0803.2892/ms.tex
1: \documentclass[12pt,preprint]{aastex}
2: 
3: \topmargin 0.5in
4: \evensidemargin 0in
5: \oddsidemargin 0in
6: 
7: \shorttitle{Cold Gas in Galactic Winds}
8: \shortauthors{Fujita et al.}
9: 
10: \begin{document}
11: \title{The Origin and Kinematics 
12: of Cold Gas in Galactic Winds: Insight from Numerical Simulations}
13: 
14: \author{Akimi Fujita\altaffilmark{1,2}, Crystal
15: L. Martin\altaffilmark{1,6,7}, Mordecai-Mark Mac
16: Low\altaffilmark{3,2,4}, Kimberly C. B. New\altaffilmark{5}, and
17: Robert Weaver\altaffilmark{5}}
18: 
19: \altaffiltext{1}{Department of Physics, University of California, Santa
20:   Barbara, CA  93106; cmartin@physics.ucsb.edu} 
21: \altaffiltext{2}{Max-Planck-Institut f\"ur Astronomie, 69117 Heidelberg,
22:   Germany} 
23: \altaffiltext{3}{Department of Astrophysics, American Museum of Natural
24:   History, New York, NY, 10024; mordecai@amnh.org}
25: \altaffiltext{4}{Institut f\"ur Theoretische Astrophysik, Zentrum f\"ur
26:   Astronomie der Universit\"at Heidelberg, 69120 Heidelberg, Germany}
27: \altaffiltext{5}{Los Alamos National Laboratory, Los Alamos, NM, 87545}
28: \vspace{0.2in}
29: \altaffiltext{6}{Packard Fellow}
30: \altaffiltext{7}{Alfred P. Sloan Foundation Fellow}
31: 
32: \begin{abstract}
33: We study the origin of Na~{\sc i} absorbing gas in ultraluminous
34: infrared galaxies motivated by the recent observations by Martin of
35: extremely superthermal linewidths in this cool gas.  We model the
36: effects of repeated supernova explosions driving supershells in the
37: central regions of molecular disks with $M_d=10^{10} M_{\odot}$,
38: using cylindrically symmetric gas dynamical simulations run with
39: ZEUS-3D.  The shocked swept-up shells quickly cool and fragment by
40: Rayleigh-Taylor instability as they accelerate out of the dense,
41: stratified disks.  The numerical resolution of the cooling and
42: compression at the shock fronts determines the peak shell density,
43: and so the speed of Rayleigh-Taylor fragmentation.  We identify
44: cooled shells and shell fragments as Na~{\sc i} absorbing gas and
45: study its kinematics along various sightlines across the grid.  We
46: find that simulations with a numerical resolution of $\le 0.2$ pc
47: produce multiple Rayleigh-Taylor fragmented shells in a given line
48: of sight that appear to explain the observed kinematics.  We suggest
49: that the observed wide Na~{\sc i} absorption lines, $\langle v
50: \rangle =320\pm120\mbox{ km s}^{-1}$ are produced by these multiple
51: fragmented shells traveling at different velocities.  We also
52: suggest that some shell fragments can be accelerated above the
53: observed average terminal velocity of $750\mbox{ km s}^{-1}$ by the
54: same energy-driven wind with an instantaneous starburst of
55: $\sim10^9~M_{\odot}$.  The mass carried by these fragments is only a
56: small fraction of the total shell mass, while the bulk of mass is
57: traveling with velocities consistent with the observed average shell
58: velocity $330\pm100\mbox{ km s}^{-1}$.  Our results show that an
59: energy-driven bubble causing Rayleigh-Taylor instabilities can
60: explain the kinematics of cool gas seen in the Na~{\sc i}
61: observations without invoking additional physics relying primarily
62: on momentum conservation, such as entrainment of gas by
63: Kelvin-Helmholtz instabilities, ram pressure driving of cold clouds
64: by a hot wind, or radiation pressure acting on dust.
65: \end{abstract}
66: 
67: \keywords{hydrodynamics, supernovae: general, ISM: bubbles, galaxies:
68:   starburst, ISM: jets and outflows, ISM: kinematics and dynamics}
69: 
70: \section{Introduction}
71: 
72: Nearly all starburst galaxies, regardless of mass, appear to drive
73: large-scale gaseous outflows, or galactic winds (Heckman et al.\ 1990;
74: Martin 1999).  Measurements demonstrate that these metal-enriched
75: winds transport interstellar gas and supernova ejecta into galactic
76: halos (Martin, et al.\ 2002). These winds are thought to influence the
77: thermal and chemical evolution of the intergalactic medium and hence
78: the formation of galaxies as well as their evolution.
79: 
80: From radio to X-ray frequencies, observations of starburst galaxies
81: reveal outflowing gas over a very broad temperature range (Martin et
82: al.\ 2002). However, all observed emission is relatively near the
83: galaxy, within a projected separation of about 10~kpc, due to the
84: radial density gradient of the wind and density-squared dependence of
85: emission processes. Absorption-line measurements are more sensitive to
86: extended, low-density gas. The number of detections of blue-shifted
87: (i.e. outflowing) interstellar absorption lines in starburst galaxy
88: spectra has grown by a large factor in recent years (Heckman et al.\
89: 2000; Rupke et al.\ 2002; Schwartz \& Martin 2004; Martin 2005). The
90: shortcoming of absorption line measurements is that they do not
91: uniquely determine the distance between the galaxy and the absorbing
92: material.
93: 
94: Numerical simulations of galactic winds can provide needed insight
95: into where the absorption originates.  Using simulations to interpret
96: observations, and observations to constrain simulations, is probably
97: the only way to really understand these complex outflows dynamically.
98: Modeling the early evolution of a galactic wind as it blows out of its
99: disk requires a numerical, rather than analytic, approach due to the
100: importance of nonlinear hydrodynamic and thermal instabilities.
101: 
102: Supershells evolve with roughly spherical geometry until they grow to
103: scales of the disk gas scale-height (Tomisaka \& Ikeuchi 1986, 1988;
104: Mac Low \& McCray 1988; Tenorio-Tagle \& Bodenheimer 1988; de Young \&
105: Heckman 1994).  The acceleration of the shell into the galactic halo
106: causes it to fragment via Rayleigh-Taylor (R-T) instabilities (Mac
107: Low et al.\ 1989).  The hot, low-density bubble interior
108: radiates inefficiently.  The wind can sweep up new shells of ambient
109: gas, that in turn fragment by R-T instability, leaving a broad region
110: containing fragments of fast-moving cool gas.
111: 
112: The swept up shell is driven by the thermal pressure of the interior
113: $P = \rho c_s^2$, where $c_s$ is the interior sound speed. After
114: blowout, the hot gas expands freely through the fragmented shell,
115: producing a supersonic, energy-driven wind with velocity $v_w$.
116: Although entrained shell fragments can still be accelerated by the ram
117: pressure of the wind $P_{ram} = \rho v_w^2$, this appears to be a
118: minor contribution to their total kinetic energy.  This can be seen by
119: comparing the velocity of a bubble expanding into a uniform medium at
120: a radius of one scale height to the final shell fragment velocities,
121: as reported, for example, by Mac Low et al.\ (1989). These are the
122: same, to within a factor of two.
123: 
124: Properties of the cool gas in starburst winds have been estimated from
125: observations of interstellar Na~{\sc i} lines in starburst galaxy spectra.
126: Estimates of the total mass of cold gas in these outflows have large 
127: uncertainties at present due to line saturation at low resolution as
128: well as corrections for ionization state and dust depletion. Nonetheless,
129: it has been emphasized that the momentum of the cool flows appear to be
130: somewhat less than the amount available from either supernova ejecta or the 
131: radiation field, at least for the most luminous starbursts (Rupke et
132: al.\ 2005; Martin 2006).  The same approximations, however, also yield kinetic
133: energies for the cool outflow that are only a few percent (up to a 
134: few tens of percent) of the supernova energy.  The same flows could also
135: be driven by energy-conserving bubbles, with only a small fraction of the 
136: total energy in the bubble going to accelerating the cold gas.
137: 
138: The standard scenario used for interpreting starburst wind absorption
139: is based on the simulation shown in Figure~11 of Heckman et al.\
140: (2000), which suggests that dense clouds are advected into the wind at
141: the interface between the low-latitude disk and the wind, by
142: Kelvin-Helmholtz instability.  With a grid resolution of 4.9 pc,
143: however, the clumps of dense gas are not fully resolved in that
144: simulation, leaving artificially large clumps that completely stop
145: fragmenting below $\sim6$ zone size.  We will show below that the R-T
146: instability is suppressed in secondary shells at that resolution, as
147: well, substantially changing the distribution of cool gas.  
148: 
149: Recently, Cooper et al.\ (2008) performed three-dimensional (3D)
150: simulations of starburst blowout through a galactic disk with a
151: fractal density distribution.  They injected energy at a rate
152: proportional to local density, rather than identifying supernova sites
153: and following the explicit evolution of their remnants.  This leads to
154: higher than physical radiative losses, so their results represent a
155: lower limit to the effects of a starburst.  They found that H$\alpha$
156: emitting gas comes from the gas dynamical stripping and fragmentation
157: of existing interstellar clouds.  This gas can reasonably also be
158: identified as a potential source of absorption, although they did not
159: address the question explicitly, nor extend the simulation to times
160: long enough to directly model the absorption.
161: 
162: We instead identify the location of the absorbing gas in fragmenting
163: shells of swept-up interstellar gas, using high-resolution
164: two-dimensional (2D) simulations with resolution as small as 0.1~pc.
165: Although in our models we compute only up to the time of blowout
166: because of the small region covered by our computational grid, we use
167: the ballistic approximation (Zahnle \& Mac Low 1995; Fujita et al.\
168: 2004) to show that gravitational deceleration does not act strongly on
169: the shells and fragments during the starburst duration (10--40 Myr) if
170: their velocities exceed $50-200 \mbox{ km s}^{-1}$ at blowout.  This
171: analysis is based, however, on an assumption that the bulk of their
172: mass remains unablated by the wind blowing past them.  Understanding
173: the full history of shell fragments and clumps will require
174: substantial further work.
175: 
176: We address the origin and kinematics of the cold wind as measured in
177: the Na~{\sc i} $\lambda\lambda$5890, 96 absorption lines.  The
178: observations pose three major questions. First, why do the absorption
179: line widths tend to greatly exceed the thermal velocity dispersion of
180: warm neutral gas?  The average full width at half maximum (FWHM) of
181: the dynamic component is $320\pm120\mbox{ km s}^{-1}$ in ultraluminous
182: infrared galaxies (ULIRGs), while the line widths range from 150 to
183: $600 \mbox{ km s}^{-1}$ in luminous infrared galaxies.  Second, why do
184: the terminal velocities of the cold gas approach the escape velocities
185: from the starburst galaxies (Martin 2005)?  Third, what do the maximum
186: and mean velocities measured in the line profiles really represent
187: physically?
188: 
189: We use our models to pursue
190: five investigations. First, we investigate
191: how the absorption properties change with viewing angle. We
192: specifically test whether multiple R-T fragmented shells along a line
193: of sight can reproduce the broad line width seen in Na~{\sc i}
194: absorption lines.  Second, we vary the 
195:    numerical 
196: resolution to demonstrate how increased resolution of radiative
197: cooling behind the shocks, and so of shell fragmentation, affect the
198: results.
199: 
200: Third, we make a more general parameter study addressing variations in
201: the properties of the outflowing cold gas with starburst luminosity,
202: the size of the starburst region, and gas surface density.  Fourth, we
203: can obtain insight into the complicated dynamics of multiphase
204: outflows, particularly their dependence on the mass-loading of the
205: wind.  We investigate mass-loading rates between
206: $\sim1.7-120~M_{\odot}\mbox{ yr}^{-1}$ and vary the mechanical
207: luminosity of the starburst between $10^{41} - 10^{43}$~erg~s$^{-1}$
208: to see what velocities are reached by the swept-up shells and their
209: fragments.  The observed X-ray temperatures vary little with starburst
210: luminosity $T \sim 10^7 K$, so the terminal wind velocities should
211: vary little with luminosity (above some critical value required for
212: blowout).
213: 
214: Finally, Heckman et al.\ (2000) argued that both Na~{\sc i} absorbing
215: and H$\alpha$ emitting gas can not originate in the swept-up shells
216: because of the lack of strong correlation between the widths of
217: Na~{\sc i} absorption lines and H$\alpha$ emission lines.  For
218: example, the outflow sources with very broad (400--600 km s$^{-1}$)
219: Na~{\sc i} absorption lines have H$\alpha$ emission-line widths
220: ranging from 145 to 1500 km s$^{-1}$.  Although a full nebular
221: emission calculation is well beyond the scope of this paper, we do
222: discuss where the ionization front might reside for various ionizing
223: photon luminosities.  We study the kinematics of the low-ionization
224: Na~{\sc i} absorbing gas and photoionized H$\alpha$ emitting gas by
225: separating them crudely, using the photoionization code of Abel et
226: al.\ (1999).
227: 
228: The acceleration of shell fragments is sensitive to how well shell
229: fragmentation is resolved.  Applying adaptive mesh refinement (AMR)
230: techniques to this problem can maintain high resolution in the shocked
231: shells and clouds. This paper is the first step toward such an
232: improved simulation.  We compare the blowout problem run on a fixed
233: grid to a similar problem run with an adaptive grid, focusing on the
234: comparison to measured properties of cold gas in galactic winds.
235: 
236: In this paper, we describe our disk and star formation models in \S~2
237: and our numerical method in \S~3.  We give the results of our
238: parameter studies in \S~4 and discuss comparisons with observations in
239: \S~5, followed by conclusions in \S~6.  In an Appendix, we show the
240: results of test simulations of blowout in a dwarf galaxy by ZEUS-3D
241: (Stone \& Norman 1992a; Clarke 1994) and SAGE (SAIC's Adaptive Grid,
242: Eulerian hydrocode; Kerbyson et al.\ 2001; Gittings et al.\ 2008).
243: 
244: \section{Disk and Star Formation Models}
245: 
246: The parameters of our starburst model are based on the properties of
247: ULIRGs to facilitate comparison with Martin (2005, 2006).  We use
248: hydrodynamic simulations to model the effects of multiple supernova
249: explosions in the central 200~pc $\times 100$~pc region of the
250: molecular disk of a ULIRG.  Our model is an extension of the blowout
251: model in dwarf galaxies described by Mac Low \& Ferrara (1999) and
252: Fujita et al.\ (2003).  Our fiducial numerical resolution is 0.2 pc,
253: sufficient to resolve cooling behind the shocks, and so the
254: fragmentation of the swept-up shells by R-T instability as well as
255: possible within a reasonable computational time. 
256:   To study the effects of numerical resolution, we use models with
257:   resolution ranging from 0.1 to 0.8 pc.
258: 
259: \subsection{Disk}
260: ULIRGs are starburst galaxies with infrared luminosity $>10^{12}
261: L_{\odot}$, and are usually found in major mergers and interacting
262: galaxies (Sanders et al.\ 1988).  They are believed to go through
263: starburst phases twice, when the gas in a galaxy with a prograde
264: orbital geometry is tidally disturbed during the first encounter with
265: another galaxy and when both galaxies meet again and finally merge
266: (Mihos \& Hernquist 1996; Murphy et al.\ 2001; Li et al.\
267: 2004).  We choose to model a molecular gas-rich spiral
268: galaxy on its first encounter with another galaxy of similar mass.
269: 
270: We set up a molecular disk with $M_g=10^{10} \mbox{ M}_{\odot}$ in a
271: dark matter halo with $M_{halo}=5\times10^{12} \mbox{ M}_{\odot}$.  CO
272: observations of ULIRGs at both first and second passages show the
273: presence of molecular disks with $M_g=0.4-1.5\times10^{10} \mbox{
274:   M}_{\odot}$ (Sanders et al.\ 1988; Solomon et al.\ 1997), which is
275: in the range found for gas-rich spiral galaxies. However, the
276: emission originates in regions a few hundred parsecs in radius,
277: yielding surface densities of $\sim0.5$--$1\times10^4~\mbox{
278:   M}_{\odot}\mbox{ pc}^{-2}$, within which the molecular mass
279: dominates the dynamical mass (Sanders et al.\ 1988; Solomon et al.\
280: 1997).  At these high surface densities, molecular hydrogen will
281: dominate (Blitz \& Rosolowsky 2006), as it can form within a few
282: million years in turbulent regions with densities over 100~cm$^{-3}$
283: (Glover \& Mac Low 2007).  The density of H$_2$ traced by CO emission
284: is $\sim500\mbox{ cm}^{-3}$, comparable to the envelope of giant
285: molecular clouds, while
286: a region of much higher density in ULIRGs is traced by HCN emission,
287: $\sim10^5\mbox{ cm}^{-3}$, comparable to star-forming cloud cores
288: (Solomon et al.\ 1992).
289: 
290: We assume that the entire interstellar medium (ISM) is a scaled-up version of
291: a normal galactic disk with the ambient densities 
292: a factor of $\sim100$ higher, making even the intercloud medium a molecular region. 
293: Thus we assume that the surface density distribution of the molecular disk is 
294: exponential, with $\Sigma(R)=\Sigma_0 exp(-R/R_d)$ where $R_d$ is a scale radius
295:  (see also $\Sigma(R)$ of Arp 220 by Scoville 1997).  
296: 
297: We choose to model a disk with a central surface density of
298: $\Sigma_{0}=10^4~\mbox{ M}_{\odot}\mbox{ pc}^{-2}$, with a disk scale
299: radius $R_d=0.7$ kpc. The disk is in hydrostatic equilibrium with a
300: Navarro, Frenk, \& White (1997; hereafter NFW) halo potential, and a
301: disk potential based on the thin disk approximation (Toomre 1963),
302: since $M_{dyn}\approx M_{g}$.  The NFW potential is
303: \begin{equation}
304: \label{nfw}
305: \Phi(x) = \frac{G M_{halo}}{R_v} ~\frac{\ln (1+cx)/x}{F(c)}, 
306: \end{equation}
307: where we set the virial radius $R_v = 326$~kpc, $x=r/R_v$, $c$ is a halo
308: concentration factor, set to $c=5$, appropriate for a large halo
309: (Jimenez et al.\ 2003), and $F(c)=\ln (1+c)-c/(1+c)$. 
310: 
311: The velocity dispersion of the molecular gas is observed to be
312: 90~km~s$^{-1}$ in Arp 220, which appears to be at the end of the
313: merging process (Scoville et al.\ 1997).  Such a high velocity dispersion
314: yields a scale height of 15 pc in its disk (Scoville et al.\ 1997).
315: Molecular clouds with such high dispersion will get destroyed by
316: colliding with other clouds, so the cooling time behind the shocks
317: must be shorter than the destruction time interval.  We assume the gas
318: is supported by turbulence with a similarly high velocity dispersion
319: $c_s=55\mbox{ km s}^{-1}$. It is lower than $90\mbox{ km s}^{-1}$
320: because gravity from our disk gas and halo can not confine the gas
321: with a higher velocity dispersion.  As a comparison, we also model a
322: disk with a higher surface density of $\Sigma_0=5\times10^4~\mbox{
323: M}_{\odot} \mbox{ pc}^{-2}$ with $R_d=0.17$ kpc and $c_s=90\mbox{ km
324: s}^{-1}$, based on the central surface density observed in Arp 220.
325: This is the highest $\Sigma_0$ of all observed ULIRGs.
326: 
327: Figure~\ref{H} shows the vertical density distributions of both disks. 
328: The exponential scale heights are rather small, 7 and 2 pc, but the
329: gas within them is very  
330: dense, $\sim5000$ and $10^5\mbox{ cm}^{-3}$ respectively. The gas density is still 
331: $\sim500\mbox{ cm}^{-3}$ at $Z\ga 4R_d$.
332: At higher altitudes, where the gas is less dense, the gas is
333: physically atomic or even ionized. We do not take that state change
334: into account in our model, though.
335: When the number density drops to $n=10^{-2}$~cm$^{-3}$ we set the gas
336: density in the halo constant as it is no longer dynamically important
337: on the length and time scales treated in our model.
338: 
339: \subsection{Star Formation}
340: \label{subsec:sf}
341: We assume a single starburst that occurs at the center of the disk,
342: and that all the kinetic energy of the starburst supernovae is
343: released in a central wind of constant mechanical luminosity.  In
344: reality, the discrete energy inputs from supernovae generate
345: blastwaves that become subsonic in the hot interior of the bubble
346: first produced by stellar winds, and hence can be treated as a
347: continuous mechanical luminosity in the study of bubble dynamics (Mac
348: Low \& McCray 1988).  These assumptions mean that a single superbubble
349: forms, evolving to produce a bipolar outflow of gas.
350: 
351: Figure~\ref{lmech} shows the evolution of mechanical luminosity
352: $L_{mech}$ as a function of time for an instantaneous starburst with
353: $10^{9}~\mbox{ M}_{\odot}$ of gas turning into stars and for
354: continuous starbursts with star formation rates of $100$ and
355: $500~\mbox{ M}_{\odot}\mbox{ yr}^{-1}$, based on the Starburst 99
356: model (Leitherer et al.\ 1999).  The Starburst~99 model uses a power
357: law initial mass function with exponent $\alpha=2.35$ between low-mass
358: and high-mass cutoff masses of $M_{low}=1~\mbox{ M}_{\odot}$ and
359: $M_{up}=100~\mbox{ M}_{\odot}$ with solar metallicity.  Star formation
360: rates in ULIRGs are estimated to be $\ga 100~\mbox{ M}_{\odot}$
361: yr$^{-1}$ based on far-infrared luminosities and the assumption of
362: continuous star formation (see Table~1 of Martin 2005; note that the
363: star formation rates given there correspond to a low mass cutoff of
364: $0.1~\mbox{ M}_{\odot}$ and must be divided by a factor of 2.55 before
365: comparison to the population synthesis models).
366: 
367: The amount of mechanical power supplied per unit stellar mass depends
368: on the star formation history. About $\sim40$ million supernovae, for
369: example, will be produced by an instantaneous starburst of
370: $M_*=10^{9}$~M$_{\odot}$.  For a continuous starburst, $L_{mech}$
371: increases until the death rate of massive stars catches up to their
372: birth rate, after about 40 Myr. The power rises particularly rapidly
373: over the first few Myr, the period modeled by our
374: simulation. Fig.~\ref{lmech}{\em a} shows the evolution for a
375: continuous starbursts with $500~\mbox{ M}_{\odot}\mbox{ yr}^{-1}$ and
376: $100~\mbox{ M}_{\odot}\mbox{ yr}^{-1}$.  For our ULIRG models, we use
377: constant mechanical luminosity winds with values $L_{mech} = 10^{43}$,
378: $10^{42}$, and $10^{41}\mbox{ erg s}^{-1}$.  The highest of these
379: corresponds to the mechanical luminosity expected from stellar winds
380: during the first 2~Myr of an instantaneous starburst with $M_*=10^9
381: \mbox{ M}_{\odot}$.  The subsequent supernovae will result in a far
382: higher mechanical luminosity, but that is likely to be vented out of
383: the galactic disk through the hole opened by the initial
384: blowout. $L_{mech} = 10^{43}$~erg~s$^{-1}$ also corresponds to a model
385: with a constant SFR of 15.9 M$_{\odot}$~yr$^{-1}$.  This
386: correspondence assumes the birth and death rates of massive stars are
387: in equilibrium, which is achieved about 40 Myr after the burst begins.
388: In this scenario, the burst would been ongoing through its initial
389: stages before our simulation starts, but gas displaced by the initial
390: feedback had been replaced by inflows, and prior feedback energy was
391: largely radiated.  Our simulation is not such a good representation of
392: this scenario.
393: 
394: We note it is an oversimplification to assume a single starburst at
395: the disk center for modeling a galactic outflow. Bipolar outflows are
396: seen in some starburst galaxies such as M82 (Strickland \& Stevens
397: 2000; Strickland et al.\ 2004) for example, although some ULIRG winds
398: appear to require starburt regions extended to $\ga1$ kpc to launch
399: the cool outflow (Martin 2006).  3D, hydrodynamic simulations of dwarf
400: starbursts have shown that extended, multiple energy sources, as well
401: as a single central energy source, form a bipolar outflow (Fragile et
402: al.\ 2004).  These simulations also demonstrated that the main effect
403: of multiple sources was to reduce the fraction of metals and energy
404: ejected from the galaxy from almost unity to around 50\%.  Our
405: assumption thus represents a reasonably strong {\em lower} bound to
406: the amount of kinetic energy that will be deposited in the observed
407: cold gas.  We therefore start with this assumption and do not expect
408: the results to differ much from those expected with more extended star
409: forming regions.
410: 
411: In addition, we neglect the effects of UV radiation on molecular
412: hydrogen in the disk.  The UV radiation from massive stars may
413: photo-dissociate some of molecular hydrogen outside star-forming cores
414: to atomic hydrogen.  However, the assumed turbulent pressure with
415: $c_{s}=55\mbox{ km s}^{-1}$ is 14 times greater than the increased
416: thermal pressure by photo-dissociation.  We thus safely neglect the
417: effects of UV radiation on the disk gas structure.
418: 
419: We define the model with $\Sigma_0=10^4~\mbox{ M}_{\odot}\mbox{
420: pc}^{-2}$ and $L_{mech}=10^{43}\mbox{ erg s}^{-1}$ as our fiducial
421: model (U1/X1). We list the parameters for all the other runs in
422: Table~\ref{run}.
423: 
424: \subsection{Bubble Dynamics}
425: 
426: In the molecular disk of a ULIRG with a very small scale height, a
427: bubble with $L_{mech}\ge10^{41}\mbox{ erg s}^{-1}$ quickly blows out
428: of the disk at $t\ll 1$ Myr.  We can only simulate the evolution of
429: the bubble up to $t\sim0.3-1$ Myr before it leaves our grid, which is
430: rather small because of the cost of high resolution. We argue in
431: \S~\ref{sec:observation} that cooled, swept-up shells, which we
432: identify as Na~{\sc i} absorbing gas, can acquire maximum velocities
433: primarily determined by when they blow out, accelerate, and fragment.
434: The question remains whether these shell fragments and clumps survive
435: as the hot interior wind streams through them.
436: 
437: Each individual shell fragment after blowout remains unstable to
438: smaller scale R-T instabilities while being further accelerated by the
439: wind, requiring extremely high resolution to fully resolve.
440: Resolution is not as critical for previously published numerical
441: studies on starburst winds that explored feedback parameters.  For
442: example, reasonable assumptions about cooling losses indicate that
443: moderate luminosity starbursts do not remove a significant fraction of
444: the galaxy's gas from the halos (Mac Low \& Ferrara 1999), and the
445: mass loss rates would only decrease further with more cooling.  A
446: large fraction of the heavy elements do escape from the halos (Mac Low
447: \& Ferrara 1999; Fujita et al.\ 2004), and this result does not depend
448: on resolution below $\sim$ a few tens of parsecs, so long as at least
449: a few R-T modes are excited in the swept-up shells to let the
450: metal-enriched gas escape.  Therefore, a major question that must be
451: answered to understand observations of cold gas at high velocity is
452: whether clumps of dense gas survive in the lower density wind (cf.\
453: Klein et al.\ 1994; Marcolini et al.\ 2005), 
454:   where they will fragment because of
455: hydrodynamic instabilities, such as R-T (strictly speaking,
456: Richtmyer-Meshkov; see Richtmyer 1960; Meshkov 1969), and Kelvin-Helmholtz.
457: 
458: Weaver et al.\ (1977) argued that the density in the hot interior of
459: bubbles in uniform gas is dominated by conductive evaporation from the
460: dense shell.  However, our model does not explicitly include thermal
461: conduction, or material ablated off of high-density molecular clouds
462: associated with the central starburst.  Instead, we add additional
463: mass to our central luminosity source to account for this process,
464: multiplying the mass input rate by a mass-loading factor $\xi$.  The
465: amount of thermally evaporated mass in a bubble expanding into a
466: uniform medium is proportional to $L^{27/35}_{mech}\rho^{-2/35}$.  
467: The internal temperature is
468: \begin{equation}
469: T_b(t)=(\gamma-1)\frac{\mu}{k_{B}}
470: \left[\frac{5}{11}\int_{0}^{t}L_{mech}(t') dt' \right] / 
471: \left[\int_0^t\xi M_{SN}(t') dt' \right].
472: \end{equation}
473: where $L_{mech}(t)$ and the mass of supernova ejecta $M_{SN}(t)$ as a
474: function of time are taken from Starburst 99 model, the adiabatic
475: index $\gamma=5/3$, mean mass per particle $\mu=14/22 m_H$, and $k_B$
476: is the Boltzmann factor.  Weaver et al.\ (1977) showed that $5/11$ of
477: the total input mechanical energy goes into the hot, pressurized
478: region bound by the inner and outer shock fronts (see their eq.~[14]).
479: This was applied to superbubbles by McCray \& Kafatos (1987).  We show
480: $T_b(t)$ in FigMac Low \& McCrayure~\ref{vtwind}{\em a}.
481: 
482: We can estimate the terminal velocity of the wind driven by such a
483: bubble by equating the bubble interior energy with the kinetic energy
484: of the mass-loaded wind, so that
485: \begin{equation}
486: \label{ewind}
487: v^2_{wind}(t)=2\left[\frac{5}{11}\int_0^t L_{mech}(t')
488:   dt'\right]/\left[\int_0^t \xi  M_{SN}(t')dt'\right],
489: \end{equation}
490: which we plot in Figure~\ref{vtwind}{\em b}.  Since the mechanical
491: luminosity $L_{mech}(t)$ and the mass of supernova ejecta $M_{SN}(t)$
492: are both linearly proportional to the amount of gas converted to
493: stars, $T_{b}$ and $v_{wind}$ are the same for all strengths of
494: starburst with the same mass-loading rate $\xi$.  We take a fiducial
495: $\xi$ value of about 8 corresponding to a mass-loading rate
496: $\dot{M_{in}}=17 (L_{mech}/10^{43})~\mbox{ M}_{\odot}\mbox{ yr}^{-1}$,
497: but run test simulations with $\xi$ varying between $\sim$1--15 to
498: explore the effects of mass-loading on the shell
499: kinematics. Comparisons to observations suggest $\xi\approx10$
500: (Suchkov et al.\ 1996 -- see M82 comparison; Martin et al.\ 2002).
501: Throughout this paper, we designate as the wind the hot interior gas
502: freely streaming outwards after blowout of the shell.
503: 
504: \section{Numerical Methods}
505: \label{sec:numerics}
506: 
507: We follow the numerical methods used by Mac Low \& Ferrara (1999) and
508: Fujita et al.\ (2003) to model a starburst in a galactic disk.  We
509: briefly summarize our methods below, but refer to the papers above for
510: more details. We compute the evolution of a starburst-driven blastwave
511: as it blows out of the molecular disk of a ULIRG with
512: ZEUS-3D\footnote{Available from the Laboratory for Computational
513: Astrophysics, http://cosmos.ucsd.edu}, an Eulerian, finite-difference,
514: astrophysical gas dynamics code (Stone \& Norman 1992a; Clarke 1994),
515: that uses second-order van Leer (1977) advection, and a quadratic
516: artificial viscosity to resolve shock fronts. We use the loop-level
517: parallelized version ZEUS-3D, in its 2D form.  Runs were
518: done on Silicon Graphics Origin 2000 machines using eight processors,
519: and typically took $\sim$1--12 days.
520: 
521: We assume azimuthal symmetry around the rotational axis of the galaxy.
522: Our fiducial grids are $1000\times500$ zones with a resolution of 0.2
523: pc, comparable to the size of star-forming cloud cores. We also run
524: the same simulations with resolution of 0.1, 0.4, and 0.8 pc to
525: examine the sensitivity of post-shock shell density and thus R-T
526: instability growth to resolution. We use reflecting boundary
527: conditions along the symmetry axis and along the galaxy midplane and
528: outfall boundary conditions on the other two axes.
529: 
530: The assumption of azimuthal symmetry limits Rayleigh-Taylor
531: instabilities to growing as rings, reducing the number of clumps below
532: what would actually be expected from 3D spikes. Mac Low et al.\ (1989)
533: compared models with azimuthal to slab symmetry to demonstrate that
534: the fixing of a central axis of symmetry did not markedly change the
535: behavior. 3D models of isolated superbubbles have only been performed
536: for the magnetized case starting with the work by Tomisaka
537: (1998). Recently Stil et al.\ (2008) have performed 3D studies of
538: unmagnetized superbubbles as calibration for a study of magnetized
539: superbubbles, but they did not extend their hydrodynamical models into
540: the R-T unstable regime. Comparison of 2D to 3D models of shocked
541: clouds by Stone \& Norman (1992b) and Xu \& Stone (1994) found little
542: difference in their dynamical evolution aside from the breakup of
543: post-shock vortex rings in 3D. Young et al.\ (2001) and Cabot (2006)
544: compared high resolution 2D and 3D models of planar, incompressible
545: R-T instability.  Cabot (2006) cautions that 2D models produce larger,
546: less-well mixed structures at late times because of the inverse energy
547: cascade that occurs in 2D flows.
548: 
549: To drive a constant luminosity wind, we add mass and energy to a
550: source region with a radius of 10 pc (50 zones).  Our fiducial mass
551: input rate is $\dot{M_{in}}=17(L_{mech}/10^{43})~\mbox{
552: M}_{\odot}\mbox{ yr}^{-1}$ which corresponds to a mass-loading factor
553: $\xi\approx8$.  For simulations with different resolutions, we keep
554: the the number of zones covering the spherical edge of the source
555: regions the same by maintaining its radius as a constant number of
556: zones rather than a constant physical size.  This is important because
557: aliasing at the edge creates density perturbations that can be
558: amplified by hydrodynamic instabilities, such as the R-T instability
559: in the swept-up shell. We use ratioed grids for lower-resolution runs
560: and we decrease the size of source region to a radius of 5 pc (50
561: zones) for a higher-resolution run.  We directly show the effects of
562: this initial noise on the development of R-T instability by running a
563: simulation with a resolution of 0.2 pc, but with a source region with
564: a radius of 5 pc (25 zones).
565: 
566: As in Mac Low \& Ferrara (1999), we use a cooling curve by MacDonald \& Bailey (1981) for
567: solar metallicity with a temperature floor of either $T_{floor}=10^2$
568: K, the temperature to which the metals can cool the gas, or $10^4$ K,
569: the temperature maintained by photoionization heating.  The shocked
570: gas in swept-up shells cools efficiently to the temperature floor set
571: in the cooling curve, since our molecular disk is very dense.  We show
572: in \S~\ref{subsec:effects-cc} that even our highest resolution runs do
573: not yet fully resolve the dense shells even for $T_{floor} = 10^4$~K,
574: so the influence of the cooling floor is not evident in our work.  We
575: also include an empirical heating function tuned to balance the
576: cooling in the background atmosphere. This is linearly proportional to
577: density, so that it is overwhelmed by cooling in compressed gas which
578: is proportional to the square of the density (Mac Low et al.\ 1989).
579: This is to prevent the background atmosphere from spontaneously
580: cooling.  We use a tracer field (Yabe \& Xiao 1993) to turn off
581: radiative cooling in the hot bubble interior, in order to prevent mass
582: numerically diffused off the dense shell from spuriously cooling the
583: interior. The cooling time of the interior is much longer than the
584: dynamical time of our bubble, so interior cooling is physically
585: unimportant to the bubble dynamics (Mac Low \& McCray 1988).
586: These adiabatic bubble interiors form energy-driven winds after blowout.
587: 
588: \section{Parameter Studies}
589: 
590: We now describe the results of parameter studies of both physical and
591: numerical variables. We begin by considering physical variables,
592: including mechanical luminosity, mass-loading of the wind, and disk
593: surface density.  We then discuss numerical variables, focusing on
594: how numerical resolution and the cooling cutoff temperature affect
595: shell density and also examining the effect of changing the size of
596: the source region.
597: 
598: \subsection{Physical Parameters}
599: \subsubsection{Mechanical Luminosities}
600: Our fiducial model (X1) has mechanical luminosity
601: $L_{mech}=10^{43}\mbox{ erg s}^{-1}$.  This model corresponds to the
602: first 2~Myr of a starburst in which $10^9~\mbox{ M}_{\odot}$ of gas
603: turns into stars instantaneously. We compare this to models with lower
604: mechanical luminosities $L_{mech}=10^{42}$ and $10^{41}$ erg s$^{-1}$
605: (models X2 and X3 respectively) in the same molecular disk.  These
606: mechanical luminosities correspond to instantaneous bursts of $10^8$
607: and $10^7$~M$_{\odot}$.  Figure~\ref{L43} shows the density
608: distribution of our fiducial model in its right panel.  This may be
609: compared to Figure~\ref{L421}, which shows the density distributions
610: of the two models with lower $L_{mech}$ at $t\approx0.49$ Myr and 0.85
611: Myr respectively.  The swept-up shells fragment due to R-T instability
612: into multiple clumps and shells.
613: 
614: Secondary Kelvin-Helmholtz instabilities ablate the sides of these
615: fragments as the hot gas streams through them. Look at the clumps, for
616: instance, at (R,Z)$\approx(40,65)$, (30,110), and (15--25,120) pc in
617: X1.  Note also that the swept-up shells in the horizontal direction
618: are also R-T unstable, because our disk gas is stratified in the
619: radial direction, too, due to its exponential surface density profile.
620: 
621: In fact, the degree of fragmentation is larger in X2 and more so in X3
622: because $\dot{M_{in}}\propto L_{mech}$ and so the density of interior
623: gas is lower.  In particular, most shell fragments in X3 are already
624: falling back to the disk.  A mechanical power of
625: $10^{42}$~erg~s$^{-1}$ is just too small in such a dense environment
626: to accelerate the bulk of the shells to the disk's escape velocity.
627: 
628: \subsubsection{Surface Density}
629: The top right panel of Figure~\ref{L421} shows the density
630: distribution of our model with $M_d=10^{10}~\mbox{ M}_{\odot}$, but
631: with a higher surface density, $\Sigma_0=5\times10^4~\mbox{
632: M}_{\odot}\mbox{ pc}^{-2}$ at $t=0.22$ Myr (V1).  With the same
633: mechanical luminosity, $L_{mech}=10^{43}\mbox{ erg s}^{-1}$, the
634: bubble blows out earlier at $t\approx0.15$ Myr, because the disk is
635: more stratified with a smaller scale height (see Figure~\ref{H}).
636: Except the time of blowout, the degree of fragmentation and the shell
637: kinematics are about the same in models with surface
638: densities different by a factor of five.
639: 
640: \subsubsection{Mass-Loading}
641: 
642: Mass-loading from thermal conduction and molecular clouds determines
643: the density of the bubble interior and wind.  Figure~\ref{w} shows the
644: density distributions of our fiducial ULIRG model with
645: $T_{floor}=10^2$ K with mass-loading rates of 1.7, 17, 49, and 120
646: $\mbox{ M}_{\odot}\mbox{ yr}^{-1}$ (models U1-A, U1, U1-B, and U1-C).
647: These mass-loading rates correspond to bubble interior temperatures,
648: $T_{b}=1.7\times10^8$, $2.2\times10^7$, $7.5\times10^6$, and
649: $3.2\times10^6$ K and wind terminal velocities expected when all the
650: thermal energy is converted to kinetic energy, $v_{wind}=2700$, 1000,
651: 600, and 250 km s$^{-1}$ respectively.
652: 
653: Figure~\ref{w} shows the bubbles just before they leave the grid, at
654: $t=0.23$, 0.27, 0.35, and 0.41 Myr respectively.  Since the input
655: mechanical luminosity is the same in all these models the bubbles
656: initially grow at about the same rate, driven by the thermal pressure
657: of the hot interior gas.  However, a bubble with a lower mass-loading
658: rate and higher wind terminal velocity expands faster into the halo
659: once the swept-up shells fragment and the hot gas blows out between
660: the fragments.  In addition, the blowout occurs earlier and in more
661: places with a lower mass-loading rate because the density of the
662: bubble interior gas is lower.  A higher density contrast between the
663: hot interior and the swept-up shells promotes shell fragmentation by
664: R-T instability, the growth of which is proportional to the density
665: contrast.  In the least dense model U1-A, all the swept-up shells
666: quickly fragment into fingers and filaments and the dense clumps at
667: their edges are subject to secondary Kelvin-Helmholtz instabilities.
668: 
669: As the hot bubble interior becomes transonic after blowout, it
670: accelerates the shells and shell fragments by ram pressure rather than
671: thermal pressure.  This low-density wind thus can accelerate the
672: shells to higher velocity after blowout. We quantitatively study the
673: effects of wind ram pressure in the next section
674: (\S~\ref{subsec:AMvelocity}).  The observed X-ray temperature $T_X$ is
675: $\sim0.67$ keV $ = 7.7 \times 10^6$~K in all kinds of starburst
676: galaxies from dwarfs to ULIRGs (Martin 1999; Heckman et al.\ 2001; Huo
677: et al.\ 2004; Grimes et al.\ 2005).  This corresponds to
678: $\xi\approx10$ in equation~\ref{ewind}.  However, the X-ray emission
679: is proportional to $n^2$ so it is biased toward high-density regions
680: such as the interface between the hot interior gas and the shells and
681: their fragments.  At this interface, conductive evaporation and
682: turbulent ablation raise the density.  The recent observations of
683: diffuse hard X-ray emission in starburst galaxies suggest the
684: existence of a very hot ($\log T\ga7.5$) metal-bearing gas
685: (e.g. Strickland et al.\ 2004; Strickland \& Heckman 2007).  
686: Then the bulk of the hot wind may
687: still be very hot $\sim10^8$ K, the temperature which Strickland \&
688: Stevens (2000) modeled for M82 with $\xi=1$.
689: 
690: \subsection{Numerical Parameters}
691: 
692: \subsubsection{Grid Resolution}
693: \label{subsec:effects-res}
694: 
695: Figure~\ref{res} shows the density distributions of our fiducial model
696: with grid resolution varying from 0.1 to 0.8~pc.  We chose
697: $T_{floor}=10^4$ K for this resolution study, because shell densities
698: are not too far from what we expect analytically with this high
699: minimum temperature.  The growth of R-T instability is significantly
700: enhanced in the highest resolution run, X1-0, and suppressed in the
701: lower resolution runs, X1-2 and X1-4. In particular, all the outermost
702: shells seen in X1 are further fragmented by R-T instability in X1-0
703: with the resolution increased by a factor of only two.  The positions
704: of outer shock fronts in the horizontal direction agree very well
705: among the four simulations, since the shells there are not subject to
706: severe hydrodynamic instabilities.
707: 
708: Figure~\ref{dres} shows the density profiles at the outer shock fronts
709: before any fragmentation occurs for the runs in our standard
710: resolution study.  The bubble in our highest resolution run X1-0 grows
711: a little slower in the beginning, because we chose the size of the
712: source region in X1-0 to be half of that in X1 in order not to
713: overproduce noise at the contact discontinuity.  We will show below
714: that this noise feeds R-T instability and must be maintained the same
715: in order to study the effects of resolution on shell fragmentation
716: alone.  Thus the density profile of X1-0 at $t=0.06$ Myr is not
717: directly comparable to those of other runs at $t=0.05$ Myr, but
718: Figure~\ref{dres} still demonstrates the trend in resolving the peak
719: shell density as a function of resolution.
720: 
721: Figure~\ref{dres} shows that the shell density is progressively better
722: resolved as the resolution increases.  However, the cooling times at
723: the shock front are typically of order $10^2$~yr (see
724: \S~\ref{subsec:effects-cc}), so a shock with cooling floor equal to
725: the background temperature of $10^4$~K may be treated as
726: isothermal. At the time displayed, the Mach number of the outer shock
727: is ${\cal M}=19$. The shell density we expect from an isothermal shock
728: propagating into a background density $\rho_{bg}$ is $\rho_{bg} {\cal
729: M}^2 = 6.1\times10^{-19}\mbox{ g cm}^{-3}$.  Figure~\ref{dres} shows
730: that the shell density $\rho_{shell}$ is still unresolved by a factor
731: of $\sim4$ even with our highest grid resolution of 0.1 pc.
732: 
733: The difference we see in the development of R-T instability develops
734: because of two factors.  First, the post-shock density of the swept-up
735: shells is not fully resolved. Increasing density contrast drives
736: faster R-T growth.  Second, the linear R-T instability grows more
737: quickly at smaller wavelengths, but the nonlinear development moves to
738: increasing larger wavelengths as competition between growing modes
739: becomes important (Youngs 1984).  At least 10--25 zones is required to
740: resolve the smallest modes, though (e.g. Mac Low \& Zahnle 1994), so
741: grid resolution matters critically for the initial development and
742: transition to nonlinearity.  However, we believe that we reached the
743: point where resolution effects are no more important than physics we
744: have not included such as magnetic fields, non-equilibrium cooling,
745: thermal conduction, and photoionization, as well as the assumption of
746: azimuthal symmetry (see \S~\ref{sec:numerics}).
747: 
748: For example, strong magnetic fields $B\sim20\mu G$ observed in the
749: Antennae merging galaxy (Chy\.zy \& Beck 2004) can potentially inhibit
750: the formation of cold, dense shells or suppress their
751: fragmentation. Our study does show that the degree of R-T
752: fragmentation is important to reproduce the observed wide range of
753: cooled shell fragments. Recall, however, that the shell density in our simulations
754: can be more than an order of magnitude underresolved. Thus the degree
755: of fragmentation will not be significantly overestimated in our
756: simulations unless the magnetic fields are strong enough to reduce the
757: shell density by more than that.
758: 
759: The density $\rho_{sh}$ expected behind an isothermal shock running
760: into a magnetic field with Alfvenic Mach number $\cal{M_A}$ and having
761: $1 \ll {\cal M}_A \ll {\cal M}$ is $\rho_{sh} = \rho_0
762: \sqrt{2}{\cal M}_A$ (Draine \& McKee 1993). Before the effects of
763: magnetization become important in limiting R-T instability, the ratio
764: between the magnetized and unmagnetized post-shock densities must be
765: of order $2^{1/2} {\cal M}_A / {\cal M}^2 < 10^{-2}$.  With
766: $\rho\sim10^{-24}\mbox{ g cm}^{-1}$ and $v_s=500-800\mbox{ km
767: s}^{-1}$, the shell density expected with $B \sim 20 \mu$G becomes
768: three order of magnitude lower than that without B.  The effect of
769: magnetic field thus becomes substantial only when a bubble grows to
770: the high-Z, low-density part of the disk, above $Z>100$ pc.  However,
771: most of the fragmentation occurs within $Z\sim100$ pc especially with
772: the highest-resolution run, Thus our results on the wide absorption
773: profiles are robust.  If anything, the suppression of fragmentation by
774: magnetic pressure at high latitude will allow the hot wind to keep
775: accelerating the outermost shells.  Magnetic pressure might thus even
776: increase the amount of cool gas with high terminal velocity.
777:  
778: We also note that resolving the shell density profile is not important
779: to following the overall dynamical evolution of bubbles driven by the
780: thermal energy of the hot, pressurized regions (Castor et al.\ 1975;
781: Weaver et al.\ 1977), but is important to follow the details of shell
782: fragmentation due to R-T instability and the fate of dense fragments
783: and clumps by shocks and following hydrodynamic instabilities.  We
784: will show below that the resolution of 0.2 pc is still not sufficient
785: to properly model hydrodynamic instabilities acting on shells and
786: clumps, but is just sufficient for the purpose of demonstrating a wide
787: range of velocities in shell fragments caused by R-T instability.
788: 
789: \subsubsection{Cooling Temperature Floor}
790: \label{subsec:effects-cc}
791: We show the density distributions of our fiducial model 
792: with different cooling curve cut-offs of $T_{floor}=10^2$ K (U1) and $10^4$ K (X1) at 
793: the time of blowout in Figure~\ref{L43}.  
794: Figure~\ref{L43} shows that the temperature floor
795: we choose for our cooling function appears to have a negligible influence on the
796: evolution of bubbles in our models.  
797: 
798: However, closer examination reveals that the density of shocked shells
799: is approximately the same in both simulations despite the difference
800: in temperature floor. For example, at $t=0.05$ Myr before any
801: fragmentation occurs, the shell density in both simulations is
802: $\sim8\times10^{-20}\mbox {g cm}^{-3}$ in the vertical direction where
803: the background disk density is $\rho_{bg}=2\times10^{-21}\mbox{ g
804: cm}^{-3}$.  This is because the density peak in the simulations is
805: limited by resolution in these models, not by the strength of cooling.
806: The shock velocity in the vertical direction at $t=0.05$~Myr is
807: $\sim230\mbox{ km s}^{-1}$, or Mach number ${\cal M} = 15$ in $10^4$~K
808: background gas. The immediate post-shock temperature is then $T =
809: 7.6\times10^5$ K. This shocked gas quickly cools to or below $10^{4}$
810: K because the exponential cooling time (e.g. Mac Low \& McCray 1988)
811: is very short,
812: \begin{equation} 
813: \label{cool}
814: \tau_{cool} = 3kT/4n_{bg}\Lambda \approx64\mbox{ yr},
815: \end{equation} 
816: with the mean mass per particle $\mu = 14/22~m_{H}$
817: for ionized gas, and $\Lambda(T)=4.1\times10^{-23}
818: \mbox{ erg cm}^3\mbox{ s}^{-1}$ from the MacDonald \& Bailey (1981) cooling curve. 
819: 
820: The shell density expected from an isothermal shock will then be
821: $\rho_{shell}=\rho_{bg} {\cal M}^2\approx 5\times10^{-19}\mbox{ g
822: cm}^{-3}$ with Mach number ${\cal M}=15$ if $T_{floor}= 10^4$ K.  If
823: $T_{floor} = 10^2$~K, the shell density will reach values even greater
824: than the isothermal value.  However, since the shell density is far
825: from being resolved even with 0.1 pc resolution in our simulations,
826: the cooling floor has a negligible influence on our models.
827: 
828: \subsubsection{Effects of Source Region on Shell Fragmentation}
829: The top left panel of Figure~\ref{L421} shows the density distribution
830: of our fiducial model with $T_{floor}=10^4$ K, but with a source
831: region of a smaller radius of 5 pc or 25 zones (S1). The density
832: structure of S1 looks very similar to that of X1, but the
833: fragmentation by R-T instability is slightly suppressed. This is
834: because a smaller number of cells is covering the edge of the
835: spherical source region in which the density is imperfect.  This
836: imperfection creates a perturbation which gets amplified by
837: hydrodynamic instabilities.  A much smaller source region will make a
838: bigger difference in the amount of fragmentation.  It is important to
839: note that we are anyway probably underestimating the degree of
840: fragmentation since our bubble sweeps up a smooth, one-phase ISM
841: instead of real ISM with density fluctuating strongly from the mean
842: (see Cooper et al.\ 2008), and we have only a small, central source
843: region instead of an extended distribution of supernovae (cp.\ Fragile
844: et al.\ 2004).
845: 
846: \section{Comparison to Observations}
847: \label{sec:observation}
848: In this section, we first use the ballistic approximation to justify
849: making a comparison between the observations and our simulations in
850: which the bubbles are evolved only up $t\ll1$ Myr.  Then, we identify
851: gas parcels likely to produce Na~{\sc i} absorption, and simulate
852: observations of this gas along sightlines towards the galactic
853: nucleus. The velocity spread, the mass-weighted velocity, and the
854: maximum velocity are compared for different viewing angles.  The
855: velocity gradient across one of these winds is also studied for
856: comparison to longslit spectra. We then compare our models to the
857: observed Na~{\sc i} absorption profiles.
858: 
859: \subsection{Ballistic Approximation}
860: \label{subsec:ballistic}
861: 
862: We use a limited grid size in order to maintain high resolution, so we
863: simulate the evolution of bubbles only up to the time of blowout.  We
864: showed in the previous section that the bubbles blow out very early,
865: at $t\ll1$ Myr, because of the small scale height of our molecular
866: disk.  For comparison, ultraluminous starbursts have ages of up to
867: 50~Myr (Murphy et al.\ 2001).  To extrapolate our computational
868: results to later times, we use the ballistic approximation (Zahnle \&
869: Mac Low 1995; Fujita et al.\ 2004) that after blowout, shell fragments
870: travel on radial ballistic orbits in the gravitational potential of
871: the galaxy, with no further accelerations by gas pressure gradients.
872: 
873: The strength of this approximation depends on the behavior of the
874: wind, and conditions in the region into which the wind
875: penetrates. This approximation was successfully used by Zahnle \& Mac
876: Low (1995) to follow the ejecta of a typical Shoemaker-Levy 9 fragment
877: falling back to Jupiter's atmosphere, and was found to give results
878: quantitatively consistent with the observations of the size and shape
879: of the infrared bright spots.  That case does differ from the
880: starburst case in having neither an ongoing wind, nor a complex
881: density distribution above the site of the explosion.  In the
882: starburst case, continuing supernova explosions at late times drive an
883: ongoing wind (see Figure~\ref{lmech}), while the mergers that drive
884: most starbursts yield a complex geometry above the site of the
885: blowout.  However, even if the stronger winds expected from supernovas
886: at late times in an instantaneous starburst do accelerate the
887: fragments further, the ballistic approximation will still yield a {\em
888:   lower} limit to their velocities.  Since the question of most
889: interest is whether cold gas can be accelerated to such high
890: velocities, a lower limit is already useful.  The complex geometry of
891: tidal tails may also not be of great concern, as models suggest that
892: they only cover a small fraction of the solid angle visible from the
893: nucleus, and so are unlikely to be dynamically dominant.
894: 
895: Under the ballistic
896: approximation, the equation of motion for a shell fragment at a
897: distance $r$ from the galactic nucleus is
898: \begin{equation}
899: v(r) = \left\{v^2(r_b)+2[\Phi(r_b)-\Phi(r)]\right\}^{1/2},
900: \label{ba}
901: \end{equation}
902: where $\Phi(r)$ is the total halo and disk potential, and $v(r_b)=v_b$
903: is the shell velocity at blowout at a position of $r_b$.
904: 
905: In Figure~\ref{ballistic}, we plot $v(r)$ for several initial
906: velocities starting at a fixed initial position, $r_b=200$ pc, just
907: above our model disk. By setting $r_b=200$ pc, we can ignore the disk
908: potential which is negligible above the disk compared to the halo
909: potential.  Thus we set $\Phi=\Phi_{halo}$ in equation~(\ref{ba}) and
910: solve the equation analytically using equation~(\ref{nfw}).  The upper
911: limit to the radius reached by a shell fragment at time $t$ is given
912: by the linear approximation $r \le v_bt$, because the shell will only
913: decelerate in the potential.  We show actual radii for different
914: initial velocities at $t=10$ Myr by vertical ticks in the left panel
915: of Figure~\ref{ballistic}.  Although Figure~\ref{ballistic} is plotted
916: as a function of radial velocity, we find similar results if total
917: velocity is used instead, because shell fragments are traveling in
918: nearly radial directions.  This figure shows that the linear
919: approximation is quite good for $v_b > 100 \mbox{ km s}^{-1}$, so we
920: use it to follow the cold gas for times of order 10~Myr.
921: 
922: For times $t\gg100$ Myr, the linear approximation fails and most gas
923: falls back to the center after reaching heights of a few hundred kpc.
924: Only gas with $v_b\ga1000\mbox{ km s}^{-1}$ can completely escape the
925: potential of the disk and halo.  (The halo alone has an escape
926: velocity $v_{esc}=800\mbox{ km s}^{-1}$ at radii less than $0.01
927: R_v$.)  We show below that very little gas actually escapes the halo,
928: with $\la0.1\%$ of the total shell mass accelerated above $v_b
929: \ga1000\mbox{ km s}^{-1}$ in our fiducial model.  Our results confirm
930: the result from previous simulations of smaller galaxies that the loss
931: of ISM mass is inefficient (e.g. Mac Low \& Ferrara 1999; D'Ercole \&
932: Brighenti 1999).  However, significant mass is circulated
933: over scales of 10 kpc, presenting a significant absorption cross
934: section as suggested by Martin (2006).
935: 
936: 
937: \subsection{Cool Gas}
938: \label{subsec:coolgas}
939: 
940: Only the coolest gas can be observed in Na~{\sc i} absorption.  We
941: chose a temperature cut-off of $T<5\times10^4$ K to trace this gas
942: because cooling remains numerically limited even with our highest
943: resolution grid.  Gas below this cut-off is so dense that the cooling
944: time (eq.~\ref{cool}) is very short, $\tau_{cool}< 0.01$ Myr, so it is
945: physically expected to reach temperatures where Na~{\sc i} is present.
946: As an example, the least dense shell fragment in Figure~\ref{line} has
947: $\rho=3.0\times10^{-24}\mbox{ g cm}^{-3}$ and $T=3.3\times10^4$~K.
948: Equation~(\ref{cool}) gives an exponential cooling time
949: $\tau_{cool}\approx 0.007$ Myr, using
950: $\Lambda(T)\approx4.2\times10^{-23}\mbox{ erg cm}^3\mbox{ s}^{-1}$
951: from the MacDonald \& Bailey (1981) cooling curve.
952: 
953: Figures~\ref{line} and~\ref{bline} show the gas density, temperature,
954: and radial velocity in models X1 and X1-0 along a line of sight
955: through the central continuum source at an angle $\theta=19^{\circ}$
956: from the vertical axis.  Notice that our temperature cut-off picks up
957: only the densest and the coolest parts of the shells excluding
958: numerically diffused interfaces between the shells and the hot gas. 
959: 
960: We compare the characteristics of this temperature-selected gas to the
961: measured properties of Na~{\sc i} absorption in ultraluminous
962: starbursts.  As argued above in \S~\ref{subsec:sf}, our assumption of
963: a central point source of wind luminosity likely represents a lower
964: limit to the amount of cold gas entrained.  Fragile et al.\ (2004)
965: found that, in galaxies where a central energy source ejects nearly
966: all its kinetic energy in the hot gas (Mac Low \& Ferrara 1999),
967: distributed supernovas deposit as much as half of the energy in the
968: cold gas.  However, much of this energy is radiated promptly rather
969: than converted to kinetic energy, so determining the significance of
970: distributed supernovas requires further work.
971: 
972: 
973: \subsubsection{Cool Gas Column Densities}
974: 
975: Figure~\ref{res} shows that the lines of sight plotted in
976: Figures~\ref{line} and \ref{bline} go through multiple distinct
977: shells, adding to a total column density of cool gas $N_H
978: \sim7\times10^{21}\mbox{ cm}^{-2}$.  The inferred column densities in
979: ultraluminous starbursts are a bit lower, exceeding
980: $10^{21}$~cm$^{-2}$ in only one out of four ultraluminous starbursts
981: (Martin 2005).  Since the model column densities fluctuate by as much
982: as an order of magnitude between neighboring sightlines, we plot the
983: column density distribution as a function of angle in
984: Figure~\ref{hcolumn}.  We find that most sightlines with $N_H > 5
985: \times 10^{22}$~cm$^{-2}$ lie within 30$^\circ$ of the galactic disk
986: where the shells are usually not subject to hydrodynamic
987: instabilities. The typical value is $N_H \sim 10^{22}$~cm$^{-2}$,
988: similar to the largest column density estimated from observations of
989: ultraluminous starbursts (Martin 2005, 2006; Rupke et al. 2002).
990: 
991: \subsubsection{Geometric Dilution}
992: 
993: The column densities found in our model are measured at a time very
994: early in the lifetime of the wind, when it has not expanded far away
995: from the galaxy. However, these winds will typically be observed at
996: later times, when the wind has expanded further.  Geometric dilution
997: will then reduce the column density along any particular line of sight even
998: if the gas simply expands radially outwards on ballistic orbits.  
999: For spherical geometry (of any opening angle) and constant velocity
1000: flow, the volume-averaged gas density declines with increasing radius.
1001: The amount of dilution is therefore determined by the radial
1002: advance of the innermost cold gas. At early times, see Figure~\ref{res}
1003: this innermost cold gas lies at a radius of $\sim 100$~pc, near the wind
1004: termination  shock.  Assuming the cold gas directly above the
1005: starburst region flows upward on a ballistic path, we would expect
1006: it to move outward by a factor of roughly 100 in the next 25~Myr,
1007: reducing the cold gas column density by the same factor.
1008: 
1009: \subsubsection{Resolution Effects}
1010: 
1011: A second reason we overestimate the cool gas column density is that
1012: the fragmentation of the shells is limited by numerical resolution, as
1013: well as our assumption of azimuthal symmetry.  However, as discussed
1014: above in \S~\ref{subsec:effects-res}, the overall kinematics described
1015: in \S~\ref{subsec:ballistic} should remain similar unless shell
1016: fragments are entirely mixed with hot gas and destroyed.  To
1017: distinguish real physical effects from numerical artifacts, we study
1018: the kinematics of the cold gas along a $\theta=19^{\circ}$ line of
1019: sight in our fiducial simulation X1 and the high-resolution simulation
1020: X1-0.
1021: 
1022: The first two fragments from the center are
1023: parts of a filament remaining from the initial fragmentation of the swept-up
1024: supershell due to R-T instability  
1025: at $t\sim0.1$ Myr. These high-density peaks are traveling with similar
1026: velocities $v\sim400-500\mbox{ km s}^{-1}$ in both simulations
1027: and dominate the column densities of cool gas in the studied sightlines. 
1028: The outermost shell in X1 keeps sweeping up high-altitude disk gas, 
1029: while in X1-0 this shell further fragments by R-T instability to the last two fragments
1030: in Figure~\ref{bline}. The coherence of the shell in X1 clearly occurs
1031: because of the lower resolution in that run.
1032: However, the shell in X1 and the outermost fragment of the two in X1-0 travel 
1033: with similarly high velocity of $\sim800-900\mbox{ km s}^{-1}$, though they 
1034: contribute very little to the total column density. 
1035: 
1036: The hot wind overruns and shocks the first fragment in X1 and the
1037: first two fragments in X1-0.  As a result, secondary R-T and K-H
1038: instabilities act on them, removing gas from the cold cloud and mixing
1039: it into the hot wind.  In both cases, these clumps of gas are resolved
1040: by about $\sim10$ cells, demonstrating that this is a minimum size
1041: below which fragments tend to survive because the secondary
1042: instabilities cannot be resolved.  Mac Low \& Zahnle (1994) show that
1043: fragments must be resolved by at least 25 zones to resolve the
1044: secondary instabilities.  Fragments smaller than that remain as
1045: artificially massive clouds in our simulations, overpredicting the
1046: cool gas column densities.
1047: 
1048: Klein et al.\ (1994) suggest that these clumps of gas should be
1049: destroyed, via hydrodynamic instabilities, over $\sim10$ shock
1050: crossing timescales
1051: \begin{equation}
1052: t_{cc} = r_c(v_w-v_c)\sqrt{\frac{\rho_{c}}{\rho_{w}}}
1053: =(0.04\mbox{ Myr})
1054: \left(\frac{r_c}{2\mbox{ pc}}\right) 
1055: \left(\frac{v_w-v_c}{500\mbox{ km s}^{-1}}\right)
1056: \left(\frac{\rho_c/\rho_w}{100}\right)^{1/2},
1057: \end{equation}
1058: where $r_c$ is the radius of the clump, $\rho_w$ and $\rho_c$ are the
1059: density, and $v_w$ and $v_c$ are the velocity of the wind and the
1060: clump, and the scaling parameters hold for typical 10 zone clumps in
1061: our model.  In reality, $t_{cc}$ may be substantially longer because
1062: the density ratios $\rho_c/\rho_w$ are probably underestimated: shell
1063: densities, and thus clump densities, are underresolved, while the wind
1064: density may be overestimated by our mass-loading scheme.  We need
1065: $\rho_c/\rho_w\approx10^{4}$ for these clumps to survive for more than
1066: $\sim5$ Myr, which may just be reachable in a rapidly diverging wind.
1067: 
1068: Moreover, recent numerical studies clumps may be stabilized against
1069: Kelvin-Helmholtz and R-T instabilities, reducing mass loss, by either
1070: thermal conduction (Marcolini et al.\ 2005, Vieser \& Hensler 2007),
1071: or even weak magnetic fields (Mac Low et al.\ 1994; Shin et al.\
1072: 2008).  Since we are not able to address the fate of shell fragments
1073: further in this study, we assume the bulk of the cool mass remains
1074: cool for the duration of starburst to be observed as Na~{\sc i}
1075: absorbing gas.  Larger, denser clumps are more likely to survive, in
1076: reality.
1077: 
1078: \subsection{Velocity Spread}
1079: \label{subsec:vspread}
1080: 
1081: The bottom left panels of Figure~\ref{line} and \ref{bline} 
1082: show the distribution of column density of cool gas $N_H$ as a function of 
1083: its radial velocity $v_r$ 
1084: at the end of the simulations. So long as the linear version of the
1085: ballistic approximation is correct, the velocity spread remains constant.
1086: Cool gas is seen over a wide range of velocities $>450\mbox{ km
1087:   s}^{-1}$ at both resolutions. Shells that fragment earlier have been
1088: accelerated less, and so their fragments travel more slowly.  
1089: 
1090: We now explore whether absorption from multiple shell fragments can
1091: explain the large absorption line widths measured for ultraluminous starbursts by
1092: considering the effects of resolution and viewing angle.
1093: Figure~\ref{lw} shows the distribution of velocity widths $\Delta v$
1094: seen in cool gas with $T<5\times10^4$ K in runs X1-0, X1, X1-2, and
1095: X1-4, with grid size increasing from 0.1 to 0.8~pc, along sightlines
1096: spaced by $1^{\circ}$ from the vertical axis.  We define $\Delta v$ as
1097: the difference between the maximum and the minimum velocities seen in
1098: the cool gas along a given sightline---that is, full width at zero.
1099: This resolution study demonstrates that the fraction of sightlines
1100: with large velocity spread increases as the resolution improves up to
1101: 0.2 pc, but then appears to converge.
1102: 
1103: Comparison to Figure~\ref{res} shows that the largest velocity widths
1104: are seen in sightlines intersecting the largest number of fragmented
1105: shells.  Better resolving post-shock densities in the swept-up,
1106: cooled, shells leads to quicker shell fragmentation and reformation,
1107: producing a larger number of shells, and thus the wider range of
1108: velocities in the fragments.
1109: 
1110: Figure~\ref{lw} also shows that most sightlines in X1 and X1-0 are above 
1111: the average observed line width $\langle v \rangle=320\pm120\mbox{ km s}^{-1}$ ({\em dashed lines}). 
1112: Observers measure the average of the sightlines toward continuum sources subtending
1113: about an arcsecond, and this corresponds to a length of
1114:   $1.84~h_{0.7}\mbox{ kpc}$ at redshift 0.1. 
1115: Assuming that the observers are likely to view ULIRGs at $t=5-10$ Myr
1116: after the onset of starburst, we can suggest all the shell fragments within
1117: $\sim20^{\circ}-40^{\circ}$ at a given sightline in our simulations will
1118: contribute to the absorption profiles.
1119: Hence, our models suggest that 
1120: observers will measure a large line width regardless of viewing angle.
1121: 
1122: To quantify this, we compute the average velocity width of all sightlines 
1123: in an axisymmetric cone within $\theta$ as 
1124: \begin{equation}
1125: \overline{\Delta v}(\le\theta) = \frac {\int_0^{\theta} \Delta
1126:   v(\alpha) \sin(\alpha) d\alpha} {\int_0^{\theta} \sin(\alpha)
1127:   d\alpha}. 
1128: \label{av}
1129: \end{equation}
1130: 
1131: The value of $\overline{\Delta v}$ is lowered by sightlines near the
1132: disk midplane where the bubbles are not blowing out and sightlines
1133: intersecting holes made by blowout with nearly zero $\Delta v$. The
1134: average velocity widths are clearly smaller for lower resolution runs,
1135: however. We find $\overline{\Delta v}(\theta\le30^{\circ})$ is
1136: 220~km~s$^{-1}$ for run X1-0, and 210~km~s$^{-1}$ for X1, but only
1137: 110~km~s$^{-1}$ for X1-4. 
1138: 
1139: \subsection{Average Velocity}
1140: \label{subsec:AMvelocity}
1141: 
1142: Figure~\ref{term} shows the distribution as a function of angle with
1143: degree spacing of mass-weighted average velocities of cool gas
1144: $v_{av}$ ({\em solid lines}) for our standard resolution study (runs
1145: X1-0 through X1-4).  The average shell velocities plotted may be
1146: misleading, since shell mass differs significantly at each sightline.
1147: Therefore, we also plot a mass-weighted average shell velocity across
1148: a $10^{\circ}$ arc centered on each sightline $v_{av,10}$.  This ought
1149: to be the quantity most directly comparable to observations of the
1150: average velocity.  
1151: 
1152: On sightlines with multiple fragments, the average mass-weighted
1153: velocity reflects the velocity of the more massive fragments.  These
1154: are fragments of the initial swept up shell, which follow ballistic
1155: orbits with little acceleration after shell fragmentation. Thus,
1156: average mass-weighted velocity tends to lie substantially below peak
1157: velocity.
1158: 
1159: The observable quantity $v_{av,10}$ shows a clear converging trend
1160: with resolution.  The converged value toward the pole appears to be
1161: under 400~km~s$^{-1}$, with correspondingly lower values at other
1162: angles, as shown in Figure~\ref{term}.  Figure~\ref{term} suggests
1163: that cool gas in shells and their fragments will be observed to be
1164: traveling with $v_{av,10}\approx200-350\mbox{ km s}^{-1}$ at angles to
1165: the pole $\theta<60^{\circ}$, the angle where blowout occurs in all
1166: models.
1167: 
1168: Our model is consistent with observations that constrain ULIRG wind
1169: geometry.  Figure~\ref{term} shows that absorption at velocities $v >
1170: 100$~km~s$^{-1}$ is detected at all angles greater than $10^{\circ}$
1171: from the disk plane.  It follows that 98 \% of all random, radial
1172: sightlines would exhibit such outflows. This result is consistent with
1173: 15 of 18 observed objects showing such outflows in Martin (2005), and
1174: a similar fraction seen by Rupke et al.\ (2005).
1175: 
1176: Figure~\ref{lwt} shows the distributions of $v_{av}$ and $v_{av,10}$
1177: for models X2 and X3 run with decreasing mechanical luminosity at the
1178: same resolution as model X1.  The mass-weighted average velocities of
1179: shells should depend on the mechanical luminosity $L_{mech}$ driving
1180: the bubble by thermal pressure. In a spherical bubble, the dependence
1181: would be $L_{mech}^{1/5}$. This dependence is actually slightly
1182: stronger in our models of blowout. We find $v_{av,10}$ dropping from
1183: 400~km~s$^{-1}$ to $\sim 100$~km~s$^{-1}$ moving from model X1 to X3,
1184: with each step having an order of magnitude lower mechanical
1185: luminosity.  This gives an empirical relation closer to
1186: $L_{mech}^{0.3}$.  Note the high spike in $v_{av,10}$ of X3 around
1187: $\sim15^{\circ}$ is biased by a small amount of cool mass present in
1188: its vicinity.  By way of comparison with observations, Figure 6 of
1189: Martin (2005) shows velocities reaching $v = 700$~km~s$^{-1}$ for an
1190: SFR of 1000~M$_{\odot}$~yr$^{-1}$. At an SFR of
1191: 1~M$_{\odot}$~yr$^{-1}$, the sparse data shown in that Figure suggest
1192: $v = 30$--40~km~s$^{-1}$, while our empirical relation would suggest a
1193: value $v \simeq 90$~km~s$^{-1}$.  More data at low SFRs and a broader
1194: range of models will be required to establish whether there is truly a
1195: discrepancy with the model results.
1196: 
1197: The mass-weighted average velocities $v_{av,10}$ in X1 and X2 agree
1198: with the observed shell velocity, but $v_{av,10}$ in X3 is well below
1199: the observed value.  In disks as massive as these, starbursts with
1200: mechanical luminosity lower than $10^{42}\mbox{ erg s}^{-1}$ produce
1201: the lower end of the outflow velocity range observed in ULIRGs, which
1202: average $v_{s,obs}=330\pm100\mbox{ km s}^{-1}$ in the Martin (2005)
1203: sample and $170\mbox{ km s}^{-1}$ in the column-density weighted
1204: average velocities from the Rupke et al. (2005) sample.
1205: Column-density weighting with velocity can be confounded by variation
1206: in covering factor with velocity, as has been shown for AGN outflows
1207: (e.g., Arav et al. 1999, 2005).  We cannot model covering factors well
1208: with our two-dimensional simulations, so we postpone consideration of
1209: the question of column-weighted velocity to future work. We do note
1210: that overpredicting acceleration is much harder to do than
1211: underpredicting it, though, so we think our model is likely to be
1212: robust. The velocity spread, which is our most important result from
1213: the simulations, is similar to that found in both ULIRG studies.
1214: 
1215: 
1216: \subsection{Terminal Velocity}
1217: \subsubsection{Resolution}
1218: \label{subsub:term-res}
1219: 
1220: Figure~\ref{term} also shows the distribution of terminal velocity of
1221: cool gas  $v_{term}$ at each sightline. We define the maximum
1222: velocity at blowout as the terminal velocity a shell will ever aquire,
1223: following the ballistic approximation
1224: (\S~\ref{subsec:ballistic}). The maximum velocity of shell fragments is
1225: very high, $v_{term}>500\mbox{ km  s}^{-1}$ at the angle $\theta<60^{\circ}$ where
1226: blowout occurs in all the runs. This high-velocity cool gas is found 
1227: in the fragments of the outermost shells in our highest resolution
1228: model X1-0. The hot wind continues to accelerate a piece of shell,
1229: sweeping up the ambient gas, until it fragments further.  As
1230: discussed above, shell fragmentation is very sensitive to resolution.
1231: Thus it is important to test the convergence of the mass of
1232: high-velocity gas.
1233: 
1234: Figure~\ref{hist}(a) shows the mass distribution of cool gas as a
1235: function of velocity in X1-0, X1, and X1-2 at $t=0.27$ Myr. The
1236: fraction of cool gas traveling at high velocities is low.  For
1237: example, the mass travelling with  $v\ge500\mbox{ km s}^{-1}$ is less
1238: than a few percent of the total shell mass, and the mass travelling
1239: faster than the observed terminal velocity of 750~km~s$^{-1}$ is
1240: $<0.1$\%. 
1241: 
1242: However, these results are best understood as upper limits. The
1243: amounts of cool gas in the high-velocity end are progressively smaller
1244: as the resolution increases.  Roughly factor of two decreases in the
1245: mass of cool gas with velocities above 500~km~s$^{-1}$ occur between
1246: runs with factor of two improvements in linear resolution.  This
1247: suggests that we have not yet converged on the actual mass of
1248: high-velocity gas, although we have set good upper limits.
1249: We think this high-velocity gas is not likely to go away entirely,
1250: even if we run a simulation with a higher resolution and with
1251: additional physics. However, we can not fully quantify its amount with
1252: our simulations.
1253: 
1254: The convergence properties of the peak velocity for cold gas
1255: $v_{term}$ are somewhat better.  The highest resolution models show $<
1256: 20$\% variations, suggesting that the general result is reasonably
1257: robust. The observed maximum value of 750~km~s$^{-1}$ is consistent
1258: with the models at angles $\theta<60^{\circ}$.
1259: 
1260: \subsubsection{Mass Loading}
1261: 
1262: Figure~\ref{termt1} shows the distributions of $v_{av}$, $v_{av,10}$,
1263: and $v_{term}$ for simulations with different mass-loading rates and
1264: so different wind terminal velocities: U1-A, U1, U1-B, and U1-C.  Cool
1265: gas has higher terminal velocities in bubbles with hotter winds that
1266: themselves have higher velocities $v_{wind}$.  The hottest wind, in
1267: model U1-A, has $v_{wind}=2700\mbox{ km s}^{-1}$, while the coldest
1268: wind, in U1-C, has $v_{wind}=250\mbox{ km s}^{-1}$.
1269: Figure~\ref{hist}(b) shows the mass distribution of cool gas as a
1270: function of velocity for the four runs at the time of blowout. The
1271: cool gas with high $v_{term}$ carries only a small amount of mass (see
1272: \S~\ref{subsub:term-res}).  For example, the fraction of cool gas with
1273: $v>500\mbox{ km s}^{-1}$ is $\la 2\%$ of the total shell mass, and the
1274: fraction with $v>750\mbox{ km s}^{-1}$ is $\la0.5\%$.  The bulk of
1275: swept-up and cooled gas is driven to $v\sim400\mbox{ km s}^{-1}$ by
1276: thermal pressure of hot sonic gas, but a small fraction of it seems to
1277: be accelerated to higher velocity by the ram pressure of the same hot
1278: gas as it accelerates to supersonic velocities during blowout.
1279: 
1280: \subsubsection{Mechanical Luminosity}
1281: 
1282: The terminal wind velocity (equation~\ref{ewind}) in our models X2 and
1283: X3 with lower $L_{mech}$ is the same as that of our fiducial model X1, 
1284: because the ejected mass $M_{SN}\propto L_{mech}$, and we keep the
1285: mass-loading factor $\xi$ constant. In all three models
1286: $v_{wind}\approx1000\mbox{ km s}^{-1}$.  
1287: 
1288: Figure~\ref{lwt} shows that the terminal velocities of cool gas are
1289: not as high as the observed average terminal velocity,
1290: $v_{t,obs}=750\mbox{ km s}^{-1}$ at most sightlines if
1291: $L_{mech}\le10^{42}\mbox{ erg s}^{-1}$. It is much harder for the wind
1292: ram pressure to accelerate the outermost shells and their fragments to
1293: very high velocity, $>500\mbox{km s}^{-1}$ if their starting velocity
1294: at blowout $v_b \la200\mbox{ km s}^{-1}$.
1295: 
1296: \subsubsection{Summary}
1297: 
1298: In summary, the terminal velocity of cool gas is determined by the
1299: combination of total mechanical power $L_{mech}$ and wind terminal
1300: velocity $v_{wind}$ (determined by the mass-loading rate $\xi$ in our
1301: model). Bubble thermal pressure accelerates the bulk of shells to
1302: their final velocity before blowout and ram pressure of the hot
1303: transonic wind after blowout accelerates a small fraction of the cool
1304: gas to much higher velocities.  Faster, less mass-loaded winds
1305: accelerate cool gas to higher velocities.  Our wind is an
1306: energy-driven wind, not a momentum-driven wind (Murray et al. 2005).
1307: The thermal energy of the hot wind is gradually turning into its
1308: kinetic energy by blowout (recall radiative cooling is turned off
1309: inside the wind: see \S~\ref{sec:numerics}).  Our model accounts for
1310: the observed $750\mbox{ km s}^{-1}$ terminal velocity of Na~{\sc i}
1311: absorbing gas without invoking any additional physics such as
1312: radiation pressure.
1313: 
1314: \subsection{Absorption Profiles}
1315: \label{subsec:profiles}
1316: 
1317: To directly compare to observed profiles, we generate Na~{\sc i}
1318: $\lambda 5890$ absorption line profiles along sightlines through our
1319: simulations, as well as generalizing this procedure to fully model the
1320: observed doublet Na~{\sc i} $\lambda,\lambda 5890,5896$. To generate the
1321: profiles, we begin with the line intensity
1322: \begin{equation}
1323: I_{\nu}=I_{\nu}(0) \exp^{-\tau_{\nu}}=I_{\nu}(0) \prod_{i=1}^N \exp^{-\tau_i}
1324: \label{I}
1325: \end{equation}
1326: where the optical depth through cell $i$ at frequency $\nu$ is
1327: $\tau_{i}$, the sightline intersects $N$ cells, and the background
1328: continuum is $I_{\nu}(0)$.  We normalize the continuum by setting
1329: $I_{\nu}(0)=1$ and compute the profile as a function of the
1330: macroscopic velocity of Na~{\sc i} absorbing gas, $v$. We compute
1331: $I_v$ in the simulations by setting $v=c(\nu-\nu_0)/\nu_0$, and
1332: computing the optical depth contributed by each cell in each of 1000
1333: velocity bins.  The optical depth in each cell
1334: \begin{equation}
1335: \tau_{i}(v)=N_{NaI}~s~\lambda_{NaI} P(\Delta v), 
1336: \label{tau}
1337: \end{equation}
1338: (Spitzer 1978) where $N_{\mbox{Na~{\sc i}},i}$ is the column density
1339: of Na~{\sc i} in a cell $i$, and the absorption cross section
1340: $s_{\nu}$ integrated over frequency $\nu$ for $h\nu\gg kT$ is $s
1341: \approx 2.654\times10^{-2}f_{5890}$ with the oscillator strength
1342: $f_{5890}=0.6$.  The Maxwellian velocity distribution function,
1343: $P(\Delta v)$, is
1344: \begin{equation}
1345: P(\Delta v)=\frac{1}{\sqrt{\pi}b}\exp{-(\Delta v^2/b^2)}, 
1346: \end{equation}
1347: with $\Delta v=v-v_i$ and $b=\sqrt{2k_B T_i/\mu m_H}$ for thermal
1348: broadening.  Our simulations do not track chemical abundances,
1349: ionization state, or dust depletion, so we do not directly predict the
1350: Na~{\sc i} column.  For the purpose of illustration, we compute the
1351: total Na~{\sc i} column from the total H~{\sc i} gas column using the
1352: conversion used by Martin (2005) to estimate $N_{\mbox{H~{\sc i}}}$
1353: from their $N_{\mbox{Na~{\sc i}}}$ measurements, $N_{\mbox{Na~{\sc
1354: i}}} = 1.122\times10^{-6}N_{\mbox{H~{\sc i}}}$.
1355: 
1356: We generated Na~{\sc i} 5890 line profiles along lines of sight
1357: through models X1 and X1-0 to show the effect of numerical resolution
1358: on the line profiles in Figure~\ref{NaI}.  Each sightline is described
1359: by its inclination from the polar axis of the simulation.  The profile
1360: shapes reflect the different structure in these simulations.  We can
1361: for example compare the absorption profiles in the middle left panel
1362: in Figure~\ref{NaI} with the density and velocity distributions of
1363: shells in Figures~\ref{line} and \ref{bline}. The sightlines at
1364: $\theta=19^{\circ}$ intersect three and four shell fragments in X1 and
1365: X1-0 respectively. Each of these fragments generates a $\Delta v \ga
1366: 50-100~\mbox{km s}^{-1}$ absorption line.  A few of these lines are
1367: optically thick with the line profile becoming completely black at the
1368: line center. As the wind expands, column densities will drop linearly
1369: with radius due to geometric dilution.  The complex of lines is spread
1370: out over $\sim400~\mbox{km s}^{-1}$.
1371: 
1372: Figure~\ref{NaI} shows that the line profiles from the high and low
1373: resolution models present very similar structure for the most part.
1374: One minor exception is the minimum velocity.  From $10\deg$ to $20\deg$, the
1375: line profiles show a sharp cut-off at a velocity $\sim500\mbox{ km
1376: s}^{-1}$ in X1, but $\sim3-40\mbox{ km s}^{-1}$ in X1-0. 
1377: This cut-off reflects the
1378: velocity of the first R-T fragments that form from the accelerating
1379: swept-up shells, which is lower in the high-resolution simulation.
1380: The overall lineshape does appear reasonably well converged.
1381: 
1382: Observers see the average over parallel sightlines subtending a few
1383: kiloparsecs of the disk.  They must also contend with the instrumental
1384: response function, and the doublet nature of the Na~{\sc i} line,
1385: which has components at 5890 \AA\ and 5896 \AA, with optical depths
1386: differing by a factor of $\sim 2$. To compare to the actual observed
1387: profiles, we generate the Na~{\sc i} 5890/5896 doublet in the frame of
1388: Na~{\sc i} 5890 by assuming the ratio of equivalent widths is only
1389: 1.3, typical of the observed lines in ULIRGs (rather than the
1390: optically thin limit of two).
1391: 
1392: We present average line profiles of the doublet over a $20\deg$
1393: wide set of radial sightlines spaced at degree intervals.  Although
1394: this is not exactly what is measured, it is comparable because these
1395: rays will subtend about 1-2~kpc of the shell when the bubble is $\sim
1396: 10\mbox{ Myr old}$, so they intersect the same region of the shell,
1397: albeit at slightly different angles.  We are actually doing the
1398: analysis at an earlier time, when the bubble is a factor of 100
1399: smaller, and correspondingly higher column density, though, so we also
1400: must apply a geometric dilution factor to the column densities.
1401: 
1402: Figure~\ref{NaI20} compares our simulated doublet line profiles, with
1403: and without geometric dilution of the column densities by a factor of
1404: 100, to five typical ULIRG spectra from Martin (2005) that have an
1405: instrumental broadening with FWHM$\simeq 65$~km~$^{-1}$. We convolve
1406: in a Gaussian of that width to simulate the broadening.  The widths of
1407: the absorption profiles are very similar between our model at
1408: $\theta\le50^{\circ}$ where blowout occurs and the observations.  The
1409: undiluted absorption in our model much exceeds that observed.  Our
1410: model overpredicts the cool gas column for two reasons (see
1411: \S\ref{subsec:coolgas}). First, because we analyzed the models at a
1412: very early time in the starburst.  At later times, the spherical
1413: expansion of the wind and the cold gas embedded within it dilutes
1414: their column densities. Second, the fragmentation of the shell is
1415: incompletely resolved, so some material remains cool that in reality
1416: would have been mixed into the hot wind. We account for the first
1417: factor by directly reducing the column densities, producing the
1418: diluted profiles shown in the Figure, which reproduce the observed
1419: intensities well.
1420: 
1421: The ULIRG spectra show absorption beginning from the systemic
1422: velocity, and extending to high velocities. Although this zero
1423: velocity absorption is absent in Figure~\ref{NaI}, it is seen in the
1424: simulated doublet profiles shown in Figure~\ref{NaI20}, as it comes
1425: primarily from contributions from the $\lambda$5896 line of the
1426: doublet blueshifted into zero velocity of the $\lambda$5890 line.
1427: (Note our model does not include the contribution from stellar
1428: absorption).
1429: 
1430: \subsubsection{Long Slit Observations}
1431: 
1432: 
1433: Measured Doppler shifts across ULIRGs show the cool outflow is extended
1434: spatially;  and some outflows present a significant velocity gradient
1435: over kpc scales (Martin 2006). 
1436: Our simulation ends long before the wind has reached such large scales
1437: and does not include the rotation of the galactic disk. The simulation 
1438: does demonstrate, however, the amplitude of the velocity gradient
1439: that arises across the minor axis due to projection effects.
1440: 
1441: 
1442: The wind is launched without any net angular momentum in our 2D simulation.  
1443: We examined whether, in the absence of rotation, the
1444: position affected the velocity much. 
1445:   Figure~\ref{30} shows terminal and mass-weighted average velocity
1446:   along slits oriented at $30^{\circ}$ from the major and minor axes
1447:   of the disk, using parallel lines of sight through a 3D rotation of
1448:   the 2D simulation. 
1449: The absorption properties along the major axis of the galaxy will be
1450: symmetric about its center.  An asymmetric velocity gradient is
1451: produced along the minor axis because one side of the outflow cone is
1452: directed more along the sightline.  Substantial variations in velocity
1453: width and average velocity are seen at $\sim40$ pc and $\pm30$
1454: pc. These occur where sightlines intersect massive, slow-moving shells
1455: at $\theta>45^{\circ}$ (see Figure~\ref{term}) as well as light,
1456: fast-moving blowout fragments at $\theta<45^{\circ}$. On the other
1457: hand, very high terminal velocities are seen on most sightlines since
1458: they intersect the fast blowout components.
1459: 
1460: The simulated  sightlines encounter both fast-moving and slow-moving gas at
1461: various latitudes, unlike the radial sightlines that we studied in previous
1462: sections. Our simulations viewed along non-radial sightlines 
1463: naturally account for a wide velocity range of cool gas starting at
1464: $v\approx0$ and a high terminal velocity.
1465: 
1466: 
1467: \subsection{Line Widths of Na~{\sc i} Absorbing and H$\alpha$ Emitting Gas}
1468: 
1469: Finally, we crudely try to separate low-ionization, Na~{\sc i} absorbing gas 
1470: and ionized H$\alpha$ emitting gas in the cooled shell fragments in our simulations. 
1471: Our aim is to compare the line widths of both components.
1472: Heckman et al.\ (2000) observed no correlation between the absorption
1473: and emission line widths for these lines, and drew the conclusion that
1474: only one of the lines could originate in the swept-up shells. 
1475: 
1476: To model the H$\alpha$ emission, we must decide whether the emitting
1477: gas is primarily photoionized by the central starburst or shock
1478: ionized in the wind.  Shock ionization dominates at least some
1479: starburst-driven winds, such as NGC~1482 (Veilleux \& Rupke 2002).
1480: Some observed ULIRGs also show extended shock-excited nebulae
1481: (Monreal-Ibero et al.\ 2006). However, the dynamics observed by
1482: Heckman et al.\ (2000) seem unlikely to come from shock ionized gas,
1483: since they see no correlation with the motions of the cold gas that
1484: presumably trace the shock.  Therefore we assume photoionization by
1485: the central starburst, choosing an ionizing photon luminosity $Q$ and
1486: a density distribution from a time the galaxy is likely to be
1487: observed.  Figure~\ref{lmech} shows the ionizing photon luminosities
1488: as a function of time for three starburst cases that we consider.
1489: Since the propagation of ionization fronts critically depends on both
1490: the densities and the positions of shells, we can not extrapolate the
1491: density distribution at blowout to $t\approx10$ Myr using the
1492: ballistic approximation. We can, however, still vary the strength of
1493: photon luminosity over a few orders of magnitude to examine the effect
1494: of attenuating the photon flux by $1/r^2$ on our existing simulations.
1495: For example, a shell will experience about three orders of magnitude
1496: less photons per unit area after it travels with $500\mbox{ km
1497:   s}^{-1}$ for 10 Myr from $r_b=200$ pc. For our purpose of
1498: demonstrating the lack of correlation between the line width of
1499: Na~{\sc i} absorbing and $H\alpha$ emitting gas, this crude method is
1500: sufficient.
1501: 
1502: To solve for the transfer of ionizing radiation across our grids, we
1503: use the photon-conserving radiative transfer code developed by Abel et al.\ (1999) in the same manner as it was used in Fujita et al.\ 
1504: (2003).  It computes the propagation of ionization fronts around a point
1505: source; in our study this is a central starburst source. This code is
1506: a post-processing step that operates on a given density distribution
1507: at a given time step in our simulation. However, the ionization fronts
1508: propagate sufficiently rapidly to adjust almost instantaneously to a
1509: changing density distribution.
1510: 
1511: Figure~\ref{ion} shows the column density as a function of velocity
1512: for Na~{\sc i} absorbing gas ({\em diamonds}) and H$\alpha$ emitting
1513: gas ({\em triangles}) in run U1 at $\theta=13^{\circ}$ with ionizing
1514: luminosities of $Q=10^{53}, 10^{54}, 1.9\times10^{54}$, and $10^{55}$
1515: photons s$^{-1}$.  With $Q\le10^{54}$~photons~s$^{-1}$, the line
1516: widths are very small for H$\alpha$ emitting gas, $\Delta v_{H\alpha}<
1517: 100\mbox{ km s}^{-1}$, but large for Na~{\sc i} absorbing gas, $\Delta
1518: v_{NaI}=480\mbox{ km s}^{-1}$.  The densest shell at $r=110$ pc traps
1519: the photons rather effectively.  As Q is increased, that first shell
1520: is ionized, but the second shell then traps the photons.  The {\em
1521:   bottom left} panel shows that the Na~{\sc i} line width is now very
1522: small, $\Delta v_{NaI}=38\mbox{ km s}^{-1}$, while $\Delta
1523: v_{H\alpha}=330\mbox{ km s}^{-1}$.  With $Q\ga 2\times10^{54}$, all
1524: the shells are ionized.  In reality, the transition from large $\Delta
1525: v_{NaI}$ to $\Delta v_{NaI}\approx0$ should not be this abrupt because
1526: many combinations of fragments and clumps are possible within the
1527: observers' $10^{\circ}\times10^{\circ}$ field of view.  As we note, we
1528: are far from presenting realistic distributions of Na~{\sc i}
1529: absorbing and H$\alpha$ emitting gas. However, Figure~\ref{ion} still
1530: demonstrates that the line widths of Na~{\sc i} absorbing and
1531: H$\alpha$ emitting gas can easily be uncorrelated although they both
1532: originate in the swept-up shells.
1533: 
1534: \section{Caveats and Summary}
1535: 
1536: \subsection{Caveats}
1537: 
1538: This study clearly is not the final word on this subject.  Rather it
1539: is a proof of the concept that a starburst wind can accelerate neutral
1540: gas up to high velocities without special circumstances, and that the
1541: resulting flow configuration can reproduce many of the observable
1542: properties of ULIRG winds.  We here recap the approximations we have
1543: made in order to treat this problem, and add some discussion of how
1544: they might be lifted.
1545: 
1546: The biggest approximation of our study is that we analyze the
1547: kinematics of shell fragments at $t\ll 1$ Myr in our small grids and
1548: extrapolate the results by the ballistic approximation including
1549: geometric dilution for comparison with the observations at $t \gg 1$
1550: Myr.  Much of the total mechanical energy from even an instantaneous
1551: burst of star formation has yet to be deposited at an age of 2~Myr
1552: when our simulation ends.  These dense clumps may further fragment and
1553: ultimately be destroyed by R-T and Kelvin-Helmholtz instability as the
1554: high-velocity flow of gas from within the bubble streams through them.
1555: Conversely, they may be dense enough to be accelerated further prior
1556: to their destruction by the ongoing starburst wind.
1557: 
1558: As the grid resolution of our models improves to 0.1 pc, we do begin
1559: to resolve the details of hydrodynamic instabilities.  Marcolini et
1560: al.\ (2005) used 0.1~pc resolution for their simulations of the
1561: dynamical shredding of radiatively cooling clouds, and seem to have
1562: reached adequate resolution.  Thus, in future work, continuing to
1563: evolve our bubbles in bigger grids with the same resolution may well
1564: improve our results.
1565: 
1566: We have made four other substantial physical approximations as
1567: well. First, the assumption of aziumthal symmetry further suppresses
1568: R-T instability. Mac Low et al.\ (1989) pointed out that the typical
1569: spike and bubble structure of the R-T instability is limited to rings
1570: in an axisymmetric blowout.  The detailed distribution of the
1571: fragments will certainly be different in three dimensions, but finer
1572: fragments will probably have a broader velocity range, and thus are
1573: unlikely to change our qualitative result.  The behavior of individual
1574: fragments will also not be qualitatively changed by the
1575: dimensionality, although two-dimensional fragments split into larger
1576: pieces initially (Stone \& Norman 1992b; Korycansky et al.\ 2002).
1577: 
1578: Second, we neglected magnetic fields in this study.  We showed that
1579: magnetic fields reduce the peak shell density, and thus the
1580: fragmentation (\S~\ref{subsec:effects-res}), but that they don't act
1581: to do so before fragmentation is essentially complete in our model, so
1582: that this effect is secondary.  Magnetic fields can, on the other
1583: hand, help preserve individual fragments from further breakup (Stone
1584: \& Norman 1992b; Mac Low et al.\ 1994; Shin et al.\ 2008), possibly
1585: even allowing their further acceleration in the continuing starburst
1586: wind.
1587:   
1588: Third, thermal conduction is not explicitly treated in our model.
1589: This acts on parsec scales, so numerical diffusion and turbulent
1590: mixing will still dominate over thermal conduction under most
1591: circumstances at the resolutions that we consider.  It is worth
1592: noting, though, that studies of individual clump fragmentation have
1593: found that thermal conduction can have stabilizing effects (Marcolini
1594: et al.\ 2005; Vieser \& Hensler 2008).  Fourth, the dynamical effects
1595: of photoionization have also been neglected.  This could heat at least
1596: low column-density shells and fragments up to $10^4$~K, reducing their
1597: density contrast with the wind and enhancing their tendency to
1598: fragment.
1599: 
1600: We also assume a single starburst at the disk center for modeling a
1601: galactic outflow. In realistic galaxies, we expect multiple star
1602: clusters to be scattered around the disks (e.g.\ Vacca 1996; Martin
1603: 1998). A more complicated structure of cool shells and their fragments
1604: is expected with a realistic distribution of star formation as shown
1605: by models of supernova-driven turbulence in our own galaxy (e.g
1606: Avillez 2000; Avillez \& Berry 2001; Joung \& Mac Low 2006). Such a
1607: distribution was modeled with a fractal density distribution by Cooper
1608: et al.\ (2008).  Fragile et al.\ (2004) showed that in dwarf galaxies,
1609: distributed supernova explosions resulted in increased transfer of
1610: energy to the cold gas compared to the centrally concentrated energy
1611: injection assumed here and by Mac Low \& Ferrara (1999), so our
1612: approximation likely gives a lower limit to the kinetic energy of the
1613: cold gas. This supports our result of a wide velocity range in Na~{\sc
1614:   i} absorbing gas.  The next obvious step is to study the effects of
1615: realistic star formation on the shell kinematics in three dimensional,
1616: AMR simulations.
1617: 
1618: Another related issue is that our models assume a low-density,
1619: uniform, gas distribution above the site of blowout. However, ULIRGs
1620: are formed in galaxy mergers. The actual situation near the nucleus of
1621: merging galaxies is likely to be more chaotic.  We rely on the idea
1622: that the large-scale tidal tails and other merger structures are
1623: generally going to be well removed from the wind generation region and
1624: will not cover a large fraction of the sky.  Ultimately, to remove
1625: this limitation, full models of merging galaxies at high resolution
1626: will be needed, but this remains some years in the future.
1627: 
1628: \subsection{Summary}
1629:     
1630: We study the origin and the kinematics of cool gas that produces
1631: Na~{\sc i} absorption lines in galactic winds, using hydrodynamic
1632: simulations of the blowout of starburst-driven superbubbles from the
1633: molecular disk of ULIRGs.  The bubbles sweep up the dense disk gas,
1634: which quickly cools to form dense, thin shells.  The cooled shells
1635: fragment by R-T instability as they accelerate in the stratified
1636: atmosphere of the disk.  This blowout occurs very early in our models,
1637: at $t \ll 1$ Myr in models with $L_{mech}\ge10^{41}\mbox{ erg
1638: s}^{-1}$.  The dense fragments left by the R-T instability lag behind
1639: the low-density, high-velocity wind.  These fragments carry most of
1640: the mass that is swept up by the bubbles, and should have the highest
1641: column of Na~{\sc i}.  The results of our numerical convergence study
1642: suggest that superbubble blowout, combined with subsequent geometrical
1643: dilution, can reproduce not just qualitative but quantitative
1644: properties of the observed lines.
1645: 
1646: As a result of R-T fragmentation, multiple shell fragments and clumps
1647: travel at different velocities. A sightline going through them
1648: reproduces the observed broad line width of the Na~{\sc i} absorption
1649: profiles: $\langle v \rangle=320\pm120\mbox{ km s}^{-1}$.  This result does not
1650: appear to depend strongly on physical parameters such as mass-loading
1651: rate, mechanical luminosity, or surface density.  However, this result
1652: requires sufficient numerical resolution to follow secondary
1653: fragmentation of the shell, so that any line of sight runs over
1654: multiple cold gas fragments.  We find that a resolution of at least
1655: $\sim0.2$ pc is required to produce this effect, if mass-loading rates
1656: remain moderate $\xi\la10$. The suggestion by Heckman et al.\ (2000) that 
1657: swept-up shells will not show a wide range in velocity is based on a hydrodynamic
1658: simulation of M82 with a resolution of 4.9 pc (Strickland \& Stevens
1659: 2000), more than an order of magnitude worse. 
1660: 
1661: The mass-weighted average velocity of cool gas in galactic outflows is
1662: found to be $\sim400$--500~km~s$^{-1}$ among all the models with
1663: $L_{mech}=10^{43}\mbox{ erg s}^{-1}$, but $\la200\mbox{ km s}^{-1}$ in
1664: models with $L_{mech}=10^{42}$ and $10^{41}\mbox{ erg s}^{-1}$.  
1665: The bulk of shells and their fragments are accelerated by the thermal
1666: pressure of the bubble interior gas, which is proportional to
1667: $L_{mech}$. No other parameters influence the results.  The
1668: mass-weighted average velocities in our high-resolution simulation
1669: with $L_{mech}=10^{42-43}$~erg~s$^{-1}$ agree with the observed
1670: value $v_{s,obs}=330\pm100\mbox{ km s}^{-1}$.  A mechanical luminosity
1671: of $L_{mech}=10^{43}\mbox{ erg s}^{-1}$ corresponds to an
1672: instantaneous burst of $M_*=10^9~\mbox{ M}_{\odot}$ or a continuous
1673: SFR$=500~\mbox{ M}_{\odot}\mbox{ yr}^{-1}$ at $t<1$ Myr after the
1674: onset of starburst.
1675: 
1676: As the swept-up shells fragment by R-T instability, the bubble
1677: interior gas blows out and becomes a low-density supersonic wind, as
1678: its thermal energy is converted to kinetic energy by expansion.  The
1679: ram pressure of this wind continues to accelerate entrained cold
1680: fragments, and to sweep up high-altitude disk gas, producing small
1681: amounts of cold gas with velocities of the order of the terminal
1682: velocity of the wind.  The correlation between wind terminal
1683: velocities and terminal velocities of cool gas in simulations with
1684: different mass-loading rates supports this picture.  If the wind
1685: velocity reaches $v_{wind}\ga1000\mbox{ km s}^{-1}$, a significant
1686: fraction of sightlines through a simulation encounters a terminal
1687: velocity close to or greater than the observed average terminal
1688: velocity of Na~{\sc i} absorption profiles $v_{t,obs}=750\mbox{ km
1689:   s}^{-1}$.  The mass of cool gas with such high velocity is less than
1690: a few percent of that of the total cool gas, however.  These outermost
1691: shells and fragments with high velocity do fragment further in our
1692: highest-resolution simulation, but their fragments are traveling with
1693: similarly high velocity.  Although the mass in the high-velocity tail
1694: decreased as the resolution is increased, we believe it is not likely
1695: to go away entirely in a higher-resolution simulation.
1696: 
1697: The clumps and fragments seen in the simulations may be observed as
1698: Na~{\sc i} absorbing gas or H$\alpha$ emitting gas, depending on the
1699: amount of ionizing photons produced from the central starburst source.
1700: To study this, we used a ray-tracing method to model the location of
1701: the ionization front in our models as a function of time, and so to
1702: trace the gas emitting in H$\alpha$.  By varying the photon
1703: luminosity, we showed that the velocity range of the ionized and
1704: neutral components do not show any correlation with each other. Thus
1705: the lack of correlation in the observed line widths of Na~{\sc i}
1706: abosorbing gas and H$\alpha$ emitting gas does not rule out the
1707: swept-up shells as the origin of both components.
1708: 
1709: The hot wind in our model is purely energy-driven. By construction the
1710: interior bubble can never become momentum-driven because radiative
1711: cooling is turned off in the bubble interior in our models.   Future work must
1712: determine whether this mechanism reproduces the empirical correlation
1713: observed between the maximum outflow velocity and the escape velocity
1714: of the host galaxy (Martin 2005). It should also predict the scaling of the
1715: mass-loss efficiency with galaxy mass, a key input in cosmological
1716: models (e.g.\ Oppenheimer \& Dav\'e 2006, 2008)
1717: 
1718: \acknowledgements We thank M. L. Norman and the Laboratory for
1719: Computational Astrophysics for use of ZEUS-3D, R. F. Coker, G. Gisler,
1720: R. M. Hueckstaedt, and C. Scovel for the help with getting started
1721: with SAGE, and M. K. R. Joung for useful discussions. We thank the
1722: referee, D.\ Rupke, for his extensive assistance in placing this work
1723: in context. Computations were performed on the SGI Origin 2000 of the
1724: Rose Center for Earth and Space, and the QSC machine of Los Alamos
1725: National Laboratory (LANL). AF was supported by a cooperative
1726: agreement between UCSB and LANL.  CLM thanks the Alfred P. Sloan
1727: Foundation and the David and Lucile Packard Foundation for support for
1728: this work.  M-MML was partly supported by NSF grants AST99-85392 and
1729: AST03-07854, and by stipends from the Max-Planck-Gesellschaft and the
1730: DAAD.
1731: 
1732: \clearpage
1733: 
1734: \appendix
1735: 
1736: \section{Appendix: Blowout in a Dwarf Galaxy with ZEUS and SAGE}
1737: 
1738: A major issue in our simulations is whether we sufficiently resolve
1739: the unstable shell during blowout.  To further examine this question,
1740: in this Appendix we describe a comparison between models of blowout
1741: from a dwarf galaxy similar to that described by Fujita et al.\ (2003)
1742: run with ZEUS-3D, and single-grid and AMR versions of SAGE.
1743: 
1744: SAGE is an AMR hydrodynamic code developed at LANL/SAIC.  It
1745: is second order accurate using a piecewise linear Godunov scheme
1746: (Kerbyson et al.\ 2001; Gittings et al.\ 2008).
1747: 
1748: The dwarf galaxy has a nominal redshift $z=8$, and disk mass $M_d=10^8
1749: \mbox{ M}_{\odot}$.  We choose the midplane density of this galactic
1750: disk so that it has an exponential surface density profile.  We set up
1751: the gas in hydrostatic equilibrium with a background dark matter halo
1752: potential, with an effective turbulent velocity of $10\mbox{ km
1753:   s}^{-1}$.  Mass and energy are injected at the center of the disk,
1754: corresponding to a mechanical energy of starburst, $10^{40} \mbox{ erg
1755:   s}^{-1}$, with a mass-loading rate of 0.1 $\mbox{ M}_{\odot} \mbox{
1756:   yr}^{-1}$.  See the details of disk and halo parameters in Fujita et
1757: al.\ (2003).  
1758: 
1759: In SAGE, the cooling is solved with an explicit method that subcycles
1760: the cooling source terms, based on a radiative cooling routine that
1761: solves for non-equilibrium chemistry (Abel et al.\ 1997).  The
1762: timesteps for each subcycle are determined as $\epsilon e/\dot{e}$
1763: where $e$ is internal energy density and $\epsilon=0.1$.  In ZEUS-3D,
1764: on the other hand, we assume that the cooling rate is a function of
1765: temperature only, using calculations of equilibrium ionization cooling
1766: rates by Sutherland \& Dopita (1993) with a cut-off at $T < 10^4$ K,
1767: and ignore inverse Compton cooling for simplicity.  The cooling curve
1768: is implemented in the energy equation with a semi-implicit method,
1769: using a Newton-Raphson root finder.
1770: 
1771: In order to avoid overcooling in the bubble interior (see \S 4), we advect two materials
1772: independently in SAGE; 1) disk and halo gas and 2) hot, metal-enriched
1773: gas injected at the starburst site, 
1774: and turn off cooling in the latter. SAGE computes hydrodynamics of a
1775: multi-material flow, 
1776: assuming all materials in a cell have the same pressure. 
1777: The time required to process mixed cells is linearly dependent on the number of materials.
1778: In ZEUS-3D we use the tracer field (Yabe \& Xiao 1993) to suppress
1779: interior cooling.
1780: 
1781: We ran simulations with and without radiative cooling,
1782: and with and without AMR.
1783: Table~\ref{sz} shows the parameters for our computations. 
1784: With SAGE we use both uniform and AMR grids 
1785: having a maximum effective resolution of 0.0693~pc.
1786: With ZEUS, we use a ratioed grid 
1787: having the same central resolution as the lowest resolution SAGE runs,
1788: and decreasing resolution outside the innermost $0.1\times0.2$
1789: kpc$^2$. 
1790: 
1791: SAGE was run on the QSC cluster at LANL using 12 Alpha processors for
1792: the UN, UC, AN, and AC runs, 24 processors for the BC run, and 48
1793: processors for the CC run, while ZEUS-3D was run on a SGI Origin 2000
1794: using 8 processors.  SAGE's performance is linear up to at least a few
1795: hundred processors (Kerbyson et al.\ 2001), while the loop-level
1796: parallelized performance of ZEUS-3D levels off after 8 processors (a
1797: massively parallel, domain-decomposed version, ZEUS-MP, has linear
1798: performance to over 512 processors).  The number of active cells used
1799: for the AMR simulations with 5 and 6 levels of refinement is less than
1800: the number of cells used for the ZEUS-3D simulations. The CPU time per
1801: cycle is much smaller with SAGE compared to that with ZEUS-3D, if
1802: cooling is included. ZEUS-3D spends the most time searching for
1803: convergence around the cut-off temperature of the cooling curve, since
1804: radiative cooling in our dense, high-redshift disk is very efficient.
1805: 
1806: Figure~\ref{compare} shows the density distributions of our models at
1807: the times when the bubbles blow out of the disks.  This time was
1808: chosen so that the positions of outer shock fronts in the horizontal
1809: direction agree.  Figure~\ref{compare} shows qualitative agreement
1810: between the results with SAGE and ZEUS.  In our no-cooling runs, the
1811: positions of outer shock fronts agree within $\sim2\%$.  Those of
1812: inner shock fronts agree within $\sim5\%$ if SAGE is run in the
1813: uniform grid and $\sim10\%$ if SAGE is run with AMR, because we chose
1814: an AMR refinement criterion that did not act in that region, while the
1815: uniform grid had the resolution of the highest AMR refinement level.
1816: 
1817: The linear piecewise advection method of SAGE seems more diffusive
1818: than the Van Leer (1977) method of ZEUS.  However, we show in
1819: Figure~\ref{compare} that the SAGE/AMR run begins resolving fine
1820: structures caused by R-T instability as well as ZEUS-3D run when one
1821: more level of refinement is added, doubling the maximum effective
1822: resolution.  Even in this case, the SAGE/AMR run still uses a smaller
1823: number of total and active cells than the ZEUS-3D run. We can save
1824: noticeable computational time with SAGE/AMR as we tackle a problem
1825: with more than a few million cells. This advantage will be bigger with
1826: 3D models.
1827: 
1828: \newpage
1829: \begin{thebibliography}{}
1830: \bibitem[Abel et al.(1997)]{aazn97} Abel, T., Anninos, P., Zhang, Y.,
1831: \& Norman, M. L. 1997, New Astron., 2, 181
1832: \bibitem[Abel et al.(1999)]{anm99} Abel, T., Norman, M. L., \& Madau,
1833: P. 1999, \apj, 523, 66
1834: \bibitem[Arav et al.(1999)]{ara99} Arav, N., Becker, R. H.,
1835:   Laurent-Muehleisen, S. A., Gregg, M. D., White, R. L., Brotherton,
1836:   M. S., \& de Kool, M. 1999, \apj, 524, 566
1837: \bibitem[Arav et al.(2005)]{ara05} Arav, N., Kaastra, J., Kriss,
1838:   G. A., Korista, K. T., Gabel, J., \& Proga, D. 2005, \apj, 620, 665
1839: \bibitem[Avillez(2000)]{a00} Avillez, M. A. 2000, \mnras, 315, 479
1840: \bibitem[Avillez \& Berry(2001)]{ab01} Avillez, M. A., \& Berry, D. L. 2001,
1841:   \mnras, 328, 708
1842: \bibitem[Blitz \& Rosolowsky(2006)]{ros06}  Blitz, L., \& Roslowsky,
1843:   E. 2006, \apj, 650, 833
1844: \bibitem[Cabot(2006)]{c06} Cabot, W. 2006, Phys. Fluids, 18, 045101
1845: \bibitem[Castor et al.(1975)Castor, McCray, \& Weaver]{cas75} Castor,
1846: J., McCray, R., \& Weaver, R. 1975, ApJ, 200, L107 
1847: \bibitem[Chy\.zy \& Beck(2004)]{chy04} Chy\.{z}y, K. T., \& Beck,
1848:   R. 2004, \aap, 417, 541
1849: \bibitem[Clarke (1994)]{cla94} Clarke, D. A. 1994, NCSA Technical
1850:   Report.
1851: \bibitem[Cooper et al.(2008)]{c08} Cooper, J. L., Bicknell, G. V.,
1852:   Sutherland, R. S., \& Bland-Hawthorn, J. 2008, \apj, 674, 157
1853: \bibitem[D'Ercole \& Brighenti (1999)]{der99} D'Ercole, A. \&
1854:   Brighenti, F. 1999, MNRAS, 309, 941
1855: \bibitem[De Young \& Heckman(1994)]{dh94} De Young, D. S., \&
1856: Heckman, T. M. 1994, \apj, 431, 598
1857: \bibitem[Draine \& McKee(1993)]{dra93} Draine, B. T., \& McKee,
1858:   C. F. 1993, \araa, 31, 373
1859: \bibitem[Fragile et al.(2004)]{fra04} Fragile, P. C., Murray, S. D, \&
1860:   Lin, D. N. C. 2004, \apj, 617, 1077
1861: \bibitem[Fujita et al.(2004)]{fu04} Fujita, A., Mac Low, M.-M.,
1862:   Ferrara, A., \& Meiksin, A. 2004,  ApJ, 613, 159
1863: \bibitem[Fujita et al.(2003)]{fu03} Fujita, A., Martin, C. L., Mac
1864: Low, M.-M., \& Abel, T. 2003, ApJ, 599, 50 (615, 1082 Erratum)
1865: \bibitem[Gittings et al.(2008)]{g08} Gittings, M., Weaver, R.,
1866: Clover, M., Betlach, T., Byrne, N., Coker, R., Dendy, E., Hueckstaedt,
1867: R., New, K., Oakes, W. R., Ranta, D., \& Stefan, R. 2008, Comput.\
1868: Sci.\ Disc., 1, 015005
1869: \bibitem[Glover \& Mac Low(2007)]{glo07} Glover, S. C. O. G., \& Mac Low,
1870:   M.-M. 2007, \apj, 659, 1317
1871: \bibitem[Grimes et al.(2005)]{gri05} Grimes, J. P., Heckman, T. M.,
1872:   Strickland, D. K., \& Ptak, A. 2005, 628, 187
1873: \bibitem[Heckman et al.(1990)]{he90} Heckman, T. M., Armus, L., \&
1874:   Miley, G. K. 1990, \apjs, 74, 833
1875: \bibitem[Heckman et al.(2000)]{he00} Heckman, T. M., Lehnert, M. D.,
1876:   Strickland, D. K., \& Armus, L. 2000,  ApJS, 129, 493
1877: \bibitem[Heckman et al.(2001)]{hec01} Heckman, T. M., Sembach, K. R.,
1878:   Meurer, G. R., Leitherer, C., Calzetti, D., \& Martin, C. L. 2001, ApJ, 558, 81
1879: \bibitem[Huo et al.(2004)]{huo04} Huo, A. Y., Xia, X. Y., Xue, S. J., Mao, S.,
1880:   \& Deng, Z. G. 2004, ApJ, 611, 208
1881: \bibitem[Jimenez et al.(2003)]{ji03} Jimenez, R., Verde, L., \& Oh,
1882:   S. P. 2003, MNRAS, 339, 243
1883: \bibitem[Joung \& Mac Low(2006)]{jm06} Joung, M. K. R., \& Mac Low,
1884:   M.-M. 2006, \apj, 653, 1266
1885: \bibitem[Kerbyson et al.(2001)]{ker01} Kerbyson, D. J., Aime, H. J.,
1886: Hoisie, A., Petrini, F., Wasserman, H. J., \& Gittings, M. 2001, in
1887: Proc. of the 2001 ACM/IEEE conference on Supercomputing,
1888: ed. G. Johnson (New York: ACM), 37
1889: \bibitem[Klein et al.(1994)]{k94} Klein, R. I., McKee, C. F., \&
1890: Colella, P. 1994, \apj, 420, 213
1891: \bibitem[Korycansky, Zahnle, \& Mac Low(2002)]{kzm02} Korycansky,
1892:   D. G., Zahnle, K. J., \& Mac Low, M.-M. 2002, Icarus, 157, 1
1893: \bibitem[Leitherer et al.(1999)]{lei99} Leitherer, C., Schaerer, D.,
1894: Goldader, J. D., Delgado, R. M. G., Robert, C., Kune, D. F., de Mello,
1895: D. F., Devost, D., \& Heckman, T. M. 1999, ApJS, 123, 3
1896: \bibitem[]{lmk04} Li, Y., Mac Low, M.-M., \& Klessen, R. S. 2004,
1897:   \apj, 614, L19
1898: \bibitem[Mac Low et al.(1989)Mac Low, McCray, \& Norman]{low89} Mac Low, M.-M., McCray, R., \& Norman, M. L. 1989, ApJ, 337, 141
1899: \bibitem[Mac Low \& McCray (1988)]{low88} Mac Low, M.-M. \& McCray,
1900:   R. 1988 ApJ, 324, 76
1901: \bibitem[Mac Low et al.(1994)]{mmml94} Mac Low, M.-M., McKee, C. F.,
1902:   Klein, R. I., Stone, J. M., \& Norman, M. L. 1994, \apj, 433, 757
1903: \bibitem[Mac Low \& Ferrara (1999)]{low99} Mac Low, M.-M. \& Ferrara, A. 1999, ApJ, 513, 142
1904: \bibitem[Mac Low \& Zahnle (1994)]{low94} Mac Low, M.-M., \& Zahnle,
1905:   K. ApJ, 434, L33 
1906: \bibitem[MacDonald \& Bailey (1981)]{mac81} MacDonald, J. \& Bailey,
1907:   M. E. 1981, MNRAS, 197, 995
1908: \bibitem[Marcolini et al.(2005)]{mar05} Marcolini, A. Strickland,
1909:   D. K., D'Ercole, A., Heckman, T. M., \& Hoopes, C. G. 2005, MNRAS,
1910:   362, 626
1911: \bibitem[Martin (1998)]{cry98} Martin, C. L. 1998, ApJ 506, 222
1912: \bibitem[Martin (1999)]{cry99} Martin, C. L. 1999, ApJ, 513, 156
1913: \bibitem[Martin (2005)]{cry05} Martin, C. L. 2005, ApJ, 621, 227
1914: \bibitem[Martin (2006)]{cry06} Martin, C. L. 2006, ApJ, 647, 222
1915: \bibitem[Martin et al.(2002)]{cry02} Martin, C. L., Kobulnicky, H. A.,
1916:   \& Heckman, T. M. 2002, \apj, 547, 663
1917: \bibitem[Meshkov(1969)]{m69} Meshkov, E. E. 1969, Sov.\ Fl.\ Dyn., 4, 101
1918: \bibitem[]{mh96} Mihos, J. C., \& Hernquist, L. 1996, \apj, 464, 641
1919: \bibitem[Monreal-Ibero et al.(2006)]{mi06} Monreal-Ibero, A., Arribas,
1920:   S., \& Colina, L. 2006, \apj, 637, 138
1921: \bibitem[Murray et al.(2005)]{mqt05} Murray, N., Quataert, E., \&
1922: Thompson, T. A. 2005, \apj, 618, 569
1923: \bibitem[Murphy et al.(2001)]{mu01} Murphy, T. W. J., Soifer, B. T., Matthews, K., \& Armus, L. 2001, ApJ, 559, 201
1924: \bibitem[Navarro et al.(1997)]{na97} Navarro, J. F., Frenk, C. S., \&
1925:   White, S. D. M. 1997, ApJ, 490, 493 (NFW)
1926: \bibitem[Oppenheimer \& Dav\'e(2006)]{od06} Oppenheimer, B. D., \&
1927:   Dav\'e, R. 2006, \mnras, 373, 1265
1928: \bibitem[Oppenheimer \& Dav\'e(2008)]{od08} Oppenheimer, B. D., \&
1929:   Dav\'e, R. 2008, \mnras, 387, 577
1930: \bibitem[Richtmyer(1960)]{r60} Richtmyer, R. D. 1960, Comm. Pure
1931:   Appl. Math, 13, 297
1932: \bibitem[]{rup02} Rupke, D.,
1933:    Veilleux, S., \& Sanders, D. B.
1934: 2002, ApJ, 570, 588 
1935: \bibitem[Rupke et al.(2005)]{rup05} Rupke, D.,
1936:    Veilleux, S., \& Sanders, D. B.
1937: 2005, ApJS, 160, 115
1938: \bibitem[Sanders et al.(1988)]{sa88} Sanders, D. B., Scoville, N. Z., Sargent, A. I., \& 
1939: Soifer, B. T. 1988, ApJ, 324, L55
1940: \bibitem[Schwartz \& Martin (2004)]{sch04} Schwartz, C. M., \& Martin,
1941:   C. L. 2004, ApJ, 610, 201
1942: \bibitem[Scoville et al.(1997)]{s97} Scoville, N. Z., Yun, M. S., \&
1943: Bryant, P. M. 1997, \apj, 484, 702
1944: \bibitem[Shin et al.(2008)]{s08} Shin, M.-S., Stone, J. M., \& Snyder,
1945:   G. F. 2008, \apj, 680, 336
1946: \bibitem[Solomon et al.(1992)]{sl92} Solomon, P. M., Downes, D., Radford, S. J. E. 1992, \apj 387, L55
1947: \bibitem[Solomon et al.(1997)]{sl97} Solomon, P. M., Downes, D., Radford, S. J. E., \& 
1948: Barrett, J. W. 1997, ApJ, 478, 144
1949: \bibitem[Scoville et al.(1997)]{sc97} Scoville, N. Z., Yun, M. S., \& Bryant, P. M. 1997, 
1950: ApJ, 484, 702
1951: \bibitem[Stil et al.(2008)]{swo08} Stil, J. M., Wityk, N. D., Ouyed, R., \&
1952:   Taylor, A. R. 2008, \apj, submitted (arXiv:0807.0057)
1953: \bibitem[Stone \& Norman (1992a)]{sto92a} Stone, J. M. \& Norman,
1954:   M. L. 1992a, ApJS, 80, 753
1955: \bibitem[Stone \& Norman (1992b)]{sto92b} Stone, J. M., \& Norman,
1956:   M. L. 1992b, \apj, 390, L17
1957: \bibitem[Strickland \& Heckman(2007)]{st07} Strickland, D. K., \&
1958:   Heckman, T. M. 2007, ApJ, 658, 258
1959: \bibitem[Strickland et al.(2004)]{st04} Strickland, D. K., Heckman,
1960: T. M., Colbert, E. J. M., Hoopes, C. G., Charles, G., \& Weaver,
1961: K. A. 2004, ApJS, 151, 193
1962: \bibitem[Strickland \& Stevens (2000)]{st00} Strickland, D. K. \&
1963:   Stevens, I. R. 2000, MNRAS, 314, 511
1964: \bibitem[Suchkov et al. (1996)]{su96} Suchkov, A. A., Berman, V. G., Heckman,
1965:   T. M., \& Balsara, D. S. 1996, ApJ, 463, 528
1966: \bibitem[Sutherland \& Dopita(1993)]{sd93} Sutherland, R. S., \&
1967: Dopita, M. A. 1993, \apjs, 88, 253
1968: \bibitem[Tenorio-Tagle \& Bodenheimer(1988)]{ten88} Tenorio-Tagle, G.,
1969:   \& Bodenheimer, P. 1988, ARA\&A, 26, 145
1970: \bibitem[Tomisaka(1998)]{t98} Tomisaka, K. 1998, \mnras, 298, 797
1971: \bibitem[Tomisaka \& Ikeuchi(1986)]{tom86} Tomisaka, K., \& Ikeuchi,
1972:   S. 1986, PASJ, 38, 697
1973: \bibitem[Tomisaka \& Ikeuchi(1988)]{tom88} Tomisaka, K., \& Ikeuchi,
1974:   S. 1988, ApJ, 330, 695
1975: \bibitem[Toomre(1963)]{t63} Toomre, A. 1963, \apj, 138, 385
1976: \bibitem[Vacca(1996)]{v96} Vacca, W. 1996, in The Interplay Between
1977: Massive Star Formation, the ISM and Galaxy Evolution, eds. D. Kunth,
1978: B. Guiderdoni, M. Heydari-Malayeri, \& T. X. Thuan (Gif-sur-Yvette:
1979: Editions Frontieres), 321
1980: \bibitem[van Leer (1977)]{van77} van Leer, B. 1977,  J. Comput. Phys.,
1981:   23, 276
1982: \bibitem[Veilleux \& Rupke(2002)]{vei02} Veilleux, S., \& Rupke,
1983:   D. S. 2002, \apj, 565, L63
1984: \bibitem[Vieser \& Hensler(2007)]{vie07} Vieser, W., \& Hensler,
1985:   G. 2007, \aap, 472, 141
1986: \bibitem[Weaver et al.(1977)]{wea77} Weaver, R., McCray, R., Castor, J., Shapiro,
1987: P., \& Moore, R. 1977, ApJ, 218, 377
1988: \bibitem[Xu \& Stone(1994)]{xs94} Xu, J. \& Stone, J. M. 1994, \apj,
1989:   454, 172
1990: \bibitem[Yabe \& Xiao(1993)]{y93} Yabe, T., \& Xiao, F. 1993,
1991:   J. Phys.Soc. Japan, 62, 2537
1992: \bibitem[Young et al.(2001)]{y01} Young, Y.-N., Tufo, H., Dubey, A., \&
1993:   Rosner, R. J. Fluid Mech. 447, 377
1994: \bibitem[Youngs(1984)]{y84} Youngs, D. L. 1984, Physica D, 12, 32
1995: \bibitem[Zahnle \& Mac Low (1995)]{z95} Zahnle, K. \& Mac Low M.-M. 1995, JGR/E, 100, 16885
1996: 
1997: \clearpage
1998: 
1999: 
2000: \end{thebibliography}{}
2001: \clearpage
2002: 
2003: \begin{figure}
2004: %\includegraphics[width=\textwidth]{scaleH.eps}
2005: \includegraphics[width=\textwidth]{f1.eps}
2006: \figcaption{The density distributions of the disks with central surface 
2007: densities of $\Sigma_0=10^4$ and $5\times10^4~\mbox{ M}_{\odot}\mbox{ pc}^{-1}$
2008:  in the vertical direction. The scale heights are $H=7$ and 2 pc respectively. 
2009: \label{H}
2010: }
2011: \end{figure} \clearpage
2012: 
2013: \begin{figure}
2014: %\includegraphics[width=0.5\textwidth]{mech.eps}
2015: %\includegraphics[width=0.5\textwidth]{Qt.eps}
2016: \includegraphics[width=0.5\textwidth]{f2a.eps}
2017: \includegraphics[width=0.5\textwidth]{f2b.eps}
2018: \figcaption{Mechanical luminosities ({\em left panel}) and ionizing photon luminosity 
2019: ({\em right panel}) as a function of time 
2020: for three starburst scenarios: instantaneous starburst with
2021: $M_*=10^9~\mbox{ M}_{\odot}$ ({\em solid line}); continuous star formation
2022: with $500~\mbox{ M}_{\odot}\mbox{ yr}^{-1}$ ({\em dashed line}); and
2023: continuous star formation with $100~\mbox{ M}_{\odot}\mbox{ yr}^{-1}$ ({\em
2024: dash-dotted line}). 
2025: Population synthesis models are from Starburst~99 (Leitherer et al.\ 1999).
2026: \label{lmech}
2027: }
2028: \end{figure} \clearpage
2029: 
2030: \begin{figure}
2031: %\includegraphics[width=0.5\textwidth]{twind.eps}
2032: %\includegraphics[width=0.5\textwidth]{vwind.eps}
2033: \includegraphics[width=0.5\textwidth]{f3a.eps}
2034: \includegraphics[width=0.5\textwidth]{f3b.eps}
2035: \figcaption{The ({\em a}) temperature of a hot bubble in
2036:   a uniform medium driven by a starburst 
2037:   with 
2038: $L_{mech}$ and the rate of supernova ejecta $M_{SN}\propto L_{mech}$ predicted by the Starburst 
2039: 99 model as a function
2040:   of time, with the given mass-loading factors $\xi$, under the
2041:   assumption of an instantaneous ({\em solid line}) or continuous
2042:   ({\em dashed line}) starburst.  The amount of mass in the hot wind
2043:   is $\xi M_{SN}$. ({\em b}) The expected terminal velocity of the
2044:   wind driven by such a bubble after its blowout from a stratified
2045:   disk. Note $T_{wind}$ and $v_{wind}$ are proportional to $L_{mech}/M_{SN}$, thus 
2046:   independent of SFR assumed. 
2047: \label{vtwind}
2048: }
2049: \end{figure} \clearpage
2050: 
2051: \begin{figure}
2052: %\includegraphics[width=\textwidth]{L43.d.eps}
2053: \includegraphics[width=\textwidth]{f4.eps}
2054: \figcaption{The density distributions of our standard model at
2055: $t=0.27$ Myr with the cooling temperature cut-offs of $T_{floor}=10^2$
2056: (U1: {\em left panel}) and $10^4$ K (X1: {\em right panel})
2057: respectively. Note that because the shell remains underresolved, the
2058: temperature floor does not substantially influence the behavior of the
2059: shell in our models.  The dark line denotes our fiducial line of
2060: sight.
2061: \label{L43}
2062: }
2063: \end{figure} \clearpage
2064: 
2065: 
2066: \begin{figure}
2067: %\includegraphics[width=\textwidth]{R4.d.eps}
2068: \includegraphics[width=\textwidth]{f5.eps}
2069: \figcaption{
2070: ({\em top}) The density distributions of two models with
2071: $L_{mech}=10^{43}\mbox{ erg s}^{-1}$: ({\em top left}) our fiducial model with a
2072: source region of only 25 zones, model S1, and ({\em top right}) model V1 with a
2073: higher surface density $\Sigma_0=5\times10^4~\mbox{ M}_{\odot}\mbox{
2074:   pc}^{-2}$. ({\em bottom})The density distributions of two models with lower mechanical
2075: luminosities:  ({\em bottom left}) model X2 with 
2076: $L_{mech}=10^{42}$ and ({\em bottom right}) model X3 with
2077: $L_{mech}=10^{41}\mbox{ erg s}^{-1}$.
2078: Each model is shown just before it exits the grid, at times of $t=0.28$, 0.22, 0.49, and 
2079: 0.85 Myr respectively.  
2080: \label{L421}
2081: }
2082: \end{figure} \clearpage
2083: 
2084: \begin{figure}
2085: %\includegraphics[width=\textwidth]{W4.d.eps}
2086: \includegraphics[width=\textwidth]{f6.eps}
2087: \figcaption{The density distributions of our model with $T_{floor}=10^2$ K 
2088: with different mass-loading rates: $\dot{M_{in}}=1.7 
2089: \mbox{ M}_{\odot}\mbox{ yr}^{-1}$ (U1-A: {\em top left}),  
2090: 17 $\mbox{ M}_{\odot}\mbox{ yr}^{-1}$ (U1: {\em top right}), 49 $\mbox{
2091:   M}_{\odot}\mbox{ yr}^{-1}$(U1-B: {\em bottom left}), 
2092: and 120 $\mbox{ M}_{\odot}\mbox{ yr}^{-1}$ (U1-C: {\em bottom right}). They are shown at $t=0.22$, 0.27, 0.35, 
2093: and 0.41 Myr respectively. 
2094: Different mass-loading rates correspond to different wind temperatures and terminal wind velocities. 
2095:  Note that the growth of hydrodynamic instabilities is suppressed 
2096: as the mass-loading rate decreases and the terminal wind velocity increases. 
2097: \label{w}
2098: }
2099: \end{figure} \clearpage
2100: 
2101: \begin{figure}
2102: %\includegraphics[width=\textwidth]{X4.d.eps}
2103: \includegraphics[width=\textwidth]{f7.eps}
2104: \figcaption{The density distributions of our standard model with $T_{floor}
2105: =10^4$ K at $t=0.27$ Myr, with 
2106: resolution of 0.1 pc (X1-0: {\em top left}), 
2107: our fiducial
2108: resolution of 0.2 pc 
2109: (X1: {\em top right}), and 
2110: resolutions of 0.4 pc (X1-2: {\em bottom left}) 
2111: and 0.8 pc (X1-4: {\em bottom right}). Note that the growth of 
2112: R-T instability is suppressed as the resolution decreases. The black
2113: lines show the typical line of sight used for line profile analysis.
2114: \label{res}
2115: }
2116: \end{figure} \clearpage
2117: 
2118: \begin{figure}
2119: %\includegraphics[width=\textwidth]{dres.eps} 
2120: \includegraphics[width=\textwidth]{f8.eps} 
2121: \figcaption{ Density
2122: profiles at the outer shock fronts in the vertical direction at
2123: $t=0.06$ Myr for 0.1~pc resolution ({\em triangles}; model X1-0), and at
2124: $t=0.05$ Myr for 0.2~pc  ({\em stars}; X1), 0.4~pc ({\em
2125: diamonds}; X1-2), and 0.8~pc ({\em crosses}; X1-4) resolution. The shell is better
2126: resolved with a higher resolution, but still not fully resolved even
2127: at 0.1 pc resolution.  (Note that the size of the highest resolution
2128: bubble is slightly smaller only because of the smaller
2129: source region.)
2130: \label{dres}
2131: }
2132: \end{figure} \clearpage
2133: 
2134: \begin{figure}
2135: %\includegraphics[width=\textwidth]{baf.eps} 
2136: \includegraphics[width=\textwidth]{f9.eps} 
2137: \figcaption{The velocity
2138: predicted by the ballistic approximation for shell fragments as they
2139: expand radially in the halo from $r_b=0.2$ kpc with initial (blowout)
2140: velocities, $v_b=50$, 100, 200, 400, 600, 800, and 1000 km s$^{-1}$.
2141: The {\em left} panel shows the behavior near the galaxy, while the
2142: {\em right} panel captures the full extent of the halo.  Radii at
2143: $t=10$ Myr ({\em left panel}) and the virial radius ({\em right
2144: panel}) are noted with {\em thick lines}.
2145: \label{ballistic}
2146: }
2147: \end{figure} \clearpage
2148: 
2149: \begin{figure}
2150: %\includegraphics[width=\textwidth]{xline.eps} 
2151: \includegraphics[width=\textwidth]{f10.eps} 
2152: \figcaption{  In the {\em bottom left} panel, the
2153: distribution of column density for the cold gas as a function of
2154: radial velocity is plotted. The other three panels plot radial
2155: profiles of radial
2156: velocity ({\em top left}),  density ({\em top right}), and 
2157: temperature ({\em bottom right}) along a line of
2158: sight through the center at an angle of 19$^{\circ}$ from the vertical
2159: axis in model X1 at the end of the simulation.  The radial profiles all use the
2160: horizontal axis labeled on the bottom right. Regions of cold gas
2161: with $T<5\times10^4$ K (which we take to be Na~{\sc i} absorbing gas)
2162: are shown in {\em diamonds}. 
2163: \label{line}
2164: }
2165: \end{figure} \clearpage
2166: 
2167: 
2168: \begin{figure}
2169: %\includegraphics[width=\textwidth]{bline.eps}
2170: \includegraphics[width=\textwidth]{f11.eps}
2171: \figcaption{The same as in Figure~\ref{line} for the same standard
2172:   simulation, but with our highest resolution of 0.1 pc (X1-0).
2173:   Again, the three radial profiles ({\em top right, top left, bottom
2174:     right}) all use the radial axis given on the bottom right.
2175: \label{bline}
2176: }
2177: \end{figure} \clearpage
2178: 
2179: \begin{figure}
2180: %\includegraphics[width=\textwidth]{column.eps}
2181: \includegraphics[width=\textwidth]{f12.eps}
2182: \figcaption{The column density distributions of cool gas at sightlines
2183: through the center as a function of angle extended from the vertical
2184: axis in models with 0.2~pc ({\em solid line}; model X1) and 0.1~pc ({\em
2185: dashed line}; X1-0) resolution. The observed range of column density
2186: inferred from observations of Na~{\sc i} absorption profiles is
2187: $1.0\times10^{19}-4.3\times10^{21}\mbox{ cm}^{-2}$, shown in {\em
2188: dotted lines}. Note that the column densities from the models will be reduced
2189: over time due to spherical expansion.
2190: \label{hcolumn}
2191: }
2192: \end{figure} \clearpage
2193: 
2194: \begin{figure}
2195: %\includegraphics[width=\textwidth]{lwX.eps}
2196: \includegraphics[width=\textwidth]{f13.eps}
2197: \figcaption{The velocity widths are shown as a function of angle from 
2198: the vertical axis at every degree for models with grid resolutions of
2199: $dx = 0.1$, 0.2, 0.4, and 0.8~pc (models X1-0, X1, X1-2, and X1-4). Note that the
2200: highest two resolutions appear to display converged behavior.
2201: The upper and lower limits of the observed average line width in Na~{\sc i} 
2202: absorption are shown in ({\em dashed lines}): $\langle v \rangle=320\pm120\mbox{ km s}^{-1}$.
2203: \label{lw}
2204: }
2205: \end{figure} \clearpage
2206: 
2207: \begin{figure}
2208: %\includegraphics[width=\textwidth]{termX.eps}
2209: \includegraphics[width=\textwidth]{f14.eps} \figcaption{The terminal
2210: velocity ({\em solid line with asterisks}) and the average
2211: mass-weighted velocity ({\em thin solid line}) of cool gas are plotted
2212: as a function of angle from the vertical axis along sightlines through
2213: the galactic center in models with grid resolutions of $dx = 0.1$,
2214: 0.2, 0.4, and 0.8~pc (models X1-0, X1, X1-2, and X1-4).  Since shell
2215: mass varies substantially with angle, we also plot the mass-weighted
2216: average velocity within a $10^{\circ}$ arc $v_{av,10}$ at each angle
2217: ({\em thick solid line}).  We show the terminal velocity of the low
2218: density wind $v_{wind}\approx1000\mbox{ km s}^{-1}$ in {\em dotted
2219: line}, the observed average shell velocity $v_{s,obs}=330\pm100\mbox{
2220: km s}^{-1}$ in {\em dashed line}, and the observed average terminal
2221: velocity $v_{t,obs}=750\mbox{ km s}^{-1}$ in {\em dash-dot-dot line}.
2222: \label{term}
2223: }
2224: \end{figure} \clearpage
2225: 
2226: 
2227: \begin{figure}
2228: %\includegraphics[width=\textwidth]{termR4.eps}
2229: \includegraphics[width=\textwidth]{f15.eps}
2230: \figcaption{Same as in Figure~\ref{term} for
2231:   lower luminosity runs with $L_{mech} = 10^{41}$~erg~s$^{-1}$ 
2232:   ({\em right}; model X3), and $10^{42}$~erg~s$^{-1}$ ({\em left}; X2). 
2233: \label{lwt}
2234: }
2235: \end{figure} \clearpage
2236: 
2237: \begin{figure}
2238: %\includegraphics[width=0.5\textwidth]{histX.eps}
2239: %\includegraphics[width=0.5\textwidth]{histW.eps}
2240: \includegraphics[width=0.5\textwidth]{f16a.eps}
2241: \includegraphics[width=0.5\textwidth]{f16b.eps} \figcaption{The mass
2242: distributionsof all the cool shells and shell fragments as a function
2243: of velocity for {\em (a)} models at $t=0.27$~Myr with increasing zone
2244: sizes $dx = 0.1$~pc ({\em dashed line}; model X1-0), 0.2~pc ({\em
2245: solid line}; X1), and 0.4 pc ({\em dashed-dot-dot line}; X1-2), and
2246: {\em (b)} models at the blowout time ($t=0.22$, 0.27, 0.33, and 0.41
2247: Myr) with increasing mass loading of the wind
2248: 1.7~M$_{\odot}$~yr$^{-1}$ ({\em dashed line}; model U1-A),
2249: 17~M$_{\odot}$~yr$^{-1}$ ({\em solid lines}; U1),
2250: 49~M$_{\odot}$~yr$^{-1}$ ({\em dashed-dot-dot line}; U1-B), and
2251: 120~M$_{\odot}$~yr$^{-1}$ ({\em dotted line}; U1-C.).
2252: \label{hist}
2253: }
2254: \end{figure} \clearpage
2255: 
2256: \begin{figure}
2257: %\includegraphics[width=\textwidth]{termt1.eps}
2258: \includegraphics[width=\textwidth]{f17.eps}
2259: \figcaption{Same as in Figure~\ref{term} for 
2260: models with increasing mass-loading of the wind 
2261: 1.7~M$_{\odot}$~yr$^{-1}$ ({\em top left}; model U1-A) at $t=0.22$
2262: Myr, 
2263:  17~M$_{\odot}$~yr$^{-1}$ ({\em top right}; U1) at $t=0.27$ Myr,  
2264:  49~M$_{\odot}$~yr$^{-1}$ ({\em bottom left}; U1-B) at $t=0.35$ Myr, 
2265: and 
2266: 120~M$_{\odot}$~yr$^{-1}$ ({\em bottom right}; U1-C) at $t=0.41$ Myr.  
2267: \label{termt1}
2268: }
2269: \end{figure} \clearpage
2270: 
2271: \begin{figure}
2272: %\includegraphics[width=\textwidth]{NaI.eps}
2273: \includegraphics[width=\textwidth]{f18.eps} 
2274: \figcaption{Simulated Na~{\sc i} 5890 absorption profiles at
2275: sightlines $\theta=5^{\circ}$, $10^{\circ}$, $19^{\circ}$,
2276: $20^{\circ}$, $30^{\circ}$, $40^{\circ}$, $50^{\circ}$, $60^{\circ}$,
2277: $70^{\circ}$, and $80^{\circ}$ ({\em clockwise from the bottom left})
2278: for models with $dx = 0.1$~pc ({\em solid line}; model X1-0) and 0.2
2279: pc ({\em dashed line}; X1).  We also show the observed range of
2280: average shell velocity in {\em shaded area} and the observed average
2281: terminal velocity in {\em dash-dot-dot} line.
2282: \label{NaI}
2283: }
2284: \end{figure} \clearpage
2285: 
2286: \begin{figure} 
2287: %\includegraphics[width=\textwidth]{NaI-dilute-0.01.eps}
2288: \includegraphics[width=\textwidth]{f19.eps} 
2289: \figcaption{Simulated Na~{\sc i} 5890/5896 doublet absorption profiles
2290:   averaged over $ª£20^{\circ}$ centered on the given (non-uniformly
2291:   distributed) angles for model X1-0 with $dx = 0.1$~pc ({\em solid
2292:     line}). The same profiles are shown after geometric dilution of
2293:   the column densities by a factor of 100 ({\em dashed line}), and
2294:   application of a Gaussian instrumental broadening with
2295:   FWHM$=65$~km~s$^{-1}$, for comparison with five observed ULIRG
2296:   spectra ({\em thin solid lines}) from Martin (2005).  Note that the
2297:   velocity frame is centered on the 5890 line; blue-shifted absorption
2298:   from the 5896 line lies at low velocities in this frame.  We also
2299:   show the observed range of average shell velocity in {\em shaded
2300:     area} and the observed average terminal velocity in {\em
2301:     dash-dot-dot} line.
2302: \label{NaI20} 
2303: } 
2304: \end{figure} \clearpage
2305: 
2306: 
2307: \begin{figure}
2308: %\includegraphics[width=\textwidth]{lwt-30.eps}
2309: \includegraphics[width=\textwidth]{f20.eps}
2310: \figcaption{The velocity width is plotted as in Figure~\ref{lw} ({\em
2311:     top panels}), and the mass-weighted average velocity and the
2312:     terminal velocity are plotted as in Figure~\ref{term} ({\em bottom
2313:     panels}) for model X1.  Parallel sightlines are chosen along a
2314:     slit oriented at $\theta=30^{\circ}$ from the axisymmetric axis
2315:     and the major axis ({\em left panels}) and the minor axis ({\em
2316:     right panels}).
2317: \label{30}
2318: }
2319: \end{figure} \clearpage
2320: 
2321: \begin{figure} 
2322: %\includegraphics[width=\textwidth]{linei.eps} 
2323: \includegraphics[width=\textwidth]{f21.eps} 
2324: \figcaption{The column density distributions as a function of radial
2325: velocity for Na~{\sc i} absorbing gas ({\em diamonds}) and H$\alpha$
2326: emitting gas ({\em triangles}) with photon luminosities of
2327: $Q=10^{53}$, $10^{54}$, $1.9\times10^{54}$, and $10^{55}$ photons
2328: s$^{-1}$ in a line of sight through the center at an angle of
2329: 13$^{\circ}$ from the vertical axis in model U1.
2330: The velocity widths of the two components vary as Q is changed.
2331: \label{ion} } \end{figure} \clearpage
2332: 
2333: \begin{figure}
2334: %\includegraphics[width=\textwidth]{t8.d.eps}
2335: \includegraphics[width=\textwidth]{f22.eps} \figcaption{ The density
2336: distributions at blowout in our model dwarf galaxy with $M_d=10^8
2337: \mbox{ M}_{\odot}$ and $L_{mech}=10^{40}\mbox{ erg s}^{-1}$ at $z=8$.
2338: The bubbles are shown at the time of blowout.  This is $t=1$ Myr for
2339: models without radiative cooling, $t=1.3$ Myr for SAGE models with
2340: radiative cooling, and $t=1.1$ Myr for ZEUS-3D models.  The top row
2341: shows models without radiative cooling using SAGE with uniform grid
2342: and 5 levels of AMR, and using ZEUS-3D.  These are models AN, UN, and
2343: RN.  The middle row shows the same grids with radiative cooling,
2344: models AC, UC, and RC.  Finally, the bottom row shows SAGE models with
2345: cooling and 6 and 7 levels of AMR, models BC, and CC.
2346: \label{compare}
2347: }
2348: \end{figure} 
2349: 
2350: \clearpage
2351: 
2352: \begin{table}
2353: \begin{center}
2354: \caption{Parameters for  
2355: Starburst Models
2356: \label{run} 
2357: }
2358: \begin{tabular}{lcccccc}
2359: \\
2360:    Model\tablenotemark{a}        & $\log_{10} \Sigma_0$\tablenotemark{b} & $\log_{10} L_{mech}$\tablenotemark{c} & 
2361: $\dot{M_{in}}$\tablenotemark{d} & $R_{SN}$\tablenotemark{e} & $T_{floor}$\tablenotemark{f} & resolution \\
2362: 
2363:                 & ($\mbox{ M}_{\odot}\mbox{ pc}^{-2}$) & (erg s$^{-1}$) & 
2364: ($\mbox{ M}_{\odot} yr^{-1}$) & (pc) & (K) & (pc) \\
2365: \hline
2366: X1    & 4 & 43 & 17  & 50 &  $10^4$ & 0.2     \\
2367: X1-0  & 4 & 43 & 17  & 25\tablenotemark{g} &  $10^4$ & 0.1     \\ 
2368: X1-2  & 4 & 43 & 17  & 50 &  $10^4$ & 0.4\tablenotemark{h}     \\ 
2369: X1-4  & 4 & 43 & 17  & 50 &  $10^4$ & 0.8\tablenotemark{h}    \\ 
2370: U1    & 4 & 43 & 17  & 50 & $10^2$ & 0.2     \\
2371: U1-A   & 4 & 43 & 1.7 & 50 &  $10^2$ & 0.2     \\
2372: U1-B   & 4 & 43 & 49  & 50 &  $10^2$ & 0.2     \\
2373: U1-C   & 4 & 43 & 120 & 50 &  $10^2$ & 0.2     \\
2374: X2    & 4 & 42 & 17  & 50 &  $10^2$ & 0.2     \\
2375: X3    & 4 & 41 & 17  & 50 &  $10^2$ & 0.2     \\
2376: S1    & 4 & 43 & 17  & 25 &  $10^2$ & 0.2    \\
2377: V1
2378: & 4.7 & 43 & 17 & 50 &  $10^2$ & 0.2     \\
2379: \end{tabular}
2380: \tablenotetext{a}{All models run with ZEUS}
2381: \tablenotetext{b}{Central surface density}
2382: \tablenotetext{c}{Mechanical luminosity}
2383: \tablenotetext{d}{Mass loading rate}
2384: \tablenotetext{e}{Size of source region where supernova energy is injected}
2385: \tablenotetext{f}{The minimum temperature floor for cooling}
2386: \tablenotetext{g}{The level of noise on the surface of the source region is kept same with that of our standard 
2387: model, X1 by setting the number of cells covering the source region the same. }
2388: \tablenotetext{h}{Ratioed grids are used (the resolution is 0.2pc within the source regions)}
2389: \end{center}
2390: \end{table}
2391: 
2392: 
2393: \begin{table}
2394: \begin{center}
2395: \caption{Parameters for Dwarf Galaxy Models 
2396: \label{sz}
2397: }
2398: \begin{tabular}{llcccccccc}
2399: \\
2400: model & code  & initial grid & res\tablenotemark{1} & cool & AMR\tablenotemark{2} & cycle & cells\tablenotemark{3} & cpu\tablenotemark{4} & procs\tablenotemark{5}  \\
2401: & & & (pc) &  & &  & active (total) & per cycle &  \\
2402: \hline
2403: UN & SAGE &  $800\times1120$                   & 0.277  & OFF & OFF & 7244  & 896       & 4.0   & 12 \\
2404: UC & SAGE &  $800\times1120$                   & 0.277  & ON  & OFF & 10877 & 896       & 4.9   & 12 \\
2405: AN & SAGE &  $50\times70$                      & 0.277  & OFF & 5   & 7469  & 111 (146) & 0.86  & 12 \\
2406: AC & SAGE &  $50\times70$                      & 0.277  & ON  & 5   & 10790 & 103 (136) & 0.84  & 12 \\
2407: BC & SAGE &  $50\times70$                      & 0.139  & ON  & 6   & 22568 & 240 (319) & 1.5   & 24 \\
2408: CC & SAGE &  $50\times70$                      & 0.0693 & ON  & 7   &
2409: 54000 &        503 (698)& 2.3   & 48 \\
2410: RN & ZEUS &  $400\times878$\tablenotemark{6}   & 0.277  & OFF & OFF & 6035  & 351.2     & ?0.51 & 8 \\
2411: RC & ZEUS &  $400\times878$\tablenotemark{6}   & 0.277  & ON  & OFF & 9036  & 351.2     & 3.7   & 8  \\ 
2412: \end{tabular}
2413: \tablecomments{We give the number of cycles, the average numbers of active (with AMR) and total cells, and 
2414: the average cpu time spent per cycle that are used
2415: to run the simulations until the bubbles blow out of the disk. 
2416: }
2417: \tablenotetext{1}{The highest resolution employed in the simulations.}
2418: \tablenotetext{2}{The level of refinement used in AMR: if OFF, uniform or ratioed grids are used.}
2419: \tablenotetext{3}{Average numbers of active and total cells used per
2420:   cycle, in thousands.}
2421: \tablenotetext{4}{The amount of cpu time spent per cycle.}
2422: \tablenotetext{5}{Number of processors used for computation.}
2423: \tablenotetext{6}{Ratioed grids are used.}
2424: \end{center}
2425: \end{table}
2426: 
2427: 
2428: \end{document}
2429: