em} \mbox{for $|z| \leq Z(x,y,t)$}, \\
520: B_{z}(x,y,z,t)\zhat \\ 
521: + B_{x}(x,y,z,t)\hat{\bl{x}}
522: + B_{y}(x,y,z,t)\hat{\bl{y}} ~~~\mbox{for $|z| > Z(x,y,t)$},  
523: \end{array}
524: \right.
525: \eeq
526: where $\Beq$ is the vertical magnetic field strength in the equatorial
527: plane. For $|z| \rightarrow \infty$, $\bl{B} \rightarrow \Bref \zhat$,
528: where $\Bref$ is a uniform, constant reference magnetic field
529: very far away from the sheet.
530: The magnetic field components above the sheet can be determined from $\Beq(x,y)$
531: at any time using the divergence-free nature
532: of the magnetic field and the current-free approximation above the sheet
533: (see CB06 for details).
534:  
535: Some simplification is obtained by integrating the physical 
536: system of equations governing the evolution (conservation of 
537: mass and momentum, Maxwell's equations, etc.) of a model cloud along 
538: the vertical axis from $z=-Z(x,y)$ to $z=+Z(x,y)$. In doing so, 
539: a ``one-zone approximation" is used, in which the density and 
540: the $x$- and $y$- components of the neutral and ion velocities, 
541: as well as the $x$- and $y$- components of the gravitational field, 
542: are taken to be independent of height within the sheet. 
543: The volume density is calculated from the vertical pressure balance
544: equation
545: \beq
546: \rho \csq = \frac{\pi}{2}G \sign^2 + \Pext + \frac{B_x^2+B_y^2}{8\pi},
547: \eeq 
548: where $\Pext$ is the external pressure on the sheet and $B_x$ and 
549: $B_y$ represent the values at the top surface of the sheet, $z=+Z$.
550: This simplification is commonly referred to as the ``thin-sheet 
551: approximation"; the motivation for and the physical reasonability
552: of it is discussed at length in Section~2 of CB06.
553: It is the nonaxisymmetric extension of the 
554: axisymmetric thin-sheet models used to study ambipolar diffusion
555: and gravitational collapse in magnetic interstellar clouds, as 
556: originally developed by \citet{CM93} and \citet{BM94}. 
557: 
558: \subsection{Basic Equations}
559: \label{s:be}
560: 
561: We solve normalized versions of the magnetic thin-sheet
562: equations as justified in CB06. The unit of velocity is taken
563: to be $\cs$, the column density unit is $\signi$, 
564: and the unit of acceleration is $2 \pi G \signi$, equal to the 
565: magnitude of vertical acceleration above the sheet. Therefore, 
566: the time unit is $t_0 = \cs/2\pi G \signi$, and the length unit is 
567: $L_0= \csq/2 \pi G \signi$. From this system we can also construct 
568: a unit of magnetic field strength, $B_0 = 2 \pi G^{1/2} \signi$. 
569: The unit of mass is $M_0 = \cs^4/(4\pi^2G^2\,\signi)$.
570: Here, $\signi$ is the uniform 
571: neutral column density of the background state, and $G$ is the gravitational
572: constant.
573: With these normalizations, the equations used to determine the
574: evolution of a model cloud are
575: \barray
576: \label{cont}
577:  \frac{{\partial \sign }}{{\partial t}} & = & - \nabla_p  \cdot \left( \sign \, \vnp \right), \\
578: \label{mom}
579:  \frac{\partial}{\partial t}(\sign \vnp)  & = & - \nabla_p \cdot (\sign \vnp \vnp) +  \bl{F}_{\rm T}
580: + \bl{F}_{\rm M} + \sign \bl{g}_p, \\
581: \label{induct}
582:  \frac{{\partial \Beq }}{{\partial t}} & = & - \nabla_p  \cdot \left( \Beq \, \vip \right) , \\
583: \label{Ftherm}
584: \bl{F}_{\rm T} & = & - \ceffsq \nabla_p \sign                   ,\\
585: \label{Fmag}
586: \bl{F}_{\rm M} & = & \Beq \, ( \bl{B}_p - Z\, \nabla_p \Beq ) + {\cal O}(\nabla_p Z), \\
587: \label{vieq}
588: \vip & = & \vnp + \frac{\tniitil}{\sign}\left(\frac{\rhoni}{\rhon}\right)^{\kion} \bl{F}_{\rm M} , \\  
589: \label{ceffeq}
590: \ceffsq & = & \sign^2 \frac{(3 \Pexttil + \sign^2)}{(\Pexttil + \sign^2)^2} , \\
591: \label{rhon}
592: \rhon & = & \frac{1}{4} \left( \sign^2 + \Pexttil + \bl{B}_p^2 \right),\\
593: \label{Zeq}
594: Z & = & \frac{\sign}{2 \rhon}, \\                                
595: \bl{g}_p & = & -\nabla_p \psi  ,\\
596: \label{gravpot}
597: \psi & = & \FTinv \left[ - \FT(\sign)/k_z \right]     ,\\
598: \bl{B}_p & = & -\nabla_p \Psi  ,\\
599: \label{magpot}
600: \Psi & = & \FTinv \left[ \FT(\Beq - \Bref)/k_z \right]  \, .
601: \earray
602: In the above equations, $\sign(x,y) = \int_{-Z}^{+Z}\rhon(x,y)~dz$ is 
603: the column density of neutrals, 
604: %%$\Beq$ is the vertical magnetic field in the equatorial plane, 
605: $\bl{B}_p(x,y) = B_x(x,y)\hat{\bl{x}} + B_y(x,y)\hat{\bl{y}}$ is the 
606: planar magnetic field at the top surface of the sheet, 
607: $\vnp(x,y) = v_x(x,y)\hat{\bl{x}} + v_y(x,y)\hat{\bl{y}}$ is the 
608: velocity of the neutrals in the plane, 
609: $\vip(x,y) = v_{{\rm i},x}(x,y)\hat{\bl{x}} + v_{{\rm i},y}(x,y)\hat{\bl{y}}$ 
610: is the corresponding velocity of the ions, and 
611: the normalized initial mass density (in units of $\signi/L_0$)
612: $\rhoni = \frac{1}{4}(1+\Pexttil)$, where $\Pexttil$ is defined below. 
613: The operator $\nabla_p = \hat{\bl{x}} \, \partial/\partial x + 
614: \hat{\bl{y}} \, \partial/\partial y$ is the gradient in the planar
615: directions within the sheet.
616: The quantities $\psi(x,y)$ and $\Psi(x,y)$ are the 
617: scalar gravitational and magnetic potentials, respectively, also in the plane
618: of the sheet. The vertical wavenumber $k_z = (k_x^2+k_y^2)^{1/2}$ is 
619: a function of wavenumbers $k_x$ and $k_y$ in the plane of the sheet,
620: and the operators $\FT$ and $\FTinv$ represent the forward
621: and inverse Fourier transforms, respectively, which we calculate 
622: numerically using an FFT technique. Terms of order 
623: ${\cal O}(\nabla_p Z)$ in $\bl{F}_{\rm M}$, the magnetic force per
624: unit area, are not 
625: written down for the sake of brevity, but are included in the numerical
626: code; their exact form is given in Sections 2.2 and 2.3 of CB06. All terms 
627: proportional to $\nabla_p Z$ are generally very small.
628: 
629: We also note that the effect of nonzero $\nabla_p Z$ and
630: external pressure $\Pext$ is accounted for in the vertically-integrated 
631: thermal pressure force per unit area, $\bl{F}_{\rm T}$, through the use of 
632: $\ceffsq$. This can be seen by noting that
633: \barray
634: \label{intP}
635: {\bl F}_{\rm T} = \int_{-Z}^{+Z} \nabla_p P~dz 
636: &=& \nabla_p \int_{-Z}^{+Z} P~dz - 2 \Pext \nabla_p Z \nonumber \\ 
637: &=& 2 \, \nabla_p (P Z - \Pext Z)~~,
638: \earray
639: where $P$ is the pressure inside the sheet. In the above expression, 
640: we have used $P=\Pext$ at the upper and lower surfaces of the sheet, 
641: %continuity of pressure across the upper and lower
642: %surfaces of the sheet, 
643: and also that $\nabla_p (+Z) = -\nabla_p (-Z)$. 
644: Using the ideal gas equation for an isothermal gas, $P=\rhon\,\csq$, the 
645: expression for half-thickness (Eq. [\ref{Zeq}]), and the normalized
646: equation for vertical
647: hydrostatic equilibrium (Eq. [\ref{rhon}], where we ignore the relatively small
648: term $\bl{B}_p^2$ for simplicity), it is straightforward to 
649: derive the normalized expression for $\ceffsq$ (Eq. [\ref{ceffeq}]).
650: %Since the dimensionless pressure $P = \rhon$ in our isothermal clouds, 
651: %and using equation (\ref{Zeq}) to substitute for $Z$, and noting that
652: %the equation for vertical hydrostatic equilibrium (\ref{rhon}) relates 
653: %$\rhon$ and $\sign$, by inserting these relations into equation 
654: %(\ref{intP}) one is able to obtain the 
655: 
656: The above equations contain the following dimensionless free parameters:
657: $\Pexttil \equiv 2 \Pext/\pi G\signi^{2}$ is the ratio of the external 
658: pressure acting on the sheet to the vertical self-gravitational stress 
659: of the reference state. The dimensionless neutral-ion collision time
660: of the reference state, $\tniitil \equiv \tnii/t_0$, 
661: expresses the effect of ambipolar diffusion. 
662: In the limit $\tniitil \rightarrow \infty$ there
663: is extremely poor neutral-ion collisional coupling, such that the ions 
664: and magnetic field have no effect on the neutrals. The opposite limit,
665: $\tniitil =0$, corresponds to the neutrals being perfectly coupled
666: to the ions due to frequent collisions, i.e. flux freezing.
667: The neutral-ion collision time of the reference state is
668: \beq
669: \tnii = 1.4 \frac{\mi + m_{{}_{\Htwo}}}{\mi} \frac{1}{\nii \sigin}\;,
670: \label{tni}
671: \eeq
672: where $\mi$ is the ion mass, which we take to be 25 a.m.u.,
673: the mass of the typical atomic ($\rm{Na}^{+}$, $\rm{Mg}^{+}$)
674: and molecular ($\rm{HCO}^{+}$) ion species in clouds, 
675: $\nii$ is the ion number density of the reference state,
676: and $\sigin$ is the neutral-ion collision rate, equal to
677: $1.69 \times 10^{-9}~{\rm{cm}}^{3}~{\rm s}^{-1}$ for 
678: $\Htwo$-${\rm{HCO}}^{+}$ collisions \citep{McDa}. 
679: The factor of 1.4 in Eq. (\ref{tni}) accounts for the fact
680: that the effect of helium is neglected in calculating the slowing-down
681: time of the neutrals by collisions with ions.
682: The parameter $\kion$ is the exponent in the power-law expression that is used 
683: to calculate the ion density $\nion$ as a function of neutral density
684: $\nn$, namely,
685: \beq
686: \label{rhoieq}
687: \nion = {\cal K} \nn^{\kion}~,
688: \eeq
689: where we adopt 
690: $\kion = 1/2$ and ${\cal K}\, (\simeq 10^{-5} {\rm cm}^{-3/2})$ for all models 
691: in this study \citep[e.g.,][]{Elme79, UN80}, but keep in mind that 
692: calculation of the ion chemistry network makes $\kion$ a function of $\nn$
693: \citep{CM98}. Finally, $\Breftil = \Bref/B_0 = \Bref/2 \pi G^{1/2} \signi$ is the 
694: dimensionless magnetic field strength of the reference state. For physical
695: clarity, we use instead the
696: dimensionless mass-to-flux ratio of the background 
697: reference state:
698: \beq
699: \label{muieq}
700: \mui \equiv 2 \pi G^{1/2} \frac{\signi}{\Bref} = 
701: \Breftil^{-1} \; ,
702: \eeq
703: where $(2 \pi G^{1/2})^{-1}$ is the critical mass-to-flux ratio for 
704: gravitational collapse in our adopted thin-sheet geometry (CB06). Models with 
705: $\mui < 1$ ($\Breftil > 1$) are subcritical clouds, 
706: and those with $\mui > 1$ ($\Breftil < 1$) are supercritical.
707: The initial mass-to-flux ratio is also related to the commonly-used plasma 
708: parameter
709: \beq
710: \beta_0 \equiv \frac{\rhoni\, \csq}{(\Bref^2/8\pi)} = \mui^2 \, (1+ \Pexttil).
711: \eeq 
712: 
713: %Expressions relating typical dimensional values of our adopted 
714: %normalizing units to the free parameters and physical quantities in the
715: %reference state of a model cloud (such as initial density, temperature,
716: %etc.) are given in equations (25) - (30) of CB06.  
717: 
718: Typical values of our units are
719: \barray
720: \label{cs}
721: \cs & = & 0.188 \, \left(\frac{T}{10 \K}\right)^{1/2} \kms,  \\
722: \label{t0}
723: t_0 & = & 3.65 \times 10^4 \left(\frac{T}{10\,\K}\right)^{1/2}
724: \left(\frac{10^{22}\,\cms}{\Nni}\right) \yr, \\
725: \label{L0}
726: L_0 & = & 7.02 \times 10^{-3} \left(\frac{T}{10 \K}\right)
727: \left(\frac{10^{22}\,\cms}{\Nni}\right) \pc \\ \nonumber
728: & = & 1.45 \times 10^3  \left(\frac{T}{10 \K}\right)
729: \left(\frac{10^{22}\,\cms}{\Nni}\right) \AU, \\
730: \label{M0}
731: M_0 & = & 9.19 \times 10^{-3} \left(\frac{T}{10\,\K}\right)^2
732: \left(\frac{10^{22}\,\cms}{\Nni}\right) \Msun, \\
733: \label{B0}
734: B_0 & = & 63.1 \left(\frac{\Nni}{10^{22}\,\cms}\right) \muG.
735: \earray
736: Here, we have used $\Nni = \signi/\mn$, where $\mn = 2.33\,\mH$ is the mean
737: molecular mass of a neutral particle for an H$_2$ gas with a 10\% 
738: He abundance by number. 
739: Furthermore, we may calculate the number density of the background state as
740: \beq
741: \nni = 2.31 \times 10^5 
742: \left(\frac{10 \K}{T}\right) 
743: \left(\frac{\Nni}{10^{22}\cms}\right)^2
744: \left(1+\Pexttil\right)\, \cmc.
745: \eeq
746: The dimensional background reference magnetic field strength for a given model is simply 
747: $\Bref = B_0/\mui$.
748: Finally, the ionization fraction ($=\nion/\nn$) in the cloud 
749: may be expressed as
750: \beq
751: \label{ioniz}
752: \xion = {\cal K} \nn^{-1/2} =  3.45 \times 10^{-8} \left(\frac{0.2}{\tniitil}\right) \left(\frac{10^5 \cmc}{\nn}\right)^{1/2} \left(1+\Pexttil\right)^{-1/2}.
753: \eeq
754: 
755: \subsection{Numerical Techniques, Boundary and Initial Conditions}
756: \label{s:techn}
757: 
758: %-describe IDL code, accuracy and run times. test cases in CB06.
759: % Flux freezing run shows no evolution
760: The system of Eqs. (\ref{cont}) - (\ref{magpot}) 
761: are solved numerically in $(x,y)$ coordinates using a multifluid 
762: non-ideal MHD code that was 
763: specifically developed for this purpose (BC04; CB06). 
764: Partial derivatives $\partial/\partial x$ and $\partial/\partial y$ are
765: replaced with their finite-difference equivalents. 
766: Gradients are approximated using three-point central 
767: differences between mesh cells, while advection of mass and magnetic 
768: flux is prescribed by using the monotonic upwind scheme of \citet{vanL}.
769: Evolution of a model is carried out 
770: within a square computational domain of size $L \times L$, spanning the
771: region $-L/2 \leq x \leq L/2$ and $-L/2 \leq y \leq L/2$. Typically, $L$
772: is taken to be several times larger (up to a factor of 4) than the characteristic
773: length scale of maximum gravitational instability $\lamgm$ 
774: (CB06; see, also, Section~\ref{s:results} below). The computational domain is then
775: divided into a set of $N^2$ equally-sized mesh cells, each having an 
776: area $L/N \times L/N$. 
777: Most of our simulations are run with $L=16\pi\,L_0$ and $N=128$, and some
778: have $L=64\pi\,L_0$ and $N=512$, so that the grid size $\Delta x =
779: \Delta y = 0.393\,L_0$ in all cases. The mass resolution is then
780: $\Delta M = 0.154\,M_0$, or $1.42 \times 10^{-3}\,\Msun$ using the standard values in
781: Eq. (\ref{M0}).
782: 
783: The numerical 
784: method of lines \citep{Schi} is employed, i.e. the first-order
785: partial differential equations
786: (\ref{cont}) - (\ref{induct}) are converted into a set of coupled 
787: ordinary differential equations (ODE's) in time, with one ODE for each
788: physical variable at each cell. Hence, the system of ODE's 
789: has the form $d\bl{\cal Y}/dt = \bl{\cal G}(\bl{\cal Y},t)$,
790: where $\bl{\cal Y}$ and $\bl{\cal G}$ are both arrays of size $V N^2$, 
791: $V$ being the number of dependent variables. Time-integration of this 
792: system of ODE's is performed by using an Adams-Bashforth-Moulton 
793: predictor-corrector subroutine \citep{Sham94}. Numerical solution of 
794: Fourier transforms and inverse transforms, necessary to calculate the 
795: gravitational and magnetic potentials $\psi$ and $\Psi$ at each time 
796: step (see Eqs. [\ref{gravpot}] and [\ref{magpot}]), is done by using 
797: fast Fourier transform techniques \citep{Press}. 
798: 
799: Periodic conditions are applied to all physical
800: variables at the boundary of the computational domain.
801: The background reference state of a model cloud is characterized
802: by a uniform column density
803: $\signi$ and magnetic field $\Beqi \zhat = \Bref \zhat$. This means that
804: the gravitational and magnetic forces are each identically
805: zero in the uniform background state. The evolution of a model cloud 
806: is started at time $t=0$ by superposing a set of perturbations 
807: $\delta\sign(x,y)$ that are random white noise with a root-mean-squared (rms)
808: value that is 3\% of $\signi$. To preserve the
809: same local mass-to-flux ratio $\sign/\Beq$ as in the uniform background
810: state, initial magnetic field perturbations 
811: $\delta\Beq = (\delta\sign/\signi)\Bref$ are also introduced.
812: 
813: Detailed tests of the accuracy of this MHD code were described in CB06. 
814: Full code runs were compared to exact linear solutions for the
815: gravitationally unstable modes of thin-sheet magnetic clouds. The 
816: code was found to be in excellent agreement with these solutions.
817: It correctly captured the temporal evolution of a model cloud in
818: the linear regime of collapse, as exemplified by the growth time of the
819: gravitational instability $\tau_{\rm g}$ for a given fragmentation 
820: length scale $\lambda$, for various values of the
821: initial parameters $\mui$, $\tniitil$, and $\Pexttil$.
822: In addition, we have run flux-freezing tests of subcritical models
823: with random perturbations and verified that no spurious gravitational
824: instability occurs in the absence of ambipolar diffusion.
825: 
826: The principal motivation for our modeling clouds as thin sheets is that
827: it significantly reduces the computational complexity of studying star 
828: formation, while still retaining many fundamental physical features 
829: necessary to understanding the dynamics of core formation and collapse 
830: within interstellar clouds. For instance, although model clouds are 
831: thin, they are not infinitesimally so, and we are able to 
832: incorporate both magnetic pressure and magnetic tension 
833: supporting forces (see Eq. [\ref{Fmag}]).
834: Additionally, the vertical integration (along the direction of the 
835: $z$-axis) that is employed to derive the system of governing equations 
836: in the thin-sheet approximation (Eqs. [\ref{cont}] - [\ref{magpot}]) 
837: has the effect of turning the fully three-dimensional gravitational 
838: collapse problem into a computationally more tractable 
839: two-dimensional problem. 
840: %Doing this greatly reduces the computational requirements for
841: %numerically studying core collapse in magnetic clouds, making it
842: %possible to have models with sufficient spatial resolution yet without
843: %a prohibitively large number of equations that need to be
844: %simultaneously integrated in time. 
845: As a result of these computational
846: savings, our numerical code is able to run efficiently on 
847: a single workstation with minimal cpu times. 
848: Depending on the initial parameters, a 
849: full simulation can be completed in as little
850: as an hour or at most a single day. Hence, we are able to quickly 
851: generate a large number of models covering the entire physically 
852: relevant range of our free parameters, and
853: produce a large quantity of models that can be used for statistical 
854: analysis. By contrast, three-dimensional MHD models
855: require a dedicated workstation or a computer cluster, and 
856: their simulation completion times are orders of magnitude greater than 
857: that needed for our thin-sheet models. 
858: Our new code is written in the IDL programming language,
859: which significantly speeds up the processes of code development,
860: debugging, and visualization.
861: 
862: We have also developed software to analyze the masses and shapes of cores
863: arising from our simulations. In any snapshot of the evolution, we first
864: isolate regions with column density (or mass-to-flux ratio in some cases)
865: above some threshold value. Thresholds are usually chosen to be high
866: enough that multiple peaks do not fall within a single contiguous
867: region above the threshold. In cases where this happens,
868: we have the option of manually isolating the cores.
869: The mass of each core is found by adding the masses of each 
870: computational zone in the isolated region above the threshold. 
871: To determine the size and shape of a core we use the MPFITELLIPSE
872: routine written in IDL by C. Markwardt, which returns the best-fit ellipse
873: to the set of zones that constitute each core. The semimajor
874: and semiminor axes of the best-fit ellipse, $a$ and $b$, respectively, are obtained
875: and used to determine the size $s = \sqrt{ab}$ and axis ratio $b/a$
876: of each core. The vertical half-thicknesses $Z$ of the zones
877: within each core are averaged to find a mean half-thickness.
878: The mean separation of fragments are found by locating,
879: for each density peak associated with a core, the nearest peak
880: of a neighboring core. These values are averaged over all cores in a simulation 
881: and over multiple model realizations. The periodic boundary 
882: conditions are also accounted for; we count any possible
883: nearest neighbor that is just across the periodic boundary.
884:  
885: 
886: \section{Results}
887: \label{s:results}
888: 
889: \subsection{Overview}
890: \label{s:over}
891: 
892: The efficiency of our two-dimensional code allows us to run a large number
893: of simulations, with various combinations of the important parameters
894: $\mui$, $\tniitil$, and $\Pexttil$. For each unique set of parameters
895: we are also able to run a multitude of independent model 
896: realizations. Each realization is distinct in specific details,
897: since the evolution is initiated by random (white noise) small-amplitude 
898: perturbations. However, the independent realizations are 
899: statistically similar, and running a large number of models allows
900: us to assess the level of randomness that contributes to 
901: distributions of various calculated quantities. 
902: 
903: Table 1 contains the parameters for each of ten models as well as
904: the predicted minimum growth time $\taugm$, the associated wavelength of 
905: maximum instability $\lamgm$, and the implied fragmentation mass
906: $\mgm \equiv \pi \signi \lamgm^2/4$, all obtained 
907: from the linear perturbation analysis of CB06.
908: The first two quantities are in normalized form, but the masses have
909: been converted to $\Msun$ using the standard values in Eq. (\ref{M0}).
910: Some insight into the numerical values of $\taugm$ and $\lamgm$ in Table 1
911: can be obtained by considering the growth rate and fragmentation
912: scale of the fastest growing mode in the limit of no magnetic
913: field, but finite external pressure. From the results of CB06, we 
914: find
915: \barray
916: \label{taug}
917: \tau_{\rm g}(\Pexttil,\Breftil=0) & = & 2 \,\frac{(1+3\Pexttil)^{1/2}}{(1+\Pexttil)} \frac{L_0}{\cs} = (1+3\Pexttil)^{1/2} \,\frac{Z_0}{\cs}, \\
918: \label{lamg}
919: \lambda_{\rm g}(\Pexttil,\Breftil=0) & = & 4 \pi \frac{(1+3\Pexttil)}{(1+\Pexttil)^2} L_0
920: = 2 \pi \left( \frac{1+3\Pexttil}{1+\Pexttil}\right) Z_0. 
921: \earray
922: In the above equations, we have used the relation
923: \beq
924: Z_0 = \frac{2L_0}{(1+\Pexttil)}.
925: \eeq
926: These results show that in the limit $\Pexttil \rightarrow 0$, the isothermal
927: sheet has effective ``Jeans length'' $\lammax = 4 \pi L_0 = 2 \pi Z_0$,
928: and growth time $\tau_{\rm T,m} = 2L_0/\cs = Z_0/\cs$. The
929: relation of these values to the sheet half-thickness
930: and sound speed is intuitively understandable; the latter is essentially
931: the dynamical time. 
932: The highly supercritical 
933: model 10, which also has low external pressure, does approach these limiting 
934: values of fragmentation scale and growth time.
935: Equations (\ref{taug})-(\ref{lamg}) also bring out the interesting property that the fragmentation
936: length and time scale of the isothermal sheet can be reduced significantly by
937: increases in $\Pexttil$, but that the fragmentation scale converges to 
938: a fixed multiple of $Z_0$ (i.e. $6 \pi Z_0$) as $\Pexttil \rightarrow \infty$.
939: This property was first noted by \citet{Elme78} and studied further by \citet{Lubo}.
940: The significantly subcritical model 1 also has a fragmentation
941: scale $\approx 2 \pi Z_0$, since instability occurs via neutral drift 
942: past near-stationary field lines; however the growth time $\taugm \approx 10\,Z_0/\cs$. 
943: The transcritical models 3 and 4
944: have values of $\taugm$ more similar to the subcritical models than the
945: supercritical ones, although their fragmentation scales $\lamgm$ are large.
946: Models 9 and 10 have have $\Pexttil=10$,
947: resulting in much smaller values of $\taugm$ and $\lamgm$ than corresponding
948: models with $\Pexttil=0.1$, although the dynamically important magnetic field
949: raises the values of both above the nonmagnetic limits.
950: 
951: Our standard simulation box is four times larger in size than
952: $\lammax$ and more than twice $\lamgm$ for most models.
953: Exceptions to this are model 3 and model 7, and these simulations are
954: carried out in larger simulation boxes of quadruple size ($L=64\,\pi L_0,N=512$)
955: when collecting information on core properties.
956: The total mass in the simulation box with 
957: $L=16\,\pi L_0 = 0.353 \,( \Nni /10^{22}\cms)^{-1} (T/10 \K)$ pc is 
958: $M = 2.53 \times 10^3\, M_0 
959: = 23.2\, ( \Nni /10^{22}\cms)^{-1}(T/10 \K)^2 \,\Msun$.
960: The expected fragmentation scales and masses for all models are 
961: significantly greater than the grid length resolution $\Delta x$ and
962: mass resolution $\Delta M$ quoted in Section \ref{s:techn}.
963: 
964: %give total mass in msun
965: %talk about resolution limit too
966: 
967: %comment out below
968: %Values of $\taugm$ and $\lamgm$ can be understood 
969: %naively in some limits by noting that the half-thickness of the
970: %reference state is $Z_0 =2L_0/(1+\Pexttil)$. Therefore, for the highly
971: %supercritical model 6 ($\mui=10, \Pexttil=0.1$), $\taugm \approx Z_0/\cs$
972: %and $\lamgm \approx 2 \pi Z_0$,
973: %while for the highly subcritical model 1 ($\mui=0.5, \Pexttil=0.1$),
974: %$\taugm \approx 10\,Z_0/\cs$ and $\lamgm \approx 2 \pi Z_0$. 
975: %In this context it is worth noting that the transcritical models 3 and 4
976: %have values of $\taugm$ more similar to the subcritical models than the 
977: %supercritical ones, although their fragmentation scales $\lamgm$ are large.
978: %Models 9 and 10 have much smaller values of $Z_0$ since $\Pexttil=10$,
979: %resulting in much smaller values of $\taugm$ and $\lamgm$ than corresponding
980: %models with $\Pexttil=0.1$. 
981: %%However, the exact numbers are not naively 
982: %%predictable since the modes are driven by external pressure more than by self-gravity. 
983: 
984: Table 2 contains model parameters and key quantities at
985: the end of each nonlinear model run.
986: %Each model with a unique set of parameters is run 
987: %about 100 times to generate average values of important physical 
988: %quantities in columns 5-11.
989: %However, models 4 and 7 need to be run in an expanded domain with
990: %$N=512$ rather than the usual $N=128$, due to their large fragmentation
991: %scales; these models are run about 25 times each to generate the averaged
992: %physical values. 
993: All runs end when $\signmax/\signi = 10$. This corresponds to 
994: a volume density enhancement $\rhon/\rhoni \approx 100$ for models with
995: $\Pexttil=0.1$ and $\rhon/\rhoni \approx 10$ for models with $\Pexttil=10$.
996: %notwithstanding a small additional effect of magnetic pinching.
997: For each model, we list representative values of $t_{\rm run}$\footnote{Experimentation
998: with different rms amplitudes of the initial perturbation (so that 
999: $\delta \sign/\signi$ is in the range 1\%-6\%), and variation of the power
1000: spectrum away from white noise but with the fixed standard rms value, reveal
1001: that the values of $t_{\rm run}$ can vary in the range 10\%-20\% from
1002: the values quoted in Table 2.},
1003: the physical time at which $\signmax=10\signi$,
1004: $\vnmax$, the maximum neutral speed in the simulated region at that time,
1005: and $\vimax$, the corresponding quantity for the ions. 
1006: In general, models with relatively large values of 
1007: $\mui, \tniitil, ~{\rm and}~ \Pexttil$ tend
1008: to evolve faster than counterparts with smaller values of these parameters.
1009: For the gravity-dominated models ($\Pexttil=0.1$), increasing values of $\mui$
1010: and/or $\tniitil$ result in greater values of $\vnmax$.
1011: %Faster evolving models also have greater values of $\vnmax$ at the 
1012: %end of the runs. 
1013: The values of $\vimax$ illustrate that the systematic motions
1014: within the nonlinearly developed cores are gravitationally driven, 
1015: so that ions lag behind neutrals somewhat. 
1016: However, the difference between the speeds of the two species 
1017: remains less than $0.1\,\cs$.
1018: We perform an analysis of core properties as described in Section \ref{s:techn},
1019: after defining a core as an enclosed region with $\sign/\signi \geq 2$
1020: at the end of the simulation.
1021: Since each simulation typically yields only a handful of cores, the
1022: models are run a large number of times to generate significant
1023: core statistics. Models 1, 2, 3, and 5 were run 100 times each, while models
1024: 4 and 7, which need to be run on an expanded grid due to very large
1025: fragmentation scales, were run 22 and 6 times, respectively.
1026: Model 8 ran 50 times and models 9 and 10 were run 25 times each.
1027: Several averaged properties of the resulting cores are presented in columns
1028: 8-12 of Table 2. These quantities are 
1029: $\lamavg$, the average distance between cores,
1030: $\Mavg$, the average mass within a core (converted to $\Msun$
1031: using Eq. [\ref{M0}]), $\savg$, the average size
1032: of a core, $\axisravg$, the average 
1033: axis ratio of a core, and $\Zavg$, the average value of the half-thickness 
1034: of a core. 
1035: 
1036: %Since each simulation typically yields only a handful of cores, the
1037: %models are run a large number of times in order to generate significant
1038: %statistics. Models 1, 2, 3, and 5 are run 100 times each, while models
1039: %4 and 7, which need to be run on an expanded grid due to very large
1040: %fragmentation scales, are run for 22 and 6 times respectively.
1041: %Model 8 is run 50 times and models 9 and 10 are run 50 times each.
1042: 
1043: %Masses in dimensional form - explain scaling
1044: 
1045: %For the purpose of analysis in our simulations, a core is defined as a 
1046: %local density peak and surrounding gas that is above a threshold
1047: %explain how thresholding and ellipse fitting is done
1048: %explain how to generate c/a
1049: 
1050: %\clearpage
1051: 
1052: %original masses, before mult. by pi/4
1053: %1.88, 2.50, 5.43, 27.7, 5.25, 1.78, 7.46, 3.68, 0.11, 0.23
1054: \begin{table}
1055: \begin{center}
1056: %\caption{Summary of Parameters and Results of Linear Theory\tablenotemark{$\dagger$}}
1057: %\caption{Summary of Parameters and Results of Linear Theory\tablenotemark{1}}
1058: \caption{Summary of Parameters and Results of Linear Theory}
1059: \end{center}
1060: \label{t:1}
1061: \begin{tabular}{crrrrrr}
1062: \hline \hline
1063: %\tablehead{
1064: %\colhead{Model} &\colhead{$\mu_0$} &\colhead{$\tniitil$} &\colhead{$\Pexttil$} & 
1065: %colhead{$t_{\rm run}$} & \colhead{$\vnmax$} & \colhead{$\lambda$} &
1066: %\colhead{Mass} & \colhead{Size} & \colhead{$b/a$} & \colhead{$Z_{\rm avg}$}
1067: %}
1068: %\startdata
1069: Model &\hspace{1em}$\mu_0$ &$\hspace{1em}\tniitil$ &\hspace{1em}$\Pexttil$ &\hspace{2em}$\taugm$ &\hspace{1em}$\lamgm$ & $\mgm$ 
1070: \\
1071: \hline
1072: 1       &0.5     &0.2    &0.1    &20.2  &14.3 &1.48
1073: \\
1074: 2       &0.8     &0.2    &0.1    &17.6  &16.5 &1.96
1075: \\
1076: 3       &1.0     &0.2    &0.1    &14.3  &24.3 &4.26
1077: \\
1078: 4       &1.1     &0.2    &0.1    &10.8   &54.9 &21.8
1079: \\
1080: 5       &2.0     &0.2    &0.1    &3.21   &23.9 &4.12
1081: \\
1082: 6       &10.0    &0.2    &0.1    &2.11   &13.9  &1.40
1083: \\
1084: 7       &1.0     &0.1    &0.1    &27.5   &28.5  &5.86
1085: \\
1086: 8       &1.0     &0.4    &0.1    &7.93   &20.0 &2.89
1087: \\
1088: 9       &0.5     &0.2    &10.0   &4.87   &3.43 &0.09
1089: \\
1090: 10      &1.0     &0.2    &10.0   &3.33   &4.99 &0.18
1091: 
1092: \\
1093: \hline
1094: %\enddata
1095: \end{tabular}
1096: \\
1097: %\tablenotetext{1}{\mbox{Times and lengths are normalized to $t_0$ and
1098: %$L_0$, respectively. Masses are converted to $\Msun$ using Eq. (\ref{M0}).}}
1099: %% Any table notes must follow the \end{tabular} command.
1100: Times and lengths are normalized to $t_0$ and
1101: $L_0$, respectively. Masses are converted to $\Msun$ using Eq. (\ref{M0}).
1102: \end{table}
1103: %\end{deluxetable}
1104: 
1105: %note original trun's are 203.7, 166.5, 120.9, 88.1, 22.8, 12.4, 261.0, 
1106: %60.9, 44.2, 22.2
1107: \begin{table}
1108: %\begin{deluxetable}{crrrrrrcrrr}[b]
1109: %\tabletypesize{\normalsize}
1110: %\tablewidth{0pt}
1111: %\tablecaption{Summary of Parameters and Main Results for Model Clouds}
1112: \begin{center}
1113: %\caption{Main Results for Model Clouds\tablenotemark{$\dagger$}\tablenotemark{$\ast$}}
1114: \caption{Main Results for Model Clouds}
1115: \end{center}
1116: \label{t:2}
1117: \begin{tabular}{crrrrrrrcrrr}
1118: \hline \hline
1119: %\tablehead{
1120: %\colhead{Model} &\colhead{$\mu_0$} &\colhead{$\tniitil$} &\colhead{$\Pexttil$} & 
1121: %colhead{$t_{\rm run}$} & \colhead{$\vnmax$} & \colhead{$\lambda$} &
1122: %\colhead{Mass} & \colhead{Size} & \colhead{$b/a$} & \colhead{$Z_{\rm avg}$}
1123: %}
1124: %\startdata
1125: Model &\hspace{1em}$\mu_0$ &$\hspace{1em}\tniitil$ &\hspace{1em}$\Pexttil$ &\hspace{2em}$t_{\rm run}$&\hspace{1em}$\vnmax$& $\hspace{1em}\vimax$ &\hspace{3em}$\lamavg$ &\hspace{1.5em}$\Mavg$ &\hspace{1.5em}$\savg$ &\hspace{1.5em}$\axisravg$ &\hspace{1.5em}$\Zavg$ 
1126: \\
1127: \hline
1128: 1       &0.5     &0.2    &0.1    &204  &0.39  & 0.35 &15.0  &\hspace{1.5em}1.26  &1.25  &0.74 
1129: &0.71\\
1130: 2       &0.8     &0.2    &0.1    &167  &0.62  &0.55 &17.7  &\hspace{1.5em}1.56  &1.49  &0.81 
1131: &0.72\\
1132: 3       &1.0     &0.2    &0.1    &121  &0.70   & 0.64 &20.1  &\hspace{1.5em}3.26  &2.09  &0.69
1133: &0.65\\
1134: 4       &1.1     &0.2    &0.1   &88   &0.95  &0.90 &47.1  &\hspace{1.5em}6.62
1135: &3.23  &0.66  &0.75\\
1136: 5       &2.0     &0.2    &0.1    &23   &1.1  &1.0 &19.1  &\hspace{1.5em}3.41  &2.08  &0.57
1137: &0.67\\
1138: 6       &10.0    &0.2    &0.1    &12   &1.2  &1.2 &12.8  &\hspace{1.5em}0.99  &1.04  &0.53
1139: &0.70\\
1140: 7       &1.0     &0.1    &0.1   &261 &0.66  &0.62 &31.4  &\hspace{1.5em}3.29
1141: &2.13  &0.77  &0.78\\
1142: 8       &1.0     &0.4    &0.1    &61  &0.73  &0.63 &18.5  &\hspace{1.5em}2.02  &1.69  &0.67
1143: &0.71\\
1144: 9       &0.5     &0.2    &10.0   &44   &0.70  &0.63 &4.0   &\hspace{1.5em}0.34  &0.60  &0.62
1145: &0.072\\
1146: 10      &1.0     &0.2    &10.0   &22   &0.50  &0.38 &5.3   &\hspace{1.5em}0.46  &0.74  &0.60
1147: &0.074\\
1148: \hline
1149: %\enddata
1150: \end{tabular}
1151: %\tablenotetext{$\dagger$}{\mbox{Times and lengths are normalized to $t_0$ and
1152: %$L_0$, respectively. Speeds are normalized to $\cs$. Masses are converted to $\Msun$ using Eq. (\ref{M0}).}}\\
1153: %\tablenotetext{$\ast$}{\mbox{Core data for models 4 and 7 are compiled from runs with $N = 512, L = 64\pi L_0$.}}
1154: Times and lengths are normalized to $t_0$ and
1155: $L_0$, respectively. Speeds are normalized to $\cs$. 
1156: Masses are converted to $\Msun$ using Eq. (\ref{M0}).
1157: Core data for models 4 and 7 are compiled from runs with $N = 512, L = 64\pi L_0$.
1158: %% Any table notes must follow the \end{tabular} command.
1159: \end{table}
1160: %\end{deluxetable}
1161: 
1162: \bfig
1163: \epsfig{file=f1.eps}
1164: \caption{Preferred fragmentation scales $\lamgm$ from linear stability 
1165: analysis \citep{CB06} compared with
1166: the average spacing of density peaks in the nonlinear simulations.
1167: The upper solid line is the calculated dependence of
1168: the wavelength with maximum growth rate $\lamgm$ versus 
1169: $\mui$ for fixed parameters $\tniitil=0.2$ and $\Pexttil=0.1$.
1170: The lower solid line is the same but for $\Pexttil = 10$. The triangles 
1171: represent the average spacing of density peaks (with peak $\sign \geq 2 
1172: \signi$) tabulated from a large number
1173: of simulations at each of $\mui = 0.5,0.8,1.0,1.1,2.0$, and $10.0$ with
1174: $\tniitil=0.2$ and $\Pexttil=0.1$. The squares represent the same but 
1175: with $\Pexttil = 10$.}
1176: \label{scales}
1177: \efig
1178: 
1179: Fig.~\ref{scales} shows the preferred fragmentation scales $\lamgm$ versus $\mui$ 
1180: from the linear analysis of CB06 for two separate values 
1181: of the dimensionless external pressure ($\Pexttil = 0.1$ on top 
1182: and $\Pexttil = 10$ below). Both lines represent models with
1183: $\tniitil = 0.2$, which is our adopted standard value based on 
1184: typical observationally-inferred ionization levels (see Eq. [\ref{ioniz}]
1185: above and Eq. [29] of CB06).
1186: Overlaid on each solid line are 
1187: average core spacings $\lamavg$ (see Table 2) calculated from  
1188: the nonlinear endpoint of our simulations. Each data point 
1189: represents an average of core spacings for a large number (20 to 100) of 
1190: simulations. 
1191: The lower line contains $\lamavg$ data for up to 25 runs each of 
1192: an additional four models
1193: with $\Pexttil=10$ that do not, for the sake of brevity, have their data 
1194: compiled in Tables 1 and 2.
1195: %For the purpose of this analysis, a 
1196: %core is defined as any isolatable region with $\sign/\signi \geq 2$
1197: %at the end of the simulation, when $\signmax/\signi = 10$.
1198: {\it These results show that the linear theory can be used with
1199: confidence to predict the average fragmentation properties of clouds even
1200: in a fully nonlinear stage of development.} 
1201: The tabulated fragmentation scales are slightly below the predictions
1202: of linear theory in the range $\mui \approx 1-2$, for $\Pexttil=0.1$.
1203: This is due to occasional subfragmentation of the initially large 
1204: (irregularly shaped) fragments as they become decidedly supercritical.
1205: We return to this issue when discussing the models with
1206: $\mui=1.1$.
1207: 
1208: %A notable discrepancy
1209: %occurs for $\mui=1.1, \Pexttil=0.1$ - the nonlinear spacing is somewhat
1210: %smaller than predicted by the linear theory. This is explained by 
1211: %considering nonlinear effects, as explained in \S\ \ref{muieffect}.
1212: 
1213: \subsection{The Effect of Varying $\mui$} 
1214: \label{s:mui}
1215: 
1216: \subsubsection{Time Evolution}
1217: \label{s:timevol}
1218: 
1219: Fig.~\ref{timevol} shows the time evolution of the maximum neutral column 
1220: density $\signmax$ and maximum mass-to-flux ratio $\mu_{\rm max}$,
1221: both normalized to their initial values, for models 1, 3, and 5, 
1222: which have $\mui= 0.5,1.0,$ and 2.0, respectively, and fixed parameters
1223: $\tniitil=0.2$ and $\Pexttil=0.1$. These three models have
1224: $\taugm/t_0=20.2,14.3,$ and 3.2, respectively. The actual time $t_{\rm run}$
1225: to runaway collapse of a core, starting from small-amplitude
1226: white-noise perturbations, is about $7-10$ times $\taugm$ for these
1227: models. Review of Table 1 and Table 2 shows that $t_{\rm run}/\taugm
1228: \approx 7-10$ is a generic feature of all our models. The clouds with
1229: initial critical and subcritical mass-to-flux ratio have a prolonged
1230: period of dormancy, compared to the supercritical models.
1231: This is due to the need for ambipolar diffusion to operate before
1232: collapse sets in for both cases. The dashed lines in Fig.~\ref{timevol} show how 
1233: much the mass-to-flux ratio changes during the evolution.
1234: The initially subcritical
1235: cloud requires a significant increase of $\mu_{\rm max}$
1236: before collapse begins. 
1237: Although $\mu_{\rm max}$ appears to be diverging
1238: at the end of the simulations, it is in fact increasing much more 
1239: slowly than $\signmax$, and will not asymptotically diverge. This can
1240: be seen in previously published work, i.e. in Fig. 2 of each of
1241: \citet{CM94} and \citet{BM94}.
1242: 
1243: 
1244: \bfig
1245: \epsfig{file=f2.eps,width=\linewidth}
1246: \caption{Time evolution of maximum values of surface density and
1247: mass-to-flux ratio in three simulations. The solid lines show the
1248: evolution of the maximum value of surface density in the simulation,
1249: $\signmax/\signi$, versus time $t/t_0$. This is shown for models
1250: 1, 3, and 5, which have initial mass-to-flux ratio values
1251: $\mui=$ 0.5, 1, and 2, respectively. For each model, a dashed line
1252: shows the evolution of the maximum mass-to-flux ratio in the simulation,
1253: $\mu_{\rm max}$, normalized to $\mui$.
1254: }
1255: \label{timevol}
1256: \efig
1257: 
1258: \subsubsection{Column Density and Velocity Structure}
1259: \label{s:cdens}
1260: 
1261: Fig.~\ref{densimgs} shows the column density map and velocity vectors 
1262: of neutrals at the
1263: end of the simulations for parameters $\mui=0.5,0.8,1.0,1.1,
1264: 2.0,~{\rm and}~10.0$, with $\tniitil=0.2$ and $\Pexttil=0.1$.
1265: All simulations end when $\signmax/\signi=10$, but the time at which 
1266: this is reached is different in each model. These times for the various
1267: models (in order of increasing $\mui$) are $t/t_0=203.7,166.5,120.9,88.1,22.8,~{\rm and}~12.4$.
1268: There is a striking variation in the spacings of cores as $\mui$ 
1269: changes. The highly subcritical case $\mui=0.5$ fragments on essentially
1270: the nonmagnetic preferred scale $\lammax = 4 \pi L_0$, since the 
1271: evolution is characterized by diffusion of neutrals past near-stationary
1272: magnetic field lines, i.e. a Jeans-like instability but on a diffusive time scale. 
1273: As predicted by the linear theory (CB06), there is a peak in the 
1274: fragmentation spacing near $\mui=1$. The peak 
1275: occurs at $\mui=1.1$ when $\tniitil=0.2$ and $\Pexttil=0.1$.
1276: The predicted fragmentation scale is $\lamgm=4.2\,\lammax$,
1277: and indeed our simulation with box width $4\,\lammax$ yields only
1278: one core, although it seems to be undergoing a secondary
1279: fragmentation into two pieces.
1280: 
1281: The velocity vectors of the neutral flow are normalized to the same
1282: scale in each frame, and the horizontal or vertical spacing of the
1283: footpoints is equal to $0.5\,\cs$. There is a monotonic increase of
1284: the typical neutral speeds as $\mui$ increases (see values of $\vnmax$
1285: in Table 2). The supercritical models, unlike their subcritical and
1286: critical counterparts, have large-scale flow patterns
1287: with velocities in the approximate range $(0.5-1.0)\,\cs$, and 
1288: maximum speeds associated with the most fully developed cores that
1289: are mildly supersonic at distances $\sim 0.1$ pc from the core centers.
1290: Interestingly, the essentially hydrodynamic model with $\mui=10$
1291: has only somewhat greater systematic speeds than the model with $\mui=2$.
1292: This is because the fragmentation scale is smaller, so that 
1293: each core has a weaker gravitational influence on its surroundings at this stage
1294: of development.
1295: However, recall that the frames are at different physical times 
1296: since the models with smaller values of $\mui$ reach $\signmax/\signi=10$
1297: at progressively later times.
1298: Spatial profiles of $\vnp$ in the vicinity of cores are presented in BC04, for
1299: some models, and we do not present them again in this paper. The trend of values
1300: of maximum neutral speed $\vnmax$ and maximum ion speed $\vimax$ (Table 2) reveal the 
1301: predictions of our models for the observable motions on the core scale. 
1302: The infall motions are gravitationally driven, so that the ion speed lags
1303: the neutral speed in all cases. The overall rms speed in the entire simulation
1304: region is always quite small for both neutrals and ions. The rms speed of
1305: neutrals, $v_{\rm n,rms}$, falls in the range $(0.05-0.21)\,\cs$ for the various
1306: models, and the corresponding quantity $v_{\rm i,rms}$ in the range
1307: $(0.04-0.20)\,\cs$.
1308: 
1309: %but reaches this stage at a much earlier time
1310: %velocity cuts in BC04
1311: 
1312:  
1313: %times are different though, and monotonically changing. Near critical
1314: %case is characterized by large spacing and longish time scale
1315: %shapes
1316: %vels not much bigger for highly supercritical case
1317: %transcritical cases distinct, but evolve slowly
1318: %tie in to large spacings observations and 2-stage fragmentation
1319: 
1320: %New par on 512^2 simulation here
1321: Since the standard box size $L =16\, \pi L_0$ allows only one fragment to 
1322: form initially in the model with $\mui=1.1$ (as seen in Fig. \ref{densimgs}), we ran
1323: another model with four times larger box size but the same resolution, 
1324: so that $L=64\, \pi L_0$ and $N=512$. This allows the formation of multiple
1325: fragments, since the preferred fragmentation scale from the linear theory
1326: is $\lamgm=54.9 L_0$. Fig.~\ref{bigimage} shows the column density and velocity vectors
1327: at the end of one such simulation. Velocity vectors are again normalized such 
1328: that the horizontal or vertical spacing between footpoints equals $0.5\,\cs$.
1329: This larger simulation does show that multiple fragments form with spacings
1330: approximately as predicted by the linear theory, but that there is also
1331: a tendency for cores to subfragment into two density peaks.
1332: An analysis of the result of 25 separate simulations with different 
1333: random realizations of the initial state reveals that the average distance
1334: between density peaks is $47.1 L_0$, which is somewhat smaller than $\lamgm=54.9 L_0$.
1335: This is explained by the occasional presence of secondary density peaks
1336: within the initially formed fragments. 
1337: While most model clouds undergo {\it single-stage fragmentation} into essentially
1338: thermal critical (Jeans-like) fragments of size $\approx \lammax$, the 
1339: transcritical clouds form
1340: first-stage fragments many times larger than $\lammax$, followed by a possible
1341: {\it second-stage fragmentation} when the mass-to-flux ratio of the fragment 
1342: becomes decidedly supercritical due to ambipolar diffusion. 
1343: This second-stage fragmentation may be favored because both the preferred
1344: fragmentation
1345: scale and growth time of gravitational instability drop precipitously as a cloud
1346: makes the transition from transcritical to supercritical (see Figs. 1{\it d} and 2 
1347: of CB06). The initially rather large fragment may itself
1348: be prone to fragmentation because of its irregular shape.
1349: 
1350: %The long-term evolution of such double-peaked cores cannot be determined 
1351: 
1352: %New par on surface plots
1353: Fig.~\ref{surfaceplots} shows an alternate view of the column density, using surface plots
1354: at the end of simulation
1355: runs, with parameters corresponding to those of models 1, 3 and 5. 
1356: These models start from different realizations of the initial state
1357: than those of the models presented in Fig.~\ref{densimgs}. Density peaks occur in 
1358: different locations but represent an equivalent outcome statistically.
1359: Animations of the time evolution of the surface plots are 
1360: available online\footnote{The animations of the models shown in Figs.~\ref{surfaceplots} and \ref{fls}
1361: reveal that they reach the final state with $\sign/\signi=10$
1362: in a time 15-20\% less than that quoted in Table 2. This is because
1363: the initial perturbation is white noise but with the smallest wavelengths
1364: ($\lambda \leq 4$ grid cells) damped out. This results in slightly more power
1365: in the longer wavelengths (for a fixed rms perturbation level), 
1366: including the preferred fragmentation scale $\lamgm$, and a consequent 
1367: quicker development of the favored mode.}.
1368: 
1369: 
1370: \bfig
1371: \vspace{-15ex}
1372: \centering
1373: \begin{tabular}{cc}
1374: \epsfig{file=f3a.eps,width=0.49\linewidth,clip=} &
1375: \epsfig{file=f3b.eps,width=0.49\linewidth,clip=} \\
1376: \epsfig{file=f3c.eps,width=0.49\linewidth,clip=} &
1377: \epsfig{file=f3d.eps,width=0.49\linewidth,clip=} \\
1378: \epsfig{file=f3e.eps,width=0.49\linewidth,clip=} &
1379: \epsfig{file=f3f.eps,width=0.49\linewidth,clip=}
1380: \end{tabular}
1381: \caption{Image and contours of column density $\sign(x,y)/\signi$, 
1382: and velocity vectors of neutrals, for six different models at the time 
1383: that $\signmax/\signi = 10$. All models have $\tniitil=0.2$ and
1384: $\Pexttil=0.1$. Top left: $\mui=0.5$. Top right: $\mui=0.8$. Middle 
1385: left: $\mui=1.0$. Middle right: $\mui=1.1$. Bottom left: $\mui=2.0$. 
1386: Bottom right: $\mui=10.0$. The color table 
1387: %in the image 
1388: is applied to the logarithm of column density and the contour lines 
1389: represent values of $\sign/\signi$ spaced in 
1390: multiplicative increments of $2^{1/2}$, having the values 
1391: [0.7,1.0,1.4,2,2.8,4.0,...].
1392: The horizontal or vertical distance between footpoints of velocity 
1393: vectors corresponds to a speed $0.5 \, \cs$.
1394: We use the normalized spatial coordinates 
1395: $x' = x/\lammax$ and $y' = y/\lammax$, where $\lammax$ is the wavelength of
1396: maximum growth rate in the nonmagnetic limit with $\Pext=0$.
1397: }
1398: \label{densimgs}
1399: \efig
1400: 
1401: \bfig
1402: \epsfig{file=f4.eps,width=\linewidth}
1403: \caption{Column density and velocity vectors as in Fig. \ref{densimgs} but
1404: with model 4 parameters ($\mui=1.1$) that is run with four times larger 
1405: computational region in each direction. The simulation
1406: region consists of $512 \times 512$ zones.
1407: The horizontal or vertical distance between footpoints of velocity 
1408: vectors corresponds to a speed $0.5 \, \cs$.
1409: }
1410: \label{bigimage}
1411: \efig
1412: 
1413: \bfig
1414: \epsfig{file=f5.eps,width=\linewidth}
1415: \caption{Surface plot of column density $\sign(x,y)/\signi$ at the end of 
1416: the simulation for models 1, 3, and 5 with (left to right) $\mui = 0.5,1.0,2.0$. 
1417: Animations of the time evolution for each model are available online.}
1418: \label{surfaceplots}
1419: \efig
1420: 
1421: \subsubsection{Magnetic Field Lines}
1422: \label{s:mfl}
1423: 
1424: For the force-free and current-free region above our model sheet, 
1425: the three-dimensional structure of the magnetic field 
1426: is obtained
1427: by solving Laplace's equation for the scalar magnetic potential
1428: $\Psim(x,y,z)$. The two-dimensional Fourier Transform of the
1429: magnetic potential at any height $z$ above the sheet is related to its
1430: known planar value $\Psi(x,y)=\Psim(x,y,z=0)$ via
1431: \beq
1432: \Psim(x,y,z) = \FTinv \left\{ \FT[\Psi(x,y)] \exp(-k_z\,|z|)\right\}.
1433: \eeq
1434: In the above expression, $\FT$ ($\FTinv$) represents the forward (inverse)
1435: Fourier Transform in a 
1436: two dimensional $(x,y)$ plane for a fixed $z$, $\Psi$ is obtained from
1437: Eq. (\ref{magpot}), and $k_z = (k_x^2 + k_y^2)^{1/2}$.
1438: We use the FFT technique to efficiently calculate $\Psim$ at any
1439: height $z$, which then leads to the various components of the magnetic
1440: field above the sheet via ${\bl B}-\Bref\zhat = -\nabla \Psim$
1441: (see CB06 for a justification of this expression).
1442: %(see CM93 and BM94 for a justification of this expression).
1443: 
1444: Fig.~\ref{fls} shows images of the final state column density for 
1445: two models with 
1446: $\mui=1.0$ and $2.0$, respectively. These are also independent realizations 
1447: and differ in specific details from models presented earlier. 
1448: Overlaid on the column density images are the magnetic field lines extending 
1449: above the sheet. These are three-dimensional images of a sheet plus field
1450: lines above, viewed from an angle of approximately $10^{\circ}$ from 
1451: the direction of the background magnetic field. Clearly, the supercritical
1452: model has more curvature in the field lines, as the contraction proceeds
1453: primarily with field-line dragging. The critical model produces its
1454: cores via a hybrid mode including both neutral-ion slip (ambipolar diffusion)
1455: and field-line dragging, hence the the lesser amount of field line curvature.
1456: The relative amounts of field line curvature in the cloud and within dense
1457: cores are quantified by calculating the quantity $\theta = \tan^{-1} (|B_p|/\Beq)$, 
1458: where $|B_p| = (B_x^2+B_y^2)^{1/2}$ is the magnitude of the planar magnetic
1459: field at any location on the sheet-like cloud. Hence, $\theta$ is the angle that
1460: a field line makes with the vertical direction at any location at the top
1461: or bottom surface of the sheet. To quantify the differences in field line
1462: bending from subcritical to transcritical to supercritical clouds,
1463: we note that models 1, 3, and 5, with $\mui=(0.5,1.0,2.0)$, have average
1464: values $\theta_{\rm av} = (1.7^{\circ}, 8.3^{\circ}, 18^{\circ})$, and maximum
1465: values (probing the most evolved core in each simulation) 
1466: $\theta_{\rm max} = (20^{\circ}, 30^{\circ}, 46^{\circ})$. 
1467: For the model with $\mui=0.5$, $\theta_{\rm max}$ is comparable to 
1468: that presented in Fig. 2 of \citet{Ciol}.
1469: 
1470: Our ability to model fragmentation with varying levels of magnetic
1471: support and neutral-ion slip opens up the possibility of making a 
1472: detailed comparison of observed hourglass morphologies of magnetic field
1473: lines, where measured \citep[e.g.][]{Schl}, with theoretical models so that 
1474: the curvature of field lines may be used as a proxy to measure the ambient magnetic 
1475: field strength. 
1476: 
1477: \bfig
1478: \centering
1479: \begin{tabular}{cc}
1480: %\epsfig{file=mu1test.eps,width=0.4\linewidth,angle=-90,clip=} &
1481: %\epsfig{file=mu2test.eps,width=0.4\linewidth,angle=-90,clip=}
1482: \epsfig{file=f6a.eps,width=0.5\linewidth,clip=} &
1483: \epsfig{file=f6b.eps,width=0.5\linewidth,clip=}
1484: \end{tabular}
1485: \caption{Image of gas column density $\sign(x,y)/\signi$ 
1486: and superposed magnetic field lines for
1487: models 3 and 5, with $\mui=1.0$ (left) and $\mui=2.0$ (right). The magnetic field
1488: lines extend above the sheet, and the image and field lines are seen
1489: from a viewing angle of about 10$^{\circ}$. Animations of the 
1490: evolution of the column density are available online. The field lines
1491: appear in the last frame of the animation.}
1492: %3D image of gas sheet and magnetic field lines above for mu0=1,2. 
1493: %Online movies accompany.}
1494: \label{fls}
1495: \efig
1496: 
1497: \subsubsection{Shapes}
1498: \label{s:shapes}
1499: 
1500: The core identification techniques
1501: described in Section~\ref{s:techn} yield core sizes, shapes, and masses, whose average
1502: values are tabulated in Table 2. Here in Fig.~\ref{shapehist} we present a more detailed 
1503: histogram of core shape distributions for each of models 1 through 6. 
1504: This isolates the effect of mass-to-flux ratio on core shapes. 
1505: The statistics are generated by running each of model 1-3 and 5-6 for 
1506: 100 distinct realizations with the usual box size ($L=16\,\pi L_0,N=128$).
1507: Model 4 is run 22 times due to the larger box size ($L=64\,\pi L_0,N=512$)
1508: necessitated by the large fragmentation scale.
1509: The number of cores generated are 
1510: 547, 367, 187, 126, 272, and 659, for models 1 through 6, respectively. The different 
1511: numbers reflect the varying numbers of cores arising for each set of parameters
1512: (see Fig. \ref{densimgs}) as well as the smaller number of runs for model 4. However, in each
1513: case we have sufficient numbers to make inferences about the differences between
1514: models.
1515: 
1516: Fig. \ref{shapehist} reveals in detail what is apparent by a visual inspection of 
1517: Fig.~\ref{densimgs}. The core shape distribution contains many circular objects
1518: (axis ratio $b/a \approx 1$) for subcritical clouds, but contains
1519: progressively more elongated cores as $\mui$ increases.
1520: All initially supercritical models have a peak axis ratio that is distinctly
1521: non-circular. All objects are also flattened in the vertical direction and usually
1522: have the shortest dimension along the magnetic field (see values of $\Zavg$ in
1523: Table 2). 
1524: The underlying physical explanation is that quasistatic formation
1525: of cores (for $\mui \leq 1$) allows for growth in all directions more
1526: equally, whereas dynamical gravity-dominated formation will strongly 
1527: accentuate the anisotropies ($k_x \neq k_y$) that are present at the outset
1528: \citep[see][]{Miya1,Miya2}.
1529: 
1530: The supercritical models have a mean axis ratio in the sheet
1531: plane $\approx 0.5$, which is in rough agreement with the
1532: observed mean {\it projected} axis ratio of dense cores \citep{Myer91}.
1533: However, a deprojection of the observed axis ratios yields
1534: intrinsic three-dimensional shapes that are inherently triaxial 
1535: \citep{JBD01,JB02,Goodw,Tass},
1536: with mean axis ratios $b/a \approx 0.9$ and $c/a \approx 0.4-0.5$.
1537: Since the direction of the smallest axis ($c$) corresponds to our
1538: preferred direction of flattening ($z$), the deprojected $b/a$ values 
1539: can be compared directly with our models. We find reasonable agreement for
1540: the subcritical models 1 and 2. The deprojected values of $c/a$ can also be
1541: compared with our thin-sheet models, in which effectively $c/a = \Zavg/a = 
1542: \Zavg \sqrt{\axisravg}/\savg$. There is reasonable agreement here again
1543: for the subcritical models 1 and 2, as well as for the highly supercritical model 6,
1544: which have, respectively, $c/a = 0.49, 0.43, ~{\rm and}~ 0.49$.
1545:   
1546: \bfig
1547: \epsfig{file=f7.eps,width=\linewidth,clip=}
1548: \caption{Histograms of axis ratios $b/a$ of best fit ellipses to dense regions 
1549: with $\sign/\signi \geq 2$, measured at the end of simulations with
1550: $\mui =  0.5,0.8,1.0.1.1,2.0,10.0$ as labeled, corresponding to models 
1551: 1 through 6 in Table 2. Each figure is the result of
1552: a compilation of results of a large number of simulations. The bin width is 0.1. }
1553: \label{shapehist}
1554: \efig
1555: 
1556: \subsubsection{Core Mass Distributions}
1557: \label{s:cmd}
1558: 
1559: In Fig. \ref{masshist} we present histograms of core mass distributions generated
1560: from the multiple runs of models 1 through 6, as 
1561: described in Section~\ref{s:shapes}. Each core is defined as an enclosed region
1562: with $\sign/\signi \geq 2$ that is present at the end of the simulation, when
1563: $\signmax/\signi = 10$. For comparison with observations,
1564: we have converted our calculated masses to
1565: $\Msun$ using an assumed background number column density $\Nni =
1566: 10^{22} \cms$ and temperature $T=10$ K (see Eq. [\ref{M0}]).
1567: 
1568: We note that the variation of the peak masses (and the average masses tabulated
1569: in Table 2) from one model to another are in qualitative agreement with 
1570: the predictions of linear theory (Table 1). Furthermore, the $\mui=0.5$ and 
1571: $\mui=10$ models generate very similar core mass distributions that are 
1572: difficult to distinguish. This is not surprising since they have such similar
1573: preferred fragmentation scales. 
1574: 
1575: The striking feature of each of the histograms is the very sharp descent
1576: at masses greater than the peak of the distribution. Gravitational fragmentation
1577: yields a very strong preferred mass scale.
1578: The peak value itself is more ambiguous and can vary according to
1579: the magnetic field strength, the background column density, the cloud temperature,
1580: and the contour level we use to define the core.
1581: In contrast, the slope on the low-mass side is much shallower. This is due to 
1582: the capture of emerging cores at the end of any simulation. Many of those cores
1583: are expected to grow in time and move over to the right by the time their 
1584: peaks undergo runaway collapse and form a star. The distribution may be described as
1585: relatively narrow and lognormal-like, but with a broader tail at the low-mass side
1586: due to the temporal spread of core ages.
1587: 
1588: The steep decline of the of the mass distributions beyond the peak 
1589: ($d\log N/d \log M \approx -5$ is typical) 
1590: in our study is in contrast to that observed for condensations in
1591: cluster forming regions \citep[e.g.][]{Mott}, where $d\log N/d \log M \approx -1.5$ 
1592: at high masses. Gravitational fragmentation under the conditions studied in
1593: this paper and at the time snapshot chosen here yields a very strong preference
1594: for a characteristic mass. We discuss possible mechanisms 
1595: of broadening the mass distribution in Section~\ref{s:disc}. 
1596: %What can account for the difference?
1597: %In the context of this model, we believe that the longer term evolution of the
1598: %molecular cloud that we have not yet calculated may yield significant and
1599: %differential growth of cores that results in a broad tail. A simplified analytic
1600: %model of this type has been presented by \citet{BJ04}. A broad tail at the high-mass
1601: %end would then be due to a temporal effect, and form in a mirror-image 
1602: %fashion to
1603: %%the broad tail at the low-mass end that we do see in our
1604: %simulations. We note that the scenario of accretional growth of cores has
1605: %some elements in common with ideas of 'competitive accretion' of protostars
1606: %themselves \citep{Bonn}. Alternatively, a broader
1607: %high-mass tail of the core mass distribution may be generated directly by
1608: %initial conditions of turbulent flows \citep[e.g.][]{Pado}. We will study the
1609: %latter effect in the context of our simulations in an upcoming paper.
1610: 
1611: \bfig
1612: \epsfig{file=f8.eps,width=\linewidth,clip=}
1613: \caption{Histograms of masses contained within regions with $\sign/\signi \geq 2$,
1614: measured at the end of simulations with $\mui =  0.5,0.8,1.0.1.1,2.0,10.0$ as labeled. 
1615: Each figure is the result of a compilation of results of a large number of 
1616: simulations. The bin width is 0.1. }
1617: \label{masshist}
1618: \efig
1619: 
1620: \subsubsection{Supercritical Cores}
1621: 
1622: For clouds that start with subcritical or transcritical initial conditions,
1623: there is available a more physical definition of a ``core'', i.e. a region that
1624: is significantly supercritical and enclosed within a subcritical common cloud envelope.
1625: Axisymmetric simulations of cores that evolve initially by ambipolar drift
1626: have shown that the contraction becomes very rapid by the time that 
1627: $\mu \approx 2$ in the central region, leaving behind a more slowly evolving
1628: and essentially subcritical envelope \citep[see, e.g.][]{FM93,CM94,BM94}.
1629: %Furthermore, the magnetic field may limit the eventual accretion rate of the
1630: %subcritical envelope onto the protostar(s) that forms from the supercritical 
1631: %region \citep{Shu04}.
1632: 
1633: Fig. \ref{mtfimgs} shows images and contour maps of $\mu(x,y)$ at the 
1634: end of the simulation
1635: for models 1 and 3, respectively. The initially subcritical ($\mui=0.5$) model
1636: 1 has peaks in $\mu(x,y)$ coinciding with the major peaks in $\sigma(x,y)$
1637: (see Fig. \ref{densimgs} upper left). However, note that the density condensations are 
1638: either largely or even entirely subcritical ($\mu < 1$) at this stage.
1639: This image shows that subcritical clouds can have observable density
1640: enhancements which may still be partially or entirely subcritical, because they
1641: are still in the process of ambipolar-drift-driven gravitational instability.
1642: The image and contours for the initially critical ($\mui=1.0$) cloud shows
1643: that the cloud naturally separates into supercritical and subcritical regions,
1644: due to ambipolar diffusion and a fixed total magnetic flux threading the cloud.
1645: The newly created supercritical regions are extended and typically contain more
1646: than one density and mass-to-flux ratio peak within them. In this case, all
1647: density peaks are associated with gas that has $\mu > 1$.
1648: 
1649: The presence of supercritical regions embedded within a common subcritical
1650: envelope allows us to define cores in a more physical way than the previous
1651: definition as regions with $\sign/\signi \geq 2$.
1652: The latter is a somewhat arbitrary designation, as indeed are all observational
1653: definitions of cores. However, the definition of supercritical cores has its own 
1654: ambiguities, as a simple definition of regions with $\mu > 1$ yields extended
1655: regions in model 3 with multiple density peaks. We find that a viable working 
1656: definition is that a core is a region with $\mu > 1.3$. This isolates individual
1657: density peaks in both models, and is consistent with the earlier axisymmetric
1658: findings that a mass-to-flux ratio somewhat {\it above} the critical value
1659: is necessary before rapid collapse and separation from the envelope becomes
1660: apparent. For example, \citet{CM93} found that $\mu > 1.23$ was required for the 
1661: absence of any available axisymmetric equilibrium state, using a similar
1662: value of $\Pexttil$ as we do.
1663: We compile data from 100 runs of each model with distinct random
1664: realizations of the initial states and present the core mass distribution for
1665: each of models 1 and 3 in Fig. \ref{scmasshist}. The conversion to dimensional 
1666: masses is done
1667: in the same manner as for Fig. \ref{masshist}. The resulting distributions 
1668: have a peak mass
1669: that is somewhat smaller than found using the different core definition used 
1670: for Fig. \ref{masshist}. However, these distributions also have a very sharp decline 
1671: at higher masses.
1672: 
1673: 
1674: \bfig
1675: \centering
1676: \begin{tabular}{cc}
1677: \epsfig{file=f9a.eps,width=0.5\linewidth,clip=} &
1678: \epsfig{file=f9b.eps,width=0.5\linewidth,clip=}
1679: \end{tabular}
1680: \caption{Image and contours of $\mu(x,y)$, the mass-to-flux ratio in units of the
1681: critical value for collapse. Regions with $\mu >1$ are displayed with a color table,
1682: while regions with $\mu <1$ are black. The contour lines are spaced in additive 
1683: increments of 0.1. Left: Final snapshot of simulation with $\mui=0.5$. Right:
1684: Final snapshot of simulation with $\mui=1.0$.}
1685: \label{mtfimgs}
1686: \efig
1687: 
1688: \bfig
1689: \epsfig{file=f10.eps,width=\linewidth,clip=}
1690: \caption{Histogram of masses contained within regions that are significantly supercritical,
1691: specifically $\mu > 1.3$, measured at the end of simulations with $\mui =  0.5$ and 
1692: $\mui=1.0$. Each figure is the result of a compilation of results of a large number of
1693: simulations. The bin width is 0.1.}
1694: \label{scmasshist}
1695: \efig
1696: 
1697: \subsection{The Effect of Varying $\tniitil$} 
1698: \label{s:tni}
1699: 
1700: Eq. (\ref{ioniz}) shows that the ionization fraction $\xion$ at a given neutral density
1701: $\nn$ increases (decreases) linearly as $\tniitil$ decreases (increases).
1702: We investigate the effect of decreasing and increasing $\tniitil$
1703: by a factor of two from its standard value 
1704: (which can be accomplished by changing the factor ${\cal K}$ in Eq. [\ref{rhoieq}])
1705: in models 7 and 8, respectively. The other two parameters
1706: are kept fixed at $\mui=1.0$ and $\Pexttil=0.1$. The 
1707: characteristic growth times of instability $\taugm$ scale approximately
1708: $\propto \tniitil^{-1} \propto x_{\rm i,0}$ (see also Table 1), where $x_{\rm i,0}$ is the
1709: ionization fraction at the background number density $\nni$.
1710: Models 7, 3, and 8 have $\tniitil=(0.1,0.2,0.4)$, $\taugm = (27.5,14.3,7.9) \times t_0$, 
1711: and $\lamgm = (28.5, 24.3, 20.0) \times L_0$, respectively.
1712: Furthermore, Table 2 reveals that the time to runaway collapse
1713: ($\signmax/\signi \geq 10$) is $\approx 10\, \taugm$ when starting from small-amplitude
1714: white noise perturbations, as generally found in our parameter study.
1715: This means that our high ionization-fraction model 7 has the largest value of
1716: $t_{\rm run}(=261\,t_0)$ in our parameter study. This also leads to the largest
1717: age spread of cores in any of our simulations. This is measured by the fact that a typical
1718: simulation, when run (for compiling statistics) in a large box with $L=64\,\pi,N=512$,
1719: has many cores that are just beginning to emerge when the first core goes into a runaway 
1720: collapse. Hence, our value for $\lamavg$ is calculated with a lower core threshold
1721: $\sign/\signi \geq \sqrt{2}$ for this one model. Our analysis reveals that
1722: the fragmentation 
1723: scales in the nonlinear phase are indeed comparable amongst models 3, 7, and 8, and
1724: in good agreement with the linear theory prediction.
1725: We conclude that the effect of varying ionization (for a fixed $\mui$) within the range studied is 
1726: primarily in the {\em rate} of evolution.
1727: 
1728: Fig.~\ref{tnidensimgs} shows images and contours of the density, as well as
1729: velocity vectors, for models 7 and 8.   
1730: The time to reach runaway collapse is about four times longer
1731: for model 7 than for model 8, consistent with its value of $\tniitil$ 
1732: being four times smaller. Model 8 has $\tniitil = 0.4$, hence poorer 
1733: neutral-ion coupling and therefore reduced magnetic support. This results
1734: in slightly
1735: greater infall speeds, slightly larger number of fragments, and cores 
1736: which are slightly more elongated. These are all consistent with its more
1737: dynamical evolution.
1738: 
1739: %talk about sensitive dependence of ionization
1740: %should redo at least 5 more runs
1741: 
1742: 
1743:  
1744: 
1745: %grinds to a halt if ionization goes up
1746: %dramatic difference in timescales but not in lengthscales
1747: 
1748: 
1749: \bfig
1750: \centering
1751: \begin{tabular}{cc}
1752: \epsfig{file=f11a.eps,width=0.5\linewidth,clip=} &
1753: \epsfig{file=f11b.eps,width=0.5\linewidth,clip=}
1754: \end{tabular}
1755: \caption{Column density and velocity vectors as in Fig. \ref{densimgs}, but for models
1756: with $\tniitil = 0.1$ (left) and $\tniitil=0.4$ (right). Both models have
1757: $\mui=1.0$ and $\Pexttil=0.1$.}
1758: \label{tnidensimgs}
1759: \efig
1760: 
1761: \subsection{The Effect of Varying $\Pexttil$} 
1762: \label{s:pext}
1763: 
1764: Our models with $\Pexttil=10$ may represent the effect of pressured environments 
1765: such as sheets brought together by the presence of shocked gas 
1766: (e.g. stellar winds or supernovae) and being embedded in or adjoining 
1767: an H {\sc II} region. These models could represent an example of 
1768: ``induced'' star formation in a 
1769: manner related but not equivalent to that of an initial turbulent flow 
1770: with high ram pressure. 
1771: 
1772: Models 9 and 10 both develop extreme clustering in comparison to the other models. 
1773: Table 2 shows that the fragmentation scales are about 1/3 to 1/4 of that for the 
1774: corresponding models with the same mass-to-flux ratio but small $\Pexttil$.
1775: The fragments grow initially through a pressure-driven mode and the spacing
1776: is in excellent agreement with the predictions of linear theory (Fig. 
1777: \ref{scales} and Table 1).
1778: However, our results show that the nonlinear instability does develop into a 
1779: gravitationally-driven runaway collapse.
1780: The maximum speeds are still subsonic at the end of our simulation, due to 
1781: the relatively weak gravitational influence of each compact core. As well as
1782: having the smallest fragmentation scales, these models also have the shortest
1783: time scales to runaway collapse. For the fiducial $\Nni = 10^{22} \, \cms$,
1784: models 9 ($\mui=0.5$) and 10 ($\mui=1.0$) have values of $t_{\rm run} =
1785: (1.6 \Myr, 0.81 \Myr)$ and $\lamavg = (5800 \AU, 7700 \AU)$, respectively.
1786: Both sets of numbers are considerably smaller than for the models 1 and 3,
1787: which have corresponding values of $\mui$ but $\Pexttil=0.1$.
1788: Fig.~\ref{pext10densimg} shows the clustering properties of model 10 at the end of the simulation, 
1789: which is very similar to the corresponding image for model 9 (not shown).
1790: This fragmentation model clearly produces a much richer cluster than in the
1791: relatively unpressured environments presented earlier. Velocity vectors are not
1792: shown in this image due to confusion arising from infall onto so many peaks.
1793: A careful inspection of the image reveals a variety of core spacings and
1794: sizes at this stage of evolution. Interestingly, the average core spacing
1795: $\lamavg = 5.3 L_0$ is in very good agreement with the preferred 
1796: wavelength in linear theory, 
1797: $\lamgm = 5.0 L_0$. These length scales are well resolved in our 
1798: simulations. However, the average core mass $\Mavg$ significantly exceeds the 
1799: linear theory value $\mgm$. Only models 9 and 10 show such a large
1800: discrepancy between these values. We attribute it 
1801: to the very small sizes of the cores (see $\savg$ values in Table 2)
1802: in these simulations. 
1803: This means that the cores themselves are barely resolved and the mass
1804: estimates should be taken as approximate values that likely represent
1805: upper limits.
1806: 
1807: \bfig
1808: \epsfig{file=f12.eps,width=\linewidth,clip=}
1809: \caption{Column density as in Fig. \ref{densimgs}, but for a model
1810: with $\Pexttil = 10$. Other parameters of this model are 
1811: $\mui=1.0$ and $\tniitil=0.2$.}
1812: \label{pext10densimg}
1813: \efig
1814: 
1815: 
1816: \section{Discussion}
1817: \label{s:disc}
1818: 
1819: Our simulations of the nonlinear development of gravitational instability
1820: under the influence of magnetic fields and ambipolar diffusion start
1821: from a background state of uniform column density and magnetic field
1822: strength. Small-amplitude white-noise perturbations initiate the evolution and
1823: eventually lead to the nonlinear growth of fragments. Averaging over a large number
1824: of simulations reveals that the average spacing of nonlinearly developed 
1825: cores is essentially that predicted from the preferred fragmentation
1826: scales in linear perturbation theory (CB06). However, the time to
1827: reach fully developed runaway collapse is up to ten times longer than
1828: that of the eigenmode with minimum growth time $\taugm$.
1829: The quantity $\taugm$ itself varies from $\approx Z_0/\cs$ 
1830: (essentially the free-fall time $\approx 1/\sqrt{G\rhoni}$ for unpressured
1831: sheets) for highly supercritical
1832: models to $\approx 10 Z_0/\cs$ for highly subcritical models (for a typical neutral-ion
1833: coupling level).
1834: The times to reach runaway collapse vary widely amongst
1835: models with different mass-to-flux ratios, ionization fractions, and
1836: external pressures. For a cloud with $\Nni=10^{22}\cms$ and $T=10\K$,
1837: the times to reach runaway growth of the
1838: first core ranges from 0.45 Myr to 9.53 Myr (see Table 2).    
1839: Since our simulations start from a flat density background, these times
1840: represent {\it upper limits} to the time that fragmentation might take
1841: for each set of parameters. However, an advantage
1842: of the uniform background density is that it allows for a self-consistent
1843: modeling of the entire core formation process, without questions
1844: about the origin of initially peaked density distributions used in earlier
1845: axisymmetric calculations \citep[e.g.][]{CM93,BM94}. 
1846: 
1847: %Our modeling of the evolution of gravitational fragmentation reveals that the
1848: %transcritical initial states take a substantially longer time to reach runaway
1849: %than do the supercritical models. In essence, they evolve on a time scale more
1850: %similar to the subcritical models. Why do they not evolve on a more intermediate
1851: %timescale? 
1852: %However, the preferred wavelength $\lamgm$ for the transcritical
1853: %modes is much larger than in either the subcritical or supercritical cases, 
1854: %due to the restoring forces associated with field-line dragging that is not
1855: %present for subcritical evolution. This makes the evolution time scale longer 
1856: %than one might anticipate.
1857: 
1858: In a medium with initial nonlinear perturbations,
1859: the time scales for all sets of parameters are indeed likely to be shorter.
1860: However, we believe that our calculated time scales are
1861: relevant if the corresponding dimensional values are obtained from higher 
1862: starting values of column density brought about in certain regions by 
1863: pre-existing (including turbulent) flows.  
1864: For example, the Taurus molecular cloud has an overall background number column 
1865: density $N \approx (1-2) \times 10^{21} \cms$ but also contains embedded
1866: dark clouds with $N \approx 5 \times 10^{21} \cms$, such as HCl 2 and L1495, 
1867: within which there are small clusters of $\sim 10-20$ YSO's and also 
1868: many dense cores \citep[see][]{Gome,Onis,Gold}. 
1869: Our periodic model may be applied to such dark clouds that may themselves
1870: have been brought together by nonlinear flows originating in the larger cloud,
1871: external triggers, or an earlier phase of gravitational fragmentation. 
1872: A second example application of our periodic model may be 
1873: within the L1688 dark cloud in Ophiuchus, which has 
1874: $N \approx 10^{22}\,\cms$ \citep{Mott}. 
1875: The mean spacing of fragments in these
1876: two regions varies \citep{Andr00}, with core edges measured to be at radii
1877: $\la 5000\, \AU$ in L1688 but at $\la 20000 \, \AU$ in the Taurus dark clouds
1878: \citep{Andr00}. While some of the difference in spacing may be due to the 
1879: different background column densities, other important aspects of spatial and
1880: kinematic structure may also arise due to 
1881: different values of $\mui$ and $\Pexttil$ (and $\tniitil$ to a lesser extent), 
1882: as demonstrated in this paper. 
1883: 
1884: %An interesting corollary of above is for transcritical fragmentation
1885: %the timescale may be long. But transcrit frag may also account for clumps
1886: %that form small groups of objects.
1887: Unlike a uniform density three-dimensional medium, our thin sheet actually
1888: has a preferred scale for gravitational fragmentation with a unique value
1889: for any given set of initial dimensionless parameters. Since our simulation
1890: region is always safely larger than this fragmentation scale, it is
1891: unlikely that
1892: the size of our system (in the $x$- and $y$-directions) influences 
1893: the final outcome, as measured by
1894: fragment spacings, time scales to runaway collapse, and core mass distributions, for
1895: example. This is supported by our tests with runs at quadruple the size of
1896: the standard simulations. 
1897: Incidentally, this is not the case for three-dimensional periodic box simulations,
1898: in which the fastest growing mode of gravitational instability is always that
1899: of the box size.
1900: We believe that a stratified medium is also a more physical and realistic 
1901: starting point
1902: than a uniform three-dimensional medium. This is because the formation process 
1903: of molecular 
1904: clouds, or magnetic fields and nonlinear flows within it, will tend to set up compressed 
1905: regions with 
1906: a characteristic scale similar to the half-thickness $Z_0 \approx
1907: \csq/(\pi G \signi)$ of our adopted background reference
1908: sheet geometry. That scale is related to the Jeans scale and effectively
1909: determines the preferred fragmentation scale, as modified by magnetic field
1910: strength, ionization fraction, and external pressure.
1911: 
1912: %-transcritical fragmentation, not as fast as one might think, closer
1913: %to subcritical case, turbulent case may be different.
1914: %refer to older axisymmetric papers. potential for 2nd fragmentation.
1915: 
1916: %-talk about high pext cases - relate to triggered core formation
1917: %-natural bndry in these models; relate to isolated models, eg BE sphere
1918: 
1919: %-CMF is not like IMF at this early time. can be different at later times
1920: %-cores can be subcritical
1921: 
1922: \bfig
1923: \epsfig{file=f13.eps,width=\linewidth,clip=}
1924: \caption{Histogram of masses from models with $\mui=0.5,0.8.1.0,1.1,2.0$ and
1925: $\tniitil=0.2,\Pexttil=0.1$. A weighted average of core masses is taken,
1926: so that an equal simulated area is assigned to each of models $1-5$.
1927: The cores are the same as the ones in the first five panels of
1928: Fig.~\ref{masshist}. The bin width is 0.1.
1929: }
1930: \label{totalmasshist}
1931: \efig
1932: 
1933: 
1934: The core mass distributions that we have compiled from our simulations
1935: are narrowly-peaked and do not match the broader distributions commonly
1936: observed in cluster-forming regions \citep[e.g.][]{Mott}.
1937: What is the missing physics that can explain the discrepancy?
1938: The explanation that is closest to the spirit of our
1939: models is that real molecular clouds start their lives with an
1940: inhomogeneous distribution of physical quantities, including the mass-to-flux
1941: ratio. We expect that the values of $\mu$ in a molecular cloud
1942: may fall within the observationally-established range of $[0.5,2]$, i.e. within
1943: a factor of two of the critical value in each direction. In that case, 
1944: we can make a simple estimate of a core mass distribution by adding up the
1945: histograms for the five models with $\mui=[0.5,0.8.1.0,1.1,2.0]$ shown
1946: in Fig.~\ref{masshist}. Since the model with $\mui=1.1$ is run fewer times but with a
1947: larger box size, we sample the number of cores necessary to give equal area
1948: weighting with the other models. The resulting histogram is presented in 
1949: Fig.~\ref{totalmasshist}. The broad variation of peaks in the individual
1950: histograms seen in Fig.~\ref{masshist} leads to a smooth power-law
1951: tail in the high mass end of the composite histogram. Fig.~\ref{totalmasshist}
1952: reveals a slope $d\log N/d\log M \approx -2$ in this region, only somewhat
1953: steeper than
1954: the value $d\log N/d\log M \approx -1.5$ measured for example by \citet{Mott}.
1955: We note that this general mechanism, arising from an initially inhomogeneous
1956: distribution of mass-to-flux ratio, is an interesting new possibility for 
1957: explaining the observed broad core mass distributions. 
1958: An important point is that a relatively {\it narrow} distribution of
1959: initial mass-to-flux ratios (factor of a few variation) can lead to a relatively
1960: broad distribution of fragment masses.
1961: This mechanism remains to be explored
1962: more generally and does not exclude the
1963: occurrence of other mechanisms like
1964: competitive accretion in a more global model \citep{Bonn}, a temporal 
1965: spread of core accretion lifetimes \citep{Myer00,BJ04}, and 
1966: turbulent fragmentation \citep[e.g.][]{Pado, Kless, Gamm,Till}. 
1967: 
1968: %The core mass functions that we have compiled from our simulations
1969: %are narrowly-peaked and do not match the broader distributions commonly
1970: %observed in cluster-forming regions \citep[e.g.][]{Mott}. Our mass
1971: %distributions represent a snapshot in time at the moment that the first
1972: %core(s) is/are undergoing runaway collapse. 
1973: %We suggest that the broader observed mass distributions may be due to any one
1974: %or a combination of the following effects: (1) subsequent temporal evolution of the
1975: %cores (i.e. differing available times for mass accumulation from the cloud environment);
1976: %(2) global effects of positioning of the cores within the overall cloud
1977: %so that their masses vary significantly,
1978: %i.e. that cores in regions of different mean density and/or different background 
1979: %mass-to-flux ratio may form cores of different masses;
1980: %and (3) directed flows creating many significantly super-Jeans
1981: %scale fragments in the first place. Future simulations will clarify these effects.
1982: %The first idea has been presented in the literature using simplified 
1983: %analytic and semi-analytic models \citep{Myer00,BJ04}. The second idea
1984: %is similar to that of competitive accretion \citep{Bonn}, but with an additional
1985: %possible twist - the effect of varying mass-to-flux ratio within the cloud.
1986: %The third idea also encompasses that of direct
1987: %turbulent fragmentation \citep[e.g.][]{Pado, Kless, Gamm}, 
1988: %which we will address in the context of our models in an upcoming paper.
1989: 
1990: The richness of the physics revealed by our models of gravitational fragmentation
1991: up to runaway collapse of the first core, under conditions of varying mass-to-flux ratio, 
1992: ionization level, and external pressure, pave the way for more extensive models 
1993: in the future. 
1994: %Fragmentation patterns under various conditions
1995: %can be compared in detail with specific observed cluster-forming regions. 
1996: %One can also investigate further the novel {\it two-stage fragmentation} of some
1997: %of the transcritical models, which form an initially super-Jeans fragment but
1998: %may break into smaller fragments as the core becomes decidedly supercritical.
1999: Adding the effects of initial cloud turbulence, implementing a technique for integrating past the
2000: formation of the first generation of stars, and including some form of energy 
2001: feedback from star formation, remain to be done.
2002: The addition of new and more complex effects will be facilitated by the fact that
2003: the thin-sheet approximation allows efficient calculation of the fragmentation
2004: process while retaining a high level of realism. In this context we point out that
2005: the recent fully three-dimensional fragmentation simulation of 
2006: \citet{Kudo07} with magnetic fields and ambipolar diffusion bears out the 
2007: main physical results presented by BC04.
2008: 
2009: \section{Summary}
2010: \label{s:summ}
2011: 
2012: We have carried out a large number of model simulations to study the 
2013: effect of initial mass-to-flux ratio, neutral-ion coupling, and external
2014: pressure on dense core formation from gravitational 
2015: fragmentation of isothermal sheet-like layers that may be embedded within larger
2016: molecular cloud envelopes. 
2017: Our simulation box is periodic in the lateral ($x,y)$ directions and 
2018: typically span four nonmagnetic (Jeans) fragmentation scales in each of these
2019: directions.
2020: The simulations reveal a wide range of outcomes, from
2021: the unique transcritical fragmentation mode into massive cores, to the 
2022: pressure-driven fragmentation into dense clusters. 
2023: We emphasize the following main results of the paper:
2024: 
2025: \begin{enumerate}
2026: 
2027: \item{\it Fragmentation Spacing.}
2028: %Core formation is initiated in a 
2029: %dynamical fashion in the highly supercritical clouds ($\mui > 2$), 
2030: %with field-line dragging, and in a quasistatic fashion in the highly 
2031: %subcritical clouds ($\mui \leq 0.5$), due to ambipolar drift with near-stationary
2032: %field lines. The intermediate case of transcritical ($\mui \approx 1$)
2033: The average spacings of nonlinearly developed fragments are generally
2034: in excellent agreement 
2035: with the preferred fragmentation scale of linear perturbation theory 
2036: \citep{CB06}, although there is
2037: definite irregularity in any simulation, with variation of fragment spacings.
2038: Both significantly subcritical {\it and} highly supercritical clouds 
2039: have average fragmentation scales $\lamavg \approx 2 \pi Z_0$, where   
2040: $Z_0$ is the half-thickness of the background state. 
2041: The transcritical ($\mui \approx 1$) models exhibit very large (super-Jeans) 
2042: average fragment 
2043: spacings, although there is evidence for nonlinear second-stage fragmentation in 
2044: some cases, which makes the average spacing slightly smaller than predicted by 
2045: linear theory.  
2046: Variation of the ionization fraction by a factor of two above and below the standard
2047: value does not have a big effect on fragment spacing. However, an external pressure
2048: dominated sheet undergoes dramatically smaller scale fragmentation to form 
2049: a dense cluster.
2050: 
2051: %nonlinear spacing agrees with linear prediction
2052: %discuss transcritical case and 2nd stage frag.?
2053: %explain for peak in fragmentation spacing
2054: %high Pext case?
2055: %transcrit may be tied to massive cores or to Taurus
2056: 
2057: \item{\it Time Evolution to Runaway.} The times $t_{\rm run}$ for various models to reach
2058: runaway collapse of the first core varies significantly for models with differing
2059: initial dimensionless mass-to-flux ratio $\mui$, neutral-ion coupling
2060: parameter $\tniitil$, and dimensionless external pressure $\Pexttil$. 
2061: Values of $t_{\rm run}$ range from 0.45 Myr to 9.53 Myr, each scaling as 
2062: $(\Nni/10^{22}\,\cms)^{-1} (T/10\K)^{1/2}$.
2063: The supercritical
2064: clouds evolve much more rapidly than the critical or subcritical clouds, with 
2065: the highly supercritical clouds evolving $\approx 10$ times more rapidly than a
2066: critical cloud, for the typical level of neutral-ion coupling. A critical cloud 
2067: in turn evolves more rapidly than a subcritical cloud, but the variation is a factor
2068: of order unity for plausible initial values of $\mui$; for example the 
2069: $\mui = 0.5$ model reaches runaway collapse in a time that is 1.7 times longer than for the
2070: $\mui=1$ model. 
2071: In all cases, the time to runaway collapse is $\approx 10 \, \taugm$ when starting with
2072: small-amplitude white noise perturbations, where
2073: $\taugm$ is the growth time of the fastest growing eigenmode mode in linear
2074: perturbation theory. 
2075: The quantity $\taugm$ itself varies from $\approx Z_0/\cs$ for highly 
2076: supercritical 
2077: models to $\approx 10 Z_0/\cs$ for highly subcritical models (for a typical neutral-ion
2078: coupling level).
2079: The effect of varying $\tniitil$ (and hence the initial ionization fraction $x_{\rm i,0}$)
2080: is that $t_{\rm run} \propto \tniitil^{-1} \propto x_{\rm i,0}$ approximately for ambipolar-drift-driven (critical or subcritical) fragmentation, so that the canonical $\taugm \approx 10 Z_0/\cs$
2081: and our calculated $t_{\rm run} \approx 100 Z_0/\cs$ are both possibly subject to significant variation.
2082: A pressure dominated cloud with $\Pexttil=10$ has $t_{\rm run}$ about 5-6 times shorter
2083: than clouds with small external pressure but other parameters held fixed.
2084: %This delay in growth occurs because the preferred mode
2085: %cannot immediately emerge from the small-amplitude white noise perturbations. 
2086: %The uniform density background also 
2087: %leads to a slower initial development of the cores than for an initially peaked density
2088: %profile, but allows a self-consistent calculation of the entire core formation process
2089: %due to dynamical contraction and/or ambipolar drift.
2090: 
2091: %all times are about 10 times linear theory prediction
2092: %upper limit since from unif background
2093: 
2094: \item{\it Velocities in the Nonlinear Regime.}
2095: Maximum infall speeds of neutrals can become supersonic on core scales in the 
2096: supercritical clouds, but remain subsonic for critical or subcritical clouds. 
2097: The latter is true even if the ionization fraction is reduced by a factor of
2098: two.
2099: The ion speeds in the cores closely follow the neutral speeds but are somewhat smaller,
2100: since the gravitationally-driven motion of the neutrals is generally opposed in
2101: the plane of the sheet by magnetic fields. 
2102: %vn > vi in plane
2103: 
2104: \item{\it Core Shapes.}
2105: An extensive compilation of core shapes shows that the distributions have a
2106: peak that is near-circular for the subcritical and critical fragmentation models. 
2107: However, the cores become more elongated in the sheet for supercritical clouds,
2108: with a mean axis ratio in the sheet plane $\approx 0.5$. The half-thickness remains
2109: smaller than either semiminor or semimajor axis in the sheet, so that the cores
2110: are triaxial and preferentially flattened along the direction of the mean 
2111: magnetic field.
2112: Preliminary comparison of our results with published deprojections
2113: of the observed axis ratios of cores shows the best agreement with our 
2114: subcritical models.
2115: 
2116: \item{\it Core Mass Distributions.}
2117: An extensive compilation of core masses shows that the peak of the distributions 
2118: are related to the preferred fragmentation mass $\mgm$ of linear theory.
2119: Transcritical fragmentation yields the largest peak masses while the highly
2120: supercritical and decidely subcritical limits have smaller (Jeans-like) peaks
2121: and similar distributions. That the peak mass is strongly selected is seen in
2122: the very sharp drops in the distributions for greater masses. This means that
2123: the observed relatively broad-tailed core mass distributions cannot be explained by
2124: a pure local gravitational fragmentation process in a medium of uniform background
2125: column density and mass-to-flux ratio. 
2126: Cores defined as significantly supercritical regions within a larger subcritical
2127: common envelope also have a very narrowly-peaked distribution. 
2128: However, a composite mass histogram that may mimic the effect of an 
2129: inhomogeneous assortment of 
2130: initial mass-to-flux ratios,
2131: does produce a broad core mass distribution that resembles its observed counterparts.
2132: %discuss sc cores
2133: 
2134: \item{\it Magnetic Field Line Structure.}
2135: Contraction of cores within a supercritical cloud yields significant
2136: curvature of the magnetic field lines and very apparent hourglass morphologies,
2137: since contraction proceeds primarily with field-line dragging. The transcritical
2138: and subcritical clouds form cores through a process driven in large part by
2139: neutral-ion slip, and result in lesser curvature in the magnetic field. 
2140: This holds out the hope of using the observed curvature of 
2141: possible future observations of hourglass magnetic fields on the core scale 
2142: as a proxy for measuring the ambient mass-to-flux ratio.
2143: 
2144: \end{enumerate}
2145: 
2146: %time of evolution, slow to rapid, 10 times linear time
2147: %times may be long but starting from uniform state
2148: %velocities at time of nonlinear development , also vn > vi
2149: %variation of shapes
2150: %CMF very peaked, requires long term evolution to settle
2151: %FL shapes?
2152: %transcritical peak - relate to isolated star formation?
2153: %second stage fragmentation for mu=1.1
2154: %sensitive dependence on ionization fraction
2155: %clusters in high pext cases
2156: 
2157: 
2158: \section*{Acknowledgements}
2159: 
2160: We thank the anonymous referee for comments which significantly improved
2161: the manuscript. We also thank
2162: Wolfgang Dapp for valuable comments on the manuscript and thank 
2163: both him and Stephanie Keating for creating 
2164: several color images and animations. 
2165: The IFRIT package, developed by Nick Gnedin, was used fruitfully 
2166: to create some color images and magnetic field line visualizations.
2167: Our IDL code benefited from the use of an Adams-Bashforth-Moulton
2168: ODE solver converted to IDL format by Craig Markwardt from the public 
2169: domain Fortran routine written by
2170: L. F. Shampine and H. A. Watts of Sandia Laboratories.
2171: SB was supported by a grant from the Natural Sciences and Engineering
2172: Research Council (NSERC) of Canada. 
2173: %GC was supported by the New York Center for the
2174: %Origins of Life (NSCORT) and the Department of Physics, Applied
2175: %Physics, and Astronomy at Rensselaer Polytechnic Institute, under
2176: %NASA Grant NAG5-7589. 
2177: JW was supported by an NSERC Undergraduate Summer Research Award. 
2178: SB would also like to thank the KITP Santa Barbara
2179: for their hospitality during the final stages of writing this paper,
2180: when this research was supported in part by the National Science 
2181: Foundation under Grant No. NSF PHY05-51164.
2182: 
2183: 
2184: \begin{thebibliography}{}
2185: 
2186: \bibitem[Alves \& Franco(2007)]{Alve}
2187: Alves, F. O., Franco, G. A. P. 2007. A\&A 470, 597.
2188: 
2189: \bibitem[Andr\'{e} et al.(1996)]{Andr} 
2190: Andr\'{e}, P., Ward-Thompson, D., Motte, F. 1996. A\&A 314, 625.
2191: 
2192: \bibitem[Andr\'{e} et al.(2000)]{Andr00}
2193: Andr\'{e}, P., Ward-Thompson, D., Barsony, M., 2000. In: Mannings, V.,
2194: Boss, A. P., Russell S. S. (Eds.), Protostars and Planets IV. University of Arizona
2195: Press, Tucson, p. 59.
2196: 
2197: \bibitem[Bacmann et al.(2000)]{Bacm}
2198: Bacmann, A., Andr\'{e}, P., Puget, J.-L., Abergel, A., Bontemps, S.,
2199: Ward-Thompson, D. 2000. A\&A  361, 555.
2200: 
2201: %\bibitem[Basu(2000)]{Basu} Basu, S. 2000. ApJ 540, L103.
2202: 
2203: \bibitem[Basu \& Ciolek(2004)]{BC04}Basu, S., Ciolek, G. E. 2004. ApJ 607, L39 (BC04).
2204: 
2205: \bibitem[Basu \& Jones(2004)]{BJ04} Basu, S., Jones, C. E. 2004. MNRAS
2206: 347, L47.
2207: 
2208: \bibitem[Basu \& Mouschovias(1994)]{BM94} Basu, S., Mouschovias, T. Ch. 
2209: 1994. ApJ 432, 720.
2210: 
2211: %\bibitem[Basu \& Mouschovias(1995)]{BM95}\ul. 1995. ApJ 453, 271.
2212: 
2213: \bibitem[Bate et al.(2003)]{Bate}
2214: Bate, M. R., Bonnell, I. A., Bromm, V. 2003. MNRAS 339, 577.
2215: 
2216: \bibitem[Benson \& Myers(1989)]{Bens} Benson, P. J., Myers, P. C. 1989.
2217: ApJS 71, 89.
2218: 
2219: %\bibitem[Bonnell et al.(1997)]{Bonn}
2220: %Bonnell I. A., Bate M. R., Clarke C. J., Pringle J. E. 1997. MNRAS, 285, 201
2221: 
2222: \bibitem[Bonnell et al.(2003)]{Bonn}
2223: Bonnell, I. A., Bate, M. R., Vine, S. G. 2003. MNRAS 343, 413.
2224: 
2225: \bibitem[Bourke et al.(2001)]{Bour} Bourke, T. L., Myers, P. C., Robinson, G.,
2226: Hyland, A. R. 2001. ApJ 554, 916.
2227: 
2228: \bibitem[Caselli et al.(2002)]{Case}
2229: Caselli, P., Walmsley, C. M., Zucconi, A., Tafalla, M., Dore, L., Myers, P. C.
2230: 2002. ApJ 565, 331.
2231: 
2232: \bibitem[Ciolek(1996)]{Ciol} 
2233: Ciolek, G. E. 1996. In: Roberge, W. G., Whittet, D. C. B.
2234: (Eds.), Polarimetry of the Interstellar Medium. ASP Conference
2235: Series Vol. 97. Astronomical Society of the Pacific, p. 542.
2236: 
2237: %\bibitem[Ciolek \& Basu(2000)]{CB00} Ciolek, G. E., Basu, S. 2000. 
2238: %ApJ 529, 925.
2239: 
2240: \bibitem[Ciolek \& Basu(2006)]{CB06} Ciolek, G. E., Basu, S. 2006. ApJ  652, 442 (CB06).
2241: %\bibitem[Ciolek \& Basu(2006)]{CB06} \ul. 2006. ApJ  652, 442 (CB06).
2242: 
2243: \bibitem[Ciolek \& Mouschovias(1993)]{CM93} 
2244: Ciolek, G. E., Mouschovias, T. Ch. 1993. ApJ 418, 774.
2245: 
2246: \bibitem[Ciolek \& Mouschovias(1994)]{CM94} 
2247: %Ciolek, G. E., Mouschovias, T. Ch. 1994. ApJ 425, 142.
2248: \ul. 1994. ApJ 425, 142.
2249: 
2250: %\bibitem[Ciolek \& Mouschovias(1995)]{CM95} \ul. 1995. ApJ 454, 194.
2251: 
2252: \bibitem[Ciolek \& Mouschovias(1998)]{CM98} \ul. 1998. ApJ 504, 280.
2253: 
2254: \bibitem[Cortes et al.(2005)]{Cort}
2255: Cortes, P. C., Crutcher, R. M., Watson, W. D. 2005. ApJ 628, 780.
2256: 
2257: \bibitem[Crutcher(1999)]{Crut99} Crutcher, R. M. 1999. ApJ 520, 706.
2258: 
2259: \bibitem[Crutcher et al.(2004)]{Crut04} Crutcher, R. M., Nutter, D. J., 
2260: Ward-Thompson, D., Kirk, J. M. 2004. ApJ 600, 279.
2261: 
2262: \bibitem[Elmegreen(1979)]{Elme79} Elmegreen, B. G. 1979. ApJ 232, 729.
2263: 
2264: \bibitem[Elmegreen(2007)]{Elme07}
2265: \ul.  2007. ApJ 668, 1064.
2266: 
2267: \bibitem[Elmegreen \& Elmegreen(1978)]{Elme78}
2268: Elmegreen, B. G., Elmegreen, D. M. 1978. ApJ 220, 1051.
2269: 
2270: \bibitem[Elmegreen \& Falgarone(1996)]{Elme96}
2271: Elmegreen, B. G., Falgarone, E. 1996. ApJ 471, 816.
2272: 
2273: \bibitem[Fatuzzo \& Adams(2002)]{Fatu}
2274: Fatuzzo, M., Adams, F. C. 2002. ApJ 570, 210.
2275: 
2276: \bibitem[Fiedler \& Mouschovias(1993)]{FM93} Fiedler, R. A., Mouschovias, T. Ch. 1993. ApJ 415, 680.
2277: 
2278: \bibitem[Folini et al.(2004)]{Foli} Folini, D., Heyvaerts, J., Waleder, R. 2004.
2279: A\&A 414, 559.
2280: 
2281: \bibitem[Fuller \& Myers(1993)]{Full} Fuller, G. A., Myers, P. C. 1993. ApJ
2282: 418, 273.
2283: 
2284: \bibitem[Gammie et al.(2003)]{Gamm}
2285: Gammie, C. F., Lin, Y.-T., Stone, J. M., Ostriker, E. C. 2003. ApJ 592, 203.
2286: 
2287: \bibitem[Goldsmith et al.(2008)]{Gold}
2288: Goldsmith, P. F., Heyer, M., Narayanan, G., Snell, R., Li, D., Brunt, C.
2289: 2008. ApJ 680, 428.
2290: 
2291: \bibitem[Gomez et al.(1993)]{Gome} 
2292: Gomez, M., Hartmann, L., Kenyon, S. J., Hewett, R. 1993. AJ 105, 1927.
2293: 
2294: \bibitem[Goodman et al.(1990)]{Good90}
2295: Goodman, A. A., Bastien, P., Myers, P. C., M\'enard, F. 1990. ApJ 359, 363. 
2296: 
2297: \bibitem[Goodman et al.(1998)]{Good98} Goodman, A. A., Barranco, J. A., 
2298: Wilner, D. J., Heyer, M. H. 1998. ApJ 504, 223.
2299: 
2300: \bibitem[Goodwin et al.(2002)]{Goodw}
2301: Goodwin, S. P., Ward-Thompson, D., Whitworth, A. P. 2002. MNRAS 330, 769.
2302: 
2303: \bibitem[Heitsch et al.(2008)]{Heit} 
2304: Heitsch, F., Hartmann, L. W., Slyz, A. D., Devriendt, J. E. G., Burkert, A. 2008. ApJ 674, 316.
2305: 
2306: %\bibitem[Hennebelle et al.(2008)]{Henn}
2307: %Hennebelle, P., Mac Low, M.-M., V\'azquez-Semadeni, E. 2008. 
2308: %In: Chabrier, G. (Ed.), Structure Formation in the Universe, Cambridge Univ. Press,
2309: %Cambridge, in press. 
2310: 
2311: \bibitem[Hennebelle et al.(2008)]{Henn}
2312: Hennebelle, P., Banerjee, R., V\'azquez-Semadeni, E., Klessen, R., Audit, E.
2313: 2008. A\&A submitted (arXiv:0805.1366).
2314: 
2315: \bibitem[Heyer et al.(2008)]{Heye}
2316: Heyer, M., Gong, H., Ostriker, E., Brunt, C. 2008. ApJ 680, 420.
2317: 
2318: \bibitem[Indebetouw \& Zweibel(2000)]{Inde} Indebetouw, R., Zweibel, E. G. 2000. ApJ 532, 361.
2319: 
2320: \bibitem[Jeans(1929)]{Jean} Jeans, J. H. 1929. Astronomy and Cosmogony.
2321: Cambridge University Press, Cambridge.
2322: 
2323: \bibitem[Jijina et al.(1999)]{Jiji} Jijina, J., Myers, P. C., Adams, F. C.
2324: 1999. ApJS 125, 161.
2325: 
2326: \bibitem[Johnstone et al.(2004)]{John} 
2327: Johnstone, D., Di Francesco, J., Kirk, H. 2004. ApJ 611, L45.
2328: 
2329: \bibitem[Jones \& Basu(2002)]{JB02} Jones, C. E., Basu, S. 2002. ApJ 569, 280.
2330: 
2331: \bibitem[Jones et al.(2001)]{JBD01} Jones, C. E., Basu, S., Dubinski, J. 2001. ApJ 551, 387.
2332: 
2333: \bibitem[Kim et al.(2002)]{Kim}
2334: Kim, W.-T., Ostriker, E. C., Stone, J. M. 2002. ApJ 581, 1080. 
2335: 
2336: \bibitem[Kirk et al.(2005)]{Kirk}
2337: Kirk, J. M., Ward-Thompson, D., Andr\'{e}, P. 2005. MNRAS 350, 1506.
2338: 
2339: \bibitem[Klessen(2001)]{Kless}
2340: Klessen, R. S. 2001. ApJ 556, 837.
2341: 
2342: \bibitem[Kudoh \& Basu(2003)]{Kudo03}Kudoh, T., Basu, S. 2003. ApJ 595, 842.
2343: 
2344: \bibitem[Kudoh \& Basu(2006)]{Kudo06}\ul. 2006. ApJ 642, 270.
2345: 
2346: \bibitem[Kudoh et al.(2007)]{Kudo07} 
2347: Kudoh, T., Basu, S., Ogata, Y., Yabe, T. 2007. MNRAS 380, 499.
2348: 
2349: \bibitem[Lada \& Lada(2003)]{Lada03}
2350: Lada, C. J., Lada, E. A. 2003. ARAA 41, 57.
2351: 
2352: \bibitem[Lada et al.(2007)]{Lada07}
2353: Lada, C. J., Alves, J. F., Lombardi, M. 2007. 
2354: In: Reipurth, B., Jewitt, D., Keil, K. (Eds.), Protostars and Planets V.
2355: University of Arizona Press, Tucson, p. 3.
2356: 
2357: \bibitem[Langer(1978)]{Lang78} Langer, W. D. 1978. ApJ 225, 95.
2358: 
2359: \bibitem[Larson(1985)]{Lars85} Larson, R. B. 1985. MNRAS 214, 379.
2360: 
2361: \bibitem[Larson(2003)]{Lars03} \ul. 2003. Rep. Prog. Phys. 66, 1651.
2362: 
2363: \bibitem[Lee et al.(2001)]{Lee} Lee, C. W., Myers, P. C., Tafalla, M. 2001. ApJS 136, 603.
2364: 
2365: %\bibitem[Li \& Nakamura(2002)]{Li} 
2366: %Li, Z.-Y., Nakamura, F. 2002. ApJ 578, 256.
2367: 
2368: \bibitem[Li \& Nakamura(2004)]{Li04} 
2369: Li, Z.-Y., Nakamura, F. 2004. ApJ 609, L83.
2370: 
2371: \bibitem[Lubow \& Pringle(1993)]{Lubo}
2372: Lubow, S. H., Pringle, J. E. 1993. MNRAS 263, 701.
2373: 
2374: \bibitem[McDaniel \& Mason(1973)]{McDa}
2375: McDaniel, E. W., Mason, E. A. 1973. The Mobility
2376: and Diffusion of Ions and Gases. Wiley, New York.
2377: 
2378: \bibitem[McKee(1989)]{McKe} McKee, C. F. 1989. ApJ 345, 782.
2379: 
2380: \bibitem[McKee(1999)]{McKe99} \ul. 1999. 
2381: In: Lada, C. J., Kylafis, N. (Eds.),
2382: The Origin of Stars and Planetary Systems. Kluwer, Dordrecht, p. 29.
2383: 
2384: \bibitem[Mac Low \& Klessen(2004)]{MacL}
2385: Mac Low, M.-M., Klessen, R. S. 2004. Rev. Mod. Phys. 76, 125.
2386: 
2387: \bibitem[Mestel \& Spitzer(1956)]{Mest} Mestel, L., Spitzer, L. Jr. 1956.
2388: MNRAS 116, 503.
2389: 
2390: \bibitem[Miyama et al.(1987a)]{Miya1}
2391: Miyama, S, Narita, S., Hayashi, C. 1987. Prog. Theor. Phys. 78, 1051.
2392: 
2393: \bibitem[Miyama et al.(1987b)]{Miya2}
2394: Miyama, S, Narita, S., Hayashi, C. 1987. Prog. Theor. Phys. 78, 1273.
2395: 
2396: %\bibitem[Morton et al.(1994)]{MMC94} Morton, S. A., Mouschovias, T. Ch., 
2397: %Ciolek, G. E. 1994. ApJ 421, 561.
2398: 
2399: \bibitem[Motte et al.(1998)]{Mott} Motte, F., Andr\'{e}, P., Neri, R. 1998. A\&A 336, 150.
2400: 
2401: %\bibitem[Mouschovias(1977)]{Mous77} Mouschovias, T. Ch. 1977. ApJ 211, 147.
2402: 
2403: \bibitem[Mouschovias(1978)]{Mous78} Mouschovias, T. Ch. 1978. 
2404: In: Gehrels , T. (Ed.), Protostars and Planets. University of Arizona 
2405: Press, Tucson, p. 209.
2406: 
2407: %\bibitem[Mouschovias \& Paleologou(1980)]{Mous80} 
2408: %Mouschovias, T. Ch., Paleologou, E. V. 1980. ApJ 237, 877.
2409: 
2410: %\bibitem[]{M87} Mouschovias, T. Ch. 1987, in Physical Processes in
2411: %Interstellar Clouds, ed. G. Morfil \& M. Scholer (Dordrecht: Reidel), 453
2412: 
2413: \bibitem[Muench et al.(2007)]{Muen} 
2414: Muench, A. A., Lada, C. J., Rathborne, J. M., Alves, J. F., Lombardi, M.
2415: 2007. ApJ 671, 1820. 
2416: 
2417: %\bibitem[]{Myers} Myers, P. C., Fuller, G. A., Goodman, A. A., \& Benson,
2418: %P. J. 1991, \apj, 376, 561
2419: 
2420: %\bibitem[Myers(2005)]{Myer05}Myers, P. C. 2005, \apj, 623, 280
2421: 
2422: \bibitem[Myers(1983)]{Myer83a} Myers, P. C. 1983. ApJ 270, 105.
2423: 
2424: \bibitem[Myers(2000)]{Myer00} \ul. 2000. ApJ 530, L119.
2425: 
2426: \bibitem[Myers \& Benson(1983)]{Myer83b} Myers, P. C., Benson, P. J. 1983. ApJ 266, 309.
2427: 
2428: \bibitem[Myers et al.(1991)]{Myer91} Myers, P. C., Fuller, G. A., 
2429: Goodman, A. A., Benson, P. J. 1991. ApJ 376, 561.
2430: 
2431: %\bibitem[Nakamura \& Hanawa(1997)]{Naka97} Nakamura, F., Hanawa, T. 1997. ApJ 480, 701.
2432: 
2433: %\bibitem[Nakamura \& Li(2002)]{Naka02} Nakamura, F., \& Li, Z.-Y. 2002, \apj, 566, L101
2434: 
2435: \bibitem[Nakamura \& Li(2005)]{Naka05} Nakamura, F., Li, Z.-Y. 2005. ApJ 631, 411. 
2436: 
2437: \bibitem[Nakamura \& Li(2008)]{Naka08} 
2438: %Nakamura, F., Li, Z.-Y. 2008. ApJ submitted (arXiv:0804.4201).
2439: \ul. 2008. ApJ submitted (arXiv:0804.4201).
2440: 
2441: \bibitem[Onishi et al.(2002)]{Onis}
2442: Onishi, T., Mizuno, A., Kawamura, A., Tachihara, K., Fukui, Y. 2002. ApJ 575, 950.
2443: 
2444: \bibitem[Padoan et al.(1997)]{Pado}
2445: Padoan, P., Nordlund, A., Jones, B. J. T. 1997. MNRAS 288, 145.
2446: %\bibitem[Padoan \& Nordlund(2002)]{Pado} 
2447: %Padoan, P., Nordlund, A. 2002. ApJ 576, 870.
2448: 
2449: \bibitem[Parker(1966)]{Park}
2450: Parker, E. 1966. ApJ 145, 811.
2451: 
2452: \bibitem[Press et al.(1996)]{Press}Press, W. H., Teukolsky, S. A., Vetterling, W. T.,
2453: Flannery, B. P. 1996. Numerical Recipes in Fortran 77: The Art
2454: of Scientific Computing (Vol. 1 of Fortran Numerical Recipes), 2nd. Ed.
2455: Cambridge, New York.
2456: 
2457: \bibitem[Price \& Bate(2008)]{Pric}
2458: Price, D. J., Bate, M. R. 2008. MNRAS 385, 1820.
2459: 
2460: \bibitem[Schiesser(1991)]{Schi} Schiesser, W. E. 1991. The Numerical 
2461: Method of Lines: Method of Integration of Partial Differential Equations. Academic Press, San Diego.
2462: 
2463: \bibitem[Schleuning(1998)]{Schl} Schleuning, D. A. 1998. ApJ 493, 811.
2464: 
2465: \bibitem[Shampine(1994)]{Sham94}
2466: Shampine, L. F. 1994. Numerical Solution of Ordinary Differential
2467: Equations. Chapman \& Hall, New York.
2468: 
2469: \bibitem[Shu et al.(1987)]{Shu87} 
2470: Shu, F. H., Adams, F. C., Lizano, S. 1987, ARA\&A 25, 23.
2471: 
2472: \bibitem[Shu et al.(1999)]{Shu99} 
2473: Shu, F. H., Allen, A., Shang, H., Ostriker, E. C., Li, Z.-Y. 1999. 
2474: In: Lada, C. J., Kylafis, N. (Eds.),
2475: The Origin of Stars and Planetary Systems. Kluwer, Dordrecht, p. 193.
2476: 
2477: %\bibitem[Shu et al.(2004)]{Shu04}
2478: %Shu, F. H., Li, Z.-Y., Allen, A. 2004, ApJ 601, 930.
2479: 
2480: \bibitem[Tafalla et al.(1998)]{Tafa} Tafalla, M., Mardones, D., Myers, P. C., Caselli, P.,
2481: Bachiller, R., Benson, P. J. 1998. ApJ 504, 900.
2482: 
2483: \bibitem[Tassis(2007)]{Tass}
2484: Tassis, K. 2007. MNRAS 379, L50.
2485: 
2486: \bibitem[Teixeira et al.(2005)]{Teix}
2487: Teixeira, P. S., Lada, C. J., Alves, J. 2005. ApJ 629, 276.
2488: 
2489: \bibitem[Tilley \& Pudritz(2007)]{Till}
2490: Tilley, D. A., \& Pudritz, R. E. 2007. MNRAS 382, 73.
2491: 
2492: %\bibitem[Tomisaka et al.(1990)]{Tomi90} 
2493: %Tomisaka, K., Ikeuchi, S., Nakamura, T. 1990. ApJ 362, 202.
2494: 
2495: \bibitem[Umebayashi \& Nakano(1980)]{UN80}
2496: Umebayashi, T., Nakano, T. 1980. PASJ 32, 405.
2497: 
2498: \bibitem[van Leer(1977)]{vanL} van Leer, B. 1977. JCP 23, 276.
2499: 
2500: \bibitem[V\'azquez-Semadeni et al.(2006)]{Vazq06}
2501: V\'azquez-Semadeni, E., Ryu, D., Passot, T., Gonz\'alez, R. F., 
2502: Gazol, A. 2006. ApJ 643, 245.
2503: 
2504: \bibitem[V\'azquez-Semadeni et al.(2007)]{Vazq07}
2505: V\'azquez-Semadeni, E., G\'omez, G. C., Jappsen, A. K., Ballesteros-Paredes, J.,
2506: Gonz\'alez, R. F., Klessen, R. S. 2007. ApJ 657, 870.
2507: 
2508: %\bibitem[Vorobyov \& Basu(2005a)]{VB05a}
2509: %Vorobyov, E. I., Basu, S. 2005a, MNRAS 360, 675.
2510: 
2511: %\bibitem[Vorobyov \& Basu(2005b)]{VB05}
2512: %Vorobyov, E. I., Basu, S. 2005. ApJ  633, L137.
2513: 
2514: %\bibitem[Vorobyov \& Basu(2006)]{VB06}
2515: %\ul. 2006. ApJ 650, 956.
2516: 
2517: %\bibitem[Vorobyov \& Basu(2007)]{VB07}
2518: %\ul. 2007. MNRAS 381, 1009.
2519: 
2520: \bibitem[Ward-Thompson et al.(1994)]{Ward94}
2521: Ward-Thompson, D., Scott, P. F., Hills, R. E., Andr\'{e}, P. 1994. MNRAS 268, 276.
2522: 
2523: \bibitem[Ward-Thompson et al.(2007)]{Ward07}
2524: Ward-Thompson, D., Andr\'{e}, P., Crutcher, R., Johnston, D., Onishi, T.,
2525: Wilson, C. 2007. In: Reipurth, B., Jewitt, D., Keil, K. (Eds.),
2526: Protostars and Planets V. University of Arizona Press, Tucson, p. 33. 
2527: 
2528: \bibitem[Williams et al.(1999)]{Will} Williams, J. P., Myers, P. C., Wilner, D. J.,
2529: DiFrancesco, J. 1999. ApJ 513, L61.
2530: 
2531: \bibitem[Zweibel(1998)]{Zwei}
2532: Zweibel, E. G. 1998. ApJ 499, 746.
2533: 
2534: \bibitem[Zweibel(2002)]{Zwei02}
2535: \ul. 2002. ApJ 567, 962.
2536: 
2537: \end{thebibliography}
2538: 
2539: 
2540: 
2541: \end{document}
2542: 
2543: