1: %%
2: %% Beginning of file 'ms.tex' Watanabe and Lin 2007 to ApJ
3: %%
4: %% Last Modified 09/07/2007
5: %%
6: %% This is a sample manuscript marked up using the
7: %% AASTeX v5.x LaTeX 2e macros.
8:
9:
10: \documentclass[12pt,preprint]{aastex}
11: %\documentclass[12pt,manuscript]{aastex}
12: % \documentclass{emulateapj}
13:
14: %% definition of my own macros
15:
16: \newcommand{\Msub}[1]{_{\rm #1}} % roman subscript
17: \newcommand{\Msup}[1]{^{\rm #1}} % roman superscript
18: \newcommand{\DP}[2]{\frac{\partial {#1}}{\partial {#2}}} % partial differential
19: \newcommand{\Sc}[0]{{\cal S}{\rm c}} % Schumidt Number
20: \newcommand{\Punit}[1]{\, \mbox{#1}} % Physical unit in the math mode
21:
22: %% You can insert a short comment on the title page using the command below.
23: %\slugcomment{}
24:
25: \shorttitle{Thermal Waves in Irradiated Protoplanetary Disks}
26: \shortauthors{Watanabe and Lin}
27:
28:
29: \begin{document}
30:
31: \title{Thermal Waves in Irradiated Protoplanetary Disks}
32:
33: \author{Sei-ichiro Watanabe\altaffilmark{1}}
34: \affil{Department of Earth and Planetary Sciences, Nagoya University,
35: Chikusa, Nagoya 464-8601, Japan}
36: \email{seicoro@eps.nagoya-u.ac.jp}
37: \altaffiltext{1}{Visiting Researcher, UCO/Lick Observatory,
38: University of California, Santa Cruz, CA 95064}
39:
40: \and
41:
42: \author{D. N. C. Lin\altaffilmark{2}}
43: \affil{UCO/Lick Observatory, University of California, Santa Cruz, CA 95064}
44: \email{lin@ucolick.org}
45: \altaffiltext{2}{KIAA, Peking University, Beijing 100871, China}
46:
47:
48: \begin{abstract}
49: Protoplanetary disks are mainly heated by radiation from the central
50: star. Since the incident stellar flux at any radius is sensitive to
51: the disk structure near that location, an unstable feedback may be
52: present. Previous investigations show that the disk would be stable to
53: finite-amplitude temperature perturbations if the vertical height of
54: optical surface is everywhere directly proportional to the gas scale
55: height and if the intercepted fraction of stellar radiation is
56: determined from the local grazing angle. We show that these
57: assumptions may not be generally applicable. Instead, we calculate
58: the quasi-static thermal evolution of irradiated disks by directly
59: integrating the global optical depths to determine the optical surface
60: and the total emitting area-filling factor of surface dust. We show
61: that, in disks with modest mass accretion rates, thermal waves are
62: spontaneously and continually excited in the outer
63: disk, propagate inward through the planet-forming domains, and
64: dissipated at small radii where viscous dissipation is dominant. This
65: state is quasi-periodic over several thermal timescales and its
66: pattern does not depend on the details of the opacity law. The viscous
67: dissipation resulting from higher mass accretion stabilizes this
68: instability such that an approximately steady state is realized throughout
69: the disk. In passive protostellar disks, especially transitional disks, these waves
70: induce significant episodic changes in SEDs, on the time scales of
71: years to decades, because the midplane temperatures can vary by a
72: factor of two between the exposed and shadowed regions. The transitory
73: peaks and troughs in the potential vorticity distribution may also
74: lead to baroclinic instability and excite turbulence in the
75: planet-forming regions.
76: \end{abstract}
77:
78:
79: \keywords{accretion, accretion disks --- circumstellar matter --- instabilities
80: --- planetary systems: protoplanetary disks --- stars: pre--main-sequence
81: --- solar system: formation}
82:
83:
84: \section{Introduction}
85: \label{SEC:Intro}
86:
87: It has become widely accepted that dusty protoplanetary disks are
88: heated by radiation from the central star, and that this heating
89: mainly determines the physical structure of the outer regions of these
90: disks. Observed infrared spectral energy distributions (SEDs) of the
91: T Tauri disks imply that their effective temperatures $T$ decreases
92: with disk radius $r$ more slowly than $T \propto r^{-3/4}$. This
93: temperature distributions is usually explained by a model in which the
94: thermal structure of the disk is assumed to be geometrically flared,
95: i.e., the surface height $z\Msub{s}$ where stellar radiation is
96: absorbed curves away from the midplane (or equivalently $z\Msub{s}/r
97: \propto r^{\gamma}$, with $\gamma > 0$) \citep{ALS87}.
98:
99: The outer regions of these disks are irradiated by the central star
100: \citep{KH87} and the flaring enables the disks to absorb more
101: radiation from the central star. In a steady state, the flaring index
102: $\gamma$ of a purely irradiated optically-thick disk can be obtained
103: from the balance between intercepted stellar flux $F\Msub{s}$ incident
104: on the surface at a low angle $\theta$ and emitted blackbody flux from
105: the disk interior, under the assumptions that 1) the surface height
106: $z\Msub{s}$ is everywhere proportional to the vertical gas scale
107: height $h$, 2) the intercepted fraction of stellar radiation is the
108: sine of the local grazing angle ($F\Msub{s} \propto \sin \theta$), and
109: 3) the central star is a point source. Under these circumstances,
110: there is a self-consistent power-law solution with $\gamma = 2/7$,
111: which corresponds to $T \propto r^{-3/7}$ \citep{KNH70, CG97} .
112:
113: In addition to this particular power-law solution, there
114: exists an one-parameter family of solutions for the disk structure,
115: including the diverging aspect-ratio solutions and the asymptotically
116: conical (in which $z\Msub{s} \propto r$) solutions \citep{D00}. Small
117: variations in the values of $z\Msub{s}$ at inner radii where the
118: integration starts can cause large differences in the disk structure
119: at large radii. \citet{D00} speculated that such a sensitive nature
120: of the steady solutions is suggestive to an intrinsic instability
121: which must be analyzed with time-dependent governing equations.
122:
123: More realistic steady solutions can be obtained numerically or
124: semi-analytically to take into account the effects of the finite
125: values of stellar radius, the disk optical depth, and the viscous
126: dissipation associated with the mass-accretion flow
127: \citep[e.g.,][]{CG97, CJC01, DDN01, THI05, DCHFS06, GL07}. The most
128: important novel feature of these series of second-generation models is
129: the assumed presence of superheated surface dust layers above and
130: below the disk midplane \citep{CG97}. Grains in these layers are
131: directly exposed to the stellar flux. Grains much smaller than the
132: peak wavelength of the self emission are superheated because of their
133: low emissivity. The disk interior is heated by the superheated dust of
134: the layers rather than directly by the central star.
135:
136: The two-layer disk model clearly explain the silicate and water-ice
137: emission bands in observed SEDs of Herbig Ae/Be stars and T Tauri
138: stars \citep{CJC01}. But, in spite of their triumph in the modeling
139: of the observed SEDs, these models\footnote{Some author
140: \citep[e.g.,][]{DDN01, THI05, GL07} determined $\chi$
141: self-consistently, but these calculations are based on the
142: grazing-angle approximation and the assumption that $\chi$ changes
143: slowly with radius.} are based on the assumption that $z\Msub{s}$ is
144: proportional to the gas scale height $h$, with a fixed constant of
145: proportionality $\chi = z\Msub{s}/h = 4$ \citep{CJC01}. While this
146: assumption has been justified by estimates which suggest that changes
147: of $\chi$ is small throughout the disk, the amount of dust in the
148: superheated layer is very sensitive function of $\chi$. This
149: dependence arises because the dust spatial density at $z\Msub{s} =
150: \chi h$ is proportional to $\exp[-\chi^2/2]$. Thus, the fixed-$\chi$
151: assumption may cause large discrepancy in determined values of
152: $z\Msub{s}$.
153:
154: In previous analyses, the magnitude $F\Msub{s}$ is directly determined
155: from the grazing angle $\theta$ \citep{CJC01}. This approximation
156: justified only in the case that the length of absorption layer (say,
157: where the optical depth to the starlight changes from 0.1 to 1) along
158: the starlight is smaller than the lengths of radial variations of
159: surface density or temperature. This condition would not be satisfied
160: if the disk surface contains fluctuations resulting from the growth of
161: short-wavelength perturbations (see \S~\ref{SEC:Simple}).
162:
163: In principle, $z\Msub{s}$ is determined by the condition that the
164: visual optical depth, which can be obtained by a direct integration
165: along the rays of starlight, is unity and the surface filling factor
166: $A\Msub{s}$ of the irradiated dust grains can be calculated through
167: the vertical (in the direction normal to the disk plane) integration
168: of geometrical opacity times mass density from $z\Msub{s}$ to
169: infinity. This global procedure yields $z\Msub{s}$ at any given radius
170: which depends not only on the local value of $h$ but also the disk
171: structure interior to that radius. Thus, $\chi = z\Msub{s}/h$ varies
172: both in space and time. Following this procedure, we can check for
173: self consistency by recalculating $z\Msub{s}$ based on the temperature
174: distribution obtained from the steady, constant-$\chi$ model. With
175: this inductive approach, we demonstrate that there are substantial
176: differences between the iterated values of $z\Msub{s}$ and the
177: initial, assumed values of $\chi h$. We also find that $A\Msub{s}$
178: calculated from the deduced values of $z\Msub{s}$ is substantially
179: different from the values of $\sin \theta$ extrapolated from the
180: constant-$\chi$ model.
181:
182: Since the irradiation heating can play such a major role in
183: determining the vertical structure of protoplanetary disks, it is
184: important to investigate the stability of such disks against the
185: excitation of ripples on their surfaces. Under the assumption that the
186: thermal timescale is much longer than the dynamical timescale in
187: protostellar disks, \citet{DCH99} investigated the thermal stability
188: of the irradiation-dominated disks, using a simple cooling equation.
189: They found the vertically isothermal disk to be stable against finite
190: amplitude perturbations. The initial temperature perturbations
191: propagate inward and damp out quickly. However, in their analysis,
192: they assumed that $\chi$ is constant throughout the disk and
193: $F\Msub{s}$ is given by $\theta$ as in the grazing angle
194: approximation. The inferred stability in that study may depend on
195: these assumptions. As we will see in \S~\ref{SEC:Simple}, changes in
196: $\chi$ may lead to an instability.
197:
198: With a linear perturbation analysis, \citet{D00} showed that the
199: flaring disk solution may become unstable to infinitesimal
200: hydrodynamic perturbations when the cooling time of the disk is much
201: shorter than the dynamical time \citep[the opposite limit
202: of][]{DCH99}. The amplitude of inwardly propagating waves grows
203: exponentially with a rate that is a decreasing function of the
204: wavelength. Subsequently, \citet{DD04a} constructed a series of
205: numerical models to examine the two-dimensional structure and
206: evolution of protoplanetary disks around Herbig Ae/Be stars. In these
207: simulations, they studied the radiative transfer process under the
208: assumption that the disk always maintains a hydrostatic
209: equilibrium. (This assumption would not be appropriate for the limit
210: that the time scale of cooling is shorter than that of dynamics.)
211: They found two sets of asymptotically steady-state results which
212: include the monotonically flaring solutions and the self-shadowed
213: solutions. For the second set of solutions, the disk has a puffed-up
214: inner rim. They stated that their iteration procedure, in which the
215: hydrostatic equilibrium and the radiative transfer are treated
216: separately in alternate steps, may not, under some circumstances, lead
217: to a set of converged solutions. In their simulations, some wave-like
218: disturbances appear to propagate over the disk from one iterative step
219: to the next and these transitory features are never damped out
220: completely. Although they limited their presentations on disks around
221: Herbig Ae/Be stars, where the perturbation on the disk structure by
222: these waves are relatively minor, they revealed that this problem
223: appears to be more serious for disks around T Tauri stars. We
224: speculate that this perturbation may be related to the above-mentioned
225: instabilities operating in the irradiation-dominated outer regions of
226: protostellar disks.
227:
228: In this paper, we attempt to address two issues: 1) Are the steady-state
229: solutions of irradiated disks constructed by the previous
230: one-dimensional models self-consistent and stable? 2) Do these regions
231: of disks tend to undergo quasi-periodic oscillations rather than
232: attain an asymptotic steady state? In principle, these questions
233: should be addressed with comprehensive two- or three-dimensional
234: numerical simulations. Such an approach is, however, fairly complicated,
235: time consuming, and often plagued with problems in the algorithm which
236: implements the radiative transfer processes \citep[e.g.,][]{DD04a}.
237: Prior to these detailed simulations, it is useful to identify the dominant
238: effects which regulate the dynamics of irradiated disks with a set of
239: one-dimensional time-dependent analyses on the thermal evolution of
240: protostellar disks.
241:
242: Since thermal timescale is much shorter than viscous diffusion
243: timescale, most previous studies adopted the assumption of thermal
244: equilibrium during the course of disks' global evolution.
245: Nevertheless, there have been a few investigations on the thermal
246: evolution of protoplanetary disks. \citet{WNN90} investigated the
247: cooling and quasi-static contraction of the protoplanetary disks from
248: an initial high-temperature state. They performed vertical
249: one-dimensional numerical calculations and found that the cooling
250: times are well estimated by a simple two-temperature (surface and
251: interior temperatures) prescription. In this paper, we utilize this
252: two-layer disk-temperature prescription to examine the stability and
253: thermal evolution of irradiated disks.
254:
255: The simplest treatments for the thermal evolution of the irradiated
256: disk are radial one-dimensional models in which the
257: vertical structure of the disk at each radii is analyzed
258: independently. In order to take into account of the irradiated
259: surface and the disk interior, we evaluate the surface height directly
260: from the location where the visual optical depth along the straight
261: lines from the star is unity. We show that such disks evolve to
262: quasi-periodic states in which thermal waves propagate inward through
263: intermediate disk radii, where planets are formed.
264:
265: A simple discussion about the nature of thermal instability is given
266: in \S~\ref{SEC:Simple}. The basic assumption of our model and its
267: governing equations are presented in \S~\ref{SEC:Basic}. The results
268: of our numerical calculations are presented in \S~\ref{SEC:Results}
269: for both simple and realistic opacities. Finally, we summarize our
270: findings and discuss some possible evolutionary scenarios.
271:
272: \section{Simple discussion about thermal instability}
273: \label{SEC:Simple}
274:
275: In this section, we discuss some potential causes for
276: irradiation-dominated regions of disks to become thermally unstable.
277: For illustration convenience, we adopt the following simplifying
278: assumptions: 1) the star is a point source, 2) the disk's internal
279: heat sources such as turbulent viscous heating as well as external
280: heating other than the stellar radiation are negligible, 3) the
281: optical depth of the disk is much larger than unity for both stellar
282: radiation and its own emission, and 4) the transport of energy in the
283: radial direction is much smaller than that in the vertical direction.
284: Note that these assumptions are adopted only in this section for the
285: purpose of pinpointing the physical process which leads to the thermal
286: instability. All of these idealized assumptions will be relaxed in
287: the numerical simulations to be presented below.
288:
289: Under these assumptions, we consider the heat balance in a
290: geometrically thin disk which is irradiated by the central star.
291: The energy equation reduces to
292: \begin{equation}
293: C \Sigma \DP{T\Msub{m}}{t} = 2 (F\Msub{s} - F\Msub{m}),
294: \label{EQ:therm}
295: \end{equation}
296: where $C$ is the specific heat per unit disk mass, $\Sigma$ is the
297: surface density of the disk, and $T\Msub{m}$ is the temperature of the
298: disk interior. Under assumption 2, the disk interior has an
299: approximately isothermal structure and $F\Msub{m}$ is the disk
300: blackbody emission given by
301: \begin{equation}
302: F\Msub{m} = \sigma T\Msub{m}^{4},
303: \end{equation}
304: and $F\Msub{s}$ is the intercepted stellar flux given by
305: \begin{equation}
306: F\Msub{s} = \frac{1}{2} \frac{L_\star}{4 \pi r^2} A\Msub{s},
307: \label{EQ:Fs1}
308: \end{equation}
309: where $\sigma$ is the Stefan-Boltzmann constant, $L_\star$ is the
310: stellar luminosity, $r$ is the cylindrical
311: radial coordinate, and $A\Msub{s}$ is the total emitting area-filling
312: factor of superheated dust grains. The factor 1/2 in the right-hand
313: side of equation~(\ref{EQ:Fs1}) comes from the fact that surface
314: irradiated dust re-radiate half of the absorbed stellar flux toward
315: the disk interior (the rest toward infinity). Assuming a homogeneous
316: mixing of gas and dust, we can obtain $A\Msub{s}$ from (see
317: Appendix~\ref{SEC:Estimate})
318: \begin{equation}
319: A\Msub{s} = \tau\Msub{G} \mbox{ erfc} \left( \frac{z\Msub{s}}
320: {\sqrt{2} h} \right)
321: = \tau\Msub{G} \mbox{ erfc} \left( \frac{\chi}{\sqrt{2}} \right),
322: \label{EQ:Assec2}
323: \end{equation}
324: where $\mbox{erfc} (x)$ is the complimentary error function,
325: $\tau\Msub{G}$ is the geometrical optical depth of the disk midplane,
326: and $z\Msub{s}$ is the surface height where stellar radiation is
327: absorbed. The ratio of the surface height $z\Msub{s}$ to the gas
328: scale height $h$ is denoted by $\chi \equiv z\Msub{s}/h$. In a
329: hydrostatic equilibrium, the gas scale height $h$ is given by
330: \begin{equation}
331: h = \frac{c\Msub{m}}{\Omega\Msub{K}} = \left(
332: \frac{k\Msub{B} T\Msub{m} r^3}{\mu m\Msub{u} G M_\star}
333: \right)^{1/2},
334: \label{EQ:h}
335: \end{equation}
336: where $c\Msub{m}$ is the disk sound speed, $\Omega\Msub{K}$ is the Keplerian
337: angular velocity, $k\Msub{B}$ is the Boltzmann constant, $\mu$ is the
338: molecular weight of disk gas, $m\Msub{u}$ is the atomic mass unit, $G$
339: is the gravitational constant, and $M_\star$ is the stellar mass.
340:
341: Figure~\ref{FIG:FmFs} displays the $F\Msub{m}$ and $F\Msub{s}$ (at $10
342: \Punit{AU}$) as functions of $T\Msub{m}$. Assuming that the initial
343: state is in a thermal equilibrium (corresponding to the point where
344: three thick lines cross) with $T\Msub{m} = T\Msub{m, eq}$, we impose a
345: small positive temperature perturbation. We consider two extreme
346: cases: 1) If $z\Msub{s}$ is determined by the local disk structure,
347: $\chi = z\Msub{s}/h$ would retain a constant value during the increase
348: of $T\Msub{m}$ such that $A\Msub{s}$ would also be constant (see
349: eq.~[\ref{EQ:Assec2}]) and $F\Msub{s}$ would not change ({\it
350: dot-dashed line}). In this case the system would be stabilized because
351: $F\Msub{m} > F\Msub{s}$ for $T\Msub{m} > T\Msub{m, eq}$. 2) If
352: $z\Msub{s}$ is determined mostly by the attenuation by dust in the
353: inner regions of the disk, $z\Msub{s}$ would remain constant despite
354: changes in the local disk temperature and $h$ such that $A\Msub{s}$
355: would increase rapidly and $F\Msub{s}$ would increase much faster than
356: $F\Msub{m}$ ({\it thick dash curve}). In this case the system would
357: be unstable because $F\Msub{m} < F\Msub{s}$ for $T\Msub{m} > T\Msub{m,
358: eq}$.
359:
360: In their stability analysis, \citet{DCH99} assumed a constant $\chi$.
361: Based on the above analysis, this assumption naturally leads to stable
362: solutions. In fact, most of the analysis on the structure of
363: irradiated disks are based on the constant-$\chi$ assumption
364: \citep[e.g.,][]{CJC01}. The usual procedure to determine $z\Msub{s}$
365: is based on a geometrical consideration, i.e,
366: \begin{equation}
367: A\Msub{s} = \sin \theta \simeq \frac{z\Msub{s}}{r}
368: \left( \frac{d \ln z\Msub{s}}{d \ln r} - 1 \right),
369: \label{EQ:Astheta}
370: \end{equation}
371: where $\theta$ is the grazing angle (i.e., the angle between the
372: starlight and the disk surface). We refer this procedure to be the
373: grazing-angle approximation. Most previous steady-state disk models
374: are constructed with equation~(\ref{EQ:Astheta}) under the assumption
375: that $\chi$ is constant throughout the disk.
376:
377: However, the magnitude of $\chi$ is generally determined by the radial
378: structure of the disk as well as its local properties. The surface
379: height $z\Msub{s}$ is determined by the optical depth integrated
380: through a ray of the stellar radiation. We derive a ray integral and
381: check the validity of equation~(\ref{EQ:Astheta}) in
382: \S~\ref{SEC:Basic} and Appendix~\ref{SEC:Estimate}. In principle,
383: equations~(\ref{EQ:Assec2}) and (\ref{EQ:Astheta}) must be resolved
384: simultaneously \citep{THI05}. However, this set of equations is
385: fairly unstable to solve numerically because they do not contain
386: contributions which may reduce any steep temperature gradients in the
387: radial direction.
388:
389: Steep temperature gradient, if present, would invalidate the constant
390: $\chi$ and the grazing-angle approximations. Physically, the radial
391: transport of heat suppresses the radial temperature gradient, but such
392: a process through the opaque regions of the disk must be analyzed with
393: multi-dimensional numerical simulations. One of the most efficient
394: process of the radial heat transport is the radiative transfer from
395: the superheated dust grains at the surface of any radial location to
396: the disk midplane at adjacent radial regions. Using a simple
397: one-dimensional model, we can take this oblique radiative transfer of
398: heat into account.
399:
400:
401: \section{Basic Equations}
402: \label{SEC:Basic}
403:
404: Following the approaches of \citet{CG97} and \citet{GL07}, we
405: construct numerical models to study the thermal evolution of a
406: protostellar accretion disk. The surface of the disk is illuminated
407: by the central star. Exposed to the stellar radiation, sub-mm dust
408: grains in the surface layers of the disk are superheated. We consider
409: the case that dust mass of the disk is so large that the disk midplane
410: (except for an innermost region where silicates are evaporated) is
411: optically thick to the stellar radiation. In contrast to the previous
412: section, the heat sources for the disk interior in these numerical
413: models include both irradiation from the superheated grains on the disk
414: surface and the viscous dissipation associated with the accretion
415: flow. We adopt a cylindrical coordinate system $(r, \phi, z)$ in
416: which the $z=0$ plane represents the disk midplane and the origin is
417: at the location of the central star. Since the star-and-disk system
418: is symmetric with respect to the midplane, we describe our results for
419: the upper half of the disk only.
420:
421: In order to simplify the problem we adopt the two-layer axisymmetric
422: disk model proposed by \citet{CG97}. In this model, the disk consists
423: of a surface superheated layer where the dust temperature is
424: $T\Msub{s} (r)$ and a disk interior where the dust and gas temperature
425: is assumed to be uniform at $T\Msub{m} (r)$. This model is simple to
426: use and includes all the essential ingredient to analyze the onset,
427: evolution, and stabilization of thermal instability in protostellar
428: disks. However, such a simplification would be invalid if the disk
429: optical depth $\tau\Msub{m}(T\Msub{m})$ to its intrinsic radiation is
430: much larger than unity and viscous heating rate is larger than surface
431: heating rate. However, the dust optical depth may be self-limited by
432: their rapid growth through cohesive collisions, so that
433: $\tau\Msub{m}(T\Msub{m}) \la 10$ throughout the disk \citep[see Fig.~6
434: in][]{THI05}. Thus, the two-layer model is valid even in the inner
435: disk where dust surface density is larger.
436:
437: The two-layer model is invalid at the inner edge of the disk, where
438: disk is irradiated not only from the top but from the radial
439: direction. The disk may have a puffed-up inner rim
440: \citep[e.g.,][]{DD04a}, but a set of two-dimensional radiative
441: transfer calculations is needed to determine the structure of the
442: innermost region. In this work we simply assume that disk within $0.1
443: \Punit{AU}$ is optically thick in the radial direction and has no
444: puffed-up rim that might cast shadows over the outer regions of the
445: disk. We confine our calculations only in the regions $r > 0.1
446: \Punit{AU}$, where the two-layer model is valid.
447:
448: The thermal timescale of the disk interior at radius $r$ is given by
449: (see eq.~[\ref{EQ:energy}])
450: \begin{eqnarray}
451: t\Msub{th} & = & \frac{(\gamma\Msub{a}+1)}{2(\gamma\Msub{a}-1)}
452: \frac{c\Msub{m}^2 \Sigma}{\sigma T\Msub{m}^4}
453: \nonumber \\
454: & \simeq & 53 \left(\frac{\Sigma_0}{\Sigma\Msub{H0}} \right)
455: \left( \frac{T\Msub{m0}}{124 \Punit{K}} \right)^{-3}
456: \left( \frac{r}{1 \Punit{AU}} \right)^{3q-p} \Punit{yr},
457: \label{EQ:tth}
458: \end{eqnarray}
459: where $\gamma\Msub{a}$ is the adiabatic exponents and $c\Msub{m}$ is
460: the sound speed of the disk interior. For evaluation, we assume here
461: power-law distributions for the total ($\mbox{gas} + \mbox{dust}$) surface
462: density $\Sigma(r) \propto r^{-p}$ and the midplane temperature
463: $T\Msub{m}(r) \propto r^{-q}$. The normalization factors, $\Sigma_0$ and
464: $T\Msub{m0}$, refer to their corresponding values at $r =1\Punit{AU}$.
465: The nominal value of surface density is given by that in the minimum mass
466: solar-nebula (MMSN) model in which $\Sigma\Msub{H} = 1.7 \times
467: 10^3 \Punit{g} \Punit{cm}^{-2}$ \citep{H81}.
468:
469: We assume that the thermal timescale $t\Msub{th}$ to be much longer
470: than the dynamical time ($\Omega\Msub{K}^{-1}$), but much shorter
471: than the viscous evolution time ($r^2/\nu$, where $\nu$ is the turbulent
472: viscosity). In this case we can regard that the whole region of the
473: disk is always in a hydrostatic equilibrium in the vertical direction
474: and has time-independent surface densities.
475:
476: The temperature $T\Msub{s}$ of superheated dust grains is given by
477: \begin{equation}
478: \frac{L_\star}{4 \pi r^2} = 4 \epsilon\Msub{s} \sigma T\Msub{s}^4,
479: \label{EQ:Ts}
480: \end{equation}
481: where $\epsilon\Msub{s}$ is the averaged emissivity of the dust grains
482: at $T\Msub{s}$. Along a ray from the surface of the star, the
483: superheated layer extends outward until the position where visual
484: optical depth has reached unity. We take into account the attenuation
485: of the stellar photons by defining the height $z\Msub{s}$ of the bottom
486: of the superheated layer with the following equation
487: \begin{equation}
488: \tau\Msub{s}(T_\star; r, z\Msub{s}(r)) = 1.
489: \label{EQ:zsdef}
490: \end{equation}
491: Here $\tau\Msub{s} (T_\star; r, z)$ is the optical depth between the
492: central star and the point $(r, z)$ to the blackbody radiation peaked
493: at the stellar effective temperature $T_\star$, given by the following
494: integration
495: \begin{equation}
496: \tau\Msub{s}(T_\star; r, z) = \int_{R_\star}^r \bar{\kappa}\Msub{s}(T_\star)
497: \rho\Msub{d}(r',\zeta r') \left( 1 + \zeta^2 \right)^{1/2} \, dr'
498: \label{EQ:taus}
499: \end{equation}
500: along a straight path from the star to the point $(r,z)$. Here
501: $R_\star$ is the stellar radius, $\zeta \equiv z/r$ is the aspect ratio,
502: $\rho\Msub{d}(r',z')$ is the spatial mass density of dust at $(r', z')$,
503: and $\bar{\kappa}\Msub{s}(T\Msub{rad})$ is the Planck mean opacity of
504: the grains interacting with the blackbody radiation peaked at
505: $T\Msub{rad}$. Note that we define here the grain opacity per unit
506: {\it dust mass}, not per unit total ($\mbox{gas} + \mbox{dust}$) mass as in the usual
507: definition, because it is convenient for the consideration of the case
508: that the dust-to-gas ratio may change vertically. The emissivity in
509: equation~(\ref{EQ:Ts}) can be given by
510: \begin{equation}
511: \epsilon\Msub{s} = \frac{\bar{\kappa}\Msub{s}(T\Msub{s})}{%
512: \bar{\kappa}\Msub{s}(T_\star)}.
513: \label{EQ:epsilons}
514: \end{equation}
515:
516: The energy equation includes heating from both stellar irradiation and
517: viscous dissipation as well as radiative losses from the disk surface
518: such that \citep[see, e.g.,][]{WNN90}
519: \begin{equation}
520: \frac{(\gamma\Msub{a} + 1)}{2(\gamma\Msub{a}-1)}
521: \frac{k\Msub{B}\Sigma}{\mu m\Msub{u}}
522: \DP{T\Msub{m}}{t} =
523: 2 \left[ F\Msub{s} - F \Msub{m} \right]
524: + \frac{3}{4 \pi} \dot{M} \Omega\Msub{K}^2,
525: \label{EQ:energy}
526: \end{equation}
527: where $\dot{M}$ is the mass accretion rate, which we assume to be
528: constant throughout the disk. Note that the steady state assumption is
529: compatible with a power-law surface density distribution for some
530: effective viscosity prescriptions \citep{CG97, DCHFS06, GL07}.
531:
532: Further, $F\Msub{s}$ and $F\Msub{m}$ are, respectively, the thermal
533: radiation fluxes downward from the superheated dust grains high up in
534: the disk atmosphere and upward from dust grains in the disk interior,
535: \begin{equation}
536: F\Msub{s}(r) = \left[ 1 - e^{-2\tau\Msub{m}(T\Msub{s})} \right]
537: \frac{L_\star}{8 \pi}
538: \left\langle \frac{A\Msub{s}}{r^2} +
539: \frac{4R_\star}{3 \pi r^3} \right\rangle,
540: \label{EQ:Fs}
541: \end{equation}
542: %
543: \begin{equation}
544: F\Msub{m}(r) = \left[ 1 - e^{-2\tau\Msub{m}(T\Msub{m})} \right] \sigma
545: T\Msub{m}^4,
546: \label{EQ:Fm}
547: \end{equation}
548: where $\tau\Msub{m}(T\Msub{s})$ and $\tau\Msub{m}(T\Msub{m})$ are the
549: optical depths of the disk interior (from $z=0$ to $z=z\Msub{s}$) to
550: the radiation from the superheated dust grains and to its own
551: emission, respectively, and $A\Msub{s}$ is the total emitting-area
552: filling-factor of superheated dust grains. We consider the effect of
553: finite radius of the central star in $F\Msub{s}$, which is important
554: in the inner part of the disk. We also consider the effects of
555: oblique radiative transfer: the angular brackets in the right-hand
556: side of equation~(\ref{EQ:Fs}) represent radial averaging of radiation
557: emitting from superheated dust within the adjacent regions (see
558: \S~\ref{SEC:Results}). The factors 2 in the exponential functions in
559: equations~(\ref{EQ:Fs}) and (\ref{EQ:Fm}) also denote oblique
560: radiative transfer in the disk interior \citep{THI05}.
561:
562: Once the dust density distribution $\rho\Msub{d}(r,z)$ of the disk is
563: specified, $A\Msub{s}$ and $\tau\Msub{m}$ can be determined from the
564: following integration:
565: \begin{equation}
566: A\Msub{s}(r) = 1 - \exp \left[ - \int_{z\Msub{s}(r)}^{\infty}
567: \tilde{\kappa}\Msub{s} (T_\star)
568: \rho\Msub{d}(r,z') \, dz' \right],
569: \label{EQ:As}
570: \end{equation}
571: %
572: \begin{equation}
573: \tau\Msub{m}(T\Msub{rad}; r) = \int_0^{z\Msub{s}(r)}
574: \bar{\kappa}\Msub{m}(T\Msub{rad})
575: \rho\Msub{d}(r,z') \, dz',
576: \end{equation}
577: where $\bar{\kappa}\Msub{m}(T\Msub{rad})$ is the Planck mean
578: opacity of midplane grains interacting the blackbody radiation peaked
579: at temperature $T\Msub{rad}$. Further details of dust opacity are
580: shown in \S \ref{SEC:Results} and Appendix \ref{SEC:Opacity}.
581:
582: We assume that the total ($\mbox{gas} + \mbox{dust}$) surface density
583: $\Sigma$ of the disk is a simple power-law distribution in the radial
584: direction:
585: \begin{equation}
586: \Sigma = \Sigma_0 \left( \frac{r}{r_0} \right)^{-p},
587: \label{EQ:Sigma}
588: \end{equation}
589: which is kept constant during the thermal evolution considered in this
590: work.
591:
592: Taking the effects of dust settling, we can obtain dust density
593: distribution $\rho\Msub{d}$ of the disk in the two-temperature model
594: adopted here (see Appendix~\ref{SEC:DustDensity}). However,
595: computationally intense iterative calculation is needed to determine,
596: self-consistently, the magnitudes of $\rho\Msub{d}$ and $z\Msub{s}$
597: simultaneously. Instead of equations~(\ref{EQ:rhodm}) and
598: (\ref{EQ:rhods}), we adopt, in most of the calculations, the following
599: simple density distribution for the dust:
600: \begin{equation}
601: \rho\Msub{d} (r,z) = \frac{\Sigma\Msub{d}}{\sqrt{2 \pi} h}
602: \exp\left( -\frac{z^2}{2 h^2} \right),
603: \label{EQ:rhodsimple}
604: \end{equation}
605: where $\Sigma\Msub{d}$ is the surface density of dust. We put
606: $\Sigma\Msub{d} = f\Msub{d} \Sigma$, where $f\Msub{d}$ is the
607: dust fraction in the surface density. We dub $f\Msub{d}$ the dust-to-gas
608: ratio. In the disk interior, dust sedimentation is not so important unless
609: the radii of dust are not so large that we can formally set the dust
610: density distribution to be equation~(\ref{EQ:rhodsimple}). In the
611: surface layer, dust settling reduces the dust density while high
612: surface temperature $T\Msub{s}$ raise it, so that
613: equation~(\ref{EQ:rhodsimple}) also gives a good estimate. In some cases
614: we compared results with more realistic dust distribution given in
615: Appendix~\ref{SEC:DustDensity} and found that they are very similar
616: unless dust sizes are not so large. We discuss the difference of the
617: results between the two distributions in \S~\ref{SEC:Discuss}.
618:
619:
620: We found that if the gradient $d \ln z\Msub{s} / d \ln r$
621: changes rapidly in $r$ direction, the approximation used to derive
622: equation~(\ref{EQ:Astheta}) is no longer justified (see
623: Appendix~\ref{SEC:Estimate}). For this reason we use
624: equations~(\ref{EQ:zsdef}) and (\ref{EQ:As})
625: instead of equation~(\ref{EQ:Astheta}).
626:
627: \section{Numerical Results}
628: \label{SEC:Results}
629:
630: We adopt the following values as fixed parameters for all models: the
631: mass, radius, and luminosity of the central star are set to be
632: $M_\star = 1 M_\sun$, $R_\star = 2.085 R_\sun$, and $L_\star =
633: 1 L_\sun$, respectively, so that its effective temperature is $T_\star
634: = 4000 \Punit{K}$. For the disk gas, we specify $\mu = 2.34$ and
635: $\gamma\Msub{a}=1.4$.
636:
637: We adopted the phenomenological MMSN model \citep{H81} for the
638: standard gas and dust surface density
639: distribution. In this model, $\Sigma_0 = \Sigma\Msub{H0} = 1.7 \times
640: 10^3 \Punit{g} \Punit{cm}^{-2}$ with $r_0 = 1 \Punit{AU}$ and $p =
641: 1.5$ in equation~(\ref{EQ:Sigma}). For comparison,
642: we also calculate a relatively flat $\Sigma$ distribution with $p =
643: 1.0$ and $\Sigma_0 = 3.54 \times 10^2 \Punit{g} \Punit{cm}^{-2}$.
644: These surface densities are kept constant with time.
645: The dust-to-gas ratio $f\Msub{d}$
646: and solid material density $\rho\Msub{mat}$ are set to be 0.014 and
647: $1.4 \Punit{g} \Punit{cm}^{-3}$, respectively. In \S~\ref{SS:CO} we
648: neglect any changes of dust surface density due to sublimation
649: and use the value of $f\Msub{d}$ throughout the disk. In \S~\ref{SS:RO}
650: we consider the effect of ice sublimation. We also vary the
651: mass accretion rate from $\dot{M} = 0$ to $ 10^{-7} M_\sun
652: \Punit{yr}^{-1}$. These steady-state accretion rates are consistent
653: with our specified surface density and temperature distributions
654: provided the magnitude of $\alpha$ is a function of the radius,
655: where $\alpha$ is the non-dimensional turbulent viscosity in the
656: so-called $\alpha$-prescription \citep{SS73}.
657:
658: The normalization time unit is the thermal timescale given by
659: equation~(\ref{EQ:tth}) with $\Sigma = \Sigma_0$ and $T\Msub{m} =
660: T\Msub{m0}$. We denote the time unit as $t\Msub{th,0}$ and the
661: non-dimensional time as $\hat{t}=t/t\Msub{th,0}$. For the standard model, we set the
662: magnitude of $\Sigma_0 = \Sigma\Msub{H0}$ and $T_0 = 124 \Punit{K}$,
663: so that $t\Msub{th,0} = 53 \Punit{yr}$. Note that the local thermal
664: timescale $t\Msub{th}$ is nearly constant with $r$ in the standard
665: disk model with $p=1.5$. The non-dimensional time step used in the numerical
666: integration is set to be $\delta \hat{t} = 0.005$. In order to verify
667: numerical convergence, we also perform several calculations with time
668: step half of the standard value and confirmed that the results have no
669: significant changes.
670:
671: The spatial grid consists of 90 points (the standard case) or 180
672: points (the high-resolution case), logarithmically distributed,
673: between $r = 0.1 \Punit{AU}$ and $100 \Punit{AU}$. Numerical
674: oscillation would be induced if there is no radial exchange of energy
675: at all. At any radius, the disk interior is exposed not only to the
676: superheated surface grains directly overhead, but also obliquely to
677: those at adjacent radial locations. Such a radial exchange of energy
678: tends to suppresses instabilities for short-wavelength perturbations.
679: In our numerical scheme, we assume that the isotropic radiation comes
680: from all radius $r'$ within $|r'-r| < z\Msub{s}(r)$ contributes to the
681: heating at radius $r$ as expressed in equation~(\ref{EQ:Fs}). This
682: implementation stabilizes the short wave-length oscillation and the
683: results are essentially independent of the numerical resolution.
684:
685: \subsection{Constant Opacity}
686: \label{SS:CO}
687:
688: In this subsection, we first illustrate the dominant features using a
689: simple opacity model. We adopt the emissivity and opacity (per unit
690: {\it dust mass}) of the grains interacting with blackbody radiation peaked
691: at temperature $T_i$ to be
692: \begin{equation}
693: \epsilon\Msub{s}(T_i) = \left( \frac{T_i}{T_\star} \right)^\beta \mbox{ and }
694: \bar{\kappa}\Msub{s}(T_i) = \bar{\kappa}\Msub{s0}
695: \left( \frac{T_i}{T_\star} \right)^\beta,
696: \end{equation}
697: where we choose the value of $\bar{\kappa}\Msub{s0} = 10^2
698: \Punit{cm}^{2} \Punit{g}^{-1}$, which approximately corresponds to the
699: commonly defined opacity per unit {\it gas mass} with the value of $1
700: \Punit{cm}^{2} \Punit{g}^{-1}$. Most of the calculation shown here is
701: for $\beta = 0$ but we also calculate some models with $\beta = 1$ for
702: comparison purposes.
703:
704: For initial conditions, we adopt the steady state solution obtained
705: from the time integration with fixing $\chi(r)$. This set of initial
706: conditions does not correspond to the asymptotic steady-state
707: solutions because the initial estimate of $\chi$ is not
708: self-consistently compatible with the actual aspect ratio
709: $\zeta\Msub{s}$ of the surface.
710: Nevertheless, the numerical calculations relax to nearby steady
711: solutions if they exist. In order to verify that our results are
712: independent of the adopted initial conditions, we calculate the
713: evolution of the disk with several different initial guesses for
714: $\chi$ and found that the system reaches to the same asymptotic state.
715:
716: We first show the results of the calculations with no mass accretion
717: ($\dot{M} = 0$). Figure~\ref{FIG:init-ck-M0-T} shows the initial
718: evolution of the midplane temperature $T\Msub{m}$ as well as the
719: surface temperature $T\Msub{s}$, which is kept constant with time. The
720: elapsed time is $\hat{t} = 1.6$, i.e., $t = 1.6 t\Msub{th,0} \simeq 85
721: \Punit{yr}$ and each curve corresponds to a time step $\Delta \hat{t}
722: = 0.2$ ($\Delta t = 10.6 \Punit{yr}$). At first, the initial state is
723: almost stable in the innermost region and the outermost region. But,
724: in the intermediate regions (0.5--$20 \Punit{AU}$), the disk becomes
725: unstable. At a typical instance of time, four local temperature peaks
726: (high $T\Msub{m}$ and $z\Msub{s}$) coexist and are amplified. These
727: peaks move inward (toward the star) as they grow. Hence, we refer
728: these propagating perturbations as waves. During the amplification of
729: the waves, they cast shadows over the outer regions of the peak. The
730: temperature $T\Msub{m}$ decreases in the shadowed regions. Each
731: fully-grown wave has a sharp slope on the inner `exposed side' of the
732: peak and a gentler decline on the outer `shadow side'. The waves
733: propagate inward with velocities about a few tenth of $r/t\Msub{th}$.
734: The outermost wave ($\sim 16 \Punit{AU}$) begins to grow just outside
735: the shadowed region of the inner adjacent wave when the shadowed
736: region is developed. This tendency shows that the outermost wave may
737: be induced by the wave ahead of it.
738:
739: Additional time integration shows that as these waves propagate inward
740: they begin to decay when they reach inside $1 \Punit{AU}$. The waves
741: are completely damped out at around $0.25 \Punit{AU}$. In contrast, new
742: waves are formed continually at the outermost region ($> 20
743: \Punit{AU}$) of the computational domain. The growth region of waves
744: gradually retreats and the maximum amplitude of each wave during its
745: propagation cycle gradually increases with time. The system reaches a
746: quasi-periodic state when $\hat{t} \sim 8$. Figure~\ref{FIG:ck-M0-T}
747: shows the evolution of $T\Msub{m}$ at this stage. Waves are
748: continuously formed and amplified at the outer disk ($>
749: 30~\Punit{AU}$), then propagate inward with nearly constant amplitude
750: through the intermediate disk regions, begin to decay at about 1~AU,
751: and are damped out completely at around $0.25 \Punit{AU}$. Temperature
752: in the innermost disk region ($r < 0.25 \Punit{AU}$) attains steady
753: values. The propagation speed of the waves is approximately given by
754: $r/t\Msub{th}$. In the intermediate disk radii (1--$20 \Punit{AU}$),
755: the peak temperature of each wave is 2--3 times higher than the bottom
756: temperature in the inner adjacent shadowed region.
757:
758: The radial profile of each wave is somewhat skewed. The half width of
759: individual waves is about 0.1--$0.2r$ on the inner side and
760: 0.2--$0.4r$ on the outer side. The wavelength is approximately twice
761: as large as $z\Msub{s}(r)$. Due to the steep radial temperature
762: gradient, the magnitude of $z\Msub{s}(r)$ is affected by the variation
763: in the thickness $h$ at the disk regions interior to $r$ on this
764: length scale. The ratio of the radii between two adjacent wave peaks
765: are about 20 if both waves are within $r > 1 \Punit{AU}$. Near the
766: inner boundary of the propagating-wave zone, the finite size of the
767: star ($R_\star$) becomes comparable to the surface height $z\Msub{s}$.
768: This time-independent contribution in equation~(\ref{EQ:Fs})
769: essentially stabilizes the innermost region of the disk.
770:
771: The change of other variables at the same epoch as
772: Figure~\ref{FIG:ck-M0-T} are shown in
773: Figures~\ref{FIG:ck-M0-zeta}--\ref{FIG:ck-M0-Pv}.
774: Figure~\ref{FIG:ck-M0-zeta} shows the time
775: evolution of $\zeta\Msub{s} = z\Msub{s}/r$. Outside $0.25
776: \Punit{AU}$, $\zeta\Msub{s}$ changes stepwise. The two or three steep
777: jumps of $\zeta\Msub{s}$ correspond to the leading inner side of the
778: thermal waves. The magnitude of increase in $z\Msub{s}$ at the leading
779: edge of each step reaches a maximum of 1.5 around $10 \Punit{AU}$.
780: The flat portions of $\zeta\Msub{s}$ correspond to the shadowed
781: regions, where $\zeta\Msub{s}$ is determined by the stellar ray which
782: passes above the peak of each wave.
783:
784: Figure~\ref{FIG:ck-M0-zovh} shows the time evolution of $\chi =
785: z\Msub{s}/h = \zeta\Msub{s}/\zeta\Msub{h}$. The local minima of
786: $\chi$ are located just ahead of the peaks in $T\Msub{m}$. The
787: decrease of $\chi$ is essential for the temperature raise. In the
788: shadowed regions $\chi$ increases because $\zeta\Msub{s}$ is kept
789: almost constant (see Fig.~\ref{FIG:ck-M0-zeta}), whereas $\zeta\Msub{h}
790: = h/r$ decreases. Changes in the value of $\chi$ produces the
791: variations in the surface filling factor $A\Msub{s}$. The time
792: evolution of $A\Msub{s}$ is shown Fig.~\ref{FIG:ck-M0-As}. Sharp
793: peaks of $A\Msub{s}$, which correspond to the minima of $\chi$ (see
794: eq.~[\ref{EQ:Aschi}]), propagate inward. Note that the variation in
795: the amplitude of $A\Msub{s}$ is very large, ranging from a few tenth
796: at the peaks to below $10^{-3}$ in the shadowed region. Such large
797: changes of $A\Msub{s}$ induce rapid heating in front of the waves and
798: rapid cooling in the shadowed regions. The unperturbed $z\Msub{s}$ in
799: the intermediate regions is comparable to the half width of the
800: propagating waves. Modest variations in $\chi$ can lead to nonlinear
801: dissipation, such that the wave amplitudes are also limited in these
802: region.
803:
804: We also plot the evolution of the distribution of the logarithmic
805: pressure gradient $d {\rm ln} P/ d{\rm ln} r$
806: (Fig.~\ref{FIG:ck-M0-dP}), where $P(r)$ is the pressure in the
807: midplane of the disk. This plot indicates that the gas pressure
808: gradient is nearly reversed just in front of the peak of the waves.
809: This inversion occurs due to the steep positive temperature gradient
810: in the exposed, leading, inner face of the waves. Consequently, the
811: velocity of the gas departs significantly from its unperturbed
812: sub-Keplerian values. In a follow-up paper, we will consider the
813: associated gas drag on grains of various sizes. Another interesting
814: quantity is the distribution of the potential vorticity (or
815: vortensity):
816: \begin{equation}
817: \frac{\Omega\Msub{ep}}{\Sigma} = \frac{1}{\Sigma r^3}
818: \frac{d}{dr} r^4 \Omega^2,
819: \end{equation}
820: where $\Omega\Msub{ep}$ is the epicyclic frequency and $\Omega$ is the
821: angular velocity of gas. Note that in a quasi-Keplerian disk
822: $\Omega\Msub{ep} \simeq \Omega\Msub{K} \propto r^{-1.5}$, so that the
823: potential vorticity is almost constant for MMSN with $p=1.5$. The
824: thermal waves disturb the potential vorticity through the change of
825: pressure gradient. Figure~\ref{FIG:ck-M0-Pv} displays the evolution
826: of the distribution of the potential vorticity. There are peaks and
827: troughs around the waves. Local extrema of this quantity can lead
828: baroclinic instabilities which may excite turbulence
829: \citep[e.g.][]{KB03} in the dead zone where the magneto-rotational
830: instability may have limited influences \citep{G96}. Further
831: investigation of this possibility will also be considered elsewhere.
832:
833: Next, we examine the dependence of wave excitation and propagation on
834: the mass accretion rate ($\dot{M}$). Other than the value of
835: $\dot{M}$ in the energy equation, we adopt the same model parameters
836: as for the no-accretion case shown in
837: Figures~\ref{FIG:init-ck-M0-T}--\ref{FIG:ck-M0-Pv}. Thus, in the
838: present context, the primary physical effects associated with the
839: accretion flow is the viscous dissipation. This internal energy source
840: (viscous dissipation) in equation~(\ref{EQ:energy}) has a greater
841: fractional contribution to the energy budget in the inner regions than
842: the outer regions of the disk \citep{GL07}.
843:
844: Although quasi-periodic oscillations are excited in all cases, the
845: radial extent where they propagate to depends on the magnitude of
846: $\dot{M}$. Figure~\ref{FIG:ck-M8-T} shows the time evolution of
847: $T\Msub{m}$ in a quasi-periodic state for the case of $\dot{M} =
848: 10^{-8} M_\sun \Punit{yr}^{-1}$. The result is similar to that
849: obtained by neglecting the viscous dissipation. In this case, waves
850: are excited in the outer region, propagate inward and are damped out
851: in the innermost region. In comparison with the no-accretion model,
852: the innermost region in this case is hotter ($T\Msub{m}(0.1
853: \Punit{AU}) \sim 600 \Punit{K}$) with larger thickness. The {\it
854: time-independent} contribution of viscous heating to the energy
855: equation provides a stabilizing effect to a slightly larger radial
856: extent ($r < 0.5 \Punit{AU}$) compared with the no-accretion
857: case ($T\Msub{m}(0.1 \Punit{AU}) \sim 500 \Punit{K}$ and
858: $r < 0.25 \Punit{AU}$).
859:
860: The quasi-periodic state, however, is drastically changed in the
861: $\dot{M} = 1 \times 10^{-7} M_\sun \Punit{yr}^{-1}$ model
862: (Fig.~\ref{FIG:ck-M7-T}). The disk becomes stable interior to about
863: $6 \Punit{AU}$. In the outer region, there are two high-temperature
864: peaks, both oscillate quasi-periodically. The positions of the two
865: peaks do not coherently propagate inward as in the case of $\dot{M} =
866: 1 \times 10^{-8} M_\sun \Punit{yr}^{-1}$, but they fluctuate to and
867: fro around 12~AU and 40~AU, respectively. The ranges of temporal
868: temperature changes are less than 2~K over the disk, so that the
869: disk is regarded to be in an approximately steady state. This result
870: shows that the disk becomes stabilized as $\dot{M}$ increases.
871:
872: Finally, we show the dependence on the disk surface density $\Sigma$.
873: Note that $\Sigma$ affects the evolution not only through the optical
874: depth of the surface layer but through the thermal timescale. We
875: perform several calculations for $\Sigma_0 = 3.54 \times 10^2
876: \Punit{g} \Punit{cm}^{-2}$ and $p = 1.0$. Note that the time unit for
877: this case is $t\Msub{th,0} = 11.0 \Punit{yr}$. We calculate the disk
878: evolution for this surface density distribution with several values of
879: $\dot{M}$ and find that the results are quite similar to the case with
880: the standard surface density distribution. The quasi-periodic wave
881: solutions are obtained for small $\dot{M}$, whereas the disk becomes
882: nearly steady for large $\dot{M}$.
883:
884: Figure~\ref{FIG:ck-M8-p10-T} shows the temporal variation of
885: $T\Msub{m}$ in the quasi-periodic state for this prescribed
886: $\Sigma(r)$ distribution with $\dot{M} = 10^{-8} M_\sun
887: \Punit{yr}^{-1}$. Comparing to the case with $p=1.5$
888: (Fig.~\ref{FIG:ck-M8-T}), the wave propagating region is relatively
889: narrow, confined to the regions between 1~AU and 20~AU, and the
890: maximum amplitude of waves is also smaller. Outermost region where $r
891: > 20 \Punit{AU}$, the amplitudes of waves are limited and do not
892: exceed about 2~K. The amplitudes grow as waves propagate from 20~AU
893: to 8~AU, then they attain a nearly constant value from 8~AU to 1.5~AU.
894: The waves are finally damped at 1.5--1~AU. In an analogous inviscid
895: model (i.e., with an identically prescribed $\Sigma (r)$
896: distribution but without viscous dissipation), the result is quite
897: similar to that in Figure~\ref{FIG:ck-M8-p10-T} except that the wave
898: propagating region extends slightly closer to the star. The results
899: for the $\dot{M} = 10^{-7} M_\sun \Punit{yr}^{-1}$ model show that the
900: disk attains an approximately steady state with only very small fluctuations
901: (amplitude is less than a few tenth~K) remaining in the outermost
902: region.
903:
904: All the above results are for $\beta = 0$. We also calculate cases
905: with $\beta =1$. Other than a modification in the distribution of
906: $T\Msub{s}$ and $T\Msub{m}$, the time dependent nature of wave
907: excitation and propagation is essentially independent of the value of
908: $\beta$. In all cases, we find that the quasi-periodic nature of the
909: inwardly-propagating thermal waves is realized for the disk with
910: $\dot{M} \la 10^{-8} M_\sun \Punit{yr}^{-1}$. The basic features of
911: this state do not strongly depend on other parameters such as $p$ and
912: $\beta$. In order to verify the universality of these results, we
913: perform further calculations with more realistic opacities.
914:
915: \subsection{Realistic Opacities}
916: \label{SS:RO}
917:
918: In this subsection, we present the results based on models with more realistic
919: grain-opacity prescription given by \citet{THI05}. In this
920: prescription, grains are assumed to be consisted of a uniform mixture
921: of H$_2$O-ice, organics, olivine, pyroxene, metallic iron, and
922: troilite, the abundances of which are given by \citet{PHB94}. For
923: relatively high temperatures ($ T > 160 \Punit{K}$), we use the dust
924: opacity of grains without ice or organics. For the low-temperature
925: ($T < 160 \Punit{K}$) state, the opacity includes the contribution
926: from ice and organics grains. Based on the single-sized monochromatic
927: opacity table made by H.~Tanaka \citep{THI05}, we calculate the Planck
928: mean of the size-averaged opacities $\bar{\kappa}\Msub{s}(T\Msub{rad})$
929: and $\bar{\kappa}\Msub{m}(T\Msub{rad})$ with an assumed power-law
930: grain-size distribution $n(s) \propto s^{-3.5}$, which is truncated
931: for a minimum and maximum range of $s\Msub{min} = 0.1 \Punit{\micron}$
932: and $s\Msub{max} = 1 \Punit{mm}$, respectively (see Appendix \ref{SEC:Opacity}).
933:
934: For illustrative models, we adopt the standard surface-density
935: distribution of the MMSN model. For the
936: gas-to-dust ratio $f\Msub{d}$, we consider two limiting cases: 1) a
937: constant ratio $f\Msub{d} = 0.014$ and 2) a ratio which includes the
938: effect of ice sublimation. In the latter case, $f\Msub{d} = 0.0043$
939: for $T > 160 \Punit{K}$ and $f\Msub{d} = 0.014$ for $T < 160
940: \Punit{K}$. In the actual implementation, these values are smoothly
941: connected around the sublimation temperature by a hyperbolic tangent
942: function. Note that even in the constant-$f\Msub{d}$ cases, the
943: opacity undergoes a transition across $T=160 \Punit{K}$.
944:
945: We perform a set of calculations with this realistic opacity. Similar
946: to the previous models with constant opacity, the disk evolves into a
947: quasi-periodic state after about 10 times of the thermal timescale.
948: First, we show the results for the case with a constant gas-to-dust
949: ratio. Figure~\ref{FIG:vk-M8-T} shows the time evolution of
950: $T\Msub{m}$ in the asymptotic quasi-periodic state of the disk with
951: $\dot{M} = 1 \times 10^{-8} M_\sun \Punit{yr}^{-1}$. The surface
952: temperature $T\Msub{s}$ is also shown in Figure~\ref{FIG:vk-M8-T}.
953: From equations~(\ref{EQ:Ts}) and (\ref{EQ:epsilons}), we note that
954: $T\Msub{s}$ is independent of time. Comparing with the
955: constant-opacity case in which $\beta = 0$ (Fig.~\ref{FIG:ck-M8-T}),
956: $T\Msub{s}$ is everywhere larger for the realistic grain
957: opacity, because of the effect of superheating. There is a small jump
958: at the radius where opacity law changes
959: ($T\Msub{s} \simeq 160 \Punit{K}$). For the same value of
960: $\dot{M}$, the structure of wave propagation region is very similar
961: to that of the constant-opacity case.
962:
963: With a sufficiently large mass-accretion rate ($\dot{M} = 1 \times
964: 10^{-7} M_\sun \Punit{yr}^{-1}$), the disk with a realistic opacity is
965: also stabilized by the effect of viscous dissipation (Fig.~\ref{FIG:vk-M7-T}).
966: There are three local maxima in the $T\Msub{m}$ distribution. The
967: innermost peak at about $6 \Punit{AU}$ corresponds to an opacity
968: transition in the disk's surface layer ($T\Msub{s} \sim 160
969: \Punit{K}$). All these local peaks do not propagate but fluctuate
970: quasi-periodically with a small amplitude.
971:
972: Finally, we show the effect of ice sublimation.
973: Figure~\ref{FIG:vk-ev-M8-T} displays the evolution of $T\Msub{m}$ for
974: the model in which the dust surface density is modified by the ice
975: sublimation. Taking into account the effect of dust's size sorting on
976: the opacity, we assume that surface dust grains has a smaller maximum
977: size $s\Msub{s,min}$ than those in disk interior $s\Msub{m,min}
978: =s\Msub{min} =1 \Punit{mm}$. We choose
979: $s\Msub{s,min} = 1 \Punit{\micron}$.
980: Owing to the increase of $A\Msub{s}$ associated with the ice
981: condensation in surface layer, the $T\Msub{m}$ distribution attains a
982: local maximum near $5 \Punit{AU}$. This peak does not propagate in
983: time. Interior to this snow line, a wave propagates to about $0.5
984: \Punit{AU}$. Outside this snow line, there is another local maximum
985: near $20 \Punit{AU}$, which may be induced by the emergence of the
986: first peak. Compare to the model in which the ice sublimation is
987: neglected (Fig.~\ref{FIG:vk-M8-T}), the ice-condensation induces the
988: formation of a local thermal maximum which prevents waves from
989: emerging at large radii and propagate inward.
990:
991: \section{Summary and Discussion}
992: \label{SEC:Discuss}
993:
994: We have performed a set of radial one-dimensional calculations to
995: examine the thermal evolution of hydrostatic disks, using the direct
996: integration of optical depths $\tau\Msub{s}(T_\star)$ to determine the
997: optical surface $z\Msub{s}$ and total emitting area-filling factor
998: $A\Msub{s}$ of a superheated layer. Our results suggest that, in
999: regions with modest and steep radial temperature gradients, the
1000: constant $\chi = z\Msub{s}/h$ assumption is incompatible with the
1001: computed height of the surface where $\tau\Msub{s}(T_\star; r,
1002: z\Msub{s}) =1$. The initial state obtained by a fixed-$\chi$
1003: iteration evolves spontaneously to the state where thermal waves
1004: grows. The disks evolve to a quasi-periodic state where thermal waves
1005: continuously propagate toward the star in the intermediate radii.
1006:
1007: The driving mechanism for this thermal instability is the intense
1008: stellar irradiation high up in the disk atmosphere. It is a
1009: consequence of a ``shadowing effect'' in which the surface where most
1010: of the stellar photons are intercepted at any given radial location
1011: may be affected by the vertical structure in the disk regions interior
1012: to that radius. This quasi-periodic state is stabilized by viscous
1013: dissipation associated with the mass accretion flow through the disk.
1014: For the cases of $\dot{M} = 10^{-7} M_\sun \Punit{yr}^{-1}$ wave
1015: excitation and propagation are suppressed and the disks reach
1016: approximately steady states. In order to eliminate the possibility of
1017: artificial dependence on the initial conditions, we perform the
1018: following numerical experiments. By setting the initial condition for
1019: an approximately steady state with $\dot{M} = 10^{-7} M_\sun
1020: \Punit{yr}^{-1}$ and by decreasing the mass accretion rate to $\dot{M}
1021: = 10^{-8} M_\sun \Punit{yr}^{-1}$, we calculate the time evolution of
1022: the disk. We find that the system evolves to a quasi-periodic state
1023: within about 10 times of thermal timescale. The asymptotic
1024: quasi-periodic state is identical to that obtained with the standard
1025: calculation for $\dot{M} = 10^{-8} M_\sun \Punit{yr}^{-1}$, which is
1026: shown in Figure~\ref{FIG:ck-M8-T}. Inversely, we also start a
1027: calculation from a quasi-periodic state with $\dot{M} = 10^{-8} M_\sun
1028: \Punit{yr}^{-1}$ and increase the mass accretion rate to $\dot{M} =
1029: 10^{-7} M_\sun \Punit{yr}^{-1}$. The disk evolves into an
1030: approximately steady state, which is almost identical to that obtained
1031: with the standard calculation for $\dot{M} = 10^{-7} M_\sun
1032: \Punit{yr}^{-1}$, which is shown in Figure~\ref{FIG:ck-M7-T}. These
1033: results show that the state of the disks are determined by the
1034: instantaneous mass accretion rate (note that viscous evolution time is
1035: much longer than thermal timescale).
1036:
1037: The result that the quasi-periodic wave-propagating states exist in
1038: disks with modest disk accretion rate is robust to variations in the
1039: disks' structural parameters such as their surface density profile,
1040: opacity law, and vertical dust distribution. Whereas these parameters
1041: weakly affect the positions of the inner and outer boundaries of the
1042: wave-propagating domains, they do not affect the basic features of
1043: the waves such as their amplitude and shape.
1044:
1045: We also perform some calculations for the constant-opacity case using
1046: a general surface density profile (eq.~[\ref{EQ:rhods}] instead of
1047: eq.~[\ref{EQ:rhodsimple}]). The results are quite similar with that
1048: of the simple density profile. The difference is even less noticeable
1049: than that due to change of opacity law.
1050:
1051: Using the calculated thermal structure of the disk, we obtain the disk
1052: SED. We assume that the disk is face-on and
1053: truncated both at inner $0.1 \Punit{AU}$ and outer $100 \Punit{AU}$.
1054: Figure~\ref{FIG:vk-ev-M8-SED} shows the evolution of the SEDs
1055: associated with the model illustrated in Figure~\ref{FIG:vk-ev-M8-T}.
1056: Contribution from the central star to the SEDs is included in this
1057: figure. The SEDs are expected to oscillate periodically within
1058: mid-infrared wavelengths. This variation correspond to the periodic
1059: propagation of the thermal waves. In this case, the period of
1060: oscillation is about $ t\Msub{th,0} \simeq 53 \Punit{yr}$. In general
1061: this time scale varies depending on the surface density distribution
1062: and accretion rate in the disk as well as the luminosity of the
1063: central star. The change of disk SED comes from not only emission from
1064: the disk interior where $T\Msub{m}$ changes but also emission from the
1065: superheated surface layer where, even if $T\Msub{s}$ is kept constant,
1066: $A\Msub{s}$ would modulate. The surface emission also contributes to
1067: the water-ice and silicate emission bands. We expect the relative
1068: heights of these emission bands in SEDs to modulate with time. Note
1069: that this calculated SED is somewhat artificial because disk is
1070: truncated at the inner and outer edges. Thus, the direct comparison
1071: with observation may be meaningless. Nevertheless, the predicted
1072: relative variations of SEDs due to propagation of the thermal waves
1073: are likely to be important in the interpretation of the observed SEDs,
1074: especially those of the so-called transitional disks taken by {\it
1075: SPITZER} Infrared Spectrograph (IRS) \citep[e.g.,][]{FHC06}.
1076:
1077: In this work, we assume $t\Msub{th}$ to be much longer than
1078: $\Omega\Msub{K}^{-1}$, so that a hydrostatic equilibrium would be
1079: quickly re-established after the passage of the thermal waves. This
1080: assumption is invalid in the outer regions (say $> 20 \Punit{AU}$,
1081: see eq.~[\ref{EQ:tth}]). But our calculations show that the wave
1082: propagating region is extended to regions interior to $1 \Punit{AU}$.
1083: This result implies that even if dynamic effect may suppress the
1084: thermal waves in the outer disk, the waves can still be excited in
1085: the inner regions. According to previous linear analysis
1086: \citep{D00}, the disk becomes unstable to infinitesimal hydrodynamic
1087: perturbations when $t\Msub{th} \ll \Omega\Msub{K}^{-1}$. This type
1088: of instabilities may also excite the thermal waves.
1089:
1090: The results presented here is based on simple radial one-dimensional
1091: analysis in which a two-temperature approximation is adopted to
1092: describe the vertical structure of the disk. We also neglected
1093: modulations in the surface density and accretion rate throughout the
1094: disk. Our next task is to relax these assumptions and to generalize
1095: our results to a set of genuine two-dimensional simulations in which
1096: the radiation and mass transfer can be considered simultaneously.
1097: Since dust growth time or radial migration time are comparable to the
1098: thermal timescale, it will also be important to consider the evolution
1099: of the dust particles \citep[e.g.,][]{TL03, DD04b}. These
1100: investigations will be carried out in the future and presented
1101: elsewhere.
1102:
1103: \acknowledgments
1104:
1105: We wish to thank the anonymous referee for valuable comments.
1106: We thank to Dr. H.~Tanaka for providing us with the single-sized
1107: monochromatic opacity table used in \citet{THI05}. We are grateful to
1108: Drs P.~Garaud and K.~Kretke for useful conversation. This work is
1109: supported in part by NASA (NAGS5-11779, NNG04G-191G, NNG06-GH45G),
1110: JPL(1270927), NSF(AST-0507424, PHY99-0794), and Grant-in-Aid of the
1111: Japanese Ministry of Education, Science, and Culture (19540239).
1112:
1113: \appendix
1114:
1115: \section{Opacities}
1116: \label{SEC:Opacity}
1117:
1118: The Planck mean opacity is given by
1119: \begin{equation}
1120: \bar{{\kappa}}_j(T\Msub{rad}) = \frac{\int_0^\infty \tilde{\kappa}_\nu (T_j)
1121: B_\nu(T\Msub{rad}) \, d\nu}{\int_0^\infty B_\nu (T\Msub{rad}) \, d\nu},
1122: \end{equation}
1123: where subscript $j$ represents ``s'' (surface) or ``m'' (disk interior),
1124: $\nu$ is the frequency, $B_\nu (T\Msub{rad})$ is the Planck function, and
1125: \begin{equation}
1126: \tilde{\kappa}_\nu (T_j) = \frac{\int_{s\Msub{min}}^{s\Msub{max}}
1127: \kappa_\nu (T_j, s) s^3 n(s) \, ds}{\int_{s\Msub{min}}^{s\Msub{max}}
1128: s^3 n(s) \, ds}.
1129: \end{equation}
1130: Here $n(s) \, ds$ is the number density of grains with radii between
1131: $s$ and $s+ds$, and the size distribution has the lower cutoff
1132: $s\Msub{min}$ and the upper cutoff $s\Msub{max}$. In addition,
1133: $\kappa_\nu (T, s)$ is the single-sized ($s$) monochromatic ($\nu$)
1134: dust opacity of grains with temperature $T$.
1135:
1136: We assume a power-law size distribution $n(s) \propto s^{-3.5}$, which
1137: is truncated for a minimum and maximum range of $s\Msub{min} = 0.1
1138: \Punit{\micron}$ and $s\Msub{max} = 1 \Punit{mm}$.
1139:
1140: The grain compositions and optical constants are adopted from
1141: \citet{THI05} and references therein. We use the resultant opacity
1142: table given by H. Tanaka. The table gives two single-sized
1143: monochromatic dust opacities as functions of $\nu$ and $s$: one for
1144: grains without ice or organics at high temperatures of $ T > 160
1145: \Punit{K}$ and the other for grains including ice or organics at $T <
1146: 160 \Punit{K}$.
1147:
1148: \section{Exact Solution of Dust Density Distribution}
1149: \label{SEC:DustDensity}
1150:
1151: The gas density distribution $\rho\Msub{g}$ of the two-temperature
1152: model is given by
1153: \begin{equation}
1154: \rho\Msub{g} (r,z) = \left\{
1155: \begin{array}{ll}
1156: \rho\Msub{g}(r,0) \exp\left( -\frac{z^2}{2 h^2} \right)&
1157: \mbox{if $|z| \le z\Msub{s}$} \\
1158: \rho\Msub{g}(r,z\Msub{s}) \exp\left( -\frac{z^2 -
1159: z\Msub{s}^2}{2 H^2} \right)
1160: & \mbox{if $|z| \ge z\Msub{s}$,}
1161: \end{array}
1162: \right.
1163: \label{EQ:rhog}
1164: \end{equation}
1165: where $h = c\Msub{m} \Omega\Msub{K}^{-1}$ (see eq.~[\ref{EQ:h}]) and
1166: $H = c\Msub{s} \Omega\Msub{K}^{-1}$ ($c\Msub{s}$ is the sound speed in
1167: the surface layer) are the gas scale height in disk interior and in
1168: the surface layer, respectively.
1169:
1170: In a steady state, the sedimentation flux of the dust grains with
1171: their terminal velocity balances the diffusive mass flux due to gas
1172: turbulence \citep[see eq.~{[30]} in][]{TL02} such that:
1173: \begin{equation}
1174: -\rho\Msub{d} \Omega\Msub{K} \hat{t}\Msub{stop} z =
1175: \frac{\rho\Msub{g} \nu}{\Sc}
1176: \frac{\partial \left(\rho\Msub{d}/\rho\Msub{g} \right)}{\partial z},
1177: \label{EQ:diffusive}
1178: \end{equation}
1179: where $\hat{t}\Msub{stop}$ is the stopping time normalized by
1180: $\Omega\Msub{K}^{-1}$, $\nu$
1181: is the turbulent viscosity, and $\Sc$ is the Schmidt number
1182: representing coupling strength between grains and gas. The nondimensional
1183: stopping time $\hat{t}\Msub{stop}$ is given by
1184: \begin{equation}
1185: \hat{t}\Msub{stop} = \frac{\rho\Msub{mat} s \Omega\Msub{K}}{\rho\Msub{g}c\Msub{t}},
1186: \label{EQ:tstop}
1187: \end{equation}
1188: where $\rho\Msub{mat}$ is the material mass density, $s$ is the dust
1189: radius, and $c\Msub{t}$ is the mean thermal velocity.
1190:
1191: Solving equation~(\ref{EQ:diffusive}) with equations~(\ref{EQ:rhog})
1192: (in the case of $|z| \le z\Msub{s}$) and (\ref{EQ:tstop}), we obtain the
1193: dust density distribution of the disk interior to be \citep[see
1194: eq.~{[31]} in][]{TL02}
1195: \begin{equation}
1196: \rho\Msub{dm} (r,z) = \rho\Msub{dm}(r,0) \exp\left[ -\frac{z^2}{2 h^2}
1197: - \frac{\Sc \hat{t}\Msub{stop,m}}{\alpha}
1198: \left( \exp \frac{z^2}{2h^2} - 1 \right)
1199: \right],
1200: \label{EQ:rhodm}
1201: \end{equation}
1202: where $\hat{t}\Msub{stop,m}$ is the stopping time in the midplane
1203: and $\alpha = \nu/(c\Msub{m} h)$ is the non-dimensional
1204: turbulent viscosity in the so-called $\alpha$-prescription \citep{SS73}. The
1205: dimensionless stopping time in the midplane is given by
1206: \begin{equation}
1207: \hat{t}\Msub{stop,m} = \frac{\pi}{2} \frac{\rho\Msub{mat} s}{\Sigma}.
1208: \end{equation}
1209: The midplane dust density $\rho\Msub{dm}(r,0)$ is determined
1210: by the vertical integration of equation~(\ref{EQ:rhodm}) to be the
1211: surface density of dust $\Sigma\Msub{d} = f\Msub{d} \Sigma$, where
1212: $f\Msub{d}$ is the dust fraction in surface density.
1213:
1214: Solving equation~(\ref{EQ:diffusive}) with equations~(\ref{EQ:rhog})
1215: (in the case of $|z| \ge z\Msub{s}$) and (\ref{EQ:tstop}), we obtain
1216: the dust density distribution in the surface layer
1217: \begin{equation}
1218: \rho\Msub{ds} (r,z) =
1219: \rho\Msub{dm}(r,z\Msub{s})
1220: \exp\left[ -\frac{z^2- z\Msub{s}^2}{2 H^2}
1221: - \frac{\Sc \hat{t}\Msub{stop,m} h}{\alpha H}
1222: \exp \left( \frac{z\Msub{s}^2}{2 h^2} \right)
1223: \left( \exp \frac{z^2-z\Msub{s}^2}{2H^2} - 1 \right)
1224: \right],
1225: \label{EQ:rhods}
1226: \end{equation}
1227: where we put $\rho\Msub{ds}(r,z\Msub{s}) = \rho\Msub{dm}(r,z\Msub{s})$
1228: and assume that $\Sc$ and $\alpha$ in the surface layer have the same values
1229: as those in the disk interior.
1230:
1231: In the disk interior, dust sedimentation is not so important unless
1232: the radii of dust are not so large that we can formally set
1233: $\rho\Msub{dm} = \rho\Msub{d}$. In the surface layer, dust settling
1234: reduces the dust density while high surface temperature $T\Msub{s}$
1235: raise it, so that $\rho\Msub{d}$ also gives a good estimate. Thus,
1236: instead of equations~(\ref{EQ:rhodm}) and (\ref{EQ:rhods}), we adopt,
1237: in most of the calculations, the simple density distribution given by
1238: equation~(\ref{EQ:rhodsimple}).
1239:
1240: \section{The validity of the grazing-angle approximation}
1241: \label{SEC:Estimate}
1242:
1243: We derive the grazing-angle
1244: approximation (eq.~[\ref{EQ:Astheta}]) and discuss its validity.
1245: Substituting $\rho\Msub{d} (r,z)$ in equation~(\ref{EQ:As}) with that
1246: in equation~(\ref{EQ:rhodsimple}) and assuming $A\Msub{s} \ll 1$, we
1247: obtain
1248: \begin{equation}
1249: A\Msub{s}
1250: = \tau\Msub{v} \mbox{ erfc} \left( \frac{z\Msub{s}}{\sqrt{2} h} \right)
1251: = \tau\Msub{v} \mbox{ erfc} \left( \frac{\chi}{\sqrt{2}} \right),
1252: \label{EQ:Aschi}
1253: \end{equation}
1254: where $\tau\Msub{v} = \tilde{\kappa}\Msub{s}(T_\star) \Sigma\Msub{d}/2$ and
1255: $\chi = z\Msub{s}/h$.
1256:
1257: We derive the relation of $A\Msub{s}$ with a path integration from
1258: the star to the point $(r,z)$. Using equations~(\ref{EQ:taus}) and
1259: (\ref{EQ:rhodsimple}), equation~(\ref{EQ:zsdef}) can be written as
1260: \begin{equation}
1261: \tau\Msub{s}(T_\star; r, z\Msub{s}(r))
1262: = \tau\Msub{v} \left( 1+\zeta\Msub{s}^2 \right)^{1/2}
1263: \sqrt{\frac{2}{\pi}} \int_{R_\star}^r \frac{\hat{\Sigma}\Msub{d}'}{h(r')}
1264: e^{-\chi'^2/2} \, dr' = 1
1265: \end{equation}
1266: with $\zeta\Msub{s} = z\Msub{s}/r$, $\hat{\Sigma}\Msub{d}' =
1267: \Sigma\Msub{d}(r')/\Sigma\Msub{d}(r)$, and $\chi' = \zeta\Msub{s} r'/h(r')$.
1268: Here we assume that $\bar{\kappa}\Msub{s}(T_\star)$ is independent of
1269: position of the path.
1270: Substituting the integral variable from $r'$ to $\chi'$ and noting that
1271: $\chi' \gg 1$ for $r' = R_\star$, we obtain
1272: \begin{equation}
1273: \tau\Msub{v} \left( 1+\zeta\Msub{s}^2 \right)^{1/2} \zeta\Msub{s}^{-1}
1274: \sqrt{\frac{2}{\pi}} \int_{\chi}^\infty \hat{\Sigma}\Msub{d}'
1275: \left( \frac{d \ln \zeta'_h}{d \ln r'} \right)^{-1} e^{-\chi'^2/2} \,
1276: d\chi' = 1,
1277: \label{EQ:taucond}
1278: \end{equation}
1279: where we define $\zeta_h=h/r$ and $\zeta'_h = h(r')/r'$.
1280: Here we assume $\chi'$ is a monotonically increasing function with $r'$.
1281: From equations~(\ref{EQ:Aschi}) and (\ref{EQ:taucond}), we obtain
1282: \begin{equation}
1283: A\Msub{s} =
1284: \left( 1 + \zeta\Msub{s}^2\right)^{-1/2} \zeta\Msub{s}
1285: \left\langle \hat{\Sigma}\Msub{d}' \left(
1286: \frac{d \ln \zeta'_h}{d \ln r'}
1287: \right)^{-1} \right\rangle_{g(\chi'),\chi}^{-1}
1288: \end{equation}
1289: with
1290: \begin{equation}
1291: g(\chi') = \left[ \mbox{erfc } \left( \chi/\sqrt{2} \right) \right]^{-1}
1292: \sqrt{2/\pi} \exp \left(-\chi'^2/2 \right),
1293: \end{equation}
1294: where $\langle X \rangle_{f(x'),x} = \int_x^\infty X(x') f(x')\, dx'$
1295: represents the weighted average with a weight function $f(x')$.
1296:
1297: If we substitute the Gaussian weight $g(\chi')$ with the delta function
1298: $\delta(\chi'-\chi)$ and assume $\zeta\Msub{s} \ll 1$, we obtain
1299: \begin{equation}
1300: A\Msub{s} = \zeta\Msub{s}
1301: \left( \frac{d \ln \zeta_h}{d \ln r} \right).
1302: \label{EQ:Asaprox}
1303: \end{equation}
1304: Except for a small difference, this equation corresponds to
1305: equation~(\ref{EQ:Astheta}).
1306: However, if the gradient $d \ln \zeta_h / d \ln r$
1307: changes rapidly in $r$ direction, the approximation used to derive
1308: equation~(\ref{EQ:Asaprox}) is no longer justified.
1309: In this case $A\Msub{s}$ is determined not only by the local gradient
1310: of $\zeta_h$ but by the gradients in inner regions
1311: because the weight function $g(\chi')$ is extended to the inner radii.
1312: For this reason we use equations~(\ref{EQ:zsdef}) and (\ref{EQ:As})
1313: instead of equation~(\ref{EQ:Asaprox}).
1314:
1315: \begin{thebibliography}{}
1316: %
1317: \bibitem[Adams, Lada, \& Shu(1987)]{ALS87} Adams, F. C., Lada, C. J., \&
1318: Shu, F. H. 1988, \apj, 312, 788
1319: \bibitem[Chiang \& Goldreich(1997)]{CG97} Chiang, E. I., \&
1320: Goldreich, P. 1997, \apj, 490, 368
1321: \bibitem[Chiang et al.(2001)]{CJC01} Chiang, E. I., Joung, M. K.,
1322: Creech-Eakman, M. J., Qi, C., Kessler, J. E., Blake, G. A.,
1323: \& van Dishoeck, E. F. 2001, \apj, 547, 1077
1324: \bibitem[D'Alessio et al.(1999)]{DCH99} D'Alessio, P.,
1325: Cant\'{o}, J., Hartmann, L., Calvet N., \& Lizano, S. 1999, \apj,
1326: 511, 896
1327: \bibitem[D'Alessio et al.(2006)]{DCHFS06} D'Alessio, P., Calvet N.,
1328: Hartmann, L., Franco-Hern\'{a}ndez \& Serv\'{i}n, H. 2006, \apj,
1329: 638, 314
1330: \bibitem[Dullemond(2000)]{D00} Dullemond, C. P. 2000, \aap,
1331: 361, L17
1332: \bibitem[Dullemond \& Dominik(2004a)]{DD04a} Dullemond, C. P. \&
1333: Dominik, C. 2004a, \aap, 417, 159
1334: \bibitem[Dullemond \& Dominik(2004b)]{DD04b} Dullemond, C. P. \&
1335: Dominik, C. 2004b, \aap, 421, 1075
1336: \bibitem[Dullemond et al.(2001)]{DDN01} Dullemond, C. P., Dominik, C.,
1337: \& Natta, A. 2001, \apj, 560, 957
1338: \bibitem[Furlan et al.(2006)]{FHC06} Furlan, E., et al.
1339: 2006 \apjs, 165, 568
1340: \bibitem[Gammie(1996)]{G96} Gammie, C. F. 1996, \apj, 457, 355
1341: \bibitem[Garaud \& Lin(2007)]{GL07} Garaud, P., \&
1342: Lin, D. N. C. 2007, \apj, 654, 606
1343: \bibitem[Hayashi(1981)]{H81} Hayashi, C. 1981, Prog.~Theor.~Phys.~Suppl.,
1344: 70, 35
1345: \bibitem[Kenyon \& Hartmann(1987)]{KH87} Kenyon, S. J., \& Hartmann, L.
1346: 1987, \apj, 323, 714
1347: \bibitem[Klahr \& Bodenheimer(2003)]{KB03} Klahr, H. H. \& Bodenheimer,
1348: P. 2003, \apj, 582, 869
1349: \bibitem[Kusaka et al.(1970)]{KNH70} Kusaka, T., Nakano, T., \&
1350: Hayashi, C. 1970, Prog.~Theor.~Phys.~Suppl., 44, 1580
1351: \bibitem[Pollack et al.(1994)]{PHB94} Pollack, J. B., Hollenbach, D.,
1352: Beckwith, S., Simonelli, D. P., Roush, T., \&
1353: Fong, W. 1994, \apj, 421, 615
1354: \bibitem[Shakura \& Sunyaev(1973)]{SS73} Shakura, N. I., \&
1355: Sunyaev, R. A. 1973, \aap, 24, 337
1356: \bibitem[Takeuchi \& Lin(2002)]{TL02} Takeuchi, T., \&
1357: Lin, D. N. C. 2002, \apj, 581, 1344
1358: \bibitem[Takeuchi \& Lin(2003)]{TL03} Takeuchi, T., \&
1359: Lin, D. N. C. 2003, \apj, 593, 524
1360: \bibitem[Tanaka et al.(2005)]{THI05} Tanaka, H., Himeno, Y., \&
1361: Ida, S. 2005, \apj, 625, 414
1362: \bibitem[Watanabe et al.(1990)]{WNN90} Watanabe, S.,
1363: Nakagawa, Y., \& Nakazawa, K. 1990, \apj, 358, 282
1364: \end{thebibliography}
1365:
1366: \clearpage
1367:
1368: \begin{figure}
1369: \plotone{f1.ps}
1370: \caption{Disk thermal emission $F\Msub{m}$ ({\it solid line}) and intercepted
1371: stellar fluxes $F\Msub{s}$ at $r = 10 \Punit{AU}$ with constant surface heights
1372: ({\it dashed curves}) of $z\Msub{s}/r = 0.12$, $0.13$, and $0.14$
1373: ({\it left to right}) and with constant $\chi = z\Msub{s}/h$ ({\it dot-dashed line})
1374: as functions of disk temperature $T\Msub{m}$. All fluxes are normalized
1375: by $\sigma T_\star^4$.
1376: \label{FIG:FmFs}}
1377: \end{figure}
1378:
1379: \begin{figure}
1380: \plotone{f2.ps}
1381: \caption{Initial stage of evolution of the midplane temperature
1382: $T\Msub{m}$ for the case of constant opacity and $\beta=0$. The thick
1383: curves cover the range $\hat{t} = t/t\Msub{th,0} = 0.0$--$1.6$ with the
1384: interval $\Delta \hat{t} = 0.2$. The surface temperature $T\Msub{s}$
1385: is represented by a thin line in the back panel. The surface density
1386: distribution is that prescribed by the MMSN model in which $p=1.5$.
1387: The viscous dissipation associated with mass accretion is neglected
1388: (i.e., $\dot{M} = 0$).
1389: \label{FIG:init-ck-M0-T}}
1390: \end{figure}
1391:
1392: \begin{figure}
1393: \plotone{f3.ps}
1394: \caption{Time evolution of $T\Msub{m}$ after it has reached a
1395: quasi-periodic state. This model is the continuation of that shown in
1396: Fig.~\ref{FIG:init-ck-M0-T} to an epoch $\hat{t} = 8.0$--$9.6$. Each
1397: curve is separated by $\Delta \hat{t} = 0.2$. The $T\Msub{s}$
1398: distribution is also shown in a thin line in the back panel. Other
1399: parameters are same as Fig.~\ref{FIG:init-ck-M0-T}.
1400: \label{FIG:ck-M0-T}}
1401: \end{figure}
1402:
1403: \begin{figure}
1404: \plotone{f4.ps}
1405: \caption{Time evolution of $\zeta\Msub{s} = z\Msub{s}/r$ in the
1406: quasi-periodic state at the same epoch as Fig.~\ref{FIG:ck-M0-T}.
1407: \label{FIG:ck-M0-zeta}}
1408: \end{figure}
1409:
1410: \begin{figure}
1411: \plotone{f5.ps}
1412: \caption{Time evolution of $\chi = z\Msub{s}/h$ in the
1413: quasi-periodic state at the same epoch as Fig.~\ref{FIG:ck-M0-T}.
1414: \label{FIG:ck-M0-zovh}}
1415: \end{figure}
1416:
1417: \begin{figure}
1418: \plotone{f6.ps}
1419: \caption{Time evolution of surface filling factor $A\Msub{s}$ in
1420: quasi-periodic state at the same epoch as Fig.~\ref{FIG:ck-M0-T}.
1421: \label{FIG:ck-M0-As}}
1422: \end{figure}
1423:
1424: \begin{figure}
1425: \plotone{f7.ps}
1426: \caption{Time evolution of the logarithmic pressure gradient
1427: $d \ln P/d \ln r$ in quasi-periodic state at the same epoch as
1428: Fig.~\ref{FIG:ck-M0-T}.
1429: \label{FIG:ck-M0-dP}}
1430: \end{figure}
1431:
1432: \begin{figure}
1433: \plotone{f8.ps}
1434: \caption{Time evolution of potential vorticity $\Omega\Msub{ep}
1435: \Sigma^{-1}$ normalized by the Keplerian value at $1 \Punit{AU}$
1436: ($\Omega\Msub{K0} \Sigma_0^{-1}$) in quasi-periodic state at the same
1437: epoch as Fig.~\ref{FIG:ck-M0-T}.
1438: \label{FIG:ck-M0-Pv}}
1439: \end{figure}
1440:
1441: \begin{figure}
1442: \plotone{f9.ps}
1443: \caption{Quasi-periodic evolution of $T\Msub{m}$ for steady
1444: accretion disk with $\dot{M} = 10^{-8} M_\sun \Punit{yr}^{-1}$ (at
1445: $\hat{t} = 14.4$--$16.0$ with $\Delta \hat{t} = 0.2$). The
1446: distribution of $T\Msub{s}$ is also shown with a thin solid line in
1447: the back panel. Other parameters are same as
1448: Fig.~\ref{FIG:ck-M0-T}.
1449: \label{FIG:ck-M8-T}}
1450: \end{figure}
1451:
1452: \begin{figure}
1453: \plotone{f10.ps}
1454: \caption{Quasi-periodic evolution of $T\Msub{m}$ for steady
1455: accretion disk with $\dot{M} = 10^{-7} M_\sun \Punit{yr}^{-1}$ (
1456: $\hat{t} = 11.2$--$14.4$ with $\Delta \hat{t} = 0.4$). The
1457: distribution $T\Msub{s}$ is also shown with a thin solid line in the
1458: back panel. Other parameters are same as Fig.~\ref{FIG:ck-M0-T}.
1459: \label{FIG:ck-M7-T}}
1460: \end{figure}
1461:
1462: \begin{figure}
1463: \plotone{f11.ps}
1464: \caption{Quasi-periodic evolution of $T\Msub{m}$ for a steady
1465: accretion disk with $p = 1.0$ and $\dot{M} = 10^{-8} M_\sun
1466: \Punit{yr}^{-1}$ ($\hat{t} = 9.6$--$12.8$ with $\Delta \hat{t} =
1467: 0.4$). The distribution of $T\Msub{s}$ is also shown with a thin
1468: solid line in the back panel. Opacities are same as
1469: Fig.~\ref{FIG:ck-M0-T}.
1470: \label{FIG:ck-M8-p10-T}}
1471: \end{figure}
1472:
1473: \begin{figure}
1474: \plotone{f12.ps}
1475: \caption{Quasi-periodic evolution of $T\Msub{m}$ for realistic
1476: opacities as well as $T\Msub{s}$ (a thin solid line in the back
1477: panel). The disk structural parameters is that of a MMSN model
1478: ($p=1.5$) with steady $\dot{M} = 10^{-8} M_\sun \Punit{yr}^{-1}$. The
1479: opacity law changes at $160 \Punit{K}$ but the effect of ice
1480: sublimation is neglected. Maximum size of surface dust grains is
1481: assumed to be $1 \Punit{mm}$ (same as that in disk interior). Times
1482: are same as Fig.~\ref{FIG:ck-M0-T}.
1483: \label{FIG:vk-M8-T}}
1484: \end{figure}
1485:
1486: \begin{figure}
1487: \plotone{f13.ps}
1488: \caption{Same as Fig.~\ref{FIG:vk-M8-T}, but for a disk with
1489: $\dot{M} = 10^{-7} M_\sun \Punit{yr}^{-1}$ and $\hat{t} = 8.0$--$11.2$
1490: with $\Delta \hat{t} = 0.4$.
1491: \label{FIG:vk-M7-T}}
1492: \end{figure}
1493:
1494: \begin{figure}
1495: \plotone{f14.ps}
1496: \caption{Same as Fig.~\ref{FIG:vk-M8-T}, but taking into account the
1497: effect of ice sublimation. The maximum size of surface dust grains is
1498: reduced to $1 \Punit{\micron}$.
1499: \label{FIG:vk-ev-M8-T}}
1500: \end{figure}
1501:
1502: \begin{figure}
1503: \plotone{f15.ps}
1504: \caption{Quasi-periodic change of SED calculated from the model
1505: shown in Fig.~\ref{FIG:vk-ev-M8-T}. Different lines represent the
1506: epochs $\hat{t} = 8.4$ ({\it thick solid curve}), $8.6$ ({\it thin solid curve}),
1507: $8.8$ ({\it dashed curve}), $9.0$ ({\it dash-dotted curve}),
1508: $9.2$ ({\it dotted curve}), and $9.4$ ({\it thick solid curve}, coincide with
1509: that at $\hat{t}=8.4$), respectively. Note that $t = 53 \hat{t} \Punit{yr}$.
1510: \label{FIG:vk-ev-M8-SED}}
1511: \end{figure}
1512:
1513: \end{document}
1514:
1515: %%
1516: %% End of file `mssei.tex'.
1517: