0810.1393/ms.tex
1: %\documentclass[useAMS,usenatbib,usegraphicx]{mn2e}
2: \documentclass{emulateapj} 
3: %\documentclass[manuscript]{aastex}
4: %\documentclass[12pt,preprint]{aastex} 
5: %\usepackage{graphicx}
6: 
7: % User defined command
8: \newcommand{\msun}{\mbox{M$_{\sun}$}}
9: \newcommand{\msunpctwo}{\mbox{\, M$_{\sun} \, {\rm pc}^{-2}$}}
10: \newcommand{\natd}[2]{\mbox{$#1 \cdot 10^{#2}$}}
11: \newcommand{\pder}[2]{\frac{\partial #1}{\partial #2}}
12: \newcommand{\pdert}[1]{\pder{#1}{t}}
13: \newcommand{\mcorr}{\bf}         % highlight corrections
14: \newcommand{\mcorrm}{\boldmath}  % highlight corrections
15: \newcommand{\bl}[1]{\mbox{\boldmath$ #1 $}}
16: \newcommand{\rhat}{\mbox{\boldmath$ \hat{r}$}} 
17: \newcommand{\phihat}{\mbox{\boldmath$ \hat{\phi}$}} 
18: 
19: \slugcomment{Submitted to Astrophysical Journal}
20: \shorttitle{Disk masses around young stellar objects}
21: \shortauthors{Vorobyov} 
22: 
23: \begin{document}
24: 
25: \title{Disk masses in the embedded and T Tauri phases of stellar evolution}
26: \author{E. I. Vorobyov\altaffilmark{1,}\altaffilmark{2}}
27: \altaffiltext{1}{Institute for Computational Astrophysics, Saint Mary's University,
28: Halifax, B3H 3C3, Canada; vorobyov@ap.smu.ca.} 
29: \altaffiltext{2}{Institute of Physics, South Federal University, Stachki 194, Rostov-on-Don, 
30: 344090, Russia.} 
31: 
32: %\maketitle
33: 
34: \begin{abstract}
35: Motivated by a growing concern that masses of circumstellar disks may
36: have been systematically underestimated by conventional observational methods, 
37: we present a numerical hydrodynamics study of 
38: time-averaged disk masses ($\langle M_{\rm d} \rangle$) around low-mass
39: Class~0, Class~I, and Class~II objects. Mean disk masses ($\overline{M}_{\rm d}$)
40: are then calculated by weighting the time-averaged disk masses according to the 
41: corresponding stellar masses using a power-law weight function with a slope 
42: typical for the Kroupa initial mass function of stars.
43: Two distinct types of disks are considered: self-gravitating disks,
44: in which mass and angular momentum are redistributed exclusively by gravitational torques,
45: and viscous disks, in which both the gravitational and viscous torques are at work.
46: We find that self-gravitating disks have mean masses that are slowly increasing along the sequence
47: of stellar evolution phases. More specifically, Class~0/I/II self-gravitating 
48: disks have mean masses $\overline{M}_{\rm d}=0.09$, $0.10$, and $0.12~M_\odot$, respectively.
49: Viscous disks have similar mean masses ($\overline{M}_{\rm d}=0.10-0.11~M_\odot$) 
50: in the Class 0/I phases but almost a factor of 2 lower mean mass in the Class II phase 
51: ($\overline{M}_{\rm d,CII}=0.06~M_\odot$). In each evolution phase, time-averaged disk masses
52: show a large scatter around the mean value.
53: Our obtained mean disk masses are {\it larger} than
54: those recently derived by Andrews \& Williams and Brown et al., regardless of the 
55: physical mechanisms of mass transport in the disk. 
56: The difference is especially large for Class II disks, 
57: for which we find $\overline{M}_{\rm d,CII}=0.06-0.12~M_\odot$ but Andrews and Williams 
58: report median masses of order $3\times 10^{-3}~M_\odot$. 
59: When Class 0/I/II systems are considered altogether, a least-squares best fit yields 
60: the following relation between the time-averaged disk and
61: stellar masses, $\langle M_{\rm d} \rangle = \left( 0.2\pm 0.05 \right) 
62: \langle M_\ast \rangle^{1.3\pm 0.15}$.  
63: The dependence of $\langle M_{\rm d} \rangle$ on $\langle M_\ast \rangle$ 
64: becomes progressively steeper along the sequence of stellar evolution phases,
65: with exponents $0.7\pm 0.2$, $1.3\pm 0.15$, and $2.2\pm 0.2$ for Class~0, 
66: Class~I, and Class~II systems, respectively.
67: \end{abstract}
68: 
69: \keywords{circumstellar matter --- planetary systems: protoplanetary disks --- hydrodynamics --- ISM: clouds ---  stars: formation} 
70: 
71: 
72: 
73: \section{Introduction}
74: It has now become evident that disks of gas and dust are present from the earliest 
75: phases of stellar evolution (Class~0 and Class~I) and last for at least several million years
76: into the late Class~II phase. Disks are observed or inferred around most T Tauri stars and 
77: even around brown dwarfs. 
78: The evidence for disks around Class~0 and Class~I sources is more indirect.
79: In this early phase of stellar evolution, the protostar/disk system is deeply 
80: embedded in an envelope -- a remnant of the cloud core from which the protostar is forming. 
81: Nevertheless, recent observations by \cite{Andrews} suggest that Class I disks
82: have a larger median mass than that of Class II disks.
83: 
84: 
85: In spite of a considerable progress in the detection of disks around young stellar objects 
86: (YSOs), an accurate determination of disk masses is still challenging.
87: It is difficult to directly determine disk masses from the spectral lines 
88: of molecular species because the brightest, easily detectable lines 
89: (i.e., the rotational transitions of CO) are optically thick and likely to be severely 
90: depleted. Therefore, disk masses are usually inferred from analyzing the spectral 
91: energy distribution of YSOs from the mid-infrared through submillimeter bands 
92: \citep[e.g.][]{Andrews,Brown}. Such measurements of disk masses suffer from large 
93: uncertainties in the normalization of dust opacities and gas-to-dust ratios,
94: which led \citet{Hartmann} to conclude that T Tauri disk masses have been 
95: systematically underestimated by conventional analyses.
96: 
97: Another complication arises from poorly known physical processes in the disk.
98: A usual assumption of optically thin circumstellar disks may significantly 
99: underestimate disk masses, particularly for objects with larger flux densities.
100: However, a self-consistent treatment of a non-negligible optical depth requires 
101: a knowledge of the radial gas surface density profile in the disk \citep{Andrews}, 
102: which may depend significantly on the dominant physical mechanism of mass and angular 
103: momentum redistribution in the disk \citep{VB4}.
104: 
105: Given large uncertainties in the measurements of disk masses, numerical simulations
106: of self-consistent formation and evolution of circumstellar disks can provide  
107: valuable information on disk masses in the early embedded and late phases 
108: of YSO evolution. It has been shown in the past that the saturation of spiral
109: gravitational instabilities at a finite amplitude in a self-gravitating, Toomre-unstable disk 
110: allows for the steady transport of mass and momentum, which eventually limits
111: disk masses \citep[e.g.][]{Adams89,Shu90,Laughlin97,Laughlin98}.
112: In this paper, we perform numerical simulations of
113: the long-term evolution of self-consistently formed circumstellar disks
114: around low-mass stars ($0.2~M_\odot \la M_\ast\la 2.0~M_\odot$).
115: We consider both the self-gravitating disks, in which radial transport of mass and
116: angular momentum is done exclusively via gravitational torques, and viscous disks, which feature 
117: gravitational torques as well as viscous ones. 
118: We seek to determine numerically the disk masses in the Class 0, Class I, and Class II phases
119: of stellar evolution.
120: 
121: The paper is organized as follows. 
122: The numerical methods and initial parameters of cloud cores are 
123: given in \S~\ref{methods}. Our obtained masses for self-gravitating and viscous disks are
124: presented in \S~\ref{SG} and \S~\ref{VD}. We compare our numerical results with observations in 
125: \S~\ref{discuss}. The model and numerical caveats are discussed in \S~\ref{caveats}. 
126: The main results are summarized in \S~\ref{summary}.
127: 
128: 
129: \section{Description of numerical methods}
130: \label{methods}
131: We use the thin-disk approximation to compute the evolution of rotating, 
132: gravitationally bound cloud cores. This allows efficient calculation of 
133: the long-term evolution of a large number of models.
134: We start our numerical integration in the pre-stellar phase, which is characterized 
135: by a collapsing {\it starless} cloud core, and continue into the main accretion phase,
136: which sees the formation of a central star and circumstellar disk.
137: We cover all major phases of the evolution of a YSO, starting from 
138: its formation and ending with the T Tauri phase. The integration ends when the age of 
139: the central star is about three million years. In some models we extend the integration
140: up to 5~Myr. We emphasize that circumstellar disks
141: are formed self-consistently in our numerical simulations, 
142: rather than being introduced as an initial parameter of the model.
143: 
144: 
145: Once the disk is formed, its mass is determined by an interplay between the efficiency 
146: of the mass and angular momentum transport in the disk\footnote{In fact, disks may also transport
147: angular momentum to the external environment due to magnetic braking. This effect will be 
148: considered in a follow-up paper.} and the infall rate of matter from the surrounding envelope onto the
149: disk. At the time of disk formation, the infall rates take values between 
150: $1.2\times 10^{-6}~M_\odot$~yr$^{-1}$ and  $7.0\times 10^{-6}~M_\odot$~yr$^{-1}$  (measured at $600$~AU)
151: for the least and most massive cloud cores, respectively, but they show a fast decline with time.
152: These values (and strong time variation) are consistent with the infall rates 
153: derived by \citet{Klessen01} using 
154: numerical models that follow molecular cloud evolution from turbulent fragmentation toward the 
155: formation of stellar clusters. We note that once the disk is formed, 
156: the infall rate of matter from the envelope onto the disk is not necessarily the same 
157: as the mass accretion rate from the disk onto the protostar. While the former shows a fast decline
158: with time, the latter is usually characterized by a much slower decline and has a strong dependence
159: on the stellar mass \citep{VB3,VB5}.
160: 
161: 
162: We use two numerical approaches: a basic approach that accounts for the radial transport due to gravitational
163: toques and a viscous approach that accounts for the radial transport due to both the gravitational
164: torques and viscosity.
165: Gravitational torques are known to efficiently redistribute mass and 
166: angular momentum in circumstellar disks \citep[e.g.][]{Lodato04,Lodato05}. 
167: They were shown to play an important role in driving the FU-Ori-like bursts
168: in the early embedded phase of disk evolution \citep{VB1,VB2}. In the late disk evolution, 
169: negative gravitational torques associated with low-amplitude azimuthal density perturbations
170: in the disk can drive mass accretion rates that are consistent with those measured in 
171: the intermediate and upper-mass T Tauri stars \citep{VB3,VB5}.
172: 
173: 
174: 
175: \subsection{Basic numerical approach}
176: \label{basic}
177: 
178: In the basic numerical approach, the collapse of a cloud core and subsequent evolution 
179: of a star/disk system 
180: is carried out by solving the basic equations of mass and momentum transport 
181: in the thin-disk approximation \citep[see e.g.][]{VB2}
182: %are derived by integrating the ideal hydrodynamics equations in the z-direction,
183: %from $z = Z(r, \phi, t)$ to $z = -Z(r, \phi, t)$, and using Leibniz's rule
184: %for differentiation of integrals and the fundamental theorem of
185: %calculus \citep[see e.g.][]{VB2}
186: \begin{eqnarray}
187: \label{cont}
188:  \frac{{\partial \Sigma }}{{\partial t}} & = & - \nabla _p  \cdot \left( \Sigma \bl{v}_p 
189: \right), \\ 
190: \label{mom}
191:  \Sigma \frac{d \bl{v}_p }{d t}  & = &  - \nabla _p {\cal P}  + \Sigma \, \bl{g}_p \, ,
192: \end{eqnarray}
193: where $\Sigma$ is the mass surface density, ${\cal P}=\int^{Z}_{-Z} P dz$ is the vertically integrated
194: form of the gas pressure $P$, $Z$ is the radially and azimuthally varying vertical scale height,
195:  $\bl{v}_p=v_r \hat{\bl r}+ v_\phi \hat{\phi}$ is the velocity in the
196: disk plane, $\bl{g}_p=g_r \hat{\bl r} +g_\phi \hat{\phi}$ is the gravitational acceleration 
197: in the disk plane, and $\nabla_p=\hat{\bl r} \partial / \partial r + \hat{\phi} r^{-1} 
198: \partial / \partial \phi $ is the gradient along the planar coordinates of the disk. 
199: The gravitational acceleration $\bl{g}_p$ is found by solving for the Poisson integral 
200: \citep[see][]{VB2}. The fact that we account for the disk self-gravity means
201: that gravitational torques arise {\it self-consistently} in our numerical simulations and
202: not imitated by some means of $\alpha$-viscosity. 
203: 
204: Taking into account the complexity of gas thermodynamics in circumstellar disks
205: (see \S~\ref{caveats}), we have adopted a barotropic 
206: equation of state that closes equations~(\ref{cont}) and (\ref{mom}) and makes 
207: a transition from isothermal to adiabatic evolution at $\Sigma = \Sigma_{\rm cr} = 
208: 36.2$~g~cm$^{-2}$
209: \begin{equation}
210: {\cal P}=c_s^2 \Sigma +c_s^2 \Sigma_{\rm cr} \left( \Sigma \over \Sigma_{\rm cr} \right)^{\gamma},
211: \label{barotropic}
212: \end{equation}
213: where $c_s$ is the isothermal sound speed, the value of which is set equal to that 
214: of the initial cloud core, and $\gamma=1.4$. 
215: Equation~(\ref{barotropic}), though neglecting detailed
216: cooling and heating processes, was shown to reproduce to a first approximation 
217: the radial temperature gradients in the disk \citep{VB3} and the density-temperature relation 
218: for collapsing cloud cores derived by \citet{Masunaga} using a detailed radiation hydrodynamics 
219: simulation \citep{VB2}.
220: 
221: The vertical scale height $Z(r,\phi,t)$ is determined assuming the local hydrostatic 
222: equilibrium in the gravitational field of a disk and central star \citep{VB4}.
223: The relevant formulas are given in the Appendix.
224: 
225: \subsection{Viscous numerical approach}
226: \label{viscmodel}
227: Viscosity is another important mechanism of angular momentum 
228: and mass redistribution in astrophysical disks. Most analytical and numerical studies 
229: of viscous evolution of thin circumstellar disks have employed 
230: the standard axisymmteric model of \citet{Lyndenbell}, in which 
231: the surface density of a Keplerian disk evolves with time according to the following diffusion equation
232: \begin{equation}
233: {\partial \Sigma \over \partial t} = {3\over r} {\partial \over \partial r} 
234: \left[ r^{1/2} {\partial \over \partial r} ( \nu \Sigma r^{1/2}) \right],
235: \end{equation}
236: where $\nu$ is the kinematic viscosity. 
237: 
238: In the present paper, we take a more fundamental approach and describe the effect of (yet unspecified)
239: viscosity in terms of the classic viscous stress tensor 
240: \begin{equation}
241: \mathbf{\Pi}=2 \mu \left( \nabla v - {1 \over 3} (\nabla \cdot v) \mathbf{e} \right),
242: \end{equation}
243: where $\nabla v$ is a symmetrized velocity gradient tensor, $\mathbf{e}$ is the unit tensor, and
244: $\mu$ is the dynamical viscosity.
245: This approach allows for a 
246: self-consistent treatment of both, self-gravity and viscosity, within the same numerical formalism.
247: The resulting mass and momentum transport equations in the viscous numerical approach are
248: \begin{eqnarray}
249: \label{visc1}
250:  \frac{{\partial \Sigma }}{{\partial t}} & = & - \nabla _p  \cdot \left( \Sigma \bl{v}_p 
251: \right), \\ 
252: \label{visc2}
253:  \Sigma \frac{d \bl{v}_p }{d t}  & = &  - \nabla _p {\cal P}  + \Sigma \, \bl{g}_p
254:  + \left( \nabla \cdot \mathbf{\Pi}\right)_p \, ,
255: \end{eqnarray}
256: where $\nabla \cdot \mathbf{\Pi}$ is the divergence of the rank-two viscous stress tensor 
257: $\mathbf{\Pi}$. The relevant components of $\left( \nabla \cdot \mathbf{\Pi} \right)_p$ 
258: are given in the Appendix.
259: Equations~(\ref{visc1}) and (\ref{visc2}) are closed with the barotropic equation of state~(\ref{barotropic}).
260: We note that the viscous approach accounts self-consistently for both the gravitational and 
261: viscous torques which may arise during numerical simulations.
262: 
263: It is evident that the practical application of equation~(\ref{visc2}) requires a knowledge
264: of the dynamical viscosity $\mu$ of the disk.
265: Unfortunately, our understanding of viscous processes in circumstellar disks is still 
266: incomplete. We know that molecular (collisional) viscosity is most certainly too low to be of 
267: practical interest. Turbulence driven by the magneto-rotational instability (MRI) 
268: is a most promising source of viscosity at present \citep{BH}, though other
269: mechanisms cannot be ruled out completely.
270: 
271: In this paper, we make no specific assumptions as to the source of viscosity and
272: define the coefficient of dynamical viscosity using the usual $\alpha$-prescription
273: of \citet{SS}
274: \begin{equation}
275: \mu=\alpha \, \Sigma \, \tilde{c}_s \, Z,
276: \end{equation} 
277: where spatially and temporally uniform $\alpha$ is set equal to 0.01 and 
278: $\tilde{c}_s=\sqrt{\partial {\cal P}/ \partial \Sigma}$ is the effective
279: sound speed. Our numerical simulations of embedded and T Tauri disks
280: indicate that $\alpha\simeq 0.001-0.01$ yields disk sizes and radial slopes 
281: that are in general agrement with observations \citep{VB4}. Lower values of $\alpha$ 
282: ($ < 10^{-3}$) have little effect on the evolution of self-gravitating disks, 
283: whereas substantially higher values ($\ga 0.1$) quickly destroy the disks.
284: \citet{Hartmann98} also predicted similar values for $\alpha$ by
285: analyzing accretion rates in T Tauri disks.
286: Since viscosity in our  model is assumed to arise due to some physical processes 
287: {\it in the disk}, we keep $\alpha$ equal to zero during the early ``pre-disk'' phase of 
288: evolution and set $\alpha$ equal to 0.01 only when a circumstellar disk forms
289: around a central star. 
290: 
291: 
292: 
293: 
294: \subsection{Initial condition}
295: 
296: The initial radial distributions of surface density $\Sigma$ and angular velocity $\Omega$ 
297: in our model cloud cores
298: are those characteristic of a collapsing axisymmetric magnetically
299: supercritical core \citep{Basu}
300: \begin{equation}
301: \Sigma={r_0 \Sigma_0 \over \sqrt{r^2+r_0^2}}\:,
302: \label{dens}
303: \end{equation}
304: \begin{equation}
305: \Omega=2\Omega_0 \left( {r_0\over r}\right)^2 \left[\sqrt{1+\left({r\over r_0}\right)^2
306: } -1\right],
307: \end{equation}
308: where $r_0$ is the radial scale length defined as $r_0 = k c_s^2 /(G\Sigma_0)$ and  $k= \sqrt{2}/\pi$.
309: These initial profiles are characterized by the important
310: dimensionless free parameter $\eta \equiv  \Omega_0^2r_0^2/c_s^2$
311: and have the property 
312: that the asymptotic ($r \gg r_0$) ratio of centrifugal to gravitational
313: acceleration has magnitude $\sqrt{2}\,\eta$ \citep[see][]{Basu}. 
314: The centrifugal radius of a mass shell initially located at radius $r$ is estimated to be
315: $r_{\rm cf} = j^2/(Gm) = \sqrt{2}\, \eta r$, where $j=\Omega r^2$ is the specific angular
316:  momentum. Since the enclosed mass $m$ is a linear function of $r$ at large radii,
317: this also means that $r_{\rm cf} \propto m$.
318: The gas has a mean molecular mass $2.33 \, m_{\rm H}$ and cloud cores are initially
319: isothermal with temperature $T=10$ K. 
320: 
321: 
322: 
323: We present results from three sets of models, each with a different value of $\eta$.  
324: The standard model has $\eta = \eta_1 = 1.2 \times 10^{-3}$ based on typical values
325: $c_s = 0.19$ km s$^{-1}$, $\Sigma_0 = 0.12$~g~cm$^{-2}$, and
326: $\Omega_0 = 1.0$~km~s$^{-1}$~pc$^{-1}$. The outer radius is taken to
327: be $r_{\rm out} = 0.04$ pc, and the total cloud mass is $0.8\,\msun$. 
328: Other models with $\eta = \eta_1$ but different mass (outer radius) 
329: are generated by varying $r_0$ and $\Omega_0$ so that their product is constant. 
330: Note that, when $r_0$ is varied, $\Sigma_0$ has to be changed accordingly.
331: All clouds are characterized by the same ratio $r_{\rm out}/r_0\approx 6.0$. 
332: To generate the second set of models, $\eta = \eta_2 = 2.3 \times 10^{-3}$, we set
333: $\Omega_0 = 1.4$~km~s$^{-1}$~pc$^{-1}$ and all other quantities the
334: same as in the standard model with $\eta= \eta_1$. Models of
335: varying mass are then generated in the same manner as for the
336: $\eta_1$ models. The third set of models, with $\eta=\eta_3
337: = 3.4 \times 10^{-3}$, are also obtained in this way,
338: by first using $\Omega_0 = 1.7$~km~s$^{-1}$~pc$^{-1}$.
339: Overall, there are 7 models with $\eta = \eta_1$, 13 models with
340: $\eta = \eta_2 \simeq 2\, \eta_1$, 
341: and 12 with $\eta = \eta_3 \simeq 3\, \eta_1$.
342: The range of initial cloud masses ($M_{\rm cl}$) amongst our models is $0.3\,\msun-2.95\,\msun$.
343: The parameters of our models are listed in Table~\ref{table1} ($\eta_1$), Table~\ref{table2}
344: ($\eta_2$), and Table~\ref{table3} ($\eta_3$).
345: We note that our model values of $\Omega_0 = (1.0-1.7)$~km~s$^{-1}$~pc$^{-1}$ are within a typical 
346: range of velocity gradients measured in dense starless cores by \citet{Caselli}.
347: 
348: 
349: \begin{table}
350: \begin{center}
351: \caption{Parameters of models with $\eta_1=1.2 \times 10^{-3}$
352: \label{table1}}
353: %\vspace{3 pt}
354: \begin{tabular}{clllll}
355: \hline\hline
356: Model & $r_0$ & $\Sigma_0$ & $\Omega_0$ & $r_{\rm out}$ & $M_{\rm cl}$  \\
357: \hline
358:  1 & 1209 & 0.13  & 1.14 & 7186  & 0.7 \\
359:  2 & 1382 & 0.12  & 1.0  & 8213  & 0.8 \\
360:  3 & 1728 & 0.093 & 0.8  & 10266 & 0.98 \\
361:  4 & 2074 & 0.077 & 0.67 & 12320 & 1.18 \\
362:  5 & 2937 & 0.055 & 0.47 & 17452 & 1.67 \\
363:  6 & 4147 & 0.039 & 0.33 & 24640 & 2.36 \\
364:  7 & 5184 & 0.031 & 0.27 & 30800 & 2.95 \\
365:  \hline
366: \end{tabular} 
367: \tablecomments{All distances are in AU, angular
368: velocities in km~s$^{-1}$~pc$^{-1}$, surface densities in g~cm$^{-2}$, and masses in $M_\odot$.}
369: \end{center}
370: \end{table} 
371: 
372: \begin{table}
373: \begin{center}
374: \caption{Parameters of models with $\eta_2=2.3 \times 10^{-3}$
375: \label{table2}}
376: %\vspace{3 pt}
377: \begin{tabular}{clllll}
378: \hline\hline
379: Model & $r_0$ &  $\Sigma_0$ & $\Omega_0$ & $r_{\rm out}$ & $M_{\rm cl}$  \\
380: \hline
381:  8 & 622  & 0.26  & 3.1  & 3696  & 0.35 \\
382:  9 & 691  & 0.23  & 2.8  & 4106  & 0.4 \\
383:  10 & 864  & 0.19  & 2.24 & 5133  & 0.5 \\
384:  11 & 1037 & 0.16  & 1.87 & 6160  & 0.6 \\
385:  12 & 1210 & 0.13  & 1.6  & 7187  & 0.7 \\
386:  13 & 1417 & 0.11  & 1.4  & 8213  & 0.8  \\
387:  14 & 1728 & 0.093 & 1.12 & 10267 & 0.98 \\
388:  15 & 2454 & 0.065 & 0.79 & 14579 & 1.4  \\
389:  16 & 2765 & 0.058 & 0.7  & 16428 & 1.57 \\
390:  17& 3456 & 0.046 & 0.56 & 20533 & 1.97 \\
391:  18& 3802 & 0.042 & 0.51 & 22587 & 2.16 \\
392:  19& 4147 & 0.038 & 0.47 & 24640 & 2.36 \\
393:  20& 4838 & 0.033 & 0.4  & 28747 & 2.75 \\
394: % 14& 5530 & 0.029 & 0.35 & 32853 & 3.15 \\
395:  \hline
396: \end{tabular} 
397: \tablecomments{All distances are in AU, angular
398: velocities in km~s$^{-1}$~pc$^{-1}$, surface densities in g~cm$^{-2}$, and masses in $M_\odot$.}
399: \end{center}
400: \end{table} 
401: 
402: 
403: \begin{table}
404: \begin{center}
405: \caption{Parameters of models with $\eta_2=3.4 \times 10^{-3}$
406: \label{table3}}
407: %\vspace{3 pt}
408: \begin{tabular}{clllll}
409: \hline\hline
410: Model & $r_0$ & $\Sigma_0$ & $\Omega_0$ & $r_{\rm out}$ & $M_{\rm cl}$  \\
411: \hline
412:  21 & 518 & 0.31  & 4.53  & 3080  & 0.3 \\
413:  22 & 691  & 0.23  & 3.4  & 4106  & 0.4 \\
414:  23 & 864  & 0.19  & 2.7  & 5133  & 0.5 \\
415:  24 & 1037 & 0.16  & 2.26 & 6160  & 0.6 \\
416:  25 & 1210 & 0.13  & 1.94 & 7187  & 0.7 \\
417:  26 & 1417 & 0.11  & 1.7  & 8213  & 0.8  \\
418:  27 & 2073 & 0.077 & 1.13 & 12320 & 1.18 \\
419:  28 & 2420 & 0.066 & 0.97 & 14373 & 1.38 \\
420:  29 & 3283 & 0.049 & 0.72 & 19506 & 1.87  \\
421:  30 & 3802 & 0.042 & 0.62  & 22587 & 2.16 \\
422:  31 & 4147 & 0.039 & 0.57 & 24640 & 2.36 \\
423:  32 & 4493 & 0.036 & 0.52 & 26693 & 2.56 \\
424: % 32& 5530 & 0.029 & 0.43 & 32853 & 3.15 \\
425: % 13& 4838 & 0.033 & 0.4  & 28747 & 2.75 \\
426: % 14& 5530 & 0.029 & 0.35 & 32853 & 3.15 \\
427:  \hline
428: \end{tabular} 
429: \tablecomments{All distances are in AU, angular
430: velocities in km~s$^{-1}$~pc$^{-1}$, surface densities in g~cm$^{-2}$, and masses in $M_\odot$.}
431: \end{center}
432: \end{table} 
433: 
434: 
435: \subsection{Numerical technique}
436:  
437: Hydrodynamic equations of the basic and viscous numerical models are solved in polar 
438: coordinates $(r, \phi)$ on a numerical grid with
439: $128 \times 128$ points. We use the method of finite differences with a time-explicit,
440: operator-split solution procedure. Advection is
441: performed using the second-order van Leer scheme.  The radial points are logarithmically spaced.
442: The innermost grid point is located at $r=5$~AU, and the size of the 
443: first adjacent cell lies in the range between 0.26~AU (model~21) 
444: and 0.36~AU (model~7). It means that the ratio $\Delta r/r$ of the cell size $\Delta r$ 
445: to radius $r$ is constant for a given cloud core and varies from $0.05$ (model~21) to 
446: $0.07$ (model~7). 
447: %It can be noticed from Fig.~\ref{figA1} that the aspect ratio $A=z/r$
448: %of the disk vertical scale height to radius is smaller than $\Delta r/r$ in the inner 70~AU, 
449: %which implies that a better radial (and angular) resolution is desirable in this inner region. 
450: %We have found that an increase in the resolution to $256 \times 256$ grid points makes 
451: %little influence on the accretion history and disk masses but helps save a considerable amount 
452: %of CPU time and, eventually, consider many more model cloud cores. We also point out that, due to
453: %the adopted log spacing in the radial direction, the resolution in the inner computational
454: %region is sufficient to fulfill the Truelove criterion in the disk.
455: 
456:  We introduce a ``sink cell'' at $r<5$~AU, 
457: which represents the central star plus some circumstellar disk material, 
458: and impose a free inflow inner boundary condition. About 95 per cent of 
459: the material crossing the inner boundary lands into the star, the
460: rest constitutes an inner circumstellar region of constant surface density.     
461: The dynamics of the inner region ($r<5$~AU) is not computed but it contributes 
462: to the total gravitational potential of the system. 
463: We do not account for a possible mass loss from the sink cell due to stellar jets, 
464: since our model is two-dimensional and
465: it is not clear how the mass ejection efficiency varies with stellar age. 
466: The outer boundary is such that the cloud has a constant mass and volume. 
467: More details on numerical techniques and relevant tests are given in \citet{VB2}.
468: 
469: 
470: 
471: 
472: \section{Masses of self-gravitating disks}
473: \label{SG}
474: In this section we present disk masses obtained using the basic numerical approach
475: described in \S~\ref{basic}. In the framework of this numerical model, disk masses 
476: are controlled by the rate of mass infall from 
477: the surrounding envelope and the radial transport of mass and angular momentum 
478: due to gravitational torques. No viscous torques are present in this case.
479: 
480: An accurate determination of disk masses in numerical simulations of collapsing cloud cores is not
481: a trivial task. Self-consistently formed circumstellar disks have a wide range of masses 
482: and sizes, which are not known {\it a priori}. However, numerical and observational 
483: studies of circumstellar disks indicate that the disk
484: surface density is a declining function of radius. Therefore, we distinguish
485: between disks and infalling envelopes using a critical surface density for the 
486: disk-to-envelope transition, for which we choose a value of $\Sigma_{\rm tr}=0.1$~g~cm$^{-2}$.
487: This choice is dictated by the fact that densest {\it starless} cores have surface
488: densities only slightly lower than the adopted value of $\Sigma_{\rm tr}$.
489: In addition, our numerical simulations indicate that self-gravitating disks have 
490: sharp outer edges and the gas densities of order $0.01-0.1$~g~cm$^{-2}$ 
491: characterize a typical transition region between the disk and envelope \citep{VB3}.
492: 
493: 
494: 
495: To compare disk masses in three distinct phases of stellar evolution
496: we need an evolutionary indicator to distinguish between Class 0, Class I, 
497: and Class II phases. 
498: We use a classification of \citet{Andre}, who
499: suggest that the transition between Class 0 and Class I objects occurs when
500: about $50\%$ of the initial cloud core is accreted onto the protostar/disk system.
501: The Class II phase is consequently defined by the time when the
502: infalling envelope clears and its total mass drops below 10\% of the initial cloud 
503: core mass $M_{\rm cl}$.
504: It should be mentioned here that there exists no unique classification scheme for protostars.
505: For instance, \citet{VB0} have proposed a classification scheme that hinges 
506: on a temporal behaviour 
507: of bolometric luminosity $L_{\rm bol}$. They identify Class~0 phase with a period of 
508: temporally increasing $L_{\rm bol}$ and Class~I phase with a later period of decreasing 
509: $L_{\rm bol}$. The peak in $L_{\rm bol}$ corresponds to the evolutionary time when 
510: $50 \pm 15$ per cent of the cloud core mass has been accreted by the protostar.
511: Observers prefer to classify protostars by their spectral energy distributions. 
512: For instance, \citet{Lada87} and \citet{Andrews} use the values of power-law index $n$ 
513: (defined by $\nu F_{\nu} \propto \nu^n$, where $F_{\nu}$ is the infrared flux density at frequency 
514: $\nu$) to distinguish between Class~I and Class~II objects. 
515: \citet{Andre} use the ratio of submillimeter to bolometric luminosity
516: $L_{\rm submm}/L_{\rm bol}$ for the same purposes. Although these classifications are
517: physically related, it is possible that they may differ from our adopted classification, 
518: especially if some of the gas is removed
519: from the cloud core and does not accrete. We acknowledge that the difference 
520: in the existing classification schemes may systematically shift our results. 
521: 
522: 
523: \subsection{Temporal evolution of self-gravitating disks}
524: \label{tempevol}
525: We start by comparing the long-term evolution of disk masses in four sample models, which are 
526: chosen to represent a full spectrum of the initial cloud core 
527: masses with a constant ratio $\eta=2.3\times 10^{-3}$ of centrifugal to gravitational 
528: acceleration. Figure~\ref{fig1} shows the disk masses (thick solid lines), stellar masses (thin
529: solid lines), and envelope masses (dashed lines) in model~8 (upper left, $M_{\rm cl}=0.35~M_\odot$),
530: model~11 (upper right, $M_{\rm cl}=0.6~M_\odot$), model~16 (lower left, $M_{\rm cl}=1.57~M_\odot$),
531: and model~20 (lower right, $M_{\rm cl}=2.75~M_\odot$). The horizontal axis is the time elapsed 
532: since the formation of a central star.
533: All masses are calculated relative to the corresponding 
534: initial cloud core mass $M_{\rm cl}$. The vertical dotted lines mark the onset 
535: of Class I (left)  and Class~II (right) phases.
536: 
537: \begin{figure}
538:   \resizebox{\hsize}{!}{\includegraphics{f1.eps}}
539: % \centering
540: %  \includegraphics{figure1.eps}
541:       \caption{Disk masses (thick solid lines), stellar masses (thin solid lines),
542:       and envelope masses (dashed lines) obtained in the framework of 
543:       the basic numerical approach (see \S~\ref{basic}) 
544:       for model~8 (upper left), model~11 (upper right), 
545:       model~16 (lower left), and model~20 (lower right).
546:       All masses are calculated relative to the corresponding 
547:       initial cloud core mass $M_{\rm cl}$. The horizontal axis shows time elapsed since the 
548:       formation of a central star. The left/right vertical dotted lines mark the onset 
549:       of Class I/II phases, respectively.}
550:          \label{fig1}
551: \end{figure}
552: 
553: Our numerical simulations demonstrate that cloud cores with constant $\eta$ (but
554: different size) form disks roughly at the same physical time after the formation 
555: of a central star but in distinct stellar evolution phases. For instance,
556: model~8 starts to build a disk in the late Class I phase, while model~20 does that in the midst
557: of Class 0 phase. Cloud cores of greater mass tend to form disks in the earlier phase of stellar 
558: evolution than their low-mass counterparts. 
559: 
560: The upper panel of Fig.~\ref{fig1} identifies cases when Class~0 stars have no disks
561: associated with them.  
562: We emphasize here that due to the use of the sink cell in our numerical code
563: we can resolve only those disks whose outer radii are larger than 5~AU. It means that even though 
564: circumstellar disks do not form around some Class 0 objects in our numerical simulations, 
565: as in models~8 and 11 
566: (top panels in Fig.~\ref{fig3}), one may suppose that such disks still exist  
567: but their size is simply smaller than that of the sink cell. 
568: Our test runs with a sink cell set to 0.5~AU (instead of 5~AU) confirm
569: that disks indeed forms earlier in the evolution but quickly expands to 5~AU and beyond.
570: Hence, the absence of disks around some Class~0 objects may be to some extent 
571: caused by a finite-size sink cell. However, it is still possible that some Class~0 
572: objects have no disks associated with them. This is particularly true for those objects that 
573: develop from cloud cores with low rates of rotation, since in this case
574: a (substantial) portion of the disk material will be characterized
575: by the centrifugal radius (estimated as $r_{\rm cf}=j^2/(Gm)$) that is
576: smaller than the radius of the protostar, $4~R_\odot$.
577: A numerical code with realistic treatment of the protostar formation is needed
578: to accurately address this issue.
579: 
580: 
581: Another interesting feature of self-gravitating disks that can be seen in Fig.~\ref{fig1} 
582: is that the disk-to-star mass 
583: ratio never exceeds some characteristic value, approximately $0.35-0.4$, irrespective of the
584: initial cloud core mass. This is rather counterintuitive. For instance, models~16 and 20 (bottom panels
585: in Fig.~\ref{fig1}) form disks in the Class 0 phase, which is characterized by envelope 
586: masses that are considerably greater than those of the protostars. As a consequence,
587: one may expect the formation of a disk with mass that is at least comparable 
588: to or even greater than that of a protostar. It turns out, however, that circumstellar disks 
589: that form in the {\it Class~0 or early Class~I phase} around stars with mass 
590: $M_\ast \ga 0.6~M_\odot$ develop vigourous gravitational 
591: instability -- a very efficient means of inward mass transport 
592: that helps keep the disk mass well below that of the protostar. 
593: Sharp drops in the disk mass (or equivalent surges in the stellar mass) 
594: seen in the upper right and bottom panels of Figure~\ref{fig1} are a manifestation of 
595: this process \citep[see e.g.][]{VB1,VB2}.  
596: \citet{Kratter} also predict that disks around stars with mass greater than $1.0~M_\odot$
597: are expected to be vigorously gravitationally unstable in the early embedded phase of
598: disk evolution. 
599: The late evolution phase (Class II) sees only an insignificant decline of disk mass with time. 
600: 
601: 
602: \begin{figure}
603:   \resizebox{\hsize}{!}{\includegraphics{f2.eps}}
604:       \caption{Time-averaged disk masses (right column) and
605:       time-averaged disk-to-star mass ratios (left column) versus time-averaged stellar masses. 
606:       Open triangles, filled squares, and plus signs show the data for Class~0, Class~I, 
607:       and Class~II 
608:       systems, respectively. In particular, top panels show the data for models 
609:       with $\eta_1 = 1.2 \times 10^{-3}$, whereas middle and bottom panels show the data for 
610:       models with $\eta_2\simeq 2\eta_1$ and $\eta_3\simeq 3\eta_1$, respectively.}
611:          \label{fig2}
612: \end{figure}
613: 
614: 
615: \subsection{Disk masses versus stellar masses}
616: We use 32 models, the parameters of which are listed in 
617: Tables~\ref{table1}-\ref{table3}, to analyse the statistical relations between time-averaged 
618: disk and stellar masses in the Class 0/I/II evolution phases. 
619: Time averaging is performed separately for each evolution phase over the duration of the phase.
620: The age of the oldest disk in our sample 
621: is about 3~Myr. Since disk masses in the Class II phase have a tendency to decline with 
622: time and the actual
623: lifetime of Class II objects may be longer, we expect that
624: we might have somewhat overestimated disk masses around Class II objects. This, obviously, does not
625: effect our estimates of Class 0/I disk masses. 
626: 
627: Figure~\ref{fig2} shows time-averaged disk masses (right column) and time-averaged 
628: disk-to-star mass ratios (left column) versus time-averaged stellar masses 
629: for Class~0 (open triangles), Class~I (filled squares) and Class~II  objects (plus signs).
630:  Several interesting conclusions can be drawn 
631: by analysing the figure. 
632: \begin{enumerate}
633:  \item Time-averaged disk and stellar masses ($\langle M_{\rm d}\rangle$ and 
634:  $\langle M_\ast \rangle$, 
635:  respectively) in models with the same ratio of rotational to gravitational 
636: acceleration fall onto a unique evolutionary track in the $\langle M_{\rm d} \rangle - \langle M_\ast
637: \rangle$  diagram. 
638: \item  Class 0 objects occupy the lower-left part of each evolutionary track. No stars with 
639: $\langle M_\ast \rangle \ga 0.9~M_\odot$ have Class 0 disks (but they have Class~I/II disks). 
640: In other words, only low-mass stars (within our range of interest, 
641: $0.2~M_\odot \la \langle M_\ast \rangle  \la 2.0~M_\odot$) can harbour Class~0 disks.
642: \item Stars of {\it equal mass} have disks with similar masses, regardless of the stellar 
643: evolution phases. 
644: \item In each stellar evolution phase, disk-to-star mass ratios tend to have greater values 
645: for stars of greater mass. However, there is a clear saturation effect for stars with
646: masses greater than $1.0~M_\odot$ -- disk masses never grow above $40\%$ 
647: of the stellar masses, even for models with the largest values of $\eta = \eta_3=3.4\times 10^{-3}$.
648: The saturation of disk-to-star mass ratios is caused by the onset of vigorous
649: gravitational instability in circumstellar disks that form in the Class~0 and early Class~I phase.
650: \end{enumerate}
651: 
652: 
653: \begin{table*}
654: \begin{center}
655: \caption{Summary of disk properties}
656: \label{table4}
657: %\vspace{3 pt}
658: \begin{tabular}{ccccccc}
659: \hline
660: \hline
661: disk type & $ \!\!\! \overline{M}_{\rm d,C0}$ & $\!\!\! \overline{M}_{\rm d,CI}$ & $\!\!\! \overline{M}_{\rm d,CII}$ & 
662: $\!\!\! \left( \langle M^{\rm min}_{\rm d,C0}\rangle : \langle M^{\rm max}_{\rm d,C0} \rangle \right)$ & 
663: $ \!\!\! \left( \langle M^{\rm min}_{\rm d,CI} \rangle : \langle M^{\rm max}_{\rm d,CI} \rangle \right) $
664:  &  $ \!\!\! \left( \langle M^{\rm min}_{\rm d,CII} \rangle : \langle M^{\rm max}_{\rm d,CII} 
665:  \rangle \right)$ \\
666: \hline
667: self-gravitating& 0.09 & 0.10 & 0.12  & $(0.04 : 0.2)$ & $(0.02 : 0.5)$  & $(0.03 : 0.7)$ \\
668: viscous & 0.10 & 0.11 & 0.06  & $(0.05 : 0.19)$  & $(0.02 : 0.55)$  & $(0.004 : 0.6)$ \\
669:  \hline
670: \end{tabular} 
671: \tablecomments{Mean masses $\overline{M}_{\rm d,C0}$, $\overline{M}_{\rm d,CI}$, and 
672: $\overline{M}_{\rm d,CII}$ of Class 0, Class I, and Class II disks, respectively, are
673: calculated according to equation~(\ref{mean}). Angle brackets $\langle \, \rangle$ denote
674: time averaging over the duration of a particular stellar evolution phase. 
675: Indices ``min'' and ``max'' refer to minimum and maximum time-averaged disk masses, respectively. 
676: All disk masses are in $M_\odot$.}
677: \end{center}
678: \end{table*} 
679: 
680: 
681: 
682: We summarize the main properties of our model self-gravitating disks in Table~\ref{table4}.
683: In order to facilitate the comparison of our numerical results with 
684: observations, we calculate mean disk masses (in each evolution phase) 
685: by weighting the time-averaged disk masses according to the corresponding stellar masses.
686: The relative importance of the stellar masses is calculated using the following power-law 
687: weight function
688: \begin{equation}
689: {\cal F}_{\rm K}\left(\langle M_\ast \rangle\right) = \left\{ \begin{array}{ll} 
690:    A \, \langle M_\ast \rangle^{-1.3} & \,\, 
691:    \mbox{if $\langle M_\ast \rangle \le 0.5~M_\odot$ } \\ 
692:    B \, \langle M_\ast \rangle^{-2.3} & \,\, 
693:    \mbox{if $\langle M_\ast \rangle > 0.5~M_\odot$ }.  \end{array} 
694:    \right. 
695:    \label{function}  
696:  \end{equation}
697: The slope of ${\cal F}_{\rm K}\left(\langle M_\ast \rangle\right)$ is typical for 
698: the Kroupa initial mass function (IMF) of stars \citep{Kroupa01}. 
699: The mean disk masses are then calculated as follows
700: \begin{equation}
701: \overline{M}_{\rm \rm d,ph} = { \sum^N_{i=1} \langle M^i_{\rm d,ph} \rangle \, \,
702: {\cal F}_{\rm K}\left(\langle M^i_{\rm \ast,ph} \rangle\right) \over 
703: \sum^N_{i=1} {\cal F}_{\rm K} \left(\langle M^i_{\rm \ast,ph} \rangle\right) },
704: \label{mean}
705: \end{equation}
706: where $\langle M^i_{\rm d,ph} \rangle$ and $\langle M^i_{\rm \ast,ph} \rangle$ are the 
707: time-averaged disk and stellar masses in the $i$-th model (see Fig.~\ref{fig2})  
708: and index ``ph'' stands for Class 0 (C0), Class I (CI), or 
709: Class II (CII) stellar evolution phases. The summation is performed separately for each 
710: evolution phase. 
711: In particular, there are $N=20$ models that have Class~0 disks and $N=31$ models that have
712: Class I disks. All 32 models have Class II disks, of course. 
713: The ratio of the  normalization constants $A/B=2$ needed to evaluate 
714: equation~(\ref{mean}) is found using the continuity condition 
715: at $\langle M_\ast\rangle=0.5~M_\odot$.
716: 
717: %The slope of the initial mass function
718: %of stars is set to 1.3 in the $0.08-0.5~M_\odot$ range and 2.3 for stars more massive than 
719: %$0.5~M_\odot$ \citep{Kroupa01}. 
720: 
721: 
722: Table~\ref{table4} indicates that the mean masses slightly increase along the stellar evolution
723: sequence from $\overline{M}_{\rm d,C0}=0.09~M_\odot$ around Class~0 objects to 
724: $\overline{M}_{\rm d,CII}=0.12~M_\odot$ around Class~II objects.
725: The minimum and maximum time-averaged disk masses 
726: ($\langle M^{\rm min}_{\rm d,ph} \rangle$ and $\langle M^{\rm max}_{\rm d,ph} \rangle$, 
727: respectively) 
728: in each evolution phase are shown in the last three columns of Table~\ref{table4}. It is 
729: seen that both the Class I and Class II disks have a similar range of masses, while 
730: Class 0 disks have a somewhat narrower range. The smallest time-averaged mass of a 
731: self-gravitating disk found in our numerical simulations is $\langle M_{\rm d} \rangle = 0.02~M_\odot$.
732: 
733: 
734:  
735: 
736: 
737: 
738: \section{Masses of viscous disks}
739: \label{VD}
740: In this section we present disk masses obtained in the framework of the viscous numerical approach
741: described in \S~\ref{viscmodel}. We seek to determine the effect of viscosity,
742: and associated viscous torques, on the masses of self-consistently formed circumstellar disks.
743: Our comparison study of viscous versus self-gravitating disks have shown that
744: viscous disks may lack a sharp transition boundary between the disk and the envelope \citep{VB4}.
745: This smearing of the disk's outer physical boundary is particularly pronounced in 
746: disks of low mass and in the late Class~II phase.
747: In this situation, it is somewhat arbitrary to adopt a value of $\Sigma_{\rm tr}=0.1$~g~cm$^{-2}$
748: for the disk-to-envelope transition. Nevertheless, we have recalculated disk
749: masses with $\Sigma_{\rm tr}=0.01$~g~cm$^{-2}$ and found that this can increase 
750: masses of Class~II disks by $25\%$ at most. Class 0/I disks are unaffected by the value of 
751: $\Sigma_{\rm tr}$, since they usually have sharp outer physical boundaries.
752: 
753: 
754: \begin{figure}
755:   \resizebox{\hsize}{!}{\includegraphics{f3.eps}}
756:       \caption{The same as Figure~\ref{fig1} but obtained in the framework of the viscous 
757:       numerical approach described in \S~\ref{viscmodel}}
758:          \label{fig3}
759: \end{figure}
760: 
761: 
762: \subsection{Temporal evolution of viscous disks}
763: In order to investigate the temporal evolution of viscous disks, we consider the same 
764: four sample model cloud cores as in \S~\ref{tempevol}. Figure~\ref{fig3}
765: presents  disk masses (thick solid lines), stellar masses (thin solid lines),
766: and envelope masses (dashed lines) in model~8 (upper left), model~11 (upper right), model~16 (lower
767: left), and model~20 (lower right). The horizontal axis is time elapsed since the formation of a central
768: star. All masses are calculated relative to the corresponding 
769: initial cloud core mass $M_{\rm cl}$.
770: The vertical dotted lines mark the onset of Class I phase (left line) and Class II phase (right line).
771: 
772: 
773: The comparison of Figures~\ref{fig1} and \ref{fig3} reveals a major difference between self-gravitating
774: and viscous disks  -- the latter have 
775: considerably smaller masses in the {\it late} Class II phase than the former. 
776: For instance, model~8 in Figure~\ref{fig1} has a final ($t=5$~Myr) stellar mass 
777: $M_\ast = 0.33~M_\odot$ and disk mass   
778: $M_{\rm d}=2.5\times10^{-2}~M_\odot$, while the same model in Figure~\ref{fig3} has  
779: $M_{\rm d}=8.5\times 10^{-4}~M_\odot$.
780: Such a large contrast, however, diminishes for stars of greater mass. For instance, model~16
781: in Figure~\ref{fig1} has $M_\ast=1.57~M_\odot$ and $M_{\rm d}=0.37~M_\odot$, while the same model 
782: in Figure~\ref{fig3} has $M_{\rm d}=0.11~M_\odot$.
783: On the other hand, both the viscous and self-gravitating disks have comparable masses in the 
784: Class 0/I phases. Even the early Class II phase sees little difference between the masses 
785: of viscous and self-gravitating disks, provided that the disks are formed from 
786: molecular cloud cores of similar mass.
787: Figure~\ref{fig3} indicates that a small portion of the viscous disk material 
788: returns to the envelope in the late disk evolution. This is because we have imposed a fixed 
789: gas column density threshold for the disk-to-envelope transition, but
790: viscous disks expand in the late evolution due to the action of viscous torques, 
791: which are in fact positive in the outer disk regions \citep{VB4}. This numerical 
792: effect may slightly reduce the resulted viscous disk masses in the Class II phase.  
793: 
794: 
795: 
796: \begin{figure}
797:   \resizebox{\hsize}{!}{\includegraphics{f4.eps}}
798:       \caption{Total negative gravitational torques (solid lines) and total negative
799:       viscous torques (dashed lines) versus time elapsed since the formation of a central star
800:       in model~8 (upper left), model~11 (upper right), model~16 (lower left), and model~20 
801:       (lower right). The vertical dotted lines mark the onset of Class I~phase (left line) 
802:       and Class~II phase (right line). The torques are in units of $8.7\times 10^{40}$
803:       ~g~cm$^{2}$~s$^{-2}$.}
804:          \label{fig4}
805: \end{figure}
806: 
807: 
808: Why does the late stellar evolution phase reveal such a noticeable difference in the 
809: disk masses (and, consequently, in disk-to-star mass ratios) between viscous and self-gravitating 
810: disks and why is there little difference in the earlier phases? The reason for that can be understood
811: if we consider the temporal evolution of viscous and gravitational torques in the disk.
812: Figure~\ref{fig4} shows total negative torques due to disk self-gravity (${\cal T}_{\rm g}$,
813: solid lines) and viscosity (${\cal T}_{\rm v}$, dashed lines) in our four typical models:
814: model~8 (upper left), model~11 (upper right), model~16 (lower left), and model~20 (lower right).
815: The horizontal axis shows time elapsed since the formation of a central star.
816: The total negative torque is computed by summing up local negative torques $\tau(r,\phi)$ 
817: in each computational zone occupied by the disk. In particular, ${\cal T}_{\rm g}$ is computed as 
818: a sum of $\tau_{\rm g}(r,\phi) = -m(r,\phi) \, \partial \Phi(r,\phi) / \partial \phi$, where $m(r,\phi)$
819: and  $\Phi(r,\phi)$ are the gas mass and gravitational potential in a cell with polar coordinates $(r,\phi)$,
820: respectively. The total negative viscous torque is calculated in a similar manner by summing up 
821: all local negative viscous torques $\tau_{\rm v}(r,\phi)=r\, (\nabla \cdot \mathbf{\Pi})_\phi  
822: \, S(r,\phi)$, where $S(r,\phi)$ is the surface area occupied by a cell with polar coordinates ($r,\phi$).
823: The local torques $\tau_{g}(r,\phi)$ and $\tau_{\rm v}(r,\phi)$ are actually the local 
824: azimuthal components of the corresponding 
825: forces, acting on a fluid element with mass $m(r,\phi)$, multiplied by the arm $r$.
826: 
827: Figure~\ref{fig4} indicates that gravitational torques always prevail over viscous 
828: torques in the Class 0 and
829: Class I phases. As a result, the disk mass in these phases is determined by
830: the interplay between the mass load from an infalling envelope and the rate of mass
831: and angular momentum transport in the disk due to gravitational torques. In the Class II phase,
832: the strength of gravitational torques diminishes, partly due to a stabilizing influence of a growing
833: central star and partly due to the exhausted mass reservoir in the infalling envelope.
834: As a result, viscous torques, the strength of which shows only a mild decline with time, 
835: take a leading role in the Class II phase. They act to further decrease the disk mass when 
836: gravitational torques fail to do so. 
837: 
838: The actual time in the Class II phase when ${\cal T}_{\rm v}$ become greater than
839: ${\cal T}_{\rm g}$ is distinct for models with different initial cloud core masses (but similar 
840: ratios of rotational to gravitational acceleration $\eta$). 
841: Low-mass cloud cores form disks late in the evolution (see e.g. model~8 in Fig.~\ref{fig3}).
842: The resulting disks have low masses, merely due to the fact that most of the cloud core material 
843: has already been accreted directly onto the protostar in the pre-disk phase. As a result, 
844: gravitational instability is underdeveloped and viscous torques quickly take over the 
845: gravitational ones in the beginning of the Class II phase. 
846: Conversely, high-mass cloud cores form disks early in the evolution 
847: (see e.g. model~16 and 20 in Fig.~\ref{fig3}). The resulting disks are characterized 
848: by considerable masses. As a consequence, gravitational torques prevail over viscous torques 
849: throughout a considerable portion of the Class II phase. The interested reader 
850: is referred to \citet{VB5,VB4} for a detailed discussion on gravitational torques and
851: appearances of self-gravitating and viscous disks.
852: 
853: 
854: \subsection{Relation between disk and stellar masses}
855: In this section we derive statistical relations between time-averaged disk masses 
856: ($\langle M_{\rm d}\rangle$), time-averaged stellar masses ($\langle M_\ast \rangle$) , 
857: and time-averaged disk-to-star mass ratios ($\langle \xi \rangle$) for viscous disks 
858: in each stellar evolution phase. Time averaging is performed separately in each evolution phase 
859: over the duration of the phase.
860: Figure~\ref{fig5} shows $\langle M_{\rm d}\rangle$ (right column) and $\langle \xi \rangle$  
861: (left column) versus $\langle M_\ast \rangle$ for Class 0 objects (open triangles),
862: Class I objects (filled squares) and Class II objects (plus signs).
863: The top-to-bottom rows show the data for models with $\eta_1=1.2\times 10^{-3}$, 
864: $\eta_2=2.3\times 10^{-3}$, and $\eta_3=3.4\times 10^{-3}$, respectively. 
865: 
866: 
867: \begin{figure}
868:   \resizebox{\hsize}{!}{\includegraphics{f5.eps}}
869:       \caption{The same as Figure~\ref{fig2} only obtained in the framework of the viscous numerical
870:       approach described in \S~\ref{viscmodel}.}
871:          \label{fig5}
872: \end{figure} 
873: 
874: 
875: 
876: The comparison of Figures~\ref{fig2} and \ref{fig5} reveals a distinct feature of viscous disks
877: around objects of equal stellar mass -- Class II disk have systematically lower masses than their 
878: Class 0/I counterparts. 
879: This is in sharp contrast to self-gravitating disks studied in \S~\ref{SG} where
880: we have found that Class O/I/II objects with equal stellar mass 
881: possess self-gravitating disks of essentially similar mass (see Fig.~\ref{fig2}). 
882: The difference in masses between Class II and Class 0/I viscous disks 
883: is considerable for low-mass stars (up to a factor of 10), but it diminishes for stars 
884: of greater mass. However, we have to keep in mind that low-mass stars are statistically more
885: important than their high-mass counterparts.
886: 
887: The dependence of $\langle \xi \rangle$ on $\langle M_\ast \rangle$
888: in Figure~\ref{fig5} is similar to that of Figure~\ref{fig2} -- stars of greater mass tend 
889: to have greater disk-to-star mass ratios. However, as for the case of disk masses, 
890: Class II viscous disks have disk-to-star mass ratios that are lower than those 
891: of self-gravitating disks. For instance, the smallest $\langle \xi \rangle$ in self-gravitation disks
892: (Fig.~\ref{fig2}) is 0.05, while viscous disks have $\langle \xi \rangle$ as small as 0.01.
893: 
894: {It is interesting to compare our results with the recent work of \citet{Kratter}, who use 
895: semi-analytical models to describe the time evolution of embedded, accreting protostellar disks.
896: As in our model, they account for disk self-gravity and MRI-induced viscosity but
897: also include the effect of stellar and envelope irradiation. They obtain maximum disk-to-star
898: mass ratios in the Class~I phases of order $30\% - 40\%$, in excellent agreement with our results.
899: The disk mass in their fiducial $1~M_\odot$ model shows a temporal behaviour similar to that 
900: shown in the lower-left panel of Fig.~\ref{fig3} -- it grows in the Class~0 phase,
901: reaches a maximum in Class~I phase, and declines afterwards.  
902: They also predict that higher mass stars have relatively more massive disks and this trend
903: is expected to extend to stellar masses much greater than those considered in our work ($\la 2.0~M_\odot$).
904: 
905: 
906: We calculate the mean masses of viscous disks using equation~(\ref{mean})
907: and summarize the main properties of our viscous disks in Table~\ref{table4}.
908: It is evident that the range of time-averaged masses and the values of mean masses are quite similar
909: in both the self-gravitating and viscous disks. The latter tend to have slightly larger
910: mean disk masses in the Class 0/I phases, which is explained by the fact that viscous
911: torques tend to oppose gravitational ones in the early disk evolution \citep{VB4}.
912: The only significant difference comes from
913: Class~II viscous disks, which have a factor of two lower mean mass, 
914: $\overline{M}_{\rm CII}=0.06~M_\odot$, than that of self-gravitating disks.
915: Class II viscous disks also feature the lowest time-averaged disk mass found in our numerical
916: simulations, $\langle M^{\rm min}_{\rm d,CII} \rangle=0.004~M_\odot$. 
917: 
918: 
919: 
920: 
921: Finally, we seek to determine the relation between time-averaged disk and stellar masses 
922: in each stellar evolution phase separately and for the three phases taken altogether.
923: %for all model cloud cores of our sample, irrespective of their initial values of $\eta$.
924: Figure~\ref{fig6} shows $\langle M_{\rm d} \rangle$ versus $\langle M_\ast \rangle$
925: for our 32 model cloud cores. 
926: In particular, the open triangles, filled squares, and plus signs 
927: show the data for the Class 0, Class I, and Class II phases, respectively. 
928: It is evident that the upper-mass stars have a considerably narrower scatter in disk masses than 
929: the lower-mass stars. The least-squares best fit to {\it all} data in Figure~\ref{fig6} 
930: (solid line) yields the following relation between time-averaged stellar and disk masses 
931: (in $M_\odot$) for the three phases taken altogether
932: \begin{equation}
933: \langle M_{\rm d} \rangle = \left( 0.2\pm 0.05 \right) \, \langle M_\ast \rangle^{1.3\pm 0.15}.
934: \label{rel1}
935: \end{equation}
936: 
937: 
938: The relation between time-averaged disk and stellar masses in each stellar evolution phase
939: is, however, quite different from that expressed by equation~(\ref{rel1}).
940: The least-squares best fits performed separately for each evolution phase yield the 
941: following relations
942: \begin{eqnarray}
943: \label{rel3}
944: \langle M_{\rm d,CO} \rangle &=& \left( 0.2\pm 0.05 \right) \, \langle M_{\rm \ast,C0} \rangle^{0.7\pm 0.2}, \\
945: \langle M_{\rm d,CI} \rangle &=& \left( 0.3\pm 0.04 \right) \, \langle M_{\rm \ast,CI} \rangle^{1.3\pm0.15}, \\
946: \langle M_{\rm d,CII} \rangle &=& \left( 0.2\pm 0.08 \right) \, \langle M_{\rm \ast,CII} \rangle^{2.2\pm0.2}, 
947: \label{rel5}
948: \end{eqnarray}
949: where indices CO, CI, and CII refer to the Class~0, Class~I, and Class~II phases, respectively.
950: The dependence of disk masses on stellar masses becomes progressively steeper
951: along the sequence of stellar evolution phases. 
952: 
953: The above relations between time-averaged stellar and disk masses were derived for 
954: viscous disks, in which both viscosity and self-gravity are 
955: at work. Purely self-gravitating disks (zero viscosity) have a similar relation 
956: between time-averaged disk and stellar masses when {\it all} stellar evolution phases 
957: are considered altogether, namely equation~(\ref{rel1}).
958: However, when each phase is taken separately, the exponents are different from those derived
959: for viscous disks in equations~(\ref{rel3})-(\ref{rel5}). To avoid confusion, we provide here 
960: the relations for viscous disks only, since we believe that some sort of viscosity should 
961: be present in circumstellar disks.
962: 
963: 
964: 
965: \begin{figure}
966:   \resizebox{\hsize}{!}{\includegraphics{f6.eps}}
967:       \caption{Time-averaged disk masses versus time-averaged stellar masses for 32~models of our sample
968:       (see Tables~\ref{table1}-\ref{table3}). Open triangles, filled squares, and plus signs
969:       show the data for Class~0, Class~I, and Class~II systems, respectively. The solid line gives
970:       a least-squares best fit to all data in the plot. }
971:          \label{fig6}
972: \end{figure}
973: 
974: 
975: Observations provide conflicting evidence as to the relation between disk and stellar masses
976: in YSOs. For instance, \citet{Natta01} found a marginal correlation between disk 
977: and stellar masses, albeit with a substantial dispersion. On the other hand, \citet{Andrews}
978: claimed no correlation. The correlation in Figure~\ref{fig6} can be broken if 
979: a large population of objects occupies both the upper-left and lower-right portions of the 
980: $\langle M_{\rm d} \rangle - \langle M_\ast \rangle$ diagram. While the latter is feasible due to
981: the incompleteness of our sample of model cloud cores (we have not considered sufficiently low 
982: values of the ratio $\eta$), the former is unlikely due to the saturation of disk masses 
983: discussed in the context of Figure~\ref{fig2} and earlier considered by \citet{Adams89} and
984: \citet{Shu90} . No stars can harbour disks with masses comparable to or greater than 
985: the stellar mass! 
986: It is likely that the uncertainty in disk measurements may introduce a considerable scatter
987: to the observational data which precludes the correlation analysis.
988: 
989: 
990: 
991: 
992: \section{Discussion}
993: \label{discuss}
994: 
995: 
996: There exist a growing concern that disk masses around YSOs
997: may be systematically underestimated by conventional observational methods.
998: The simplest argument in favour of this conjecture is a high frequency of detection
999: of Jupiter-like exosolar planets. Both the core accretion and gravitational instability 
1000: models require disk masses of order $0.1~M_\odot$ for Jupiter-like planets to form, 
1001: which is about an order of magnitude 
1002: higher than median disk masses inferred by \citet{Andrews} (hereafter, AW).
1003: Another argument in favour of fairly massive disks is the fact that resolved disks 
1004: show rich non-axisymmetric structures, such as {\it multiple} spiral arms or arcs in 
1005: the case of AB Aurigae \citep{Fukagawa, Grady99} and HD 100546 \citep{Grady01}. Such structures 
1006: are difficult to sustain in low-mass disks. To remind the reader, a companion star or
1007: giant planet would  most likely trigger a two-armed spiral response in the disk of a target and 
1008: not a flocculent multi-arm structure. 
1009: 
1010: As was discussed in the Introduction, measurements of disk masses suffer from the uncertainties
1011: in dust opacities and disk radial structure, which could lead to a substantial underestimate
1012: of disk masses \citep[e.g.][]{Hartmann}. Alternatively, the disk may mask itself, especially late in
1013: the evolution, by locking a significant
1014: amount of its mass in the form of large dust grains or planetesimals which are inefficient emitters
1015: at submillimeter wavelengths \citep{Andrews}. 
1016: 
1017: Our numerical modeling supports the supposition of \citet{Hartmann} that circumstellar disks are 
1018: more massive than currently inferred from observations. The mean masses of disks 
1019: in our models range between $0.10-0.11~M_\odot$ for Class~I objects and $0.06-0.12~M_\odot$
1020: for Class II objects, well in excess of the AW's median masses for either 
1021: Class I disks ($3\times 10^{-2}~M_\odot$) or, especially, Class II disks ($3\times 10^{-3}~M_\odot$).
1022: Much the same, our Class 0 disks have mean masses $0.09-0.10~M_\odot$, while 
1023: five Class 0 disks in a sample of \citet{Brown} have a mean mass of order $0.01~M_\odot$,
1024: though the statistics in the case of Class~0 objects is inadequate to draw any firm conclusions.
1025: 
1026: A closer look at the data provided by AW reveals that our numerical modeling fails to
1027: reproduce only the lower bound on the measured disk masses but succeeds at reproducing the
1028: upper bound. For instance, figure~15 of AW
1029: indicates that Class I disk masses lie in the range $M_{\rm d,CI}=0.001-0.6~M_\odot$,
1030: while our Class I disks have masses in the range 
1031: $\langle M_{\rm d,CI} \rangle =0.02-0.55~M_\odot$. 
1032: Similarly, AW's Class II disk masses lie in the range $M_{\rm d,CII}=5\times 
1033: 10^{-4} - 0.5~M_\odot$, while our disks have masses in the range 
1034: $\langle M_{\rm d,CII} \rangle = 4\times 10^{-3}-0.7~M_\odot$.
1035: If disk masses are indeed systematically underestimated by conventional observational methods, then
1036: low-mass disks should suffer from this problem to a greater extent than do upper-mass disks.
1037: The reason for this is not well understood.
1038: 
1039: It is worth noting that our Class 0 disk have masses that are similar to those 
1040: of Class I disks. This is in agreement with the modeling of circumstellar
1041: envelope dust emission of 6 deeply embedded systems by \citet{Looney}, who have found 
1042: that Class 0 systems do not 
1043: have disks that are significantly more massive than those in Class I systems.
1044: This led them to conclude that the disk in Class 0 systems must quickly and
1045: efficiently process $\sim 1.0~M_\odot$ of material from the envelope onto the protostar.
1046: Our numerical modeling corroborates their conclusion -- Class 0/I disks develop 
1047: vigorous gravitational instability that helps keep the disk mass well below that of the protostar.
1048: 
1049: 
1050: To what extent can our numerical modeling overestimate disk masses?
1051: The reason for the overestimate may be twofold.
1052: First, Class II disk masses in Figures~\ref{fig2} and \ref{fig5} were
1053: derived from disks which have ages of order 3~Myr but real disks may be older.
1054: To test this possibility, we ran a few models for 5~Myr and compared the resulting 
1055: Class II disk masses with those obtained for 3~Myr-old disks. We found that longer ages of disks 
1056: could only cause a factor of 2 (at most) decrease in the time-averaged Class II disk masses,
1057: still not enough to reconcile our obtained disk masses with those of AW.
1058: Second, we might have not taken into account some important physical
1059: mechanisms that help reduce disk masses. 
1060: Indeed, dust disks and accretion disappear on time scales of about 6~Myr and
1061: this points to the existence of additional disk clearing mechanisms, especially 
1062: relative to the non-viscous disk models.
1063: Magnetic braking is known to be efficient 
1064: at transporting disk angular momentum directly to the external environment. 
1065: However, disks are known to have low ionization fractions and 
1066: ambipolar diffusion may strongly moderate the effect of magnetic braking.
1067: We plan to explore this mechanism in a follow-up paper.
1068: The formation of a binary system can also decrease the resulting disk masses
1069: but the fraction of binary systems is under debate and it may be too low to reconcile our
1070: model disk masses with observations \citep{Lada}. 
1071: Photoevaporation of disks due to the external ultraviolet radiation may somewhat decrease 
1072: disk masses but it operates only late in the evolution of Class II disks and is expected to
1073: have little influence on the time-averaged disk masses.
1074: Clearly,  the statistics on Class~II disks must depend on the 
1075: existence of additional disk clearing mechanism and the adopted end time in our numerical and
1076: more work is needed to resolve the cause of disparity between observationally and numerically
1077: derived disk masses.
1078: 
1079: \section{Numerical and model caveats}
1080: \label{caveats}
1081: \subsection{Gas thermodynamics}
1082: It has recently become evident that
1083: the gas thermodynamics plays an important role in regulating gravitational instability 
1084: and fragmentation in circumstellar disks \citep[e.g.][]{Pickett03,Johnson03,Rice03,Boley06,Cai08}. 
1085: While most researches agree that circumstellar disks are susceptible to gravitational instability, 
1086: particularly in the early phase of stellar evolution, 
1087: there still no consensus on whether fragmentation
1088: ever occurs and, if it does, whether any clumps that form will become bound protoplanets.
1089: For instance, \citet{Johnson03} have considered 
1090: geometrically thin disks that cool by radiatively transporting thermal energy (generated in the midplane)
1091: to the disk surface. They have found that disks are susceptible to fragmentation only if 
1092: the effective cooling time is comparable to or smaller than the dynamical time. Similar 
1093: conclusions have been made by \citet{Rice03} and \citet{Mejia04} using global SPH and grid-based
1094: numerical simulations, respectively. Fragmentation can be stabilized in the inner disks by slow cooling
1095: \citep[e.g.][]{Rafikov, Stamatellos} and in the outer disks by stellar and envelope 
1096: irradiation \citep{Matzner05,Cai08}.
1097: On the outer hand, disk fragmentation can be aided by the infall of matter from an external envelope,
1098: certainly in the early phase of disk evolution \citep{VB1,VB2,Krumholz,Kratter}. 
1099: 
1100: The accurate implementation of gas thermodynamics in numerical simulations
1101: of self-consistent disk formation and long-term evolution is a very difficult task.
1102: This is because our numerical simulations capture several physically distinct regimes 
1103: that see a substantial change with time in the chemical composition, opacities, and dust properties.
1104: Numerical techniques allowing for the accurate treatment of thermal physics 
1105: in such numerical simulations are only starting to emerge \citep[e.g.][]{Stamatellos07b}.
1106: Taking into account the complexity of gas thermodynamics in circumstellar disks, 
1107: we use the barotropic equation of state.
1108: 
1109:  It should be stressed here that circumstellar disks described by the 
1110: barotropic equation of state with $\gamma=1.4$ are susceptible to fragmentation and 
1111: formation of stable clumps in the early embedded phase of evolution. These clumps are 
1112: later driven onto the protostar and produce bursts of mass accretion that can help explain 
1113: FU Ori eruptions. This phase of protostellar accretion is known as the burst phase, 
1114: it operates in disk-star systems with mass $\ga 0.6~M_\odot$ 
1115: and it is very efficient at transporting the disk material onto the protostar \citep{VB1,VB2}. 
1116: Circumstellar disks described by $\gamma=5/3$ are hotter and less susceptible to
1117: fragmentation but the clump production is not completely suppressed \citep{VB2}.
1118: %A more realistic treatment of the thermal physics
1119: %using radiation transfer tends to produce hotter disks and create less clumps 
1120: %in the early embedded phase of disk evolution, though not completely terminating 
1121: %their production. 
1122: Recent numerical and semi-analytic modeling using a more accurate prescription for the energy
1123: balance in the disk (including radiation transfer) also predict
1124: clump formation in disks (particularly, in their outer parts) 
1125: around stars with mass equal to or more massive than one solar mass
1126: \citep{Krumholz,Mayer07,Stamatellos07,Boss08,Kratter}.
1127: 
1128: 
1129: In order to examine if a higher disk temperature can affect the accretion history and, consequently,
1130: disk masses, we stiffened the barotropic equation of state by rasing $\gamma$ from 1.4 to $5/3$.
1131: In the case of self-gravitating disks, this affects most of the disk material
1132: because most of the disk is optically thick and is characterized by surface densities 
1133: considerably larger than $\Sigma_{\rm crit}=36.2$~g~cm$^{-2}$ \citep{VB3}. On the other hand, 
1134: Class II viscous disks may have surface densities comparable to or lower than
1135: $\Sigma_{\rm crit}$, particularly in the late evolution phase \citep{VB4}. Even in this case, the increase in $\gamma$ is expected to affect
1136: a large portion of the disk material because equation~(\ref{barotropic}) 
1137: makes a gradual transition between the optically thin $\Sigma << \Sigma_{\rm crit}$ and
1138: optically thick $\Sigma>>\Sigma_{\rm crit}$ regimes. 
1139: The detailed numerical simulations are presented in \citet{VB4}, here we provide only the main 
1140: results. We find that while an increase in $\gamma$ raises the disk temperature 
1141: from 50~K (at 10~AU) to about 100~K and  makes the disk less susceptible to fragmentation 
1142: and clump formation, the amount of accreted material changes insignificantly. In fact,
1143: the mass accretion rates time-averaged over $\sim 10^4$~yr are very similar in models with $\gamma=1.4$
1144: and $\gamma=5/3$, while the instantaneous rates may be quite different.
1145: This is because a modest increase in disk temperature acts to 
1146: suppress higher order spiral modes $m > 2$, which are rather inefficient at transporting mass 
1147: radially inward and tend to produce 
1148: more fluctuations and cancellation in the net gravitational torque. At the same time,
1149: the growth of lower order modes $m \le 2$ is promoted, 
1150: which are efficient agents for mass and angular momentum redistribution. 
1151: A similar effect was reported by \citet{Cai08} for circumstellar disks heated by a strong
1152: envelope irradiation. Thus, the relative strength of lower order spiral modes increases 
1153: in hotter disks, which compensates an apparent decrease
1154: in the amount of accreted gas due to the less efficient burst phase of accretion. 
1155: 
1156: \subsection{Numerical resolution and central point mass}
1157: % The radial points are logarithmically spaced.
1158: %The innermost grid point is located at $r=5$~AU, and the size of the 
1159: %first adjacent cell lies in the range between 0.26~AU (model~21) 
1160: %and 0.36~AU (model~7). 
1161: The use of a logarithmically spaced numerical grid in the radial direction means 
1162: that the ratio $\Delta r/r$ of the cell size $\Delta r$ 
1163: to radius $r$ is 
1164: constant for a given cloud core and varies from $0.05$ (model~21) to 
1165: $0.07$ (model~7). It can be noticed from Fig.~\ref{figA1} that the aspect ratio $A=z/r$
1166: of the disk vertical scale height to radius is smaller than $\Delta r/r$ in the inner 70~AU, 
1167: which implies that a better radial (and angular) resolution is desirable in this inner region. 
1168: We have found that an increase in the resolution to $256 \times 256$ grid points makes 
1169: little influence on the accretion history and disk masses but helps save a considerable amount 
1170: of CPU time and, eventually, consider many more model cloud cores. We also point out that, due to
1171: the adopted log spacing in the radial direction, the resolution in the inner computational
1172: region is sufficient to fulfill the Truelove criterion in the disk.
1173: 
1174: In our numerical simulations the position of the central star is fixed in 
1175: the coordinate center. In reality, however, the star may wobble around the center 
1176: of mass in response to the gravity force of the disk. This
1177: can promote the growth of non-axisymmetric spiral modes (especially the m=1 mode) 
1178: in massive disks and, consequently, help limit disk masses in the early phase of stellar evolution
1179: \citep{Adams89,Shu90}. We plan to explore this potentially important mechanism in 
1180: a follow-up paper.
1181: 
1182: \subsection{The weight function}
1183: The mean disk masses listed in Table~\ref{table4} may depend on the 
1184: form of the weight function. Our adopted weight function 
1185: ${\cal F}_{\rm K}$ [see eq.~(\ref{function})] assumes that stellar masses in each
1186: stellar evolution phase are distributed according to the Kroupa IMF \citep{Kroupa01},
1187: which actually refers to the initial masses of stars that have already formed.
1188: To determine the extent to which our mean disk masses depend on the particular form
1189: of the weight function, we also consider 
1190: weight functions with a slope typical for the Salpeter IMF, 
1191: ${\cal F}_{\rm S} \left( \langle M_\ast \rangle \right)=A\, 
1192: \langle M_\ast\rangle^{-2.35}$ \citep{Salpeter}, and Miller-Scalo IMF \citep{Miller}
1193: \begin{equation}
1194: {\cal F}_{\rm MS}\left(\langle M_\ast \rangle\right) = \left\{ \begin{array}{lll} 
1195:    A \, \langle M_\ast \rangle^{-1.25} & \,\,\, 
1196:    \mbox{if $0.1~M_\odot < \langle M_\ast \rangle \le 1.0~M_\odot$ } \\ 
1197:    B \, \langle M_\ast \rangle^{-2.0} & \,\,\,
1198:    \mbox{if $1.0~M_\odot < \langle M_\ast \rangle \le 2.0~M_\odot$ } \\
1199:    C \, \langle M_\ast \rangle^{-2.3} & \,\,\,
1200:    \mbox{if $\langle M_\ast \rangle > 2.0~M_\odot$. } 
1201:     \end{array} 
1202:    \right. 
1203:    \label{MillerScalo}  
1204:  \end{equation}
1205: The resulted mean disk masses are listed in Table~\ref{table5}. It is evident that
1206: the use of different weight functions (indicated in parentheses) yields mean values that are 
1207: different from those derived from the Kroupa law by $30\%$ at most. 
1208: 
1209: 
1210: \begin{table}
1211: \begin{center}
1212: \caption{Mean disk masses with different weight functions}
1213: \label{table5}
1214: %\vspace{3 pt}
1215: \begin{tabular}{cccc}
1216: \hline
1217: \hline
1218: disk type & $ \!\!\! \overline{M}_{\rm d,C0}$ & $\!\!\! \overline{M}_{\rm d,CI}$ & $\!\!\! \overline{M}_{\rm d,CII}$  \\ 
1219: \hline
1220: self-gravitating (${\cal F}_{\rm S}$) & 0.08 & 0.08 & 0.10  \\
1221: viscous (${\cal F}_{\rm S}$) & 0.09 & 0.09 & 0.05  \\
1222: self-gravitating (${\cal F}_{\rm MS}$) & 0.09 & 0.12 & 0.15 \\
1223: viscous (${\cal F}_{\rm MS}$) & 0.10 & 0.13 & 0.08 \\
1224:  \hline
1225: \end{tabular} 
1226: \tablecomments{All disk masses are in $M_\odot$. The weight functions used to derive
1227: the mean disk masses are shown in parentheses.}
1228: \end{center}
1229: \end{table} 
1230: 
1231: 
1232: \section{Summary}
1233: \label{summary}
1234: We have considered numerically the long-term evolution of self-consistently formed 
1235: self-gravitating and viscous circumstellar disks around low-mass stars ($0.2~M_\odot \la 
1236: M_\ast\la 2.0~M_\odot$).
1237: We seek to determine time-averaged disk masses in the Class~0 ($\langle M_{\rm
1238: d, C0}\rangle$), Class~I ($\langle M_{\rm d,CI}\rangle$), 
1239: and Class~II ($\langle M_{\rm d,CII} \rangle$) stellar evolution phases. 
1240: Time averaging in each evolution phase is done over the duration of the phase.
1241: The mean disk masses ($\overline{M}_{\rm d}$) are then derived from these time-averaged masses 
1242: by weighting them over the corresponding stellar masses using a power-law function
1243: with a slope typical for the Kroupa initial mass function of star (see eq.~[\ref{mean}]). 
1244: In self-gravitating disks the radial transport of mass and angular momentum is performed 
1245: exclusively by gravitational torques, while in viscous disks both the gravitational and viscous torques
1246: are at work. Our numerical simulations yield the following results.
1247: 
1248: \begin{enumerate}
1249: \item Both the self-gravitating and viscous disks have Class~I
1250: mean masses $\overline{M}_{\rm d,CI}=0.10-0.11~M_\odot$ that are slightly more
1251: massive than those of Class~0 disks $\overline{M}_{\rm d,C0}=0.09-0.10~M_\odot$.
1252: However, viscous Class II disks have a mean mass ($\overline{M}_{\rm d,CII}=0.06~M_\odot$) that
1253: is a factor of 2 lower than that of self-gravitating Class II disks.
1254: This is explained by the fact that gravitational torques prevail through the Class 0 and Class I 
1255: phases but succumb to viscous torques through much of the Class II phase.
1256: 
1257: \item Our obtained mean disk masses are {\it larger} than
1258: those derived by AW for 153 YSOs in the Taurus-Auriga star formation region, 
1259: regardless of the physical mechanisms of mass transport in the disk.
1260: The difference is especially large for Class II disks, for which we find
1261: $\overline{M}_{\rm d,CII}=0.06-0.12~M_\odot$ but AW report median masses 
1262: of order $3\times 10^{-3}~M_\odot$.
1263: Our Class~I disks have almost a factor of 4 larger mean disk mass than those of AW.
1264: 
1265: 
1266: \item Time-averaged disk masses have a considerable scatter around mean values in each evolution phase.
1267: For instance, Class~0 and Class~I disk masses lie in the range $\langle M_{\rm d,C0} \rangle = 0.04-0.2~M_\odot$
1268: and $\langle M_{\rm d,CI} \rangle=0.02-0.55~M_\odot$, respectively, 
1269: while Class II disks have a much wider range in
1270: masses $M_{\rm d,CII}=0.004-0.7~M_\odot$.
1271: 
1272: 
1273: \item When the three stellar evolution phases are considered {\it altogether}, 
1274: we find a near-linear relation between time-averaged disk and stellar masses, 
1275: $\langle M_{\rm d} \rangle = \left( 0.2 \pm 0.05 \right) \, \langle M_\ast \rangle^{1.3\pm 0.15}$.
1276: This relation can potentially become somewhat shallower if cloud cores with very low
1277: ratios of rotational-to-gravitational acceleration are considered, but cannot be broken completely.
1278: However, when each phase is considered {\it separately}, the relation between disk and stellar 
1279: masses becomes progressively steeper along the sequence of stellar evolution phases.
1280: In particular, the corresponding relations for Class 0, Class I, and Class II objects have
1281: exponents $0.7\pm 0.2$, $1.3\pm 0.15$, and $2.2 \pm 0.2$, respectively.
1282: 
1283: \item  In each stellar evolution phase, time-averaged disk-to-star mass ratios 
1284: $\langle \xi \rangle$ tend to have greater values 
1285: for stars of greater mass. However, there is a clear saturation effect -- $\langle \xi \rangle$ 
1286: never exceed $40\%$, regardless of the stellar mass. 
1287: {\it This means that no star can harbour a disk with mass comparable to or greater than
1288: that of the star}.
1289: This saturation effect is caused by the onset of vigorous
1290: gravitational instability in circumstellar disks that form in the Class~0 or early Class~I phase
1291: \citep[see also][]{Adams89,Shu90}.
1292: 
1293: \item Only low-mass objects $M_\ast \le 0.9~M_\odot$ are expected to have Class~0 disks.
1294: Some Class~0 objects (particularly those formed from slowly rotating cloud cores)
1295: may have no disks and outflows associated with them. 
1296: %We predict that protostellar outflows do not form around these objects.
1297: 
1298: \end{enumerate}
1299: 
1300: 
1301: 
1302: \acknowledgments
1303:    The author is thankful to the referee for suggestions and comments that 
1304:    helped improve the manuscript and 
1305:    Prof. Shantanu Basu for stimulating discussions. The author
1306:    gratefully acknowledges present support 
1307:    from an ACEnet Fellowship. Numerical simulations were done 
1308:    on the Atlantic Computational  Excellence Network (ACEnet). Some of the simulations
1309:    we done on the Shared Hierarchical Academic Research Computing Network (SHARCNET)
1310:    while the author was a CITA National Fellow at the University of Western Ontario.
1311: 
1312: \appendix
1313: 
1314: \section{Disk scale height}
1315: We derive the disk vertical scale height $Z$ at each computational cell 
1316: via the equation of local vertical pressure balance 
1317: \begin{equation}
1318: \rho \, \tilde{c}_s^2 = 2\int_0^Z \rho \left( g_{z,\rm gas}+g_{z,\rm st} \right) dz,
1319: \label{eq1}
1320: \end{equation}
1321: where $\rho$ is the gas volume density, $g_{z,\rm gas}$ and $g_{z,\rm st}$ are the {\rm vertical} 
1322: gravitational accelerations due to self-gravity of a gas layer and gravitational pull of 
1323: a central star, respectively. 
1324: Assuming that $\rho$ is a slowly varying function of vertical distance $z$ between $z=0$ (midplane)
1325: and $z=Z$ (i.e. $\Sigma=2\, Z \,\rho$) and using the Gauss theorem, one can show that
1326: \begin{eqnarray}
1327: \int_0^Z \rho \, g_{z,\rm gas} \,  dz &=& {\pi \over 4} G \Sigma^2 \label{eq2} \\
1328: \int_0^Z \rho \, g_{z,\rm st} \, dz &=& {G M_\ast \rho \over r} 
1329: \left\{  1-\left[ 1+ \left({\Sigma\over 2 \rho r} \right) \right]^{-1/2} \right\},
1330: \label{eq3}
1331: \end{eqnarray}
1332: where $r$ is the radial distance and $M_\ast$ is the mass of the central star.
1333: Substituting equations~(\ref{eq2}) and (\ref{eq3}) back into equation~(\ref{eq1}) we obtain
1334: \begin{equation}
1335: \rho \, \tilde{c}_s^2 = {\pi \over 2} G \Sigma^2 + {2 G M_\ast \rho \over r} 
1336: \left\{  1-\left[ 1+ \left({\Sigma\over 2 \rho r} \right) \right]^{-1/2} \right\}.
1337: \label{height1}
1338: \end{equation}
1339: This can be solved for $\rho$ given the model's known $\tilde{c}_s^2$, $\Sigma$, and $M_\ast$
1340: using Newton-Raphson iterations.
1341: The vertical scale height is finally derived as $Z=\Sigma/(2\rho)$.
1342: 
1343: 
1344: Finally, we compare our model values of $Z$ with those  predicted 
1345: from detailed vertical structure models of irradiated accretion disks around T Tauri stars 
1346: by \citet{DAlessio}.  Their figure~1b (dashed curve) yields the following relation 
1347: between the aspect ratio $A=Z/r$ of the local scale height to radial distance and  radial distance
1348: $r$
1349: \begin{equation}
1350: A = 0.1 \left( r/100\,{\rm AU} \right)^{1/4},
1351: \label{height2}
1352: \end{equation}
1353: which is plotted by the dashed line in Figure~\ref{figA1}.
1354: The solid line in Figure~\ref{figA1} shows our obtained aspect ratio $A=Z/r$ for model~13 at t=0.3~Myr.
1355: It is evident that our $A$ is somewhat underestimated in the
1356: inner disk and overestimated in the outer disk but the disagreement between the two curves 
1357: is within a factor of unity. 
1358: 
1359: 
1360: \begin{figure}
1361: \centering
1362:   \includegraphics[width=10 cm]{f7.eps}
1363:       \caption{Aspect ratio of the disk vertical scale height to radius (Z/r)
1364:       as a function of radius (r). The thick solid line shows Z/r
1365:       calculated using equation~(\ref{height1}), while the dashed line does that
1366:       for equation~(\ref{height2}).}
1367:          \label{figA1}
1368: \end{figure}
1369: 
1370: 
1371: \section{Divergence of the viscous stress tensor}
1372: The components of $\nabla \cdot {\bl \Pi}$ in polar coordinates ($r,\phi$) are
1373: \begin{eqnarray}
1374: \left( \nabla \cdot {\bl \Pi} \right)_r &=& {1\over r} {\partial \over \partial r} r \Pi_{rr} +
1375: {1 \over r}  {\partial \over \partial \phi} \Pi_{r\phi} - {\Pi_{\phi\phi} \over r}, \\
1376: \left( \nabla \cdot {\bl \Pi} \right)_\phi &=& {\partial \over \partial r} \Pi_{\phi r}
1377: + {1\over r} {\partial \over \partial \phi} \Pi_{\phi\phi} + 2 {\Pi_{r\phi} \over r},
1378: \end{eqnarray}
1379: where we have neglected the contribution from off-diagonal components $\Pi_{rz}$ and $\Pi_{\phi z}$.
1380: The components of the viscous stress tensor $\bl \Pi$ in polar coordinates ($r,\phi$) can be found in
1381: e.g. \citet{VT}.
1382: 
1383: 
1384: \begin{thebibliography}{}
1385: 
1386: \bibitem[\protect\citeauthoryear{Adams et al.}{1989}]{Adams89}
1387: Adams, F. C., Ruden, S. P., \& Shu F. H. 1989, ApJ, 347, 959
1388: 
1389: \bibitem[\protect\citeauthoryear{Andr\'e et al.}{1993}]{Andre}
1390: Andr\'e, P., Ward-Thompson, D., \& Barsony, M. 1993, ApJ, 406, 122
1391: 
1392: \bibitem[\protect\citeauthoryear{Andrews \& Williams}{2005}]{Andrews}
1393: Andrews, S. M., \& Williams, J. P. 2005, ApJ, 631, 1134  
1394: 
1395: \bibitem[\protect\citeauthoryear{Balbus \& Hawley}{1991}]{BH}
1396: Balbus, S. A., \& Hawley, J. F. 1991, ApJ, 376, 214
1397: 
1398: %\bibitem[\protect\citeauthoryear{Basu \& Mouschovias}{1994}]{Basu94}
1399: %Basu, S., \& Mouschovias, T. Ch. 1994, ApJ, 432, 720
1400: 
1401: \bibitem[\protect\citeauthoryear{Basu}{1997}]{Basu}
1402: Basu, S. 1997, ApJ, 485, 240  
1403: 
1404: \bibitem[\protect\citeauthoryear{Boley et al.}{2006}]{Boley06}
1405: Boley, A. C., Mej\'ia, A. C., Durisen, R. H., Cai, K., Pickett, M. K., D\'Alessio, P.
1406: 2006, ApJ, 651, 517
1407: 
1408: \bibitem[\protect\citeauthoryear{Boss}{2008}]{Boss08}
1409: Boss, A. 2008, ApJ, 677, 607
1410: 
1411: \bibitem[\protect\citeauthoryear{Brown et al.}{2000}]{Brown}
1412: Brown, D. W., et al. 2000, MNRAS, 319, 154
1413: 
1414: \bibitem[\protect\citeauthoryear{Cai et al.}{2008}]{Cai08}
1415: Cai, K., Durisen, R. H., Boley, A. C., Pickett, M. K., Mej\'ia, A. C. 2008,
1416: ApJ, 673, 1138
1417: 
1418: \bibitem[\protect\citeauthoryear{Caselli et al.}{2002}]{Caselli}
1419: Caselli, P., Benson, P. J., Myers, P. C., \& Tafalla, M. 2002, 572, 238
1420: 
1421: \bibitem[\protect\citeauthoryear{D'Alessio et al.}{1999}]{DAlessio}
1422: D'Alessio, P., Calvet, N., Hartmann, L., Lizano, S., \& Cant\'o, J. 1999, ApJ, 527,
1423: 893
1424: 
1425: \bibitem[\protect\citeauthoryear{Fukagawa et al.}{2004}]{Fukagawa}
1426: Fukagawa, M., et al. 2004, ApJ, 605, L53
1427: 
1428: %\bibitem[\protect\citeauthoryear{Galli et al.}{2006}]{Galli06}
1429: %Galli, D., Lizano, s., Shu, F. H., Allen, A., 2006, ApJ, 647, 374
1430: 
1431: \bibitem[\protect\citeauthoryear{Grady et al.}{1999}]{Grady99}
1432: Grady, C. A., et al. 1999, ApJL, 523, L151
1433: 
1434: \bibitem[\protect\citeauthoryear{Grady et al.}{2001}]{Grady01}
1435: Grady, C. A., et al. 2001, AJ, 122, 3396
1436: 
1437: \bibitem[\protect\citeauthoryear{Hartmann et al.}{1998}]{Hartmann98}
1438: Hartmann, L., Calvet, N., Gullbring, E., \& D\'Alessio, P. 1998, ApJ, 495,385
1439: 
1440: \bibitem[\protect\citeauthoryear{Hartmann et al.}{2006}]{Hartmann}
1441: Hartmann, L., D'Alessio, P., Calvet, N., \& Muzerolle, J. 2006, ApJ, 648, 484 
1442: 
1443: \bibitem[\protect\citeauthoryear{Johnson \& Gammie}{2003}]{Johnson03}
1444: Johnson, B. M., \& Gammie C. F. 2003, ApJ, 597, 131
1445: 
1446: \bibitem[\protect\citeauthoryear{Klessen}{2001}]{Klessen01}
1447: Klessen, R. S. 2001, ApJ, 550, 77
1448: 
1449: \bibitem[\protect\citeauthoryear{Kratter et al.}{2008}]{Kratter}
1450: Kratter, K. M., Matzner, Ch. D., Krumholz, M. R. 2008, ApJ, 681, 375
1451: 
1452: \bibitem[\protect\citeauthoryear{Kroupa}{2001}]{Kroupa01}
1453: Kroupa, P. 2001, MNRAS, 322, 231
1454: 
1455: \bibitem[\protect\citeauthoryear{Krumholz et al.}{2007}]{Krumholz}
1456: Krumholz, M. R., Klein, R. I., \& McKee, C. F. 2007, ApJ, 656, 959
1457: 
1458: \bibitem[\protect\citeauthoryear{Lada}{1987}]{Lada87}
1459: Lada, Ch. J. 1987, in IAU Symp. 115, Star Forming Regions, editors
1460: M. Peimbert \& J. Jugaku (Dordrecht: Reidel), p. 1
1461: 
1462: \bibitem[\protect\citeauthoryear{Lada}{2006}]{Lada}
1463: Lada, Ch. J. 2006, ApJL, 640, 63L
1464: 
1465: \bibitem[\protect\citeauthoryear{Laughlin et al.}{1997}]{Laughlin97}
1466: Laughlin, G., Korchagin, V., \& Adams, F. C. 1997, ApJ, 477, 410
1467: 
1468: \bibitem[\protect\citeauthoryear{Laughlin et al.}{1998}]{Laughlin98}
1469: Laughlin, G., Korchagin, V., \& Adams, F. C. 1998, ApJ, 504, 945
1470: 
1471: %\bibitem[\protect\citeauthoryear{Lin \& Pringle}{1990}]{Lin90}
1472: %Lin, D. N. C., \&  Pringle, J. E. 1990, ApJ, 358, 515
1473: 
1474: \bibitem[\protect\citeauthoryear{Lodato \& Rice}{2004}]{Lodato04}
1475: Lodato, G.,  \& Rice, W. K. M. 2004, MNRAS, 351, 630
1476: 
1477: \bibitem[\protect\citeauthoryear{Lodato \& Rice}{2005}]{Lodato05}
1478: Lodato, G., \&  Rice, W. K. M. 2005, MNRAS, 358, 1489
1479: 
1480: \bibitem[\protect\citeauthoryear{Looney et al.}{2003}]{Looney}
1481: Looney, L. W., Mundy, L. G., \& Welch, W. J. 2003, ApJ, 592, 255
1482: 
1483: \bibitem[\protect\citeauthoryear{Lynden-Bell \& Pringle}{1974}]{Lyndenbell}
1484: Lynden-Bell, D., \& Pringle, J. E. 1974, MNRAS, 168, 603
1485: 
1486: %\bibitem[\protect\citeauthoryear{Nakano \& Nakamura}{1978}]{Nakano}
1487: %Nakano, T., Nakamura, T., 1978, PASJ, 30, 671
1488: 
1489: %\bibitem[\protect\citeauthoryear{Masunaga et al.}{1998}]{Masunaga98}
1490: %Masunaga, H., Miyama, S. M., \& Inutsuka, S. 1998, ApJ, 495, 346
1491: 
1492: \bibitem[\protect\citeauthoryear{Masunaga \& Inutsuka}{2000}]{Masunaga}
1493: Masunaga, H., \& Inutsuka, S. 2000, ApJ, 531, 350
1494: 
1495: \bibitem[\protect\citeauthoryear{Matzner \& Levin}{2005}]{Matzner05}
1496: Matzner, C. D., \& Levin, Yu. 2005, ApJ, 628, 817
1497: 
1498: \bibitem[\protect\citeauthoryear{Mayer et al.}{2007}]{Mayer07}
1499: Mayer, L., Lufkin, G., Quinn, T., \& Wadsley, J. 2007, ApJL, 661, 77
1500: 
1501: \bibitem[\protect\citeauthoryear{Mej\'ia}{2004}]{Mejia04}
1502: Mej\'ia, A. C. 2004, Ph.D. Dissertation, Indiana University
1503: 
1504: \bibitem[\protect\citeauthoryear{Miller \& Scalo}{1979}]{Miller}
1505: Miller, G. E., \& Scalo, J. M. 1979, ApJS, 41, 513
1506: 
1507: \bibitem[\protect\citeauthoryear{Natta et al.}{2001}]{Natta01}
1508: Natta, A., Grinin, V. P., Mannings, V. 2001, in Protostars and Planets IV,
1509: ed. V. Mannings, A. P. Boss, S. S. Russell (Tucson: Univ. Arizona Press),
1510: 559
1511: 
1512: \bibitem[\protect\citeauthoryear{Pickett et al.}{2003}]{Pickett03}
1513: Pickett, B. K., Mejia, A. C., Durisen, R. H., Cassen, P. M., Berry, D. K., 
1514: \& Link R. P. 2003, ApJ, 590, 1060
1515: 
1516: \bibitem[\protect\citeauthoryear{Rafikov}{2005}]{Rafikov}
1517: Rafikov, R. R. 2005, ApJ, 621, 69
1518: 
1519: \bibitem[\protect\citeauthoryear{Rice et al.}{2003}]{Rice03}
1520: Rice, W. K. M., Armitage, P. J., Bate, M. R., \& Bonnell, I. A. 2003,
1521: MNRAS, 339, 1025
1522: 
1523: \bibitem[\protect\citeauthoryear{Salpeter}{1955}]{Salpeter}
1524: Salpeter, E. 1955, ApJ, 121, 161
1525: 
1526: \bibitem[\protect\citeauthoryear{Shakura \& Sunyaev}{1973}]{SS}
1527: Shakura, N. I., \& Sunyaev, R. A. 1973, A\&A, 24, 337
1528: 
1529: \bibitem[\protect\citeauthoryear{Shu et al.}{1990}]{Shu90}
1530: Shu F. H., Tremaine S., Adams F.C., \& Ruden S. P. 1990, ApJ, 358, 495
1531: 
1532: \bibitem[\protect\citeauthoryear{Stamatellos et al.}{2007a}]{Stamatellos07}
1533: Stamatellos, D., Hubber, D. A., \& Whitworth, A. P. 2007, MNRAS, 382, L30
1534: 
1535: \bibitem[\protect\citeauthoryear{Stamatellos et al.}{2007b}]{Stamatellos07b}
1536: Stamatellos, D., Whitworth, A. P., Bisbas, T., \& Goodwin, S. 2007, A\&A, 475, 37
1537: 
1538: \bibitem[\protect\citeauthoryear{Stamatellos \& Whitworth}{2008}]{Stamatellos}
1539: Stamatellos, D., \& Whitworth, A. P. 2008, A\&A, 480, 879
1540: 
1541: 
1542: 
1543: %\bibitem[\protect\citeauthoryear{Velikhov}{1959}]{Velikhov}
1544: %Velikhov, E. P., 1959, Soviet Phys.-JETP, 36, 995 (JETP Letters 35, 1398)
1545: 
1546: \bibitem[\protect\citeauthoryear{Vorobyov \& Basu}{2005a}]{VB0}
1547: Vorobyov, E. I., \& Basu, S. 2005, MNRAS, 360, 675
1548: 
1549: \bibitem[\protect\citeauthoryear{Vorobyov \& Basu}{2005b}]{VB1}
1550: Vorobyov, E. I., \& Basu, S. 2005, ApJL, 633, L137
1551: 
1552: \bibitem[\protect\citeauthoryear{Vorobyov \& Basu}{2006}]{VB2}
1553: Vorobyov, E. I., \&  Basu, S. 2006, ApJ, 650, 956
1554: 
1555: \bibitem[\protect\citeauthoryear{Vorobyov \& Theis}{2006}]{VT}
1556: Vorobyov, E. I., \& Theis, Ch. 2006, MNRAS, 373, 197
1557: 
1558: \bibitem[\protect\citeauthoryear{Vorobyov \& Basu}{2007}]{VB3}
1559: Vorobyov, E. I., \& Basu, S. 2007, MNRAS, 381, 1009
1560: 
1561: \bibitem[\protect\citeauthoryear{Vorobyov \& Basu}{2008a}]{VB5}
1562: Vorobyov, E. I., \& Basu, S. 2008a, ApJ, 676, 139
1563: 
1564: \bibitem[\protect\citeauthoryear{Vorobyov \& Basu}{2008b}]{VB4}
1565: Vorobyov, E. I., \& Basu, S. 2008b, submitted to MNRAS
1566: 
1567: 
1568: 
1569: \end{thebibliography}
1570: 
1571: 
1572: 
1573: 
1574: 
1575: \end{document}
1576: 
1577: