astro-ph0304330/ms.tex
1: \documentclass[12pt,preprint]{aastex}
2: 
3: \usepackage{natbib}
4: \usepackage{epsfig}
5: 
6: \newcommand{\rhill}{r_{\mathrm{H}}}
7: \newcommand{\msun}{\mathrm{M}_{\sun}}
8: \newcommand{\mearth}{\mathrm{M}_{\oplus}}
9: \newcommand{\AU}{\:\mathrm{AU}}
10: \newcommand{\figref}[1]{Figure \ref{#1}}
11: \newcommand{\eqref}[1]{equation (\ref{#1})}
12: \newcommand{\Eqref}[1]{Equation (\ref{#1})}
13: 
14: \shorttitle{Radiative Transfer in Protoplanetary Disks}
15: \shortauthors{Jang-Condell \& Sasselov}
16: %\slugcomment{DRAFT \today}
17: \slugcomment{submitted to ApJ February 18, 2003 -- revised \today}
18: 
19: \begin{document}
20: \newcommand{\etal}{et al.~}
21: \newcommand{\eg}{e.g.~}
22: \title{Radiative Transfer on Perturbations in Protoplanetary Disks}
23: \author{Hannah Jang-Condell\altaffilmark{1}
24: \and Dimitar D.~Sasselov\altaffilmark{2}}
25: \affil{Harvard-Smithsonian Center for Astrophysics}
26: \affil{60 Garden St., Cambridge, MA 02138}
27: \altaffiltext{1}{hjang@cfa.harvard.edu}
28: \altaffiltext{2}{dsasselov@cfa.harvard.edu}
29: 
30: \newcommand{\tunpert}{133.6\:\mathrm{K}}	
31: % 1:       130.683      135.475      133.596      135.475
32: \newcommand{\tsmmin}{130.7\:\mathrm{K}}
33: \newcommand{\tsmmax}{135.5\:\mathrm{K}}
34: \newcommand{\tsmd}{5\:\mathrm{K}}
35: \newcommand{\losm}{2.2}
36: \newcommand{\hism}{1.4}
37: \newcommand{\dsm}{4}
38: % 10:      123.678      142.308      133.596      157.056
39: \newcommand{\tlgmin}{123.7\:\mathrm{K}}
40: \newcommand{\tlgmax}{142.3\:\mathrm{K}}
41: \newcommand{\tlgd}{19\:\mathrm{K}}
42: \newcommand{\lolg}{7.4}
43: \newcommand{\hilg}{6.5}
44: \newcommand{\dlg}{14}
45: % 20:      122.036      146.574      133.596      153.426
46: \newcommand{\txlmin}{122.0\:\mathrm{K}}
47: \newcommand{\txlmax}{146.6\:\mathrm{K}}
48: \newcommand{\txld}{25\:\mathrm{K}}
49: \newcommand{\hixl}{8.7}
50: \newcommand{\loxl}{9.7}
51: \newcommand{\dxl}{18}
52: 
53: \begin{abstract}
54: We present a method for calculating the radiative 
55: tranfer on a protoplanetary disk perturbed by a protoplanet.  
56: We apply this method to determine the effect on the temperature 
57: structure within the photosphere of a passive circumstellar disk 
58: in the vicinity of a small protoplanet of up to 20 Earth masses.  
59: The gravitational potential of a protoplanet induces a compression 
60: of the disk material near it, resulting in a decrement in the density 
61: at the disk's surface.  Thus, an isodensity contour at the height 
62: of the photosphere takes on the shape of a well.  When such a well is 
63: illuminated by stellar irradiation at grazing incidence, 
64: it results in cooling in a shadowed 
65: region and heating in an exposed region.  For typical stellar and 
66: disk parameters relevant to the epoch of planet formation, 
67: we find that the temperature variation due to a protoplanet at 
68: 1 AU separation from its parent star is about \dsm\% ($\tsmd$) for a planet 
69: of 1 Earth mass, about \dlg\% ($\tlgd$) for planet of 10 Earth masses, 
70: and about \dxl\% ($\txld$) for planet of 20 Earth masses, 
71: We conclude that even such relatively small protoplanets can induce 
72: temperature variations in a passive disk.  Therefore, many of the processes 
73: involved in planet formation should not be modeled with a locally 
74: isothermal equation of state.
75: \end{abstract}
76: 
77: \keywords{planetary systems: protoplanetary disks --- 
78: planetary systems: formation --- radiative transfer}
79: 
80: \section{Introduction}
81: 
82: Planetary systems are formed from rotating protoplanetary disks,
83: which are often modeled
84: in steady-state, in vertical hydrostatic equilibrium, with gas and 
85: dust fully mixed and thermally coupled \citep{kenyon}.  
86: Such disks describe very well the observed properties of
87: T Tauri disks of age $\sim$1~Myr and typical mass accretion rates
88: of $\leq 10^{-8} M_{\odot} {\rm yr}^{-1}$ \citep{hartmann,vertstruct}. 
89: They are passive disks in the sense modeled
90: by \eg \citet{CG}, where the main source of photospheric heating 
91: is irradiation from the parent star, although viscous heating is 
92: still important at small stellar distances and near the midplane.  
93: The temperature in such disks
94: is computed under the assumption that the upper surface of the disk
95: is perfectly concave and smooth at all radii, which is a very good
96: description of an unperturbed disk because thermal and
97: gravitational instabilities are damped very efficiently \citep{thermstab}.
98: 
99: We are interested in planet formation inside such disks, which
100: means that we might be forced to abandon the above assumption of
101: a perfectly smooth surface. The newly formed planet core will
102: distort it, affect the heating and cooling of the disk
103: locally, and could have significant consequences for the further
104: growth and migration of protoplanets. As discussed in \citet{sas_lec}, 
105: the distortion need only be large enough compared to
106: the grazing angle at which the starlight strikes the disk. This
107: small angle has a minimum
108: at $0.4\AU$ and increases significantly only at very large distances.
109: The depth of the depression due to the additional mass of the planet, $m_p$,
110: will be proportional to $(\rhill/h)^3$, where $\rhill$ is the Hill radius,
111: and $h$ is the local scale height, with
112: a shade area dependent on the grazing angle \citep{sas_lec}.
113: 
114: Generally planet formation has been treated numerically in 2-dimensional
115: disks; to date, there is only one recent study of the 3-dimensional
116: effect due to planets embedded in a protoplanetary disk \citep{bate}.
117: Bate \etal model the 3-dimensional response of a gaseous viscous disk
118: with a simple locally isothermal equation of state, without taking 
119: into account irradiation from the central star.  
120: 
121: In this paper we study in detail the development of a compression
122: in a standard passive disk and the radiative transfer in it. 
123: We consider this to be the first step in developing a global
124: 3-dimensional simulation, similar to the one by \citet{bate},
125: but with a realistic equation of state, especially near and inside
126: the protoplanet's Roche lobe.
127: We are explicitly interested in the planetary growth process before a 
128: gap in the disk is formed, hence our approach is limited to 
129: $m_p\lesssim30\:\mearth$, which results in a small perturbation to the 
130: irradiated surfaces of a passive disk.
131: 
132: \section{Heating Sources}
133: 
134: A schematic of our adopted disk model is shown in \figref{flareddisk}.  
135: The disk is flared, so that disk height increases faster than 
136: distance to the star, consistent with 
137: observational evidence and theoretical models
138: \citep{kenyon,CG,vertstruct}.  
139: We assume that the dust is well mixed with the gas 
140: and is the primary source of 
141: opacity at wavelengths characteristic of stellar emission.  
142: The dust absorbs incident stellar radiation at the surface 
143: and re-emits at longer wavelength, where the opacity is much lower.  
144: We are interested in the region of the disk that is optically thick, 
145: at distances of $\lesssim10$ AU \citep{dalessio2}.  
146: This gives rise to three different layers with smooth transitions 
147: in the disk: 
148: the uppermost is optically thin to both stellar and disk radiation, 
149: the interior is optically thick to both stellar and disk radiation, 
150: and the middle layer is a transition region that is optically thick 
151: to stellar radiation, but still optically thin to disk radiation.  
152: We shall refer to this optically thick/thin region as the photosphere.  
153: The reprocessing of stellar radiation in this region 
154: is important for determining the interior temperature.  
155: 
156: \citet{CG} calculate the temperature of only the thin/thin and
157: thick/thick regions using energy conservation considerations, without
158: detailing the temperature structure of the photosphere.
159: \citeauthor{vertstruct} (\citeyear{vertstruct},\citeyear{dalessio2})
160: calculate the detailed temperature structure of the
161: photosphere in disks around young stars in a self-consistent way,
162: including radiative transport and convection. 
163: They find that radiative transfer is the dominant mechanism for energy 
164: transport in the disk photosphere, and the primary heating sources are 
165: stellar irradiation and viscous
166: heating, as shown by comparison to \citet{calvet}, who use the
167: Milne-Strittematter treatment of the superposition of solutions
168: \citep{milne,strittmatter}.  We shall consider stellar radiation and
169: viscous heating to be the primary sources of heating, where the total
170: temperature is arrived at by calculating the sum of the viscous and
171: radiative fluxes, as \(T^4 = T_v^4 + T_r^4\), where $T_v$ is
172: the viscous temperature and $T_r$ is stellar radiation heating
173: temperature.  
174: 
175: \subsection{Viscous Heating}
176: 
177: The temperature due to viscous heating for a constant flux, gray atmosphere is 
178: \begin{equation}
179: T_v^4 = \frac{3F_v}{4\sigma_B}(\tau_d+2/3)
180: \end{equation}
181: where $\sigma_B$ is the Stefan-Boltzmann constant, 
182: $\tau_d$ is the optical depth of the disk's own radiation,
183: and the viscous flux $F_v$ at a distance $a$, for a star of effective 
184: temperature $T_{\star}$, mass $M_{\star}$, and radius $R_{\star}$, 
185: accreting at a rate $\dot{M}_a$ is 
186: \begin{equation}
187: F_v = \frac{3GM_{\star}\dot{M}_a}{4\pi a^3}
188: 	\left[1-\left(\frac{R_{\star}}{a}\right)^{1/2}\right]
189: \end{equation}
190: \citep{pringle}.
191: 
192: 
193: \subsection{Stellar Irradiation}
194: 
195: We shall use a modified form of the Milne-Strittmatter
196: treatment to calculate the effect of heating by stellar irradiation 
197: on a three-dimensional
198: perturbation on a passive accretion disk.
199: We assume that the dust is well-mixed with the gas in the disk and that 
200: they are thermodynamically coupled so that the dust temperature 
201: and gas temperature are the same.  In general, there is no guarantee
202: that these two temperatures are the same, or that the dust is either 
203: well mixed or uniform in composition.
204: 
205: Following the Milne-Strittmatter treatment, the incident radiation 
206: from the star is considered to be at some characteristic wavelength, 
207: typically in the visible, and the radiation emitted by the 
208: disk will be at a longer characteristic wavelength, typically 
209: in the infrared.  
210: The incident, short wavelength radiation will be indicated by the 
211: subscript $s$ (stellar or scattered), and the re-emitted, long wavelength 
212: radiation will be indicated by the subscript $d$ (disk or diffuse).
213: Some fraction of the energy of the incident stellar 
214: radiation, $\sigma$, is scattered 
215: at the same frequency.  The remaining fraction, $\alpha = 1-\sigma$, is 
216: absorbed by dust in the disk and re-emitted at a longer characteristic 
217: wavelength, typically in the infrared.  This gives the 
218: equation of radiative transfer for the 
219: stellar short wavelength radiation,
220: \begin{equation}
221: \frac{1}{\kappa_s \rho} \mathbf{\hat{k}} \cdot \nabla I_s =
222: 	I_s - \frac{\sigma E_0 e^{-\tau_s}}{4\pi}
223: 	- \frac{\sigma \int I_s d\Omega}{4\pi} 
224: \end{equation}
225: where $\rho$ is the density, $\kappa_s$ is the opacity at short wavelengths, 
226: and $E_0$ is the incident energy flux given by 
227: \begin{equation}
228: E_0 = \sigma_B T_{\star}^4 \left(\frac{R_{\star}}{a}\right)^2.
229: \end{equation}
230: The zeroth moment of this equation is 
231: \begin{equation}\label{s0}
232: \frac{1}{\kappa_s \rho} \nabla \cdot \mathbf{F}_s 
233: = \alpha 4\pi J_s - \sigma E_0 e^{-\tau_s},
234: \end{equation}
235: and the first moment, making use of the Eddington approximation, is
236: \begin{equation}\label{s1}
237: \frac{1}{\kappa_s \rho} \nabla J_s = \frac{3\,\mathbf{F}_s}{4\pi}. 
238: \end{equation}
239: 
240: The radiative transfer equation for the radiation absorbed 
241: and re-emitted within the disk is 
242: \begin{equation}
243: \frac{1}{\kappa_d\rho} \mathbf{\hat{k}} \cdot \nabla I_d = I_d - B
244: \end{equation}
245: where $\kappa_d$ is the optical depth at disk radiation wavelengths.  
246: The zeroth and first moments (using the Eddington approximation) are 
247: \begin{equation}\label{d0}
248: \frac{1}{\kappa_d\rho} \nabla \cdot \mathbf{F}_d = 4\pi (J_d - B),
249: \end{equation}
250: \begin{equation}\label{d1}
251: \frac{1}{\kappa_d\rho}\,\nabla J_d = \frac{3\,\mathbf{F}_d}{4\pi}.
252: \end{equation}
253: 
254: 
255: \subsubsection{Plane-parallel Disk}
256: \label{plane-parallel}
257: 
258: In a plane parallel disk atmosphere, this problem is reduced 
259: to one dimension, and we can express everything as a function of 
260: optical depth normal to the surface, $\tau_d$.  We can substitute 
261: \[ \frac{1}{\kappa_d\rho}\nabla \rightarrow \frac{d}{d\tau_d}
262: \quad \mbox{and} \quad
263: \mathbf{F}_{s,d} \rightarrow F_{s,d}. \]
264: If the angle of incidence 
265: of the stellar radiation is $\cos^{-1}\mu_0$, then 
266: \begin{equation}
267: \tau_d = \tau_s\mu_0/q
268: \end{equation}
269: where $q$ is the ratio of opacities
270: \begin{equation}
271: q \equiv \frac{\kappa_s}{\kappa_d}.
272: \end{equation}
273: The condition of zero net flux relates the short and long wavelength 
274: fluxes as 
275: \begin{equation}\label{sd}
276: F_s + F_d = E_0 \mu_0 \exp(-q\tau_d/\mu_0)
277: \end{equation}
278: and the boundary conditions at the surface are that of an isotropic 
279: radiation field, 
280: \begin{equation}
281: 2\pi\,J_s(0) = F_s(0) \quad \mbox{and} \quad
282: 2\pi\,J_d(0) = F_d(0).
283: \end{equation}
284: 
285: Now, equations (\ref{s0}), (\ref{s1}), (\ref{d0}), (\ref{d1}), and 
286: (\ref{sd}) are a closed system of linear differential equations 
287: that we can solve directly for the temperature, 
288: $T_r = (\pi B/\sigma_B)^{1/4}$.  The solution is given in 
289: \citet{calvet} as 
290: \begin{equation}
291: \label{radtemp}
292: B = \frac{\alpha E_0 \mu_0}{4\pi} 
293: [C_1' + C_2' \exp(-q\tau_d/\mu_0) + C_3' \exp(-\beta q\tau_d)]
294: \end{equation}
295: where $\beta\equiv\sqrt{3\alpha}$, 
296: \begin{eqnarray}
297: C_1' &=& (1+C_1)\left(2+\frac{3\mu_0}{q}\right) 
298: 	+ C_2\left(2+\frac{3}{\beta q}\right), \\
299: C_2' &=& \frac{(1+C_1)}{\mu_0} \left(q-\frac{3\mu_0^2}{q}\right), \\
300: C_3' &=& C_2 \beta \left(q - \frac{3}{q\beta^2}\right),
301: \end{eqnarray}
302: and 
303: \begin{eqnarray}
304: C_1 &=& -\frac{3\sigma\mu_0^2}{1-\beta^2\mu_0^2}, \\
305: C_2 &=& \frac{\sigma(2+3\mu_0)}{\beta(1+2\beta/3)(1-\beta^2\mu_0^2)}.
306: \end{eqnarray}
307: 
308: \figref{t_unperturbed} shows how temperature varies with 
309: $\mu_0$ at the surface ($\tau_d=0$) and in the interior 
310: ($\tau_d\rightarrow\infty$).  The temperatures are normalized 
311: to an effective radiant temperature $T_0 = (E_0/\sigma_B)^{1/4}$, 
312: which is the blackbody temperature corresponding to the incident energy flux.  
313: Note that at the surface the temperature remains relatively constant 
314: with $\mu_0$ and is 
315: greater than $T_0$ due to the absorptive properties of the dust, 
316: in accordance with the ``superheated'' surface layer proposed in \citet{CG}.
317: In contrast, the interior temperature is sensitive to the incident angle, 
318: especially at grazing incidence.  Thus, perturbations in the surface 
319: that change the angle of incidence can significantly affect 
320: the interior temperature.
321: 
322: By definition, $\tau_d=2/3$ is where the disk becomes optically 
323: thick to its own radiation.  
324: For $q\gtrsim1$, the temperature at $\tau_d=2/3$ is 
325: can be approximated by the interior temperature, since as 
326: $q\tau_d$ becomes large, the exponentials in \eqref{radtemp}
327: vanish and $B$ becomes independent of $\tau_d$.  
328: This indicates that in the absence of viscous heating, 
329: the disk interior is vertically isothermal for $\tau_d\geq2/3$.
330: 
331: 
332: \subsubsection{Disk with Perturbation}
333: 
334: Now we consider radiative transfer on a perturbed disk.  
335: A point $P(x,y,z)$ within 
336: the disk ``sees'' radiative flux coming from the surface of the disk.
337: Each area element $\delta A$ contributes a 
338: solid angle $\delta\Omega$ of flux to $P$.
339: \figref{angles} shows a schematic of such an area element.  
340: The angle of incidence of radiation on this surface element is 
341: $\cos^{-1}\mu$.  If $\mathbf{k}$ is the vector from $\delta A$ to $P$ and 
342: $\mathbf{\hat{n}}$ is the unit normal to $\delta A$, then 
343: $\nu$ is the cos of the angle between $\mathbf{\hat{n}}$ and $\mathbf{k}$.
344: 
345: We approximate the flux contribution from $\delta A$ as if 
346: it were a surface emitting isotropically with intensity 
347: \[ I = \frac{F_d(\tau_d,\mu)}{\pi}. \]
348: and the corresponding contribution to the total flux is 
349: $ d\mathbf{F} = I \,\mathbf{\hat{k}}\; d\Omega.$
350: Then the total flux at $P$ over all surface 
351: contributions is 
352: \begin{equation}
353: \mathbf{F}_{\mathrm{tot}} = 
354: \frac{1}{\pi} \int F_d \,\mathbf{\hat{k}}\; \delta\Omega.
355: \end{equation}
356: Since $\nabla$ is a linear operator,
357: \begin{equation}
358: \frac{1}{\kappa_d\rho}\nabla\cdot\mathbf{F}_{\mathrm{tot}} =
359: 	\frac{1}{\pi} \int \frac{d F_d}{d\tau_d} \,
360: 		\mathbf{\hat{n}} \cdot \mathbf{\hat{k}}\; \delta\Omega =
361: 	4\pi\left(
362: 	\frac{1}{\pi} \int J_d\, \nu \; \delta\Omega -
363: 	\frac{1}{\pi} \int B\, \nu \; \delta\Omega \right).
364: \end{equation}
365: Defining 
366: \begin{eqnarray}
367: J_{\mathrm{tot}} &=& \frac{1}{\pi} \int J_d(\tau_d,\mu)\,\nu\;\delta\Omega 
368: \quad \mbox{and} \\
369: B_{\mathrm{tot}} &=& \frac{1}{\pi} \int B(\tau_d,\mu)\,\nu\;\delta\Omega 
370: \label{btot}
371: \end{eqnarray}
372: we see that $\mathbf{F}_{\mathrm{tot}}$, $J_{\mathrm{tot}}$, 
373: and $B_{\mathrm{tot}}$ satisfy equations (\ref{d0}) and (\ref{d1}), 
374: and we can calculate the perturbed temperature as 
375: \begin{equation}
376: T_r = \left(\frac{\pi B_{\mathrm{tot}}}{\sigma_B}\right)^{1/4}.
377: \end{equation}
378: 
379: For a plane-parallel surface, $\tau_d$ and $\mu$ are constant, and 
380: $\nu = \cos \theta$,
381: where $\theta$ is the angle with respect to the surface normal.
382: We then recover the solution for the plane-parallel case.  
383: 
384: 
385: This method of calculating radiative transfer can be applied to 
386: any three-dimensional disk configuration, including perturbations 
387: induced by protoplanets.  Sufficiently massive protoplanets will 
388: open a gap in the disk, in which case the effects of shadowing and 
389: illumination across the gap can be calculated for all azimuthal angles.  
390: In the following section, we will calculate radiative transfer on 
391: small perturbations induced by 
392: planets too small to open a gap, but large enough to act on the 
393: disk locally.  
394: 
395: 
396: \section{Hydrostatic Equilibrium}
397: 
398: Now we consider a perturbation induced by the gravitational potential 
399: of a protoplanet on a circumstellar disk.  We shall consider only 
400: the gravitational effects of the protoplanet on hydrostatic 
401: equilibrium and assume that any resonant effects are washed out by 
402: gas drag.  
403: 
404: \subsection{Vertical Density Profile}
405: 
406: In order to find the vertical structure of a non-self-gravitating 
407: gaseous disk orbiting a central star, 
408: we solve the equation of hydrostatic equilibrium:
409: \begin{equation}
410: \nabla P = - \rho \nabla \Phi
411: \end{equation}
412: where $P$ is the gas pressure of the disk and $\Phi$ is the gravitational 
413: potential of the central star.  Since we are interested in the vertical 
414: structure, we consider only the $z$ components and assume that the disk is 
415: isothermal in the $z$ direction.  For a disk without a perturbing 
416: protoplanet, \( \Phi = GM_{\star}/a \) where $M_{\star}$ and $a$ 
417: are the mass and distance of the star, respectively.  
418: If $c_s$ is the isothermal sound speed, then 
419: \begin{equation}
420: c_s^2\,\frac{d\, \ln \rho}{dz} = 
421: - \frac{GM_{\star}}{a^3} z.
422: \label{thediffeq}
423: \end{equation}
424: If we assume that $a\gg z$ and define 
425: \( h \equiv c_s\,a^{3/2}/( GM_{\star})^{1/2}
426: = (c_s/v_{\phi}) a, \)
427: then 
428: \begin{equation}
429: \rho(z) =  \rho_0 \exp\left(-\frac{z^2}{2 h^2}\right),
430: \label{unperturbed}
431: \end{equation}
432: i.e.~the density has a Gaussian distribution with a scale height 
433: of $h$ which is determined by the local temperature. 
434: 
435: The insertion of a planet into a passive disk adds an additional term 
436: to $\Phi$, representing the potential due to the planet:
437: \[ \Phi = GM_{\star}/a + G m_p/r_p \]
438: where $m_p$ and $r_p$ are the mass and distance of the planet.
439: Assuming that the planet is in the midplane ($z=0$),
440: Eq.~(\ref{thediffeq}) becomes
441: \begin{equation}
442: \frac{d\, \ln \rho}{dz} =  -\frac{z}{h^2}
443: 	- \left(\frac{M_p a^3}{M_{\star} h^2}\right) \frac{z}{r_p^3}.
444: \end{equation}
445: The Hill radius is defined as 
446: \( \rhill \equiv (M_p/3M_{\star})^{1/3} a,\)
447: and this equation integrates to 
448: \begin{equation}
449: \rho(z) = \rho_1 \exp\left[-\frac{z^2}{2 h^2}
450: 	+ \frac{3\rhill^3}{h^2 \, r_p}
451: 	\right]
452: \end{equation}
453: where \(r_p = \sqrt{x^2 + y^2 + z^2}\), 
454: with the planet as the coordinate origin.
455: So long as $\rhill \ll h$ and $|z|>\rhill$, 
456: the effect of the planet is a small 
457: perturbation on the density profile.  
458: 
459: We normalize this equation by matching 
460: the density at $z=0$ to the unperturbed density, thus
461: \begin{equation}
462: \rho_p(z) = \rho_0 \exp\left[-\frac{z^2}{2 h^2}
463:         + \frac{3\rhill^2}{h^2}
464: 	  \left(\frac{\rhill}{\sqrt{x^2+y^2+z^2}}
465: 	-\frac{\rhill}{\sqrt{x^2+y^2}}\right)\right]
466: \label{pertdens}
467: \end{equation}
468: 
469: Note that there is a singularity ar $r_p = 0$.   \figref{zprof} shows 
470: the shape of this density profile. 
471: In \figref{zprof}a, we hold $\rhill/h$ fixed at $0.5$ and vary
472: $\sqrt{x^2+y^2}$.  
473: For a disk with $h=0.04\AU$ at $1\AU$ around a central star of
474: $0.5\:\msun$, this corresponds to a protoplanetary mass of
475: $4\:\mearth$.  For
476: $(x^2+y^2)\gtrsim\rhill^2$, the perturbed density profile does not
477: deviate significantly from the unperturbed density profile.  For
478: $(x^2+y^2)\lesssim\rhill^2$, the density profile develops a sharper
479: profile at $z<h$.  
480: In \figref{zprof}b, we hold $\sqrt{x^2+y^2}$ fixed at the Hill radius
481: and vary the mass of the protoplanet, effectively varying $\rhill$.
482: Larger protoplanets have the effect of decreasing the overall density 
483: profile, whereas position controls the shape of the profile, 
484: particularly close to the protoplanet.  
485: Since the flow of gas within the Roche lobe cannot
486: be adequately described hydrostatically, we exclude the region
487: $\sqrt{x^2+y^2} \lesssim \rhill$ from our analysis.
488: 
489: The normalization on \eqref{pertdens} is somewhat arbitrary, but 
490: at the disk heights that are of interest to us, the overall normalization 
491: has small effect.  We demonstrate this as follows: 
492: as $z\gg\rhill$, 
493: \eqref{pertdens} becomes
494: \begin{equation}
495: \rho_p(z\gg\rhill) \approx \rho_0
496: \exp\left(-\frac{z^2}{2 h^2}-\frac{3\rhill^3}{h^2\sqrt{x^2+y^2}}\right),
497: \end{equation}
498: that is, a Gaussian with respect to $z$ with some normalization 
499: that depends on $\rhill$ and $\sqrt{x^2+y^2}$.
500: The height of the photosphere, $H$, is generally several times the 
501: pressure scale height.  If we take $H/h=5$, then the change in 
502: density with vertical distance is 
503: \begin{equation}
504: \label{drho2}
505: \frac{\Delta\rho}{\rho} = 
506: \left|\frac{\Delta z}{\rho}\frac{d\rho}{dz}(z=H)\right|
507: = \frac{25\Delta z}H.
508: \end{equation}
509: So, a density change by a factor of $1/2$ corresponds to a shift in 
510: vertical height of 2\%.  In other words, the density gradient in the 
511: photosphere is so high that changes in the normalization of the 
512: density will not significantly affect our results. 
513:  
514: \subsection{Photosphere Height}
515: \label{photosphere}
516: We calculate the height of the perturbed photosphere from the 
517: perturbed density profile as the isodensity contour at the 
518: density of the unperturbed photosphere.
519: That is, if $H_0$ is the unperturbed photosphere 
520: height and $H=H_0(1-\epsilon)$ is the new photosphere height, then 
521: we set 
522: \[
523: \rho_d(H_0) = \rho_p(H)
524: \]
525: or
526: \begin{equation}
527: -\frac{H_0^2}{2} =
528: -\frac{H^2}{2} + 3\rhill^3 
529: 	  \left(\frac{1}{\sqrt{x^2+y^2+H^2}}-\frac{1}{\sqrt{x^2+y^2}}\right).
530: \end{equation}
531: Setting \( \sqrt{x^2+y^2} = \gamma H_0 \), then
532: to first order in $\epsilon$
533: 
534: \begin{equation}
535: H \approx H_0 \left\{1 - 
536: 	3\left(\frac{\rhill}{H_0}\right)^3
537: 	\left[ \frac{1}{\gamma} - \frac{1}{\sqrt{\gamma^2+1}} \right] 
538: 	\right\} 
539: \end{equation}
540: so that the depth of the perturbation scales as $(\rhill/H_0)^3$.  
541: For typical disk models, the ratio of the photosphere height to the 
542: thermal scale height, $H_0/h$, is between 3 and 5, depending on the 
543: distance from the star and is often taken to be a constant 
544: \citep{kenyon,CG,vertstruct}.  Restating the scaling of the 
545: depth of the perturbation as $\propto(\rhill/h)^3$, we see that this 
546: agrees with the estimate made in \citet{sas_lec}.
547: 
548: The shape of this perturbed photosphere is shown in \figref{hprof} 
549: for varying values of $\rhill$.  The perturbation is small for $r>\rhill$, 
550: and there is a singularity at $r=0$.  As the Hill radius increases, 
551: the depth of the perturbation also increases.  
552: Note that if $H_0/h=5$, then $\rhill/H_0=0.2$ means that the Hill 
553: radius is comparable to the disk pressure scale height and
554: the perturbation is no longer small.  Protoplanets this massive are 
555: likely to open a gap in the disk rather than simply inducing a 
556: perturbation to the photosphere.
557: 
558: 
559: 
560: \section{Disk Shear}
561: 
562: The differential rotation of the disk causes the disk material 
563: near the protoplanet to move through the perturbation at a rate 
564: that depends on the velocity with respect to the protoplanet's 
565: orbit.  Material moving along a given streamline in the disk 
566: may pass through the perturbation too quickly to experience 
567: significant heating and cooling, thereby diminishing the effect 
568: of the perturbation on the temperature structure of the disk.
569: 
570: The velocity of the disk material with respect to 
571: the protoplanet is 
572: $v \simeq a(\Omega-\Omega_p)$ where $\Omega$ and $\Omega_p$ are 
573: the orbital angular velocities of the disk material and the protoplanet,
574: respectively, and $a$ is the orbital radius of the protoplanet.
575: The orbital angular velocity of a gaseous disk is 
576: \begin{equation}\label{streamvel}
577: r^2\Omega^2 = \frac{GM_{\star}}{r} + \frac{1}{\rho}\frac{dP}{dr}.
578: \end{equation}
579: In thin disks, such as we are investigating, the pressure gradient 
580: term is small so the orbital velocity is nearly equal to the 
581: Keplerian velocity.  
582: 
583: The heating/cooling rate of the disk material can be expressed as 
584: \begin{equation}
585: C \frac{\partial T}{\partial t} = F - \sigma T^4
586: \end{equation}
587: where we take $F = \sigma(T_v^4+T_r^4)$ and $C$ is the specific 
588: heat per unit surface area of the disk.
589: We shall adopt a specific heat of $C = k\Sigma/m$ where $k$ is the 
590: Boltzmann constant, $\Sigma$ is the total surface density, and 
591: $m$ is the mean molecular mass.  
592: 
593: We shall assume that the streamlines of the disk material follow 
594: \eqref{streamvel}, and that they are not perturbed by the 
595: protoplanet's graviational potential.  This is an adequate approximation 
596: for disk material outside the protoplanet's Hill radius.
597: Along each streamline, we calculate the total radiative flux, $F$, 
598: at a given position, and using the velocity of the streamline 
599: with respect to the protoplanet along with heating/cooling 
600: rate, we can calculate the steady state temperature at each position 
601: in the disk. 
602: 
603: 
604: \section{Model Parameters}
605: 
606: Dust properties can be parametrized by $\alpha$, $\kappa_s$, and $q$.
607: Then, the temperature profile in a plane-parallel disk is determined 
608: completely by $F_v$, $E_0$, $\mu_0$, and $\tau_d$, where $\tau_d$ is 
609: measured normal to the surface of the photosphere.  
610: Stellar properties determine $F_v$ and $E_0$, 
611: and the disk profile determines $\mu_0$.
612: If $H_0 \propto r^{\xi}$, then the incident angle is determined by 
613: \begin{equation}
614: \mu_0 = \frac{ (\xi-1) H_0/a }{
615: [1+(H_0/a)^2][1+\xi^2(H_0/a)^2]}
616: \end{equation}
617: 
618: A perturbation changes the local angle of incidence and optical depth.  
619: In particular, without plane parallel symmetry, there is no single 
620: optical depth that parametrizes the distance below the surface.  
621: Instead, we sum over different lines of sight, using the optical depth 
622: along each line of sight 
623: to determine the contribution to the disk flux as in \eqref{btot}.  
624: The local angle of incidence depends only on the geometry of the 
625: perturbed surface, while the optical depth also depends on the 
626: density structure along the line of sight.  
627: The shape of the perturbed surface depends on $\rhill$ and $H_0$, 
628: and the density structure depends on the disk pressure scale height, $h$.  
629: 
630: 
631: 
632: \subsection{Fiducial Model}
633: 
634: Motivated by observations of T Tauri stars, we will assume 
635: for our fiducial model that the central star 
636: has mass $M_{\star} = 0.5\;\msun$, radius $R_{\star} = 2 R_{\odot}$, 
637: and effective temperature $T_{\star} = 4000$ K, 
638: and that the protoplanet is at $a = 1\AU$ from the star.  
639: We assume an accretion rate of 
640: $\dot{M}_a = 10^{-8}\;\msun\,\mbox{yr}^{-1}$, so that 
641: $T_v = 72$ K at $\tau_d = 2/3$ at 1 AU.  
642: With these parameters, heating in the photosphere will be dominated 
643: by stellar irradiation rather than viscous heating.  
644: Near the midplane, however, viscous heating will dominate.  
645: Since we only consider the photosphere in our model, 
646: the calculation of the midplane temperature is outside the scope 
647: of this paper.  
648: 
649: We will assume that the fraction of scattered radiation is $\alpha = 0.28$, 
650: with opacity in the optical of 
651: $\kappa_s = 400\;\mbox{cm}^2\mbox{g}^{-1}$, and 
652: ratio of opacities $q = 20$.
653: 
654: We shall assume that $H_0 = 0.16\AU (r/1\AU)^{9/7}$,
655: consistent with \eg \citet{dalessio2}.  
656: In calculating the detailed density structure in the photosphere, 
657: we shall assume that $h=3.4\times10^{-2}\AU$ so that $H_0/h = 5$.  
658: A protoplanet with mass $m_p = 1\;\mearth$ ($10\:\mearth$) will have a 
659: Hill radius of $1.26\times10^{-2}\AU$ ($2.72\times10^{-2}\AU$).  
660: As shown by \citet{bate}, planets with masses $\lesssim 30\:\mearth$ 
661: are not massive enough to open a gap, and our perturbative 
662: treatment is valid.  We have calculated models with 
663: $\:1\mearth$, $10\:\mearth$, and $20\:\mearth$ planets, and present the 
664: results below.
665: 
666: 
667: \subsection{Numerical details} 
668: 
669: We define our coordinate axes so that 
670: $x$ is aligned with the radial direction, 
671: $y$ is aligned with $\phi$ (i.e.~the direction of planet's orbit), 
672: and $z$ is perpendicular to the orbital plane.
673: The coordinate origin is set at the surface of the disk above 
674: the planet position, so that the planet's coordinates are 
675: $(0,0,-H_0)$.  
676: 
677: We calculate the shape of the disk surface in the $x$ and $y$ directions 
678: over a range \( -10\rhill \leq x< 10\rhill \) and 
679: \( -10\rhill \leq y< 10\rhill \), 
680: and use this surface to numerically 
681: integrate temperatures within the photosphere, to a depth of 
682: $z(\tau_d=2/3) = -0.011\AU$.
683: The temperature is calculated at each point within a 
684: $64\times64\times64$ grid.
685: 
686: When summing over surface elements to find $B_\mathrm{tot}$ as in 
687: \eqref{btot}, we can approximate 
688: $\delta\Omega\approx\nu\delta A/\ell^2$ 
689: for $\delta A \ll \ell^2$, where $\ell$ is the distance between 
690: the point $P$ and $\delta A$.  However, 
691: for points close to the surface, this approximation breaks down.  
692: In addition, $\nu$ may have large excursions over $\delta A$.  
693: For this reason, we assume that each $\delta A$ is approximately planar, 
694: so that $B(\tau_d,\mu)$ does not vary, but we calculate 
695: $\int_{\delta\Omega} \nu\: d\Omega$ analytically from the limits 
696: of $\delta A$.  In other words, we calculate
697: \begin{equation}
698: B_\mathrm{tot}
699:  = \frac{1}{\pi} \sum B(\tau_d,\mu) \int\limits_{\delta\Omega} \nu\: d\Omega
700: \end{equation}
701: and solve for the temperature.
702: 
703: 
704: \section{Results}
705: 
706: Figures \ref{planet01}, \ref{planet10},  and \ref{planet20} 
707: show the spatial distribution of temperatures 
708: in the photosphere of the fiducial model for $1\:\mearth$ 
709: $10\:\mearth$, and $20\:\mearth$ planets,
710: respectively.  The colors represent temperatures in Kelvin, 
711: as indicated by the colorbars.
712: The horizontal axis indicates increasing radial distance from the star 
713: so that the photosphere is illuminated from the left.
714: The cross-section is taken at the $\tau_d=2/3$ surface. 
715: The bottommost row and leftmost column show the
716: unperturbed temperature of $\tunpert$ for comparison.  This temperature
717: is also indicated in the colorbar by a dotted white line.  
718: 
719: The direction of the planet's orbit is upward on the plots, so 
720: the interior disk material moves faster, and the 
721: exterior disk material moves slower than the planet.  
722: This causes the heated and cooled regions to shear out.  The 
723: positions of minimum and maximum temperatures are offset from 
724: the planet's position since it takes some time for the 
725: heated/cooled material to come back to the equilibrium 
726: temperature. 
727: 
728: For a $1\:\mearth$ protoplanet, 
729: the shadowing and illumination at the surface results in a slight
730: temperature change, with the area to the right of the
731: planet reaching a maximum temperature of $\tsmmax$, and the area to 
732: the left cooled to a minimun of $\tsmmin$ 
733: -- a change of $-\losm\%$ to $+\hism\%$.  
734: Thus, we see that an Earth-mass planet is too small
735: significantly affect the interior temperature of the disk.
736: 
737: A larger planet will create a larger perturbation, which will lead to
738: a greater temperature change.  
739: \figref{planet10} and \figref{planet20} show the temperature distribution
740: in the photosphere for a $10\:\mearth$ planet and $20\:\mearth$ planet,
741: respectively.  
742: The plots are scaled to the Hill radius, which goes as 
743: $m_p^{1/3}$, so the physical range in Figures \ref{planet10} and 
744: \ref{planet20} is actually larger than that in \figref{planet01}.  
745: The high temperatures close to the Hill radius are due to the low
746: density in this region, which gives a low optical depth to the
747: illuminated surface.  However, since this region is not in hydrostatic
748: equilibrium, we exclude it from our analysis.  Outside the Hill 
749: radius, the temperatures at $\tau_d=2/3$ for a $10\:\mearth$ protoplanet 
750: range from $\tlgmin$ to $\tlgmax$, or by $-\lolg\%$ to $+\hilg\%$.  
751: For a $20\:\mearth$ protoplanet, which is close to but not at the limit 
752: where a gap in the disk can form, the temperature range increase 
753: slightly, with a minimum of $\txlmin$ and maximum of $\txlmax$, 
754: or by $-\loxl\%$ to $+\hixl\%$.  
755: Thus, a protoplanet with mass 10 $\mearth$ 
756: or larger may have a significant effect on the local
757: temperature structure of a protoplanetary disk.
758: 
759: \section{Discussion}
760: 
761: As planet-forming disks are likely to be passive \citep{calvet2002}, 
762: we have explored
763: stellar irradiation effects due to a disk surface perturbation
764: induced by an embedded protoplanet. We find that even a relatively small
765: protoplanet (in the range from 1 to 20 Earth mass) produces enough
766: of a perturbation to cause shadowing. 
767: In particular, we find that in a standard passive disk accreting 
768: at $10^{-8}\:\msun\mathrm{yr}^{-1}$, 
769: the temperature variation induced by a $1\:\mearth$ planet at 1 AU 
770: is $\tsmd$, or only about $\dsm\%$.  However, we find that even though 
771: a protoplanet may not be massive enough to open a gap in a disk, it may 
772: still be able to significantly change the temperature structure in 
773: a the disk material near it.  
774: A $10\:\mearth$ ($20\:\mearth$) planet can induce a 
775: temperature variation of $\tlgd$ ($\txld$), or $\dlg\%$ ($\dxl\%$).
776: 
777: This results in a local temperature
778: perturbation which is large enough to affect the local properties of
779: the disk and the accretion rate, and possibly the migration rate
780: of the protoplanet.  
781: Although the perturbation in temperature may be 
782: too subtle to directly observe, it can have serious consequences for 
783: planet growth and migration.
784: 
785: \citet{wardA} has shown that 
786: the rate of Type I migration is strongly dependent upon the temperature 
787: gradient of the disk.  If the temperature decreases with distance from 
788: the star, then the total net torques result in inward migration 
789: of the secondary.  However, if $k = -d\ln T/d\ln r \lesssim -1$, then 
790: the opposite is true and the secondary will migrate outward.  
791: The temperature perturbation near a protoplanet locally 
792: decreases the value of $k$, and may result in halting or even 
793: reversing Type I migration.  This may resolve the conundrum that 
794: Type I migration timescales are typically much less than observed 
795: disk lifetimes of $\sim10^7$ years \citep{hartmann}.
796: Yet to date, no simulation of planet-disk interactions accounts 
797: for local temperature pertubations as described here.  
798: 
799: We have not yet considered 
800: the behavior of the dust (as a function of size) in
801: the perturbed region.  We expect the dust to remain coupled
802: to the gas and compress its scale height accordingly. In
803: fact crude estimates of dust sedimentation in the vicinity
804: of the planet confirm that. However, a detailed estimate of
805: the sedimentation timescale is necessary in order to evaluate
806: the amount of residual dust, \eg swept by stellar radiation
807: from the inner edge, and its total opacity. These are issues 
808: we plan to address in future work.  
809: 
810: In a forthcoming paper, we will undertake a parameter study 
811: of our model for radiative transfer on a perturbed disk.  
812: By varying attributes of the disk, star, and protoplanet, we 
813: will investigate the possibility that a relatively small protoplanet 
814: which is not large enough to open a gap in a disk might still be 
815: able to affect disk structure.  
816: 
817: 
818: 
819: \acknowledgements
820: We thank Nuria Calvet for numerous helpful comments in the preparation 
821: of this paper.  
822: 
823: \bibliographystyle{apj}
824: \bibliography{apj-jour,/home/hjang/Thesis/Papers/planets}
825: 
826: 
827: \begin{figure}
828: \plotone{f1.eps}
829: \caption{\label{flareddisk}Schematic of a flared disk. 
830: Proportions are not to scale.  The labels thick and thin 
831: refer to optical depth in short and long wavelengths.
832: }
833: \end{figure}
834: 
835: \begin{figure}
836: \plotone{f2.eps}
837: \caption{\label{t_unperturbed}The temperature due solely to stellar 
838: irradiation at the surface of the disk (dashed line) 
839: and in the interior (solid line) 
840: for varying values of the angle of
841: incidence, $\mu_0$, where $T_0 = (E_0/\sigma_B)^{1/4}$.  
842: }
843: \end{figure}
844: 
845: \begin{figure}
846: \epsscale{0.5}
847: \plotone{f3.eps}
848: \caption{\label{angles}Angles involved in calculation of perturbed surface.  
849: See text for details.}
850: \epsscale{1.0}
851: \end{figure}
852: 
853: \begin{figure}
854: \plottwo{f4a.eps}{f4b.eps}
855: \caption{\label{zprof}Perturbed vertical density profile for varying 
856: parameters.  Density is normalized to the midplane density.  Both 
857: plots are to the same scale and show the unperturbed density profile 
858: as a solid line.
859: (a) Density profile for fixed Hill radius and 
860: varying values of $(x^2+y^2)^{1/2}$ as indicated.
861: (b) Density profile for $(x^2+y^2)^{1/2}$ equal to the 
862: Hill radius and varying planetary masses.
863: }
864: \end{figure}
865: 
866: \begin{figure}
867: \plotone{f5.eps}
868: \caption{\label{hprof}Perturbed photosphere height as a function of 
869: distance in the $xy$-plane, $r$, for varying Hill radii.  The solid vertical 
870: line indicates where $r=\rhill$.
871: }
872: \end{figure}
873: 
874: \begin{figure}
875: \plotone{f6.eps}
876: \caption{\label{planet01}Temperature profile of the disk photosphere 
877: near a protoplanet of mass 1 $\mearth$ 
878: at $\tau_d = 2/3$ in the $xy$-plane.  
879: The colors indicate the temperature in Kelvin, scaled as shown in 
880: the colorbar. 
881: The cross in the center represents planet position, 
882: with the arrow indicating the direction of the planet's orbit.
883: The blacked-out circle indicates the extent of the Hill radius.  
884: The dotted white line 
885: in the color bar indicates the unperturbed temperature at 
886: $\tau_d=2/3$.
887: }
888: \end{figure}
889: 
890: \begin{figure}
891: \plotone{f7.eps}
892: \caption{\label{planet10}Same as \figref{planet01}, for a 
893: protoplanet of mass 10 $\mearth$.  Note that the dimensions 
894: on the figures scale with the Hill radius, so the physical sizes of 
895: the plots are $10^{1/3}$ larger than in \figref{planet01}.
896: }
897: \end{figure}
898: 
899: \begin{figure}
900: \plotone{f8.eps}
901: \caption{\label{planet20}Same as \figref{planet01}, for a 
902: protoplanet of mass 20 $\mearth$.}
903: \end{figure}
904: 
905: \end{document}
906: