1: %\documentstyle[12pt,aasms4]{article}
2: %\documentstyle[12pt,amssym,aaspp4]{article}
3: \documentstyle[aas2pp4]{article}
4:
5: % The eqsecnum style changes the way equations are numbered. Normally,
6: % equations are just numbered sequentially through the entire paper.
7: % If eqsecnum appears in the \documentstyle command, equation numbers will
8: % be sequential through each section, and will be formatted "(sec-eqn)",
9: % where sec is the current section number and eqn is the number of the
10: % equation within that section. The eqsecnum option can be used with
11: % any substyle.
12:
13: %\documentstyle[11pt,eqsecnum,aaspp4]{article}
14:
15: % Authors are permitted to use the fonts provided by the American Mathematical
16: % Society, if they are available to them on their local system. These fonts
17: % are not part of the AASTeX macro package or the regular TeX distribution.
18:
19: %\documentstyle[12pt,amssym,aasms4]{article}
20:
21: % Here's some slug-line data. The receipt and acceptance dates will be
22: % filled in by the editorial staff with the appropriate dates. Rules will
23: % appear on the title page of the manuscript until these are uncommented
24: % out by the editorial staff.
25:
26: %\received{4 August 1988}
27: %\accepted{23 September 1988}
28: %\journalid{337}{15 January 1989}
29: %\articleid{11}{14}
30:
31: \slugcomment{}
32:
33: % Authors may supply running head information, if they wish to do so, although
34: % this may be modified by the editorial offices. The left head contains a
35: % list of authors, usually three allowed---otherwise use et al. The right
36: % head is a modified title of up to roughly 44 characters. Running heads
37: % are not printed.
38:
39: \lefthead{Wada, Spaans and Kim}
40: \righthead{Multi-phase gas dynamics}
41:
42: % This is the end of the "preamble". Now we wish to start with the
43: % real material for the paper, which we indicate with \begin{document}.
44: % Following the \begin{document} command is the front matter for the
45: % paper, viz., the title, author and address data, the abstract, and
46: % any keywords or subject headings that are relevant.
47: \newcommand{\beq}{\begin{equation} }
48: \newcommand{\eeq}{\end{equation}}
49: \newcommand{\bea}{\begin{eqnarray} }
50: \newcommand{\eea}{\end{eqnarray}}
51: \def\mbf#1{\mbox{\boldmath ${#1}$}}
52: \newcommand \hi {H\,{\sc i}}
53:
54: \begin{document}
55: \title{Formation of Cavities, Filaments, and Clumps by
56: the Non-linear Development of Thermal and Gravitational
57: Instabilities
58: in the Interstellar Medium under Stellar Feedback}
59: \author{Keiichi Wada\altaffilmark{1}, Marco Spaans\altaffilmark{2}, and
60: Sungeun Kim\altaffilmark{3}}
61: \affil{$^1$National Astronomical Observatory, Mitaka, 181-8588, Japan\\
62: Email: wada.keiichi@nao.ac.jp}
63: \affil{$^2$Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138}
64: \affil{$^3$Astronomy Department, University of Illinois at
65: Urbana-Champaign, Urbana, IL 61801}
66:
67: % Notice that each of these authors has alternate affiliations, which
68: % are identified by the \altaffilmark after each name. The actual alternate
69: % affiliation information is typeset in footnotes at the bottom of the
70: % first page, and the text itself is specified in \altaffiltext commands.
71: % There is a separate \altaffiltext for each alternate affiliation
72: % indicated above.
73:
74: %\altaffiltext{1}{National Astronomical Observatory, Mitaka, 181, Japan (email: wada@th.nao.ac.jp)}
75: %\altaffiltext{2}{Space Telescope Science Institute, Baltimore, MD 21218 (email: norman@stsci.edu)}
76: %\altaffiltext{3}{Space Telescope Science Institute, Baltimore, MD 21218 (email: norman@stsci.edu)}
77:
78: % The abstract environment prints out the receipt and acceptance dates
79: % if they are relevant for the journal style. For the aasms style, they
80: % will print out as horizontal rules for the editorial staff to type
81: % on, so long as the author does not include \received and \accepted
82: % commands. This should not be done, since \received and \accepted dates
83: % are not known to the author.
84:
85: \begin{abstract}
86: %\input{abst.tex}
87: Based on our high resolution, two-dimensional hydrodynamical
88: simulations, we propose that large cavities may be formed by the nonlinear development of
89: the combined thermal and gravitational instabilities, without need for
90: stellar energy injection in a galaxy modeling the Large Magellanic Clouds (LMC).
91: Our numerical model of the star formation allows us to
92: follow the evolution of the blastwaves due to supernovae in
93: the inhomogeous, multi-phase, and turbulent-like media self-consistently.
94: Formation of kpc-scale inhomogeneity, such as cavities, observed \ion{H}{1} map of the LMC, is suppressed by
95: frequent supernovae (average supernova rate for the whole disk
96: is $\sim 0.001$ yr$^{-1}$). However the supernova explosions are necessary for
97: the hot component ($T_g > 10^{6-7}$ K).
98: %Alternatively the kpc-scale cavities
99: % can be formed due to nonlinear evolution of
100: %the gravitational and thermal instability in the rotating disk.
101: Position-velocity maps show that kpc-scale shells/arcs formed
102: through the nonlinear evolution in a model without stellar
103: energy feedback has similar kinematics to explosional phenomena, such as
104: supernovae.
105: We also find that dense clumps and filamentary structure are
106: formed due to a natural consequence of the non-linear evolution of the
107: multi-phage ISM. Although the ISM in a small scale looks turbulent-like and
108: transient, the global structure of the ISM is quasi-stable. In the quasi-stable
109: phase, the volume filling factor of the hot, warm, cold components are
110: $\sim 0.2, \sim 0.6$, and $\sim 0.2$, respectively.
111: We compare the observations of \ion{H}{1} and molecular gas of the
112: LMC with the numerically obtained \ion{H}{1} and CO brightness temperature distribution.
113: The morphology and statistical properties
114: of the numerical \ion{H}{1} and CO maps are discussed.
115: We find that the cloud mass spectrum of our models
116: represent a power-law shape, but their slopes change between models with and
117: without the stellar energy injection, and also the slope depends on the
118: threshold brightness temperature of CO.
119:
120: \end{abstract}
121:
122: % The different journals have different requirements for keywords. The
123: % keywords.apj file, found on aas.org in the pubs/aastex-misc directory,
124: % contains a list of keywords used with the ApJ and Letters. These are
125: % usually assigned by the editor, but authors may include them in their
126: % manuscripts if they wish.
127:
128: \keywords{ISM: structure, kinematics and dynamics --- galaxies:
129: structure --- individual: LMC --- method: numerical}
130:
131: \section{INTRODUCTION}
132:
133: The topology of the neutral
134: interstellar medium (ISM) can be studied in great detail
135: by the spatial and velocity structures in the neutral \ion{H}{1} gas.
136: A recent
137: high-resolution \ion{H}{1} survey of the Large Magellanic Cloud (LMC) reveals
138: that the structure of the neutral atomic inter stellar gas is dominated by
139: numerous holes and shells as well as complex filamentary structure (\cite{KIM98}). These features are commonly seen in recent high-resolution \ion{H}{1} images of
140: nearby galaxies obtained with radio synthesis interferometers
141: (\cite{DD90}; \cite{P92}; \cite{SS97}; \cite{WB99}; \cite{ST99}).
142: In general, the shell-like and hole structure seen in
143: \ion{H}{1} has been understood as the cumulative effect of
144: stellar winds from massive stars and supernova explosions evacuating the cool
145: ISM (\cite{TT88}; \cite{VDH96}; \cite{OEY96} ; \cite{OC97}). However, the extensive study of \ion{H}{1} shells in the LMC shows that
146: there is a relatively weak correlation between the \ion{H}{1} shells and the
147: ionized gas traced out by the \ion{H}{2} regions and \ion{H}{2} filaments
148: (\cite{KI99}). Furthermore, the correlation between the \ion{H}{1}
149: shells and 122 OB stellar associations in the LMC is not very tight (\cite{KI00}). Rhode et al. (1999) claimed that there is no remnant
150: star clusters at the center of the \ion{H}{1} holes in Holmberg II,
151: and it is inconsistent with the SNe hypothesis.
152: Moreover an energy source generating
153: the kpc-scale supergiant \ion{H}{1} holes is a puzzle.
154: These issues raise an interesting question about whether \ion{H}{1}
155: shells/holes have been formed by the interaction between stars and the ISM or not.
156: %Even though the lifetime of the \ion{H}{1} shells can be longer than the
157: %lifetime of OB stars.
158:
159: Recent hydrodynamical simulations by Wada \& Norman (1999) demonstrate that
160: a gravitationally and thermally unstable disk, which models an ISM in
161: galaxies, can generate the cold, dense
162: clumps and filaments surrounded by hot, diffuse medium.
163: They show that porous structure is a natural consequence
164: of the non-linear evolution of the ISM.
165: This result strongly
166: suggests that some fraction of \ion{H}{1} shells, supershells or
167: holes seen in galaxies does not relate to the
168: interaction between stellar activities and the ISM.
169:
170: The other important component of the ISM is the molecular gas.
171: Molecular clouds are potential sites of star formation,
172: but their formation mechanism, and the relationship between their evolution
173: and star formation have been poorly understood.
174: Dense molecular hydrogen is traced by the rotational transition of
175: CO, and a recent high resolution survey of $^{12}$CO (J$=$1-0) in
176: the LMC with NANTEN, which is
177: 4m millimeter-wave telescope at Las Campanas Observatory, established
178: a comprehensive view of the giant molecular clouds in the LMC (\cite{FU98}). Since the gas clumps, whose size is typically $10-100$ pc,
179: in the simulations by Wada \& Norman (1999)
180: are dense ($n > 1000$ cm$^{-3}$) and cold ($T=10-100 $ K), they are
181: counterparts of the observed giant molecular clouds in the LMC or other galaxies.
182:
183: Using numerical simulations,
184: Feitzinger et al. (1981) and Gardiner, Turfus, \& Putman (1998)
185: have studied global dynamics of the ISM and star formation in
186: the LMC.
187: Although the numerical
188: methods they used are different, the stochastic self-propagating
189: star formation model (e.g. Seiden \& Gerola 1984) in Feitzinger et al. (1981) and
190: sticky particle method in Gardiner, Turfus, \& Putman (1998),
191: are both phenomenological concerning
192: the structure and dynamics of the
193: ISM and star formation processes.
194: Unfortunately the spatial resolution (100 pc in Feitzinger et al. 1981)
195: and mass resolution (6$\times 10^4 M_\odot$, which corresponds
196: to $\sim 200$ pc for $n\sim 1$ cm$^{-3}$ and $H \sim 100$ pc, in Gardiner, Turfus,
197: \& Putman 1998)
198: in these simulations are not good enough for comparison between
199: models and the recent high resolution observations ($\sim$ 15 pc for HI
200: (\cite{KIM98}) and $\sim $ 40 pc for CO (J$=$1-0) (\cite{FU98})).
201:
202: In this paper, we apply the numerical scheme used in Wada \& Norman
203: (1999) to a LMC-type model galaxy, and we
204: conduct two-dimensional hydrodynamical simulations of
205: the multi-phase ISM in a LMC-like model galaxy,
206: taking into account self-gravity of the gas, radiative cooling and
207: various heating processes, such as supernova explosions.
208: High spatial resolution (7.8 pc) and
209: a modern hydrodynamical scheme allow us to model the star formation and
210: its feedback less phenomenologically, and therefore it needs less assumptions
211: than the previous semi-analytic and numerical approaches (e.g. a review
212: by \cite{SF95}).
213: In contrast to the model used in Wada \& Norman (1999),
214: the rotation curve assumed here is nearly rigid as suggested in the LMC
215: or other LMC-type dwarf galaxies.
216: We derive the \ion{H}{1} and CO brightness map from
217: the simulations. Then we compare these simulation results with the \ion{H}{1}
218: and CO observations of the LMC.
219:
220: In \S 2, we describe our numerical method and
221: models. In \S 3, the numerical results with and without star formation
222: are discussed on morphology and statistical structure of the ISM, then they are compared with the observations.
223: Position-Velocity diagrams derived from the numerical results
224: are also discussed.
225: Comparison with past numerical simulations and
226: other implications are discussed in
227: \S 4, and conclusions are presented in \S 5.
228: %
229: \section{NUMERICAL METHOD AND MODELS}
230: %
231: Taking into account the multi-phase and inhomogenous nature of the ISM is
232: crucial for realistic simulations of global star formation in galaxies.
233: Stars are formed preferentially in cold molecular gas,
234: and supernovae produce
235: low density, high temperature gas. Interaction between such different
236: phases are also impornant (e.g. \cite{MO}; \cite{IH}).
237: In order to simulate dynamics of the inhomogeous, multi-phase
238: interstellar matter and star formation, our numerical method has
239: following features.
240: (1) High spatial resolution (7.8 pc), for diffuse
241: gas to high density gas. We use an Euler mesh code with
242: 1024$^2$ Cartesian grid points.
243: (2) The simulations are global in order to study effects of galactic
244: rotation and other kpc scale phenomena to the local structure of the ISM.
245: (3) Self-gravity of the gas is calculated.
246: (4) Radiative cooling for gas whose temperature is between $10^8$ K and
247: 10 K is taken into account.
248: (5) Various heating processes are included. Here we assume photoelectric heating due to dust grains and UV radiation, SN explosions, and stellar wind from massive stars.
249: (6) High numerical accuracy for shocks is achieved with
250: a modern hydrodynamical scheme. This is crucial,
251: because the ISM is usually supersonic, and SNe produce
252: strong shocks.
253:
254: \setcounter{footnote}{0} We solve the following hydrodynamical
255: equations and the Poisson equation numerically
256: in two dimensions to simulate the evolution of a rotating gas disk in
257: a stellar potential.
258:
259: \begin{equation}
260: \frac{\partial \rho}{\partial t} + \nabla \cdot (\rho \mbf{v}) = 0,
261: \label{eqn: rho}
262: \end{equation}
263:
264: \begin{equation}
265: \frac{\partial \mbf{v}}{\partial t} + (\mbf{v}
266: \cdot \nabla)\mbf{v} +\frac{\nabla p}{\rho} + \nabla \Phi_{\rm ext} +
267: \nabla \Phi_{\rm sg} = 0, \label{eqn: rhov}
268: \end{equation}
269:
270: \begin{equation}
271: \frac{\partial
272: E}{\partial t} + \frac{1}{\rho} \nabla \cdot [(\rho E+p)\mbf{v}] =
273: \Gamma_{\rm UV} + \Gamma_{\rm \star}-\rho \Lambda(T_g), \label{eqn: en}
274: \end{equation}
275:
276: \begin{equation}
277: \nabla^2 \Phi_{\rm sg} = 4 \pi G \rho, \label{eqn: poi}
278: \end{equation}
279:
280: where, $\rho,p,\mbf{v}$ are the density, pressure, and velocity of
281: the gas, and the specific total energy $E \equiv |\mbf{v}|^2/2+
282: p/(\gamma -1)\rho$, with $\gamma= 1.4$. We assume a time-independent
283: external potential $\Phi_{\rm ext} \propto v_c^2/(R^2+
284: a^2)^{1/2}$, where $a=2.5$ kpc is a core radius of the potential and
285: $v_c = 63$ km s$^{-1}$ is
286: the maximum rotational velocity, and they are determined to
287: mimic the rotation curve of the LMC (\cite{KIM98}).
288: Since the total gas mass is about 10 \% of the total dynamical mass (see
289: below), the effect of self-gravity of the gas to the
290: rotation curve is not significant. In fact the rotation curve after the system
291: evolves is close to the rigid rotation (see \S 3.3 and
292: the position-velocity diagrams in Fig. 10).
293:
294: We also assume a cooling function
295: $\Lambda(T_g) $ ($10 < T_g < 10^8$ K) (\cite{SN}).
296: The cooling processes taken into account are, (1) recombination of H, He, C, O, N, Si and Fe, (2) collisional excitation of HI, CI-IV and OI-IV, (3)
297: hydrogen and helium bremsstrahlung, (4) vibrational and rotational excitation of H$_2$ and (5) atomic and molecular cooling due to
298: fine-strucure emission of C, C$^+$ and O, and rotational
299: line emission of CO and H$_2$.
300: %\footnote{ In previous simulations on the
301: %multi-phase ISM, the minimum temperature was assumed as 300 K
302: %(\cite{RB93}; \cite{RB95}; \cite{VZ95a}; \cite{VZ95b}) or 100 K
303: %(\cite{VZ96}).}
304: As a heating source, $\Gamma_{\rm UV}$,
305: we assume a uniform UV radiation field (\cite{GI}), which is
306: normalized to the local interstellar value, and photoelectric
307: heating by grains and PAHs. The bulk of the heating for the $10^2-10^4$
308: K gas is provided by photo-emission of UV irradiated dust grains
309: (\cite{BT94}).
310:
311: We consider two feedback
312: effects of massive stars on the gas dynamics, namely stellar
313: winds and supernova explosions, although our results show that
314: the former is less
315: effective than the latter.
316: We first identify cells which satisfy
317: criteria for star formation. The criteria are a surface density
318: threshold $(\Sigma_g)_{i,j} > \Sigma_c$ and a
319: critical temperature $(T_g)_{i,j} < T_c$ below which star formation is
320: allowed. The surface density is defined as $\Sigma_g =2 H \rho$,
321: where $H$ is the scale height and assumed to be constant (100 pc).
322: In these simulations we take
323: $\Sigma_c = 40 M_\odot$ pc$^{-2}$ and $T_c = 15$ K (model `Star
324: Formation 1', hereafter `SF1') and ten times larger threshold density,
325: $\Sigma_c = 400 M_\odot$ pc$^{-2}$ and $T_c = 15$ K (model SF2).
326: These criteria are chosen to satisfy the condition, $L_J < 0.1 \Delta$
327: or $L_J < 0.01 \Delta$, where $L_J$ and $\Delta$ are the Jeans length
328: and the size of each cell.
329: The volume filling factors of the star-forming cells to the
330: total volume are typically
331: $\sim 5\times 10^{-3}$ and $\sim 5\times 10^{-4}$ for model SF1 and
332: SF2, respectively. Namely the star forming sites are cold, clumpy
333: regions. We also assume the star-forming criteria must last during
334: $10^5$ yr for each cell
335: before the star formation is initiated.
336: Since the spatial resolution is fine enough compared to the GMC size,
337: we do not have to assume any
338: global criteria for gravitational instability to identify star forming
339: sites. Assuming the Salpeter IMF with $m_u = 120 M_\odot$ and $m_l = 0.2
340: M_\odot$, we create test particles representing massive stars
341: ($\geq 8 M_\odot$) in the star forming cell.
342: The kinematics of the test particles in the external potential and
343: the self-gravity potential of the gas are traced by the second-order
344: time-integration method.
345: The stars (test particles) inject energy
346: due to stellar winds during their lifetime which is approximately $\sim
347: 10^7$ yr (\cite{LE98}).
348: When one of stars represented by a test particle explodes as
349: supernova, an energy of $10^{51}$ ergs is injected into the cell as
350: thermal energy where the test particle is located at that moment.
351: The cooling procedure is not used for such
352: cells, but the cells adjoining the supernova cell are treated
353: normally. Evolution of the supernova remnant is very dependent on
354: its environment. In contrast to past numerical studies of the ISM
355: with supernova explosions on a galaxy scale, we do not introduce
356: simple evolutionary models for the SNR, such as the Sedov solution, and
357: heating efficiency for the ISM due to a supernova.
358: With our code, two-dimensional evolution of blast waves caused by
359: supernovae in an inhomogeneous and
360: turbulent medium with global rotation is simulated
361: explicitly. Therefore, we can trace consistently the thermal and
362: dynamical evolution of the ISM around the star forming regions and the
363: associated supernovae remnants and superbubbles (c.f. \cite{NI})
364:
365: %In the previous numerical models including the star formation and
366: %feedback, it is impossible to follow the evolution of supernova
367: %remnants due to the poor spatial resolution.
368: %Therefore it is unclear how much energy and momentum from a supernova
369: %is transferred to the ISM. On the contrary, we can follow
370: %the dynamical and thermal
371: %evolution of supernova remnants and their interaction with
372: %the inhomogeous ISM can be followed. As a result, we do not
373: %need parameters, such as the 'heating efficiency' and
374: %any analytical models such as the Sedov-solution.
375:
376: The hydrodynamic part of the basic equations is solved by the third-order AUSM
377: (Advection Upstream Splitting Method) (\cite{LS}).
378: % with a van
379: % Leer-type flux splitting process (Liou 1996).
380: After testing this
381: code for various hydrodynamical 1-D and 2-D problems, we find that
382: AUSM is as powerful a scheme for astrophysical problems as are the
383: PPM (\cite{WC}) and Zeus (\cite{SN92}) codes.
384: % We achieve third-order
385: % spatial accuracy with MUSCL (\cite{VL}).
386: %To satisfy the TVD
387: % condition using MUSCL, we use the minmod limiting function.
388: More details about our numerical code and test results are described in
389: Wada \& Norman (2000, in preparation).
390:
391: We use $1024^2$ Cartesian grid points covering a 8 kpc $\times$ 8 kpc
392: region. The spatial resolution is 7.8 pc.
393: A periodic Green function is used to calculate the self-gravity for the
394: 8 kpc $\times$ 8 kpc region with $2048^2$ grid points (\cite{HE}). The
395: second-order leap-frog method is used for the time integration. We
396: adopt implicit time integration for the cooling term in
397: equation (3).
398:
399: The initial disk is an axisymmetric and rotationally supported
400: ($R=3.7$ kpc) with uniform surface density,
401: $\Sigma_g = 12 M_\odot $ pc$^{-2}$,
402: and the total gas mass is $5\times 10^8 M_\odot$ (\cite{KIM98}).
403: Random density and temperature fluctuations are added to the
404: initial disk. Amplitude of the initial fluctuations
405: is less than 5 \% of the unperturbed
406: values and have an approximately white noise distribution. The initial
407: temperature is set to $10^4$ K ($R \leq 3.7$ kpc)
408: and $10^2$ K ($R > 3.7$ kpc). The reason why we chose the low
409: temperature in the outer region is to avoid the numerical artifact of
410: the boundaries, i.e. reflection and generation of waves or shocks,
411: for the initial evolution of the gas disk.
412: In ghost zones at the boundaries of the
413: calculating region (i.e. $8\times 8$ kpc$^2$), all physical quantities remain at their initial
414: values during the calculations. From test runs we found that this
415: boundary condition is much better than `outflow' boundaries, because
416: the latter cause strong unphysical reflection of waves at the
417: boundaries.
418:
419: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
420: %
421: \section{RESULTS}
422: %
423: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
424: \subsection{Evolution and Structure of the ISM}
425: \subsubsection{Model without Star Formation}
426:
427: Figure 1 shows the time evolution of density and temperature of
428: a model without star formation and its feedback
429: (hereafter we call this model ``model NSF'').
430: Due to gravitational and thermal instability in the gas
431: disk, clumpy fluctuations evolve in the first $10^8$ yr, and
432: then the clumps merge and form larger structure.
433: They are deformed by the local tidal field and global shear
434: in the non-linear phase,
435: and as a result, filamentary structure is formed.
436: Higher density clumps ($ \Sigma_g >10^3 M_\odot$ pc$^{-2}$)
437: are formed in the filaments due to the
438: gravitational instability, or collisions between the filaments.
439: The high temperature cavities ($T>10^5$ K) are
440: formed due to shock heating.
441: The dynamical time scale for the large ($L \sim 200$ pc)
442: high temperature ($T \sim 10^5$ K)
443: region seen in the temperature panel at $t=800$ Myr is
444: $\tau_{\rm dyn} \sim L /\Delta v \sim 200$ pc$/$20 km s$^{-1} \sim 10^7$
445: yr. The cooling time, $\tau_{\rm cool}$ on the other hand, for this low density region
446: ($n\sim 10^{-3}$) is $\sim 2\times 10^7$ yr (Spitzer 1977),
447: which is comparable to
448: $\tau_{\rm dyn}$. In fact the high temperature cavity is a tentative
449: structure that lasts $\sim 10^7$ yr.
450: Photo-electric heating and UV irradiation contribute to form $T \sim 10^4$ K gas.
451: In the high density ($ \Sigma_g >10^2 M_\odot$ pc$^{-2}$) filaments and
452: clouds, the temperature is less than 100 K because the radiative cooling
453: is effective.
454: The global strucure of the disk does not change significantly
455: after 400 Myr, and
456: it reaches a quasi-steady state.
457: Figure 2 is time evolution of the volume filling factor, $f_v$, of
458: each temperature level and time evolution of
459: the maximum density.
460: This also represents that the global ($>$ kpc scale)
461: system reaches a quasi-stationally state after $t\sim 300$ Myr.
462: This period is comparable to the local free fall time $t_{ff}$ for
463: the initial density, $t_{ff} \sim 200 (\Sigma_g/12 M_\odot$
464: pc$^{-2})^{-1/2}$ Myr.
465: At the global quasi-steady state, the warm gases ($T_g = 100-10^4$ K) occupy
466: a large volume ($\sim$ 60 \% of the whole calculating region), and
467: $f_v \sim $ 20 \% for the cold ($T_g < 100$ K) and $f_v \sim $ 20 \% for
468: hot gas ($T_g = 10^4-10^5$ K).
469: The maximum density in the same plot
470: show slower evolution after $t\sim 300$ Myr than
471: the non-linear evolutional phase.
472:
473: One should note, however, that
474: the globally stable state does not mean the local filamentary and porous
475: structure whose scale is less than 1 kpc.
476: It is rather quite dynamic and transient, which
477: is similar to the past turbulent ISM models
478: (\cite{BL80}; \cite{CP85}; \cite{CB88}; \cite{RB95};
479: \cite{VZ95a}; \cite{VZ95b}; \cite{GP99}).
480: Clump mass is increasing during the initial linear and non-linear
481: evolutional phase ($t < 200$ Myr)
482: due to the mass accretion and merging
483: with other clumps or filaments, but in the quasi-stationally phase,
484: disruption processes of the clumps, such as local tidal field (e.g.
485: interaction between clumps, or shear) or shocks,
486: prevent monotonic increase of the mass of each clump.
487: Since we introduce the cutoff temperature for the
488: cooling (i.e. 10 K), the gaseous pressure prevent
489: the dense clouds further collapsing towards singularities.
490: The angular momentum also supports the clouds (see \S 3.3).
491:
492: The density and temperature structures (Fig. 1) are
493: similar to those in the model of Wada \& Norman
494: (1999), where a smaller disk is investigated (the radius is 1
495: kpc), but the present model shows kpc-scale inhomogeneity and
496: a more asymmetric distribution against the galactic center.
497: This difference is probably caused by the difference of the rotation
498: curves used in the two models:
499: a rigid rotation or differential
500: rotation. With a rigid rotation,
501: random and turbulent motion
502: dominates the circular rotation in the central region.
503: With the global shear, on the other hand, global spirals develop toward
504: the galactic center.
505:
506:
507: \subsubsection{Models with Star Formation}
508:
509: Figure 3 represents density and temperature maps of the model SF2
510: at $t= 833$ Myr. The model SF1, which has ten times smaller threshold
511: density for the star formation criterion (see \S 2),
512: shows very similar density morphology
513: and temperature structure.
514: The red regions in the temperature map are
515: hot gaseous regions where $T_g >10^6 $ K. They are young ($<10^6$ yr) supernova
516: remnants.
517: Figure 3 also shows that most young supernova remnants are not axisymmetric. This is caused by that the background ISM is highly inhomogeous, and the radiative cooling at
518: the dense filaments is so effective, and that it prevents the blastwaves
519: expanding axi-symmetrically.
520: Typical size of the hot cavities is less than 500 pc.
521: Some bubbles together form kpc scale `super bubbles'. The hot region around
522: $(-3,-1)$ is one of the examples. Note that the size of the cavities
523: is expected to change away from the galactic plane in a 3-D model
524: (see also \S 4).
525:
526: Although the filamentary and clumpy structures of the gas
527: which result in the star forming models
528: are similar to those of model NSF, the kpc scale inhomogeneity seen in the
529: the model (Fig. 1) is not apparent in the SF models. The large scale
530: inhomogeneity in the model SF1 and SF2 is less prominent than that of the
531: model NSF. This is because that the time scale for
532: making the kpc-scale holes,
533: which is a dynamical time scale, $\sim 10^8$ yr, is
534: much longer than the the time scale for supernova explosions
535: and evolution of the blastwaves, $\lesssim 10^6$ yr.
536: Approximately $10^3$ supernovae per kpc$^2$
537: explode in this model during $\sim 10^8$ yr. The kpc-scale
538: low density cavities, which is formed due to evacuation of gas
539: by the effect of the gravitational and thermal instabilities,
540: cannot evolve under such frequent supernova explosions.
541:
542: We find that the supernova rate is fluctuate in a time scale
543: $\sim 10^7$ yr, but it stays in a range
544: 4-10$\times 10^{-4}$ yr$^{-1}$ during 6$\times 10^8$ yr in model SF2
545: (Fig. 4). This behavior also appears in model SF1,
546: but the SN rate becomes smaller for the larger threshold density
547: ($\sim $ 8-12$\times 10^{-4}$ yr$^{-1}$).
548: ROSAT observed 46 supernova remnants and more candidates
549: in the LMC (\cite{HP99}) . If we assume the
550: ages of SNRs are between $\sim $ 10000 yrs and $\sim$ 30000 yrs, then
551: we have SN explosion every 250 yrs or 750 yrs. This indicates
552: about 0.0013/yr or 0.004/yr for SN rate in the LMC.
553: If we use the larger threshold density for the star formation in our
554: model than that in model SF2, this results in
555: smaller supernova rate than $\sim 0.001$ yr$^{-1}$.
556: Therefore the threshold density much larger than $400 M_\odot$ pc$^{-2}$
557: would be excluded for modeling the LMC,
558: if we assume that the star formation
559: efficiency is 10 \% or less and the standard IMF.
560:
561: % of $\sim 10^{6-7}$ yr, which is about the life span of the massive
562: %stars. When the massive stars die as SNe, they generate blastwaves and trigger
563: %star formation in ambient dense clumps.
564: %The newly formed massive stars explode after $\sim 10^6$ yr, and
565: %again trigger star formation.
566: %As a result, recursive bursts of SNe are observed.
567:
568: In Figure 5, the star particles at $t=245$ Myr in model SF1 are plotted,
569: and one may compare it to the density distribution (the right panel)
570: of the same snapshot. Massive stars are not uniformly distributed, but
571: they form clusters (`OB associations'). It is notable that the
572: distribution of the star clusters does not necessarily correlate to
573: the inhomogenous gaseous structure. In other words, cavities are
574: not necessarily associate with the `OB associations'.
575: This is also seen in simulations by other groups, for example
576: 2-D simulations of a highly compressible isobaric, non-selfgravitational
577: fluid with star formation (\cite{SC99}).
578:
579: \subsection{Line Emission Maps and Comparison with the Observations}
580:
581: With the numerical simulations, we have density, temperature, and
582: velocity fields (spatial resolution: 7.8 pc).
583: Using this information as an input, we compute \ion{H}{1} 21 cm brightness map
584: and CO (J=1-0) line map for model NSF and SF1 and SF2.
585: These maps can be directly compared with
586: the recent high resolution surveys of the LMC: \ion{H}{1} with ATCA (Kim et al. 1998)
587: and CO (J=1-0) with NANTEN (\cite{FU98}).
588:
589: %\subsubsection{HI 21 cm}
590:
591: With the numerical results described above, the following procedure is
592: followed to compute a \ion{H}{1} 21 cm brightness map.
593: It is assumed that atomic hydrogen is in mostly neutral form in those
594: regions where the temperature is at least a factor of 2 less than 8000 K,
595: a value typical of (mostly ionized) HII regions, and motivated by measurements
596: of the kinetic temperatures of HI clouds and HI intercloud gas.
597: It is also assumed that the \ion{H}{1} level populations of interest
598: follow a thermal distribution in those regions along the line of
599: sight, and that this mostly neutral gas dominates the integrated emissivity.
600: The line of sight is face-on, i.e., perpendicular to the grid used for the
601: hydrodynamic simulations. The two-dimensional density indeed is used as a
602: column density in the radiative transfer calculations. The latter are done
603: in three dimensions by assuming a scale height for the neutral gas of
604: H=100 pc, thus converting the column density in a (constant) local density
605: for each point along the line of sight. It is this three-dimensional grid,
606: with two-dimensional hydrodynamic information, that is
607: used to determine the \ion{H}{1} central brightness temperature in the Monte Carlo
608: procedure (Spaans 1996), where the ambient radiation field and its interactions
609: with matter is represented by a discrete number of ``photon packages''.
610: This method is three-dimensional and explicitly includes optical depth effects
611: as well as the detailed velocity field of the hydrodynamical simulations
612: for photons that travel along lines of sight that are not face-on.
613: These latter photon trajectories need to be incorporated when the
614: level populations are not in thermal equilibrium, i.e., for CO.
615: A local velocity dispersion $\Delta V$ of $1.29\times 10^4 T^{1/2}$ cm s$^{-1}$
616: is adopted for a kinetic temperature $T$, with a minimum of $0.5$ km s$^{-1}$
617: due to micro turbulent motions. For each grid point in the hydrodynamical
618: simulation, the Monte Carlo radiative transfer then yields the integrated
619: \ion{H}{1} 21 cm intensity along the line of sight.
620:
621: %\subsubsection{CO (J$=$1-0)}
622:
623: For the CO (J$=$1-0) line, the approach is similar to the \ion{H}{1} case, with the
624: appropriate correction factor in $\Delta V$ for the different atomic
625: weight of the CO molecule.
626: Furthermore, a carbon chemistry is added to compute the abundance of
627: CO (\cite{SV97}), for a dust abundance equal to $1/5$ of
628: Solar. This chemistry is well understood (\cite{VB88}),
629: and care has been taken to include, for each computed line of sight and
630: corresponding column density, the important self-shielding transitions of
631: H$_2$ and CO (c.f. \cite{ST94}). The ambient average interstellar
632: radiation field, required for the ambient chemical balance, is determined by
633: scaling with the LMC B band surface brightness in mag per square arcsecond
634: with respect to Galactic. This typically yields an enhancement of a factor
635: of $3-5$ in the mean LMC energy density compared to Galactic. It is assumed,
636: because the hydrodynamical simulations are two-dimensional, that, to compute
637: the local CO emissivity and the ambient chemical equilibrium, the gas density
638: is given by the particular line of sight column density divided by a scale
639: height of $H=100$ pc for the neutral gas. The latter number $H$ does not
640: strongly influence the qualitative features in the presented maps.
641:
642: Figure 6 presents \ion{H}{1} 21 cm and CO (J=1-0) line brightness temperature maps
643: calculated from the numerical data of the NSF model at $t=800$ Myr,
644: using the procedure described above.
645: Figure 7 is the same plot as Fig. 6, but for model SF2 at $t=834$ Myr.
646: Size of the \ion{H}{1} filaments, shells, and holes in model NSF is
647: $\sim$ 1 kpc. The filaments and cavities in the outer region ($R > 2$
648: kpc) of model SF2 are also kpc-scale.
649: We would like to emphasize
650: that large cavities ($>$ kpc) and filaments in model NSF are NOT caused by a direct dynamical effect
651: of star forming activities, but by the non-linear development of
652: gravitational
653: and thermal instabilities.
654:
655: The CO emission is localized in many clumps (size $\sim$ 10-100 pc) or
656: clouds complexes (size $\sim$ 0.1-1 kpc) in model NSF.
657: Such CO ``cloud'' complexes could be sites for active star forming regions, such as
658: the 30 Dor star forming region in the LMC.
659: Model SF2 shows more uniform distribution of ``stars'' than in model NSF.
660:
661: Figure 8 (a) and (b) are mass spectra of {\it molecular clouds} obtained from
662: the CO brightness temperature ($T_B$) distribution of
663: models NSF and SF2.
664: We identify {\it clouds} and their mass by the following procedure.
665: Assuming a threshold $T_B$ for the CO map (Figs. 6 and 7), then
666: we have many {\it islands} of CO emission. Most {\it islands}
667: are not round in shape, but elongated or filament-like shape.
668: We derive the mass of the {\it islands} using the surface
669: density of the simulation data. Therefore the mass in Fig. 8 is the
670: total gas mass of each {\it island},
671: not the virial mass. Here we plot three histograms for
672: three different thresholds.
673: The mass spectra shows a power-law, which is
674: roughly $dN_c/dM_c \propto M_c^{-1.7}$ for model NSF, where $dN_c$ is
675: number of clouds between the mass $M_c$ and $M_c +dM_c$, but this slope does
676: not significantly depend
677: on the threshold brightness.
678: On the other hand, the spectrum of
679: the star forming model steeper than the model without the energy
680: feedback especially for small threshold:
681: $dN_c/dM_c \propto M_c^{-1.7}$, $M_c^{-2.3}$, and
682: $M_c^{-2.7}$ for model SF2 with
683: $T_B=$ 100, 50, and 30 K, respectively.
684: The behavior of the slope to the threshold $T_B$ in the model NSF
685: and SF2 means that the stellar energy feedback changes structure of
686: low density envelope of the dense clumps.
687:
688:
689: The NANTEN survey discovered about hundred molecular clouds in
690: the LMC, and the mass spectrum of them is
691: $dN/dM_{\rm vir} \propto M_{\rm vir}^{-1.5\pm0.1}$ for a range
692: between $10^{5-6} M_\odot$ (\cite{FU98}), where $M_{\rm vir}$ is
693: the virial mass of the cloud estimated by using the observed line width.
694: In our model, smaller clouds ($M_c < 10^5 M_\odot$) tend to have smaller
695: mass compared to their virial mass estimated from their internal velocity dispersions.
696: In other words, the smaller clouds are not in equilibrium, but
697: rather in a transient phase (\cite{VZ95a}).
698: We find a rather weak but positive correlation between the cloud mass $M_c$ and
699: the virial mass $M_{\rm vir}^{1/2}$ in our model.
700: The mass spectrum of the model SF2 would be
701: $dN_c/dM_{\rm vir} \propto M_{\rm vir}^{-1.4}$, if we use the lowest
702: threshold $T_B=30$ K.
703: A similar discussion for the interpretation of the cloud mass spectra based on
704: two-dimensional numerical simulations of the interstellar medium has been given by
705: V$\acute{\rm{a}}$zquez-Semadeni et al. (1997) (see \S 4).
706:
707: The maximal mass of the clouds is approximately $10^{6.5} M_\odot$
708: and $10^{5.5-6} M_\odot$ in models NSF and models with the star formation
709: (SF1 and SF2).
710: This implies that very massive clouds ($> 10^6 M_\odot$) are difficult to grow under the frequent
711: supernova explosions.
712: The maximal masses of the observed molecular clouds are
713: about $10^{6.5} M_\odot$ (\cite{FU98}), which seems to prefer
714: the model NSF.
715: However, this does not mean that the star formation and its energy
716: feedback are not important for shaping the ISM
717: in the LMC, but it implies that the largest clouds
718: could evolve in preferentially an environment
719: where SNe are not frequent.
720: Though one should be careful to make comparison between the
721: observed and numerical mass spectrum, because the definition and
722: identification of the ``cloud'' is not exactly identical, and
723: it is related to noise level in observations.
724:
725: We also find that the \ion{H}{1} map of the computational model shows
726: similar statistical properties to the observed \ion{H}{1} map.
727: Figure 9 shows the \ion{H}{1} luminosity function, i.e. the
728: histogram of the observed and numerical \ion{H}{1} map of the LMC.
729: %The shape of \ion{H}{1} luminosity function is consistent with
730: %a Log-Normal like distribution for both
731: %observed and numerical data. and their coincidence is remarkable.
732: %The \ion{H}{1} luminosity function from the observations has been derived by
733: %$\Sigma $ $T_i$ $v_i$/$\Sigma$ $T_i$
734: %for the centroid velocity in each line of sight ($T_i$ and $v_i$
735: %are the brightness temperature and velocity corresponding to the $i^{th}$ channel,
736: %and summations has been taken for the \ion{H}{1} full line widths).
737: The \hi\ brightness temperature has been computed from $T_B = {S_\nu
738: \lambda^2}/{2k_B\Omega_{sb}}$. $S_\nu$ is the \hi\ flux density, $k_B$ is
739: the Boltzmann constant, and $\Omega_{sb}$ is the solid angle of the
740: synthesized beam of ATCA mosaiced map. The \hi\ intensity is determined from
741: the integral of the brightness temperatures $\int T_B dv$ over the peak \hi\
742: line profile, where $dv$ is the channel width in kilometers per second.
743: The \ion{H}{1}
744: luminosity functions appear to be log-normal like
745: distribution for both observed and model NSF. The distributions of
746: the \ion{H}{1} luminosity function of the model NSF and SF2 for lower
747: surface brightness ($T_B < 1000$ K)
748: are similar to the observed one.
749: The model SF2, on the other hand, shows excess above $T_B > 1000$ K than
750: the observations. These high brightness regions are originated from
751: shock compressed, dense gas.
752: %
753: \subsection{Position-Velocity Diagrams}
754: %
755: Position-Velocity (PV) diagrams give us information on
756: the kinematics of the ISM.
757: In Fig. 10 (a), we show the PV diagram, in which y-component of
758: velocity ($v_y$) is integrated through $x$-positions,
759: for the gas $\Sigma_g < 10^2 M_\odot$ pc$^{-2}$ in model NSF.
760: Many arc-like structures in Fig. 10 (a)
761: look like expanding shells originating from explosional phenomena in the ISM.
762: However, they are NOT caused by
763: explosions because there is no energy input due to supernovae in
764: model NSF.
765: The arcs in the PV diagram
766: are actually caused by the gases that form filament-like structure seen
767: in the density map (Fig. 1).
768: The filaments and shells exhibit non-circular motion of the order of
769: 10 km s$^{-1}$, and their shape continuously changed.
770: Here we would like to emphasize that,
771: it is hard to distinguish, on the PV diagrams, between
772: expanding shells and such shell-like
773: structures which are changing in shape due to local random motion.
774:
775: From Fig. 10 (b),
776: which is the same as Fig. 10(a), but for
777: $\Sigma_g > 10^2 M_\odot$ pc$^{-2}$,
778: we find that the dense and compact clumps, rotating with
779: $\sim 10-30$ km s$^{-1}$, which are recognized
780: as steep `dotted lines' on the
781: PV-diagram. Here we can
782: identify about 80 clumps, and about half of them show retrograde rotation
783: against the sense of galactic rotation. A representative one can be seen
784: at $(x,v_y) = (-0.5, -50)$ and less prominent one is at $(1.5,25)$ and
785: $(3,75)$.
786: Rotation of the clouds is important for the internal structure,
787: motion, and star formation in the molecular clouds.
788: Unfortunately, the spatial resolution in the NANTEN survey ($\sim 40 $
789: pc) is insufficient to resolve the rotation of each molecular
790: clouds of the LMC.
791: High resolution observations and statistical analysis of kinematics of
792: molecular clouds in external galaxies are necessary to understand
793: the formation of molecular clouds and star formation processes.
794:
795: Figure 10 (c) shows the similar features shown in Fig. 10 (a), but for model SF1 at $t=$ 610
796: Myr. Arc-like structures seen in the PV map for the NSF are not
797: clearly shown in this map. This means that
798: the line-of-sight velocity field is much more random and chaotic
799: than the NSF model, and
800: there is no prominent large-scale coherent motion
801: in the ISM of the model with stellar energy feedback.
802:
803: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
804: %
805: \section{DISCUSSION}
806: %
807: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
808: The comparison of our model calculations with the observations of the LMC
809: indicates that the star formation models are more consistent with the formation
810: of small ($\ll$ kpc ) and hot bubbles ($T_g$ $>$ 10$^7$ K), which have been detected
811: by X-ray diffuse emission. The power law slope of cloud mass spectrum of the
812: models is close to that of observed one. On the other hand, the model calculations
813: without the stellar feedback show even better agreement with global
814: inhomogeneity
815: of the HI and CO gas in
816: the LMC, and also
817: the HI distribution function is well represented in the model NSF.
818: Therefore,
819: it is most likely that
820: energy feedback to the ISM from massive stars and
821: supernovae explosions are important processes for producing hot gas,
822: but they do not necessarily dominate dynamics and
823: formation processes for the large scale inhomogeneity observed
824: in \ion{H}{1} and
825: CO gas in the LMC.
826: However, one should note that the initial condition of the
827: models (i.e. axisymmetric and uniform
828: density distributions) causes nearly uniform star formations
829: in the whole disk.
830: If we begin the simulations from
831: a more inhomogeous disk, stars will be formed non-uniformly, and
832: effects of the stellar feedback are different, depending on location of
833: the disk.
834: We suspect that the more realistic model of
835: the ISM in the LMC-type galaxy would be
836: between our two extreme cases, but that can not be achieved by
837: changing the threshold density of the star forming model.
838:
839: In our models with stellar feedback,
840: typical sizes of the hot cavities are less than 500 pc.
841: Some `superbubbles' can be
842: formed in a outer disk region, and their
843: linear sizes are $\sim$ kpc (see Fig. 3).
844: Rosen \& Bregman (1995) revealed in their two-dimensional, two-fluid (stars
845: and gas) simulations in a disk galaxy that hot bubbles are formed by
846: stellar activities, and their sizes depend on the energy injection rate.
847: In their simulations, the linear size of the
848: bubbles is up to $\sim 500-1000$ pc for the supernova rate is 0.0075-0.03
849: yr$^{-1}$. This rate is 1-2 orders of magnitude larger than that of
850: our star forming models.
851: Local, three dimensional MHD simulations of the
852: ISM with supernova explosions by
853: Korpi et al. (1999) show that the linear size of the superbubbles are
854: typically 200-400 pc, which is consistent with our SF models.
855: Gazol-Patino \& Passot (1999) compute the evolution of the ISM
856: in a region of the Galactic plane of size 1 kpc$^2$ in two-dimensional periodic
857: domain. They found that superbubbles, and the largest one has linear scale
858: is $\sim$ 500-1000 pc, due to about 3000 supernova explosions
859: in $\sim 10^7$ yr. The background density of their simulation
860: is equivalent to that in the outer region of our simulations.
861: The local ($<$ kpc scale) structure of the ISM of our
862: global simulations, where hot bubbles and cold filaments coexist,
863: is also similar to the local, 2-D simulations of the ISM with
864: a periodic boundary condition (\cite{VZ95a}).
865: In a conclusion, the past local models with supernova
866: explosions, in which superbubble formation is reported, consistent with
867: our global models with stellar energy feedback concerning the size of the
868: superbubbles.
869:
870: The cloud mass spectra of the model SF1 and SF2 are
871: steeper than that of the model NSF (Fig. 8).
872: In other words, the stellar energy input affects the evolution of
873: molecular clouds.
874: %The observed mass spectrum of CO in the LMC is similar to the
875: %model NSF. This may be interpreted as that
876: %the stellar energy feedback for the dense gas is not a dominant physics for
877: %evolution of dense clouds in the LMC. However it should be noted that
878: %the statistical errors of the observed
879: %mass spectrum are significant, and it is not straightforward to define
880: %`clouds' in observations with a single line, such as CO (J$=$1-0).
881: V$\acute{\rm{a}}$zquez-Semadeni, Balleseros-Paredes, \& Rodriguez (1997) (hereafter VBR97) analyzed their
882: 2-D hydromagnetic simulations and showed that the mass
883: spectra have the form $dN_c/dM_c \propto M_c^{-1.44\pm0.1}$
884: (c.f. their Fig. 3 and our Fig. 8), and they reported that the mass
885: spectra from observations are
886: consistent with that from simulations in which the
887: density filed had been shaped by stellar activity.
888: Note that our mass spectrum of the star formation models, the slope is
889: shallower for smaller clouds ($ M_c < 10^5 M_\odot$) than that for
890: massive clouds ($dN_c/dM \propto M^{-2}$).
891: Therefore our results might be consistent with the results of
892: VBR97 (see also \S 3.2).
893: However, direct comparison between these two results
894: should not be straightforward,
895: because there are number of differences between our simulations and
896: theirs, on the numerical scheme (AUSM vs. pseudo spectral method), the
897: boundary condition (global simulations vs. local periodic boundary),
898: and the cooling curve especially for $T_g < 100$ K ($\Lambda \ne 0$
899: vs. $\Lambda = 0$).
900: The maximum density contrast is about 5000 in their model, but
901: it is about $2\times 10^6$ in our model.
902: This difference is caused probably due to the difference of
903: the cooling curve for
904: the cold gas, and also due to the numerical scheme.
905: In their numerical method, they added a mass
906: diffusion term to the continuity equation in order to smooth out
907: the density gradients (see also \cite{VZ95a}).
908: This could affect the structure of shocks and dense gas, i.e, the cloud mass
909: spectrum.
910: The recipe for the stellar energy feedback are also different.
911: VBR97 assumed that once a star is formed, then it remains
912: fixed with respect to the numerical grid. In our models, the star
913: particles are orbited in the self-gravitational potential of the gas and
914: the external fixed potential.
915: The last point, however, would not be important, if number of stars is large
916: enough, or the ISM is fully turbulent.
917: VBR97, on the other hand, include the magnetic field in their models.
918: Therefore, again, one should be careful to make direct comparison between
919: results in the present paper and VBR97.
920:
921: Our result implies that there are two mechanisms of cavity formation.
922: One is the pure evacuation of gas from the low-density regions by the
923: effect of the gravitational and thermal instabilities. Similar
924: evacuation phenomena (\cite{EG94}; \cite{VZ96}) are also observed in the simulations of the turbulent ISM
925: without SNe (\cite{VZ95b}). The other is the formation of cavities by
926: the effect of SNe, and especially by the synchronized explosions of
927: stars in OB associations. The two processes are essential to produce
928: the whole structure of the ISM. The latter is necessary for hot ($T_g > 10^6$ K)
929: component of the ISM. These two type of low-density regions are also observed
930: in the two-dimensional, local simulations of the turbulent ISM with SNe
931: by Gazol \& Passot (1999).
932:
933: As mentioned in \S 1, the origin of supergiant \ion{H}{1} holes in
934: galaxies has been controversial (see also \cite{WB99}).
935: If the observed kpc-scale holes and shells are caused by explosional
936: phenomena only, one needs
937: highly energetic events, like Gamma Ray Bursts (\cite{LP99}).
938: However, our numerical simulations give an alternative explanation
939: about the origin of the large scale holes:
940: the non-linear evolution of the multi-phase gas disk.
941: In the multi-phase ISM, most of the gas mass is concentrated
942: in the cold, dense clumps and filaments, but volume filling
943: factor of such component is much smaller than that of
944: hot, diffuse gas.
945: The hot regions are surrounded by dense filaments as seen in
946: Fig. 1 and Fig. 3.
947: Therefore the multi-phase ISM in a quasi-steady state is
948: naturally porous. Our results suggest that if the energy feedback
949: from massive stars are not effective,
950: kpc-scale inhomogeneity can be evolved in a disk in about several $10^8$ yr.
951: Dense filaments are changing in shape due to
952: local random velocity field as well as the global shear, and often
953: show kinematics features similar to those of ``expanding shells'' as seen in
954: the position-velocity map (\S 3.3).
955: We suspect that many supergiant holes and shells
956: in dwarf galaxies do not have an explosion origin, but
957: that supergiant shells far outside the disk plane
958: can not be formed without explosional events.
959:
960: The work presented here yields the model for the ISM in a LMC-type galaxy.
961: Nevertheless there are couple of things that one should consider for constructing
962: a complete numerical model of the LMC including the interaction with
963: the Small Magellanic Cloud,
964: the optical bar, and non-uniform UV radiation field due to
965: active star forming regions, such as the 30 Dor regions.
966: The interactions with the SMC
967: are important events for the star formation history
968: in the LMC and the Magellanic Stream (\cite{MF80}; \cite{VA96}; \cite{GN96}).
969: The off-center optical bar could also perturb the global structure of
970: the ISM and star formation in the LMC (\cite{GA98}). However the
971: \ion{H}{1} mapping (\cite{KIM98}) does not show clear evidence of the
972: stellar bar, namely off-set shocks which are often seen in barred galaxies.
973: It would be interesting to investigate the effect of the
974: off-center stellar bar in our model.
975: This might contribute to the formation of the active star forming
976: region, such as 30 Dor region.
977:
978: In the present paper, we have not solved the vertical structure
979: of the ISM.
980: %We believe that cold, dense gases will form filamentary structure
981: %in three dimensional space.
982: It is expected that the hot component behaves differently in
983: three dimensions. The hot gas, which is above $10^6$ K, cannot
984: be confined in the disk plane (\cite{RB95}; \cite{DV99}),
985: and it would probably been blown out
986: from the disk plane.
987: The scale-height of the
988: hot gas should be larger than the cold gas,
989: and, as a result, the volume filling factor of
990: the hot gas would be
991: larger away from the disk plane.
992: Thus the radiative cooling would be less effective
993: in the hot gas, in three dimensions.
994: This may affect interaction between cold and hot components,
995: and the feedback process on the ISM. For example,
996: it is more difficult to form the supergiant holes by SNe.
997: In a subsequent paper, we will extend our method to
998: three-dimensional modeling, and investigate these problems.
999: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1000: %
1001: \section{CONCLUSIONS}
1002: %
1003: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1004: Using high resolution hydrodynamical simulations, we have
1005: computed that the global dynamics and structure of the multi-phase
1006: ISM in a LMC-type galaxy.
1007: Due to gravitational and thermal instability in the gas
1008: disk, clumpy fluctuations evolve, and
1009: then the clumps merge and form filamentary structure in the non-linear phase.
1010: Higher density clumps are formed in the filaments due to the
1011: gravitational instability, or collisions between the filaments.
1012: Our numerical model with the star formation allows us to provide
1013: the evolution of the blastwaves due to supernovae explosions in
1014: the rotating, inhomogeneous, multi-phase, and turbulent-like media.
1015: We find that the supernova rate in the model with stellar energy
1016: feedback is typically of the order of
1017: 0.001 yr$^{-1}$ during several hundred Myrs, but fluctuates rapidly
1018: (time scale $\sim$ 10 Myr) by a factor of three or four.
1019: The model also shows that kpc-scale low density cavities seen
1020: in the observed \ion{H}{1} map (\cite{KIM98}) are difficult to be formed under
1021: frequent supernovae, but SNe are necessary to form
1022: hot bubbles where the gaseous temperature is
1023: greater than $10^{6-7}$ K.
1024: Our result suggests that there are two possible causes of
1025: low density regions in the ISM.
1026: The kpc-scale inhomogeneity and arcs can be formed
1027: as a natural consequence of non-linear evolution of the multi-phase interstellar medium
1028: in a LMC-type galaxy. We find in the PV-diagram of the numerical models
1029: that filamentary structure in the non-star forming model, which
1030: are caused by the gravitational instability, has similar kinematics
1031: to the structure formed by SN explosions. The dense clouds are rotating,
1032: and about half of of them show retrograde rotation
1033: against the sense of galactic rotation.
1034: Using the Monte Carlo radiative transfer code,
1035: we have computed \ion{H}{1} and CO brightness
1036: temperature distributions, and compared them with those from the
1037: recent observations.
1038: The CO cloud mass spectrum in the model with stellar energy feedback is
1039: similar to the observed one (\cite{FU98}), but
1040: \ion{H}{1} distribution function is well fitted by the model without
1041: SNe.
1042: Therefore we conclude that the small scale structure and dynamics of the ISM
1043: in the LMC is mainly affected by the stellar activities, but the
1044: gravitational instability significantly contributes to the global morphology
1045: and dynamics of the interstellar matter in a kpc-scale.
1046:
1047: \acknowledgments
1048:
1049: We are grateful to Colin Norman for stimulating discussions.
1050: We also thank E. V$\acute{\rm{a}}$zquez-Semadeni (the referee)
1051: for his fruitful comments and suggestions.
1052: Numerical computations were carried out on VPP300/16R at the
1053: Astronomical Data Analysis Center of the National Astronomical
1054: Observatory, Japan, VPP700 at the SUBARU observatory, Hawaii, and
1055: VPP500 at RIKEN, Japan.
1056: We would like to thank R. Ogasawara for his help at the SUBARU observatory .
1057: KW is supported in part by
1058: Grant-in-Aids for Scientific Research (no. 1113421)
1059: of Japanese the Ministry of Education, Culture, Sports and Science,
1060: and Foundation for Promotion of Astronomy, Japan.
1061: MS is supported by Hubble Fellowship grant HF-01101.01-97A, awarded by
1062: STScI.
1063:
1064: \newpage
1065: %%%%% Figure Captions %%%%%%%%%%%%%%%%%%
1066: \figcaption{Evolution of density and temperature distributions of a
1067: model without star formation and energy feedback (model ``NSF'').
1068: The density and temperature are Log-scaled, and their units
1069: are $M_\odot$ pc$^{-2}$ and K, respectively.
1070: Time is shown at each panel in an unit, $10^8$ yr.
1071: Note: Although the gray scale is
1072: assigned for $1000-0.01 M_\odot$ pc$^{-2}$ in this plot, the maximum density is
1073: much grater than 1000 $M_\odot$ pc$^{-2}$ (see Fig. 2).}
1074:
1075: \figcaption{Evolution of the volume filling factor for
1076: four different phases, i.e. $T \leq 100$ K (filled circles), $100<T \leq 9000$ K
1077: (open
1078: circles),
1079: $9000<T < 10^5$ K (open diamonds), and $T \geq 10^5$ K (filled diamonds).
1080: The maximum density of the gas is also plotted (thin solid line).}
1081:
1082: \figcaption{Same as Fig. 1, but for
1083: a model with star formation and energy feedback (model ``SF2'') at
1084: $t=834$ Myr. }
1085:
1086: \figcaption{Evolution of a total supernova rate in model SF2.}
1087:
1088: \figcaption{Star particles and density distribution at $t=245$ Myr in
1089: model SF1.}
1090:
1091: \figcaption{Brightness temperature ($T_B$ (K)) maps of
1092: \ion{H}{1} (left) and CO (J=1-0) (right) computed from data of model
1093: NSF at $t=800$ Myr. The intensity of \ion{H}{1} is Log-scaled. }
1094:
1095: \figcaption{Same as Fig. 6, but for model SF2 at $t=834$ Myr}
1096:
1097: \figcaption{(a) Mass spectrum of CO clouds for model NSF for three
1098: thresholds of the brightness temperature, $T_B = 30$ (solid line), 50 (dashed line), and 100 K (dotted line).
1099: The thick line shows $dN_c/dM_c \propto M^{-1.7}$.
1100: (b) Same as (a), but for model SF2.}
1101:
1102: \figcaption{Probability distribution function (pdf) of \ion{H}{1}. Thick
1103: line shows pdf of model NSF ($N (T_B)$ is the number of cells for the
1104: brightness temperature, $T_B$, and $N_0$ is the total number of cells).
1105: The stars show the histogram from the \ion{H}{1}
1106: synthesis observation, with N$^{1/2}$ error bars, which is normalized
1107: for $\log(T_B) =2.8$ K of the model pdf.}
1108:
1109: \figcaption{Position-Velocity maps for (a) low density gas in model NSF,
1110: (b) high density gas in model NSF, and (c) low density gas in model SF1.}
1111:
1112: %%%%% END Figure Captions %%%%%%%%%%%%%%%%%%
1113: \newpage
1114: \begin{thebibliography}{l}
1115: \bibitem[Avillez 1999]{DV99} Avillez, M. A. D. D. 1999,
1116: Stromlo Workshop on High-Velocity Clouds, eds. Gibson, B.K. \& Putman,
1117: M.E., ASP Conference Series Vol. 166, p. 103, 103
1118: \bibitem[Bakes \& Tielens 1994]{BT94} Bakes, E. L. O. \&
1119: Tielens, A. G. G. M. 1994, \apj, 427, 822
1120: \bibitem[Bania \& Lyon 1980]{BL80} Bania, T. M. \& Lyon, J. G. 1980,
1121: \apj, 239, 173
1122: \bibitem[Chiang \& Bregman 1988]{CB88} Chiang, W. -H. \&
1123: Bregman, J. N. 1988, \apj, 328, 427
1124: \bibitem[Chiang \& Prendergast 1985]{CP85} Chiang, W. -H. \&
1125: Prendergast, K. H. 1985, \apj, 297, 507
1126: %\bibitem[Braunsfurth \& Feitzinger 1983]{BF83} Braunsfurth,
1127: %E. \& Feitzinger, J. V. 1983, \aap, 127, 113
1128: \bibitem[Deul \& den Hartog 1990]{DD90} Deul, E.R. \& den Hartog,
1129: R.H. 1990, \aap, 229, 362
1130: \bibitem[Elmegreen 1994]{EG94} Elmegreen, B. G. 1994,
1131: \apj, 433, 39
1132: \bibitem[Feitzinger et al. 1981]{F81} Feitzinger, V., Glassgold, A.E.,
1133: Gerola, H., \& Seiden, P.E. 1981, A\&Ap, 98, 371
1134: \bibitem[Fukui et al. 1999]{FU98} Fukui, Y., et al. 1999, PASJ, 51, 745
1135: \bibitem[Gardiner \& Noguchi 1996]{GN96} Gardiner, L. T. \& Noguchi, M. 1996, \mnras, 278, 191
1136: \bibitem[Gardiner, Turfus, \& Putman 1998]{GA98} Gardiner,L.T.,
1137: Turfus, C., \& Putman, M.E., 1998, ApJ, 507, L35
1138: \bibitem[Gazol \& Passot 1999]{GP99} Gazol-Patino, A., \& Passot, T., 1999, ApJ 518, 748
1139: \bibitem[Gerritsen \& Icke 1997]{GI} Gerritsen, J.P.E., Icke, V., 1997, A\&Ap, 325,972
1140: %\bibitem[Haberl Pietsch \& Motch 1999]{HP99} Haberl, F.,
1141: %Pietsch, W. \& Motch, C. 1999, \aap, 351, L53
1142: \bibitem[Haberl \& Pietsch 1999]{HP99} Haberl, F. \&
1143: Pietsch, W. 1999, \aaps, 139, 277
1144: \bibitem[Hockney \& Eastwood 1981]{HE} Hockney, R. W., \& Eastwood, J. W. 1981, Computer Simulation Using Particles (New York : McGraw Hill)
1145: \bibitem[Ikeuchi, Habe, \& Tanaka 1984]{IH} Ikeuchi, S., Habe, A., \& Tanaka,
1146: Y.D. 1984, \mnras, 207, 909
1147: %\bibitem[Kamphuis, Sancisi \& Van Der Hulst 1991]{KSV91} Kamphuis, J., Sancisi, R. \& Van Der Hulst, T. 1991, \aap, 244, L29
1148: %\bibitem[Kim et al. 1998]{KIM98b} Kim, S., Staveley-Smith, L.,
1149: %Sault, R. J., Kesteven, M. J., MCConnell, D., Dopita, M. A. \& Bessell, M.
1150: %1998, Publications of the Astronomical Society of Australia, 15, 132
1151: \bibitem[Kim et al. 1998]{KIM98} Kim, S., Staveley-Smith, L., Dopita,
1152: M.A., Freeman, K.C., Sault, R.J., Kesteven, M.J., \& McConnell, D.,
1153: 1998, ApJ, 503, 674
1154: \bibitem[Kim et al. 1999]{KI99} Kim, S., Dopita, M.A., Staveley-Smith,
1155: L., Bessell, M. 1999, AJ, 118, 2797.
1156: \bibitem[Kim \& Chu 2000]{KI00} Kim, S. \& Chu, Y.-H. 2000, in preparation.
1157: %\bibitem[Kennicutt 1989]{KN89} Kennicutt, R., 1989, \apj, 344, 685
1158: \bibitem[Korpi et al. 1999]{KO99}
1159: Korpi, M. J., Brandenburg, A., Shukurov, A. \& Tuominen, I. 1999, \aap,
1160: 350, 230
1161: %\bibitem[Liou 1996]{LI} Liou, M., 1996, J.Comp.Phys., 129,364
1162: \bibitem[Liou \& Steffen 1993]{LS} Liou, M., Steffen, C. 1993,
1163: J.Comp.Phys., 107, 23
1164: \bibitem[Loeb \& Perna 1998]{LP99} Loeb, A. \& Perna, R. 1998, \apjl,
1165: 503, L35
1166: \bibitem[Leitherer, Robert, \& Drissen 1992]{LE98} Leitherer, C., Robert, C. \& Drissen, L. 1992, \apj, 401, 596
1167: \bibitem[McKee \& Ostriker 1977]{MO} McKee, C.F. \& Ostriker, J.P. 1977, ApJ,
1168: 218,148
1169: \bibitem[Murai \& Fujimoto 1980]{MF80} Murai, T. \& Fujimoto, M. 1980, \pasj, 32, 581
1170: %\bibitem[Nordlund \& Padoan 1998]{NP} Nordlund, A., Padoan, P., 1998,
1171: %in Interstellar Turbulence, ed. J.Franco \& A. Carraminana, Cambridge,
1172: %Cambridge Univ. Press, in press
1173: \bibitem[Norman \& Ikeuchi 1989]{NI} Norman, C.A., \& Ikeuchi, S. 1989,
1174: ApJ, 345, 372
1175: \bibitem[Oey 1996]{OEY96} Oey, M. S. 1996, \apj, 467, 666
1176: \bibitem[Oey \& Clarke 1997]{OC97} Oey, M. S. \& Clarke, C.
1177: J. 1997, \mnras, 289, 570
1178: \bibitem[Passot, V$\acute{\rm{a}}$zquez-Semadeni, \& Pouquet
1179: 1995]{VZ95b} Passot, T., V$\acute{\rm{a}}$zquez-Semadeni, E., \&
1180: Pouquet, A. 1995, ApJ, 455, 536
1181: %\bibitem[Padoan, Nordlund, \& Jones 1997]{PN} Padoan,P., Nordlund, A.,
1182: %\& Jones,B.J.T. 1997, MNRAS, 288,145
1183: \bibitem[Puche et al. 1992]{P92} Puche, D.
1184: , Westpfahl, D. , Brinks, E. \& Roy, J. -R. 1992, \aj, 103, 1841
1185: \bibitem[Rhode et al. 1999]{RH99} Rhode, K.L., Salzer, J.J., Westpfahl,
1186: D.J., \& Radice, L.A. 1999, AJ in press (astro-ph/9904065)
1187: \bibitem[Rosen \& Bregman 1995]{RB95} Rosen, A., Bregman, J.N., 1995,\apj, 440, 634
1188: %\bibitem[Rosen, Bregman, \& Norman 1993]{RB93} Rosen, A., Bregman, J.N., Norman, M.L. 1993,\apj, 413, 137
1189: %\bibitem[Ryu et al. 1993]{RY} Ryu, et al., 1993,\apj, 414,1
1190: %\bibitem[Scalo et al. 1998]{SV} Scalo, J.,
1191: %V$\acute{\rm{a}}$zquez-Semadeni, E., Chappell, D., Passot, T., 1998,
1192: %ApJ, 504,835
1193: \bibitem[Scalo \& Chappell 1999]{SC99} Scalo, J., Chappell, D., 1999,
1194: ApJ 510, 258
1195: %\bibitem[Shlosman, Begelman, \& Frank 1990]{SBF} Shlosman, I.,
1196: %Begelman, M.C. \& Frank, J. 1990, Nature, 345, 679
1197: \bibitem[Seiden \& Gerola 1982]{SG82} Seiden, P. E. \&
1198: Gerola, H. 1982, Fundamentals of Cosmic Physics, 7, 241
1199: \bibitem[Shore \& Ferrini 1995]{SF95} Shore, S. N. \&
1200: Ferrini, F. 1995, Fundamentals of Cosmic Physics, 16, 1
1201: \bibitem[Spaans et al. 1994]{ST94} Spaans, M. , Tielens, A. G. G. M., Van Dishoeck, E. F. \& Bakes, E. L. O.
1202: 1994, \apj, 437, 270
1203: \bibitem[Spaans 1996]{SP96} Spaans, M. 1996, \aap, 307, 271
1204: \bibitem[Spaans \& Norman 1997]{SN} Spaans, M., Norman C., 1997, \apj,483,87
1205: \bibitem[Spaans \& Van Dishoeck 1997]{SV97} Spaans, M. \&
1206: Van Dishoeck, E. F. 1997, \aap, 323, 953
1207: \bibitem[Spitzer 1978]{SP97} Spitzer, L. 1978, John Wiley \& Sons, Inc., New York, p. 139
1208: \bibitem[Stanimirovic et al. 1999]{ST99} Stanimirovic,
1209: S., Staveley-Smith, L., Dickey, J. M., Sault, R. J. \& Snowden, S. L. 1999,
1210: \mnras, 302, 417
1211: \bibitem[Stone \& Norman 1992]{SN92} Stone, J., \& Norman, M.N. 1992, APJS, 80,753
1212: \bibitem[Staveley-Smith et al. 1997]{SS97} Staveley-Smith,
1213: L., Sault, R. J., Hatzidimitriou, D., Kesteven, M. J. \& MCConnell, D.
1214: 1997, \mnras, 289, 225
1215: \bibitem[Tenorio-Tagle 1988]{TT88} Tenorio-Tagle, G. 1988, ARA\&A, 26, 145
1216: %\bibitem[van Leer 1977]{VL} van Leer, B., 1977, J.Comp.Phys., 32,101
1217: %\bibitem[V$\acute{\rm{a}}$zquez-Semadeni 1994]{VZ94} V$\acute{\rm{a}}$zquez-Semadeni, E. 1994, ApJ, 423, 681
1218: \bibitem[V$\acute{\rm{a}}$zquez-Semadeni et al. 1995]{VZ95a}
1219: V$\acute{\rm{a}}$zquez-Semadeni, E., Passot, T. \& Pouquet, A. 1995, ApJ,
1220: 441, 702
1221: \bibitem[V$\acute{\rm{a}}$zquez-Semadeni, Passot, \& Pouquet 1996]{VZ96}
1222: V$\acute{\rm{a}}$zquez-Semadeni, E., Passot, T. \& Pouquet, A. 1996,
1223: ApJ, 473,881
1224: \bibitem[V$\acute{\rm{a}}$zquez-Semadeni et al. 1997]{VZ97}
1225: V$\acute{\rm{a}}$zquez-Semadeni, E., Balleseros-Paredes, J., \&
1226: Rodriguez, L.F., 1997, ApJ, 474, 292
1227: \bibitem[van der Hulst 1996]{VDH96} van der Hulst, J.M. 1996, in ASP
1228: Conf: Ser. Vol. 106: The Minnesota Lectures on Extragalactic Neutral
1229: Hydrogen, ed. E.D. Skillman (ASP: San Francisco), 47
1230: \bibitem[van Dishoeck \& Black 1988]{VB88} Van Dishoeck, E. F. \& Black, J. H. 1988, \apj, 334, 771
1231: \bibitem[Vallenari et al. 1996]{VA96} Vallenari, A., Chiosi, C., Bertelli, G. \& Ortolani, S. 1996, \aap, 309, 358
1232: %\bibitem[Wada 1994]{WA94} Wada, K., 1994, PASJ, 46, 165
1233: \bibitem[Wada \& Norman 1999]{WN99} Wada, K., \& Norman, C., 1999, ApJ,
1234: 516, L13
1235: %\bibitem[Wada, Sakamoto, \& Minezaki 1998]{WSM98} Wada, K., Sakamoto,
1236: %K., \& Minezaki, T. 1998, ApJ, 494, 236
1237: %\bibitem[Wada \& Habe 1992]{WH92} Wada, K., Habe, A., 1992, \mnras, 258,
1238: %82
1239: %\bibitem[Wada \& Habe 1995]{WH95} Wada, K., Habe, A., 1995, \mnras, 277,
1240: %433
1241: \bibitem[Walter \& Brinks 1999]{WB99} Walter, F. \& Brinks,
1242: E. 1999, \aj, 118, 273
1243: \bibitem[Woodward \& Colella 1984]{WC} Woodward, P. R., \& Colella,
1244: P. 1984, J.Comp.Phys., 54, 115
1245:
1246: \end{thebibliography}
1247:
1248: \end{document}
1249:
1250: