0804.0139/ms.tex
1: %\documentclass{emulateapj} \usepackage{apjfonts}
2: \documentclass[12pt, preprint]{aastex}
3: %
4: %===============================================================
5: 
6: \newcommand\cm{{\;\rm cm}}
7: \newcommand\pcc{{\;\rm cm}^{-3}}
8: \newcommand\Kel{{\;\rm K}}
9: \newcommand\cs{c_s}
10: \newcommand\dmax{\rho_{\rm max}}
11: \newcommand\dmid{\rho_{00}}
12: \newcommand\dz{\rho_0 (z)}
13: \newcommand\vA{v_{\rm A}}
14: \newcommand\yr{{\;\rm yr}}
15: \newcommand\Myr{{\;\rm Myr}}
16: \newcommand\Msun{{\;\rm\,M_\odot}}
17: \newcommand\Lsun{{\;\rm\,L_\odot}}
18: \newcommand\kms{{\;\rm km\,s^{-1}}}
19: \newcommand\mh{m_{\rm H}}
20: \newcommand\torb{t_{\rm orb} }
21: \newcommand\tcro{t_{\rm cross} }
22: \newcommand\pc{{\;\rm\,pc}}
23: \newcommand\kpc{{\;\rm kpc}}
24: \newcommand\Kcm{{\;\rm K\,cm^{-3}}}
25: \newcommand\simgt{\lower.5ex\hbox{$\; \buildrel > \over \sim \;$}}
26: \newcommand\simlt{\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}}
27: \newcommand\xhat{\hat{\mathbf{x}} }
28: \newcommand\yhat{\hat{\mathbf{y}} }
29: \newcommand\zhat{\hat{\mathbf{z}} }
30: \newcommand\Rhat{\hat{R}}
31: \newcommand\phihat{\hat{\phi}}
32: \newcommand\vel{\mathbf{v}}
33: \newcommand\ergs{{\rm \;erg\,s^{-1}}}
34: 
35: %===============================================================
36: \shorttitle{SPIRAL SHOCKS WITH THERMAL INSTABILITY}
37: \shortauthors{KIM, KIM, \& OSTRIKER}
38: %===============================================================
39: \begin{document}
40: 
41: \title{Galactic Spiral Shocks with Thermal Instability}
42: 
43: \author{Chang-Goo Kim\altaffilmark{1}, Woong-Tae Kim\altaffilmark{1},
44: and Eve C.\ Ostriker\altaffilmark{2}}
45: \affil{$^1$Department of Physics \& Astronomy, FPRD,
46: Seoul National University, Seoul 151-742, Republic of Korea}
47: \affil{$^2$Department of Astronomy, University of Maryland, 
48: College Park, MD 20742, USA}
49: \email{kimcg@astro.snu.ac.kr, wkim@astro.snu.ac.kr, ostriker@astro.umd.edu}
50: \slugcomment{Accepted for Publication in \apj}
51: %\slugcomment{Last Modified: \today}
52: 
53: \begin{abstract}
54: Using one-dimensional hydrodynamic simulations including 
55: interstellar heating, cooling, and thermal conduction,
56: we investigate nonlinear evolution of gas flow across galactic spiral
57: arms.  We model 
58: the gas as a non-self-gravitating, unmagnetized fluid, and follow
59: its interaction with a stellar spiral potential in a local frame 
60: comoving with the stellar pattern.
61: Initially uniform gas with density $n_0$ in the range 
62: $0.5\pcc\leq n_0 \leq 10\pcc$
63: rapidly separates into warm 
64: and cold phases as a result of thermal instability (TI), and also
65: forms a quasi-steady shock that prompts phase transitions.
66: After saturation, the flow follows a recurring cycle: 
67: warm and cold phases in the interarm region are 
68: shocked and immediately cool to become a denser cold medium in the
69: arm; post-shock expansion reduces the mean density
70: to the unstable regime in the transition zone and TI subsequently 
71: mediates evolution back into warm and cold interarm phases.
72: For our standard model with $n_0=2\pcc$, the gas resides in 
73: the dense arm, thermally-unstable transition zone, and interarm
74: region for 14\%, 22\%, 64\% of the arm-to-arm crossing time. 
75: These regions occupy 1\%, 16\%, and 83\% of the arm-to-arm distance,
76: respectively.  Gas at intermediate temperatures (i.e.\ neither warm
77: stable nor cold states) represents $\sim25$-30\% 
78: of the total mass, similar to the fractions estimated from \ion{H}{1}
79: observations (larger interarm distances could reduce this
80: mass fraction, whereas other physical processes associated with 
81: star formation could increase it).
82: Despite  transient features and multiphase structure, 
83: the time-averaged shock profiles 
84: can be matched to that of a diffusive isothermal medium
85: with temperature $1,000\Kel$ (which is $\ll T_{\rm warm}$) and 
86: ``particle'' mean free path of $l_0=100\pc$.
87: Finally, we quantify numerical conductivity associated with translational 
88: motion of phase-separated gas  on the grid, and show that 
89: convergence of numerical results requires the numerical conductivity 
90: to be comparable to or smaller than the physical conductivity.
91: \end{abstract}
92: \keywords{galaxies: ISM --- instabilities --- ISM: kinematics and dynamics
93: --- methods: numerical --- stars: formation}
94: 
95: \section{Introduction}
96: 
97: Spiral arms are the most prominent features in disk galaxies.
98: As the interstellar medium (ISM) passes through the moderate
99: gravitational potential 
100: well of the stellar spiral arms, it is strongly compressed
101: and shocked, producing narrow dust lanes in optical images.
102: Active star formation is subsequently triggered in high-density
103: clouds inside the arms, resulting in downstream optical arms that 
104: contain OB associations and giant \ion{H}{2} regions 
105: distributed in a ``beads on a string'' fashion 
106: (e.g., \citealt{baa63,elm83,2006ApJ...642..158E,she07}). 
107: Other arm substructures include filamentary gaseous spurs (or feathers) 
108: seen in optical extinction, IR emission from dust, and
109: H$\alpha$ emission from star formation
110: (e.g. 
111: \citealt{1980ApJ...242..528E,scorec01,sco01,ken04,2004ApJS..154..222W,
112: lav06,gor07}), 
113: and giant molecular associations and atomic superclouds seen 
114: in CO and \ion{H}{1} radio observations
115: (e.g., \citealt{elm83,vog88,ran90,kna93}).
116: The locations of these arm substructures downstream from the primary dust
117: lanes indicates that the shock compression represents the first 
118: step in an evolutionary sequence that begins with diffuse ISM gas and
119: ends with star formation (for strongly-bound cores) 
120: and dispersal (for more weakly self-gravitating structures),
121: although it is uncertain whether spiral arms actually enhance 
122: star formation rate or just organize it
123: (e.g., \citealt{ger78,elm86,sle96,sei02}).
124: 
125: Studies of galactic spiral shocks date back to \citet{rob69}, 
126: who used a semi-analytic approach to 
127: obtain one-dimensional, stationary shock profiles 
128: as functions of the distance perpendicular to the shocks
129: (see also \citealt{fuj68,rob70,shu73}).  
130: \citet{woo75} used time-dependent calculations to show that spiral
131: shocks in local models indeed develop within one or two crossings 
132: of 
133: the background arm potential.  This and subsequent work
134: (e.g. \citealt{kim02}) suggests spiral arm 
135: shocks in the one-dimensional approximation are highly stable
136: for a range of the arm strength.  On the other hand, spiral shocks
137: have been shown to be intrinsically unstable when the vertical dimension is 
138: included \citep{mar98,gom02,gom04,bol06,kim06}. 
139: Since the arm-to-arm crossing periods are in general incommensurable 
140: with the vertical oscillation periods, the gas streamlines are not closed,
141: giving rise to shock flapping motions that dump a significant amount of 
142: random kinetic energy in the gas  \citep{kko06}.  Under certain
143: (strong compression) conditions, two-dimensional {\it in-plane} 
144: spiral shocks can also become unstable due to strong shear within the arm
145: \citep{wad04,dob06}.  However, these in-plane modes are stabilized by
146: moderate magnetic fields \citep{she06,dob08}, and suppressed in fully
147: three-dimensional models due to vertical dynamics \citep{kim06}.
148: 
149: Inclusion of gaseous self-gravity tends to enhance the arm response and 
150: symmetrize the density profile \citep{lub86}, and causes the shock front
151: to move downstream relative to the minimum in the potential \citep{kim02}.  
152: High post-shock density enhances the growth of self-gravitating
153: perturbations 
154: within spiral arms, although postshock flow expansion can limit this growth 
155: \citep{bal85,bal88}.
156: Using two-dimensional simulations with both self-gravity and magnetic 
157: fields, 
158: \citet{kim02} demonstrated that magneto-Jeans instability (in which magnetic
159: tension forces counterbalance the stabilizing Coriolis forces) leads
160: to the formation of both arm spurs and GMAs/GMCs with realistic properties
161: (see also \citealt{lyn66,elm94}). Subsequent studies including
162: three-dimensional effects \citep{kim06} and global spiral structure 
163: \citep{she06} have confirmed these findings.
164: 
165: While recent work has improved our understanding of
166: galactic spiral shocks and their larger 
167: substructures, these studies have oversimplified
168: the ISM thermodynamics, usually adopting 
169: an isothermal equation of state.
170: This ignores  potential consequences of thermal instability (TI) 
171: (\citealt{fie65}; see also \citealt{mee96} for review), which
172:  changes an otherwise homogeneous ISM to clumpy, multi-phase gas. 
173: (e.g., \citealt{fie69,hei01,wol03}).  In the classical
174: two-phase picture of the ISM,  cold dense clouds are in pressure equilibrium
175: with warm intercloud media that surround them \citep{fie69}.
176: Supernovae lead to a hot, diffuse third phase
177: \citep{1974ApJ...189L.105C,mo77}, but because massive star formation
178: is spatially
179: correlated and much of the hot gas produced is vented away, most of
180: the volume remains relatively unaffected (e.g. 
181: \citealt{1998ApJ...503..700F,2004A&A...425..899D}).
182: Since the cold clouds and the warm intercloud gas 
183: differ in density and temperature
184: by about two orders of magnitudes, their respective responses to 
185: spiral shocks and downstream expansion flows will be much different 
186: from the isothermal case.
187: When realistic thermal processes are considered, for instance,
188: warm rarefied gas in the interarm region can be converted 
189: via shocks  to cold dense gas in the arm regions. 
190: The reverse phase transition can then occur downstream for some
191: fraction of the mass, yielding  a quasi-steady cyclic exchange.
192: 
193: \citet{shu72} were the first to study the effects of gas cooling and 
194: heating on spiral shocks.  By considering a mixture of the comoving two 
195: stable phases and allowing for phase transitions, they calculated 
196: steady-state shock profiles for both cold and warm phases.  However, 
197: they employed a pressure-density relation, instead of solving 
198: the time-dependent energy equation, based on the assumption of
199: instantaneous thermal equilibrium; this precluded the possibility 
200: of unstable-phase gas in their calculations.
201: 
202: More recent years have seen 
203: a few numerical studies of spiral shocks with explicit heating and cooling, but
204: most of these suffer from strong numerical diffusion.
205: \citet{bak74} argued that allowance 
206: for thermal phase changes produces ``accretion fronts/waves'' 
207: instead of spiral shocks, in which the inflowing material radiates
208: its energy away.  As they mentioned, however, 
209: this result could  be due to large numerical diffusion;
210: we will indeed show below that in a moving medium, 
211: numerical conductivity can be large enough to 
212: suppress TI. On the other hand,
213: \citet{tub80} and \citet{mar83} have shown that some models
214: develop spiral shocks in which phase transitions from warm interarm gas 
215: to cold arm clouds occur, although insufficient resolution in their models
216: made the cloud sizes and separations significantly overestimated 
217: and prevented the transition regions from cold to warm phases 
218: from being resolved.
219: Very recently, \citet{dob07} and \citet{dob08} studied 
220: the effects of the warm phase on a pre-existing cold phase using
221: particle simulations, but they did not allow for phase transitions 
222: that are crucial in spiral shocks with TI.
223: 
224: In this paper, we initiate a study of galactic spiral shocks 
225: subject to ISM heating and cooling, using 
226: very high-resolution numerical hydrodynamic simulations.
227: We consider one-dimensional models that represent slices
228: perpendicular to the arm.  
229: We ignore the gaseous self-gravity and magnetic fields here,
230: deferring studies of these effects to future work.  
231: Our primary objectives are to determine overall shock structures under TI,
232: to explore where and how the transitions among the cold, warm, and 
233: unstable phases occur, and to find statistical properties 
234: such as temperature distributions, mass fractions, and velocity dispersions. 
235: 
236: The remainder of this paper is organized as follows: 
237: in \S2 we describe the basic equations we solve and present 
238: our model parameters and numerical methods.
239: In \S3 we test our numerical code and quantify the
240: diffusion (due to translational motion over the grid)
241: in terms of a numerical conductivity.
242: In \S4 we address the evolution  of cold and warm 
243: phases as they traverse spiral shocks, and provide statistical
244: measures to quantify the exchange cycle that develops. 
245: Finally, we summarize our results and discuss their implications in \S5. 
246: 
247: \section{Numerical Methods}
248: 
249: \subsection{Basic Equations}
250: 
251: We study  galactic gas flows and thermodynamic evolution in response
252: to an external
253: stellar spiral potential, which is assumed to be tightly wound
254: with a pitch angle $i\ll1$, and rotating at a constant pattern speed $\Omega_p$
255: with respect to an inertial frame.
256: For local simulations, it is advantageous to
257: set up a frame corotating with the spiral pattern, 
258: centered at the position $(R, \phi) =(R_0, \Omega_p t)$.
259: The local frame is tilted by an angle $i$ relative to the radial direction 
260: in such a way that the two orthogonal axes correspond to the directions 
261: perpendicular ($x$-axis) and parallel ($y$-axis) to the local arm segment, 
262: respectively \citep[][]{rob69}.
263: We assume that all physical variables depend only on the $x$-coordinate,
264: while allowing nonzero velocity in the $y$-direction.
265: Since the independent variable in our local models is a projection of 
266: the position on a streamline in a large-scale flow onto the $x$-axis,
267: the distance on the $x$-axis divided by $\sin i$ represents the 
268: distance that the flow has traversed in the azimuthal direction
269: along the streamline.  Therefore, the {\it temporal} interval between 
270: one arm crossing and the next crossing is the same as it would be for a 
271: global model. 
272: 
273: In the local arm frame, the background velocity due to galactic rotation 
274: is approximately given by
275: \begin{equation}\label{eq:v0}
276:  \vel_0=R_0(\Omega_0-\Omega_p)\sin i\xhat
277: +[R_0(\Omega_0-\Omega_p)-q_0\Omega_0 x]\yhat,
278: \end{equation}
279: where $\Omega_0$ is the angular velocity of gas at $R_0$ in the inertial frame
280: and $q_0\equiv-(d\ln\Omega/d\ln R)|_{R_0}$ is the local shear rate 
281: in the background flow in the absence of the spiral potential 
282: \citep[][]{kim02,kim06}.
283: Assuming that the motions induced by the stellar potential are much smaller 
284: than $R_0\Omega_0$, the basic equations of ideal hydrodynamics expanded in
285: the local frame read
286: \begin{equation}\label{eq:cont}
287:  \frac{\partial\rho}{\partial t}+\nabla\cdot(\rho \vel_T)=0,
288: \end{equation}
289: \begin{equation}\label{eq:mom}
290: \frac{\partial\vel_T}{\partial t}+\vel_T\cdot\nabla\vel_T=
291: -\frac{1}{\rho}\nabla P -q_0\Omega_0 v_{0x}\yhat-2\mathbf{\Omega}_0\times
292: \vel-\nabla\Phi_{\rm ext},
293: \end{equation}
294: \begin{equation}\label{eq:energy}
295: \frac{\partial e}{\partial t} +\vel_T\cdot\nabla e =
296: -\frac{\gamma}{\gamma-1}P\nabla\cdot\vel_T -\rho\mathcal{L}+\nabla\cdot(\mathcal{K}\nabla T),
297: \end{equation}
298: \citep[see][]{rob69, shu73, bal88, kim02,pio04}, 
299: where $\vel_T\equiv \vel_0+\vel$ is the total velocity in the local frame,
300: $\Phi_{\rm ext}$ is the external stellar spiral potential, 
301: $\rho\mathcal{L}(\rho, T)$ is the net cooling function,
302: and $\mathcal{K}$ is the thermal conductivity. 
303: Other symbols have their usual meanings. 
304: We adopt an ideal gas law $P=(\gamma-1)e$ with $\gamma=5/3$.
305: 
306: For the stellar spiral potential, we consider a 
307: simple sinusoidal shape:
308: \begin{equation}\label{eq:ephi}
309:  \Phi_{\rm ext}=\Phi_{\rm sp}\cos\left(\frac{2\pi x}{L_x}\right),
310: \end{equation}
311: analogous to a
312: logarithmic potential of \citet{rob69} and \citet{shu73}.
313: In equation (\ref{eq:ephi}), $\Phi_{\rm sp}$ denotes the amplitude of the 
314: spiral potential, while $L_x =2\pi R_0\sin i/m$ 
315: is the arm-to-arm separation for an $m$-armed spiral.  
316: We take the size of the simulation domain equal to $L_x$;
317: since $x$ varies from $-L_x/2$ to $L_x/2$ and $\Phi_{\rm sp}<0$, 
318: $\Phi_{\rm ext}$ attains its minimum at the center ($x=0$).
319: We parametrize the spiral arm strength using 
320: \begin{equation}\label{eq:F}
321: F\equiv \frac{m}{\sin i}\left(\frac{|\Phi_{\rm sp}|}{R_0^2\Omega_0^2}\right),
322: \end{equation}
323: which measures the maximum force due to the spiral 
324: potential relative to the the mean axisymmetric gravitational force 
325: \citep{rob69}.  
326: 
327: The net cooling function per unit volume is given by 
328: $\rho\mathcal{L}\equiv n(n\Lambda[T]-\Gamma)$, 
329: where $n=\rho/(\mu m_{\rm H})$ is the gas number density 
330: and $\mu=1.27$ is the the mean molecular weight per particle.
331: For the heating and cooling rates of the atomic ISM, 
332: we take the fitting formulae
333: \begin{equation}\label{eq:heat}
334:  \Gamma = 2.0\times10^{-26}\ergs,
335: \end{equation}
336: \begin{equation}\label{eq:cool}
337:  \frac{\Lambda(T)}{\Gamma}=10^7\exp\left(\frac{-1.184\times10^5}{T+1000}\right)
338:  +1.4\times10^{-2}\sqrt{T}\exp\left(\frac{-92}{T}\right) \cm^3,
339: \end{equation}
340: suggested by \citet{koy02} (see also \citealt{vaz07}).  
341: Under the adopted net cooling curve, the minimum and maximum
342: pressures for the coexistence of the classical warm/cold phases 
343: in a static equilibrium are 
344: $P_{\rm min}/k_B=1600\Kcm$ and $P_{\rm max}/k_B=5000\Kcm$.
345: The corresponding transition temperatures
346: $T_{\rm max}=5012\Kel$ and $T_{\rm min}=185\Kel$  
347: define cold ($T<T_{\rm min}$), warm ($T>T_{\rm max}$),
348: and intermediate-temperature phases ($T_{\rm min}<T<T_{\rm max}$).
349: To resolve the length scales of TI numerically 
350: (e.g., \citealt{koy04,pio04}), 
351: we include a constant value of thermal conductivity 
352: $\mathcal{K}_0=10^5\ergs\cm^{-1}{\, \rm K}^{-1}$.\footnote{While thermal 
353: conductivity is proportional to $T^{1/2}$ for neutral hydrogen at 
354: kinetic temperatures below $4.5\times 10^4\Kel$  \citep{par53},
355: we choose for simplicity a fixed value corresponding to 
356: thermal equilibrium at $T=1500\Kel$ for our standard density $n_0=2\pcc$.} 
357: The associated Field length below which thermal conduction 
358: erases temperature perturbations completely is defined by
359: \begin{equation}\label{eq:field}
360:  \lambda_F=2\pi\left\{\frac{\rho^2\Lambda}{\mathcal{K}_0T}
361:  \left[1-\left(\frac{\partial \ln \Lambda}{\partial \ln T}
362:  \right)\right]\right\}^{-1/2}
363: \end{equation}
364: \citep{fie65}.  In our models, $\lambda_F$ typically 
365: amounts to $\sim 0.18\pc$. 
366: 
367: \subsection{Model Parameters \& Numerical Methods}
368: 
369: We consider a simulation box in which the gas is initially homogeneous 
370: with density $n_0$ and pressure
371: $P_0$ when a spiral perturbation is absent.
372: Other than thermal processes involving TI, overall dynamics and
373: structures of spiral shocks in our models are completely characterized by 
374: the arm-to-arm distance $L_x$ and the flow speed $v_{0x}$ relative to 
375: the perturbing stellar potential in 
376: the $x$-direction (as well as $\sin i$, $q_0$, $F$, and $\Omega_p/\Omega_0$).
377: The azimuthal wavenumber $m$ of spiral arms is arbitrary,  and
378: the box location and the gaseous angular speed 
379: relative to the arms can then be specified as
380: $R_0=mL_x/(2\pi\sin i)$ and 
381: $\Omega_0-\Omega_p=2\pi v_{0x}/(mL_x)$, respectively.
382: To achieve the numerical resolution sufficient to resolve the Field length, 
383: we consider a small box with $L_x=628$ pc. 
384: For the relative velocity, we choose $v_{0x}=13\kms$  corresponding to 
385: the rotational velocity of $R_0\Omega_0=260\kms$ with a flat rotation curve 
386: ($q_0=1$).  
387: The corresponding arm-to-arm crossing time is $\tcro\equiv L_x/v_{0x} = 
388: 4.7\times10^7\yr$, which we choose as the fiducial time unit in our 
389: presentation.  
390: For spiral arm parameters, we take pattern speed $\Omega_p=\Omega_0/2$,
391: pitch angle $\sin i=0.1$, and strength $F=5\%$ in all the models. 
392: We note that driven by numerical requirement, 
393: spiral arms in our models have a small separation and thus a short 
394: dynamical time, so that some of our numerical results 
395: (e.g., mass fractions) that depend 
396: on the ratio of cooling time to dynamical time may not be 
397: applicable to spiral arms with a much larger arm-to-arm crossing time.
398: 
399: To simulate spiral shocks with varying total gas contents, 
400: we consider 12 models that have the same initial thermal pressure 
401: $P_0/k_B$ but differ in the initial density $n_0$;  
402: we adopt the solar neighborhood value at $P_0/k_B=3000\Kcm$ 
403: based on the observational (e.g., \citealt{fer01,hei03,jen07}) and theoretical 
404: (e.g., \citealt{wol03}) arguments.
405: Note that although some models are out of thermal equilibrium initially,
406: they immediately tend towards equilibrium owing to rapid
407: heating and cooling, with the equilibrium pressure
408: depending on $n_0$. 
409: Table~\ref{tbl:model} lists the model parameters and simulation outcomes.
410: Column (1) labels each run; while the models with the prefix SU
411: rapidly undergo TI even with $F=0$, the
412: SW and SC models would stay warm or cold throughout
413: were it not for the spiral perturbations. 
414: Column (2) lists $n_0$.  Columns (3)--(8) give the width
415: of, and time spent in, the arm, transition, and interarm regions,
416: respectively,  in each model. 
417: The mass and volume fractions of the cold, warm, and intermediate-temperature 
418: phases are given in columns (9)-(14), respectively.
419: We take model SU2 with $n_0=2\pcc$ as our fiducial model; at this density,
420: the total surface density would be $\sim 12\Msun\pc^{-2}$
421: for vertical scale height of $\sim100\pc$.
422: 
423: We integrate the time-dependent partial differential equations 
424: (\ref{eq:cont})--(\ref{eq:energy})
425: using a modified version of the Athena code \citep{gar05}. 
426: Athena implements a single step, directionally unsplit Godunov scheme for 
427: compressible hydrodynamics in multi-spatial dimensions and
428: allows a variety of spatial reconstruction methods and approximate Riemann 
429: solvers;
430: the version we use here employs the piecewise linear method (PLM) with 
431: the Roe Riemann solver.\footnote{
432: Although the piecewise parabolic method (PPM) is known to provide 
433: in general less diffusive spatial reconstruction than the PLM, 
434: our experiments have shown that the PPM often produced negative 
435: internal energy in the presence of strong radiative cooling
436: inside spiral shocks.} 
437: We implement the shearing-periodic boundary condition at the 
438: $x$-boundaries \citep{haw95}.
439: Because of the very short cooling time, energy updates from net 
440: cooling are made implicitly based on Newton-Raphson iteration,
441: while the conduction term is solved fully explicitly. 
442: For stable and accurate results, we ensure the time step is kept 
443: smaller than the CFL condition for thermal conduction as well as the 
444: cooling/heating time (see \citealt{pio04}).
445: Our standard models employ $N=16,384$ zones, corresponding to 
446: grid spacing of $\Delta x = 0.04\pc$, which satisfies the condition
447: $\Delta x < \lambda_F/3$ for convergence of numerical results 
448: \citep{koy04}; we also ran models with different
449: grid sizes in order to study the effects of numerical resolution. 
450: 
451: \section{Code Test and Numerical Conductivity\label{sec:numc}}
452: 
453: The Athena code we use has been verified on a wide variety of test 
454: problems including hydrodynamic shock tubes, advection of a square box
455: in a background shear flow, and one-dimensional propagation of sound waves
456: in a rotating, shearing medium.
457: For simulations involving cooling/heating and conduction terms, 
458: it is crucial to check if
459: the numerical scheme employed can resolve the length and times scales of 
460: the fastest growing TI modes.  This test is of particular importance
461: for the current work since the gas in our models is non-static, moving 
462: in the $x$-direction at an average speed of $v_{0x}=13\kms$, so that
463:  numerical diffusion associated with advection and zone averaging 
464: may reduce the growth rates of TI at small scales.
465: In this section, we describe the test results 
466: of our numerical code on the development of TI 
467: in static, rotating, and moving media 
468: in the absence of spiral potential perturbations, and 
469: provide a quantitative measure of the
470: numerical diffusion in terms of numerical conductivity.
471: 
472: \subsection{Linear Dispersion Relation}
473: 
474: We test our code by comparing the numerical growth rate of a 
475: particular TI mode with the corresponding analytic prediction.  
476: To this end, we derive a dispersion relation for local, axisymmetric TI 
477: in a homogeneous medium that is rotating, shearing, and undergoing uniform 
478: translational motion with $\vel_{0}$.  
479: We linearize equations (\ref{eq:cont})--(\ref{eq:energy}) 
480: (dropping the external potential), assuming 
481: plane-wave disturbances $\propto  e^{nt - i k x}$, 
482: where $n$ and $k$ are the growth rate and
483: wavenumber of the disturbances in the $x$-direction, respectively.
484: We follow the same steps as in \citet{fie65}, except that we include
485: the non-zero background motions and the indirect forces arising from 
486: galaxy rotation.  The resulting dispersion relation for local disturbances
487: is given by
488: \begin{eqnarray}\label{eq:disp}
489:   \tilde n^3   
490: + \tilde n^2 a\left(k_T+\frac{k^2}{k_\mathcal{K}}\right) 
491: + \tilde n(\kappa^2+k^2 a^2) 
492: + a\left[\frac{a^2k^2}{\gamma}\left(k_T-k_\rho+\frac{k^2}{k_\mathcal{K}}\right)
493:    +\kappa^2\left(k_T+\frac{k^2}{k_\mathcal{K}}\right)\right]=0,
494: \end{eqnarray} 
495: where 
496: $\tilde n \equiv n - i kv_{0x}$ is the Doppler-shifted growth rate,
497: $a\equiv(\gamma k_B T/\mu m_H)^{1/2}$ is the adiabatic speed of sound,
498: $\kappa\equiv( R^{-3}d(R^4\Omega^2)/dR|_{R_0})^{1/2}=(4-2q_0)^{1/2}\Omega_0$
499: is the local epicyclic frequency,
500: and $k_T, k_\rho, k_\mathcal{K}$ are the wavenumbers defined by
501: \begin{eqnarray}
502: k_T=\frac{\gamma(\gamma-1)}{a^3} 
503:  \left(\frac{\partial \mathcal{L}}{\partial\ln T}\right)_\rho, \;\;\; 
504: k_\rho=\frac{\gamma(\gamma-1)}{a^3} 
505:  \left(\frac{\partial \mathcal{L}}{\partial\ln \rho}\right)_T, \;\;\;
506: k_\mathcal{K} = \frac{a^3\rho}{\gamma(\gamma-1)\mathcal{K} T},
507: \end{eqnarray}
508: (e.g., \citealt{fie65}).
509: In the limit of $q_0=0$ and $v_{0x}=0$, 
510: equation (\ref{eq:disp}) recovers the dispersion relation given by 
511: \citet{fie65} for TI in a rigidly-rotating medium with
512: no translational motion.  The fact that $v_{0x}$ occurs in the
513: dispersion relation only through $\tilde n$ implies that 
514: a constant translational motion does not change 
515: the growth rates of TI if measured in a frame moving with $v_{0x}$,
516: analogous to Galilean invariance in mechanics.
517: 
518: For isobaric TI to occur, the term in the square 
519: brackets in equation (\ref{eq:disp}) must be negative. 
520: This of course necessitates $k_T - k_\rho < 0$, the 
521: Field criterion for isobaric TI \citep{fie65}.
522: Thermal conduction and rotation suppress short and long wavelength 
523: perturbations against TI, respectively, reducing the unstable 
524: range of wavelengths to $k_1 < k < k_2$, where
525: $k_1$ and $k_2$ are two positive roots of 
526: $k^4 + (k_\mathcal{K}[k_T - k_\rho] + \gamma \kappa^2/a^2) k^2
527: + \gamma \kappa^2k_Tk_\mathcal{K}/a^2=0$.
528: One can show that 
529: $k_2 \rightarrow (k_\mathcal{K}[k_\rho - k_T])^{1/2}=2\pi/\lambda_F$
530: for weakly- or non-rotating systems ($\kappa \rightarrow 0$), 
531: while $k_1^2 \rightarrow \gamma \kappa^2 k_T /(a^2[k_\rho - k_T])$
532: in the limit of vanishingly small conductivity 
533: ($k_\mathcal{K} \rightarrow \infty$).
534: Figure~\ref{fig:grate} plots as various lines sample growth rates 
535: $\tilde n$ of TI calculated from equation (\ref{eq:disp})
536: for cases $\mathcal{K}=\Omega_0=0$ (\textit{dotted line}),
537: $\mathcal{K}=\mathcal{K}_0$ with $\Omega_0=0$ (\textit{solid line}),
538: and $\mathcal{K}=\mathcal{K}_0$  with $\Omega_0=130\kms\kpc^{-1}$ 
539: (\textit{dashed line}).
540: The density, pressure, and shear parameter are taken to be $n_0=2\pcc$, 
541: $P_0/k_B=3000\pcc$, and $q_0=1$, corresponding to our fiducial model SU2.
542: Stabilization of TI by rotation and conduction are apparent at large
543: and small scales, causing the wavelengths of the most unstable 
544: disturbances to occur at $\lambda\sim3\pc$ in between the
545: cut-off wavelengths.
546: 
547: \subsection{Numerical Conductivity} 
548: 
549: For our code tests, we consider three kinds of models depending
550: on $\Omega_0$ and $\Omega_p$:
551: (1) a static disk ($\Omega_0=\Omega_p=0$);
552: (2) a rotating disk without translational motion 
553: ($\Omega_0=\Omega_p=130\kms\kpc^{-1}$, yielding $v_{0x}=0$
554: from eq.\ [\ref{eq:v0}]) 
555: (3) a rotating disk with translational motion 
556: ($\Omega_0=2\Omega_p=130\kms\kpc^{-1}$, so that $v_{0x}=13\kms$).
557: The same conductivity $\mathcal{K}=\mathcal{K}_0$ is adopted
558: for all the models.
559: Other parameters including $n_0$, $P_0$, and $q_0$ are 
560: the same as in model SU2.
561: In each model, we initialize an eigenmode of TI with wavelength $\lambda$ 
562: and vary the box size to fit it in, while keeping the grid
563: spacing $\Delta x=0.04\pc$ fixed.  We monitor the evolution of the maximum
564: density and measure its growth rate numerically in the linear regime.
565: Figure~\ref{fig:grate} plots the resulting growth rates from 
566: the runs as cross, diamond, and plus symbols for the first,
567: second, and third types of models, respectively.
568: Evidently, the numerical and analytic results are in good agreement for 
569: models without the translational motion, confirming the performance of our
570: implementation of the heating/cooling and conduction terms.
571: While the translational motion of the gas at a level of $v_{0x}=13\kms$ 
572: does not affect large-scale modes much,
573: it significantly reduces the numerical growth rates for small-scale modes 
574: with $\lambda < 2\pc$.
575: The discrepancy between the numerical and analytic results 
576: appear to be due to numerical diffusion that causes the thermal energy 
577: to spread out as cooling/heating regions are advected 
578: with the background flow.
579: 
580: To quantify the strength of numerical heat diffusion in our code, 
581: we solve the linear dispersion relation (\ref{eq:disp}) 
582: for a given set of parameters to find the effective conductivity 
583: $\mathcal{K}_{\rm eff}$ that yields an analytic growth rate equal 
584: to the numerical value obtained from the model simulation with 
585: the same parameters except for the 
586: physical conductivity $\mathcal{K}=\mathcal{K}_0$.
587: We then calculate the numerical conductivity from
588: $\mathcal{K}_n=\mathcal{K}_{\rm eff}-\mathcal{K}$.
589: In the case of models with $v_{0x}=13\kms$ 
590: and $\lambda=2.56$ and 1.28 pc shown 
591: in Figure~\ref{fig:grate}, for instance,
592: $\mathcal{K}_n=3.4\times10^{5}$ and
593: $1.2\times10^{6}\ergs\cm^{-1}\Kel^{-1}$, 
594: respectively,
595: about 3.4 and 12 times larger than $\mathcal{K}_0$.
596: 
597: In order to find the parametric dependences of $\mathcal{K}_n$ on 
598: the translational velocity, grid size, and perturbation wavelength,
599: we ran a suite of models varying $v_{0x}$ from 
600: 2.6 to $26\kms$, $\Delta x$ from 0.02 to $0.16\pc$,
601: and $\lambda$ from 0.32 to $10.2\pc$.  
602: We also explored the cases with different levels of physical conductivity at
603: $\mathcal{K}=0$, $\mathcal{K}_0$, or $4\mathcal{K}_0$.
604: The numerical conductivity is calculated only when 
605: the numerical growth rate is lower than the analytic value at
606: the same $\mathcal{K}$ by more than 1\%.
607: Figure~\ref{fig:knum} plots as various symbols the resulting numerical
608: conductivity based on the PLM scheme for spatial reconstruction,
609: showing that $\mathcal{K}_n$ is well fitted by
610: \begin{equation}\label{eq:numk}
611:  \mathcal{K}_n=10^{9.4}\left(\frac{v_{0x}}{1\kms}\right)
612:                        \left(\frac{\Delta x}{1\pc}\right)^3
613:                        \left(\frac{\lambda}{1\pc}\right)^{-2}
614: \ergs\cm^{-1}{\, \rm K}^{-1}.
615: \end{equation}
616: This suggests that the numerical conductivity in a moving medium 
617: depends rather sensitively on the numerical resolution. 
618: Equation (\ref{eq:numk}) states that our models with $\Delta x=0.04\pc$
619: have numerical conductivity comparable to the physical conductivity
620: for the fastest growing modes,
621: so that the effect of numerical diffusion on the results presented
622: in this paper is not significant.
623: 
624: We repeated the same calculations using the higher-order PPM
625: reconstruction scheme, and found
626: the numerical conductivity is linearly proportional to 
627: $v_{0x}\Delta x^{4}\lambda^{-3}$.
628: Generalizing these results,  we rewrite equation (\ref{eq:numk}) as
629: $\mathcal{K}_n\propto v_{0x}\Delta x(\Delta x/ \lambda)^{p}$, 
630: where $p$ is the order of the spatial reconstruction scheme 
631: ($p=2$ and $3$ for PLM and PPM, respectively).  This implies that 
632: the numerical conductivity can be viewed as 
633: the diffusion coefficient ($\propto v_{0x}\Delta x$) modified by the
634: accuracy of the interpolation scheme used in spatial reconstruction 
635: ($\propto[\Delta x/\lambda]^{p}$).
636: 
637: \section{Nonlinear Simulations}
638: 
639: \subsection{Standard Model\label{sec:evol}}
640: 
641: We now study the nonlinear evolution of thermally-unstable gas flows 
642: under an imposed spiral potential.  In this subsection, we 
643: focus on model SU2 with $n_0=2\pcc$.  We initially apply density
644: perturbations created by a Gaussian random field with a
645: power spectrum $|\rho_k|^2 \propto k^{-5/3}$ for 
646: $1 \leq 2\pi k/L_x \leq 128$ 
647: and zero power for $2\pi k/L_x>128$ in the Fourier space, 
648: corresponding to a one-dimensional Kolmogorov spectrum for sonic disturbances. 
649: The standard deviation of the density perturbations is fixed to 
650: be 1\% in physical space.
651: In order to suppress rapid motions of the gas 
652: caused by an abrupt introduction of the spiral potential,
653: we slowly turn it on, reaching the full level $F=5\%$ at $t/\tcro\sim2.4$.
654: 
655: Figure~\ref{fig:tevol} shows density distributions of model SU2 
656: at $t/\tcro=0$, 1, 2.4, and 4.0 together with scatter plots of pressure 
657: versus density overlaid on the equilibrium cooling curve.
658: The gas initially has $T=1500\Kel$ and is thus thermally unstable
659: (Fig.\ \ref{fig:tevol}a).  It evolves rapidly to 
660: cold ($n>n_{\rm max}=8.6\pcc$ and $T<T_{\rm min}=185\Kel$) and 
661: warm ($n<n_{\rm min}=1\pcc$ and $T>T_{\rm max}=5012\Kel$)
662: phases within $0.5\tcro$.  The TI soon saturates, 
663: and random cloud motions aided by epicyclic shaking
664: cause cold clouds to collide and merge together, or split sometimes, 
665: resulting in, on average, 70 cold clouds with a mean cloud separation 
666: of $\sim 0.8\pc$, which are   
667: in rough pressure equilibrium with the surrounding
668: warm gas at $P/k_B=1900\Kcm$ (Fig.\ \ref{fig:tevol}b).  
669: At this time, the spiral potential remains weak and  
670: most gas in the unstable temperature range corresponds to the boundaries 
671: of the cold clouds.  
672: As the amplitude of the potential grows, the gas is gathered 
673: toward the potential minimum, forming a spiral shock near $x=0$. 
674: The shock reaches maximum strength at around $t/\tcro=2.5$ 
675: shortly after $F$ attains the full strength.
676: Both cold and warm phases in the interarm regions continually enter the shock 
677: front and are compressed to become cold gas with higher density.  
678: They subsequently 
679: expand and become thermally unstable as they leave the spiral arm regions, 
680: returning back to the cold and warm interarm phases (see below). 
681: At about $t/\tcro=3$, the overall shock structure reaches a quasi-steady 
682: state in the sense that the mass and volume fractions of each phase do
683: not change appreciably with time, although the shock oscillates slightly
684: around an equilibrium position and cold clouds shift as they follow 
685: galaxy rotation (projected onto the $\hat x$ direction). 
686: 
687: Figure~\ref{fig:sat} plots the distributions
688: of physical variables in model SU2 at $t/\tcro=4$ after this quasi-steady
689: state has been reached; many spikes and discontinuities in density as well as 
690: sawtooth-like velocity profiles are evident.
691: Figure~\ref{fig:schematic} schematically illustrates the evolutionary
692: tracks of the cold and warm phases in the $n$--$P$ plane.
693: Regions marked with A, B, C, and D correspond to interarm, 
694: immediate post-shock, spiral arm, and thermally-unstable 
695: transition zones, respectively;
696: the transition zone refers to the window downstream
697: from the arm (between $x/L_x\sim0.03-0.2$) in Figure \ref{fig:sat}b, 
698: where most of the gas has temperatures in the unstable range between
699: $T_{\rm min}$ and $T_{\rm max}$.
700: The subscripts 1 and 2 denote the warm and cold phases, respectively,
701: of interarm gas, and their respective immediate post-shock counterparts.
702: 
703: In the interarm regions (either $x/L_x<0$ or $x/L_x>0.2$
704: in Fig.\ \ref{fig:sat}), 
705: warm (A$_1$) and cold (A$_2$) phases are in rough equilibrium 
706: in terms of the total (= thermal + ram) pressure.
707: While the cold interarm clouds typically have slightly lower thermal
708: pressure than the warm interarm medium, their ram pressure is of
709: comparable magnitude, i.e. $P/k_B\sim 2000-4000\Kcm$. 
710: Some of the cold interarm 
711: clouds exhibit ``double-horned'' 
712: structure, a consequence of merging with neighbors or splitting into
713: two pieces (see e.g., Fig.\ \ref{fig:sat}a).
714: As the warm and cold interarm phases enter the shock front, they experience 
715: a strong compression, jumping to B$_1$ and B$_2$, respectively.
716: While the density jump is by a factor of $4$ for both cold and warm phases, 
717: the pressure jump for the cold phase is about 10 times larger than
718: in the warm phase since the former is about $100$ times colder.\footnote{
719: Strictly speaking, this holds true only for one-dimensional shocks.
720: In two or three dimensions, small clouds can experience enhanced 
721: compression as the shock wrapping around them is able to propagate 
722: towards the cloud centers. They may also be subject to dynamical effects 
723: such as Kelvin-Helmholtz and Rayleigh-Taylor instabilities 
724: (e.g., \citealt{woo76}).}
725: The shocked gas in states B$_1$ and B$_2$ is not in thermal balance
726: and subsequently undergoes strong post-shock cooling
727: (e.g., \citealt{muf74}), moving almost 
728: isobarically to C$_1$ and C$_2$. The transition from A to C is
729: essentially instantaneous (cooling time $\sim 10^3$ yrs).\footnote{
730: Because of its very large post-shock density, the transition 
731: from A$_2$ to C$_2$ occurs over an extremely short cooling scale
732: ($\sim 0.02$ pc) that is not resolved in our models.}
733: 
734: One of the characteristics of galactic spiral shocks is that gas
735: accelerates after the maximum shock compression 
736: (e.g., \citealt{rob69,bal88,kim02}).
737: Because of this post-shock expansion, the cold gas inside the arm 
738: ($0<x/L_x<0.03$) 
739: becomes progressively less dense, evolving from C to D.  
740: Since the dynamical time scale is much longer than the cooling time 
741: (due to the reduced velocity inside the arm), the cold gas either
742: at C$_1$ or C$_2$ moves all the way down to the transition zone D 
743:  following  the equilibrium curve in the $n$--$P$ plane.  
744: When the expanding gas reaches region D ($0.03<x/L_x<0.2$), it
745: becomes thermally unstable and turns back into either the warm (A$_1$) 
746: or cold (A$_2$) interarm phase.
747: For model SU2, it typically takes about
748: $\sim 0.14\tcro$ from C to D stages and $\sim 0.22\tcro$ from D to A stages.
749: Gas is in the interarm regime A for the balance of the cycle
750: ($\sim0.64\tcro$).
751: When averaged over $t/\tcro=5-8$, 
752: the arm, transition, and interarm regions in model SU2 occupy 
753: approximately 1, 16, 83\% of the spatial domain, respectively. 
754: 
755: Our models, as well as the real ISM, contain many interarm
756: clouds which would be too small to be detected individually in 
757: extra-galactic radio observations.
758: In modeling gaseous  spiral shocks,
759: \citet{lub86} did not explicitly solve for the cloudy structure.
760: Instead, they adopted an isothermal equation of state and treated 
761: the effects of cold clouds (and their
762: collisions) by including a viscous term parameterized by a mean
763: free path $l_0$ for the fluid.
764: They found that viscosity renders arm profiles smoother 
765: and more symmetric. 
766: In order to compare our models 
767: with a  single-phase viscous counterpart, 
768: we take a temporal average over $t/\tcro=5-8$ of the density 
769: distributions in model SU2. We then take a boxcar average 
770: of the time-averaged profile, with a window of $8$ pc.
771: Figure~\ref{fig:vis} plots the resulting mean density profile 
772: as a solid line.  For comparison, we have run isothermal models that 
773: include an explicit viscosity term in the momentum equation (\ref{eq:mom}) 
774: in a manner similar to in \citet{lub86}.  Selected results are shown 
775: in Figure~\ref{fig:vis}.
776: Evidently, larger viscosity tends to produce 
777: a weaker spiral shock, with a peak that is more symmetric and 
778: farther downstream.
779: In terms of the strength and placement of the spiral arm, 
780: an isothermal model with $T=1000\Kel$ and $l_0=100\pc$ provides a
781: fairly good match for the mean density profile when TI is included. 
782: The isothermal viscous model, however, has a slightly broader arm
783: compared to the time average of the multiphase model.
784: Our results thus demonstrate that (for the purposes of obtaining a
785: lower-resolution ``beam averaged'' profile) the effects of clouds 
786: embedded in a warm medium can be effectively modeled by viscosity
787: \citep{cow80,gam96}, but only if the medium's 
788: temperature and the mean free path are chosen appropriately.
789: We note that if a temperature comparable to that of the warm medium
790: were adopted for an isothermal counterpart, the shock would generally
791: be too weak.
792: 
793: \subsection{Effects of Initial Number Density\label{sec:ndep}}
794: 
795: All the SU models we consider start from a thermally-unstable initial state.
796: Since the TI growth time is short compared to the time over which we
797: turn on the spiral potential, these models 
798: first evolve into a thermally-bistable
799: state before developing shocks.  
800: Figure~\ref{fig:high_n} plots a density profile 
801: at $t/\tcro=4$ in model SU5 with $n_0=5\pcc$.
802: Comparing to model SU2 as shown in Figure~\ref{fig:tevol},
803: Figure~\ref{fig:high_n} shows that the densities (and temperatures) 
804: of the cold and warm phases after TI saturates
805: are insensitive to the initial gas density, 
806: although of course 
807: models with higher $n_0$ produce more  cold clumps.\footnote{ 
808: Mass conservation requires the number of cold clouds $N_c$ to be 
809: given approximately by $N_c\approx n_0 L_x/(ln_c)$, 
810: where $l$ and $n_c$ are the mean size and number density of the clouds. 
811: From simulations with differing $n_0$, we
812: empirically found that $l\sim 0.8\pc\;(n_0/2\cm^{-3})^{0.3}$,
813: i.e. the mean cloud size (presumably set by merging
814: and splitting), depends only 
815: weakly on $n_0$.  This gives $N_c \propto n_0^{0.7}$.}
816: The ensuing development of spiral shocks and evolutionary tracks in
817: the $n$--$P$ plane  are 
818: also qualitatively similar to those described in the previous subsection.
819: 
820: Perhaps the most notable difference in the late-time states
821: with different $n_0$ is the size of the postshock transition zone.
822: Columns (3)-(8) of Table~\ref{tbl:model} show that 
823: the transition zone (and also the arm region, to a lesser extent)
824: widens with increasing $n_0$.  One can easily see this by 
825: comparing Figures~\ref{fig:sat}d and \ref{fig:high_n}.
826: This trend is of course because models with larger $n_0$
827: have a larger fraction of the total mass in the cold phase that
828: evaporates to maintain an unstable state.
829: In addition, the mean separation of cold clouds inside the arm is smaller in 
830: models with higher $n_0$, allowing for
831: merging with neighboring clouds more often during the postshock
832: expansion stage. This reduces the expansion rate effectively, 
833: thereby extending the size of zones with unstable density
834: toward downstream.  
835: 
836: Unlike the SU models, 
837: the SW and SC models initially have a low- or high-enough density
838: that they are thermally stable and the 
839:  early development of a spiral shock takes place in a 
840: single-phase medium.  Nevertheless, the shock compression and/or
841: post-shock expansion can still produce thermally unstable gas 
842: if the initial density is not far from the unstable range. 
843: For example, the shock in model SW0.5 drives 
844: the thermal pressure  above $P_{\rm max}$, and the gas evolves rapidly
845: via TI into  a cold stable phase
846: inside the arm. 
847: In model SC10, on the other hand, strong post-shock expansion 
848: drives the pressure of the initially cold gas below $P_{\rm min}$,
849: after which a 
850: fraction of the gas expands to become a warm phase. 
851: Consequently, the resulting structures at late times in models 
852: SW0.5 and SC10 are similar to those in the SU models.
853: In models with $n_0$ further away from the unstable values, however, 
854: changes in the gas pressure due to shock compression and  
855: postshock expansion are insufficient to trigger phase transitions. 
856: Figure~\ref{fig:single} plots equilibrium shock profiles in 
857: models SW0.1 and SC20, which start with $n_0=0.1$ and $20\pcc$, respectively.
858: The gas  in both of these models consists only of the warm or cold phase
859: and the profiles are smooth everywhere except at the shock front.
860: 
861: \subsection{Temperature Distribution}\label{sec:temp}
862: 
863: Figure~\ref{fig:pdf} plots the volume-weighted and mass-weighted 
864: temperature probability distribution functions (PDFs) 
865: for models SW0.5, SU2, SU5, and SC10, averaged over $t/\tcro=5-8$. 
866: The vertical dotted lines in each panel indicate 
867: $T_{\rm min}$ and $T_{\rm max}$, 
868: marking the cold, intermediate-temperature, and warm phases.
869: For the standard model SU2,
870: the mass-weighed temperature PDF is characterized by 
871: a broad cold peak at $T\sim 20-150\Kel$ and a narrow warm peak 
872: at $T\sim 6000-8000\Kel$, although there also exists a substantial 
873: amount of the intermediate-temperature gas. 
874: In the broad cold peak, the portion of gas with $T\simlt 50\Kel$
875: corresponds to the very dense arm population immediately behind the shock,
876: while the portion with $50\Kel \simlt T < T_{\rm min}$ 
877: includes both arm and interarm cold clouds.
878: Since a larger initial density implies a larger initial cold fraction
879: even before
880: the spiral potential is applied, the cold mass
881: fractions in both arm and interarm regions increase in models with 
882: larger $n_0$. 
883: For model SW0.5, on the other hand,  
884: introduction of the spiral potential allows only a small fraction 
885: of the gas to exceed $P_{\rm min}$ and undergo TI,
886: resulting in a lower cold peak and a higher warm peak than in model SU2.
887: For all the models, the cold peak is fractionally much broader
888: (i.e. larger $\Delta T/T$) than the warm peak.
889: 
890: Of the total intermediate-temperature phase, about $70\%$ is found in the 
891: post-shock transition zone for models with TI,
892: while the remaining portion resides in boundary layers between the cold and 
893: warm phases in the interarm region.
894: Figure~\ref{fig:pdf} shows that the
895: distribution of the intermediate phase is almost flat for model
896: SU2 and increasingly favors lower temperature as $n_0$ increases.  
897: This is because models with larger $n_0$ have a slower post-shock expansion 
898: rate, and thus more gas near  $T_{\rm min}$.
899: The mean density and density-weighted mean temperature of the 
900: intermediate-temperature gas are found to scale as 
901: $n_i \approx 0.41 n_0 + 1.29$ and $T_i \approx 4458/(n_0 + 1.81)\Kel$
902: for $0.5 \simlt n_0 \simlt 10$, where $n_i$ and $n_0$ are in units 
903: of $\pcc$.  For the cold and warm phases, the mean values 
904: are $n_c\approx23\pcc, T_c\approx86\Kel$, and $n_w\approx0.39\pcc, 
905: T_w\approx6400\Kel$,
906: almost independent of $n_0$. These points lie slightly above $P_{\rm min}$
907: on the cold and warm branches of the thermal equilibrium curve.
908: 
909: \subsection{Mass and Volume Fractions\label{sec:mf}}
910: 
911: For models with  thermal instability,  we calculate the mass and volume 
912: fractions of three phases averaged over $t/\tcro=5-8$.  
913: Figure~\ref{fig:conv} plots against numerical resolution the mean 
914: mass fractions in model SU2, along with the standard deviations as errorbars. 
915: While the warm mass fraction is fairly insensitive to the number of
916: grid points $N$, the cold (intermediate-temperature) 
917: mass fraction increases (decreases) with increasing resolution for 
918: $N \simlt 10^4$.  This is mainly because of the numerical conductivity 
919: associated with a large zone size for small $N$, 
920: as explained in \S\ref{sec:numc}.  When numerical conductivity is large,
921: the boundary layers in between the cold and warm phases thicken in 
922: proportion to the Field length.  This increase in the
923: intermediate-temperature mass (in the interarm) occurs primarily 
924: at the expense of the cold phase \citep{beg90,fer93,ino06}.
925: In addition,  large numerical conductivity tends to suppress 
926: TI in the post-shock transition zone, yielding less cold gas. 
927: 
928: Overall, the broadening of interarm cloud interfaces in low resolution models
929: is responsible for 80\% of the increase in the intermediate-temperature mass,
930: while the remaining 20\% is due to the suppression of TI in the post-shock
931: transition zone.  There is no interarm cold gas in models with $N\simlt 512$,  
932: while arm cold gas still exists in these lowest resolution models since
933: the post-shock regions are, when fully resolved, denser and broader than 
934: interarm clouds, and thus are less affected by zone averaging.
935: As long as $N \simgt 10^4$, 
936: the numerical conductivity becomes comparable to, or
937: smaller than, the physical conductivity, and the results are
938: numerically well resolved.
939: Because most of the difference between moderate and high resolution
940: models is at warm/cold interfaces, the extremely high resolutions 
941: that we find are needed for accurate measurement of the
942: thermally-unstable mass fraction would not be required for studies
943: that focus primarily on shock dynamics.
944: We find that the overall shock structure (in terms of the breadth of
945: the transition zone, and the time-averaged profiles) are comparable
946: to those in the converged models provided $N\simgt 10^3$.  At these
947: moderate resolutions, edges of individual clouds are smeared out, but the
948: physically-important transition zone downstream from the arm is still
949: recovered. 
950: 
951: In the case of model SU2, the converged values of the mass fractions
952: are 59\%, 14\%, and 27\% for the cold, warm, and 
953: intermediate-temperature phases, respectively.  
954: By volume the cold takes up 4\% of the total, 
955: while the warm and intermediate-temperature media occupy 70\% and 26\%. 
956: Since these proportions are 80\%, 10\%, and 10\% by mass
957: and 10\%, 80\%, and 10\% by volume when the spiral potential is weak
958: or absent (e.g., Fig.\ \ref{fig:tevol}b), this
959: implies that spiral shocks and subsequent post-shock expansion 
960: zone are important for populating the 
961: intermediate-temperature portion of the phase plane.
962: 
963: Columns (9)--(14) of Table~\ref{tbl:model} give the converged values 
964: of the mass and volume fractions of each phase in models 
965: with multiple phases.  
966: Figure~\ref{fig:mf_ndep} plots these proportions as functions of the initial 
967: number density $n_0$.  Interestingly,
968: the fraction of mass in the intermediate-temperature phase 
969: has a substantial value of 
970: $f_i \approx 0.28$, almost independent of $n_0$.\footnote{
971: By running models with differing $F$ (not listed in Table~\ref{tbl:model}),
972: we found that $f_i\approx 0.040 + 0.045 F(\%)$ 
973: for $3\%\leq F \leq 7\%$, but is insensitive to $n_0$ for the fixed 
974: arm parameters.}
975: The volume fraction of the intermediate phase increases with $n_0 \simlt 6\pcc$ 
976: and becomes flat at large $n_0$.  
977: Assuming that the cold, warm, and intermediate-temperature phases 
978: in each model are represented by the characteristic densities 
979: $n_c$, $n_w$, and $n_i$, respectively, mass conservation requires the mass 
980: fractions in the cold and warm phases to be 
981: $f_c = (1-n_w/n_0) - (1-n_w/n_i)f_i$ and
982: $f_w = n_w/n_0 - (n_w/n_i)f_i$, respectively,
983: where $n_c/n_w \gg 1$ is assumed (cf.\ \citealt{pio04}).
984: Dotted lines in Figure~\ref{fig:mf_ndep} plot these theoretical 
985: $f_c$ and $f_w$ using the empirical results $f_i=0.28$, $n_w=0.39\pcc$, and   
986: $n_i = 0.41 n_0 + 1.29\pcc$. These estimates are overall 
987: in good agreement with the simulation results.
988: 
989: \subsection{Velocity Dispersions\label{sec:veld}}
990: 
991: Finally, we quantify the level of random gas motions driven in our models
992: due to TI and spiral potential perturbations.
993: Spiral arms produce gas streaming  motions that are ordered but vary
994: perpendicular to the shock front.  Since streaming velocities are much
995: larger in amplitude than the true random motions, care is needed in
996: measuring the latter.  In addition, while the overall shock profiles reach
997: a quasi-steady state after $t/\tcro=5$, several other effects make it
998: difficult to measure turbulent amplitudes in multiphase models:
999: randomly-placed cold clouds move with varying 
1000: speeds in the interarm region, new structures are continuously developing 
1001: in the transition zone, and the spiral shocks themselves
1002: undergo small-amplitude oscillations perpendicular to the 
1003: shock front. 
1004: To separate out the background streaming from the total velocity 
1005: as cleanly as possible, 
1006: we first construct velocity template profiles $\langle v_x \rangle$ and 
1007: $\langle v_y \rangle$ for each model, where 
1008: the angle brackets denote a time average over
1009: $t/\tcro=5-8$.
1010: Dashed lines in Figure~\ref{fig:sat}d,e show sample
1011: $\langle v_x \rangle$ and $\langle v_y \rangle$ profiles in model SU2.
1012: We then calculate the density-weighted velocity dispersions using
1013: $\sigma_i^2\equiv \int \rho(v_i-\langle v_i\rangle)^2dx /\int\rho dx$
1014: (with $i=x$ or $y$) in the arm, interarm, and postshock transition zones
1015: separately.
1016: Figure~\ref{fig:veld_tevol} plots $\sigma_i(t)$ for model SU2,
1017: while Figure~\ref{fig:veld_ndep} draws
1018: the mean values $\bar\sigma_i \equiv \langle \sigma_i^2 \rangle^{1/2}$ 
1019: along with the standard deviations $\Delta \sigma_i$ 
1020: for models with multi-phase spiral shocks.
1021: 
1022: Figure~\ref{fig:veld_tevol} shows that the velocity dispersions of
1023: the cold phase in the arm region exhibit large-amplitude temporal 
1024: fluctuations, with characteristic periods of $\sim 0.4\tcro$ and $\sim0.2\tcro$
1025: for $\sigma_x$ and $\sigma_y$, respectively.  
1026: These variations of $\sigma_x$ and $\sigma_y$ are due to 
1027: small-amplitude oscillations offsetting the spiral  shock front 
1028:  from its  mean position. 
1029: Since streaming varies strongly within the arm (Fig. \ref{fig:sat}d,e), 
1030: the offset of the shock position leads to large differences between
1031: instantaneous and mean streaming velocities within the arm; these
1032: contaminate the measured velocity dispersion at locations near the
1033: shock front.  
1034: During the quasi-periodic oscillations of the spiral shock front,
1035: $\sigma_x$ attains its minimum value when the shock is maximally displaced
1036: upstream, while
1037: $\sigma_y$ is smallest when the instantaneous shock 
1038: front coincides with its mean position. 
1039: Therefore, the local minima in the time series of 
1040: $\sigma_x$ and $\sigma_y$ correspond to 
1041: the upper limits to the level of random gas motions inside the arm.
1042: For model SU2, the average values of the local minima in 
1043: $\sigma_x(t)$ and $\sigma_y(t)$ are $\sim 2$ and $\sim 3\kms$, respectively,
1044: which roughly equals  $\bar\sigma_i - \Delta\sigma_i$. As 
1045: Figure~\ref{fig:veld_ndep} shows, $\bar\sigma_i - \Delta\sigma_i$ 
1046: does not vary much with $n_0$, suggesting that the true velocity
1047: dispersion within the arm is relatively independent of mean density. 
1048: 
1049: Unlike their behavior in the arm region, the time-averaged velocity 
1050: profiles in the interarm and transition zones are relatively smooth, so that 
1051: the velocity dispersions as calculated 
1052: in these regions are relatively  free of
1053: streaming motions. As Figures~\ref{fig:veld_tevol} and 
1054: \ref{fig:veld_ndep} show, the velocity dispersions in the interarm and 
1055: transition zones are similar, 
1056: amounting to $\bar\sigma_x\sim1.3\kms,\bar\sigma_y\sim0.8\kms$ 
1057: for model SU2, and decreasing slowly with increasing $n_0$.
1058: The ratios of the velocity dispersions in the $x$- to $y$-directions
1059: for all regions are consistent with predictions from epicyclic analysis
1060: (e.g., \citealt{bin87}),
1061: $\bar\sigma_y^2/\bar\sigma_x^2 = \kappa^2/(4\Omega_0^2) = 1 - q/2$.  Here,
1062: the local shear rate
1063: $q\equiv 2 - (2-q_0)n/n_0$ is modified from the background value 
1064: $q_0\equiv -d\ln\Omega_0/d\ln R$, due to 
1065: the constraint of potential vorticity conservation 
1066: \citep{gam96,gam01,kim01,kim02}. 
1067: For a flat rotation curve ($q_0=1$), 
1068: local shear is reversed inside the arm where the local density exceeds
1069: $2n_0$; this in turn increases 
1070: $\bar\sigma_y /\bar\sigma_x \approx (n/2n_0)^{1/2}$ above $1/\sqrt{2}=0.7$. 
1071: On the other hand, the whole interarm region and most of the unstable 
1072: region have $n<n_0$, and thus the ratio 
1073: $\bar\sigma_y /\bar\sigma_x$ is smaller than the prediction for a disk
1074: without spiral structure.
1075: 
1076: \section{Summary \& Discussion}
1077: 
1078: \subsection{Summary}
1079: 
1080: Galactic spiral shocks in disk galaxies play an important role in 
1081: structural and chemical evolution  by forming 
1082: spiral-arm substructures and triggering star formation. 
1083: Spiral shocks inherently involve large variations 
1084: in the background density, 
1085: while cooling and heating processes that determine the ISM density and
1086: temperature 
1087: depend rather sensitively on the local state.  The interplay between
1088: these processes
1089: may significantly alter
1090: the shock strengths and structures, 
1091: compared to those computed under an isothermal approximation. 
1092: In particular,  large-scale compressions and expansions
1093: across spiral arms may trigger
1094: TI, thereby 
1095: regulating transitions among the different ISM phases.
1096: In this paper, we have used high-resolution
1097: numerical simulations to
1098: investigate the dynamics and thermodynamics of this highly nonlinear
1099: process.  Our models include heating and cooling terms appropriate for
1100: atomic gas 
1101: explicitly in the energy equation, and thus naturally 
1102: allow for transitions among cold, warm, and intermediate-temperature phases. 
1103: The current investigation employs a one-dimensional model in which
1104: all the physical quantities vary only in the in-plane 
1105: direction perpendicular to
1106: a local segment of a spiral arm. We allow for 
1107: gas motions parallel to the arm, and include 
1108: galactic differential 
1109: rotation.  The effects of magnetic fields and self-gravity are neglected
1110: in the present paper.  
1111: 
1112: Our main results are summarized as follows:
1113: 
1114: 1. Background flow over the grid, represented by $v_{0x}$ in our
1115: models, may result in a significant level of numerical diffusion  
1116: in  finite-difference schemes dealing with the cooling/heating and conduction 
1117: terms if the numerical resolution is not high enough.
1118: We quantify the numerical diffusion of our implementation 
1119: in terms of an effective numerical conductivity, and measure it using
1120: growing modes of the thermal instability.  By comparing growth rates
1121: with predictions from the linear dispersion relation including conduction,
1122: we find that the numerical diffusion in the Athena code we use behaves 
1123: as $\mathcal{K}_n\propto (v_{0x}\Delta x)(\Delta x/\lambda)^{p}$, 
1124: where $\Delta x$ is the grid width, $p$ is the order of 
1125: spatial reconstruction method ($p=2$ for PLM and 3 for PPM), and
1126: $\lambda$ is the spatial wavelength.
1127: For typical values of
1128: $v_{0x}=13\kms$, $\Delta x=0.04\pc$, and $\lambda = 3.5\pc$
1129: in our models, the numerical conductivity
1130: with the PLM amounts to $\mathcal{K}_n 
1131: \sim 2.3\times 10^{5}\ergs\cm^{-1}\Kel^{-1}$, 
1132: comparable to the physical conductivity 
1133: $\mathcal{K}_0 =10^5 \ergs\cm^{-1}\Kel^{-1}$
1134: adopted in the current work.
1135: 
1136: 
1137: 2. Stellar spiral potential perturbations induce shocks that reach a
1138: quasi-steady state after a few orbits.  The resulting flow contains
1139: phase transitions provided the mean gas density is in the range 
1140: $0.5\pcc \leq n_0 \leq 10 \pcc$, for the parameters considered 
1141: in the present work.  Models with $n_0 \leq 0.1\pcc$ or $n_0 \geq 20\pcc$
1142: yield single-phase spiral shock profiles.
1143: We divide the flow into three distinct zones based on thermal regime: 
1144: arm, interarm, and transition.
1145: The ``arm'' refers to the highly-compressed postshock region filled with 
1146: cold gas at $T<T_{\rm min}=185\Kel$, while in the ``interarm'' 
1147: region far from the shock most of the volume is occupied 
1148: by warm gas with $T>T_{\rm max}=5012\Kel$.  The ``transition'' zone
1149: corresponds to the expanding region downstream from the arm where  
1150: intermediate-temperature gas ($T_{\rm min} <T<T_{\rm max}$) undergoes TI.
1151: Figure~\ref{fig:schematic} summarizes the evolutionary cycle in the 
1152: density--pressure plane: the warm/cold interarm gas (A$_1$/A$_2$) is
1153: shocked (B$_1$/B$_2$) and immediately cools 
1154: to become the denser cold arm gas (C$_1$/C$_2$); this 
1155: subsequently enters the unstable transition zone (D)  
1156: and evolves back into the warm and cold interarm phases. 
1157: For our standard model SU2 with mean density $n_0=2\pcc$,
1158: the duration of the arm, transition, and interarm stages are 
1159: approximately 14\%, 22\%, and 64\% of the cycle, 
1160: respectively, 
1161: occupying roughly 1\%, 16\%, and 83\% of the simulation domain.
1162: 
1163: 3.  At late times, instantaneous profiles in models with TI 
1164: show many  density spikes representing 
1165: cold clouds. Time-averaged profiles
1166: are, however, relatively smooth, and the density peaks representing
1167: the arm are more symmetric than in non-diffusive, isothermal models.
1168: We find that a viscous isothermal model with 
1169: $T=1,000\Kel$ and mean free path of $l_0=100\pc$ yields a similar peak
1170: density and arm width to the average profile from 
1171: model SU2.  This confirms the notion that for purposes where detailed
1172: ISM knowledge is not needed, multiphase effects 
1173: on shocks can be approximately treated via a viscosity 
1174: modeling excursions and collisions of dense clouds, as suggested by 
1175: \citet{cow80} and \citet{lub86}.  Simulations such as those we have
1176: performed are needed in order to calibrate the viscous/isothermal
1177: model parameters, however.
1178: 
1179: 4.  For models with multi-phase spiral shocks, 
1180:  intermediate-temperature gas amounts to 
1181: $\sim 0.25-0.3$ of the total by mass, insensitive to $n_0$.  
1182: Of this, about 70\% is found in the transition zone, 
1183: while the remaining 30\% lies at interfaces between the cold and
1184: warm media.  This suggests that the postshock expanding flows, an
1185: inherent feature of galactic spiral structure, is 
1186: important for producing intermediate-temperature gas.  
1187: The mean density and density-weighted mean temperature of the 
1188: intermediate-temperature phase are found to 
1189: be $n_i=0.41 n_0+1.29\pcc$ and $T_i = 4458/((n_0/1\pcc) + 1.81)\Kel$.
1190: The fractions of the cold and warm phases 
1191: are 59\% and 14\% by mass and 4\% and 70\% by volume for
1192: model SU2, respectively, and vary with $n_0$ according to simple
1193: expectations based on mass conservation with a prescribed density in
1194: each phase.
1195: 
1196: 
1197: 5. We find that one-dimensional spiral shocks with multi-phase gas 
1198: produces non-negligible random
1199: gas motions.  At late times, the gas in both interarm and 
1200: transition zones has 
1201: typical 
1202: density-weighted velocity dispersions of 
1203: $\sigma_x\sim1.3\kms$ and $\sigma_y\sim0.8\kms$  in the directions
1204: perpendicular and parallel to the spiral arms, respectively. 
1205: This is trans-sonic with respect to the cold medium, and subsonic with
1206: respect to the warm medium.
1207: The cold gas in the arms is estimated to 
1208: have slightly larger values $\sigma_x\sim 2\kms$ and 
1209: $\sigma_y\sim 3\kms$, although true turbulence levels may be slightly
1210: lower, because it is difficult to fully subtract streaming motions 
1211: in the arm region.
1212: 
1213: 
1214: \subsection{Discussion}
1215: 
1216: Since the growth rates of pure TI are maximized at the smallest available
1217: wavelengths, growth of grid-scale noise dominates numerical
1218: simulations if it is not suppressed.  To prevent this numerical problem,
1219: it is customary to include thermal conduction that 
1220: preferentially stabilizes small-scale perturbations 
1221: (e.g., \citealt{pio04,pio05}), and also broadens the transition
1222: layer between cold and warm phases (e.g., \citealt{beg90}). 
1223: \citet{koy04} studied TI in an initially static medium
1224: and showed that numerical results converge only if these conductive
1225: interfaces  are well resolved.
1226: They found this requires the cell size to be less than one third of 
1227: the Field length.  
1228: In modeling  systems with large turbulent 
1229: (e.g., \citealt{gaz05}) or ordered (this work) velocity 
1230: flows over the grid, numerical conductivity may be 
1231: considerable if resolution is inadequate.
1232: The convergence study presented in \S\ref{sec:mf} 
1233: suggests that the convergence criterion of \citet{koy04} is
1234: valid in a moving medium, too, with the condition that 
1235: the grid spacing must be smaller than the {\it effective}
1236: Field length based on the total (physical $+$ numerical) 
1237: conductivity.
1238: 
1239: In earlier work,
1240: \citet{tub80} and \citet{mar83} performed time-dependent simulations of
1241:  galactic spiral shocks that included cooling and heating, 
1242: but their models appear to suffer from large numerical
1243: conductivity associated with insufficient resolution. 
1244: While they showed that phase transitions from warm gas to cold clouds
1245: do occur in some models at the shock locations, they were unable to 
1246: resolve the post-shock transition zone.  In particular, the 
1247: $n_0=0.5\pcc$ model shown in Figure 7 of \citet{mar83} exhibits 
1248: a smooth density profile without any indication of TI although much of 
1249: the region is occupied by gas with density and temperature 
1250: in the unstable range.\footnote{The heating and cooling functions adopted by
1251: \citet{mar83} have the critical densities demarcating the warm, 
1252: intermediate-temperature,
1253: and cold phases about 20 times smaller than those in the present work. 
1254: Thus, their model with $n_0=0.5\pcc$ corresponds to our model SC10, which
1255: develops multi-phase structure.} 
1256: This is presumably because large numerical conductivity 
1257: (due to low resolution) renders
1258: the time and/or length scales of TI 
1259: in their simulations 
1260: longer than the duration and/or width of the post-shock transition zone.
1261: Indeed, we find that when we run our own models at 
1262: low resolution (e.g. with 512 zones or less for model 
1263: SC10), the numerical conductivity is
1264: large enough that 
1265: the gas in the transition zone does not undergo TI, and instead smoothly
1266: converts to interarm warm phase.
1267:  
1268: In terms of their time-averaged properties, the overall dynamics 
1269: and flow characteristics of spiral shocks with TI remain similar 
1270: to those of isothermal spiral shocks (e.g., \citealt{rob69,shu73,woo75,
1271: kim02}).  The salient features of spiral shocks with thermal 
1272: evolution  include 
1273: the facts that they allow phase transitions at the shock front 
1274: (from warm to cold)
1275: and in the post-shock transition zone (from cold to warm), 
1276: and that there are
1277: cold clumps in the interarm region.  Given the
1278: very high density and pressure within the arm, 
1279: cold gas downstream from the shock front
1280: would transform to molecular clouds if self-gravity and chemical reactions 
1281: for molecule formation were included.
1282: By running SPH simulations with separate 
1283: cold and warm ``particles'', 
1284: \citet{dob07} indeed found that the density of the cold component inside 
1285: arms is sufficient to form molecules
1286: (see also \citealt{dob08}).  These authors did not allow for phase 
1287: transitions between the cold and warm components, however, which 
1288: is an essential aspect of the evolution (cf. \citealt{shu72}).
1289: 
1290: A post-shock transition zone, where the gas is predominantly in
1291: the intermediate-temperature range, is a necessary feature of any
1292: quasi-steady state. 
1293: About 40\% of the atomic 
1294: gas in the local Milky Way is observed to be CNM
1295: \citep{hei01,hei03}, which is consistent with our results for mean
1296: density $n_0=1\pcc$, similar to the local Milky Way value.\footnote{The
1297: value of $n_0$ that produces the numerical results consistent with 
1298: observations may differ in models with different $L_x$ since the 
1299: fractional size of the transition region
1300: is likely to 
1301: depend on the ratio of cooling time to arm-to-arm crossing time.}
1302: \citet{hei03} also find that about half of the remaining atomic gas is
1303: consistent with being thermally-stable WNM, while the balance is in
1304: the unstable temperature range 500-5000K.  These fractions are also
1305: comparable to what we find for $n_0=1\pcc$.  The presence of
1306: thermally-unstable gas has been interpreted as owing to 
1307: dynamical effects which act on timescales comparable to the heating
1308: and cooling times.  Dynamical processes that have been investigated include
1309: magnetorotational instability \citep{pio04,pio05,pio07}, colliding flows 
1310: \citep{aud05,hei05,hei06,vaz06}, energy injection from OB stars
1311: \citep{vaz00,gaz01}, and supernova explosions \citep{avi05,mac05}.
1312: Here, we have shown that even without these additional small-scale 
1313: energy sources, a significant amount of intermediate-temperature 
1314: gas forms as a natural product of large-scale spiral shocks with TI,
1315: primarily in the expanding region downstream from the arm.
1316: 
1317: 
1318: Many numerical studies of TI have shown  that turbulent amplitudes driven
1319: by ``pure TI'' {\it alone} are quite small.  For example, 
1320: \citet{pio04} showed that pure TI in an initially static, uniform medium 
1321: in thermal equilibrium 
1322: produces only a modest level of random gas motions, 
1323: $\sigma \sim  0.2-0.3\kms$.  
1324: \citet{kri02a,kri02b} similarly found that TI of gas starting from
1325: thermal equilibrium, or with time-dependent heating, results in 
1326: only subsonic turbulence, $\sigma \sim0.2\kms$ when the cold gas dominates.
1327: The velocity dispersions $\sigma\sim1.5\kms$ of the gas found 
1328: in the interarm and transition zones in our one-dimensional models 
1329: are significantly larger than those from pure TI.  
1330: These velocities are much lower than those
1331: we found in previous (isothermal) simulations \citep{kim06,kko06} 
1332: in which the vertical shock structure is resolved, however.
1333: Evidently, shock flapping  offers a more effective means of converting
1334: galactic rotation to turbulence than simple in-plane oscillations. 
1335: Work is currently underway to see how TI interacts with
1336: vertically-resolved spiral shocks -- and differential buoyancy of cold
1337: and warm gas -- to generate turbulence and structure
1338: in the ISM.
1339: 
1340: 
1341: \acknowledgments
1342: We thank A.\ Kritsuk for drawing our attention to earlier work of
1343: Marochnik et al. on spiral shocks with thermal instability.
1344: This work was supported by the Korea Research Foundation Grant funded
1345: by the Korean Government (MOEHRD) (KRF -- 2007 -- 313 -- C00328).  The work of 
1346: E.~C.~O on this project was supported by the U.~S.\ National Science
1347: Foundation under grant AST0507315.
1348: The numerical computations presented in this work were performed on the
1349: Linux cluster at KASI (Korea Astronomy and Space Science Institute)
1350: built with funding from KASI and ARCSEC.
1351: 
1352: 
1353: \begin{thebibliography}{}
1354: \bibitem[de Avillez \& Breitschwerdt(2004)]{2004A&A...425..899D} 
1355:  de Avillez, M.~A., \& Breitschwerdt, D.\ 2004, \aap, 425, 899
1356: \bibitem[Avillez \& Breitschwerdt(2005)]{avi05}
1357:    de Avillez, M.\ A., \& Breitschwerdt, D.\ 2005, \aap, 436, 585
1358: \bibitem[Audit \& Hennebelle(2005)]{aud05}
1359:    Audit, E., \& Hennebelle, P.\ 2005, \aap, 433, 1
1360: \bibitem[Baade(1963)]{baa63}
1361:    Baade, W.\ 1963, in The Evolution of Stars and Galaxies,
1362:    ed.\ C.\ Payne-Gaposchkin (Cambridge: Harvard Univ.\ Press), 218
1363: \bibitem[Balbus(1988)]{bal88}
1364:    Balbus, S.\ A.\ 1988, \apj, 324, 60
1365: \bibitem[Balbus \& Cowie(1985)]{bal85}
1366:    Balbus, S.\ A., \& Cowie, L.\ L. 1985, \apj, 297, 61
1367: \bibitem[Baker \& Barker(1974)]{bak74}
1368:    Baker, P.\ L., \& Barker, P.\ K.\ 1974, \aap, 36, 179
1369: \bibitem[Begelman \& McKee(1990)]{beg90}
1370:    Begelman, M.\ C., \& McKee, C.\ F.\ 1990, \apj, 358, 375
1371: \bibitem[Binney \& Tremaine(1987)]{bin87}
1372:    Binney, J., \& Tremaine, S.\ 1987,
1373:    Galactic Dynamics (Princeton: Princeton Univ.\ Press)
1374: \bibitem[Boley \& Durisen(2006)]{bol06}
1375:    Boley, A.~C., \& Durisen, R.~H.\ 2006, \apj, 641, 534
1376: \bibitem[Cowie(1980)]{cow80}
1377:    Cowie, L.\ L.\ 1980, \apj, 236, 868
1378: \bibitem[Cox \& Smith(1974)]{1974ApJ...189L.105C} Cox, D.~P., \& Smith, 
1379: B.~W.\ 1974, \apjl, 189, L105 
1380: \bibitem[Dobbs \& Bonnell(2006)]{dob06}
1381:    Dobbs, C.\ L., \& Bonnell, I.\ A.\ 2006, \mnras, 367, 873
1382: \bibitem[Dobbs \& Bonnell(2007)]{dob07}
1383:    Dobbs, C.\ L., \& Bonnell, I.\ A.\ 2007, \mnras, 376, 1747
1384: \bibitem[Dobbs \& Price(2008)]{dob08}
1385:    Dobbs, C.\ L., \& Price, D.\ J.\ 2008, \mnras, in press; astro-ph/710.3558
1386: \bibitem[Dobbs et al.(2006)]{dbp06}
1387:    Dobbs, C.\ L., Bonnell, I.\ A., \& Pringle, J.\ E.\ 2006, \mnras, 371, 1663
1388: \bibitem[Elmegreen(1994)]{elm94} Elmegreen, B.\ G.\ 1994, \apj, 433, 39
1389: \bibitem[Elmegreen \& Elmegreen(1983)]{elm83} 
1390:    Elmegreen, B.\ G., \& Elmegreen, D.\ M.\ 1983, \mnras, 203, 31
1391: \bibitem[Elmegreen \& Elmegreen(1986)]{elm86}
1392:    Elmegreen, B.\ G., \& Elmegreen, D.\ M.\ 1986, \apj, 311, 554
1393: \bibitem[Elmegreen(1980)]{1980ApJ...242..528E} Elmegreen, D.~M.\ 1980, 
1394: \apj, 242, 528 
1395: \bibitem[Elmegreen et al.(2006)]{2006ApJ...642..158E} Elmegreen, D.~M., 
1396: Elmegreen, B.~G., Kaufman, M., Sheth, K., Struck, C., Thomasson, M., \& 
1397: Brinks, E.\ 2006, \apj, 642, 158 
1398: \bibitem[Ferriere(1998)]{1998ApJ...503..700F} Ferriere, K.\ 1998, \apj, 
1399: 503, 700
1400: \bibitem[Ferriere(2001)]{fer01} Ferriere, K.\ 2001, Rev. Mod.\ Phys., 2001,
1401:   73, 1031
1402: \bibitem[Ferrara \& Shchekinov(1993)]{fer93}
1403:    Ferrara, A., \& Shchekinov, Y.\ 1993, \apj, 417, 595
1404: \bibitem[Field(1965)]{fie65}
1405:     Field, G.\ B.\ 1965, \apj, 142, 531
1406: \bibitem[Field et al.(1969)]{fie69}
1407:     Field, G.\ B., Goldsmith, D.\ W., \& Habing, H.\ J.\ 1969, \apj, 155, L149
1408: \bibitem[Fujimoto(1968)]{fuj68}
1409:     Fujimoto, M.\ 1968, in IAU Symp.\ No.\ 29, 453 
1410: \bibitem[Gammie(1996)]{gam96} Gammie, C.\ F.\ 1996, \apj, 462, 725
1411: \bibitem[Gammie(2001)]{gam01} Gammie, C.\ F.\ 2001, \apj, 553, 174
1412: \bibitem[Gardiner \& Stone(2005)]{gar05}
1413:     Gardiner, T.\ A., \& Stone, J.\ M.\ 2005, J.\ Comp.\ Phys., 205, 509
1414: \bibitem[Gazol et al.(2001)]{gaz01}
1415:     Gazol, A., V\'{a}zquez-Semadeni, E., S\'{a}nchez-Salcedo, F.\ J., \& Scalo, J.\ 2001, \apj, 557, L121
1416: \bibitem[Gazol et al.(2005)]{gaz05}
1417:     Gazol, A., V\'{a}zquez-Semadeni, E., \& Kim, J.\ 2005, \apj, 630, 911
1418: \bibitem[Gerola \& Seiden(1978)]{ger78}
1419:     Gerola, H., \& Seiden, P.\ E.\ 1978, \apj, 223, 129
1420: \bibitem[G\'{o}mez \& Cox(2002)]{gom02}
1421:     G\'{o}mez, G.\ C., \& Cox, D.\ P. 2002, \apj, 580, 235
1422: \bibitem[G\'{o}mez \& Cox(2004)]{gom04}
1423:     G\'{o}mez, G.\ C., \& Cox, D.\ P. 2004, \apj, 615, 744
1424: \bibitem[Gordon(2007)]{gor07} Gordon, K. (2007), 
1425: {\it Spitzer} press release at \\
1426: {\tt 
1427: http://gallery.spitzer.caltech.edu/Imagegallery/image.php?image\underline{\ }name=sig07-025
1428: }
1429: \bibitem[Hawley, Gammie, \& Balbus(1995)]{haw95}
1430:    Hawley, J.\ F., Gammie, C.\ F., \& Balbus, S.\ A.\ 1995, \apj, 440, 742
1431: \bibitem[Heiles(2001)]{hei01}
1432:    Heiles, C.\ 2001, \apj, 551, L105
1433: \bibitem[Heiles \& Troland(2003)]{hei03}
1434:    Heiles, C., \& Troland, T.\ H.\  2003, \apj, 586, 1067
1435: \bibitem[Heitsch et al.(2005)]{hei05}
1436:    Heitsch, F., Burkert, A., Hartmann, L.\ W., Slyz, A.\ D., \& Devriendt, 
1437:    J.\ E.\ G.\ 2005, \apj, 633, L113
1438: \bibitem[Heitsch et al.(2006)]{hei06}
1439:    Heitsch, F., Slyz, A.\ D., Devriendt, J.\ E.\ G., Hartmann, L.\ W., 
1440:    \& Burkert, A.\ 2006, \apj, 648, 1052
1441: \bibitem[Inoue et al.(2006)]{ino06}
1442:    Inoue,  T., Inutsuka, S., \& Koyama, H.\ 2006, \apj, 652, 1331
1443: \bibitem[Jenkins \& Tripp(2007)]{jen07}
1444:    Jenkins, E. B., \& Tripp, T.\ M.\ 2007,
1445:    IAU Symposium: Triggered Star Formation in a Turbulent ISM, 
1446:    eds.\ B.\ G.\ Elmegreen \& J.\ Palous, 237, 53
1447: \bibitem[Kennicutt(2004)]{ken04}
1448:    Kennicutt, R.\ C.\  2004, Spitzer press release at  \\
1449:    {\tt http://www.spitzer.caltech.edu/Media/releases/ssc2004-19/ssc2004-19a.shtml}
1450: \bibitem[Kim, Kim, \& Ostriker(2006)]{kko06}
1451:     Kim, C.-G., Kim, W.-T., \& Ostriker, E.\ C.\ 2006, \apj, 649, L13
1452: \bibitem[Kim \& Ostriker(2001)]{kim01}
1453:     Kim, W.-T., \& Ostriker, E.\ C.\ 2001, \apj, 559, 70 
1454: \bibitem[Kim \& Ostriker(2002)]{kim02}
1455:     Kim, W.-T., \& Ostriker, E.\ C.\ 2002, \apj, 570, 132
1456: \bibitem[Kim \& Ostriker(2006)]{kim06}
1457:     Kim, W.-T., \& Ostriker, E.\ C.\ 2006, \apj, 646, 213
1458: \bibitem[Knapen et al.(1993)]{kna93}
1459:    Knapen, J.\ H., Cepa, J., Beckman, J.\ E., Soledad del Rio, M.,
1460:    \& Pedlar, A.\  1993, \apj, 416, 563
1461: \bibitem[Kritsuk \& Norman(2002a)]{kri02a}
1462:     Kritsuk, A. G., \& Norman, M.\ L.\ 2002a, \apj, 569, L127
1463: \bibitem[Kritsuk \& Norman(2002b)]{kri02b}
1464:     Kritsuk, A. G., \& Norman, M.\ L.\ 2002b, \apj, 580, L51
1465: \bibitem[Koyama \& Inutsuka(2002)]{koy02}
1466:     Koyama, H., \& Inutsuka, S.\ 2002, \apj, 564, L97
1467: \bibitem[Koyama \& Inutsuka(2004)]{koy04}
1468:     Koyama, H., \& Inutsuka, S.\ 2004, \apj, 602, L25    
1469: \bibitem[La Vigne et al.(2006)]{lav06}
1470:     La Vigne, M.\ A., Vogel, S.\ N., \& Ostriker, E.\ C.\ 2006, \apj, 650, 818
1471: \bibitem[Lubow et al.(1986)]{lub86}
1472:     Lubow, S.\ H., Balbus, S.\ A., \& Cowie, L.\ L.\ 1986, \apj, 309, 496
1473: \bibitem[Lynden-Bell(1966)]{lyn66} Lynden-Bell, D.\ 1966, Observatory, 86, 57
1474: \bibitem[Mac Low et al.(2005)]{mac05}
1475:     Mac Low, M.-M., Balsara, D.\ S., Kim, J., \& Avillez, M.\ A.\ 2005, \apj, 626, 864 
1476: \bibitem[Marochnik et al.(1983)]{mar83}
1477:     Marochnik, L.\ S., Berman, B.\ G., Mishurov, YU.\ N., Suchkov, A.\ A.\ 1983, \apss, 89, 171
1478: \bibitem[Martos \& Cox(1998)]{mar98}
1479:     Martos, M.\ A., \& Cox, D.\ P. 1998, \apj, 509, 703
1480: \bibitem[McKee \& Ostriker(1977)]{mo77}
1481:    McKee, C.\ F., \& Ostriker, J.\ P.\ 1977, \apj, 218, 148
1482: \bibitem[Meerson(1996)]{mee96}
1483:    Meerson, B.\ 1996, Rev.\ Mod.\ Phys.\ 68, 215
1484: \bibitem[Mufson(1974)]{muf74}
1485:     Mufson, S.\ L.\ 1974, \apj, 193, 561 
1486: \bibitem[Parker(1953)]{par53}
1487:    Parker, E.\ N.\ 1953, \apj, 117, 431
1488: \bibitem[Piontek \& Ostriker(2004)]{pio04}
1489:    Piontek, R.\ A., \& Ostriker, E.\ C.\ 2004, \apj, 601, 905
1490: \bibitem[Piontek \& Ostriker(2005)]{pio05}
1491:    Piontek, R.\ A., \& Ostriker, E.\ C.\ 2005, \apj, 629, 849
1492: \bibitem[Piontek \& Ostriker(2007)]{pio07}
1493:    Piontek, R.\ A., \& Ostriker, E.\ C.\ 2007, \apj, 663, 183
1494: \bibitem[Rand \& Kulkarni(1990)]{ran90}
1495:    Rand, R.\ J., \& Kulkarni, S.\ R.\ 1990, \apj, 349, L43
1496: \bibitem[Roberts(1969)]{rob69}
1497:    Roberts, W.\ W. 1969, \apj, 158, 123
1498: \bibitem[Roberts \& Yuan(1970)]{rob70} 
1499:    Roberts, W.\ W., \& Yuan, C.\ 1970, \apj, 161, 887
1500: \bibitem[Scoville et al.(2001)]{sco01}
1501:    Scoville, N.\ Z., Polletta, M., Ewald, S., Stolovy, S.\ R.,
1502:    Thompson, R., \& Rieke, M.\ 2001 \aj, 122, 3017
1503: \bibitem[Scoville \& Rector(2001)]{scorec01}
1504:    Scoville, N.\ \& Rector  T.\ 2001, HST press release at \\
1505:   {\tt http://oposite.stsci.edu/pubinfo/PR/2001/10/index.html}
1506: \bibitem[Seigar \& James(2002)]{sei02}
1507:    Seigar, M.\ S., \& James, P.\ A.\ 2002, \mnras, 337, 1113
1508: \bibitem[Shetty \& Ostriker(2006)]{she06}
1509:    Shetty, R., \& Ostriker, E.\ C.\ 2006, \apj, 647, 997
1510: \bibitem[Shetty et al.(2007)]{she07}
1511:    Shetty, R., Vogel, S.\ N., \& Ostriker, E.\ C., \& Teuben, P.\ T.\ 
1512:    2007, \apj, 665, 1138 
1513: \bibitem[Shu et al.(1972)]{shu72}
1514:    Shu, F.\ H., Milione, V., Gebel, W., Yuan, C., Goldsmith, D.\ W., \& Roberts, W.\ W.\
1515:    1972, \apj, 173, 557
1516: \bibitem[Shu et al.(1973)]{shu73}
1517:    Shu, F.\ H., Milione, V., \& Roberts, W.\ W.\ 1973, \apj, 183, 819
1518: \bibitem[Sleath \& Alexander(1996)]{sle96}
1519:    Sleath, J.\ P., \& Alexander, P.\ 1996, \mnras, 283, 353
1520: \bibitem[Tubbs(1980)]{tub80}
1521:     Tubbs, A.\ D.\ 1980, \apj, 239, 882
1522: \bibitem[V\'azquez-Semadeni et al.(2000)]{vaz00}
1523:     V\'azquez-Semadeni, E., Gazol, A., \& Scalo, J.\ 2000, \apj, 540, 271
1524: \bibitem[V\'azquez-Semadeni et al.(2006)]{vaz06}
1525:     V\'azquez-Semadeni, E., Ryu, D., Passot, T., Gonz\'alez, R. F., 
1526:     \& Gazol, A.\ 2006, \apj, 643, 245
1527: \bibitem[V\'azquez-Semadeni et al.(2007)]{vaz07}
1528:     V\'azquez-Semadeni, G\'omez, G.\ C., Jappsen, A.-K., Ballesteros-Paredes, 
1529:     J., Gonz\'alez, R.\ F., \& Klessen, R.\ S.\ 2007, \apj, 657, 870
1530: \bibitem[Vogel, Kulkarni, \& Scoville(1988)]{vog88} Vogel, S.\ N.,
1531:    Kulkarni, S.\ R., \& Scoville, N.\ Z.\ 1988, Nature, 334, 402
1532: \bibitem[Wada \& Koda(2004)]{wad04}
1533:     Wada, K., \& Koda, J.\ 2004, \mnras, 349, 270
1534: \bibitem[Willner et al.(2004)]{2004ApJS..154..222W} Willner, S.~P., et al.\ 
1535: 2004, \apjs, 154, 222
1536: \bibitem[Wolfire et al.(2003)]{wol03}
1537:    Wolfire, M.\ G., McKee, C.\ F., Hollenbach, D., Tielens, A.\ G.\ G.\ M.\
1538:    2003, \apj, 587, 278
1539: \bibitem[Woodward(1975)]{woo75}
1540:    Woodward, P.\ R.\ 1975, \apj, 195, 61  
1541: \bibitem[Woodward(1976)]{woo76}
1542:    Woodward, P.\ R.\ 1976, \apj, 207, 484 
1543: \end{thebibliography}
1544: 
1545: \clearpage
1546: 
1547: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1548: %TABLE1
1549: \begin{deluxetable}{lccclcclcclccclccc}
1550: \rotate
1551: \tabletypesize{\footnotesize}
1552: \tablecaption{Summary of model parameters and simulation results\label{tbl:model}} \tablewidth{0pt}
1553: \tablehead{
1554: \colhead{Model} &
1555: \colhead{$n_0$} &
1556: \multicolumn{2}{c}{Arm} &
1557: \colhead{} &
1558: \multicolumn{2}{c}{Transition} &
1559: \colhead{} &
1560: \multicolumn{2}{c}{Interarm} &
1561: \colhead{} &
1562: \multicolumn{3}{c}{Mass Fractions (\%)} &
1563: \colhead{} &
1564: \multicolumn{3}{c}{Volume Fractions (\%)} \\
1565: \cline{3-4}\cline{6-7}\cline{9-10}\cline{12-14}\cline{16-18}
1566: \colhead{}&
1567: \colhead{}&
1568: \colhead{$\!\!\Delta L/L_x\!\!\!\!$} &
1569: \colhead{$\!\!\Delta t/\tcro\!\!\!\!$} &
1570: \colhead{}&
1571: \colhead{$\!\!\Delta L/L_x\!\!\!\!$} &
1572: \colhead{$\!\!\Delta t/\tcro\!\!\!\!$} &
1573: \colhead{}&
1574: \colhead{$\!\!\Delta L/L_x\!\!\!\!$} &
1575: \colhead{$\!\!\Delta t/\tcro\!\!\!\!$} &
1576: \colhead{}&
1577: \colhead{Cold} &
1578: \colhead{$\!\!\!\!$Intermediate$\!\!\!\!$} &
1579: \colhead{Warm} &
1580: \colhead{}&
1581: \colhead{Cold} &
1582: \colhead{$\!\!\!\!$Intermediate$\!\!\!\!$} &
1583: \colhead{Warm} \\
1584: \colhead{(1)}&
1585: \colhead{(2)}&
1586: \colhead{(3)} &
1587: \colhead{(4)} &
1588: \colhead{}&
1589: \colhead{(5)} &
1590: \colhead{(6)} &  
1591: \colhead{}&
1592: \colhead{(7)} &
1593: \colhead{(8)} &
1594: \colhead{}&
1595: \colhead{(9)} &
1596: \colhead{(10)} &
1597: \colhead{(11)} &
1598: \colhead{}&
1599: \colhead{(12)} &
1600: \colhead{(13)} &
1601: \colhead{(14)} 
1602: }
1603: \startdata
1604:   SW0.1 &    0.1 & \nodata & \nodata & & \nodata & \nodata & & \nodata & \nodata & &  0 &    0 &      100 &  &      0 &    0 &     100 \\
1605:   SW0.5 &    0.5 &  0.00 & 0.05 & & 0.04 & 0.14 & & 0.96 & 0.82 & &    18 &     26 &     55 &  &      0 &     10 &     90 \\
1606:     SW1 &    1.0 &  0.00 & 0.10 & & 0.10 & 0.21 & & 0.89 & 0.69 & &    43 &     25 &     31 &  &      2 &     16 &     83 \\
1607:     SU2 &    2.0 &  0.01 & 0.14 & & 0.16 & 0.22 & & 0.83 & 0.64 & &    59 &     27 &     14 &  &      4 &     26 &     70 \\
1608:     SU3 &    3.0 &  0.01 & 0.16 & & 0.22 & 0.24 & & 0.77 & 0.60 & &    62 &      30 &     8 &  &      7 &     35 &     58 \\
1609:     SU4 &    4.0 &  0.02 & 0.19 & & 0.26 & 0.25 & & 0.72 & 0.56 & &    65 &      29 &     5 &  &     11 &     40 &     49 \\
1610:     SU5 &    5.0 &  0.03 & 0.20 & & 0.28 & 0.25 & & 0.69 & 0.54 & &    66 &      30 &     3 &  &     14 &     43 &     43 \\
1611:     SU6 &    6.0 &  0.04 & 0.22 & & 0.32 & 0.28 & & 0.63 & 0.50 & &    67 &      31&     2 &  &     18 &     48 &     34 \\
1612:     SU7 &    7.0 &  0.05 & 0.23 & & 0.35 & 0.30 & & 0.60 & 0.47 & &    70 &      29&      2 &  &     23 &     47 &     30 \\
1613:     SU8 &    8.0 &  0.06 & 0.24 & & 0.35 & 0.29 & & 0.59 & 0.47 & &    73 &      26 &     1 &  &     29 &     45 &     26 \\
1614:    SC10 &   10.0 &  0.08 & 0.27 & & 0.40 & 0.32 & & 0.52 & 0.41 & &    76 &      24 &     1 &  &     37 &     46 &     17 \\
1615:    SC20 &   20.0 & \nodata & \nodata & & \nodata & \nodata & & \nodata & \nodata & &   100 &      0 &      0 &  &    100 &      0 &      0 \\
1616: \enddata
1617: \tablecomments{Col.\ (1): Model name;
1618: Col.\ (2): Initial number density in units of $\cm^{-3}$;
1619: Cols.\ (3)-(8): Width and duration of the arm, transition, 
1620: and interarm regions.
1621: Cols.\ (9)-(14): Mass and volume fractions of the cold, 
1622: intermediate-temperature, and warm phases
1623: }
1624: \end{deluxetable}
1625: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1626: %FIGURE1
1627: \begin{figure}
1628:  \epsscale{1.}
1629:  \plotone{fig1.eps}
1630:  \caption{Analytic growth rate of TI versus perturbation wavelength
1631: (eq.\ [\ref{eq:disp}]) for cases 
1632: $\mathcal{K}=\Omega_0=0$ (\textit{dotted}), 
1633: $\mathcal{K}=\mathcal{K}_0$ with $\Omega_0=0$ (\textit{solid}),
1634: and $\mathcal{K}=\mathcal{K}_0$ with $\Omega_0=130\kms\kpc^{-1}$ 
1635: (\textit{dashed}), showing that rotation and conduction stabilize
1636: perturbations at large and small scales, respectively.  
1637: Symbols indicate the numerical growth 
1638: rates measured from test simulations, all with
1639: $\mathcal{K}=\mathcal{K}_0$: for a static medium (\textit{cross}),
1640: for a rotating medium without translational motion (\textit{diamond}), 
1641: and for a medium with both rotational and translational motion
1642: (\textit{plus}).  Note that numerical diffusion associated with 
1643: translational motion over the grid reduces the growth rates at small scales
1644: significantly.  See text for details.}
1645:  \label{fig:grate}
1646: \end{figure}
1647: 
1648: %FIGURE2
1649: \begin{figure}
1650:  \epsscale{1.}
1651:  \plotone{fig2.eps}
1652:  \caption{Dependence of numerical conductivity $\mathcal{K}_n$ on
1653: the translational velocity $v_{0x}$, the grid size $\Delta x$,
1654: and the perturbation wavelength $\lambda$ of TI, from
1655: the test simulations with physical conductivity 
1656: $\mathcal{K}=0$ (\textit{cross}), 
1657: $\mathcal{K}=\mathcal{K}_0$ (\textit{diamond}),
1658: and $\mathcal{K}=4\mathcal{K}_0$ (\textit{plus}).
1659: Note that $v_{0x}$ is expressed in units of
1660: $\kms$, while $\Delta x$ and $\lambda$ are in parsecs.
1661: The solid line, represented by equation (\ref{eq:numk}), 
1662: gives the best fit to the test results based on 
1663: the second-order PLM for spatial reconstruction in the Athena code
1664: we use.
1665: }
1666:  \label{fig:knum}
1667: \end{figure}
1668: 
1669: %FIGURE3
1670: \begin{figure}
1671:  \epsscale{1.0}
1672:  \plotone{fig3.eps}
1673:  \caption{\textit{Left}: 
1674: Density snapshots at 
1675: $t/\tcro=0.0$, $1.0$, $2.4$, $4.0$ from model SU2.
1676: Dotted lines labeled by $n_{\rm max}$ and $n_{\rm min}$ demarcate
1677: the cold and warm phases.
1678: \textit{Right}: Corresponding scatter plots of $n$ versus $P/k_B$ 
1679: together with the equilibrium cooling curve (\textit{solid line}).
1680: Dot-dashed lines marked by $T_{\rm max}$ and $T_{\rm min}$ 
1681: divide the warm ($T>T_{\rm max}$), intermediate-temperature
1682: ($T_{\rm min}<T<T_{\rm max}$), and cold ($T<T_{\rm min})$ phases.
1683:  }
1684:  \label{fig:tevol}
1685: \end{figure}
1686: 
1687: %FIGURE4
1688: \begin{figure}
1689:  \epsscale{1.0}
1690:  \plotone{fig4.eps}
1691:  \caption{Profiles of number density, temperature, 
1692: pressure, and perturbation components of the 
1693: velocities perpendicular ($v_x = v_{T,x}-v_{0x}$) 
1694: and parallel ($v_y=v_{T,y}-v_{0y}$) to the arm,
1695: for  model SU2 at $t/\tcro=4.0$.
1696: In (a) to (c), dotted lines demark the transitions between
1697: the cold, intermediate-temperature, and warm phases.
1698: Dashed lines in (d) and (e) show the time-averaged 
1699: velocity (streaming) profiles over $t/\tcro=5-8$.
1700:  }
1701:  \label{fig:sat}
1702: \end{figure}
1703: 
1704: %FIGURE5
1705: \begin{figure}
1706:  \epsscale{1.}
1707:  \plotone{fig5.eps}
1708: \caption{Schematic evolutionary track in the $n-P$ plane (\textit{solid lines
1709: with arrows}) of gas cycling through a spiral pattern with TI.
1710: The dotted line is the thermal equilibrium curve, 
1711: while dot-dashed lines mark $T_{\rm max}$ and $T_{\rm min}$.
1712: The warm and cold phases in the interarm regions (A) are shocked
1713: (B), undergo radiative cooling (C), become thermally unstable 
1714: due to postshock expansion (D), and return to the 
1715: interarm two-phase state again (A).  See text for details.}
1716:  \label{fig:schematic}
1717: \end{figure}
1718: 
1719: 
1720: %FIGURE6
1721: \begin{figure}
1722:  \epsscale{1.0}
1723:  \plotone{fig6.eps}
1724:  \caption{Density profile averaged over
1725: $t/\tcro=5-8$ from model SU2 (\textit{solid}), compared to
1726: stationary shock profiles for isothermal models. \textit{Dotted curve}:  
1727: $T=1000\Kel$ non-viscous model; 
1728: \textit{dot-dashed curve}: $T=500\Kel$ 
1729: viscous models with $l_0=200\pc$; 
1730: \textit{dashed curve}:
1731: $T=1000\Kel$ viscous model with $l_0=100\pc$.
1732:  }
1733:  \label{fig:vis}
1734: \end{figure}
1735: 
1736: 
1737: %FIGURE7
1738: \begin{figure}
1739:  \epsscale{1.}
1740:  \plotone{fig7.eps}
1741:  \caption{
1742: Density profile of model SU5 with $n_0=5\pcc$ 
1743: at $t/\tcro=4$.   Note that
1744: the postshock transition zone in model SU5 is wider than that of
1745: model SU2 shown in Figure~\ref{fig:sat}.
1746: }
1747:  \label{fig:high_n}
1748: \end{figure}
1749: 
1750: %FIGURE8
1751: \begin{figure}
1752:  \epsscale{1.}
1753:  \plotone{fig8.eps}
1754:  \caption{Density profiles and scatter plots in the
1755: density--pressure plane of models SW0.1 with $n_0=0.1\pcc$ 
1756: (\textit{upper frames}) and 
1757: SC20 with $n_0=20\pcc$ (\textit{lower frames}) 
1758: at $t/\tcro=2.4$.
1759: Note that the spiral shocks in these models are smooth and 
1760: each case contains only a single phase of gas.  The shock compression is
1761: much stronger for all-cold than all-warm gas, because of the higher
1762: Mach number.}
1763:  \label{fig:single}
1764: \end{figure}
1765: 
1766: %FIGURE9
1767: \begin{figure}
1768:  \epsscale{1.}
1769:  \plotone{fig9.eps}
1770:  \caption{Mass-weighted ({\it thick lines}) 
1771: and volume-weighted ({\it thin lines}) temperature PDFs 
1772: averaged over $t/\tcro=5-8$ for models 
1773: SW0.1, SU2, SU5, and SC10.  In each panel, vertical 
1774: dotted lines indicate $T_{\rm min}$ and $T_{\rm max}$ to demark 
1775: the cold, intermediate-temperature, and warm phases.
1776: Of the cold component, the gas with $T\simlt 50\Kel$ is located in 
1777: the arm region, while cold clouds with $50\Kel\simlt T\simlt T_{\rm min}$
1778: are found in the interarm or transition zones.
1779:  }
1780:  \label{fig:pdf}
1781: \end{figure}
1782: 
1783: %FIGURE10
1784: \begin{figure}
1785:  \epsscale{1.}
1786:  \plotone{fig10.eps}
1787:  \caption{Mass fractions of the cold, intermediate-temperature,
1788: and warm phases in model SU2 versus numerical resolution.
1789: The numerical results are not affected by numerical conductivity
1790: as long as the number of zones is larger than $10^4$.
1791:  }
1792:  \label{fig:conv}
1793: \end{figure}
1794: 
1795: %FIGURE11
1796: \begin{figure}
1797:  \epsscale{1.}
1798:  \plotone{fig11.eps}
1799:  \caption{Mass and volume fractions of the cold (\textit{square}),
1800: warm (\textit{triangle}), and intermediate-temperature (\textit{diamond}) 
1801: phases averaged over $t/\tcro = 5-8$,
1802: as functions of the initial number density $n_0$. 
1803: Errorbars indicate the standard deviations of the measurements.
1804: Dotted lines show the theoretical estimates computed by adopting a
1805: constant value $f_i=0.28$ for the intermediate-temperature of the 
1806: mass fraction.
1807:  }
1808:  \label{fig:mf_ndep}
1809: \end{figure}
1810: 
1811: %FIGURE12
1812: \begin{figure}
1813:  \epsscale{1.}
1814: \plotone{fig12.eps}
1815: \caption{Density-weighted velocity dispersions $\sigma_x$ and $\sigma_y$, 
1816: relative to time-averaged template values, of the gas in the arm 
1817: (\textit{solid}), transition (\textit{dotted}), interarm (\textit{dashed}) 
1818: regions of model SU2.  The large-amplitude fluctuations of 
1819: the velocity dispersions
1820: in the arm region are caused
1821: by incomplete subtraction of the arm streaming motions.  
1822: The velocity dispersions in both unstable and interarm regions
1823: are not subject to this effect, and have mean values of $\sim1.3\kms$
1824: and $\sim 0.8\kms$ in the $x$- and $y$-directions, respectively.
1825:  }
1826:  \label{fig:veld_tevol}
1827: \end{figure}
1828: 
1829: %FIGURE13
1830: \begin{figure}
1831:  \epsscale{1.}
1832: \plotone{fig13.eps}
1833:  \caption{Mean values (\textit{symbols}) and standard 
1834: deviations (\textit{errorbars}) of the density-weighted velocity 
1835: dispersions of each phase during the time span $t/\tcro=5-8$ 
1836: for models with multi-phase spiral shocks.
1837:  }
1838:  \label{fig:veld_ndep}
1839: \end{figure}
1840: 
1841: \end{document}
1842: