0704.3269/hj.tex
1: \documentclass[12pt,preprint]{aastex}
2: \usepackage{epsfig,natbib,lscape,fullpage,indentfirst,color}
3: \shorttitle{Hot Jupiters}
4: \newcommand{\pdt}[1]{{{\partial #1}\over {\partial t}}}
5: \newcommand{\pdr}[1]{{{\partial #1}\over {\partial r}}}
6: \newcommand{\Mp}{M_p}
7: 
8: \begin{document}
9: \title{Atmospheric Dynamics of Short-period Extra Solar Gas Giant Planets I:}
10: \title{Dependence of Night-Side Temperature on Opacity} 
11: \author{Ian Dobbs-Dixon$^1$ \& D.N.C. Lin$^{1,2}$}
12: \affil{$^1$Department of Astronomy and Astrophysics,
13: University of California, Santa Cruz, CA 95064, USA}
14: \affil{$^2$Kavli Institute of Astronomy \& Astrophysics, Peking
15: University, Beijing, China}
16: 
17: \begin{abstract}
18: More than two dozen short-period Jupiter-mass gas giant planets have
19: been discovered around nearby solar-type stars in recent years,
20: several of which undergo transits, making them ideal for the detection
21: and characterization of their atmospheres. Here we adopt a
22: three-dimensional radiative hydrodynamical numerical scheme to
23: simulate atmospheric circulation on close-in gas giant planets. In
24: contrast to the conventional GCM and shallow water algorithms, this
25: method does not assume quasi hydrostatic equilibrium and it
26: approximates radiation transfer from optically thin to thick regions
27: with flux-limited diffusion.  In the first paper of this series, we
28: consider synchronously-spinning gas giants. We show that a full
29: three-dimensional treatment, coupled with rotationally modified flows
30: and an accurate treatment of radiation, yields a clear temperature
31: transition at the terminator. Based on a series of numerical
32: simulations with varying opacities, we show that the night-side
33: temperature is a strong indicator of the opacity of the planetary
34: atmosphere. Planetary atmospheres that maintain large, interstellar
35: opacities will exhibit large day-night temperature differences, while
36: planets with reduced atmospheric opacities due to extensive grain
37: growth and sedimentation will exhibit much more uniform temperatures
38: throughout their photosphere's. In addition to numerical results, we
39: present a four-zone analytic approximation to explain this dependence.
40: \end{abstract}
41: \maketitle
42: 
43: \section{Introduction}
44: Among the over 200 extra-solar planets discovered through a variety of
45: observational techniques including Doppler shifted spectral lines,
46: transit searches, and gravitational lensing, approximately $9\%$ have
47: orbital periods less then $4$ days. One of the more puzzling and
48: unexpected results to come from planet detections, these planets,
49: known as hot-Jupiter's or Pegasus planets, receive approximately
50: $10^4$ times more energy from stellar irradiation than from internal
51: heating. This energy input greatly modifies the thermal structure and
52: dynamics of the atmosphere. Given the short orbital period of these
53: planets, tidal forces also play a strong role, quickly circularizing
54: their orbits and synchronizing their spin and orbital frequencies
55: \citep{goldreich1966, dobbsdixon2004}. Consequently, one side of the
56: planet perpetually faces the host star, reaching equilibrium
57: temperatures of over $1000 \mathrm{ K}$, while night-side temperatures
58: are determined by the ability of the planet to redistribute energy. In
59: the absence of heat redistribution, the night-side of the planet would
60: remain at $\approx 100\mathrm{ K}$ after a few Gyr of cooling and
61: Kelvin-Helmholtz contraction. For comparison, Jupiter has a rotational
62: period of $11$ hours and its internal luminosity is approximately
63: equal to the energy it receives from the Sun. Its photospheric
64: temperature is quite uniform in both longitude and latitude.
65: 
66: The small semi-major axis of hot-Jupiter's provide several new methods
67: for exploring their structural parameters and atmospheres including
68: primary transits \citep{charbonneau2007}, secondary eclipses
69: \citep{deming2005, charbonneau2005, deming2006}, and spectroscopy
70: \citep{charbonneau2002, brown2002, deming2005_2, grillmair2007,
71: richardson2007, swain2007}.  Detection of a primary eclipse, as the
72: planet passes between the star and the Earth, yields measurements of
73: planetary radii and masses that can be used to infer structural
74: parameters of the planets. The secondary eclipse, caused by the
75: passage of the planet behind its host star, provides a direct
76: measurement of the day-side temperature of the objects by measuring
77: the decrement in infrared flux. As of the writing of this paper, $14$
78: planets have been detected via the primary transit method, while $3$
79: of those have also been detected via their secondary eclipse. The
80: day-side temperatures of the objects found from secondary eclipse
81: measurements are found to be $T = 1130\pm150 \mathrm{K}$ for HD209458b
82: \citep{deming2005}, $T = 1060\pm50 \mathrm{K}$ for TrES-1
83: \citep{charbonneau2005}, and $T = 1117\pm42 \mathrm{K}$ for HD189733b
84: \citep{deming2006}. Measurements of HD209458b and HD189733b were done
85: using the Spitzer Space Telescopes $24\mu m$ and $16\mu m$ band-passes
86: respectively. The estimate of temperature is then sensitive to both
87: the ratio of stellar and planetary radii and stellar temperature.
88: TrES-1 was observed in both the $4.5\mu m$ and $8\mu m$ bands, and
89: temperatures were derived by assuming the planet emits as a
90: blackbody. In addition, recent non-transiting observations of
91: $\nu-$Andromeda b with the Spitzer Space Telescope, indicate a
92: substantial orbital phase dependence in the flux at $24\mu m$
93: \citep{harrington2006}.
94: 
95: A number of groups have used one-dimensional, plane-parallel radiative
96: transfer models to calculate the emergent spectra of these
97: hot-Jupiter's \citep{fortney2005, burrows2005, seager2005, barman2005,
98: fortney2006_2, burrows2006_2, fortney2006}. These approaches
99: simultaneously solve the equations of radiative transfer, radiative
100: equilibrium, and hydrostatic balance to determine the pressure,
101: temperature, and spectral energy distribution as a function of radial
102: position in the atmosphere. The input parameters include the
103: temperatures at the bottom of the atmosphere and the incident stellar
104: energy spectrum at the top. The temperature at the bottom is
105: determined primarily by the internal luminosity, and ideally it should
106: be taken from one-dimensional evolutionary and structural
107: models. However, given the dominate influence of the incident energy
108: near the surface, the choice of the bottom does not noticeably
109: influence the emergent spectra.
110: 
111: In hot-Jupiter's, the intense irradiation of the day-side drives
112: strong thermal winds toward the night-side. The resulting temperature
113: in the upper atmosphere depends on the ability of these winds to
114: redistribute the stellar irradiation, and in general should be a
115: function of both longitude and latitude. In the absence of dynamical
116: models, this redistribution of incident energy at the upper layers is
117: usually set to be some fraction of the incident irradiation, to
118: represent the degree of energy re-distribution. The value of this
119: parameter is a major uncertainty in these radiative models, and
120: authors have computed a number of cases with varying degrees of
121: re-distribution. In addition, such parameterization neglects the role
122: of advection and radiative transfer within lower levels of the
123: atmosphere, which are also important in determining the final
124: pressure-temperature profiles.
125: 
126: In an attempt to address the dynamical redistribution of incident
127: stellar energy within the atmosphere, there have been a number of
128: dynamical models of the atmospheres of hot-Jupiter's. These models
129: utilize a variety of methods employing various simplifications,
130: including solving the primitive equations
131: \citep{showman2002,cooper2005}, the shallow water equations
132: \citep{cho2003,menou2003,langton2007}, the equivalent-barotropic
133: equations \citep{cho2006}, and two-dimensional hydrodynamical
134: equations \citep{burkert2005}. These models predict a wide range of
135: behavior. More details on these models are presented in
136: \S\ref{sec:comparison}, in comparison with the model presented in this
137: paper. One group \citep{fortney2006} has attempted to couple the
138: spectral and dynamical models by utilizing pressure and temperature
139: profiles derived directly from the simulations of
140: \citet{cooper2005}. Unlike previous models, these atmospheres are not
141: iterated to achieve radiative-convective equilibrium.
142: 
143: In this paper, we present the results obtained from a three-
144: dimensional, hydrodynamic simulation based on flux-limited radiative
145: transfer models of hot-Jupiter's for a variety of rotation rates and
146: opacities. In \S\ref{sec:method}, we present the basic equations,
147: numerical methods, and initial conditions. In \S\ref{sec:results} we
148: present our results and analysis for both rotating and non-rotating
149: flows. We also study the effects of changing opacity on the dynamics
150: and heat distribution. In \S\ref{sec:comparison} we include a detailed
151: comparison with previous dynamical models in an attempt to highlight
152: the consequences of making certain simplifying assumptions. We conclude
153: in \S\ref{sec:discussion} with a discussion.
154: 
155: \section{Numerical Method}
156: \label{sec:method}
157: 
158: \subsection{Flux-Limited Radiative Hydrodynamical Model}
159: \label{section:numerical}
160: We cast our numerical model in spherical coordinates
161: $\left(r,\phi,\theta\right)$, where $\phi$ is the azimuthal angle (or
162: longitude) and $\theta$ is the meridional angle (or latitude) measured
163: from the equator. Including both the Coriolis ($2{\bf \Omega\times
164: u}$) and centrifugal forces (${\bf\Omega\times} \left({\bf\Omega\times
165: r} \right)$), the equations of motion for the fluid can be written as
166: \begin{equation}
167: \pdt{\rho} + \nabla\cdot\left(\rho{\bf u}\right) = 0
168: \label{eq:continuity}
169: \end{equation}
170: 
171: \begin{equation}
172: \pdt{{\bf u}} + \left({\bf u}\cdot\nabla\right) {\bf u}= -
173: \frac{1}{\rho}\nabla{P} + \frac{1}{\rho}\nabla{\Phi} -2{\bf
174: \Omega\times u} - {\bf\Omega\times}\left({\bf\Omega\times r}\right)
175: \label{eq:momentum}
176: \end{equation}
177: 
178: The rotation frequency is given by $\Omega$, and the gravitational
179: potential, $\Phi=-{G\Mp\over r}$, varies only in the radial
180: direction. We neglect explicit viscosity, but some degree of numerical
181: viscosity is inevitable. We also include the curvature terms in
182: $\left({\bf u}\cdot\nabla\right) {\bf u}$. We solve equations
183: (\ref{eq:continuity}) and (\ref{eq:momentum}) on a stagged grid, where
184: scalars are defined in the center of cells and vectors on cell
185: boundaries. This method yields second-order spatial accuracy. Given
186: the decreasing grid size near the pole, our computational domain is
187: limited to $\theta =\pm 70^{\circ}$. Although excluding this region
188: neglects an avenue for energy re-distribution, even a modest amount of
189: rotation (approximately $2$ to $4\mathrm{ days}$ for the
190: hot-Jupiter's) will cause the dominate flow patterns to be
191: concentrated near the equator. We simulate the entire azimuth of the
192: planet, instituting periodic boundary conditions at $\phi = 0$ and
193: $2\pi$.  The radial extent of the domain extends from $7.95\times10^9
194: \mathrm{ cm}$ to $8.65\times 10^9 \mathrm{ cm}$, corresponding to
195: $1.06$ to $1.2 R_J$.  The pressure scale height on the day-side is
196: approximately $340$ km.
197: 
198: Our numerical radiative transfer scheme is capable of following the
199: temperature and radiation energy independently, linking them with a
200: given heating/cooling function. However, this degree of sophistication
201: is not necessary for calculations in which the gas is close to thermal
202: equilibrium. Instead, we use a one-fluid approximation where radiation
203: energy density is a simple function of temperature, $E = a T^4$. For
204: the temperatures considered here, the radiative energy density is much
205: lower than the thermal energy density. The internal and radiation
206: energy equations reduce to
207: \begin {equation}
208: \left[ \pdt{\epsilon} + ({\bf u}\cdot\nabla) \epsilon \right] = - P
209: \, \nabla \cdot {\bf u} - \nabla \cdot {\bf F}
210: \label{eq:energy}
211: \end{equation}
212: where $\epsilon=c_{v} \rho T$ is the internal energy density, $T$ is
213: the temperature, $c_v$ is the specific heat, and ${\bf F}$ is the
214: radiative flux.
215: 
216: To calculate the radiative flux we use the flux limited radiation
217: transfer approximation of \citet{levermore1981},
218: \begin {equation}
219: {\bf F} = - \lambda {\frac{c}{\rho\kappa}} \nabla E,
220: \end {equation}
221: where $\lambda$ is a non-constant flux limiter which prescribes the
222: relationship between the radiative flux and the radiative energy
223: gradient. We use the flux limiter developed by \citet{levermore1981},
224: given by
225: \begin{equation}
226: \lambda = \frac{2+R}{6+3R+R^2},
227: \label{lambda}
228: \end {equation}
229: where
230: \begin {equation}
231: R = \frac{1}{\rho \kappa} \frac{| \nabla E |}{E}
232: \label{Rdef}
233: \end {equation}
234: compares the scale height of the radiation energy to its mean free
235: path. This overall procedure allows for an accurate treatment of both
236: optically thick and thin regions; in the optically thick {\it
237: diffusion} limit $R \rightarrow 0$, $\lambda \rightarrow 1/3$, and
238: ${\bf F} \rightarrow - \, c \nabla E / 3 \rho \kappa $, while in the
239: optically thin {\it streaming} limit $R \rightarrow \infty$, $\lambda
240: \rightarrow 1/R$, and ${\bf F} \rightarrow c E$. While the flux
241: limiter $\lambda$ is an approximation, it is quite accurate in both
242: the optically thick and thin limits.
243: 
244: In order to follow the evolution on a long time-scales, the radiative
245: portion of the energy equation is solved {\it implicitly}. For
246: quasi-static radiative conditions such as those considered here, the
247: equations can be advanced much more rapidly than if they were
248: restricted by a radiative time-step. We use the successive
249: over-relaxation method (SOR) to solve the $\nabla \cdot {\bf F}$
250: portion of equation (\ref{eq:energy}), alternately updating even and
251: odd grid cells. The limiting factor for the numerical time-step then
252: becomes the Courant condition.
253: 
254: The advection scheme, described in \citet{hawley1984} and
255: \citet{kley1987}, is an extension of the first-order van Leer
256: \citep{vanleer1977} method known as the 'mono-scheme'. It employs
257: operator-splitting, where the finite difference equations are split
258: into parts, which are then solved separately always using the latest
259: values of the variables. The scheme yields semi-second order temporal
260: accuracy and allows for the accurate resolution of discontinuities in
261: the fluid flow with limited diffusion.
262: 
263: Finally, we use an ideal gas equation of state for the pressure,
264: $p=({\sf R}_{G}\rho T)/\mu$, with specific heat $c_{v}={\sf R}_{G} /
265: (\mu (\gamma-1))$ where ${\sf R}_{G}$ is the gas constant, $\gamma
266: =\frac {7}{5}$, and the mean molecular weight is fixed at $\mu =
267: 2.3$. Although the temperatures in the hydrodynamic models do become
268: hot enough to dissociate hydrogen molecules in some locations within
269: the planet, the region which we are most concerned with is well
270: described by a constant molecular weight. Radiative opacities are
271: found using the tables of \citet{pollack1985} for lower temperatures
272: coupled with \citet{alexander1994} for higher temperatures. These are
273: Rosseland mean opacities and include the effects of atomic, molecular,
274: and solid particulate absorbers and scatters. The opacity is one of
275: the largest uncertainties. The effect of composition, clouds,
276: settling, wavelength dependence, grains' condensation, sublimation, 
277: collisional growth, and sedimentation are but a few of the parameters
278: that alter the magnitude of the opacity. To address this uncertainty,
279: we explore the effect of varying opacity in \S\ref{sec:opacity}.
280: 
281: Our initial density and temperature profiles are taken from a
282: one-dimensional planetary evolutionary code. For details on the model,
283: see model B1 of \citet{bodenheimer2001}. Use of this model implies
284: that the simulations are initially spherically symmetric. Although
285: unrealistic, we allow simulations to run for sufficient time to relax
286: into equilibrium configurations; the details of the initial state are
287: lost. Figure (\ref{fig:initmod}) shows the initial temperature and
288: density profiles from the one-dimensional model near the top of the
289: planet. The model follows a $0.63M_J$ planet for $4.5
290: \mathrm{Gyr}$. The upper boundary was held at a fixed temperature of
291: $1200\mathrm{K}$ to simulate the effects of irradiation from the
292: central star. As noted in \citet{cho2006}, a quiescent initial start
293: neglects the effects of pre-established small-scale structures such as
294: eddies and jets. Given that we lack the resources to simultaneously
295: study these small scale effects and full three dimensional effects, we
296: assume that the dynamics will be dominated by the large scale
297: anisotropic heating imposed by the stellar irradiation.  In contrast
298: to the shallow-water approximation, specific vorticity can be
299: generated in our full 3D radiative hydrodynamic simulations.
300: Therefore, baroclinic instabilities can be excited spontaneously and
301: can lead to the generation of structure down to the resolution length
302: scales.
303: 
304: \begin{figure}
305: \plotone{./figs/initprof.ps}
306: \caption{Spherically symmetric temperature and density profiles from
307: the one-dimensional evolutionary models of \citet{bodenheimer2001}
308: used as an initial condition for the simulations presented here. The
309: models were run for $4.5 \mathrm{Gyr}$, during which time the
310: temperature at the upper boundary was held at $1200\mathrm{K}$ to
311: simulate the irradiation from the central star.}
312: \label{fig:initmod}
313: \end{figure}
314: 
315: To represent the impinging radiation, we set the temperature at the
316: upper boundary to $T = \mathrm{max}\left[1200\mathrm{K}
317: \left(cos\left(\phi\right)cos\left(\theta\right)\right)^{1/4},
318: 100\mathrm{K}\right]$,
319: approximating the stellar radiation field with a maximum at the
320: sub-stellar point, and a night-side held at $100\mathrm{ K}$. The
321: night-side temperature is chosen to be consistent with the
322: photospheric temperature derived in non-irradiated planetary evolution
323: models. This imposed temperature distribution will increase the scale
324: height of the atmosphere on the day-side, while cooling and
325: contracting it on the night-side. An interesting result is a planet
326: that is no longer spherical; the scale-height on the day-side is
327: somewhat larger than on the night-side. Given the computational
328: difficulties of modeling the extremely low density regions in the
329: upper atmosphere, we impose a movable upper boundary at the location
330: where $\rho<10^{-9}\mathrm{g/cm^3}$. At the bottom boundary we specify
331: the temperature flux taken from our initial one-dimensional models. As
332: mentioned above, the energy input from the star overwhelms the planets
333: intrinsic luminosity in the upper atmosphere, so the choice of this
334: boundary condition is not critical. We have run several simulations
335: with fixed temperature to verify the validity of this approximation,
336: and found little difference in the resulting energy distribution or
337: dynamics.
338: 
339: 
340: \section{Results of Numerical Simulations}
341: \label{sec:results}
342: 
343: \subsection{The Non-Rotating Model}
344: As a reference, we first present the results from an idealized and
345: artificial non-rotating model in which the gas giant is 
346: subject to one-sided stellar irradiation. The azimuthal pressure
347: gradient associated with the imposed temperature contrast drives
348: strong winds toward the night-side of the planet. The flow pattern is
349: symmetric with respect to the planet-star line, and fluid flows to the
350: night-side around both sides of the planet. As the material moves to
351: the night-side, it achieves maximum velocities of $\sim 3 \mathrm{
352: km/s}$ near the terminators ($\phi=\frac{\pi}{2}$ and
353: $\phi=\frac{3\pi}{2}$). This velocity corresponds to Mach numbers of
354: up to $2.7$. A substantial portion of energy is released in these
355: shocked regions. By the time the two symmetric flows converge on the
356: night-side they have both undergone substantial cooling and sink
357: radially inward, initiating a return flow at depth. In Figure
358: (\ref{fig:norot_teq}) we show the temperature distribution in the
359: equatorial plane. Despite the high winds, a clear day/night difference
360: persists. Also evident is the cool region at depth, which is due to
361: the confluence of the negative temperature gradient from the interior,
362: the positive temperature gradient near the surface, and the cool flow
363: returning to the day-side, completing the two, approximately
364: symmetric, azimuthal convection cells.
365: 
366: Figure (\ref{fig:norot_tnvphoto}) shows the temperature and velocity
367: at the photosphere ($\tau=\frac{2}{3}$) of the planet. Given the
368: changing scale height near the surface of the planet, these plots do
369: not represent a constant radial surface. Moving from the day to
370: night-side, a clear increase in the velocity near the terminator is
371: evident, followed by a drop at the convergent point. It should also be
372: noted that the convergence point is quite dynamical, oscillating in
373: both longitude and latitude. However, despite these oscillations, two
374: distinct convective cells remain. Night-side temperatures range from
375: $300-550 \mathrm{K}$, with an average night-side temperature
376: ($\frac{\pi}{2}<\phi<\frac{3\pi}{2}$) of $414\mathrm{K}$. Also evident
377: in Figures (\ref{fig:norot_teq}) and (\ref{fig:norot_tnvphoto}) is an
378: area of increased temperature near $180^{\circ}$. As the two
379: converging flows meet on the night-side, compressional heating drives
380: up the temperatures, and thus the local scale height. The resulting
381: increase is small, but can be seen both at depth and near the
382: photosphere.
383: 
384: \begin{figure}
385: \plotone{./figs/norot_teq.ps}
386: \caption{The temperature distribution at the equator of a non-rotating
387: hot-Jupiter. The inner boundary is assumed to be spherically
388: symmetric, with an outward energy flux fixed from the initial
389: one-dimensional model. The outer black area represents regions with
390: $\rho < 10^{-9} g/cm^3$, outside our movable boundary.}
391: \label{fig:norot_teq}
392: \end{figure}
393: 
394: \begin{figure}
395: \begin{center}
396: \includegraphics[height=8.0cm,width=12.0cm]{./figs/norot_tph.ps}
397: \includegraphics[height=8.0cm,width=12.0cm]{./figs/norot_vph.ps}
398: \end{center}
399: \caption{The temperature (upper panel) and velocity (lower panel) at
400: the photosphere of a non-rotating planet. A strong temperature
401: gradient between the day-side ($\phi = 0^\circ$) and the night-side
402: ($\phi=180^\circ$) is evident. The regions of largest $\nabla T$
403: correspond to fluid motions of $> 3 km/s$.}
404: \label{fig:norot_tnvphoto}
405: \end{figure}
406: 
407: \subsection{A Planetary Model with 3-Day Rotation}
408: It is widely believed that tidal forces within the atmospheres and the
409: envelope of hot-Jupiter's drive them to tidally locked spin
410: configurations on time-scales much shorter than the main sequence life
411: span of their host stars.  In this synchronous state, the planets'
412: spin frequency equals their orbital frequency. With this assumption,
413: hot-Jupiter's spin with periods on the order of $3\mathrm{-days}$. In
414: comparison to our own giant planets, this is a relatively slow spin
415: rate. Nevertheless, the associated Coriolis force significantly alters
416: the resulting flow dynamics and may have implications concerning the
417: ability of the planet to fully synchronize its spin. Another crucial
418: consideration that came to light during this study was the effects
419: associated with the initialization of the rotation. The
420: one-dimensional initial models described in Section
421: (\ref{section:numerical}) were non-rotating models, and the rate that
422: we chose to turn on the rotation had observable effects. We will
423: explore this effect in a subsequent paper.
424: 
425: \subsubsection{The Day-Side Isothermal Surface}
426: In Figure (\ref{fig:3d_teq}) we show the temperature distribution in
427: the equatorial plane for a simulation rotating with a period of $3
428: \mathrm{-days}$. The sub-solar point on the day-side is characterized
429: by an radially-extended nearly isothermal region with an effective
430: day-side temperature $T_d \sim 1200K$.  Upon adjusting to a
431: hydrostatic equilibrium, a slightly negative temperature gradient is
432: established so that the reprocessed stellar radiation can penetrate
433: into the planetary envelope.  Nevertheless the day-side photosphere is
434: essentially isothermal with a density profile
435: \begin{equation}
436: \rho (r) = \rho (r_b) {\rm exp}\left[ - (g \mu / {\sf R}_{G} T_d)
437: (r-r_b)\right],
438: \end{equation}
439: where $r_b$ is a planet's radius at the base of the isothermal region
440: and $g = G M_p/ r_b^2$ is the surface gravity of the planet.  The
441: imposed stellar heating falls off as a function of longitude, thus the
442: radial extent of this isothermal region decreases with the inclination
443: angle between the local zenith and the position of the host star
444: overhead.
445: \begin{figure}
446: \plotone{./figs/3d_teq.ps}
447: \caption{The temperature distribution at the equator of a planet
448: rotating with a period of $3 \mathrm{ days}$. The day-side is
449: characterized by a large isothermal area near the top, the extent of
450: which falls off with increasing longitude. A cool region at depth is
451: also evident due to the combined effects of a negative temperature
452: gradient from the internal heating, a positive temperature gradient
453: from the irradiation, and a cooling return flow from the night-side.}
454: \label{fig:3d_teq}
455: \end{figure}
456: 
457: \subsubsection{Azimuthal Effective-Temperature Distribution}
458: Below the isothermal photosphere, a cooler region at lower depths is also
459: evident on the day-side, associated in part with a cool return flow
460: from the night-side. Figure (\ref{fig:3d_tph}) shows temperature
461: distribution both throughout the entire photosphere, and focusing on
462: structure on the night-side. Despite the added effect of rotation, a
463: clear day-night delineation is still apparent, with the night-side
464: characterized by effective temperatures $T_n$ from $310$ to
465: $500\mathrm{ K}$. The average night-side temperature is $380\mathrm{
466: K}$, slightly smaller then the non-rotating simulation with the same
467: opacities. This slight decrease in average temperature is due to
468: increased cooling associated with rotationally modified flows
469: discussed in the next sub-section. Slightly hotter regions near the
470: terminators, associated with jets from the day-side, are apparent with
471: temperatures reaching $\sim 500\mathrm{ K}$.
472: 
473: \begin{figure}
474: \begin{center}
475: \includegraphics[height=8.0cm,width=12.0cm]{./figs/3d_tph.ps}
476: \includegraphics[height=8.0cm,width=12.0cm]{./figs/3d_tph_night.ps}
477: \end{center}
478: \caption{The temperature at the photosphere of a planet rotating with
479: a period of $3\mathrm{ days}$. The upper panel shows the distribution
480: over the entire planet, while the lower panel highlights the
481: temperature structure on the night-side from $\phi=\frac{\pi}{2}$ to
482: $\phi=\frac{3\pi}{2}$. A clear day-night delineation persists, despite
483: complicated dynamical structure, due to substantial radiation near the
484: terminators.}
485: \label{fig:3d_tph}
486: \end{figure}
487: 
488: 
489: \subsubsection{Thermal Current and the Corilois Effect}
490: The upper panel of Figure (\ref{fig:v_coriolis_ph}) shows the velocity
491: magnitude at the photosphere ($\left|v\right| = \sqrt{v^2_\phi+
492: v^2_\theta}$). As in the non-rotating model, material is moving quite
493: rapidly, reaching speeds of $\sim 4\mathrm{ km/s}$ near the
494: terminators.  Eastward (prograde or $v_\phi >0$, {\it i.e.} in the
495: same direction as the unperturbed spin) moving material appears to be
496: funneled from the sub-tropical latitudes ($\vert \theta \vert > 0$)
497: into an equatorial jet near $\phi=\frac{\pi}{2}$, while westward
498: (retrograde or $v_\phi <0$) moving material is pushed from the
499: subtropical zone toward the poles near $\phi=\frac{3\pi}{2}$. It is
500: these flow structures, rapidly advecting energy from the day-side,
501: that account for the hotter regions seen on the night-side in Figure
502: (\ref{fig:3d_teq}). To understand this structure, we must evaluate the
503: ${\bf\hat{\theta}}$ component of the Coriolis force, given by
504: $-2\Omega v_{\phi}sin\left(\theta\right)$, which is shown in the lower
505: panel of Figure (\ref{fig:v_coriolis_ph}). Although material near the
506: equator doesn't feel any Coriolis force in this direction, it is clear
507: that material at higher and lower (sub-tropical) latitudes does. The
508: asymmetry imposed by the rotation ({\it i.e.} the azimuthal
509: velocities) causes the fluid moving in eastward and westward
510: directions to behave significantly different then in the non-rotating
511: case.
512: 
513: The approximate magnitude of velocity can be estimated from equation
514: (\ref{eq:momentum}), considering only the pressure gradient
515: term. Assuming an approximately constant acceleration around to the
516: night-side of the planet given by $a \simeq -\frac{1}{\rho}\nabla P$,
517: the velocity at the terminator should be given by
518: \begin{equation}
519: v_T \simeq \left[\frac{2 \gamma {\sf R}_{G}}{\mu}
520: \left(T_d-T_n\right)\right]^{1/2}.
521: \label{eq:vel_terminator}
522: \end{equation}
523: Given a day-side temperature of $T_d \simeq 1200K$ and an average
524: night-side temperature of $T_n \simeq 350K$, flows should achieve Mach
525: numbers of $\sim 2$ near the terminator. The sound-speed 
526: $( \gamma {\sf R}_{G}
527: T/\mu )^{1/2} \sim 1.7$ km/s at the terminator, yielding a local Mach
528: number, as predicted, of $\sim 2$.
529: 
530: 
531: Also evident in the plot of velocity magnitude in Figure
532: (\ref{fig:v_coriolis_ph}) is the marked decrease in velocity where the
533: eastward and westward flows converge. Neither flow is able to
534: instigate circumplanetary flow at the surface. Figure
535: (\ref{fig:3d_veq_high}) shows an equatorial slice of the azimuthal
536: velocity $v_{\phi}$ at the equator (upper panel) and at higher
537: latitudes (lower panel). It is evident from this plot that the
538: eastward moving flow does continue around the planet at depth near the
539: equator, while the westward moving fluid continues around the planet
540: at higher (and lower) latitudes. Because of the effects of rotation
541: shown in Figure (\ref{fig:v_coriolis_ph}), the convergence point is
542: near $\phi=\frac{5\pi}{4}$ for the equatorial flow, and near
543: $\phi=\frac{3\pi}{4}$ for flows at higher and lower latitude. This
544: flow pattern implies that, upon converging, one of the two flows has
545: undergone substantially more cooling then its counterpart. Thus, at
546: the equator, when the eastward flow encounters the westward flow near
547: $\phi=\frac{5\pi}{4}$ (past ``mid night''), the former is cooler and
548: sinks below. The opposite is true at higher latitudes, with the
549: westward flow experiencing more cooling and sinking below the eastward
550: flow. This cooling trend can be seen in the lower panel of Figure
551: (\ref{fig:3d_tph}). Lastly, very little motion is apparent deeper in
552: the planet, as the flows are confined to a relatively small region
553: near the top of the planet, supporting our assumption of a spherically
554: symmetric inner boundary condition.
555: 
556: 
557: \begin{figure}
558: \begin{center}
559: \includegraphics[height=8.0cm,width=12.0cm]{./figs/3d_vph.ps}
560: \includegraphics[height=8.0cm,width=12.0cm]{./figs/3d_coriolis.ps}
561: \end{center}
562: \caption{The upper panel shows the magnitude of the velocity at the
563: photosphere, given by $\left|v\right| =
564: \sqrt{v^2_\phi+v^2_\theta}$. The lower panel shows the latitudinal
565: component of the Coriolis effect ($-2\Omega
566: v_{\phi}sin\left(\theta\right)$) at the photosphere of a planet
567: spinning at $3\mathrm{ days}$. The direction of the latitudinal
568: Coriolis force is different for eastward and westward moving material,
569: causing eastward flows to be focused toward the equator, while
570: westward flow is funneled toward the poles.}
571: \label{fig:v_coriolis_ph}
572: \end{figure}
573: 
574: \begin{figure}
575: \begin{center}
576: \includegraphics[height=8.0cm,width=9.0cm]{./figs/3d_veq.ps}
577: \includegraphics[height=8.0cm,width=9.0cm]{./figs/3d_vhigh.ps}
578: \end{center}
579: \caption{The velocity distribution at the equator (upper panel) and a
580: latitude of $\theta=35^{\circ}$ (lower panel) of a planet rotating
581: with a period of $3 \mathrm{ days}$. At the equator, eastward flow
582: (red) is able to circumnavigate the planet at depth, while at higher
583: and lower latitudes, it is westward flow (blue) that is able to
584: traverse all the way around the planet. The sinking of one flow under
585: the other is due to different degrees of cooling due to rotationally
586: altered dynamics.}
587: \label{fig:3d_veq_high}
588: \end{figure}
589: 
590: 
591: \subsubsection{Sub-surface Thermal Stratification}
592: Another method of visualizing the structure of the planet is through
593: pressure-temperature profiles. Of crucial importance when calculating
594: the emergent spectra, the pressure-temperatures at four different
595: longitudes are shown in Figure (\ref{fig:3d_PT}) for both the equator
596: and $\theta=35^{\circ}$. All profiles agree well at pressures above
597: $0.1 \mathrm{bar}$, again supporting the assumption of spherical
598: symmetry at depth.  However, below this pressure (or above this
599: height), their behaviors are quite different.
600: 
601: The day-side profile undergoes a significant temperature inversion
602: ({\it i.e.} temperature begins to increase with radius and decrease
603: with pressure). This temperature gradient allows radiative diffusion
604: from the photosphere to the planetary interior. This excess flux is
605: advected to the night-side deep down in the planetary envelope.
606: Comparing to Figure (\ref{fig:3d_teq}), it is evident that the lowest
607: temperature region, $\sim 650\mathrm{K}$, is associated in part with
608: the cool, return flow. This transitional region is analogous to the
609: thermocline in the terrestrial ocean which separates the surface and
610: deep water layers.
611: 
612: Near the upper atmosphere (where the pressure is low), the temperature
613: distribution contains a large, approximately isothermal region on the
614: day-side, clearly dominated by the stellar irradiation. In contrast,
615: the temperature of the night-side ($\phi=\pi$) monotonically decreases
616: throughout the entire atmosphere, reaching a temperature of $\sim 300
617: \mathrm{K}$ at the photosphere. While the day-side is fully radiative,
618: there is a region near the photosphere on the night-side that is
619: convectivly unstable in the radial direction.  As a consequence of
620: efficient convective transport, the $P-T$ distribution on the night
621: side is approximately adiabatic.
622: 
623: The profiles near the terminators (at both $\phi=\frac{\pi}{2}$ and
624: $\phi=\frac{3\pi}{2}$) exhibit more complex structures due to
625: considerable differences in the advective transport of heat. They also
626: show differences between the equatorial values (left-hand panel) and
627: higher latitudes (right-hand panel). At the equator, the temperature
628: at $\phi=\frac{\pi}{2}$ exhibits an isothermal region ranging from
629: $0.1$ to $10^{-3}$ bars. This terminator is associated with the
630: prograde flow. By the time this flow reaches $\phi=\frac{3\pi}{2}$ it
631: has cooled substantially. A slightly hotter region is evident from the
632: westward flow at lower pressures (larger radius) near the
633: photosphere. For the profiles from $\theta=35^{\circ}$, it is at
634: $\phi=\frac{3\pi} {2}$ where an approximately isothermal region exists
635: at depth. The $\phi=\frac{\pi}{2}$ profile at high latitudes decreases
636: monotonically, as very little heat is advected eastward.
637: 
638: \begin{figure}
639: \includegraphics[height=8.0cm,width=9.0cm]{./figs/3d_PT_eq.ps}
640: \includegraphics[height=8.0cm,width=9.0cm]{./figs/3d_PT_high.ps}
641: \caption{The pressure temperature profiles within a planet spinning
642: with a $3 \mathrm{day}$ period. The left-hand panel shows profiles at
643: the equator ($\theta=0$), while the right-hand panel shows profiles
644: from a latitude of $\theta=35^{\circ}$. The individual lines
645: represent $\phi=0$ (solid line), $\phi=\frac{\pi}{2}$ (dotted),
646: $\phi=\pi$ (dashed), and $\phi=\frac{3\pi}{2}$ (dash-dot). The dot on
647: each profile denotes the location of the photosphere.}
648: \label{fig:3d_PT}
649: \end{figure}
650: 
651: \subsection{Opacity Effects}
652: \label{sec:opacity}
653: The effect of opacity can not be understated. Opacities regulate the
654: efficiency of both the absorption of the incident stellar irradiation
655: on the day-side and the re-radiation from the night-side. As noted by
656: previous authors, major uncertainties in composition, metallicity, and
657: chemistry all cause significant changes to the opacity. Furthermore,
658: the inclusion of clouds, characterized by models of particle size
659: distributions and vertical extent, tends to smooth out wavelength
660: dependent opacities, resulting in spectral energy distributions that
661: more closely approximate blackbodies. In our current models, opacities
662: are found using the tables of \citet{pollack1985} for lower
663: temperatures coupled with \citet{alexander1994} for higher
664: temperatures. These are Rosseland mean opacities and include the
665: effects of atomic, molecular, and solid particulate absorbers and
666: scatters.
667: 
668: In light of the uncertainties associated with opacity, we have studied
669: the effect of uniformly changing the opacity by some multiplicative
670: factor. More detailed studies of specific, temperature and density
671: dependent augmentations to the opacity will be presented
672: elsewhere. Figure (\ref{fig:T_opc_ph}) shows the temperature
673: distribution across the photosphere of a planet with opacities reduced
674: by both a factor of $100$ (upper panel) and a factor of $1000$ (lower
675: panel). In comparison to Figure (\ref{fig:3d_tph}), night-side
676: temperatures are significantly higher for both models with lower
677: opacities, and distributions are smoother. In addition, due to changes
678: in the flow detailed below, the hottest spot is displaced slightly
679: from the sub-solar point. At the equator, the displacement is $\sim
680: 10^\circ$ and $\sim 20^{\circ}$ eastward (in the direction of
681: rotation) for opacity reductions of $100$ and $1000$ respectivly. This
682: displacement is largest at the equator, with maximum temperatures at
683: higher/lower latitudes occurring closer to the sub-solar longitude.
684: 
685: \begin{figure}
686: \begin{center}
687: \includegraphics[height=8.0cm,width=12.0cm]{./figs/low_100_Tph.ps}
688: \includegraphics[height=8.0cm,width=12.0cm]{./figs/low_1000_Tph.ps}
689: \end{center}
690: \caption{The temperature at the photosphere of a model with the
691: opacities reduced by a factor of $100$ (upper panel) and $1000$ (lower
692: panel). Lower opacity fluid absorbs the incident stellar irradiation
693: deeper in the atmospheres. This higher density material is able to
694: advect the energy to the night-side more efficiently, leading to larger
695: night-side temperatures.}
696: \label{fig:T_opc_ph}
697: \end{figure}
698: 
699: In a previous analysis (\citet{burkert2005}), we derived a formula for
700: the night-side temperature by equating the radiative timescale with
701: the crossing timescale.  This two-zone (day-night) model assumes that:
702: the advective heat flux is larger then the heat flux from the
703: interior, the heat carried by a day-night thermal current is
704: determined by the amount of radiative diffusion during the hemispheric
705: circulation, and the night-side radiates all the heat advected to its
706: proximity as a black body. With these assumptions, the night-side
707: temperature can be estimated by
708: \begin{equation}
709: T_n = \left(\frac{4vc_d^2}{3\pi\kappa_d\sigma R_p}\right)^{1/4}
710: \label{eq:Tburkert}
711: \end{equation}
712: which is an decreasing function of opacity. In this formula, $v$ is
713: the average advective speed which is on the order of the day-side
714: sound-speed $c_d$, and $\kappa_d$ is the day-side opacity. The
715: increase of night-side temperature with decreasing opacity reflects
716: the depth that incident stellar irradiation is deposited on the
717: day-side. If the atmosphere contains grains with an abundance and size
718: distribution comparable to that of the interstellar medium, only
719: shallow heating occurs on the day-side, and the circulation does not
720: effectively transmit heat to the night-side, which then cools well
721: below the day-side. However, as the abundance of grains in the
722: atmosphere is reduced, the stellar radiative flux penetrates more
723: deeply into the atmosphere on the day-side, and the higher density
724: atmospheric circulation carries a larger flux of heat over to the
725: night-side.
726: 
727: For the parameters used here, equation (\ref{eq:Tburkert}) predicts
728: $T_n \sim 250\mathrm{K}$ for the interstellar opacity simulation,
729: while for the lower opacity simulations, the formula predicts $\sim
730: 750\mathrm{K}$ and $\sim 950\mathrm{K}$ for reductions of $100$ and
731: $1000$ respectively. Inspection of Figures (\ref{fig:3d_tph}) and
732: Figure (\ref{fig:T_opc_ph}) show average night-side temperatures of
733: $\sim 380\mathrm{K}$, $\sim 700\mathrm{K}$, and $\sim 890\mathrm{K}$
734: for the same three cases. These results clearly indicate that the
735: night-side temperature decreases with the magnitude of the opacity.
736: The predicted values from equation (\ref{eq:Tburkert}) generally agree
737: with those from the simulation and it provides a framework for
738: understanding the global heat flow. The differences in these results
739: can be attributed to several factors; the surface heat flux carried by
740: the thermal current from the day to night-side is not entirely
741: radiated on the night-side, but rather cools as it travels and
742: advection at depth plays an important role in transporting heat. These
743: effects can be incorporated into a more comprehensive four-zone
744: (day-night and interior-photosphere) model where we examine the energy
745: transfer within the optically thick regions below the planetary
746: photosphere.
747: 
748: Before the presentation of the four-zone model, it is useful to
749: analyze the results of the numerical simulation.  These calculations
750: indicate that changing the opacity of the atmosphere not only alters
751: the night-side temperature, but it also modifies both the flow
752: dynamics and interior structure. For large opacity (our standard
753: case), the isothermal region on the day-side is relatively shallow,
754: and thus the increased cooling leads to lower night-side
755: temperatures. The large temperature differential promotes a fast flow
756: velocities around the planet. As the opacities are decreased, and the
757: night-side temperature increases and the velocities decrease.  In the
758: lowest opacity simulation, the velocity remains subsonic throughout
759: the entire simulation. A more uniform temperature across the planet
760: surface also allows for circumplanetary flow near the equator by
761: reducing the pressure gradient. Figure (\ref{fig:vph_opc_ph}) shows
762: the velocity at the photosphere of the two low-opacity
763: simulations. These results should be compared to the standard-value
764: results shown in Figure (\ref{fig:v_coriolis_ph}). As in the standard
765: case, flow at higher and lower latitudes still travels westward at the
766: $\frac{3\pi}{2}$ terminator and circumplanetary flow at the surface is
767: suppressed due to increased cooling times.
768: 
769: In Figure (\ref{fig:lowopc_PT}) we show the pressure-temperature
770: profiles at the equator for the two reduced opacity simulations. In
771: contrast to Figure (\ref{fig:3d_PT}) where the night-side was fully
772: convective, both lower opacity simulations exhibit isothermal regions
773: in the upper atmospheres around the entire planet. Below the
774: photosphere, there is a slightly negative temperature gradient so that
775: the day-night advective heat flux deep beneath the photosphere can
776: radiatively diffuse to the planet's photosphere. As it is to be
777: expected, the radial extent of the nearly isothermal atmosphere is
778: largest with the lowest opacity. Both simulations retain a convective
779: regions below the isothermal regions on the night-side.
780: 
781: \begin{figure}
782: \begin{center}
783: \includegraphics[height=8.0cm,width=12.0cm]{./figs/low_100_vph.ps}
784: \includegraphics[height=8.0cm,width=12.0cm]{./figs/low_1000_vph.ps}
785: \end{center}
786: \caption{Velocity at the photosphere of models with opacities reduced
787: by a factor of $100$ (upper panel) and $1000$ (lower
788: panel). Velocities, determined by both the day-night temperature
789: differential and the cooling efficiency, decrease with deceasing
790: opacity.}
791: \label{fig:vph_opc_ph}
792: \end{figure}
793: 
794: 
795: \begin{figure}
796: \includegraphics[height=8.0cm,width=9.0cm]{./figs/low_100_PT_eq.ps}
797: \includegraphics[height=8.0cm,width=9.0cm]{./figs/low_1000_PT_eq.ps}
798: \caption{The pressure temperature profiles at the equator of a planet
799: spinning with a $3 \mathrm{day}$ period. The left-hand panel shows
800: profiles for the run with opacities reduced by a factor of $100$,
801: while the right-hand panel shows profiles for a simulation with
802: opacities reduced by a factor of $1000$. The individual lines
803: represent $\phi=0$ (solid line), $\phi=\frac{\pi}{2}$ (dotted),
804: $\phi=\pi$ (dashed), and $\phi=\frac{3\pi}{2}$ (dash-dot). The dot on
805: each profile denotes the location of the photosphere, considerably
806: deeper then in the standard opacity simulation.}
807: \label{fig:lowopc_PT}
808: \end{figure}
809: 
810: 
811: 
812: \subsubsection{A Four-Zone Model for the Sub-Surface Thermal Structure of Planetary Atmospheres}
813: Discussions in the previous section clearly indicate that thermal
814: currents not only transport heat from the day to night, but also from
815: the night to day-side at other radii. In addition, the day-side
816: atmosphere is radiative and isothermal, while the night-side is
817: convective and adiabatic. Finally, quasi hydrostatic equilibria is
818: maintained in most regions except the subduction zones which separate
819: the thermal currents.  We now take these effects into consideration in
820: the determination of the night-side temperature and the sub-surface
821: thermal structure.
822: 
823: The two-zone approximation in equation (\ref{eq:Tburkert}) is based on
824: the assumption that the radiative and crossing time-scales in the
825: planetary atmosphere are nearly equal. We show in the next section
826: (Figure (\ref{fig:3d_rad_cross})) that this assumption is only
827: marginally satisfied on the night-side near the photosphere.  In the
828: standard opacity model, the radiative timescale around the terminators
829: is several orders of magnitude shorter than the crossing
830: timescale. This allows for significant cooling, yielding much lower
831: night-side temperatures then would be expected from equation
832: (\ref{eq:Tburkert}). Similar behavior is seen in the lower opacity
833: models.
834: 
835: In order to analytically account for the transport of heat well below
836: the photosphere on either side of the planet, we divide it into four
837: zones, representing the day-night and interior-photosphere regions.
838: There are then 4 sets of thermodynamic variables including: $P_{DS}$,
839: $P_{DI}$, $P_{NS}$, $P_{NI}$, $T_{DS}$, $T_{DI}$, $T_{NS}$, and
840: $T_{NI}$ where $P$ and $T$ are pressure and temperature, and the
841: subscripts $D$, $N$, $S$, and $I$ represent day, night, photospheric
842: surface and interior respectively. There are three additional
843: quantities which connect these regions: the velocity between the day
844: and night interiors, $v_{adv}$, the radial distance between the
845: photospheric surface and the planetary interior, $l_D$, and the
846: density of the gas at the planetary interior on the day-side
847: $\rho_{DI}$. These quantities can be solved simultaneously with the
848: following 11 equations.
849: 
850: Although the total stellar incident flux $F=L_\ast/4 \pi D^2$ (where
851: $L_\ast$ is the stellar luminosity and $D$ is the distance between the
852: star and the planet), irradiates only on the day-side, the condition
853: for thermal equilibriums implies that
854: \begin{equation}
855: 2 \sigma (T_{DS}^4 + T_{NS}^4)  = F.
856: \label{eq:4zone1}
857: \end{equation}
858: Hydrostatic equilibrium at the photosphere of both the day and night
859: sides gives:
860: \begin{equation}
861: P_{DS} = \frac{2}{3}\frac{g}{\kappa_{DS}}, \ \ \ \
862: P_{NS} = \frac{2}{3}\frac{g}{\kappa_{NS}}.
863: \label{eq:4zone2}
864: \end{equation}
865: For the simplicity of analytic approximation, we represent opacity as
866: \begin{equation}
867: \kappa = \kappa_0 T^\beta.
868: \label{eq:4zone3}
869: \end{equation}
870: In the temperature and density ranges which are relevant to the
871: present investigation, the opacity is roughly independent of density,
872: thus we adopt the approximation $\kappa_0 = 0.0391$ and $\beta =
873: 0.641$.
874: 
875: Our numerical results indicate that the pressure may change by more
876: then two orders of magnitude between the planetary surface and the
877: temperature inversion layer. In contrast, the temperature changes by
878: less than a factor of two. In the spirit of analytic simplicity, we
879: adopt an isothermal approximation for the hydrostatic envelope when we
880: determine the pressure at the planetary interior on the day-side, such
881: that
882: \begin{equation}
883: P_{DI} = P_{DS}exp\left[\frac{l_d g}{c_{s,d}^2}\right],
884: \label{eq:4zone4}
885: \end{equation}
886: where $c_{s,d} = (\gamma{\sf R}_{G} T_{DS}/\mu)^{1/2}$ is the sound
887: speed.  However, to calculate the temperature variation, we adopt a
888: radiative diffusion approximation in the radial direction. Including
889: the azimuthal advective transport, we have
890: \begin{equation}
891: 1 - \left( \frac{T_{DI}}{T_{DS}} \right)^{4-\beta} \simeq \left(4 -
892: \beta\right) \left[ \left(1 - {\rho_{DI} v_{adv} c_{s, d}^2 \over
893: F}\right) exp\left[\frac{g l_d}{c_{s,d}^2}\right] - 1 \right] \left( 1
894: + {T_{NS}^4 \over T_{DS}^4} \right),
895: \label{eq:4zone5}
896: \end{equation}
897: where 
898: \begin{equation}
899: \rho_{DI} = \mu P_{DI} / {\sf R}_{G} T_{DI}.
900: \label{eq:4zone6}
901: \end{equation}
902: We choose $l_D$ to be the depth where advection carries half of the
903: incident flux, {\it i.e.} where
904: \begin{equation}
905: \rho_{DI} v_{adv} c_{s, d}^2 \simeq F/2.
906: \label{eq:4zone7}
907: \end{equation}
908: The advective velocity at depth is driven by the pressure
909: differential, such that
910: \begin{equation}
911: v_{adv}^2 \simeq \frac{2 {\sf R}_{G} T_{DI}}
912: {\mu}ln\left(\frac{P_{DI}}{P_{NI}}\right).
913: \label{eq:4zone8}
914: \end{equation}
915: We assume the consequence of the advective transport is to thoroughly
916: mix the gas so that planetary interior becomes isothermal with
917: \begin{equation}
918: T_{NI} = T_{DI}
919: \label{eq:4zone9}
920: \end{equation}
921: 
922: For the standard opacity model, the night-side remains fully
923: convective. Assuming convection is efficient, the envelope of the
924: night-side is expected to be adiabatic, such that
925: \begin{equation}
926: \frac{T_{NI}}{T_{NS}} = \left(\frac{P_{NI}}{P_{NS}}\right)^{\frac{\gamma-1}
927: {\gamma}}.
928: \label{eq:4zone10}
929: \end{equation}
930: The energy equation for the night-side can be written as
931: \begin{equation}
932: \left(1-\frac{P_{DI}}{P_{NI}}\right) = -\frac{v_{conv}}
933: {v_{adv}}\frac{g\mu\pi R_p}{{\sf R}_{G} T_{NI}}
934: \label{eq:4zone11}
935: \end{equation}
936: where the convective velocity $v_{conv}$ can be obtained from the
937: mixing length theory. For the low opacity models, the upper regions of
938: the night-side become stabilized against convection. We can then
939: replace the energy equation with the diffusion approximation and
940: recalculate the thermal structure accordingly.
941: 
942: With these equations, we can provide a set of algebraic equations
943: which essentially reproduce the behavior of our numerical
944: simulation. These equations are also more comprehensive than that in
945: equation (\ref{eq:Tburkert}). For opacity similar to that of the
946: interstellar medium ({\it i.e.} no reduction in $\kappa_0$), the
947: night-side remains convective so that
948: equations(\ref{eq:4zone1})-(\ref{eq:4zone11}) are valid. In this
949: limit, $T_{DS}$ is sufficiently larger than $T_{NS}$ such that
950: equations (\ref{eq:4zone1}) and (\ref{eq:4zone2}) lead to complete
951: information on the surface layer of the day-side,
952: \begin{equation}
953: T_{DS}^4 = F/2 \sigma, \ \ \ \ \ \ P_{DS} = 2 g / 3 \kappa_0 T_{DS}^\beta.
954: \end{equation}
955: In comparison with the large variations in equation (\ref{eq:4zone4}),
956: the day-side is nearly isothermal with $T_{DI} \sim T_{DS}$. At the
957: depth where half of the incident stellar flux is advected from the day
958: to night-side, equation (\ref{eq:4zone7}) yields
959: \begin{equation}
960: P_{DI} v_{adv} \simeq {F \over 2}, \ \ \ \ \ l_d \simeq {c_{s, d} ^2 \over g}
961: {\rm ln} \left( {3 F \kappa_0 T_{DS}^\beta \over 4 g v_{adv} }\right).
962: \end{equation}
963: In the fully adiabatic night-side, we find from equations
964: (\ref{eq:4zone2}), (\ref{eq:4zone4}), (\ref{eq:4zone8}), and
965: (\ref{eq:4zone10}) that
966: \begin{equation}
967: P_{NS} = P_{DS} (T_{NI}/T_{NS})^\beta,
968: \end{equation}
969: and
970: \begin{equation}
971: {T_{NS} \over T_{DS}} = \left[ {\rm exp} \left(
972: { v_{adv}^2 - 2 g l \over 2 c_{s, d}^2 } \right) \right]^{(\gamma-1)/
973: (\gamma+ \beta (\gamma-1))}.
974: \label{eq:4zone12}
975: \end{equation}
976: From equations (\ref{eq:4zone8}), (\ref{eq:4zone9}), and
977: (\ref{eq:4zone11}), we find
978: \begin{equation} 
979: {\rm exp}\left[{v_{adv}^2 \over 2 c_{s, d} ^2}\right] - {v_{conv} \over v_{adv} }
980: {g \pi R_p \over c_{s, d}^2 } = 1. 
981: \label{eq:4zone13}
982: \end{equation}
983: In the limit that $v_{adv} < c_{s, d}$, equation (\ref{eq:4zone13}) reduces to
984: \begin{equation}
985: v_{adv} \sim ( 2 v_{conv} g \pi R_p) ^{1/3}.
986: \end{equation}
987: The convective speed can be estimated from the mixing length theory
988: which is mostly determined by $T_{NI} \sim T_{DS}$ and $\rho_{NI}$
989: These quantities have very little or no dependence on $\kappa_0$ so
990: that $v_{adv}$ does not change significantly as a function of
991: $\kappa_0$. In this limit, equation (\ref{eq:4zone12}) becomes
992: \begin{equation}
993: {T_{NS} \over T_{DS}} \propto \left[ {4 g v_{adv} \over 3 F \kappa_0
994: T_{DS}^\beta} \right]^{((\gamma-1) -1)/ (\gamma+ \beta (\gamma-1))}
995: \propto \kappa_0 ^{-1/4}.
996: \label{eq:4zone14}
997: \end{equation}
998: Thus, the four-zone model generates a similar $T_{NS}$ dependence on
999: $\kappa_0$ as equation (\ref{eq:Tburkert}). In comparison with
1000: numerical results, equation (\ref{eq:4zone14}) does reasonably well
1001: for high opacity convective simulations, but expression would be
1002: improved by considering the possibility that the night-side may also
1003: become radiative in the limit of very low $\kappa_0$.
1004: 
1005: \section{Analysis and Model Comparisons}
1006: \label{sec:comparison}
1007: A number of groups, including \citet{showman2002}, \citet{cho2003},
1008: \citet{burkert2005}, \citet{cooper2005}, \citet{cho2006}, and
1009: \citet{langton2007} have carried out non-linear numerical simulations
1010: studying the dynamics of hot-Jupiter atmospheres. Both the methods and
1011: results vary considerably. In this section, we attempt to compare
1012: their methodology to that presented here, concentrating on the
1013: assumptions that lead to differing results.
1014: 
1015: Both \citet{showman2002} and \citet{cooper2005} solve the primitive
1016: equations, the former using the EPIC code developed by
1017: \citet{dowling1998} and the later using the ARIES/GEOS Dynamical Core
1018: model initially developed by \citet{suarez1995}. The primitive
1019: equations, cast in isobaric coordinates, are widely used in
1020: meteorology. When deriving them from the full Navier-Stokes equation,
1021: a key assumption of hydrostatic equilibrium is invoked. Assumption of
1022: hydrostatic equilibrium explicitly links the thickness of a layer to
1023: its local temperature. Radial motion can only occur in conjunction
1024: with divergence along isobars, the magnitude determined by that which
1025: maintains the condition of hydrostatic equilibrium. With this
1026: assumption, the continuity equation and hydrostatic condition take the
1027: place of the radial momentum equation, and can be written as
1028: \begin{equation}
1029: \nabla_p {\bf \cdot v} + \frac{\partial \omega}{\partial p} = 0,
1030: \end{equation}
1031: where ${\bf v}$ is the horizontal velocity and $\omega$ is the
1032: vertical velocity with respect to pressure coordinates. The resulting
1033: radial velocity is slow in comparison to the horizontal (or isobaric)
1034: motion. Although an excellent assumption for thin terrestrial
1035: atmospheres, it neglects critical flows present in thicker atmospheres
1036: such as those of hot-Jupiter's. Although deviation from hydrostatic
1037: equilibrium provides the dominate radial acceleration, high velocity
1038: azimuthal flows on the night-side give rise to non-negligible radial
1039: forces neglected in the primitive equations that contribute to the
1040: ability of the converging flows to pass under one another as seen in
1041: Figure(\ref{fig:3d_veq_high}).
1042: 
1043: Our numerical results and analytic approximation clearly indicate
1044: the dependence of the night-side temperature on the efficiency
1045: of radiation transfer.  The heat flux carried by the thermal current
1046: is determined by the penetration depth of the incident stellar radiation.
1047: Radiative diffusion and convection are essential processes which 
1048: regulate the heat diffusion well below the photosphere of the planet.
1049: In conjunction with the primitive equations, both \citet{showman2002}
1050: and \citet{cooper2005} utilize a Newtonian radiative scheme to
1051: approximate stellar irradiation. In principle, this scheme is only
1052: appropriate for energy deposition and emission in optically thick 
1053: regions. The Newtonian radiative scheme
1054: provides a heating/cooling term to the energy equation that relaxes
1055: the temperature toward some pre-defined equilibrium distribution on
1056: some radiative timescale. The forcing term can be expressed
1057: \citep{cooper2005} as,
1058: \begin{equation}
1059: \frac{q}{c_v} =
1060: -\frac{T\left(\theta,\phi,p,t\right)-T_{eq}\left(\theta,\phi,p\right)}{\tau_{eq}\left(p\right)}.
1061: \end{equation}
1062: For the radiative timescale, $\tau_{eq}$, \citet{showman2002} assume a
1063: constant value given by that at $5 \mathrm{ bars}$, while
1064: \citet{cooper2005} use the calculations of \citet{iro2005} to set a
1065: radiative relaxation timescale that is dependent on the local
1066: pressure. As noted by the authors, this approximation is crude, but
1067: allows rapid computation of a large number of models. In order for the
1068: Newtonian approximation to be viable, the radiative timescale
1069: ($\tau_{rad}\approx\frac{E_T}{F}$) must be much longer than the
1070: crossing timescale ($\tau_{x}\approx\frac{\pi R_p}{2
1071: \left|v\right|}$). In this limit, the temperature distribution will be
1072: determined by the dynamics, rather than the assumed equilibrium
1073: distribution. This assumption becomes problematic in the upper
1074: atmosphere. In Figure (\ref{fig:3d_rad_cross}) we plot the ratio of
1075: timescales $\frac{\tau_{rad}}{\tau_{x}}$ at the photosphere of the
1076: planet.
1077: 
1078: \begin{figure}
1079: \begin{center}
1080: \includegraphics[height=8.0cm,width=12.0cm]{./figs/3d_rad_cross.ps}
1081: \end{center}
1082: \caption{The ratio of radiative timescale ($\tau_{rad}$) to the
1083: crossing timescale ($\tau_{x}$) at the photosphere of the simulation
1084: with interstellar opacities. Near the terminator, the primary region
1085: of interest, the radiative timescale is $\sim 5$ orders of magnitude
1086: shorter then the crossing timescale. This rapid cooling accounts for
1087: the sharp temperature gradient near the terminator seen in Figure
1088: (\ref{fig:3d_tph}).}
1089: \label{fig:3d_rad_cross}
1090: \end{figure}
1091: 
1092: It is clear from Figure (\ref{fig:3d_rad_cross}) that $\tau_{rad} <
1093: \tau_{x}$ throughout the photosphere. The most significant deviation
1094: occurs on the day-side, where there is very little motion, and
1095: $\tau_{rad}$ exceeds $\tau_{x}$ at the planetary photosphere 
1096: by many orders of magnitude. Near the terminator, where the stellar 
1097: irradiation falls drastically, the surface 
1098: radiative timescale is also much shorter then the dynamical timescale;
1099: it is here that the deviations from a Newtonian radiative scheme are
1100: most important. The flow radiates a significant portion of its energy
1101: before it flows to the night-side, allowing for the sharp edges in the
1102: temperature distribution seen in Figure (\ref{fig:3d_tph}), despite
1103: high velocity flows. In contrast, the Newtonian assumption that
1104: $\tau_{rad} >> \tau_{x}$, implies that the fluid carries a significant
1105: quantity of heat, leading to an overall distribution that will be more
1106: uniform with longitude. Finally, despite substantial radiation near
1107: the terminators, the near equality of $\tau_{rad}$ and $\tau_{x}$ at
1108: $\phi=\pi$ demonstrates that the temperature at the back-side is
1109: determined purely by the advection of energy by the flow.
1110: 
1111: To again highlight the effect of opacity, Figure
1112: (\ref{fig:radtime_opc_ph}) shows the ratio of radiative and crossing
1113: timescales, $\frac{\tau_{rad}}{\tau_{x}}$, for the runs with varying
1114: opacity. The upper and lower panels show the simulations with
1115: opacities reduced by a factors of $100$ and $1000$
1116: respectively. Again, the most important area in determining the
1117: behavior is near the terminators. In comparison to Figure
1118: (\ref{fig:3d_rad_cross}), it is obvious that lower opacity fluid is
1119: able to advect energy to the night-side more efficiently due to
1120: increased cooling times. As mentioned above, lower opacity on the
1121: day-side, allows heat to penetrate further into the planet, and
1122: re-radiation near the terminator is somewhat suppressed. Although
1123: Figure (\ref{fig:radtime_opc_ph}) shows a marked increase in
1124: $\frac{\tau_{rad}}{\tau_{x}}$ from the interstellar opacity
1125: simulation, the Newtonian approximation ($\tau_{rad} >> \tau_{x}$) is
1126: never satisfied. The largest values, $\frac{\tau_{rad}}{\tau_{x}}\sim
1127: 3$, only occur on the night-side in the high velocity circumplanetry
1128: jet.
1129: 
1130: \begin{figure}
1131: \begin{center}
1132: \includegraphics[height=8.0cm,width=12.0cm]{./figs/low_100_rad.ps}
1133: \includegraphics[height=8.0cm,width=12.0cm]{./figs/low_1000_rad.ps}
1134: \end{center}
1135: \caption{The ratio $\frac{\tau_{rad}}{\tau_{x}}$ at the photosphere of
1136: models with opacities reduced by a factor of $100$ (upper panel) and
1137: $1000$ (lower panel). Although lowering the opacity decreases the
1138: ability of the fluid to radiate its thermal energy, the radiative
1139: timescale remains several orders of magnitude shorter then the
1140: crossing timescale for most of the photosphere.}
1141: \label{fig:radtime_opc_ph}
1142: \end{figure}
1143: 
1144: The studies of \citet{cho2003}, \citet{menou2003}, \cite{cho2006}, and
1145: \cite{langton2007} take a different approach, concentrating on the
1146: effect of eddies and waves on the overall dynamics. \citet{cho2003}
1147: \citet{menou2003} and \cite{langton2007} use the shallow-water
1148: equations, while \cite{cho2006} solves the more generalized
1149: equivalent-barotropic equations. The equivalent-barotropic equations
1150: can be derived by vertically integrating the primitive equations
1151: described above. The resulting equations describe fluid flow in a
1152: single, isentropic layer whose scale height can vary. Stellar heating
1153: is prescribed by a deflection of the scale-height as a function of
1154: position. Concentrating on a single layer allows for the high
1155: resolution necessary to study the effect of turbulent eddies and
1156: waves. However, allowing for only a single layer implies that the
1157: atmosphere is radial isothermal. Although not a bad assumption in the
1158: upper regions on the day-side, the rest of the atmosphere exhibits
1159: significant radial temperature structure. As noted above, cooling in
1160: the upper regions of the night-side allows material to sink radial and
1161: initiate a return flow. Disallowing this flow would significantly
1162: alter the resulting dynamics.
1163: 
1164: The final dynamical study of hot-Jupiter's was done by
1165: \citet{burkert2005}. The numerical model they used is quite similar to
1166: the model presented in this paper. They solve the full hydrodynamical
1167: equations, given by equations (\ref{eq:continuity}),
1168: (\ref{eq:momentum}) and (\ref{eq:energy}) together with flux-limited
1169: radiation diffusion. However, they restrict their attention to the
1170: $r-\phi$ plane from $\phi=0\rightarrow\pi$, neglect the curvature
1171: terms in equation (\ref{eq:momentum}), and ignore the effects of
1172: rotation. Comparing our non-rotating results to \citet{burkert2005}
1173: 'Case 1' with our 'standard' opacity, we find that the two models
1174: agree quite well. The only substantial differences are that our
1175: backside temperature is slightly lower ($\sim 415 \mathrm{K}$ compared
1176: to their $480 \mathrm{K}$), and the convergence point is seen to
1177: oscillate in the simulations presented here. Our cooler temperatures
1178: can easily be explained by considering the increased compressional
1179: heating that the fluid in \citet{burkert2005} simulations experiences
1180: as it hits the boundary at $\phi=\pi$, and our extension into
1181: three-dimensions. Flows that are able to spread in latitude will cool
1182: more then those confined to the equator. Given that our non-rotating
1183: simulation results in two symmetric convective cells suggests that by
1184: restricting attention to $\phi=0\rightarrow\pi$, they did not miss any
1185: fundamental physics for that scenario. However, the addition of
1186: rotation significantly changes flow patterns and allows for increased
1187: cooling.
1188: 
1189: In summary, four fundamental differences are included in the models
1190: presented here, that are not included in some way in the previous
1191: methods: self-consistent radiative transfer, solution of the full
1192: radial momentum equation, $3$-dimensions, and rotation. Noticeable
1193: changes occur due to the inclusion of each. Self-consistent radiative
1194: transfer allows significant cooling as the fluid travels around to the
1195: terminators and night-side of the planet. Coriolis forces alter the
1196: structure of the flows, alternately compressing and diverging
1197: streamlines moving around the planet in different directions, and the
1198: multi-dimensional aspect changes the resulting temperature
1199: distribution. Finally, the full treatment of the radial momentum
1200: equation yields significant, non-hydrostatic radial structure and
1201: motion. Disallowing efficient cooling, and limiting radial motion,
1202: while subjecting the planet to the continual stellar energy input,
1203: will lead to much more uniform temperature distributions across the
1204: entire planet then those that we observe here.
1205: 
1206: \section{Discussion}
1207: \label{sec:discussion}
1208: In this paper we have considered the flow dynamics and heat
1209: redistribution in a tidally locked hot-Jupiter orbiting a solar-type
1210: star with an orbital period of $3 days$. We utilized a
1211: three-dimensional model that solves the full hydrodynamical equations
1212: and models radiative transfer the flux-limited radiation diffusion. We
1213: show that the temperature distribution across the planetary atmosphere
1214: is a sensitive function of its opacity. Our models exhibit strong
1215: day-night temperature contrasts, despite strong winds, the size of
1216: which increases with increasing atmospheric opacity. Large temperature
1217: differences are due, in a large part, to significant cooling of the
1218: flow near the terminators of the planet. Rotational effect
1219: significantly alter the flow pattern, allowing for increased cooling
1220: and suppressing surface circumplanetary flow in the higher opacity
1221: atmospheres.
1222: 
1223: In our standard model, opacities are calculated using the tables of
1224: \citet{pollack1985} for lower temperatures coupled with
1225: \citet{alexander1994} for higher temperatures. These are interstellar
1226: Rosseland mean opacities and include the effects of atomic, molecular,
1227: and solid particulate absorbers and scatters. Although there is
1228: significant room for improvement in our treatment of opacity, we
1229: explore the effect of uniformly reducing them by factors of $100$ and
1230: $1000$. In agreement with the two-dimensional simulations of
1231: \citet{burkert2005}, we find that lower opacities yield higher
1232: night-side temperatures. Lower opacities allow incident stellar
1233: irradiation to be deposited at larger depths on the day-side,
1234: decreasing the cooling rate as this energy is advected to the
1235: night-side, and increasing night-side temperatures. The relation
1236: between the night-side effective temperature and the opacity in the
1237: planetary atmosphere is verified by a more comprehensive four-zone
1238: analytic model which takes detailed account of the effect of radiation
1239: transfer on both the day and night-side of the planet. This model
1240: highlights the importance of appropriately treating radiation transfer
1241: in the simulation of atmospheric dynamics on close-in gas giant
1242: planets.
1243: 
1244: The actual value of opacity in the atmospheres of these planets is
1245: highly uncertain. However, the day-side temperature of short period
1246: gas giant planets around solar type stars is $T_{DS} \sim 1200$K,
1247: which is below the grain destruction temperature. We note that the
1248: flow returning to the day-side is dredged up from a cooler interior
1249: ($T_{DI} < T_{DS}$) at the sub-solar point. Thus, throughout the
1250: thermal circulation, refractory magnesium-rich grains are
1251: preserved. Other species of less refractory silicate grains will
1252: sublimate near the planetary photosphere on the day-side, condense on
1253: night-side surface, and be carried along with the returning current
1254: beneath the surface. These thermal currents become turbulent as they
1255: generate specific vorticity and excite instabilities. Small grains are
1256: well coupled to the gas and turbulence induces them to collide and
1257: coagulate. In a static atmosphere, particles with sizes $s_p$ and
1258: density $\rho_p$ attain attain terminal velocities of
1259: \begin{equation}
1260: v_{t} \sim (g \rho_p s_p / \rho_g)^{1/2}
1261: \end{equation}
1262: in a background gas with density $\rho_g$. Since the distance over
1263: which these particle attain their terminal speed, $L_{term} \sim
1264: \rho_p s_p / \rho_g$, is smaller than the density scale height $\sim
1265: c_{s, d}^2 / g$, particles larger than $\sim 1 mm$ cannot be carried
1266: by the upwelling current on the day-side, and are left well beneath
1267: the planet's photosphere. This potential channel for grain-gas
1268: separation implies that it is very likely that the opacity in the
1269: planet's atmosphere is well below that of the interstellar medium. A
1270: detailed analysis of the grain evolution in the atmosphere of
1271: short-period gas giants will be presented elsewhere.
1272: 
1273: We also presented a detailed comparison of our model to previous
1274: dynamical models of \citet{showman2002}, \citet{cho2003},
1275: \citet{burkert2005}, \citet{cooper2005}, \citet{cho2006}, and
1276: \citet{langton2007}. The approaches among these groups varies
1277: greatly, and we have attempted to highlight several of the crucial
1278: differences. The most significant differences that we include are
1279: solving the full fluid equations in all three-dimensions, rotation,
1280: and our treatment of radiation via the flux-limited diffusion
1281: method. Flux-limited radiation transport allowed for a self-consistent
1282: treatment of the flow of radiation throughout the planet. This
1283: treatment produced significantly shorter radiative timescales then
1284: previously calculated, yielding much lower night-side temperatures.
1285: 
1286: The dynamics and heat distribution within the atmospheres giant
1287: planets allow us to probe fundamental questions surrounding planet
1288: formation. The diversity of planetary sizes must ultimately result
1289: from interior structure variations arising from formation and
1290: evolutionary processes. These atmospheres not only serve as a valuable
1291: observable links to the interior, but may also play a role in
1292: regulating planetary contraction rates. As suggested in
1293: \citet{burrows2000} and \citet{burrows2006} an isothermal atmosphere
1294: will result in decreased heat transfer and increased planetary
1295: contraction timescales. However, a strong day-night temperature
1296: difference may allow an avenue for cooling via the night-side. Based
1297: on the simulations presented here, it is clear that the efficiency of
1298: redistribution decreases with increasing opacity. High opacity
1299: atmospheres, similar to interstellar values, will have cool night-side
1300: temperatures, providing an avenue for internal heat loss and radial
1301: contraction. If, as is expected, there is a significant amount of
1302: grain growth and sedimentation, thus reducing the atmospheric opacity,
1303: the planets night-side will also contain a large isothermal region,
1304: possibly allowing for the retention of internal energy. Diversity in
1305: atmospheric opacity may lead to diversity in planetary radii. With the
1306: promise of improved observational techniques, including transit
1307: spectroscopy and full phase light-curves, the relation between opacity
1308: and temperature differential should be testable in the near future.
1309: 
1310: \acknowledgements We thank Geoff Bryden, Peter Bodenheimer, Katherine
1311: Kretke, and Gordon Ogilvie for supplying the original version of our
1312: numerical scheme, initial structural model, and many helpful
1313: discussions. We would also like to acknowledge the use of NASA's High
1314: End Computing Program computer systems. This work is partially
1315: supported by NASA (NAGS5-11779, NNG04G-191G, NNG06-GH45G), JPL
1316: (1270927), NSF(AST-0507424, PHY99-0794) and IGPP.
1317: 
1318: 
1319: \bibliographystyle{aa}
1320: \bibliography{ian}
1321: \end{document}
1322: 
1323: