1: \documentclass[12pt,preprint]{aastex}
2: %\documentclass[12pt,manuscript,doublespace]{aastex}
3: \doublespace
4: \slugcomment{73 pages, including 12 figures and 7 tables}
5: \shorttitle{solid surface density of solar nebula}
6: \shortauthors{Dodson-Robinson et al.}
7: \begin{document}
8:
9: \title{Ice Lines, Planetesimal Composition and Solid Surface Density in
10: the Solar Nebula}
11:
12: \author{Sarah E. Dodson-Robinson}
13: \affil{NASA Exoplanet Science Institute, California Institute of
14: Technology \\
15: MC 100-22 Pasadena, CA 91125}
16: \email{sdr@ipac.caltech.edu}
17:
18: \author{Karen Willacy}
19: \affil{Jet Propulsion Laboratory, California Institute of Technology \\
20: 4800 Oak Grove Dr, Pasadena, CA 91109}
21:
22: \author{Peter Bodenheimer}
23: \affil{ Department of Astronomy and Astrophysics, University of
24: California at Santa Cruz \\
25: 1156 High St, Santa Cruz, CA 95064}
26:
27: \author{Neal J. Turner}
28: \affil{Jet Propulsion Laboratory, California Institute of Technology \\
29: 4800 Oak Grove Dr, Pasadena, CA 91109}
30:
31: \author{Charles A. Beichman}
32:
33: \affil{NASA Exoplanet Science Institute/Jet Propulsion Laboratory \\
34: California Institute of Technology \\
35: MC 100-22, Pasadena, CA 91125}
36:
37: %Proposed running head: Solid Surface Density in the Solar Nebula
38:
39: %\bigskip
40:
41: %Please direct correspondence to:
42:
43: %\begin{center}
44: %\begin{tabular}{l}
45: %Sarah Dodson-Robinson \\
46: %NASA Exoplanet Science Institute \\
47: %California Institute of Technology \\
48: %Mail Code 100-22 \\
49: %Pasadena, CA 91125 \\
50: %sdr@ipac.caltech.edu \\
51: %\end{tabular}
52: %\end{center}
53:
54: %\clearpage
55:
56: %\begin{center}
57: %{\bf \large ABSTRACT}
58: %\end{center}
59:
60: \begin{abstract}
61: To date, there is no core accretion simulation that can successfully
62: account for the formation of Uranus or Neptune within the observed
63: 2-3~Myr lifetimes of protoplanetary disks. Since solid accretion rate
64: is directly proportional to the available planetesimal surface density,
65: one way to speed up planet formation is to take a full accounting of all
66: the planetesimal-forming solids present in the solar nebula. By
67: combining a viscously evolving protostellar disk with a kinetic model of
68: ice formation, which includes not just water but methane, ammonia, CO
69: and 54 minor ices, we calculate the solid surface density of a possible
70: giant planet-forming solar nebula as a function of heliocentric distance
71: and time. Our results can be used to provide the starting planetesimal
72: surface density and evolving solar nebula conditions for core accretion
73: simulations, or to predict the composition of planetesimals as a
74: function of radius.
75:
76: We find three effects that favor giant planet formation by the core
77: accretion mechanism: (1) a decretion flow that brings mass from the
78: inner solar nebula to the giant planet-forming region, (2) the fact that
79: the ammonia and water ice lines should coincide, according to recent lab
80: results from \cite{collings04}, and (3) the presence of a substantial
81: amount of methane ice in the trans-Saturnian region. Our results show
82: higher solid surface densities than assumed in the core accretion models
83: of Pollack et al. (1996) by a factor of 3--4 throughout the
84: trans-Saturnian region. We also discuss the location of ice lines and
85: their movement through the solar nebula, and provide new constraints on
86: the possible initial disk configurations from gravitational stability
87: arguments.
88: \end{abstract}
89:
90: {\bf Keywords:} solar nebula; planetary formation; ices
91:
92: \clearpage
93:
94: \section{Introduction}
95: \label{introduction}
96:
97: Previous core accretion simulations of the formation of Uranus and
98: Neptune require the solar nebula lifetime to be at least 10~Myr longer
99: than the observed median protostellar disk lifetime (2--3 Myr; Haisch et
100: al. 2001). Similarly, the simulation of Saturn's formation by Alibert
101: et al. (2001) allows the planet to form in 2.5~Myr, but only if it
102: migrates from 11.9~AU to 9.5~AU during formation.
103: %and incorporates 100\% of the available
104: %planetesimals into its core.
105: This scenario is incompatible with the
106: outward migration of Saturn predicted by the Nice model of planetary
107: dynamics (Tsiganis et al. 2005)
108: %and the presence of $\sim 40 M_{\oplus}$
109: %of comets scattered into the Oort cloud by Jupiter and Saturn (Weissman
110: %1996).
111: This research was motivated by the need to create a model of the
112: solar nebula with enough solid mass to form Jupiter, Saturn, Uranus and
113: Neptune via core accretion at heliocentric distances of $5 < R < 20$~AU.
114:
115: For core accretion simulations, planetesimal mass available for building
116: giant planet cores is usually calculated by invoking the scaled
117: minimum-mass solar nebula with surface density $\Sigma \propto
118: R^{-3/2}$, obtained from spreading out the current mass of the planets
119: (Weidenschilling 1977), and assuming a gas/solid ratio of $\sim 70$.
120: However, a single planet core embedded in a disk of planetesimals grows
121: at the rate of
122: \begin{equation}
123: {dM_{\rm core} \over dt} = \pi a_{\rm core}^2 \Sigma_{\rm solid}
124: \Omega F_g ,
125: \label{planetgrowth}
126: \end{equation}
127: where $F_g$ is a factor taking into account gravitational focusing by
128: the core (Safronov 1969), $\Sigma_{\rm solid}$ is the solid surface
129: density of planetesimals, $\Omega$ is the orbital angular frequency of
130: the core, and $a_{\rm core}$ is the core's effective capture radius.
131: For Keplerian orbits, we have
132: \begin{equation}
133: \Omega = \left ( {GM_* \over R^3} \right )^{1/2} .
134: \label{kepler}
135: \end{equation}
136: The $R^{-3/2}$ dependence of $\Omega$ means that even planets forming
137: in a disk with uniform surface density should, in the end, reach masses
138: that suggest $\Sigma \propto R^{-3/2}$ if growth rate is not taken
139: into account. A near-uniform surface density distribution, in which
140: mass is not concentrated in the inner disk, gives higher surface
141: densities in the outer disk and may speed up the formation
142: of Uranus and Neptune.
143:
144: Furthermore, the canonical gas/solid ratio of 70, which is based on the
145: composition of Comet Halley (Jessberger et al. 1989), includes only one
146: ice in the solid inventory---H$_2$O. However, there were certainly
147: other ices present in the solar nebula: the Deep Impact ejecta from
148: Tempel 1 contained $\sim 1000$~tons of excavated CO, CO$_2$ and CH$_3$OH
149: (A'Hearn 2008). Taking a full accounting of the solids in the solar
150: nebula may reveal other volatile mass sources besides H$_2$O that
151: condense in the outer solar nebula and improve the formation prospects
152: of Uranus and Neptune.
153:
154: In this work, we calculate the time-evolving solid surface density
155: available for giant planet formation. We make two updates to previous
156: solar nebula simulations: (1) we relax the assumption of a steady-state
157: disk and allow for viscous redistribution of mass, which tends to drive
158: the surface density profile toward uniformity, and (2) we add a chemical
159: reaction network to trace the formation and freezeout of volatiles.
160:
161: This paper is organized as follows. In \S \ref{overview}, we outline
162: our method for calculating solid surface density in the solar nebula:
163: which codes we use, how they are linked, and how they differ from
164: previous solar nebula models. In \S \ref{alphadisk}, we describe our
165: protostellar disk model and use it to calculate the total surface
166: density (gas+solid) and temperature evolution of the disk. \S
167: \ref{chemmodel} details the physics behind our chemical reaction
168: network, the abundances of the major ice species, and the location of
169: the ice lines. We calculate the solid surface density available for
170: planet formation in \S \ref{planetesimal} and present our conclusions in
171: \S \ref{conclusions}.
172:
173: \section{Simulation Overview}
174: \label{overview}
175:
176: This simulation consists of two independent, non-interacting codes used
177: in sequence. The first code is a non-steady-state, 1+1-dimensional
178: $\alpha$-disk model that simulates the dynamical evolution of the solar
179: nebula. Code 1 gives the surface density distribution $\Sigma(R,t)$,
180: along with the vertical and radial temperature and volume density
181: distributions $T(R,z,t)$ and $\rho(R,z,t)$.
182:
183: The results from Code 1 fill an important niche in the body of work on
184: protostellar disk evolution. Currently available are 2-d, static disk
185: models that include stellar insolation and radiogenic heating (e.g.
186: D'Alessio et al. 2006); quasi-steady-state 1+1-d models where
187: $\dot{M}(t)$ follows an imposed prescription (e.g. Hersant et al. 2001);
188: and non-steady-state models with only one zone in the vertical direction
189: (Ruden and Pollack 1991). The evolving protostellar disks of Boss
190: (2007) include 3-d hydrodynamics and radiative diffusion, but cover a
191: timespan of $< 4000$ years.
192:
193: In this work, we couple full vertical structure models with the
194: time-dependent, non-steady-state surface density evolution of the disk
195: and present the first long-term solar nebula simulation with viscous
196: angular momentum transport determined self-consistently with vertical
197: structure. For a description of how the resulting dynamical evolution
198: of our disk differs from previous simulations, see \S \ref{massflux}.
199:
200: Code 2 is a non-equilibrium chemical model, including gas reactions,
201: grain-surface reactions, freezeout and desorption. This code traces the
202: abundance of 211 species, including 58 ices, as the disk cools and the
203: ice lines move inward. Chemical reaction, freezeout and desorption
204: rates are governed by the temperature and density of the solar nebula
205: gas and grains. We use the midplane temperatures and densities
206: calculated by Code 1 to generate reaction rate coefficients for Code 2.
207:
208: %Even though the energy release during freezeout and
209: %exothermic reactions can heat the disk and affect its structure and
210: %evolution, we do not yet have the capacity to integrate a complex
211: %chemical-reaction network into a long-term dynamical simulation.
212:
213: Previous chemical models of the solar nebula include those of Aikawa et
214: al. (1996), Willacy and Langer (2000), and Ilgner and Nelson (2006).
215: \cite{aikawa96} extended the minimum-mass solar nebula (MMSN; Hayashi
216: 1981) to 800~AU to predict the observed depletion of gaseous CO in the
217: outer regions of T-Tauri disks. Their calculation included gas-phase
218: reactions of CO and other C, H, O, N and S-based molecules, freezeout of
219: CO and its reaction products, formation of H$_2$, and electron-ion
220: recombination on grain surfaces. The density and temperature of the
221: underlying disk were assumed to stay constant with time, while the
222: time-evolving midplane CO abundance was calculated with an independent
223: sequence of radial one-zone models.
224:
225: Willacy and Langer (2000) calculated the vertical distribution and
226: column density of common interstellar molecules as a function of time in
227: the flared protoplanetary disk of Chiang and Goldreich (1997). They
228: used a similar species set to Aikawa et al. (1996) but added
229: grain-surface reactions and photodesorption to the chemistry. The
230: simulation consisted of a 2-d grid of noninteracting one-zone models in
231: the $R-z$ plane. The underlying temperature and density distribution of
232: the protostellar disk remained constant with time.
233:
234: Finally, Ilgner and Nelson (2006) embedded a five-species chemical
235: reaction network in an evolving $\alpha$-disk model, fully coupling the
236: chemical and dynamical evolution of the disk. They used the chemical
237: reaction network to trace the ionization fraction, which is affected by
238: charge-transfer reactions, electron-ion recombination, and viscous
239: diffusion. The simulation was computationally tractable due to the
240: small reaction network (5 species, 4 reactions), and the limited extent
241: of the protostellar disk (0.1--10~AU).
242:
243: Code 2 of this work is the first to combine an extensive chemical
244: reaction network, including ice formation and grain-surface reactions,
245: with a time-evolving protostellar disk model (as used by Ilgner and
246: Nelson 2006). Our chemical model consists of a radial sequence of
247: noninteracting one-zone models, as in Aikawa et al. (1996). The disk
248: model in Code 1 therefore functions as a stand-alone input to the
249: chemical model of Code 2. Our models differ from those of Aikawa et al.
250: (1996) and Willacy and Langer (2000) in that we allow the disk
251: temperature and density to change with time, as prescribed by the
252: results of Code 1. Although Code 1 simulates the full vertical
253: structure of the protostellar disk, we apply Code 2 only to the
254: midplane, which is where solids settle and build planetesimals.
255:
256: Code 2 calculates the abundance of each ice and gas species as a
257: function of heliocentric radius and time. Combining these results with
258: information from the literature about the composition of refractory
259: grains, we calculate the gas/solid mass ratio in the solar nebula. This
260: ratio tells us what fraction of the total surface density calculated by
261: Code 1 is solid and can build giant planet cores. The end result of
262: both codes is the solid surface density distribution $\Sigma_{\rm
263: solid}(R,t)$.
264:
265: Since most molecules in the solar nebula formed in a molecular cloud, we
266: first run Code 2 for molecular cloud conditions ($T = 10$~K, hydrogen
267: number density $2 \times 10^4 \; {\rm cm}^{-3}$, grain radius $0.258 \mu
268: m$). The molecular cloud model begins with atomic gas of solar
269: composition (except for hydrogen, 99\% of which is in H$_2$) and bare
270: refractory grains at the standard interstellar abundance $n_g = 10^{-12}
271: (n_{\rm H} + 2 n_{\rm H2})$, where $n_g$ is the grain abundance and
272: $n_{\rm H}$ and $n_{\rm H2}$ are the abundances of H and H$_2$,
273: respectively. We run the molecular cloud model for $10^6$~yr and use
274: the resulting ice-gas mixture as the starting composition of the solar
275: nebula.
276:
277: \section{Disk Model}
278: \label{alphadisk}
279:
280: \subsection{Overall Model Formulation}
281: \label{framework}
282:
283: The goals of this simulation are to calculate the time-evolving
284: temperature and density in the solar nebula, which determine the
285: reaction rates for our kinetic model of ice formation, and to follow the
286: viscous evolution of the surface density profile. Although from a
287: planet formation perspective we are primarily concerned with the ice
288: inventory at the solar nebula midplane, an accurate dynamical
289: description of the midplane requires understanding the vertical energy
290: balance in the disk. To simplify the problem of long-term protostellar
291: disk evolution, we use the following assumptions:
292:
293: \begin{enumerate}
294:
295: \item The disk is axisymmetric and symmetric about the midplane.
296:
297: \item The disk is geometrically thin: $H / R << 1$.
298:
299: \end{enumerate}
300:
301: The disk is represented in cylindrical polar coordinates $R$
302: (heliocentric distance) and $z$ (height above midplane), where the
303: $z$-axis is the rotation axis of the nebula. Assumption 1 reduces the
304: physical 3-dimensional disk to a 2-dimensional quadrant with zero flux
305: at the midplane, allowing us to suppress the azimuthal coordinate
306: $\theta$. Assumption 2 allows the vertical and radial dimensions of the
307: disk to be decoupled to form a 1+1-dimensional framework: at each radial
308: gridpoint resides an independent vertical structure model. At each
309: timestep, energy balance between viscous heating and radiative cooling
310: is solved within in a single annulus.
311:
312: \subsection{Radial Diffusion}
313: \label{raddiff}
314:
315: We calculate the radial motion of mass using the surface density
316: diffusion equation of Lynden-Bell and Pringle (1974):
317: \begin{equation}
318: {\partial \Sigma \over \partial t} = {3 \over R} {\partial \over
319: \partial R} \left [ R^{1/2} {\partial \over \partial R} \left ( \Sigma
320: \nu R^{1/2} \right ) \right ] .
321: \label{diffusion}
322: \end{equation}
323: Eq. \ref{diffusion} is valid when the protostar's accretion rate is
324: small, so the sun is almost fully assembled at the beginning of the
325: simulation. This assumption is consistent with the T-Tauri phase of
326: disk evolution. We take the viscosity $\nu$ governing radial diffusion
327: to be the midplane value.
328:
329: To calculate $\nu_{\rm midplane}$, we make the following
330: assumptions about angular momentum transport in the disk:
331:
332: \begin{enumerate}
333:
334: \item {\it Viscous stresses follow the standard $\alpha$-viscosity
335: prescription} (Shakura and Syunyaev 1973). We assume the
336: magnetorotational instability (Balbus and Hawley 1991; hereafter MRI)
337: provides the requisite turbulent viscosity.
338:
339: \item {\it The grains and gas are well mixed.} This assumption allows us
340: to track the disk's solid surface density evolution by modeling only a
341: single fluid.
342:
343: \end{enumerate}
344:
345: The turbulent viscosity is
346: \begin{equation}
347: \nu = \frac{2}{3} \alpha c_s H ,
348: \label{visc}
349: \end{equation}
350: where $c_s$ is the isothermal sound speed and $H$ is the modified
351: pressure scale height, softened into a nonsingular form (Milsom et al.
352: 1994):
353: \begin{equation}
354: H = { c_s / \Omega \over \sqrt{ 1 + \left ( {2
355: z^2 \Omega^2 / c_s^2} \right ) }} .
356: \label{scale}
357: \end{equation}
358: The angular velocity, neglecting disk self-gravity, is
359: \begin{equation}
360: \Omega = \left [ {G M_* \over \left ( R^2 + z^2
361: \right )^{3/2} } \right ]^{1/2} ,
362: \label{omega}
363: \end{equation}
364: and the isothermal sound speed is
365: \begin{equation}
366: c_s^2 = {R_g \over \mu} T .
367: \label{soundspeed}
368: \end{equation}
369: In parts of the disk where $H > z_{\rm surf}$, the disk surface height,
370: we substitute $z_{\rm surf}$ for $H$ in Eq. \ref{visc}. We take a
371: mean molecular weight of $\mu = 2.33$~g~mol$^{-1}$ (Ruden and Pollack
372: 1991). Based on recent global-disk MHD simulations (Lyra et al. 2008),
373: we choose $\alpha = 2 \times 10^{-3}$. We further discuss the interplay
374: between $\alpha$, disk mass and surface density profile, along with our
375: constraints on these parameters, in \S \ref{initial}. All free
376: parameters in the disk model are listed in Table \ref{freedisk}.
377:
378: %If MRI is the source of viscous turbulence, the grains must be small
379: %enough to be coupled to the gas, but not so small as to remove charge
380: %from the gas and lower the ionization fraction. We assume a mean grain
381: %radius of $0.258 \; \mu m$, slightly above the $0.1 \; \mu m$
382: %minimum required to keep enough free charge in the gas to sustain MRI
383: %turbulence (Turner et al. 2007).
384:
385: Although we assume turbulence is caused by MRI, we do not include the
386: radial variation of the viscosity coefficient $\alpha$ caused by the
387: dead zone (e.g. Kretke and Lin 2007, Reyes-Ruiz et al. 2003).
388: %We will
389: %investigate the effect of the dead zone on disk evolution in a
390: %forthcoming paper.
391: However, we do account for the cessation of MRI
392: turbulence in the tenuous outer regions of the disk, where ions and
393: neutrals may be only weakly coupled. \cite{hawley98} find that neutrals
394: participate fully in MRI turbulence only when the ion-neutral collision
395: timescale $T_{\rm coll}$ is much less than the orbital timescale;
396: $T_{\rm coll} \leq 0.01 / \Omega$. Ions and neutrals decouple and MRI
397: turbulence ceases when $T_{\rm coll} \geq 100 / \Omega$. In the
398: transition region between coupled and decoupled neutral-ion interaction,
399: where $0.01 \leq T_{\rm coll} \Omega \leq 100$, we decrease $\log
400: \alpha$ linearly from its fiducial value.
401:
402: To calculate $T_{\rm coll}$, we find the mean free time of a neutral
403: particle traveling through a gas of moving ions. The neutrals and ions
404: collide at an average angle of $90^{\circ}$ and thus a mean impact
405: velocity of
406: \begin{equation}
407: v_{\rm coll} = \sqrt{v_n^2 + v_i^2} = v_n \left (1 + {\mu_n \over \mu_i}
408: \right )^{1/2} ,
409: \label{vcoll}
410: \end{equation}
411: where $\mu_n$ and $\mu_i$ are the molecular weights of neutrals and
412: ions, respectively. The mean collision time is then
413: \begin{equation}
414: T_{\rm coll} = {1 \over \pi \left ( a_n + a_i \right )^2 n_i v_n
415: \left (1 + \mu_n / \mu_i \right )^{1/2}} ,
416: \label{tcoll}
417: \end{equation}
418: where $a_n$ and $a_i$ are the radii of neutrals and ions, $v_n =
419: \sqrt{3} c_s$ is the mean velocity of a neutral and $n_i$ is the number
420: density of ions:
421: \begin{equation}
422: n_i = {f_i \rho \over m_i}
423: \label{ni}
424: \end{equation}
425: We assume an ionization fraction of $f_i = 10^{-9}$ and an ion mass of
426: $m_i = 30$~AMU. $f_i$ and $m_i$ are listed along with other free
427: parameters in the simulation in Table \ref{freedisk}.
428:
429: \subsection{Vertical Structure}
430:
431: Since the disk midplane viscosity is determined by the balance between
432: viscous energy generation and radiative energy losses at the
433: photosphere, we calculate full vertical structure models at each
434: radial grid point. We use the following assumptions about the mass and
435: energy balance in the vertical direction:
436:
437: \begin{enumerate}
438:
439: \item {\it The disk stays in hydrostatic equilibrium}. Even though the
440: mass distribution evolves, the vertical sound crossing timescale is much
441: less than the radial diffusion timescale.
442:
443: \item {\it We neglect stellar irradiation as an energy source, as it has
444: little effect on the midplane temperature}. In section \ref{initial} we
445: show that the disk is self-shadowed ($H/R$ decreases with $R$), making
446: this assumption self-consistent with our simulation results.
447:
448: \end{enumerate}
449:
450: A system of three coupled differential equations specifies the vertical
451: structure of the disk. First, the disk is in hydrostatic equilibrium:
452: \begin{equation}
453: {\partial P \over \partial z} = -\rho \Omega^2 z .
454: \label{hydrostatic}
455: \end{equation}
456: Second, the viscous energy generation rate per unit volume (Pringle
457: 1981) is
458: \begin{equation}
459: {\partial F \over \partial z} = \frac{9}{4} \nu \Omega^2 \rho .
460: \label{flux}
461: \end{equation}
462: Third, where vertical energy transport is regulated by radiative
463: diffusion and energy transport in the radial direction is negligible
464: (thin disk assumption), the energy equation is
465: \begin{equation}
466: {\partial T \over \partial z} = {\partial P \over \partial z} \nabla {T
467: \over P} .
468: \label{tempgrad}
469: \end{equation}
470:
471: To calculate the thermodynamic gradient $\nabla = d \ln T / d \ln P$, we
472: use the Schwarzschild criterion for stability against convection
473: (Kippenhahn and Weigert 1994):
474: \begin{equation}
475: \nabla = \left \{ \begin{array}{l} \nabla_{\rm rad},
476: \nabla_{\rm rad} \le \nabla_{\rm ad} \\ \nabla_{\rm conv},
477: \nabla_{\rm rad} > \nabla_{\rm ad} \end{array} \right. .
478: \label{nabla}
479: \end{equation}
480: In Eq. \ref{nabla}, $\nabla_{\rm ad}$ is the adiabatic gradient,
481: which is $2/7$ for a diatomic gas; $\nabla_{\rm conv}$ is the convective
482: gradient; and $\nabla_{\rm rad}$ is the radiative gradient:
483: \begin{equation}
484: \nabla_{\rm rad} = \frac{3}{4} {\kappa P F \over a c \Omega^2 z T^4} ,
485: \label{nablarad}
486: \end{equation}
487: where $a$ is the radiation density constant and $\kappa$ is the local
488: Rosseland mean opacity (for a description of the opacity tables used in
489: this calculation, see \S \ref{opsec}).
490:
491: We use mixing-length theory to calculate $\nabla_{\rm conv}$. We begin
492: the convection treatment by defining two dimensionless quantities:
493: \begin{equation}
494: W \equiv \nabla_{\rm rad} - \nabla_{\rm ad}
495: \label{convw}
496: \end{equation}
497: and
498: \begin{equation}
499: U \equiv {3 a c T^3 \over c_P \rho^2 \kappa \ell_m^2}
500: \sqrt{8 \varepsilon} .
501: \label{convu}
502: \end{equation}
503: In Equation \ref{convu}, $c_P$ is the specific heat at constant pressure
504: and $\ell_m$ is the mixing length. As in \cite{milsom94}, we define
505: $\ell_m$ as the minimum of the convection zone top or the pressure scale
506: height $H$ (Equation \ref{scale}). $\varepsilon$ is an analytical
507: softening of the ratio $H / \Omega^2 z$, eliminating the singularity at
508: $z = 0$:
509: \begin{equation}
510: \varepsilon = \sqrt{1 \over \Omega^4 (1 + z^2 / H^2)} .
511: \label{varparm}
512: \end{equation}
513: Defining a new variable $\xi$ as
514: \begin{equation}
515: \xi \equiv \nabla_{\rm conv} - \nabla_{\rm ad} + U^2 = \sqrt{\nabla_{\rm
516: conv} - \nabla_e} + U
517: \label{xieqn}
518: \end{equation}
519: (where $\nabla_e$ is the thermodynamic gradient of an individual
520: convective element), we can then solve the cubic equation
521: \begin{equation}
522: \left ( \xi - U \right )^3 + \frac{8U}{9} \left ( \xi^2 - U^2 - W \right
523: ) = 0
524: \label{cubic}
525: \end{equation}
526: to find $\nabla_{\rm conv}$. For the derivations of Equations
527: \ref{xieqn} and \ref{cubic} see Chapter 7 of \cite{kippenhahn94}.
528:
529: Finally, we close the system of equations with the ideal gas equation of
530: state,
531: \begin{equation}
532: P = \left ({R_g \over \mu} \right ) \rho T .
533: \label{idealgas}
534: \end{equation}
535:
536: \subsection{Computational Methods and Boundary Conditions}
537:
538: We begin the simulation with a specified surface density profile for the
539: disk, $\Sigma \propto r^{-3/2}$ (Hayashi 1981). The total disk mass
540: used to normalize $\Sigma(r)$ and the viscosity coefficient $\alpha$ are
541: determined by constraints of gravitational stability
542: % presence of
543: %crystalline silicates in comets
544: and the need for enough mass to form
545: giant planets. For full details of how the $t = 0$ disk is constructed,
546: see \S \ref{initial}. We also assume the disk is embedded in a medium
547: with an ambient temperature $T_{\rm amb}$ (for example, a molecular
548: cloud core). This places a firm lower limit on the nebula temperature
549: and forces inactive regions of the disk to assume an isothermal vertical
550: structure.
551:
552: As in \cite{bell97}, we place the disk surface at
553: optical depth $\tau = 0.03$. We begin the vertical structure
554: calculation with guesses for the free parameters $T_{\rm surf}$ and
555: $\rho_{\rm surf}$, the density and temperature at the $\tau = 0.03$
556: surface.
557:
558: The net flux leaving the disk surface is determined by the accretion
559: temperature $T_{\rm acc}$, generated solely by viscous stresses:
560: \begin{equation}
561: F = \sigma T_{\rm acc}^4 .
562: \label{teff}
563: \end{equation}
564: We require a value of $T_{\rm acc}$ at the disk surface to to start
565: integrating Equations \ref{flux} and \ref{tempgrad}. The free parameter
566: $T_{\rm surf}$ is the flux sum of the accretion temperature $T_{\rm
567: acc}$ and the ambient temperature $T_{\rm amb}$, invoking the Eddington
568: approximation:
569: \begin{equation}
570: T_{\rm surf}^4 = T_{\rm amb}^4 + \frac{3}{4} \left [ \tau + f \left (
571: \tau \right ) \right ] T_{\rm acc}^4 .
572: \label{fluxsum}
573: \end{equation}
574: In equation \ref{fluxsum}, $f(\tau)$ is the Hopf function, with a value
575: of 0.601242 for $\tau = 0.03$ (Bell et al. 1997; Mihalas 1978).
576:
577: After calculating $P$ from Eq. \ref{idealgas}, the variables of
578: integration $T_{\rm acc}$, $P$ and $F$ are all specified. We find the
579: height of the $\tau = 0.03$ surface by solving the following equation
580: for z:
581: \begin{equation}
582: { 1 \over \Omega^2 \, z } = {\tau \over \kappa(\rho, T) \, P} .
583: \label{zsurfeqn}
584: \end{equation}
585: Integration proceeds inward from the disk surface to the midplane using
586: the Runge-Kutta method with adaptive stepsize control (Press et al.
587: 1992). We repeat the integration, adjusting $T_{\rm surf}$ and
588: $\rho_{\rm surf}$ with the Newton-Raphson algorithm (Press et al. 1992),
589: until a solution is found that satisfies
590: \begin{equation}
591: F_{\rm midplane} = 0
592: \label{fluxconv}
593: \end{equation}
594: (following Assumption 1, \S \ref{framework}) and
595: \begin{equation}
596: 2 \int_{z=0}^{z_{\rm surf}} \rho \: dz = \Sigma .
597: \label{sdconv}
598: \end{equation}
599:
600: After the vertical structure is solved for each radius, we use the
601: resulting $\nu_{\rm midplane}$ to update $\Sigma(r)$. We implement
602: implicit finite differencing to solve Eq. \ref{diffusion},
603: adjusting the timestep $\Delta t$ so that surface density varies by a
604: maximum of 1.5\% during a single timestep (e.g. Ruden and Pollack 1991).
605:
606: The radial grid is equispaced in $r^{1/2}$ and contains 600 cells
607: between the inner boundary at $0.3$~AU and the initial outer boundary at
608: $30$~AU. We implement a zero viscous-stress inner boundary condition,
609: so matter falls directly onto the star from the innermost gridpoint.
610: Although the actual accretion shock is at a much smaller radius than
611: $0.3$~AU, \cite{ruden86} find that the location of the inner boundary
612: $R_{\rm in}$ is unimportant as long as $R_{\rm out} >> R_{\rm in}$. We
613: allow the disk to expand freely from the original outer boundary, so new
614: radial grid cells are added to the disk as the simulation progresses.
615:
616: \subsection{Opacities}
617: \label{opsec}
618:
619: Solving Eq. \ref{nablarad} requires knowing local mean opacity.
620: Semenov et al. (2003) calculated Rosseland mean opacities due to icy
621: grains for temperatures down to 5K. The authors modeled several
622: different grain types, including spheres, aggregates, and porous grains.
623: We use the 5-layered sphere topology, where each grain consists of a
624: silicate/iron nucleus covered with successive layers of volatiles as the
625: temperature decreases. Following Semenov et al., we assume the grains
626: condensed from gas of solar composition, with the abundances given by
627: Helling et al. (2000). These abundances are primarily derived from the
628: catalog of Anders and Grevesse (1989). We list the abundances of the
629: most common elements in Table \ref{solarcomp}, along with the types of
630: solids they can form: rock, metal, ice, or refractory CHON (e.g.
631: graphite or kerogen; see Jessberger et al. 1988).
632:
633: At temperatures above 1200K, iron and silicates sublimate and the main
634: opacity source switches from dust to molecular gas. Semenov et al.
635: (2003) caution that the gas opacities in their model are only
636: approximate in the temperature range where dust and gas opacity are
637: comparable. For temperatures above 1000K, we use the opacities of
638: Ferguson et al. (2005), updated for our chosen set of solar abundances
639: (J. Ferguson, private communication). It should be noted that the
640: number of atomic and molecular lines included in the calculation of an
641: opacity table has a greater impact on the resulting opacities than minor
642: changes in assumed solar composition.
643:
644: Regions where opacity is a strong function of temperature can cause
645: convergence problems for the Newton-Raphson algorithm, which will
646: oscillate between two solutions with similar temperatures but quite
647: different opacities. To mitigate this problem, we smoothed the raw
648: opacity tables with an averaging filter. Figure \ref{opplot} shows
649: opacity as a function of temperature and density in our model.
650:
651:
652: \subsection{Initial Conditions}
653: \label{initial}
654:
655: Constraints on the early stages of the solar nebula come from
656: observations of Class 1 protostars and T-Tauri stars and the presence
657: and orbital configuration of the planets. We require our $t = 0$ disk
658: to meet two criteria:
659:
660: %analysis of the
661: %chemical composition of comets and other pristine bodies,
662:
663: \begin{enumerate}
664:
665: \item The disk is gravitationally stable to axisymmetric perturbations
666: at all radii.
667:
668: %\item The temperature in the inner nebula, $r \leq 2$~AU, is hot enough
669: %to produce crystalline silicates ($> 800$~K; Gail 1998).
670:
671: \item The disk is massive enough to produce Jupiter and Saturn by core
672: accretion within the observed mean protostellar disk lifetime of
673: 2--3~Myr (Haisch et al. 2001).
674:
675: \end{enumerate}
676:
677: To test whether a particular disk meets Criterion 1, we calculate the
678: Toomre Q parameter at every radial gridpoint. Gravitational stability
679: requires that
680: \begin{equation}
681: Q = {c_s \Omega \over \pi G \Sigma} > 1 .
682: \label{toomreq}
683: \end{equation}
684: Gravitationally unstable disks will fragment and form stellar or
685: substellar clumps on a $10^3$ year timescale. The lack of a binary
686: companion for the sun and the near-circular, non-chaotic orbits
687: displayed by all the planets makes gravitational fragmentation anytime
688: during solar nebula evolution unlikely.
689:
690: Our $Q > 1$ criterion determines the combinations of viscosity parameter
691: $\alpha$, mass distribution and total disk mass that create viable
692: accretion disks. High values of $\alpha$ act to stabilize the disk by
693: increasing energy generation and therefore sound speed. Large disk
694: masses destabilize the disk by increasing $\Sigma$. Surface density
695: profiles that decline steeply with radius damp gravitational instability
696: by loading the mass preferentially toward the inner disk, where
697: dissipative shear stresses are strongest.
698:
699: Hersant et al. (2001; hereafter HGH) used an evolving $\alpha$-disk
700: model and measured solar-system D/H ratios to constrain the structure of
701: the solar nebula. They chose the surface density profile $\Sigma
702: \propto R^{-1}$. Even with an extremely high turbulent energy
703: generation rate, $\alpha \geq 0.01$, disks with this surface density
704: distribution are gravitationally stable only if $M_{\rm disk} \leq 0.08
705: M_{\odot}$. However, Criterion 3 requires that there be sufficient
706: solid mass to form Jupiter and Saturn within $\sim 3$~Myr. Assuming a
707: gas/solid mass ratio of 70 (Pollack et al. 1996), a disk with $\Sigma
708: \propto R^{-1}$ out to 30~AU and $M_{\rm disk} = 0.08 M_{\odot}$ has a
709: solid surface density of only 6~g~cm$^{-2}$ in Saturn's feeding zone at
710: 9.5~AU. \cite{hubickyj05} found that a planet forming from
711: 6~g~cm$^{-2}$ of solids at 5~AU takes more than 13~Myr to initiate rapid
712: gas accretion, necessary for forming a massive atmosphere. Since planet
713: formation proceeds more slowly at larger heliocentric distances (Eq.
714: \ref{planetgrowth}), this disk is unlikely to be able to form Saturn.
715: We thus require a more massive disk, with a steeper surface density
716: profile to keep the outer regions gravitationally stable.
717:
718: We choose a disk mass $M_{\rm disk} = 0.12 M_{\odot}$ and a surface
719: density $\Sigma \propto R^{-3/2}$ (Hayashi 1981). We take $\alpha = 2
720: \times 10^{-3}$, the approximate azimuthally and vertically averaged
721: value at Jupiter's heliocentric distance in the global MHD turbulence
722: simulations of Lyra et al. (2008). Our solar nebula has inner and outer
723: radii of $R_{\rm in} = 0.3$ AU and $R_{\rm out} = 30$~AU, covering the
724: range of heliocentric distances where planets exist today. It is worth
725: noting that studies of the relative motions of dust and gas predict a
726: much larger outer radius: the gas drag on cm-sized bodies can cause them
727: to be quickly lost to the Sun unless they have a very large distance to
728: migrate (Stepinski and Valageas [1997]; Ciesla and Cuzzi [2006]).
729: Mitigating the deleterious effect of gas drag on large grains may
730: require that the nebula has an additional low-density component outside
731: $R = 30$~AU, which we do not model in this work.
732:
733: We assume our $t = 0$ disk coincides with the beginning of the sun's
734: T-Tauri phase. Most of the sun's mass is already in place: $M_* = 0.95
735: M_{\odot}$ at the start of our simulation. We assume an ambient
736: temperature of 20~K, approximating a disk embedded in the remnants of a
737: molecular cloud. All free parameters in the disk model are listed in
738: Table \ref{freedisk}.
739:
740: Although our model neglects solar irradiation, we find that the giant
741: planet-forming region of our t=0 disk is in fact shadowed. The maximum
742: aspect ratio $H / R$ of the $\tau = 2/3$ photosphere occurs at $R = 2$
743: AU. Beyond this radius, no part of the disk receives direct sunlight.
744: The shadow persists until $t \sim 1$~Myr, when regions with $R > 20$~AU
745: attain an aspect ratio greater than the local maximum at 2~AU. In this
746: way, our disk evolution is broadly consistent with model T-Tauri disk
747: SEDs, which require flaring to explain observed spectra (e.g. Chiang et
748: al. 2001), yet the giant planet-forming region ($\sim 5$--20~AU) still
749: receives no solar heating during the icy planetesimal formation epoch.
750:
751: \subsection{Disk Model Results}
752: \label{viscous}
753:
754: \subsubsection{Surface Density}
755: \label{surfacedensity}
756:
757: We run the solar nebula model for 2~Myr, about the median age of
758: observed protostellar disks (Haisch et al. 2001). Figure \ref{sdplanet}
759: shows the evolution of the disk surface density profile in the giant
760: planet-forming region, 5-20~AU (Tsiganis et al. 2005). The surface
761: density profile at $t = 0$ is shown on the upper left. Subsequent
762: panels show snapshots of $\Sigma(R)$ as the simulation progresses
763: plotted in black. For reference, surface density profiles from
764: preceding plot panels are retained in gray. Figure \ref{sdwholedisk}
765: shows the surface density evolution of the entire disk as it expands
766: from 30~AU at $t = 0$ to $> 80$~AU after 2~Myr.
767:
768: From examining Fig. \ref{sdplanet}, we can divide the evolution of the
769: solar nebula's giant planet-forming region into two epochs. In Epoch 1,
770: the surface density profile flattens as mass is redistributed from the
771: inner disk to the outer disk. At $t = 0$, the net mass flow is outward
772: everywhere except at $R < 1$~AU, which is accreting inward toward the
773: star. A wave of local surface density enhancement consequently
774: propagates outward, reaching $\sim 10$~AU at $t = 3 \times 10^4$ yr and
775: dissipating near $\sim 17$~AU at $t = 3 \times 10^5$ yr. At the end of
776: Epoch 1, at $t \sim 5 \times 10^5$~yr, the solar nebula has assumed a
777: surface density profile flatter than $\Sigma \propto R^{-1}$ (equal mass
778: in each annulus) in all but the inner $0.5$~AU of the disk. The effect
779: of this mass redistribution epoch, therefore, is to drive the surface
780: density distribution toward uniformity.
781:
782: In Epoch 2, $t > 5 \times 10^5$~yr, surface density decreases with time
783: everywhere in the disk. In Fig. \ref{sdwholedisk}, we see the solar
784: nebula grow more tenuous as it expands: after 2 Myr, the total disk mass
785: is 0.067~$M_{\odot}$ spread over 80~AU, as opposed to 0.12~$M_{\odot}$
786: within 30 AU at the beginning of the simulation. The protosun reaches a
787: final mass of 1.003~$M_{\odot}$.
788:
789: Figure \ref{sdann} shows surface density as a function of time at 5~AU,
790: 10~AU, 15~AU and 20~AU. At each heliocentric distance, we see the
791: temporary mass growth as the density wave passes through. The largest
792: relative surface density increase $\Delta \Sigma / \Sigma_0$ occurs at
793: 11.5~AU, where the initial surface density of 442~g~cm$^{-2}$ grows by
794: 28\% to 558~g~cm$^{-2}$ in 85,000~yr. {\it This surface density
795: enhancement moving outward through the giant planet-forming region
796: strongly favors solid planet-core formation.}
797:
798: At the beginning of the sun's T-Tauri phase, before substantial grain
799: growth occurs, gas and grains are well mixed: accreting gas drags grains
800: with it. If gas moves inward with a constant mass flux, surface density
801: decreases with time everywhere in the disk and solids are lost into the
802: sun. Our net outward gas flow deposits small grains in the 5--15~AU
803: region within the planetesimal formation timescale of $\sim 5 \times
804: 10^4$~yr (Hubbard and Blackman 2006). {\it These grains provide surface
805: area for ice deposition and add solid mass to the giant planet feeding
806: zones, speeding up solid core formation.}
807:
808: The existence of Epoch 1 and consequent driving of the surface density
809: distribution toward uniformity is a robust result that we see even using
810: $t = 0$ surface density profiles with power-law indices of 0.7 and 1.
811: Our chosen fiducial disk remains stable to axisymmetric perturbations
812: (c.f. Eq. \ref{toomreq}) throughout its evolution even as mass
813: flows outward.
814:
815: \subsubsection{Mass Flux}
816: \label{massflux}
817:
818: In their study of the similarity solutions of viscously evolving disks,
819: \cite{lyndenbell74} found that the natural behavior of an accretion disk
820: is to diffuse and spread out, creating a net inward flow in the inner
821: disk, a turnaround radius, and an outward decretion flow in the outer
822: disk. \cite{ruden86} uncovered the same behavior in numerical
823: simulations of the evolution of the minimum-mass solar nebula. Why
824: haven't many other $\alpha$-disk models predicted the outward mass flow
825: into the trans-Saturnian region that could so improve the prospects for
826: giant planet formation? Most previous evolving disk models such as HGH
827: and the pseudoevolutionary sequence of \cite{bell97} have required the
828: disk to be in steady state. In steady-state models, each annulus
829: receives an equal inward-driven mass flux with the sun serving as the
830: sink particle. There is therefore no chance for any part of the disk
831: interior to $R_{\rm out}$ at $t = 0$ to gain mass. Note that the
832: assumption of steady-state, net inward accretion is still consistent
833: with viscous expansion of the disk from its original outer boundary, as
834: occurs in the HGH models and this work.
835:
836: The non-steady-state evolving disk model of \cite{ruden91} assumed
837: turbulent stresses were caused solely by convection, and set the
838: viscosity to zero in optically thin regions where radiative transport is
839: efficient. Far from the star, the shallow potential well does not allow
840: the disk to attain high volume densities, so reaching an optical depth
841: of $\tau > 1$ requires more mass than in the inner disk. At $R =
842: 30$~AU, the outer radius at the beginning of the Ruden and Pollack
843: calculation, convection ceases almost immediately and the disk becomes
844: quiescent. The inward accretion flow then removes enough mass from
845: progressively interior annuli to make them optically thin, leaving
846: behind a remnant surface density profile where $\Sigma$ {\it increases}
847: with $R$.
848:
849: In the Ruden and Pollack simulation, mass is not redistributed from the
850: inner disk to the outer disk---the slope of $\Sigma(R)$ changes from
851: negative to positive because the inner parts of the disk remain
852: optically thick for the longest, participate in convection-driven
853: accretion the longest, and therefore lose the most mass. However, in
854: their analysis of the coupling between the kinetic stress tensor
855: governing hydrodynamic stability, the magnetic stress tensor, and the
856: shear field, \cite{hawley99} find that MRI is uniquely capable of
857: providing turbulent angular momentum transport in disks. By relaxing
858: the steady-state condition, but still permitting angular momentum
859: transport by MRI in optically thin regions (though MRI ceases if ions
860: and neutrals decouple; see \S \ref{raddiff}), we allow outward flows to
861: develop and viscous stress to redistribute mass from the inner disk to
862: the giant planet-forming region.
863:
864: %Indeed, the immediate cessation of viscous
865: %evolution at $R \sim 30$~AU found by Ruden and Pollack is inconsistent
866: %with the work of HRH, who compare their disk model results with
867: %observed D/H ratios in planet atmospheres,
868: %satellites and comets to argue that the solar nebula
869:
870: Figure \ref{mdot} shows time snapshots of the mass accretion rate as a
871: function of radius. The location with $\dot{M} = 0$ is the turnaround
872: between inward accretion and outward ``decretion.'' This turnaround
873: radius moves outward over time, so that more and more of the disk is
874: accreting toward the star. This behavior is required to open a hole in
875: the inner disk and make the transition from classical T-Tauri to
876: non-accreting, weak-lined T-Tauri star. Instead of accreting onto the
877: star, the outwardly decreting material ($R > 15$~AU) will eventually be
878: lost to photoevaporation (Alexander et al. 2006).
879:
880: %Our prediction of a decretion disk is consistent with the analysis of Desch (2007), who
881: %combined the masses and starting locations of the giant planets in the Nice model
882: %(Tsiganis et al. 2005) with an a net outward mass flow combined with photoevaporative
883: %truncation of the disk.
884:
885: \subsubsection{Midplane Temperature}
886:
887: Figure \ref{midplanetemp} shows shapshots of the temperature of the
888: protostellar disk midplane at $R < 30$~AU as the disk evolves. Outside
889: of 30~AU, the entire disk (midplane and above) is isothermal at the
890: ambient temperature throughout the simulation (20~K, plotted with a
891: dash-dotted line). The first three panels of Fig. \ref{midplanetemp}
892: show that during Epoch 1, areas of the disk that increase in mass are
893: correspondingly heated: the temperature at 10~AU increases from 40~K at
894: $t = 0$ to 49~K at $3 \times 10^4$~yr.
895:
896: We first evaluate the suitability of our model disk for delivering
897: crystalline silicates to comets, as detected by \cite{keller06} in comet
898: 81P/Wild 2 dust captured by the Stardust spacecraft. The dashed line in
899: Fig. \ref{midplanetemp} shows the minimum temperature for silicate
900: crystallization of 800~K (Gail 1998). (Note that forsterite formation
901: requires $T > 1067$~K; Hallenbeck et al. [2000].) The inner nebula
902: retains temperatures above 800~K for $10^5$~yr. \cite{dullemond06}
903: calculated the formation and transport rates of crystalline silicates in
904: a 1-d evolving $\alpha$-disk model with infall from a collapsing cloud.
905: Relaxing the steady-state assumption, they found that viscous spreading
906: of the disk, rather than turbulent radial mixing, could relocate
907: crystalline silicates produced early on in the inner, hot ($T \ga
908: 800$~K) parts of the disk to distances of $R > 15$~AU, where comets
909: formed. A gradual outward push of crystalline silicates is consistent
910: with the nebula flow pattern observed in this work, and may explain the
911: anticorrelation between disk age and crystallinity in the inner disk ($R
912: < 10$~AU) discussed by \cite{vanboekel04}. However, the danger that
913: crystalline silicates transported outward through the midplane would be
914: swept up by protoplanets before ever reaching the comet-forming region
915: led \cite{ciesla07} to favor more rapid transport than the decretion
916: flow in this model.
917:
918: %One danger when crystalline silicates are formed in the inner nebula and
919: %transported outward through the disk midplane is that they will be swept
920: %up by forming protoplanets and never incorporated into comets.
921: %Fortunately, there are other possibilities for both the formation and
922: %transport of silicates.
923: %Scott and Krot (2005) inferred high cooling rates of chondritic
924: %meteorites, on the order of 10--1000 K~hr$^{-1}$, favoring localized,
925: %brief heating (as from a shock) as the source of crystalline silicates
926: %in asteroids. In a study of melilite mantle formation on type B CAIs,
927: %Richter et al. (2006) find that temperatures greater than $\sim 1300$~K
928: %were sustained for only $\sim 10^4$~yr---not long enough to erase
929: %melilite-\aa kermanite zoning.
930:
931: %Grains annealed in the quiescent disk could be transported to
932: %protocometary material at 5--20~AU by the decretion flow described in \S
933: %\ref{massflux} that occurred in the crystalline silicate-forming region
934: %during the first $2 \times 10^4$~yr.
935:
936: %Turbulent mixing with $\alpha \ge 4
937: %\times 10^{-4}$ could then deliver crystalline silicates to the
938: %asteroid- and comet-forming regions within a few Myr (Desch 2007).
939:
940: %The need for a hot, long-lived inner
941: %solar nebula with efficient radial mixing is far from clear.
942:
943: The dotted line on Fig. \ref{midplanetemp} shows the H$_2$O ice line
944: at 160~K (Collings et al. 2004). Although we require a chemical kinetic
945: model to determine the exact position and width of the ice line (since
946: ice deposition is not instantaneous but proceeds at a finite rate
947: limited by the gas velocity of H$_2$O and the available grain surface
948: area), we can nevertheless estimate its location and compare with
949: previous disk models. In our simulation, the H$_2$O ice line is always
950: inside the formation zones of Saturn, Uranus and Neptune, and crosses
951: Jupiter's feeding zone (5.2~AU) at $5 \times 10^4$~yr. All giant
952: planet cores, therefore, should have formed from icy grains.
953:
954: %In the HGH fiducial model, the H$_2$O ice line crosses Jupiter's feeding
955: %zone at $\sim 5 \times 10^5$~yr, with a total surface density of $\sim
956: %300$~g~cm$^{-2}$. Assuming a gas/solid mass ratio of 70 for icy regions
957: %(Pollack et al. 1996), the solid surface density available for forming
958: %Jupiter would be only $\sim 4.2$~g~cm$^{-2}$ in the HGH model.
959: %\cite{hubickyj05} find that forming Jupiter from 6~g~cm$^{-2}$ of solids
960: %takes 13~Myr. The HGH disk, therefore, must be extraordinarily
961: %long-lived in order to form Jupiter.
962:
963: After 2~Myr, the midplane outside 15~AU has become isothermal at the
964: ambient temperature. This temperature, therefore, has a profound effect
965: on the ice inventory of the solar nebula. The most volatile ices (CO,
966: CH$_4$ and N$_2$) have sublimation temperatures below 45~K. Simply
967: raising the ambient temperature above this value would ensure that none
968: of these ices could freeze on grains. However, C/H enrichments of
969: 20--40 over the solar value measured in Uranus and Neptune's atmospheres
970: indicate that direct gas accretion from the nebula cannot explain these
971: planets' composition (Baines et al. 1995). We therefore postulate that
972: the outer solar nebula midplane contained CO and CH$_4$ ice, making our
973: choice of ambient temperature appropriate. Further evidence for a cold
974: solar nebula comes from the observed noble gas enrichment in Jupiter's
975: atmosphere---trapping argon in icy planetesimals requires temperatures
976: of 30~K (Owen et al. 1999).
977:
978: %Late-stage accretion of
979: %refractory CHON grains similar to those observed in Halley's comet dust
980: %by \cite{jessberger88} cannot provide such a large carbon overabundance:
981: %a mechanism of tapping into the massive CO and CH$_4$ reservoirs is
982: %required.
983:
984: In the next section, we describe our chemical model of the solar nebula
985: midplane. We will use these results, along with the surface density
986: curves shown in Fig. \ref{sdplanet}, to predict the solar nebula's
987: solid mass distribution in \S \ref{planetesimal}.
988:
989: \section{Chemical Model}
990: \label{chemmodel}
991:
992: \subsection{Reaction Set}
993:
994: The foundation of our chemical model is the UMIST database RATE95
995: (Millar et al. 1997). We follow the chemistry of 211 species, 153
996: gas-phase and 58 ices. Our reaction network contains 2479 reactions,
997: including thermal desorption, gas-grain reactions and grain surface
998: reactions. No non-thermal desorption processes (e.g. cosmic ray
999: heating) are included.
1000:
1001: We make the simplifying assumption that radial and vertical motions of
1002: gas and dust are slow compared to chemical reaction timescales, and do
1003: not track the motion of gas and dust through the nebula. This
1004: assumption is true for grain surface reactions and freezeout/desorption,
1005: which have short timescales, but gas-phase chemistry proceeds more
1006: slowly. Furthermore, inward migration of cm-size particles can deplete
1007: the outer solar nebula of ice and concentrate each species at its
1008: condensation front (e.g. Ciesla and Cuzzi 2006). In this paper, we
1009: ignore relative motions of solid and gas.
1010:
1011: Since protostellar disks have a high optical depth to UV radiation, we
1012: neglect photoionization in the disk midplane model. However, we include
1013: cosmic rays at the standard interstellar flux in parts of the disk with
1014: a half-plane surface density lower than the penetration depth of
1015: 150~g~cm$^{-2}$ (Umebayashi and Nakano 1981). The cosmic-ray
1016: penetration depth is listed in Table \ref{freechem} with the other free
1017: parameters in the chemical model.
1018:
1019: The following sections describe our treatment of gas-grain and grain
1020: surface reactions and the binding energies we use to calculate those
1021: reaction rates.
1022:
1023: \subsubsection{Grain Surface Reactions}
1024:
1025: Following \cite{willacy06} and \cite{garrod06}, we use rate equations to
1026: calculate grain surface reaction rates. Rate equations work in the
1027: mean-field approximation, neglecting the stochastic variation of
1028: abundances on different grains. The accuracy of rate equations is
1029: compromised when a species reacts on the grain surface faster than it
1030: can adsorb or desorb: this drives the reactant abundance per grain to
1031: less than unity and the mean-field approximation is no longer valid
1032: (Caselli et al. 1998). However, unlike stochastic methods that account
1033: for the discrete nature of the grains, solving rate equations is
1034: computationally tractable even for large-scale simulations. For
1035: example, a model that uses direct integration of the master equation
1036: (the probability distribution governing reaction rates) has the number
1037: of equations in the simulation increasing exponentially with the number
1038: of species (Barzel and Biham 2007).
1039:
1040: The species most likely to violate the mean-field approximation is
1041: atomic hydrogen, which can scan grain surfaces and find reaction
1042: partners quickly due to its low atomic weight. While heavier atoms move
1043: across grain surfaces by thermal hopping, Hornekaer et al. (2003) find
1044: that H atoms often tunnel between adjacent vacancies in the grain
1045: surface lattice. This result conflicts with previous work by Katz et
1046: al. (1999), who found that the quantum tunneling mechanism is
1047: improbable. Although the question of whether quantum tunneling or
1048: thermal hopping is primarily responsible for hydrogen mobility on grain
1049: surfaces is not resolved, Ruffle and Herbst (2000) show that rate
1050: equation simulations limiting H motion to thermal hopping only---and
1051: thus slowing down reactions involving atomic H---show good agreement
1052: with stochastic methods. We therefore assume thermal hopping regulates
1053: all grain surface reaction rates.
1054:
1055: The thermal hopping rate to an adjacent vacancy on the grain surface
1056: lattice is
1057: \begin{equation}
1058: t_{\rm hop}^{-1} = \nu_0 \exp \left ( -0.3 E_D / k T_g \right ) ,
1059: \label{hop}
1060: \end{equation}
1061: where $E_D$ is the binding energy, $\nu_0$ is the oscillation frequency
1062: between the ice and the grain surface, and $T_g$ is the grain
1063: surface temperature (assumed to be in equilibrium with the gas). The
1064: oscillation frequency is
1065: \begin{equation}
1066: \nu_0 = \sqrt{2 n_s E_D / \pi^2 m} ,
1067: \label{osc}
1068: \end{equation}
1069: where $n_s \sim 1.5 \times 10^{15} {\rm cm}^{-2}$ is the surface density
1070: of lattice vacancies (Hasegawa et al. 1992; Table \ref{freechem}) and
1071: $m$ is the reactant mass. Typical oscillation frequencies are
1072: 1--3$\times 10^{12}$~s$^{-1}$ (Herbst et al. 2005). Following
1073: \cite{willacy06}, we assume that only atoms can move across grain
1074: surfaces: molecules, which have higher binding energies, are stationary.
1075: All grain surface reactions thus involve at least one atom.
1076:
1077: The grain-surface reaction rate ($s^{-1}$) is given by
1078: \begin{equation}
1079: t_{\rm reac}^{-1} = { t_{\rm hop,1}^{-1} + t_{\rm hop,2}^{-1}
1080: \over 4 \pi a^2 n_s n_g } ,
1081: \label{surfrate}
1082: \end{equation}
1083: where $t_{\rm hop,1}$ and $t_{\rm hop,2}$ are the thermal hopping rates
1084: of reactants 1 and 2, $a$ is the grain radius and $n_g$ is the number
1085: density of grains. As in the molecular cloud model, we assume a mean
1086: grain radius of $0.258 \; \mu m$ (Table \ref{freechem}). We take a
1087: grain number density of $n_g / (n_{\rm H} + 2 n_{\rm H2}) = 10^{-12}$
1088: (Table \ref{freechem}), consistent with our assumption of well-mixed
1089: grains and gas.
1090:
1091: \subsubsection{Gas-Grain Reactions}
1092:
1093: We assume that all species can freeze out on grains except He, which has
1094: a very low binding energy, $E_D / k = 100$~K (Tielens and Hagen 1982).
1095: The freeze-out rate is
1096: \begin{equation}
1097: k_f = S_x \left \langle \pi a^2 n_g \right \rangle C_i v_x n_x ,
1098: \label{freezerate}
1099: \end{equation}
1100: where $S_x$ is the sticking coefficient (assumed to be 0.3 for H atoms
1101: and 1 for all other species; Table \ref{freechem}), $v_x$ is the
1102: gas-phase velocity of species $x$, and $n_x$ is the gas-phase number
1103: density of $x$. The factor $C_i$ takes into account the attraction of
1104: positively charged ions to grains which, on average, are negatively
1105: charged (Umebayashi and Nakano 1980):
1106: \begin{equation}
1107: \left \{ \begin{array}{l} C_i = 1 \; {\rm for \; neutrals} \\ C_i = 1 +
1108: e/(akT) \; {\rm for \; single-charged \; positive \; ions.} \end{array}
1109: \right.
1110: \label{ci}
1111: \end{equation}
1112: In Eq. \ref{ci}, $k$ is the Boltzmann constant and $e$ is the
1113: electron charge. Once an ion hits a grain, we assume it is neutralized
1114: by an adsorbed free electron. The thermal desorption rate ($s^{-1}$) at
1115: which ice mantle species are returned to the gas is
1116: \begin{equation}
1117: k_t = \nu_0 \exp \left ( E_D / k T_g \right ) .
1118: \label{therm}
1119: \end{equation}
1120:
1121: %Although formation of $> 10 \mu m$ grains proceeds rapidly
1122: %in microgravity experiments (Blum et al. 2000), we assume the continued
1123: %presence of small grains and keep the mean grain radius fixed throughout
1124: %our simulation.
1125:
1126: \subsubsection{Binding Energies}
1127:
1128: The binding energies $E_D$ for each species depend on the surface onto
1129: which the molecules adsorb and the mixture of ices on that surface. If
1130: each ice were a pure solid, $E_D$ would simply be the sublimation energy
1131: and desorption physics could be calculated nonkinetically using vapor
1132: pressure. However, molecules in mixed ices may have substantially
1133: different binding energies than those in pure ices, so the sublimation
1134: energy approximation is inadequate.
1135:
1136: The most abundant astrophysical ice is H$_2$O, so the binding energy of
1137: each species under protostellar disk conditions is primarily determined
1138: by how it behaves on/in an amorphous water-ice matrix.
1139: \cite{collings04} group binding energies into three categories based on
1140: their behavior when co-deposited with H$_2$O: CO-like, water-like, and
1141: intermediate. The extremely volatile CO-like species include CO, N$_2$,
1142: O$_2$, and CH$_4$. Molecules of these species do not bond efficiently
1143: either with themselves or with water: they quickly diffuse through an
1144: amorphous H$_2$O matrix and desorb mainly at their sublimation
1145: temperatures, which are quite low ($<40$~K), with minor amounts
1146: remaining trapped in the H$_2$O ice.
1147:
1148: Molecules of water-like species, including NH$_3$ and CH$_3$OH, bind
1149: more strongly to water than to each other or to a metal substrate.
1150: These polar molecules form hydrogen bonds with water that keep them in
1151: the solid phase at higher temperatures than possible in their pure form.
1152: The data of \cite{collings04} show that ammonia-water mixtures are
1153: substantially less volatile than pure ammonia, which has a binding
1154: energy of only 1110~K (Hasegawa and Herbst 1993). NH$_3$ therefore takes
1155: on the binding energy of H$_2$O, 5700~K.
1156:
1157: Finally, intermediate species are trapped by H$_2$O to some extent, but
1158: do not have the same mobility as CO inside a water-ice matrix. Some
1159: co-desorption with H$_2$O occurs, but most desorption takes place at the
1160: sublimation energy. These species include H$_2$S, OCS and CH$_3$CN. We
1161: have not included the partial co-desorption with water and treat each
1162: molecule as having only one specific desorption tmeperature. Table
1163: \ref{binding} gives the binding energy data used in our model.
1164:
1165: \subsection{Preprocessing in Molecular Cloud}
1166:
1167: The ices in protostellar disks have undergone significant chemical
1168: processing before giant planet formation begins. To simulate this
1169: preprocessing, we model $10^6$ years of molecular cloud evolution to
1170: derive input abundances for the protostellar disk. The molecular cloud
1171: model uses the same species set and reactions as our disk model, with
1172: the addition of photoionization due to the interstellar UV field.
1173:
1174: The atoms that form the species in our simulation are H, He, C, N, O,
1175: and S. We also include Si adsorption, desorption and ionization as a
1176: tracer of the metal population. Except for hydrogen, we assume all
1177: material enters the molecular cloud in atomic form, as if coming from a
1178: recent supernova. We take $n_{\rm H} / (n_{\rm H} + 2 n_{\rm H2}) =
1179: 10^{-2}$ (Table \ref{freechem}), so that 99\% of hydrogen atoms are in
1180: H$_2$. We assume a density of $2 \times 10^4$~cm$^{-3}$, a temperature
1181: of 10~K and an extinction $A_V = 10$~mag.
1182:
1183: Not all of the atomic CHONS material is available to form ices: some
1184: will help form the solid grains where ice mantles deposit. To figure
1185: out the fraction of C, N, O, and S locked away in grains, we use the
1186: atomic abundance ratios in Halley's comet dust grains reported by
1187: \cite{jessberger88}: C:N:O:Si:S = 814:42:890:185:72. Assuming only
1188: trace amounts of silicon remain in the gas phase (gas/grain $< \:
1189: 0.001$), we calculate grain fractions for C, N, O and S as
1190: \begin{equation}
1191: n_{\rm grain} / n_{\rm gas} = \left ( {N_{\rm Si} \over N_x} \right
1192: )_{\odot} \left ( {N_x \over N_{\rm Si}} \right )_{\rm comet} .
1193: \label{dustfrac}
1194: \end{equation}
1195:
1196: Using the solar composition of \cite{helling00}, as in the dynamical
1197: disk model, we find $F_{\rm grain}$ = 0.4, 0.1, 0.2 and 0.85 for C, N, O
1198: and S respectively (Table \ref{freechem}). This portion of each atom's
1199: inventory does not participate in chemistry and is excluded from the
1200: simulation. The gas and ice abundances of major species after 1~Myr of
1201: molecular cloud evolution are listed in Table \ref{cloudoutput}. These
1202: abundances form the chemical initial conditions for our protostellar
1203: disk model.
1204:
1205: The major ice species formed in the molecular cloud are H$_2$O, NH$_3$,
1206: CO, and CH$_4$. More than 90\% of the C, N and O atoms that aren't part
1207: of refractory grains are in these compounds. Sulfur is sequestered in
1208: solid H$_2$S after the molecular cloud phase, but this composition is
1209: subsequently modified during disk evolution. Minor CNO carriers are
1210: HCN, N$_2$ and NO. Figure \ref{pies} shows the division of solid C, N
1211: and O atoms among the species in our model.
1212:
1213: Our molecular cloud model predicts total ${\rm CO / CH_4} = 0.14$. Both
1214: of these species reside in ice mantles. Abundances of major ices
1215: relative to H$_2$O are listed in Table \ref{iceratios}. Our CO/CH$_4$
1216: ice ratio is consistent with the \cite{aikawa08} models of prestellar
1217: cores, which give ${\rm CO / CH_4} = 0.43$ in an embedded core and ${\rm
1218: CO / CH_4} = 0.1$ in an isolated core. Likewise, Lodders (2003)
1219: concluded that the reactions
1220: \begin{equation}
1221: \begin{array}{ll}
1222: 3 {\rm H_2 + CO} \rightarrow {\rm CH_4 + H_2O} \\
1223: {\rm CO} \rightarrow {\rm C} \; ({\rm graphite}) + {\rm CO_2}
1224: \end{array}
1225: \label{noco}
1226: \end{equation}
1227: would deplete CO from presolar gas (note, however, that in our model
1228: CH$_4$ forms on grains and not in the gas). Lodders predicts ${\rm CH_4
1229: / H_2O} = 0.58$, whereas we find a methane-to-water ratio of 0.38 (Table
1230: \ref{iceratios}). This discrepancy reflects the fact that we have a
1231: higher input oxygen abundance (to match the opacity tables calculated by
1232: Semenov [2003]).
1233:
1234: We note that the dominant components of interstellar ice mantles are
1235: typically H$_2$O, NH$_3$, CO, CO$_2$, and CH$_3$OH (Charnley and Rodgers
1236: 2008), which suggests carbon oxide formation in the ISM should be
1237: favored and hydrocarbon formation kinetically inhibited. One possible
1238: way to reduce the ${\rm CH_4 / CO}$ ratio is to lower the assumed atomic
1239: hydrogen abundance, which would reduce the efficiency of hydrogenated
1240: molecule formation on the grains at early times when atomic carbon is
1241: abundant. However, the purpose of this work is to determine the ice
1242: mass available for giant planet formation. Since CH$_4$ and CO have such
1243: similar condensation temperatures (41 and 35 K, respectively) and we
1244: assume ${\rm O/C} > 1$, so CO cannot incorporate all available oxygen,
1245: ${\rm CO/CH_4}$ is not critical to our final result.
1246:
1247: We find that the most abundant nitrogen-carrying molecule is ammonia,
1248: with ${\rm N_2 / NH_3} = 2.5 \times 10^{-3}$ (both species reside in ice
1249: mantles). We calculate ${\rm NH_3 / H_2O} = 0.14$, which matches the
1250: input N/O ratio.
1251: %This result is consistent
1252: %with observed abundances in the multiple protostellar system IRAS 16293
1253: %(Charnley and Rodgers 2008).
1254: Lodders (2003) also predicted that hydrated ammonia was the major
1255: nitrogen reservoir in the solar nebula (see \S \ref{ammonia} for a
1256: discussion of hydrated ammonia in Jupiter and Saturn's feeding zones).
1257: However, \cite{womack92} infer ${\rm N_2 / NH_3} = 4$ in the
1258: comet-forming region of the solar nebula from models of elemental
1259: nitrogen depletion in Comet Halley.
1260:
1261: Overall, our model molecular cloud favors the formation of hydrogenated
1262: C, N, and O compounds over oxides or diatomic molecules (except for
1263: H$_2$). We next discuss the subsequent chemical evolution of the solar
1264: nebula.
1265:
1266: \subsection{Chemical Model Results: Ice Lines}
1267: \label{icelinesec}
1268:
1269: In this section, we report the locations of important condensation
1270: fronts and the relative abundances of the most common ices. We use the
1271: gas and ice mixture resulting from the molecular cloud model as the
1272: starting point for a radial series of midplane chemical models that span
1273: 2~Myr of disk evolution. These models track the ice inventory of
1274: the disk midplane as a function of heliocentric distance and time.
1275:
1276: The ices in the disk can be subdivided into two categories: (1) those
1277: that form in the molecular cloud and freeze out from the disk gas, and
1278: (2) those that form or are destroyed in the warm disk midplane. All
1279: Category 1 molecules are highly stable both as gases and solids: in the
1280: radiation-shielded disk midplane, their reaction probabilities are very
1281: low. The stability and ease of formation of these molecules makes them
1282: highly abundant, $> 99$\% of the solar nebula ice mass. In Category 1
1283: are water, methane, CO, ammonia, and HCN. In Category 2 are all
1284: sulfur-carrying molecules, simple hydrocarbons, N$_2$, and NO.
1285:
1286: \subsubsection{Category 1: Stable and Abundant}
1287:
1288: Figure \ref{freezeabun} shows time snapshots of the abundance of
1289: Category 1 ice abundances (solid phase only) as a function of radius.
1290: Note that the ammonia and water condensation fronts coincide; we will
1291: discuss the implications of this further in \S \ref{ammonia}. The most
1292: abundant ice is H$_2$O, followed by methane, ammonia, CO and HCN. These
1293: five ices account for 98\% of the ice mass in regions where $T \leq
1294: 35$~K, the sublimation temperature of CO.
1295:
1296: In Fig. \ref{freezeabun}, we see the H$_2$O condensation front sweeping
1297: from 6~AU at $t = 0$ to 1.5~AU after 2~Myr. Observations of the
1298: asteroid belt place the H$_2$O ice line at 2.7~AU (Abe et al. 2000;
1299: Rivkin et al. 2002), a position it reaches at $t = 5 \times 10^5$~yr in
1300: our model. The slow motion of the snow line through the inner solar
1301: system could perhaps account for the H$_2$O abundance gradient observed
1302: in the asteroid belt (Barucci et al. 1996): we find that icy grains are
1303: present for $\sim 10^5$ years longer at 3~AU than at 2~AU. The snow
1304: line crosses Jupiter's heliocentric distance, 5.2~AU, after only $5
1305: \times 10^4$ yr, leaving ample time for the formation of Jupiter's core
1306: from icy planetesimals.
1307:
1308: If we assume Saturn formed {\it in situ} at 9.5~AU, primordial methane
1309: accretion onto both the giant planet and its satellites is possible. We
1310: find that grain mantles in Saturn's feeding zone should have ${\rm CH_4
1311: / H_2O} \ge 0.01$ after $2.3 \times 10^5$~yr. CH$_4$ is 90\% solid at
1312: 9.5~AU after $5.4 \times 10^5$~yr. We note that the C/H ratio in
1313: Saturn's atmosphere has recently been revised upward by a factor of 2
1314: (Flasar et al. 2005).
1315:
1316: We find that if Uranus and Neptune formed outside 12~AU, their
1317: core-forming planetesimals contained both CH$_4$ and CO. The
1318: composition of both planets strongly suggests that they accreted
1319: carbonaceous ices: ${\rm (C/H) / (C/H)}_{\odot}$ is 41 for Neptune and
1320: $\leq 260$ for Uranus (Lodders and Fegley 1994). Furthermore, Grundy et
1321: al. (2002) find that methane ice is widely distributed across the
1322: surface of Triton.
1323:
1324: \subsubsection{Category 2: Chemically Active}
1325:
1326: In Fig. \ref{freezeabun} we see that the typical behavior of an ice
1327: line is to move inward as the nebula cools. This can happen only if the
1328: molecule is relatively inert. Sulfur species are missing from Fig.
1329: \ref{freezeabun} because H$_2$S, the dominant sulfuric ice at $t = 0$
1330: with 99.6\% of the sulfur atoms, does not survive in the gas phase.
1331: Likewise, while C$_3$H$_n$ chains are stable and freezeout-dominated in
1332: our model, acetylene ice partially hydrogenates to form ethylene and
1333: ethane. Gaseous NO and N$_2$ participate in ion exchange reactions in
1334: the gas, reducing the abundance available for ice formation. In this
1335: section, we consider the sulfuric, nitric and hydrocarbon ice systems.
1336:
1337: As soon as our midplane chemical model begins, the reaction
1338: \begin{equation}
1339: {\rm H_2S + H \rightarrow HS + H_2}
1340: \label{h2s}
1341: \end{equation}
1342: starts destroying gaseous H$_2$S where temperature is above $\sim 50$~K.
1343: The product HS then initiates a network of reactions that form OCS and
1344: H$_2$CS. The reactions linking the main sulfur-carrying molecules are
1345: shown in Fig. \ref{sulfurnetwork}. We do not find any depletion of
1346: H$_2$CS once formed on grains. Figure \ref{sulfuric} shows time
1347: snapshots of the abundances (solid phase only) of the main
1348: sulfur-carrying molecules H$_2$S, OCS and H$_2$CS. After 2~Myr, we see
1349: a banded structure where H$_2$S dominates at $R > 8$~AU, OCS is most
1350: abundant between 5 and 8~AU and H$_2$CS dominates at $R < 5$~AU.
1351:
1352: H$_2$S remains in the outer solar nebula because the gas at $R > 9$~AU
1353: never gets hotter than the H$_2$S sublimation temperature in our model.
1354: This scenario may be physically unrealistic: infalling molecular cloud
1355: material is likely heated in an accretion shock during disk formation.
1356: However, if we examine the total abundance of icy sulfur atoms, $N({\rm
1357: H_2S}) + N({\rm OCS}) + N({\rm H_2CS})$, we see that it is conserved
1358: after OCS formation finishes at $3 \times 10^5$ yr. From Fig.
1359: \ref{sulfuric}, we can imagine a sulfurous ice line sweeping through the
1360: solar nebula, moving from 8.5~AU at $t = 0$ to 2.5~AU after 2~Myr.
1361:
1362: Figure \ref{sulfuric} also shows the abundances of solid NO and N$_2$.
1363: NO is chemically active during the late stages of the solar nebula, $T
1364: \geq 10^6$~yr. Even when the solar nebula temperature drops below the
1365: nominal sublimation temperature of NO, some thermal desorption still
1366: takes place, albeit at a much-reduced rate (see Eq. \ref{therm}).
1367: For most species, re-freezing after desorption occurs faster than
1368: reactions in the gas and no ice molecules are lost. However, NO can be
1369: destroyed in the gas by reactions with ${\rm H^+}$. Once the half-plane
1370: surface density drops below the cosmic ray penetration depth at late
1371: times, the production rate of ${\rm H^+}$ increases by 33\% due to the
1372: reaction of ${\rm H_2}$ with cosmic ray-produced photons:
1373: \begin{equation}
1374: {\rm H_2} + \gamma_{CR} \rightarrow {\rm H^+ + H + e^-} .
1375: \label{makehplus}
1376: \end{equation}
1377: Thermally desorbed NO atoms are then efficiently removed from the gas
1378: before refreezing by the reaction sequence
1379: \begin{equation}
1380: \begin{array}{c}
1381: {\rm NO + H^+ \rightarrow NO^+ + H} \\
1382: {\rm then} \\
1383: {\rm NO^+} \stackrel{\rm freeze}{\longrightarrow} {\rm N + O} \\
1384: {\rm or} \\
1385: {\rm NO^+ + e^- \rightarrow N + O} . \\
1386: \end{array}
1387: \end{equation}
1388:
1389: Laboratory work by \"{O}berg et al. (2005) indicates that N$_2$ is
1390: extremely volatile, with $E_D / k = 790$~K. This puts its
1391: sublimation temperature near the assumed ambient temperature of our
1392: protostellar disk simulation, 20~K. The smooth shape of the N$_2$
1393: abundance curve at $t = 1$~yr reflects the midplane temperature's
1394: asymptotic approach to isothermality at 20~K. As the disk evolves,
1395: the N$_2$ ice line moves inward, but the overall N$_2$ abundance
1396: declines due to the gas-phase reaction
1397: \begin{equation}
1398: {\rm He^+ + N_2 \rightarrow N^+ + N + He} .
1399: \label{dinitrogen}
1400: \end{equation}
1401: The ${\rm N^+}$ ion then begins a sequence of reactions with H$_2$ that
1402: produce ${\rm NH^+}$, ${\rm N_2H^+}$, ${\rm N_3H^+}$ and finally the
1403: ammonium ion ${\rm N_4H^+}$. Ammonium recombines with an electron to
1404: form NH$_3$ + H, so the net effect is to convert gaseous N$_2$ into
1405: NH$_3$.
1406:
1407: Our model includes the small aliphatic hydrocarbons C$_2$H$_2$,
1408: C$_2$H$_4$, C$_2$H$_6$, C$_3$H$_2$ and C$_3$H$_4$. We do not include
1409: longer chains in our model---we merely use these five species to find
1410: the location of the ``oil line'' of icy hydrocarbons. Lodders (2004)
1411: suggested that tarry compounds, rather than water, provided the bulk of
1412: the solid mass for Jupiter's core. This work does not truly test that
1413: hypothesis, since Lodders' scenario requires a substantial amount of
1414: hydrocarbons that are more refractory than water (our light oils have
1415: sublimation temperatures of 55-70~K). PAHs would be suitable tar
1416: candidates, but we do not include them in our model for lack of accurate
1417: binding energy measurements. We plan on incorporating PAH freezeout
1418: into future work, but we note that Jupiter falls inside the water-ice
1419: line for all but the first 50,000~yr of disk evolution.
1420:
1421: %Broadly speaking, our results imply that
1422: %unsaturated hydrocarbons are preferred over saturated ones, but
1423: %concatenation of small hydrocarbons is likely.
1424:
1425: Figure \ref{hydrocarbon} shows the evolution of the hydrocarbon ice
1426: lines. The most abundant hydrocarbons in our simulation are C$_3$H$_n$
1427: chains, which are at the end of our reaction sequence. This suggests
1428: that concatenation of small carbon chains is efficient. The small
1429: hydrocarbons in our model have similar sublimation temperatures,
1430: creating an ``oil line'' that moves from 8~AU at $t = 0$ to 3~AU after
1431: 2~Myr.
1432:
1433: For the C$_2$H$_n$ hydrocarbons, acetylene dominates over ethylene and
1434: ethane in the trans-Saturnian region, $R > 10$~AU. The activation
1435: barrier $E/k = 1210$~K for the reaction ${\rm C_2H_2 + H \rightarrow
1436: C_2H_3}$ (Hasegawa et al. 1992), which has to break the strong C-C
1437: triple bond, means that at the cold temperatures required for acetylene
1438: to deposit on grain surfaces, hydrogenation proceeds slowly. Only when
1439: the disk is near the acetylene sublimation temperature ($\sim 55$~K)
1440: does a significant amount of ethane form. Throughout most of the disk,
1441: ethylene is more abundant than ethane. However, an ``ethylene gap''
1442: forms between 7 and 10~AU after $3 \times 10^4$~yr. We explain the
1443: ethylene gap as follows: between $10^4$ and $10^5$~yr, 7--10~AU region
1444: takes on mass due to an outward-moving density wave (recall the
1445: discussion in \S \ref{surfacedensity}). This surface density increase
1446: causes a temperature increase of $\sim 15$~K, increasing the rate of
1447: ethylene hydrogenation.
1448:
1449: %A secondary acetylene gap and ethane peak forms at late times ($t >
1450: %1$~Myr) near 4~AU. The temperature at the start of the secondary
1451: %acetylene gap formation is also $\sim 60$~K.
1452:
1453: %C$_2$H$_n$ hydrocarbons move toward
1454: %saturation everywhere in the disk that
1455:
1456: \subsubsection{The Snow-Ammonia Line}
1457: \label{ammonia}
1458:
1459: Since we have adapted lab results showing co-desorption of water and
1460: ammonia (Collings et al. 2004) for our model, we find that the ammonia
1461: and water condensation fronts coincide. This statement may seem
1462: controversial, and indeed, further lab experiments are needed to verify
1463: that hydrogen bonding between the species is strong enough to govern
1464: ammonia desorption. However, the composition of Jupiter, Saturn and
1465: their satellites provides strong evidence for primordial ammonia-rich
1466: planetesimals between 5 and 10 AU.
1467:
1468: Lopes et al. (2007) examined the rheological properties of a
1469: cryovolcanic flow on Titan with Synthetic Aperture Radar imaging and
1470: found a slurry composition most consistent with an
1471: ammonia-water-methanol mixture (recall that methanol is another ice
1472: found to co-desorb with water in lab experiments). The presence of
1473: ammonia in cryovolcanic outflows strongly suggests primordial ammonia
1474: accretion: late-stage volatile delivery by comets would provide a small
1475: surface reservoir of ammonia but not embed it in the geologically active
1476: subsurface. Based on the low temperatures ($T < 75$~K) necessary for
1477: trapping N$_2$ in water ice, \cite{owen00} argued that Titan's massive
1478: nitrogen atmosphere originally supplied as NH$_3$, {\it unmodified from
1479: its relative abundance in the outer solar nebula.}
1480:
1481: %Similarly, Tobie et al. (2005) examined tidal
1482: %dissipation on Titan and found that a primordial liquid water shell {\it
1483: %with a few percent of ammonia} is required to maintain Titan's high
1484: %orbital eccentricity.
1485:
1486: Other Saturnian moons also show evidence of primordial ammonia
1487: accretion: Prentice (2007) calculates that Iapetus is 27\% ammonia by
1488: mass, assuming homologous contraction of the Saturnian subnebula and
1489: placing Enceladus just at the stability point of liquid water. The
1490: source material for the rock/ice Iapetus cannot be solar nebula gas, nor
1491: could N$_2$ have provided the initial nitrogen budget: with 34\% rock
1492: and 34\% H$_2$O, there would have been no free hydrogen reservoir for
1493: ammonia formation. Freeman et al. (2007) suggest that an ammonia-water
1494: ice mantle is necessary to explain the subsurface ocean driving
1495: Enceladus' south pole geyser. Finally, Prentice (2006) finds that
1496: the lack of compressional features on the surface of Rhea, which are
1497: expected to result from phase II of H$_2$O crystalline ice, favors an
1498: ammonia-rich satellite composition (25\% NH$_3$ by mass), since ammonia
1499: inhibits phase II H$_2$O ice formation.
1500:
1501: In the Jovian system, conducting subsurface oceans are required to
1502: explain the magnetic fields of Europa, Ganymede, and Callisto. One way
1503: to lower water's melting point and maintain liquid oceans is to mix in
1504: ammonia (Hussmann et al. 2006). Spohn and Schubert (2003) use
1505: equilibrium heat transfer models to confirm the likelihood of subsurface
1506: ammonia-water oceans on all Galilean satellites except Io.
1507:
1508: Perhaps the best piece of evidence for ammonia-containing planetesimals
1509: between 5 and 10 AU is the nitrogen enrichment in Jupiter and Saturn's
1510: atmospheres. ${\rm (N/H) / (N/H)}_{\odot} = 3.3$ for Jupiter and 2--4
1511: for Saturn (Owen and Encrenaz 2003). Such a large enrichment suggests a
1512: non-gaseous origin for nitrogen in both planets. Trapping N$_2$ or
1513: NH$_3$ in clathrate hydrates is possible, but \cite{hersant04} find that
1514: the ammonia clathrate cannot form until 0.9~Myr after solar nebula
1515: formation: the ammonia clathrate is more volatile than a co-deposited
1516: ammonia-water ice mixture. Based on the upward revision of Saturn's C/H
1517: ratio, Hersant et al. (2008) suggest that ammonia was present as a
1518: hydrate, rather than a clathrate, in Saturn's feeding zone. {\it We
1519: recommend that the canonical ``snow line'' be reconceived as a water
1520: ice-ammonia line, and that NH$_3$ be added to the volatile inventory of
1521: Jupiter- and Saturn-forming planetesimals.}
1522:
1523: %We note, however, that self-shielding of $^16$CO and preferential
1524: %photodestruction of heavier isotopes altered the isotopic composition of
1525: %refractory material throughout the solar system (Lyons and Young 2005).
1526: %Even though the Category 1 ices can be considered inert for the purpose
1527: %of predicting bulk planetesimal composition, they do react in the upper
1528: %layers of the disk and reaction products subsequently diffuse to the
1529: %midplane.
1530:
1531: \section{Solid Surface Density}
1532: \label{planetesimal}
1533:
1534: Knowing the solid-phase abundance of each ice species and the location
1535: of its condensation front, we can now calculate the solid surface
1536: density available for planet formation everywhere in the disk. We first
1537: calculate the gas/solid mass ratio
1538: \begin{equation}
1539: G / S = {\sum_g \mu_g n_g \over \sum_s \mu_s n_s} ,
1540: \label{gsr}
1541: \end{equation}
1542: where $\mu$ is the mean molecular weight of a species, $n$ is its
1543: abundance, and the subscripts $g$ and $s$ denote gas and solid species,
1544: respectively. We include Na, Mg, Si, Fe, Ni, Al and Ca in the solid
1545: inventory in solar proportions and assume these atoms form refractory
1546: compounds. Finally, we calculate the solid surface density as
1547: \begin{equation}
1548: \Sigma_{\rm solid} = {\Sigma_{\rm total} \over 1 + G / S} .
1549: \label{sigmasolid}
1550: \end{equation}
1551:
1552: Figure \ref{solids} shows the evolution of the solid surface density
1553: distribution in the solar nebula. Note that as the disk evolves, the
1554: solids should grow and decouple from the gas, whereas these solid
1555: surface density calculations assume grains and gas are well mixed. We
1556: see the large bump caused by the snow-ammonia line beginning at 6~AU and
1557: moving inward to 1.5~AU after 2~Myr. In addition, methane adsorption
1558: creates a second local maximum in the solid surface density profile at
1559: 11~AU, which moves inward to 6.5~AU by the end of the simulation.
1560:
1561: Panels 1 and 2 in Fig. \ref{solids}, $t = 1$ and $3 \times 10^4$~yr,
1562: show the effect of the decretion flow on the solid surface density
1563: profile. The solid mass between 6 and 14~AU actually increases during
1564: the early stages of disk evolution, at the expense of the inner 4~AU.
1565: The next panel, $10^5$~yr, shows the further solid surface density
1566: increase between 12 and 18~AU. Disk annuli that increase in surface
1567: density also heat up, and the methane bump consequently moves outward
1568: for the first $10^5$~yr.
1569:
1570: \cite{pollack96} presented core accretion models of Jupiter, Saturn, and
1571: Uranus, beginning with solid surface densities of 10, 3, and
1572: 0.75~g~cm$^{-2}$. While these simulations showed the plausibility of
1573: giant planet formation by core accretion, the planet formation
1574: timescales calculated were too long: Jupiter, Saturn and Uranus took 8,
1575: 9.6, and 16~Myr to form, respectively. Since runaway solid core growth
1576: rate is directly proportional to surface density (Eq.
1577: \ref{planetgrowth}), we see that having a high surface density speeds up
1578: giant planet formation.
1579:
1580: If one assumes that nearly 100\% of the available solids were
1581: incorporated into planetesimals, our results give starting solid surface
1582: densities that are far more conducive to giant planet formation than
1583: those from \cite{pollack96}. (The $\Sigma_{\rm solid}$ of
1584: \cite{pollack96} refers specifically to 100-km planetesimals, not
1585: smaller bodies or grains.) Table \ref{sdtable} lists solid surface
1586: density as a function of radius and time in the solar nebula. Note that
1587: these results assume a substantial population of small grains is always
1588: available for ice deposition.
1589:
1590: We consider $1.5 \times 10^5$~yr the extreme upper limit of the
1591: planetesimal formation timescale, which Hubbard and Blackman [2006]
1592: predict to be $\le 5 \times 10^4$~yr. However, we continue the disk
1593: simulation out to 2~Myr in order to calculate nebular gas temperature,
1594: density and composition, which are necessary for models of giant planet
1595: formation that build on this work (see, for example, the new Saturn
1596: formation simulation by Dodson-Robinson et al. [2008]). At Saturn's
1597: heliocentric distance, 9.5~AU, the solid surface density never drops
1598: below 8.7~g~cm$^{-2}$ during the first $1.5 \times 10^5$~yr of disk
1599: evolution. This large solid surface density should provide more than a
1600: factor of 3 speedup in Saturn's solid growth phase over the Pollack et
1601: al. results.
1602:
1603: For Uranus, the relative increase in solids is even more pronounced: at
1604: 20~AU, the minimum solid surface density during the first $1.5 \times
1605: 10^5$~yr is 3.3~g~cm$^{-2}$, for a factor of 4 speedup in runaway core
1606: growth. Moving Uranus' starting position inward to 16~AU, in accordance
1607: with the Nice model of planet migration (Tsiganis et al. 2005), provides
1608: a starting $\Sigma_{\rm solid}$ of not less than 4.9~g~cm$^{-2}$, in
1609: addition to the speedup given by a higher angular-frequency orbit. The
1610: prospects for Neptune to form rapidly at $\sim 12$~AU, as predicted by
1611: the Nice model, are quite good: $\Sigma_{\rm solid} \ge
1612: 7.1$~g~cm$^{-2}$.
1613:
1614: \cite{hubickyj05} provided an update to the Pollack et al. simulations
1615: of Jupiter's growth, this time assuming rapid grain settling and
1616: sublimation would lower opacities of the gaseous protoplanetary envelope
1617: and speed up its contraction. (Even if we assume low-opacity envelopes
1618: for Saturn and Uranus, they still require high solid surface densities
1619: to form quickly, as their solid growth stages both take $\ge 1.5$~Myr in
1620: the Pollack et al. models.) Hubickyj et al. found that, given a
1621: planetesimal surface density of 10~g~cm$^{-2}$, Jupiter could reach its
1622: present mass in 2.3~Myr. Our work shows that this starting condition
1623: for Jupiter is not only viable but likely: we predict solid surface
1624: densities not less than 8.2~g~cm$^{-2}$ with refractory material only
1625: and as high as 14.5~g~cm$^{-2}$ after the ice line moves through at $5
1626: \times 10^4$~yr.
1627:
1628: \section{Conclusions}
1629: \label{conclusions}
1630:
1631: By combining a viscously evolving protostellar disk model with a kinetic
1632: model of ice formation, we have calculated the time-evolving solid
1633: surface density available for giant planet formation. We find three
1634: results that are highly favorable for gas- and ice-giant formation:
1635:
1636: \begin{enumerate}
1637:
1638: \item The total (gas+solid) surface density distribution evolves toward
1639: uniformity, with an outward-moving wave depositing mass in the giant
1640: planet-forming region between 5 and 20~AU,
1641:
1642: \item The ammonia and water ice lines coincide, providing a mass
1643: enhancement of 7\% at the snow line over pure water condensation, and
1644:
1645: \item Methane and CO condensation beyond 12~AU add mass to Uranus and
1646: potentially Neptune's feeding zones.
1647:
1648: \end{enumerate}
1649:
1650: The next step is to use a core-accretion model to calculate how fast
1651: Saturn, Uranus and Neptune can form with our enhanced surface densities.
1652: The same methods used by \cite{hubickyj05} should work for Saturn.
1653: However, an updated method of tracking planetesimal scattering may be
1654: required for feeding zones with $R > 15$~AU, where orbital velocities of
1655: planetesimals decrease to $\sim 10$--50 times their escape velocities
1656: and self-stirring begins to affect the dynamics of the planetesimal
1657: disk. By calculating solid surface density as a function of radius and
1658: time self-consistently with viscous evolution of the protostellar disk,
1659: instead of assuming a single gas/solid ratio covers the entire giant
1660: planet-forming region of the solar nebula, the research presented here
1661: allows core accretion simulations to be fully deterministic.
1662:
1663: One area that requires more work is the viscous evolution of an
1664: MRI-turbulent disk with a dead zone, where $\alpha$ varies with radius.
1665: Annuli with an inactive midplane would have a lower column-averaged
1666: value of $\alpha$, which could slow or stop the movement of mass from
1667: the inner solar nebula to the giant planet-forming region. However,
1668: Turner and Sano (2008) find that a toroidal magnetic field component
1669: produced in the midplane by shear drives a laminar accretion flow even
1670: through the nonturbulent dead zone.
1671:
1672: Since our focus is on increasing the solid mass available for giant
1673: planet cores, we have neglected clathrate hydrates in our chemical
1674: model, preferring to tap into the large mass reservoir provided by ices.
1675: However, noble gas enrichment in the giant planet atmospheres suggests
1676: that clathration is very important, if not for building planet cores,
1677: then at least for determining atmospheric composition during the late
1678: stages of accretion (Hersant et al. 2004). The presence of PAHs in
1679: carbonaceous meteorites, cometary dust and Iapetus implies that these,
1680: too, might be an important solid mass reservoir. It is not yet known
1681: whether PAHs survive in the ISM or form primarily in the warm disk.
1682: Woods and Willacy (2007) present a mechanism for benzene formation in
1683: the inner 3~AU of the solar nebula. We are investigating the
1684: possibility of including PAH chemistry in future work.
1685:
1686: Finally, we note that the ALMA Design Reference Science Plan contains a
1687: proposal for detecting the snow line in nearby protoplanetary disks. We
1688: submit that the CO ice line would be an equally valuable detection,
1689: since this work strongly suggests that CO ice should be part of the
1690: solid inventory of the Uranus- and Neptune-building planetesimals. In a
1691: 1 Myr-old solar nebula analog in Taurus (140 pc), the CO ice line would
1692: lie 6 resolution units from the star when observing the CO ($J = 3
1693: \rightarrow 2$) transition.
1694:
1695: We thank Julie Moses, Mark Marley and Doug Lin for helpful discussions
1696: about the project design. S.D.R. was supported by grants from the NSF
1697: Graduate Research Fellowship program and the Achievement Rewards for
1698: College Scientists Foundation. K.W. and N.T. were supported by the JPL
1699: Research and Technology Development Program. P.B. received funding from
1700: NSF Grant AST-0507424 and NASA Origins Grant NNX08AH82G.
1701: % G.L. was
1702: %supported by NSF Career Grant AST-0449986 and NASA Planetary Geology and
1703: %Geophysics Program Grant NNG04GK19G.
1704:
1705: \begin{thebibliography}{}
1706:
1707: \bibitem[Abe et al.(2000)]{abe00} Abe, Y., Ohtani, E., Okuchi, T.,
1708: Righter, K., Draker, M.\ 2000.\ Water in the Early Earth.\ Origin of the
1709: earth and moon, edited by R.M.~Canup and K.~Righter and 69 collaborating
1710: authors. Tucson: University of Arizona Press, p. 413-433
1711:
1712: \bibitem[A'Hearn(2008)]{ahearn08} A'Hearn, M.~F.\ 2008.\ Deep Impact and
1713: the Origin and Evolution of Cometary Nuclei.\ Space Science Reviews 56.
1714:
1715: \bibitem[Aikawa et al.(1996)]{aikawa96} Aikawa, Y., Miyama, S.~M.,
1716: Nakano, T., Umebayashi, T.\ 1996.\ Evolution of Molecular Abundance in
1717: Gaseous Disks around Young Stars: Depletion of CO Molecules.\
1718: Astrophysical Journal 467, 684.
1719:
1720: \bibitem[Aikawa et al.(2008)]{aikawa08} Aikawa, Y., Wakelam, V., Garrod,
1721: R.~T., Herbst, E.\ 2008.\ Molecular Evolution and Star Formation: From
1722: Prestellar Cores to Protostellar Cores.\ Astrophysical Journal 674, 984.
1723:
1724: \bibitem[Alexander et al.(2006)]{alexander06} Alexander, R.~D., Clarke,
1725: C.~J., Pringle, J.~E.\ 2006.\ Photoevaporation of protoplanetary
1726: discs - II. Evolutionary models and observable properties.\ Monthly
1727: Notices of the Royal Astronomical Society 369, 229-239.
1728:
1729: \bibitem[Allamandola and Sandford(1994)]{allamandola94} Allamandola,
1730: L.~J., Sandford, S.~A.\ 1994.\ The First Symposium on the Infrared
1731: Cirrus and Diffuse Interstellar Clouds, 58, 302
1732:
1733: \bibitem[Allen and Robinson(1997)]{allen77} Allen, M., Robinson, G.~W.
1734: 1977. The molecular composition of dense interstellar clouds.\
1735: Astrophysical Journal 212, 396-415.
1736:
1737: \bibitem[Anders and Grevesse(1989)]{anders89} Anders, E., Grevesse, N.
1738: 1989.\ Abundances of the elements - Meteoritic and solar.\ Geochimica et
1739: Cosmochimica Acta 53, 197-214
1740:
1741: \bibitem[Balbus and Hawley(1991)]{balbus91} Balbus, S.~A., Hawley,
1742: J.~F. 1991.\ A powerful local shear instability in weakly magnetized
1743: disks. I - Linear analysis. II - Nonlinear evolution.\ Astrophysical
1744: Journal 376, 214-233.
1745:
1746: \bibitem[Barucci et al.(1996)]{barucci96} Barucci, M.~A., Fulchignoni,
1747: M., Lazzarin, M.\ 1996.\ Water ice in primitive asteroids? Planetary and
1748: Space Science 44, 1047-1049.
1749:
1750: \bibitem[Barzel and Biham(2007)]{barzel07} Barzel, B., Biham, O. 2007.\
1751: Efficient stochastic simulations of complex reaction networks on
1752: surfaces.\ Journal of Chemical Physics 127, 4703.
1753:
1754: \bibitem[Bell et al.(1997)]{bell97} Bell, K. R., Cassen, P. M., Klahr,
1755: H. H., Henning, Th. 1997.\ The Structure and Appearance of Protostellar
1756: Accretion Disks: Limits on Disk Flaring.\ Astrophysical Journal 486,
1757: 372.
1758:
1759: %\bibitem[Blum et al.(2000)]{blum00} Blum, J., et al. 2006, PhRvL, 85,
1760: %2426
1761:
1762: \bibitem[Boss(2005)]{boss05} Boss, A.~P. 2005.\ Evolution of the Solar
1763: Nebula. VII. Formation and Survival of Protoplanets Formed by Disk
1764: Instability.\ Astrophysical Journal 629, 535-548.
1765:
1766: %\bibitem[Botta et al.(2005)]{botta05} Botta, O., Bada, J. L., and
1767: %Ehrenfreund, P. 2002, Asteroids, Comets, and Meteors: ACM 2002, 500, 925
1768:
1769: \bibitem[Caselli et al.(1998)]{caselli98} Caselli, P., Hasegawa, T.~I.,
1770: Herbst, E. 1998.\ A Proposed Modification of the Rate Equations for
1771: Reactions on Grain Surfaces.\ Astrophysical Journal 495, 309.
1772:
1773: \bibitem[Cazaux and Tielens(2002)]{cazaux02} Cazaux, S., Tielens,
1774: A.~G.~G.~M. 2002. Molecular Hydrogen Formation in the Interstellar
1775: Medium. Astrophysical Journal 575, L29-L32.
1776:
1777: \bibitem[Chiang and Goldreich(1997)]{chiang97} Chiang, E.~I., Goldreich,
1778: P.\ 1997.\ Spectral Energy Distributions of T Tauri Stars with Passive
1779: Circumstellar Disks.\ Astrophysical Journal 490, 368.
1780:
1781: \bibitem[Ciesla(2007)]{ciesla07} Ciesla, F.~J.\ 2007.\ Outward Transport
1782: of High-Temperature Materials Arount the Midplane of the Solar Nebula.\
1783: Science 318, 613.
1784:
1785: \bibitem[Ciesla and Cuzzi(2006)]{ciesla06} Ciesla, F.~J., Cuzzi, J.~N.\
1786: 2006.\ The evolution of the water distribution in a viscous
1787: protoplanetary disk.\ Icarus 181, 178-204.
1788:
1789: \bibitem[Charnley and Rodgers(2008)]{charnley08} Charnley, S.~B.,
1790: Rodgers, S.~D.\ 2008.\ Interstellar Reservoirs of Cometary Matter.\
1791: Space Science Reviews 40.
1792:
1793: \bibitem[Collings et al.(2004)]{collings04} Collings, M.~P., Anderson,
1794: M.~A., Chen, R., Dever, J.~W., Viti, S., Williams, D.~A., McCoustra,
1795: M.~R.~S.\ 2004.\ A laboratory survey of the thermal desorption of
1796: astrophysically relevant molecules.\ Monthly Notices of the Royal
1797: Astronomical Society 354, 1133-1140.
1798:
1799: \bibitem[D'Alessio et al.(2006)]{dalessio06} D'Alessio, P., Calvet, N.,
1800: Hartmann, L., Franco-Hern\'{a}ndez, R., Serv\'{i}n, H.\ 2006.\ Effects
1801: of Dust Growth and Settling in T Tauri Disks. Astrophysical Journal
1802: 638, 314-335.
1803:
1804: \bibitem[Desch(2007)]{desch07} Desch, S.~J.\ 2007.\ Mass Distribution
1805: and Planet Formation in the Solar Nebula. Astrophysical Journal 671,
1806: 878-893.
1807:
1808: \bibitem[Dodson-Robinson et al.(2008)]{dodsonrobinson08}
1809: Dodson-Robinson, S.~E., Bodenheimer, P., Laughlin, G., Willacy, K.,
1810: Turner, N.~J., Beichman, C.~A.\ 2008.\ Saturn Forms by Core Accretion in
1811: 3.4~Myr.\ ArXiv e-prints arXiv:0810.0228 (accepted for publication in
1812: Astrophysical Journal Letters).
1813:
1814: \bibitem[Dullemond et al.(2006)]{dullemond06} Dullemond, C.~P., Apai,
1815: D., Walch, S.\ 2006.\ Crystalline Silicates as a Probe of Disk Formation
1816: History.\ Astrophysical Journal 640, L67-L70.
1817:
1818: \bibitem[Ferguson et al.(2005)]{ferguson05} Ferguson, J.~W., Alexander,
1819: D.~R., Allard, F., Barman, T., Bodnarik, J.~G., Hauschildt, P.~T.,
1820: Heffner-Wong, A., Tamanai, A.\ 2005.\ Low-Temperature Opacities.
1821: Astrophysical Journal 623, 585-596.
1822:
1823: %\bibitem[Fischer and Valenti(2005)]{fischer05} Fischer, D.~A., and
1824: %Valenti, J. 2005, \apj, 622, 1102
1825:
1826: \bibitem[Flasar et al.(2005)]{flasar05} Flasar, F.~M., and 45
1827: colleagues. 2005.\ Temperatures, Winds, and Composition in the Saturnian
1828: System.\ Science 307, 1247-1251.
1829:
1830: \bibitem[Fraser et al.(2001)]{fraser01} Fraser, H.~J., Collings, M.~P.,
1831: McCoustra, M.~R.~S., Williams, D.~A.\ 2001.\ Thermal desorption of water
1832: ice in the interstellar medium. Monthly Notices of the Royal
1833: Astronomical Society 327, 1165-1172
1834:
1835: \bibitem[Freeman et al.(2007)]{freeman07} Freeman, J., Stegman, D.,
1836: May, D.\ 2007.\ The Role of Ammonia in the Evolution of Enceladus.\ AGU
1837: Fall Meeting Abstracts 7.
1838:
1839: \bibitem[Garrod and Herbst(2006)]{garrod06} Garrod, R. T., Herbst, E.\
1840: 2006.\ Formation of methyl formate and other organic species in the
1841: warm-up phase of hot molecular cores. Astronomy and Astrophysics 457,
1842: 927-936.
1843:
1844: \bibitem[Grundy et al.(2002)]{grundy02} Grundy, W.~M., Buie, M.~W.,
1845: Spencer, J.~R.\ 2003.\ Spectroscopy of Pluto and Triton at 3-4 Microns:
1846: Possible Evidence for Wide Distribution of Nonvolatile Solids.
1847: Astronomical Journal 124, 2273-2278.
1848:
1849: \bibitem[Haisch et al.(2001)]{haisch01} Haisch, K.~E., Jr., Lada, E.~A.,
1850: Lada, C.~J.\ 2001.\ Disk Frequencies and Lifetimes in Young Clusters.\
1851: Astrophysical Journal 553, L153-L156.
1852:
1853: \bibitem[Hallenbeck et al.(2000)]{hallenbeck00} Hallenbeck, S.~L., Nuth,
1854: J.~A. III, Nelson, R.~N.\ 2000.\ Evolving Optical Properties of
1855: Annealing Silicate Grains: From Amorphous Condensate to Crystalline
1856: Mineral.\ Astrophysical Journal 535, 247-255.
1857:
1858: \bibitem[Harker and Desch(2002)]{harker02} Harker, D.~E., Desch, S.~J.\
1859: 2002.\ Annealing of Silicate Dust by Nebular Shocks at 10~AU.
1860: Astrophysical Journal 565, L109-L112.
1861:
1862: \bibitem[Hasegawa and Herbst(1993)]{hasegawa93} Hasegawa, T.~I.,
1863: Herbst, E.\ 1993.\ New gas-grain chemical models of quiescent dense
1864: interstellar clouds - The effects of H2 tunnelling reactions and cosmic
1865: ray induced desorption. Monthly Notices of the Royal Astronomical
1866: Society 261, 83-102
1867:
1868: \bibitem[Hasegawa et al.(1992)]{hasegawa92} Hasegawa, T.~I., Herbst, E.,
1869: Leung, C.~M.\ 1992.\ Models of gas-grain chemistry in dense interstellar
1870: clouds with complex organic molecules.\ Astrophysical Journal Supplement
1871: Series 82, 167-195.
1872:
1873: \bibitem[Hawley and Stone(1998)]{hawley98} Hawley, J.~F., Stone, J.~M.\
1874: 1998.\ Nonlinear Evolution of the Magnetorotational Instability in
1875: Ion-Neutral Disks.\ Astrophysical Journal 501, 758.
1876:
1877: \bibitem[Hawley et al.(1999)]{hawley99} Hawley, J.~F., Balbus, S.~A.,
1878: Winters, W.~F.\ 1999.\ Local Hydrodynamic Stability of Accretion Disks.\
1879: Astrophysical Journal 518, 394-404.
1880:
1881: \bibitem[Hayashi (1981)]{hayashi81} Hayashi, C.\ 1981.\ Structure of the
1882: Solar Nebula, Growth and Decay of Magnetic Fields and Effects of
1883: Magnetic and Turbulent Viscosities on the Nebula.\ Progress of
1884: Theoretical Physics Supplement 70, 35-53.
1885:
1886: \bibitem[Helling et al.(2000)]{helling00} Helling, Ch., Winters, J.~M.,
1887: Sedlmayr, E.\ 2000.\ Circumstellar dust shells around long-period
1888: variables. VII. The role of molecular opacities.\ Astronomy and
1889: Astrophysics 358, 651-664.
1890:
1891: \bibitem[Herbst et al.(2005)]{herbst05} Herbst, E., Chang, Q., Cuppen,
1892: H.~M.\ 2005.\ Chemistry on interstellar grains.\ Journal of Physics
1893: Conference Series, 6, 18-35.
1894:
1895: \bibitem[Hersant et al.(2001)]{hersant01} Hersant, F., Gautier, D.,
1896: Hur\'{e}, J.-M.\ 2001.\ A Two-dimensional Model for the Primordial
1897: Nebula Constrained by D/H Measurements in the Solar System: Implications
1898: for the Formation of Giant Planets.\ Astrophysical Journal 554, 391-407.
1899:
1900: \bibitem[Hersant et al.(2004)]{hersant04} Hersant, F., Gautier, D.,
1901: Lunine, J.~I.\ 2004.\ Enrichment in volatiles in the giant planets of
1902: the Solar System.\ Planetary and Space Science 52, 623-641.
1903:
1904: %\bibitem[Hersant et al.(2008)]{hersant08} Hersant, F., Gautier, D.,
1905: %Tobie, G., Lunine, J.\ 2008.\ Interpretation of the carbon abundance in
1906: %Saturn measured by Cassini.\ Planetary and Space Science 56, 1103-1111.
1907:
1908: \bibitem[Hornekaer et al.(2003)]{hornekaer03} Hornekaer, L., Baurichter,
1909: A., Petrunin, V.~V., Field, D., Luntz, A.~C.\ 2003.\ Importance of
1910: Surface Morphology in Interstellar H$_2$ Formation.\ Science 302,
1911: 1943-1946.
1912:
1913: \bibitem[Hubbard and Blackman(2006)]{hubbard06} Hubbard, A., Blackman,
1914: E.~G.\ 2006.\ Planetesimal growth in turbulent discs before the onset of
1915: gravitational instability.\ New Astronomy 12, 246-263.
1916:
1917: \bibitem[Hubickyj et al.(2005)]{hubickyj05} Hubickyj, O., Bodenheimer,
1918: P., Lissauer, J.~J.\ 2005.\ Accretion of the gaseous envelope of Jupiter
1919: around a 5-10 Earth-mass core.\ Icarus 179, 415-431.
1920:
1921: \bibitem[Hussmann et al.(2006)]{hussmann06} Hussmann, H., Sohl, F.,
1922: Spohn, T.\ 2006.\ Subsurface oceans and deep interiors of medium-sized
1923: outer planet satellites and large trans-neptunian objects. Icarus 185,
1924: 258-273.
1925:
1926: \bibitem[Ilgner and Nelson(2006)]{ilgner06} Ilgner, M., Nelson, R.~P.\
1927: 2006.\ On the ionisation fraction in protoplanetary disks. II. The
1928: effect of turbulent mixing on gas-phase chemistry.\ Astronomy and
1929: Astrophysics 445, 223-232.
1930:
1931: \bibitem[Jessberger et al.(1988)]{jessberger88} Jessberger, E. K.,
1932: Christoforidis, A., Kissel, J.\ 1988.\ Aspects of the major element
1933: composition of Halley's dust.\ Nature 332, 691-695.
1934:
1935: \bibitem[Keller et al.(2006)]{keller06} Keller, L.~P., and 32
1936: colleagues. 2006.\ Infrared Spectroscopy of Comet 81P/Wild 2 Samples
1937: Returned by Stardust.\ Science 314, 1728.
1938:
1939: \bibitem[Kippenhahn and Weigert(1994)]{kippenhahn94} Kippenhahn, R.,
1940: Weigert, A.\ 1994, Stellar Structure and Evolution (Berlin:
1941: Springer-Verlag)
1942:
1943: \bibitem[Kretke and Lin(2007)]{kretke07} Kretke, K.~A., Lin, D.~N.~C.\
1944: 2007.\ Grain Retention and Formation of Planetesimals near the Snow Line
1945: in MRI-driven Turbulent Protoplanetary Disks.\ Astrophysical Journal
1946: 664, L55-L58.
1947:
1948: \bibitem[Lodders(2003)]{lodders03} Lodders, K.\ 2003.\ Solar System
1949: Abundances and Condensation Temperatures of the Elements.\ Astrophysical
1950: Journal 591, 1220-1247.
1951:
1952: \bibitem[Lodders(2004)]{lodders04} Lodders, K.\ 2004.\ Jupiter Formed
1953: with More Tar than Ice.\ Astrophysical Journal 611, 587-597.
1954:
1955: \bibitem[Lodders and Fegley(1994)]{lodders94} Lodders, K., Fegley, B.\
1956: 1994.\ The origin of carbon monoxide in Neptune's atmosphere.\ Icarus,
1957: 112, 368-375.
1958:
1959: \bibitem[Lopes et al.(2007)]{lopes07} Lopes, R.~M.~C., and 43
1960: colleagues 2007.\ Cryovolcanic features on Titan's surface as revealed
1961: by the Cassini Titan Radar Mapper.\ Icarus 186, 395-412.
1962:
1963: \bibitem[Lynden-Bell and Pringle(1974)]{lyndenbell74} Lynden-Bell, D.,
1964: Pringle, J.~E.\ 1974.\ The evolution of viscous discs and the origin of
1965: the nebular variables.\ Monthly Notices of the Royal Astronomical
1966: Society 168, 603-637.
1967:
1968: %\bibitem[Lyons and Young(2005)]{lyons05} Lyons, J. R., and Young, E. D.
1969: %2005, \nat, 435, 317
1970:
1971: \bibitem[Lyra et al.(2008)]{lyra08} Lyra, W., Johansen, A., Klahr, H.,
1972: and Piskunov, N.\ 2008.\ Global magnetohydrodynamical models of
1973: turbulence in protoplanetary disks. I. A cylindrical potential on a
1974: Cartesian grid and transport of solids.\ Astronomy and Astrophysics 479,
1975: 883-901.
1976:
1977: \bibitem[Mihalas (1978)]{mihalas78} Mihalas, D. 1978, Stellar
1978: Atmospheres (San Francisco: Freeman)
1979:
1980: \bibitem[Millar et al.(1997)]{millar97} Millar, T.~J., Farquhar, P.~R.
1981: A., Willacy, K.\ 1997.\ The UMIST Database for Astrochemistry 1995.\
1982: Astronomy and Astrophysics Supplement Series 121, 139-185.
1983:
1984: \bibitem[Milsom et al.(1994)]{milsom94} Milsom, J. A., Chen, X.,
1985: Taam, R.\ 1994.\ The vertical structure and stability of accretion disks
1986: surrounding black holes and neutron stars.\ Astrophysical Journal 421,
1987: 668-676.
1988:
1989: \bibitem[\"{O}berg et al.(2005)]{oberg05} \"{O}berg, K.~I., van
1990: Broekhuizen, F., Fraser, H.~J., Bisschop, S.~E., van Dishoeck, E.~F.,
1991: Schlemmer, S.\ 2005.\ Competition between CO and N$_2$ Desorption from
1992: Interstellar Ices.\ Astrophysical Journal 631, L33-L36.
1993:
1994: \bibitem[Owen(2000)]{owen00} Owen, T.\ 2000.\ On the origin of Titan's
1995: atmosphere. Planetary and Space Science 48, 747-752.
1996:
1997: \bibitem[Owen and Encrenaz(2003)]{owen03} Owen, T., Encrenaz, T.\ 2003.\
1998: Element Abundances and Isotope Ratios in the Giant Planets and Titan.\
1999: Space Science Reviews 106, 121-138.
2000:
2001: \bibitem[Podolak(2003)]{podolak03} Podolak, M.\ 2003.\ The contribution
2002: of small grains to the opacity of protoplanetary atmospheres.\ Icarus
2003: 165, 428-437.
2004:
2005: \bibitem[Pollack et al.(1996)]{pollack96} Pollack, J.~B., Hubickyj, O.,
2006: Bodenheimer, P., Lissauer, J.~J., Podolak, M., Greenzweig, Y.\ 1996.\
2007: Formation of the Giant Planets by Concurrent Accretion of Solids and
2008: Gas.\ Icarus 124, 62-85.
2009:
2010: \bibitem[Prentice(2007)]{prentice07} Prentice, A.~J.\ 2007.\ Iapetus: a
2011: Prediction for Bulk Chemical Composition, Internal Physical Structure
2012: and Origin.\ AGU Fall Meeting Abstracts 1432.
2013:
2014: \bibitem[Pringle(1981)]{pringle81} Pringle, J.~E.\ 1981.\ Accretion
2015: discs in astrophysics.\ Annual Review of Astronomy and Astrophysics 19,
2016: 137-162.
2017:
2018: \bibitem[Press et al.(1992)]{press92} Press, W. H., Teukolsky, S. A.,
2019: Vetterling, W. T., Flannery, B. P. 1992, Numerical Recipes in Fortran
2020: 77, Second Edition (Cambridge: Cambridge University Press)
2021:
2022: \bibitem[Reyes-Ruiz et al.(2003)]{reyesruiz03} Reyes-Ruiz, M.,
2023: P\'{e}rez-Tijerina, E., S\'{a}nchez-Salcedo, F.~J.\ 2003.\ The
2024: Magneto-Rotational Instability in Protoplanetary Disks.\ Revista
2025: Mexicana de Astronomia y Astrofisica Conference Series 18, 92-96.
2026:
2027: %\bibitem[Richter et al.(2006)]{richter06} Richter, F.~M., Mendybaev, R.~
2028: %A., Davis, A.~M.\ 2006.\ Conditions in the protoplanetary disk as seen
2029: %by the type B CAIs.\ Meteoritics and Planetary Science, 41, 83-93.
2030:
2031: \bibitem[Rivkin et al.(2000)]{rivkin00} Rivkin, A.~S., Howell, E.~S.,
2032: Vilas, F., Lebofsky, L.~A.\ 2002.\ Hydrated Minerals on Asteroids: The
2033: Astronomical Record.\ Asteroids III 235-253.
2034:
2035: %\bibitem[Robinson et al.(2006)]{robinson06} Robinson, S. E., Laughlin,
2036: %G., Bodenheimer, P., and Fischer, D. 2006, \apj, 643, 484
2037:
2038: \bibitem[Ruffle and Herbst(2000)]{ruffle00} Ruffle, D.~P., Herbst, E.\
2039: 2000.\ New models of interstellar gas-grain chemistry - I. Surface
2040: diffusion rates.\ Monthly Notices of the Royal Astronomical Society 319,
2041: 837-850.
2042:
2043: \bibitem[Ruden and Lin(1986)]{ruden86} Ruden, S.~P., Lin, D.~N.~C.\
2044: 1986.\ The global evolution of the primordial solar nebula.\
2045: Astrophysical Journal 308, 883-901.
2046:
2047: \bibitem[Ruden and Pollack(1991)]{ruden91} Ruden, S.~P., Pollack,
2048: J.~B.\ 1991.\ The dynamical evolution of the protosolar nebula.\
2049: Astrophysical Journal 375, 740-760.
2050:
2051: \bibitem[Safronov(1969)]{safronov69} Safronov, V.~S. 1969, In Russian,
2052: English translation: NASA-TTF-677, 1972
2053:
2054: \bibitem[Sandford and Allamandola(1990)]{sandford90} {Sandford, S.~A.,
2055: Allamandola, L.~J.\ 1990.\ The volume- and surface-binding energies of
2056: ice systems containing CO, CO2, and H2O.\ Icarus 87, 188-192.
2057:
2058: %\bibitem[Scott and Krot(2005)]{scott05} Scott, E.~D.~R., Krot, A.~N.\
2059: %2005.\ Thermal Processing of Silicate Dust in the Solar Nebula: Clues
2060: %from Primitive Chondrite Matrices.\ Astrophysical Journal 623, 571-578.
2061:
2062: \bibitem[Semenov et al.(2003)]{semenov03} Semenov, D., Henning, Th.,
2063: Helling, Ch., Ilgner, M., Sedlmayr, E.\ 2003.\ Rosseland and Planck mean
2064: opacities for protoplanetary discs.\ Astronomy and Astrophysics 410,
2065: 611-621.
2066:
2067: \bibitem[Shakura and Syunyaev(1973)]{shakura73} Shakura, N.~I.,
2068: Syunyaev, R.~A.\ 1973.\ Black holes in binary systems. Observational
2069: appearance.\ Astronomy and Astrophysics 24, 337-355
2070:
2071: \bibitem[Spohn and Schubert(2003)]{spohn03} Spohn, T., Schubert, G.\
2072: 2003.\ Oceans in the icy Galilean satellites of Jupiter? Icarus 161,
2073: 456-467.
2074:
2075: \bibitem[Stamatellos and Whitworth(2008)]{stamatellos08} Stamatellos,
2076: D., Whitworth, A.~P.\ 2008.\ Can giant planets form by gravitational
2077: fragmentation of discs? Astronomy and Astrophysics 480, 879-887.
2078:
2079: \bibitem[Stepinski and Valageas(1997)]{stepinski97} Stepinski, T.~F.,
2080: Valageas, P.\ 1997.\ Global evolution of solid matter in turbulent
2081: protoplanetary disks. II. Development of icy planetesimals.\ Astronomy
2082: and Astrophysics 319, 1007-1019.
2083:
2084: \bibitem[Tielens and Allamandola(1987)]{tielens87} Tielens, A.~G.~G.~M.,
2085: Allamandola, L.~J.\ 1987.\ Composition, structure, and chemistry of
2086: interstellar dust.\ Interstellar Processes 134, 397-469.
2087:
2088: \bibitem[Tielens and Hagen(1982)]{tielens82} Tielens, A.~G.~G.~M.,
2089: Hagen, W.\ 1982.\ Model calculations of the molecular composition of
2090: interstellar grain mantles.\ Astronomy and Astrophysics 114, 245-260.
2091:
2092: %\bibitem[Tobie et al.(2005)]{tobie05} Tobie, G., Grasset, O., Lunine, J.
2093: %I., Mocquet, A., Sotin, C.\ 2005.\ Titan's internal structure inferred
2094: %from a coupled thermal-orbital model.\ Icarus 175, 496-502.
2095:
2096: \bibitem[Tsiganis et al.(2005)]{tsiganis05} Tsiganis, K., Gomes, R.,
2097: Morbidelli, A., Levison, H.~F.\ 2005.\ Origin of the orbital
2098: architecture of the giant planets of the Solar System. Nature 435,
2099: 459-461.
2100:
2101: \bibitem[Turner and Sano(2008)]{turner08} Turner, N.~J., Sano, T.\
2102: 2008.\ Dead Zone Accretion Flows in Protostellar Disks.\ Astrophysical
2103: Journal 679, L131-L134.
2104:
2105: %\bibitem[Turner et al.(2007)]{turner07} Turner, N.~J., Sano, T.,
2106: %Dziourkevitch, N.\ 2007.\ Turbulent Mixing and the Dead Zone in
2107: %Protostellar Disks.\ Astrophysical Journal 659, 729-737.
2108:
2109: \bibitem[Umebayashi and Nakano(1980)]{umebayashi80} Umebayashi, T.,
2110: Nakano, T.\ 1980.\ Recombination of Ions and Electrons on Grains and the
2111: Ionization Degree in Dense Interstellar Clouds.\ Publications of the
2112: Astronomical Society of Japan 32, 405.
2113:
2114: \bibitem[Umebayashi and Nakano(1981)]{umebayashi81} Umebayashi, T.,
2115: Nakano, T.\ 1981.\ Fluxes of Energetic Particles and the Ionization Rate
2116: in Very Dense Interstellar Clouds.\ Publications of the Astronomical
2117: Society of Japan 33, 617.
2118:
2119: \bibitem[van Boekel et al.(2004)]{vanboekel04} van Boekel, R., and 22
2120: colleagues. 2004.\ The building blocks of planets within the
2121: `terrestrial' region of protoplanetary disks.\ Nature 432, 479-482.
2122:
2123: \bibitem[Weidenschilling(1977)]{weidenschilling77} Weidenschilling,
2124: S.~J.\ 1977.\ The distribution of mass in the planetary system and solar
2125: nebula.\ Astrophysics and Space Science 51, 153-158.
2126:
2127: %\bibitem[Weissman(1996)]{weissman96} Weissman, P.~R.\ 1996.\ The Oort
2128: %Cloud.\ Completing the Inventory of the Solar System, 107, 265-288.
2129:
2130: \bibitem[Willacy and Langer(2000)]{willacy00} Willacy, K., Langer,
2131: W.~D.\ 2000.\ The Importance of Photoprocessing in Protoplanetary
2132: Disks.\ Astrophysical Journal 544, 903-920.
2133:
2134: \bibitem[Willacy et al.(2006)]{willacy06} Willacy, K., Langer, W.,
2135: Allen, M., Bryden, G.\ 2006.\ Turbulence-driven Diffusion in
2136: Protoplanetary Disks: Chemical Effects in the Outer Regions.\
2137: Astrophysical Journal 644, 1202-1213.
2138:
2139: \bibitem[Woods and Willacy(2007)]{woods07} Woods, P.~M., Willacy, K.
2140: 2007.\ Benzene Formation in the Inner Regions of Protostellar Disks.\
2141: Astrophysical Journal 655, L49-L52.
2142:
2143: \bibitem[Womack et al.(1992)]{womack92} Womack, M., Wyckoff, S.,
2144: Ziurys, L.~M.\ 1992.\ Estimates of N2 abundances in dense molecular
2145: clouds.\ Astrophysical Journal 401, 728-735.
2146:
2147: \end{thebibliography}
2148:
2149: %Tables Here
2150:
2151: \include{freedisk}
2152: \include{solarcomp}
2153: \include{freechem}
2154: \include{binding}
2155: \include{cloudoutput}
2156: \include{iceratios}
2157: \include{sdtable}
2158:
2159: \clearpage
2160:
2161: %Figures Here
2162:
2163: %\begin{figure}
2164: %\centering
2165: %\includegraphics{taucont.jpg}
2166: %\caption{Optical depth as a function of disk height and radius.}
2167: %\label{opdepth}
2168: %\end{figure}
2169:
2170: \begin{figure}
2171: \centering
2172: \includegraphics[scale=0.9]{smooth.jpg}
2173: \caption{Opacity as a function of temperature and density. The black
2174: line shows the unsmoothed tables of Ferguson et al. (2005) ($T >
2175: 1000$K)and Semenov et al. (2003) ($T < 700$K). In the range $700 < T <
2176: 1000$K, we take a weighted average in log-T space of the two tables.
2177: Finally, the green curve shows the smoothed opacity functions we use in
2178: the model.}
2179: \label{opplot}
2180: \end{figure}
2181:
2182: \begin{figure}
2183: \centering
2184: \includegraphics[scale=0.9]{sdplanet.jpg}
2185: \caption{Surface density as a function of radius and time in the giant
2186: planet-forming region. Each panel shows a snapshot of the disk surface
2187: density profile in black, with the profiles from previous epochs
2188: retained in gray. During the first $5 \times 10^5$~yr of evolution, a
2189: wave of local surface density enhancement propagates outward, depositing
2190: mass in the giant planet region of the solar nebula.}
2191: \label{sdplanet}
2192: \end{figure}
2193:
2194: \begin{figure}
2195: \centering
2196: \includegraphics[scale=0.9]{sdwholedisk.jpg}
2197: \caption{Surface density as a function of radius and time throughout the
2198: whole disk. Each panel shows a snapshot of the disk surface density
2199: profile in black, with the profiles from previous epochs retained in
2200: gray. The disk expands from $30$~AU to $> 80$~AU over the course of the
2201: simulation.}
2202: \label{sdwholedisk}
2203: \end{figure}
2204:
2205: \begin{figure}
2206: \centering
2207: \includegraphics[scale=0.9]{sdann.jpg}
2208: \caption{Surface density as a function of time at 5, 10, 15 and 20~AU.
2209: An outward-propagating wave adds mass to each annulus as it passes
2210: through. The largest relative surface density increase, $\Delta \Sigma
2211: / Sigma_0$, occurs at 10~AU, near Saturn's feeding zone.}
2212: \label{sdann}
2213: \end{figure}
2214:
2215: \begin{figure}
2216: \centering
2217: \includegraphics[scale=0.9]{mdot.jpg}
2218: \caption{Time snapshots of accretion rate as a function of radius.
2219: Negative values of $\dot{M}$ indicate ``decretion'' regions of the disk
2220: with net outward mass flow.}
2221: \label{mdot}
2222: \end{figure}
2223:
2224: \begin{figure}
2225: \centering
2226: \includegraphics[scale=0.9]{midplanetemp.jpg}
2227: \caption{Midplane temperature as a function of radius and time in the
2228: inner 30~AU of the disk. Outside of 30~AU, the midplane is isothermal
2229: at 20~K, the assumed ambient temperature. The dashed line at 800~K
2230: shows the minimum temperature for silicate crystallization (Gail 1998);
2231: the dotted line at 160~K shows the sublimation temperature of H$_2$O
2232: from the temperature-programmed desorption experiments of
2233: \cite{collings04}; and the dash-dotted line at 20~K shows the assumed
2234: ambient temperature.}
2235: \label{midplanetemp}
2236: \end{figure}
2237:
2238: \begin{figure}
2239: \centering
2240: \begin{tabular}{c}
2241: \includegraphics[scale=0.5]{carbon.jpg} \\
2242: \includegraphics[scale=0.5]{nitrogen.jpg} \\
2243: \includegraphics[scale=0.5]{oxygen.jpg} \\
2244: \end{tabular}
2245: \caption{Distribution of C, N and O atoms among dust and most common ice
2246: species. Proportions are rounded to the nearest 1\%.}
2247: \label{pies}
2248: \end{figure}
2249:
2250: \begin{figure}
2251: \centering
2252: \includegraphics[scale=0.9]{freezeabun.jpg}
2253: \caption{Time snapshots of ice abundance as a function of radius for
2254: species formed mainly by gas freezeout. Note that the condensation
2255: fronts are not step functions: the region of partial ice freezeout can
2256: cover up to 3 AU.}
2257: \label{freezeabun}
2258: \end{figure}
2259:
2260: \begin{figure}
2261: \centering
2262: \includegraphics[scale=0.6]{sulfurnetwork.jpg}
2263: \caption{Reactions linking the most abundant sulfuric ices, H$_2$S,
2264: H$_2$CS and OCS. The label ``G'' in front of a compound name denotes
2265: that the species is solid, adsorbed on the grains. The notation $H_2S
2266: \stackrel{H}{\rightarrow} HS$ denotes the reaction $H_2S + H \rightarrow
2267: HS + H_2$; all other reaction pathways can be read analogously to this
2268: example. Non-sulfuric reaction by-products, such as H$_2$ in the
2269: aforementioned reaction, are not shown. An unlabeled double arrow, such
2270: as the one connecting GOCS and OCS, denotes a phase change. Reaction
2271: pairs are not in equilibrium.}
2272: \label{sulfurnetwork}
2273: \end{figure}
2274:
2275: \begin{figure}
2276: \centering
2277: \includegraphics[scale=0.9]{sulfuric.jpg}
2278: \caption{Abundances of solid NO, N$_2$, and sulfur-carrying species as a
2279: function of radius and time. We find a banded sulfur ice distribution
2280: in which H$_2$S dominates outside $\sim 7$~AU, OCS is most abundant
2281: between 5 and 7~AU, and H$_2$CS is dominant in the inner nebula.}
2282: \label{sulfuric}
2283: \end{figure}
2284:
2285: \begin{figure}
2286: \centering
2287: \includegraphics[scale=0.9]{hydrocarbon.jpg}
2288: \caption{Hydrocarbon ice abundances as a function of radius and time.
2289: We group all C$_3$H$_n$ compounds together because our reaction sequence
2290: ends at C$_3$H$_4$, so we cannot differentiate between different levels
2291: of C$_3$ saturation.}
2292: \label{hydrocarbon}
2293: \end{figure}
2294:
2295: \begin{figure}
2296: \centering
2297: \includegraphics[scale=0.9]{solids.jpg}
2298: \caption{Time snapshots of solid surface density available for planet
2299: formation as a function of radius. In each panel, the current solid
2300: surface density is plotted in black and previous time snapshots are
2301: retained in gray.}
2302: \label{solids}
2303: \end{figure}
2304:
2305:
2306: \end{document}
2307: