1: %\documentclass[12pt,preprint]{aastex}
2:
3: \documentclass{emulateapj}
4:
5: %\documentclass[10pt]{aastex}
6: %\usepackage{emulateapj5}
7:
8: %\renewcommand{\thefootnote}{\fnsymbol{footnote}}
9: %\setcounter{footnote}{-1}
10:
11: \def\mj{$M_{\rm J}\ $}
12: \def\rj{$R_{\rm J}\ $}
13: \def\etal{{et~al.\,}}
14: \def\mo{M$_\odot$}
15: \def\ro{R$_\odot$}
16: \def\lo{L$_\odot\,$}
17: \def\mp{M$_{\rm p}$}
18: \def\rp{R$_{\rm p}\,$}
19: \def\lbol{L$_{bol}\,$}
20: \def\mstar{M$_{\ast}$}
21: \def\rstar{R$_{\ast}$}
22: \def\teff{$T_{\rm eff}\,$}
23: \def\teffs{$T_{\rm eff}$s$\,$}
24: \def\mic{$\mu$m$\,$}
25: \def\undertext#1{$\underline{\smash{\hbox{#1}}}$}
26: \def\sles{\lower2pt\hbox{$\buildrel {\scriptstyle <}
27: \over {\scriptstyle\sim}$}}
28: \def\sgreat{\lower2pt\hbox{$\buildrel {\scriptstyle >}
29: \over {\scriptstyle\sim}$}}
30:
31: %\null\voffset=+0.0pc % +0.0pc
32:
33: \slugcomment{Accepted to Ap.J.}
34:
35: \begin{document}
36:
37: \title{Theoretical Spectra and Light Curves of Close-in
38: Extrasolar Giant Planets and Comparison with Data}
39:
40: \author{A. Burrows\altaffilmark{1}, J. Budaj\altaffilmark{1,2} \& I. Hubeny\altaffilmark{1}}
41:
42: \altaffiltext{1}{Department of Astronomy and Steward Observatory,
43: The University of Arizona, Tucson, AZ \ 85721;
44: burrows@zenith.as.arizona.edu, budaj@as.arizona.edu, hubeny@aegis.as.arizona.edu}
45:
46: \altaffiltext{2}{Astronomical Institute, Tatranska Lomnica, 05960 Slovak Republic}
47:
48: \begin{abstract}
49:
50:
51: We present theoretical atmosphere, spectral, and light-curve models for
52: extrasolar giant planets (EGPs) undergoing strong irradiation for which
53: {\it Spitzer} planet/star contrast ratios or light curves have been published (circa June 2007).
54: These include HD 209458b, HD 189733b, TrES-1, HD 149026b, HD 179949b, and $\upsilon$ And b.
55: By comparing models with data, we find that a number of
56: EGP atmospheres experience thermal inversions and have stratospheres.
57: This is particularly true for HD 209458b, HD 149026b, and $\upsilon$ And b.
58: This finding translates into qualitative changes in the planet/star contrast
59: ratios at secondary eclipse and in close-in EGP orbital light curves.
60: Moreover, the presence of atmospheric water in abundance is fully consistent with
61: all the {\it Spitzer} data for the measured planets. For planets
62: with stratospheres, water absorption features invert into emission features
63: and mid-infrared fluxes can be enhanced by a factor of two. In addition,
64: the character of near-infrared planetary spectra can be radically
65: altered. We derive a correlation between the importance of such stratospheres
66: and the stellar flux on the planet, suggesting that close-in EGPs bifurcate
67: into two groups: those with and without stratospheres. From the finding
68: that TrES-1 shows no signs of a stratosphere, while HD 209458b does,
69: we estimate the magnitude of this stellar flux breakpoint.
70: We find that the heat redistribution parameter, P$_n$, for the
71: family of close-in EGPs assumes values from $\sim$0.1 to $\sim$0.4.
72: This paper provides a broad theoretical context for the future direct
73: characterization of EGPs in tight orbits around their illuminating stars.
74:
75:
76: \end{abstract}
77:
78:
79: \keywords{stars: individual (HD 209458, HD 189733, TrES-1, HD 149026,
80: $\upsilon$ And, HD 179949)---(stars:) planetary systems---planets and satellites: general}
81:
82:
83:
84: \section{Introduction}
85: \label{intro}
86:
87: To date, more than 250 extrasolar planets have been discovered and
88: more than 29 of them are transiting their primary star\footnote{See J. Schneider's
89: Extrasolar Planet Encyclopaedia at http://exoplanet.eu, the Geneva Search Programme at
90: http://exoplanets.eu, and the Carnegie/California compilation at http://exoplanets.org}.
91: One transiting planet is a ``Neptune" (GJ 436b), but the rest are giant planets
92: with an impressively wide range of masses
93: and radii that speak to the heterogeneity of the family of close-in
94: EGPs ({\bf E}xtrasolar {\bf G}iant {\bf P}lanets). Table \ref{t1} lists
95: these transiting EGPs and some of their relevant properties, along with
96: many of the references to the observational and discovery papers from which
97: these data were taken. Table \ref{t2} lists useful data for the corresponding
98: primary stars, including their masses, luminosities, radii, and approximate
99: ages. Both tables are in order of increasing planetary semi-major axis and,
100: considering the pace of the field, both should be considered provisional.
101: Not shown are the eccentricities, which are generally small, but which for HAT-P-2b, GJ 436b, and
102: XO-3b are $\sim$0.507, $\sim$0.14, and $\sim$0.22, respectively. For these three
103: close-in EGPs, significant tidal heating and, perhaps, forcing by an unseen
104: companion are implied.
105:
106: Radial-velocity measurements for a non-transiting EGP provide a lower
107: limit to its mass, but little else. However, the transiting EGPs yield
108: radii as well, and resolve the $\sin{i}$ ambiguity to reveal the planets' masses.
109: These data provide physical constraints with which detailed evolutionary and structural
110: models that incorporate irradiation and migration can be tested (see, e.g.,
111: Burrows et al. 2007a; Guillot et al. 2006). With superb photometric accuracy, the wavelength-dependence
112: of the transit radii can in principle provide a measure of a planet's atmospheric
113: composition (Fortney et al. 2003). In this way, sodium has been
114: detected in HD 209458b (Charbonneau et al. 2002) and water has been identified
115: in both HD 189733b (Tinetti et al. 2007, but see Ehrenreich et al. 2007) and HD 209458b (Barman 2007).
116: Moreover, high-precision optical photometry has constrained
117: (perhaps, measured) the geometric albedo of the close-in EGP HD 209458b
118: (Rowe et al. 2006, 2007). HD 209458b's optical albedo is very low ($\sim$3.8$\pm4.5$\%), in keeping with
119: the predictions of Sudarsky et al. (2000) when the alkali metals, and not
120: clouds, dominate absorption in the atmosphere and Rayleigh scattering
121: dominates scattering.
122:
123: Nevertheless, using current technology, transit measurements have
124: limited utility in characterizing the atmospheres and compositions of these
125: planets. Astronomers require more direct detections of the planet's spectrum to
126: probe its chemistry and atmospheric properties. This is in the tradition of
127: remote sensing in the solar system and of the Earth. Until recently, it had
128: been thought that the light from an extrasolar planet had to be separated from
129: under the glare of its parent star using high-angular resolution, extremely high-contrast
130: imaging. This is still the case in the optical for the cool ``wide-separation" EGPs
131: (Burrows et al. 2004; Sudarsky et al. 2005; Burrows 2005) and terrestrial planets,
132: for which the planet-star contrast ratio is $\sim$10$^{-9}$ to $\sim$10$^{-10}$,
133: but such performance has not yet been demonstrated.
134:
135: However, for the hot close-in
136: EGPs, the planet-star contrast ratios in the mid-infrared are
137: much more favorable (Burrows, Sudarsky, \& Hubeny 2003,2004),
138: oftimes exceeding 10$^{-3}$. This capability has led to a breakthrough
139: in the study of exoplanets. With the infrared space telescope
140: {\it Spitzer} (Werner \& Fanson 1995), using its IRAC and MIPS cameras
141: and the IRS spectrometer, one can now measure the summed light of the
142: planet and the star in and out of secondary eclipse, and thereby,
143: from the difference, determine the planet's spectrum at superior conjunction.
144: Moreover, for a subset of the closest EGPs it is possible to use {\it Spitzer} to measure
145: their flux variations with planetary phase between transit and secondary eclipse.
146: Hence, for the close-in EGPs in the near- to mid-infrared, and without
147: the need to separately image planet and star, the direct detection of
148: planetary atmospheres via low-resolution spectroscopy and precision infrared (IR) photometry
149: is now a reality.
150:
151: The secondary eclipse fluxes have now been measured for five transiting EGPs (HD 189733b,
152: TrES-1, HD 209458b, HD 149026b, GJ 436b), but not yet in all {\it Spitzer} bands.
153: In addition, using the IRS spectrometer, spectra between $\sim$7.5
154: $\mu$m and $\sim$15 $\mu$m of two transiting EGPs at
155: secondary eclipse have been obtained (HD 189733b [Grillmair et al. 2007]
156: and HD 209458b [Richardson et al. 2007]). Though at very low resolution, these are the
157: first measured spectra of any extrasolar planet.
158: Furthermore, light curves have been measured for three EGPs ($\upsilon$ And b
159: [Harrington et al. 2006: at 24 $\mu$m], HD 179949b [Cowan et al. 2007: at
160: 8 $\mu$m], and HD 189733b [Knutson et al. 2007b: at 8 $\mu$m]). Only one of these (HD 189733b) is transiting
161: and has an absolute calibration. For none of the latter three are there
162: extant light-curve measurements for more than one {\it Spitzer} band;
163: for some of these EGPs only upper limits in a few of the other bands have been
164: determined. Table \ref{t3} summarizes all the direct
165: detection data for the EGP family obtained to date (circa June 1, 2007), along
166: with associated references, comments, and table notes. Clearly, in the next
167: year or two we can expect a great deal more secondary eclipse and light-curve
168: data in the various {\it Spitzer} IRAC and MIPS bands. However,
169: there has already been significant progress in measuring
170: EGP atmospheres.
171:
172: This paper is a continuation of our series of interpretative studies (e.g., Burrows et al. 2005, 2006, 2007b)
173: of the direct measurements of close-in EGPs. Here, we analyze the secondary eclipse and light-curve
174: data summarized in Table \ref{t3} for the EGPs HD 189733b, HD 209456b, TrES-1,
175: HD 149026b, HD 179949b, and $\upsilon$ And b and make theoretical predictions in support of
176: future {\it Spitzer} planet measurement campaigns. Importantly, by fitting the current data we
177: extract physical information concerning the atmospheres, compositions, and thermal
178: profiles of these first six objects listed in Table \ref{t3}. We have explored the
179: dependence of the spectra and light curves on the heat redistribution factor
180: P$_n$ (Burrows et al. 2006), on atmospheric metallicity, and on the possible
181: presence of a stratospheric absorber. The recent analysis by Burrows et al. (2007b)
182: of the IRAC data of Knutson et al. (2007c) indicates that HD 209458b boasts a thermal inversion
183: that radically alters the {\it Spitzer} fluxes and their interpretation. In fact,
184: in Burrows et al. (2007b), we speculate that thermal inversions and stratospheres
185: may play a role in the atmospheres of many close-in EGPs and are a new feature in the
186: study of transiting planets. A similar conclusion was reached by Fortney et al. (2006),
187: specifically in the context of HD 149026b.
188:
189: We find that we can fit all the secondary eclipse and light-curve data,
190: except for the nightside flux of HD 189733b and its day/night contrast.
191: While we can fit its dayside secondary eclipse flux, we suspect that
192: HD 189733b will require a more sophisticated day/night
193: redistribution model than we now employ (\S\ref{tech}).
194: We note that Knutson et al. (2007b) conclude that the dimmest and
195: brightest spots on HD 189733b reside on the same
196: hemisphere and that the dimmest spot is shifted from the anti-stellar
197: point by as much as $\sim$30$^{\circ}$. Our current light curve models
198: are symmetric about the peak.
199:
200: We find that the degree of longitudinal
201: heat redistribution (P$_n$) may vary from planet to planet, hinting at a variety
202: of meteorological conditions and day/night contrasts within the family of close-in
203: EGPs. Moreover, as also concluded in Burrows et al. (2005), we can not obtain
204: good fits at secondary eclipse without the presence of water in abundance in the
205: atmospheres of these irradiated EGPs. This is particularly true for
206: TrES-1 and HD 209458b. Though we will not dwell in this paper
207: on metallicity, we find that the metallicity dependence of the secondary-eclipse fluxes
208: is not strong, but that it is in principle measurable, and that the metallicity
209: dependence of the variation of the planetary flux with phase is only modest.
210: Importantly, we also conclude that upper-atmosphere absorption in the optical
211: by an as-yet unknown molecule, and the concommitent thermal inversions
212: and stratospheres, provide better fits to some of the data.
213:
214: In \S\ref{tech}, we describe our numerical techniques, the new heat redistribution model,
215: and how we generate stratospheres. This section is supported with
216: Appendices \S\ref{redist}, \S\ref{app2}, \S\ref{app3}, and \S\ref{app5},
217: in which we provide details concerning the heat redistribution model and derive
218: some analytic formulae concerning atmospheric physics with day-night
219: coupling. In particular, in \S\ref{app5}, we address the enhancement at
220: secondary eclipse in the integrated planetary flux at Earth over and above
221: what would be expected if the planet emitted isotropically. For a radar antenna,
222: this would be its ``gain" factor.
223: In \S\ref{tp}, we discuss the derived temperature-pressure
224: profiles on the day and the night sides for all six EGPs highlighted in this investigation. We show
225: that the $\tau = 2/3$ decoupling layers for the {\it Spitzer} IRAC and MIPS (24-$\mu$m)
226: band fluxes are all above the isothermal region of an irradiated EGP's atmosphere
227: and, hence, that {\it Spitzer} does not probe these deeper regions. For the same reasons, we
228: find that the presence of a thermal inversion at altitude and of a stratosphere in
229: some EGP atmospheres can significantly alter these {\it Spitzer} fluxes and their relative strengths.
230: In \S\ref{tp}, we also provide representative planet spectra to demonstrate that most of the planet's
231: flux emerges at shorter wavelengths than are accessible to {\it Spitzer}, and, hence,
232: that {\it Spitzer} probes only a small tail of the emergent flux distribution.
233: This may be of relevance when JWST is available to follow up on the
234: {\it Spitzer} EGP data and, even earlier, as the JWST exoplanet campaign is being designed.
235:
236: In \S\ref{planetstar}, we present the best-fit planet-star contrast ratios at secondary
237: eclipse for four of the transiting EGPs for which these have been measured (all
238: but GJ 436b, for which see Deming et al. 2007 and Demory et al. 2007), as well
239: as various comparison models to gauge a few parameter dependences.
240: Then, in \S\ref{curves} we match our theoretical phase light curves with
241: the three measured light curves and derive approximate planetary parameters.
242: The paper is brought to a close in \S\ref{conclusions} with a synopsis
243: of our results and a general discussion of the issues raised.
244:
245:
246: \section{Numerical Techniques}
247: \label{tech}
248:
249: Our model atmospheres are computed using the updated code {\sc CoolTLUSTY}, described
250: in Sudarsky, Burrows, \& Hubeny (2003), Hubeny, Burrows, \& Sudarsky (2003),
251: and Burrows, Sudarsky, \& Hubeny (2006), which is a variant of the
252: universal spectrum/atmosphere code {\sc TLUSTY} (Hubeny 1988; Hubeny \& Lanz 1995).
253: The molecular and atomic opacities are taken from Sharp \& Burrows (2007)
254: and the chemical abundances, which include condensate rainout, are derived
255: using the thermochemical model of Burrows \& Sharp (1999), updated as described
256: in Sharp \& Burrows (2007) and in Burrows et al. (2001).
257:
258: To handle convection, we use standard mixing-length theory,
259: with a mixing length equal to the pressure scale height. The stellar
260: irradiation boundary condition is numerically challenging, and has
261: not been done properly by some workers in the past. To ensure
262: an accurate numerical solution with a non-zero incoming specific intensity,
263: we use the formalism discussed in Hubeny, Burrows, \& Sudarsky (2003).
264: The stellar spectral models are taken from Kurucz (1994) for the six stars
265: listed in Table \ref{t2} that are the primaries of the EGPs upon which we
266: focus in this paper. The day and night sides are approached differently,
267: with the nightside, quite naturally, experiencing no incident flux, but
268: receiving heat from the irradiated dayside using the new algorithm described in
269: Appendix \S\ref{redist}. An important additional feature of our new heat redistribution
270: formalism is the capacity to match both the entropy and the gravity at the base
271: of both the day- and the night-side atmospheres. Since the inner convective zone,
272: which constitutes most of the planet, is isentropic, this is the physically
273: correct procedure. For a dayside calculation, we can assume a given interior
274: flux effective temperature, $T_{\rm int}$ (a standard number could be 75 K).
275: For a given gravity and irradiation regime, this leads to an atmosphere
276: solution on the dayside. This solution incorporates an entropy in
277: the convective zone. For the nightside atmosphere, we can adjust the nightside
278: $T_{\rm int}$ until the entropy in its convective interior matches that
279: found in the dayside convective zone. One product of this procedure is a connection
280: between dayside and nightside $T_{\rm int}$ that has a bearing on overall
281: planet cooling and shrinkage (Burrows et al. 2007a). However, since for a given
282: measured planet radius, this mapping does not have a significant effect on the
283: close-in planet's spectrum, we do this here only approximately and leave
284: to a later paper a general discussion of this topic.
285:
286: The simple parametrization we use to simulate the effects of
287: an extra stratospheric absorber entails placing an absorber
288: with constant opacity, $\kappa_{\rm e}$, in the optical frequency
289: range $(\nu_0, \nu_1)$ = ($3\times 10^{14}$ Hz, $7\times 10^{14}$ Hz)
290: and high up at altitude, where the pressure is below a prescribed value,
291: generally take to be 0.03 bars. Hence, $\kappa_{\rm e}$
292: is the most important parameter to be adjusted. We could
293: easily introduce a specified frequency and/or depth
294: dependence, but this would add free parameters
295: which we feel are not justified at this stage. We have also
296: generated models in which TiO and VO are allowed to assume their
297: chemical-equilibrium upper-atmosphere abundances (Sharp \& Burrows 2007),
298: uncorrected for the cold-trap effect (\S\ref{conclusions}; Burrows et al.
299: 2007b), and to generate a stratosphere. These TiO/VO models produce
300: qualitatively the same effects as do our ad hoc models. However, in this paper
301: we prefer the flexibility of the $\kappa_{\rm e}$ prescription.
302:
303:
304: In Burrows et al. (2006), once the day- and night-side atmospheres were
305: calculated, we used a 2D radiative transfer code to determine the
306: integrated emissions ``at infinity" at a given viewing angle from
307: the planet-star axis for the day- and night-side hemispheres. These numbers were then
308: transformed into a light curve as a function of wavelength and planetary phase angle using the
309: methods described in Sudarsky et al. (2000, 2005). However, we have found that
310: since the detailed shape of the light curve connecting the day- and the night-side fluxes
311: is likely to be only poorly constrained for the foreseeable future and since our model is
312: symmetric about secondary eclipse, it is inappropriate to invest a disproportionate
313: amount of effort in performing expensive 2D transfer calculations. Rather, we
314: invest our efforts in obtaining state-of-the-art day- and the night-side fluxes
315: at secondary and primary eclipse and then connect them with a simple, though well-motivated, curve.
316: Therefore, our light curve model is:
317: \begin{equation}
318: {\cal{C}} = \frac{D + N}{2} + \frac{D - N}{2}\cos{\alpha}\sin{i}\, ,
319: \label{phasel}
320: \end{equation}
321: where $\cal{C}$ is the planet/star flux ratio, $D$ is the dayside flux (see Fig. \ref{fig4}), $N$ is the
322: nightside flux, $\alpha$ is the phase angle, and $i$ is the inclination angle ($\sim{90^{\circ}}$
323: if in transit). This is the form adopted in Cowan et al. (2007). Using this simpler
324: approach, one does not imply more precision than is warranted at this preliminary
325: stage of inquiry.
326:
327: We have revisited the question of what type of averaging of
328: the incoming radiation from the star over the surface of the
329: planet is best suited to describe the planetary spectrum close
330: to the secondary eclipse. We had demonstrated earlier (Sudarsky
331: et al. 2005) that detailed 2D phase-dependent spectra averaged
332: over the phase are equal, within a few percent, to the spectrum
333: computed for a representative model atmosphere that is
334: constructed assuming that the incoming flux is distributed
335: evenly over the surface of the dayside.
336: In the usual terminology, this corresponds to the flux distribution
337: factor $f=1/2$ (Burrows et al. 2000)\footnote{We allow the planet to be irradiated
338: by the full flux received from the star; it is only deeper in the
339: atmosphere where the energy is transported to the nightside. See Appendix \ref{redist}.}.
340:
341: However, the spectrum of a planet observed close to the secondary
342: eclipse should be biased toward a higher flux than that obtained
343: using the $f=1/2$ model. This is because the hottest part of the
344: planet, the substellar point, is seen as emerging from the planet
345: perpendicularly to the surface, and, thus, with the lowest
346: amount of limb darkening. Therefore, the contribution of the
347: hottest part of the planet is maximized. To study this effect,
348: we have computed a series of model atmospheres corresponding to
349: a number of distances from the substellar point, and have integrated
350: the individual contributions to get the flux received by an
351: external observer at a phase close to superior conjunction. It
352: turns out that the flux is very close to that computed for
353: $f=2/3$, which is the value we subsequently use in all simulations
354: presented here. There is a simple analytic argument
355: why $f$ should be approximately equal to $2/3$ which we present
356: in Appendix \S\ref{app5}.
357:
358: The formalism for $D$ and $N$ is the best we have fielded to date. Nevertheless,
359: a 3D general circulation model (GCM) that incorporates state-of-the-art opacities,
360: compositions, and radiative transfer will be needed to properly address
361: day-night heat redistribution, the vortical and zonal mass motions, and the positions
362: of the hot and cold spots. Such a model is not yet within reach, but there have been
363: preliminary attempts to treat this physics (Showman \& Guillot 2002; Cho et al. 2003;
364: Menou \etal 2003; Williams et al. 2006; Cooper \& Showman 2005; Lunine \& Lorenz 2002). The proper GCM
365: physics remains the major uncertainty in current planetary secondary eclipse and
366: light-curve modeling.
367:
368: \section{Temperature$-$Pressure Profiles}
369: \label{tp}
370:
371: We have used the techniques outlined in \S\ref{tech} and in
372: Appendix \S\ref{redist} and the data in Tables \ref{t1} and \ref{t2} to create models
373: of six of the close-in EGPs in Table \ref{t3} for which there are {\it Spitzer}
374: secondary eclipse or phase light curve data. These planets are HD 209458b, HD 189733b, TrES-1,
375: HD 179949b, HD 149026b, and $\upsilon$ And b. The product of our
376: investigation is an extensive collection of atmosphere models, with associated
377: spectra, for many combinations of planet, P$_n$, metallicity, and values of $P_0$/$P_1$ (\S\ref{redist}).
378: We have, however, settled on presenting in this paper only the central and
379: essential results for each planet, in the knowledge that the data are not yet
380: exquisitely constraining.
381:
382: We focus on models with solar-metallicity (Asplund,
383: Grevesse, \& Sauval 2006) opacities, ($P_0, P_1$) = (0.05, 0.5) bars,
384: and an interior flux $T_{\rm int}$ of 75 K. These are our baseline model parameters.
385: For a given measured planet radius, the dependence of the models on
386: $T_{\rm int}$ is extremely weak. We find that the specific pressure range ($P_0, P_1$)
387: in which most of the heat carried from the dayside to the nightside is conveyed
388: plays a role in the planet-star flux ratios, but a subtle one. Therefore, in
389: lieu of a comprehensive, and credible, 3D climate model, we prefer not to claim too much
390: concerning the details of atmospheric circulation and heat redistribution.
391: We also explore the effects of a stratospheric absorber with an optical
392: opacity of $\kappa_{\rm e}$ (see also Burrows et al. 2007b).
393: We find that such models will be most important for close-in EGPs with the greatest
394: stellar insolation and guided by this principle, particularly relevant for
395: HD 209458b, HD 149026b, and $\upsilon$ And b, we explore the consequences.
396:
397: Figure \ref{fig1} portrays in six panels the temperature-pressure ($T/P$) profiles
398: of a representative collection of dayside and nightside models of the six close-in EGPs of this study.
399: For the dayside, the different curves correspond to different values of P$_n$ from
400: 0.0 (no redistribution) to 0.5 (full redistribution) and to
401: models with and without stratospheric absorbers. For the nightside, P$_n$ ranges
402: from 0.1 to 0.5. For all models, the radiative-convective boundaries are identified
403: and are quite deep (on the far right of each panel). When $\kappa_{\rm e} \ne 0$, the $T/P$
404: profiles show distinct thermal inversions.
405:
406: There are quite a few generic features in evidence on these panels. The first is that
407: the atmospheres are never isothermal. Since the opacities in the optical, where most of
408: the stellar irradiation occurs, and in the infrared, where most of the reradiation occurs,
409: are very different, a quasi-isothermal inner region interior to $\sim$1 bar is always bounded by
410: an outer region in which the temperature decreases (Hubeny, Burrows, \& Sudarsky 2003).
411: As Fig. \ref{fig1} indicates, the magnitude of the temperature decrease
412: from the plateau to the $\sim$10$^{-5}$ bar level for dayside models
413: without stratospheres is $\sim$1000 K. With a stratosphere, the outward increase
414: from a pressure of $\sim$0.1 bars can be correspondingly large. For our nightside models,
415: the monotonic decrease is $\sim$500$-$1000 K. Models with temperature inversions due
416: to a strong absorber at altitude clearly stand out in the panels of Fig. \ref{fig1} and may
417: result from the presence of a trace species, TiO/VO, or a non-equilibrium species
418: (Burrows et al. 2007b). The possible effect of such upper-atmosphere absorbers
419: on the $T/P$ profiles and the resultant dayside spectra are exciting new features
420: of the emerging theory of irradiated EGPs.
421:
422: The discussion above is made more germane when we note that the decoupling surfaces for the
423: IRAC and MIPS (24-$\mu$m) channels, the effective photospheres where $\tau_{\lambda} \sim 2/3$,
424: are all in the outer zone. Figure \ref{fig1} indicates their positions for the dayside P$_n = 0.3$
425: model of TrES-1. They are at similar pressures for all other models. The photospheres
426: for shorter wavelengths not accessible to {\it Spitzer} are deeper in.
427: Figure \ref{fig2} portrays these ``formation," ``brightness," or photospheric temperatures as a
428: function of wavelength for three models of TrES-1 with different values of P$_n$,
429: and for both the dayside and nightside, and illustrates this fact. The
430: approximate wavelength intervals of the {\it Spitzer} bands are superposed.
431: Though the IRAC 1 flux can decouple at interesting depths, the photospheres for the $Y$, $J$, $H$,
432: and $K$ bands are generally deeper. The photospheres in the far-IR beyond $\sim$10 $\mu$m
433: are high up at altitude and we repeat that {\it Spitzer} photometry does not probe the isothermal
434: region so characteristic of theoretical close-in EGP atmospheres. Moreover, since the {\it Spitzer}
435: observations are probing the outer regions of the atmosphere most affected by stellar
436: irradiation and which can have inversions, the treatment of the outer boundary
437: condition due to incoming stellar flux must be accurate. Slight errors or uncertainties
438: in the outer boundary condition of the transfer solution, or in the upper-atmosphere opacities, can
439: translate into significant errors in the predicted {\it Spitzer} fluxes. This is particularly true
440: longward of $\sim$10 $\mu$m. As a result, measured fluxes in both the near-IR and
441: mid-IR are useful diagnostics of upper-atmosphere absorbers and thermal
442: inversions (Hubeny, Burrows, \& Sudarsky 2003; Burrows et al. 2006; Burrows
443: et al. 2007b; Knutson et al. 2007c). All these caveats and points must be
444: borne in mind when interpreting the {\it Spitzer} EGP data.
445:
446: As a prelude to our discussions in \S\ref{planetstar} and \S\ref{curves} of the {\it Spitzer}
447: planet-star flux ratios at secondary eclipse and during an orbital traverse,
448: and to emphasize the fact that {\it Spitzer} does not comprehensively probe the
449: irradiated planet's atmosphere, we plot in Fig. \ref{fig3} theoretical
450: dayside spectra ($\lambda$F$_{\lambda}$ versus log$_{10}$($\lambda$)) for three models of TrES-1
451: at zero phase angle (superior conjunction). Superposed on the plot are the positions
452: of the near-IR, IRAC, and MIPS bands. Such a figure allows one to determine at a glance
453: the wavelengths at which most of the flux is radiated (at least, theoretically).
454: As Fig. \ref{fig3} suggests, most of the planet's flux
455: emerges in the near-IR, not in the IRAC or MIPS channels. In fact, depending
456: on the planet, no more than one fifth to one third of the planet's flux comes
457: out longward of $\sim$3.6 $\mu$m (IRAC 1), while no more than one twentieth to one
458: tenth emerges longward of $\sim$6.5 $\mu$m, the ``left" edge of the IRAC 4 channel.
459: Since much of the best EGP data have been derived in IRAC channel 4, one must acknowledge
460: that they may represent very little of the total planetary emissions.
461:
462: Finally, we call the reader's attention to the slight bumps (on the nightside)
463: and depressions (on the dayside) between 0.05 and 1.0 bars in the $T/P$ profiles depicted
464: in Fig. \ref{fig1}. This region is near where we imposed heat redistribution using
465: the formalism described in \S\ref{redist}. The actual shapes of these profiles
466: are determined by this mathematical procedure and other algorithms will produce
467: different local thermal profiles. Note that with this formalism it is possible
468: at the higher P$_n$s ($\ge 0.35$) for the nightside to be hotter than the dayside
469: at the same pressure levels in the redistribution region. While this
470: may seem at odds with thermodynamics, what is essential is that energy
471: is conserved and is redistributed at optical depths that are not either
472: too low or too high. If the former, the absorbed stellar heat would be reradiated
473: before it can be carried to the nightside. If the latter, then the stellar radiation
474: can not penetrate to the conveyor belt. For our default choice of $P_0$ and $P_1$, $\tau_{\rm Rossland}$
475: is generally between $\sim$0.3 and $\sim$6. These depths are not unreasonable, but
476: our redistribution algorithm is clearly only a stopgap until a better GCM
477: can be developed and justified.
478:
479:
480: \section{Planet-Star Flux Ratios$-$Comparison with Data}
481: \label{planetstar}
482:
483: We discuss below and in turn model fits for each transiting EGP at secondary eclipse.
484: However, first we present our results collectively and in summary fashion.
485: Figure \ref{fig4} in four panels portrays for the four transiting EGPs the correspondence between
486: the secondary eclipse data and representative models of the planet-star
487: flux ratio as a function of wavelength from 1.5 $\mu$m to 30 $\mu$m. This figure
488: summarizes our major results. The models are for values of P$_n$ of
489: 0.1, 0.3, and 0.5 and various values of $\kappa_{\rm e}$. The data
490: include 1-$\sigma$ error bars and can be found in Table \ref{t3}.
491: As Fig. \ref{fig4} indicates, we can fit all the published
492: data. The P$_n$ dependence for both stratospheric models and models
493: without inversions is strongest in the $K$ band and in IRAC 1. In fact,
494: in the near-IR, models with inversions depend very strongly on P$_n$.
495: Fig. \ref{fig5} for HD 209458b in the near-IR indicates this most clearly. This finding
496: implies that measurements at these shorter IR wavelengths are good diagnostics of
497: P$_n$, particularly if inversions are present.
498:
499: Importantly, including a non-zero $\kappa_{\rm e}$
500: and generating a stratosphere results in a pronounced enhancement
501: longward of IRAC 1, particularly in IRAC 2 and 3, but also at MIPS/24 $\mu$m and at
502: the 16-$\mu$m peak-up point of {\it Spitzer}/IRS. Hence, fluxes at the longer IR wavelengths
503: might be good diagnostics of thermal inversions. The models in Fig. \ref{fig4} for
504: HD 209458b and HD 149026b demonstrate this feature best.
505:
506: No attempt has been made to achieve refined fits, but the correspondence between
507: theory and measurement, while not perfect, is rather good for all the planets. Moreover,
508: different EGPs seem to call for different values of P$_n$ and $\kappa_{\rm e}$, and, hence, perhaps,
509: different climates, degrees of heat redistribution, compositions, and upper-atmospheric physics.
510: The light-curve analyses in \S\ref{curves} also suggest this. A goal is to relate
511: these measured differences with the physical properties of the star and planet and these
512: infrared secondary eclipse data allow us to begin this program in earnest.
513:
514: We note that comparisons between model and data must actually be
515: made after the band-averaged flux-density ratios of the detected
516: electrons are calculated. Performing this calculation slightly mutes
517: the predicted variation from channel to channel in the IRAC regime.
518: This is particularly true when comparing IRAC 1 and IRAC 2,
519: even if a pronounced spectral bump at and near $\sim$3.6 $\mu$m
520: obtains, as it does for models with modest or no thermal
521: inversion. However, to avoid the resultant clutter and
522: confusion, we do not plot these bandpass-averaged predictions
523: on Figs. \ref{fig4} and \ref{fig5}. We now turn to case-by-case
524: discussions of the secondary eclipse measurements and models.
525:
526:
527: \subsection{HD 209458b}
528: \label{hd209sub}
529:
530: The first transiting EGP discovered was HD 209458b (Henry et al. 2000; Charbonneau et al. 2000)
531: and it has since been intensively studied. The direct-detection data of relevance
532: to this paper are summarized in Table \ref{t3}. The most relevant data are the geometric albedo
533: constraints in the optical from MOST (Rowe et al. 2006,2007), a $K$-band upper limit
534: using IRTF/SpeX from Richardson, Deming, \& Seager (2003), a MIPS/24-$\mu$m photometric point
535: from Deming et al. (2005) (and its possible update), a low-resolution {\it Spitzer}/IRS spectrum from
536: Richardson et al. (2007), and, importantly, photometric points in IRAC channels 1 through 4
537: from Knutson et al. (2007c). These data collectively provide useful information
538: on the atmosphere of HD 209458b.
539:
540: Motivated by the recent data of Knutson et al. (2007c),
541: Burrows et al. (2007b) provide partial theoretical explanations for HD 209458b's
542: atmosphere. Much of the discussion in the current paper concerning HD 209458b is taken
543: from Burrows et al. (2007b), so we refer the reader to both the Burrows et al. (2007b)
544: and Knutson et al. (2007c) papers for details. However, here we expand upon
545: the discussion in those works where it is necessary to put the HD 209458b findings in the broader
546: context of the EGPs listed in Table \ref{t3}. The major conclusion of Burrows et al. (2007b) is
547: that the atmosphere of HD 209458b has a thermal inversion and a stratosphere, created
548: by the absorption of optical stellar flux by a strong absorber at altitude,
549: whose origin is currently unknown. This converts absorption features into emission
550: features, while still being consistent with the presence of water in abundance.
551:
552: All relevant data, save the albedo constraint in the optical, are displayed in the upper-left panel
553: of Fig. \ref{fig4}. Figure \ref{fig5} includes the Knutson et al. (2007c) IRAC 1
554: point and the Richardson, Deming, \& Seager (2003) upper limit in $K$ and focuses on the
555: near-IR. Also provided on both figures are models for
556: P$_n$ = 0.1, 0.3, and 0.5, without and with an extra stratospheric absorber.
557: The latter is implemented using the formalism in outlined in \S\ref{tech}
558: and a $\kappa_{\rm e}$ of 0.1 cm$^2$/g.
559:
560: Figure \ref{fig4} shows that the low upper limit of Richardson, Deming, \& Seager (2003) in the $K$ band
561: that was problematic in the old default theory (Burrows, Hubeny, \& Sudarsky 2005; Fortney et al. 2005;
562: Barman, Hauschildt, \& Allard 2005; Seager et al. 2005; Burrows et al. 2006) is consistent with the models with an extra
563: upper-atmosphere absorber in the optical, particularly for higher values of P$_n$.
564: This is more clearly seen in Fig. \ref{fig5}.
565: Moreover, the theoretical peak near the IRAC 1 channel ($\sim$3.6 $\mu$m) in the old
566: model without an inversion is reversed with the extra absorber into a deficit
567: that fits the Knutson et al. (2007c) point. The theory without an extra absorber
568: at altitude predicts that the planet-star flux ratio in the IRAC 2 channel should be lower than
569: the corresponding ratio at IRAC 1. However, with the extra absorber the relative strengths in
570: these bands are reversed, as are the Knutson et al. (2007c) points. This reversal
571: is a direct signature of a thermal inversion in the low-pressure regions of
572: the atmosphere, and an indirect signature of the placement of the heat redistribution
573: band (see Appendix \S\ref{redist}). The top-left panel of Fig. \ref{fig1} depicts the corresponding
574: temperature-pressure profiles and the thermal inversion at low pressures
575: introduced by the presence of an extra absorber in the optical
576: that is indicated by the data.
577:
578: As Fig. \ref{fig4} also demonstrates,
579: there is a significant difference in the IRAC planet-star
580: flux ratios between the old default model without an inversion
581: and the new models with a stratospheric absorber, and
582: that the models with a stratosphere fit the IRAC
583: channel 1, 2 and 4 flux points much better. However, the height of the
584: IRAC 3 point near 5.8 $\mu$m is not easily fit, while maintaining the good
585: fits at the other IRAC wavelengths and consistency with the $K$-band limit.
586: Theoretically, the positions of the IRAC 3 and IRAC 4 photospheres should be close to one
587: another, so this discrepancy is surprising. Nevertheless, the IRAC 2, 3, and
588: 4 data together constitute a peak, whereas in the default theory
589: an absorption trough was expected.
590:
591: The 24-$\mu$m MIPS point obtained by Deming et al. (2005) is lower
592: than the prediction of our best-fit model. However, the flux at this point
593: is being reevaluated and may be closer to $\sim$0.0033$\pm{0.0003}$
594: (D. Deming, private communication). If the new number supercedes the old
595: published value, then our best-fit model(s)
596: with inversions fit at this mid-IR point as well (Fig. \ref{fig4}). Higher
597: planetary fluxes longward of $\sim$10 $\mu$m are generic features of stratospheric
598: inversions.
599:
600: The 1-$\sigma$ optical albedo limit from Rowe et al. (2007) is a very low 8.0\%. For comparison, the
601: geometric albedos of Jupiter and Saturn are $\sim$40\%. However, such a low number was predicted
602: due to the prominence in the optical of broadband absorptions by the alkali metals sodium
603: and potassium in the hot atmospheres of irradiated EGPs (Sudarsky et al. 2000;
604: their ``Class IV"). The associated planet-star flux ratios are $\sim$10$^{-5}$$-$10$^{-6}$.
605: This low albedo is consistent with the identification of sodium
606: in the atmosphere of HD 209458b using HST/STIS transit spectroscopy (Charbonneau
607: et al. 2002). Both these datasets suggest that any clouds that might reside in the
608: atmosphere of HD 209458b are thin. A thick cloud layer would reflect light
609: efficiently, leading to a high albedo. If the extra stratospheric absorber
610: is in the gas phase, and there is no cloud, then our new thermal inversion models
611: are easily consistent with the low albedo derived by Rowe et al. (2006,2007).
612: If the extra absorber is a cloud, the cloud particles must have a low scattering albedo
613: and can not be very reflecting. This rules out pure forsterite, enstatite, and iron
614: clouds.
615:
616: The IRS data are noisy, but their flattish shape is consistent with
617: our model(s) with thermal inversions and a stratosphere. Richardson et al. (2007) suggest
618: that there is evidence in the IRS data for two spectral features: one near
619: 7.78 $\mu$m and one near 9.67 $\mu$m. However, we think the data are too noisy to
620: draw this conclusion. Richardson et al. (2007) also suggest that
621: the flatness and extension of their data to shorter wavelengths implies
622: the near absence of water, since previous theoretical models predicted
623: a spectral trough between $\sim$4 $\mu$m and $\sim$8 $\mu$m. However,
624: if there is an outer thermal inversion, as we here and
625: in Burrows et al (2007b) argue is the case for HD 209458b,
626: a trough is flipped into a peak for the same water abundance.
627: This renders moot the use of the spectral slope at the edge
628: of the IRS spectrum to determine the presence or absence of water.
629: One of our major conclusions, implicit in Figs. \ref{fig4} and \ref{fig5},
630: is that water is not depleted at all in the atmosphere of HD 209458b.
631:
632: The recent controversies surrounding such an interpretation occasion
633: the following remarks. One thing to bear in mind concerning the use of these IRS spectra
634: to infer compositions is that they are very low resolution. The use of classical
635: astronomical spectroscopy to identify constituents stems from the ability
636: at much higher resolution to see characteristic features at precise wavelengths
637: and patterns of absorption or emission lines to high accuracy. This allows one to make
638: element and molecule identifications in a narrow wavelength range without
639: a global view across the whole spectrum. However, at the low resolution of the IRS,
640: no individual water features are accessible. There are water features near $\sim$10 $\mu$m,
641: but they would require a $\lambda/\Delta\lambda$ is excess of $\sim$2000 to identify. Otherwise,
642: all one sees is the collective effect of millions of lines and the resulting pseudo-continuum
643: (the band structure). Clearly, when the data are low-resolution, a {\it global} photometric and spectral fit
644: is necessary to address the issue of composition. The signature of water's presence comes from the goodness
645: of the global fit across the entire spectrum from the optical to the mid-IR.
646: The good fit we obtain in Fig. \ref{fig4},
647: lead us to conclude that the IRS, IRAC 4, and MIPS data for HD 209458b are
648: consistent with the presence of water in abundance.
649:
650: Note that if the $T/P$ profile were entirely flat (but see Fig. \ref{fig1}), whatever
651: the opacity and molecular abundances the emergent spectrum would be a perfect black body
652: and would give no hint concerning composition. Transit spectra would then be our
653: only reliable means of determining atmospheric composition (Fortney et al. 2003;
654: Barman 2007; Tinetti et al. 2007; Ehrenreich et al. 2007). However, as we have argued, there is every indication that
655: the $T/P$ profiles of strongly irradiated EGPs are not flat (Fig. \ref{fig1}). As a result,
656: spectral measurements of irradiated EGPs can be usefully diagnostic of
657: both composition and non-trivial thermal profiles.
658:
659:
660: \subsection{HD 189733b}
661: \label{hd189}
662:
663: Models for HD 189733b at secondary eclipse, with and without an
664: extra upper-atmosphere absorber, are portrayed in the upper right-hand panel of
665: Fig. \ref{fig4}. They include $\kappa_{\rm e}$ = 0.0 cm$^2$/g models with P$_n$ = 0.1, 0.3, and 0.5,
666: and one $\kappa_{\rm e}$ = 0.04 cm$^2$/g model with P$_n$ = 0.3. The IRAC 4 data at
667: 8 $\mu$m from Knutson et al. (2007b) [brown], the IRS peak-up
668: point at 16 $\mu$m obtained by Deming et al. (2006) [gray],
669: and the IRS spectrum between $\sim$7.5 $\mu$m and $\sim$13.5 $\mu$m
670: from Grillmair et al. (2007) [gold] are superposed on the figure.
671: Though data in the other IRAC channels and at 24 $\mu$m have been
672: taken and reduced, they have yet to be published.
673:
674: As this panel indicates, the IRAC 4 point can be fit by models that include
675: most values of P$_n$, with a very slight preference for lower values from 0.1
676: to 0.3. The IRS data are not well-calibrated, but evince the slight turndown
677: at the shorter wavelengths characteristic of atmospheres with weak or no stratospheric
678: absorber. This turndown is in contrast with the behavior of the Richardson et al. (2007)
679: IRS data for HD 209458b, and reinforces the conclusion that a thermal inversion,
680: if present in HD 189733b, is very slight (see the
681: upper-right panel of Fig. \ref{fig1}). However, the 16-$\mu$m point of
682: Deming et al. (2006) is a bit higher than models with $\kappa_{\rm e}$ = 0,
683: whatever the value of P$_n$. This suggests that there may be some extra heating in the
684: upper atmosphere of HD 189733b, but that it is weaker than in the atmosphere of HD 209458b.
685: The $\kappa_{\rm e}$ = 0.04 cm$^2$/g model shown in the HD 189733b panels of
686: Figs. \ref{fig1} and \ref{fig4} indicates the possible magnitude of such stratospheric
687: heating, if present. Note that to fit the 16-$\mu$m point we require
688: a smaller value of $\kappa_{\rm e}$ than used to fit the IRAC channel data for
689: HD 209458b ($\kappa_{\rm e}$ = 0.1 cm$^2$/g). This may not be surprising, since,
690: as Table \ref{t1} indicates, the stellar flux at the substellar point of
691: HD 189733b is lower by more than a factor of two than the corresponding
692: number for HD209458b. Perhaps, this indicates a systematic trend for the family
693: of strongly irradiated transiting EGPs (see Tables \ref{t1} and \ref{t3}), with
694: planets with the higher values of F$_{p}$ possessing stratospheres and atmospheres with
695: pronounced inversions.
696:
697: Be that as it may, we can predict, using the logic employed in \S\ref{hd209sub},
698: that if the IRAC 1 to IRAC 2 ratio turns out to be greater than or close to one,
699: any thermal inversion in the atmosphere of HD 189733b is either not pronounced,
700: or is absent. Under these circumstances, we certainly would then expect the
701: ``brightness" temperature at IRAC 1 to be demonstrably higher than
702: that at IRAC 2 (see, e.g., Fig. \ref{fig2}). Conversely, if the IRAC 1 to IRAC 2
703: ratio is much less than one (as for HD 209458b), then a thermal inversion in the
704: atmosphere of HD 189733b would be strongly suggested. The same can be said of
705: the IRAC 4 to IRAC 3 ratio. If the IRAC channel 3 planet-to-star flux
706: ratio is higher than the Knutson et al. (2007b) point
707: at 8 $\mu$m, then a stratosphere would be indicated for HD 189733b.
708: Since the irradiation regime of HD 189733b is a bit more benign than that
709: of HD 209458b, and given the contrast in the short-wavelength behavior
710: of the IRS data for each EGP, we hypothesize that HD 189733b does
711: not boast much of a stratosphere. Note that our HD 189733b models all
712: have water in abundance and that if the unpublished IRAC 1, 2, and 3 data for HD 189733b
713: prove to decrease monotonically with wavelength shortward of IRAC 4, this
714: would be fully consistent with the presence of the water band between
715: $\sim$4 $\mu$m and $\sim$8 $\mu$m in absorption.
716:
717:
718:
719: \subsection{TrES-1}
720: \label{tres1}
721:
722: Charbonneau et al. (2005) obtained IRAC 2 ($\sim$4.5 $\mu$m)
723: and IRAC 4 ($\sim$8.0 $\mu$m) data for TrES-1 and these data were
724: analyzed by Burrows et al. (2005). A major conclusion of that paper
725: was that water is indeed seen in absorption. Our new models, depicted in
726: the bottom-left panel of Fig. \ref{fig4} with the two IRAC data points
727: superposed, reinforce this finding. As can be seen in the figure, the
728: models with different values of P$_n$ (here, all with $\kappa_{\rm e}$ = 0) can not
729: easily be distinguished using these two IRAC points. This fact emphasizes
730: the need to obtain more {\it Spitzer} photometric data to help better
731: constrain the properties of the atmosphere of TrES-1. However, it is
732: clear from the significant drop in planet/star flux ratio from IRAC 4
733: to IRAC 2 that the atmosphere of TrES-1 is qualitatively different from
734: that of HD 209458b. In particular, this behavior is a signature of a strong
735: water absorption trough. There are no signatures of a thermal inversion in
736: the atmosphere of TrES-1, or of water in emission, and the old, default models
737: with a monotonic temperature profile (see left-middle panel of Fig. \ref{fig1}) are
738: perfectly suitable. Given this, we predict that the IRAC 1 point
739: (when obtained) will be slightly higher than the IRAC 2 point.
740:
741: %%%%% Update figure for T/P of HD 189733b to include $\kappa_{\rm e} \ne 0$ model.
742:
743:
744: \subsection{HD 149026b}
745:
746:
747: As is suggested by the relative values of F$_p$ found in Table \ref{t1} for HD
748: 209458b, HD 189733b, and TrES-1 and the different thermal profiles
749: inferred for their atmospheres, there appears to be a correlation between
750: the character of a planet's atmosphere and its
751: value of F$_p$, or a related quantity (UV insolation?).
752: This possibility is intriguing, but not yet explained. In particular,
753: we have yet to identify the extra stratospheric absorber in that subset
754: of close-in EGPs ``clearly" manifesting thermal inversions. However, 1) HD 149026b's F$_p$
755: is almost twice that of HD 209458b, 2) it has one of the hottest atmospheres
756: among those listed in Table \ref{t1} (see middle-left panel
757: of Fig. \ref{fig1}), and 3) we can not fit the IRAC 4 data point obtained
758: by Harrington et al. (2007) without a strong temperature inversion.
759: The latter conclusion agrees with that of Fortney et al. (2006),
760: who predicted a mid-infrared flux for HD 149026b near the value actually
761: measured by Harrington et al. (2007) by allowing TiO/VO to reside at low
762: pressures for the hottest atmospheres (Hubeny et al. 2003). The lower-right
763: panel of Fig. \ref{fig4} depicts three models for HD 149026b, one of
764: which has $\kappa_{\rm e} = 0.64$ cm$^2$/g. This is much larger than the $\kappa_{\rm e}$
765: employed to fit HD 209458b. The stratospheric model shown has P$_n = 0.0$,
766: which minimizes the value of $\kappa_{\rm e}$ necessary to fit the lone
767: Harrington et al. (2007) data point, and it is the only model
768: among the three depicted in Fig. \ref{fig4} that does fit. Therefore, the trend
769: in ``inversion" strength" with F$_p$, seen in the sequence TrES-1, HD 189733b,
770: and HD 209458b, continues with HD 149026b. Not only do the EGP atmospheres
771: grow hotter with F$_p$ (a not unexpected result), but the importance
772: of a thermal inversion and a stratosphere in explaining the extant data
773: increases with it as well. Given the current paucity
774: of data for HD 149026b, we urge that HD 149026b be a priority target
775: so as to help discriminate the various models only partially represented on
776: the HD 149026b panel of Fig. \ref{fig4}. We predict that the pattern
777: of the four IRAC flux ratios for HD 149026b will mimic that found for
778: HD 209458b, and that its flux ratios from $\sim$10 $\mu$m to $\sim$30 $\mu$m
779: will comfortably exceed those of models without obvious thermal inversions,
780: perhaps by large margins.
781:
782:
783: \section{Light Curves$-$Comparison with Data}
784: \label{curves}
785:
786: Measuring the infrared planet-star contrast ratio as a function of orbital
787: phase, i.e., the planet's light curve, provides the best constraints
788: on the longitudinal distribution of planetary emissions. In principle,
789: phase-dependent light curves at different wavelengths can be inverted to determine the
790: ``brightness" temperature and composition distributions over the surface of the planet,
791: including its night side. Contrast ratios obtained not just at secondary eclipse
792: ($\alpha = 0^{\circ}$), but also at other phase angles, help reveal and quantify
793: zonal winds and establish their role in redistributing stellar energy
794: (i.e., P$_n$) and matter around the planet. They can help identify transitions
795: at the terminator (Guillot \& Showman 2002; Showman \& Guillot 2002),
796: shifts in the substellar hot spot (Cooper \& Showman 2005; Williams et al. 2006), asymmetries
797: in the thermal distributions (Knutson et al. 2007b), and persistent atmospheric structures.
798: In sum, light curve measurements probe both atmospheric dynamics and the
799: planet's climate and are the key to the bona fide remote sensing of exoplanets.
800:
801: Having said this, since full light curves require many more pointings and much more
802: telescope time to obtain, and mostly address the dimmer phases of a planet's orbital traverse,
803: obtaining them is much more difficult than measuring the contrast ratio at secondary eclipse.
804: As a result, to date there are only three published light curves for irradiated EGPs
805: (for $\upsilon$ And b, HD 179949b, and HD 189733b), despite numerous observational forays.
806: All of these are for only one {\it Spitzer} waveband each and none
807: covers a complete orbit. There do exist recent upper limits (e.g., Cowan et al. 2007),
808: but these are not usefully constraining and we will not address them here\footnote{However, a few of these
809: limits are listed in Table \ref{t3}.}.
810:
811: Below, we discuss the three systems for which light curves, however sparse,
812: have been obtained and try to extract physical information by comparison with our
813: light-curve models (eq. \ref{phasel}). Before we do so, we note the following. Classic light curve studies
814: are in the optical and measure geometric albedos, phase functions (Sudarsky et al. 2005),
815: and polarizations, i.e. they measure reflected stellar light. Albedos and polarizations
816: are significantly affected by the presence of clouds, and so these traditional optical
817: campaigns focus on reflection by condensates or surfaces. The planet-star flux ratios in the optical
818: range from $\sim$10$^{-10}$ for EGPs at AU distances to $\sim$10$^{-5}$$-$10$^{-6}$
819: for the close-in EGPs near $\sim$0.05 AU (Table \ref{t1}). However,
820: in the near- and mid-IR, the planet-star contrasts are around $\sim$10$^{-3}$ (see Figs.
821: \ref{fig4} and \ref{fig5}). These larger numbers are why {\it Spitzer}
822: IR measurements, rather than optical measurements, have assumed center stage
823: in the direct study of EGPs. The light seen is not reflected stellar light, but reprocessed
824: stellar flux, emitted predominantly in the near- and mid-IR (Fig. \ref{fig3}) at the lower
825: temperatures ($\sim$1000$-$2000 K) of the resulting planetary atmospheres. The ratio
826: between the optical and IR components, and thus the relative advantage of IR measurements,
827: is roughly the square of the ratio between the orbital distance and the
828: stellar radius, a number near $\sim$10$^2$ for most of the planets listed in Table \ref{t1}.
829:
830:
831: \subsection{$\upsilon$ And b}
832:
833: Harrington et al. (2006) have measured the phase variation at 24 $\mu$m of the planet-star
834: contrast for the close-in EGP $\upsilon$ And b (Butler et al. 1997). Since this planet is not transiting,
835: we know neither \mp\ nor $\sin{(i)}$, but only the combination M$_{\rm p}$$\sin{(i)}$ (= 0.69 \mj).
836: Moreover, without a transit we don't have a measurement of \rp. In fact, the models
837: used to fit the five (!) data points obtained by Harrington et al. (2006), which are
838: not anchored by absolute calibration, depend on P$_n$, $\kappa_{\rm e}$, \rp, and $\sin{(i)}$ (see eq. \ref{phasel}).
839: In addition, all interpretations hinge upon only the two extreme points in the Harrington et al. (2006) dataset.
840: Therefore, we have too many degrees of freedom to allow us to draw strong conclusions
841: concerning planetary and atmospheric parameters and must make do with limits and general
842: correlations.
843:
844: Figure \ref{fig6} portrays in eight panels comparisons of theoretical 24-$\mu$m phase curves
845: with the $\upsilon$ And b data. The left panels contain models with $\kappa_{\rm e}$ = 0,
846: and the right panels contain models with $\kappa_{\rm e}$ = 0.2 cm$^2$/g. The models in the top four panels
847: have P$_n$ = 0.0, and those in the bottom four panels have P$_n$ = 0.3. Inclinations of both 45$^{\circ}$
848: and 80$^{\circ}$ (near eclipse) are employed. On each panel, we provide models
849: with a wide range of planetary radii. Since the data have no absolute calibration,
850: we are free to move the data points up and down, as long as their relative values
851: are maintained, and we have done so in an attempt to provide on each panel the best fit
852: to the overall shape and the day/night difference. The corresponding $T/P$ profiles
853: at $\alpha = 0^{\circ}$ (day) and $\alpha = 180^{\circ}$ (night) are displayed in the lower-left
854: panel of Fig. \ref{fig1}.
855:
856: From figures like Fig. \ref{fig6}, we can extract general trends and limits. In their discovery
857: paper, the authors noted that the shift of the hot spot away from the substellar point was
858: small. They also remarked on the large difference from peak to trough ($\sim$0.002). Both
859: observations suggested that there is not much heat redistribution from the dayside
860: to the nightside and that P$_n$ is small, perhaps near zero. While at this stage this
861: conclusion can not be refuted, our models suggest that there is a broader range of
862: possible interpretations. Importantly, as we have noted in \S\ref{planetstar}, the presence
863: of a stratosphere can enhance dayside planetary fluxes in the mid-IR, and certainly near 24 $\mu$m,
864: even for modest values of P$_n$ which would otherwise decrease $D/N$ (eq. \ref{phasel}).
865: $\upsilon$ And b experiences a substellar flux, F$_p$, of $\sim$1.3$\times$10$^9$ erg cm$^{-2}$ s$^{-1}$,
866: and this is larger than that impinging upon HD 209458b. Therefore, we expect that
867: the atmosphere of $\upsilon$ And b will have a thermal inversion as well, thereby enhancing
868: the mid-IR day/night contrasts even for values of P$_n$ near 0.3\footnote{It is possible that
869: $\sin{(i)}$ is small, and, therefore, that the gravity and \mp\ are large. It is also possible that
870: a large gravity can shift the breakpoint between EGP atmospheres with and without
871: inversions. However, we suspect that F$_p$, more than gravity, is the crucial
872: parameter in determining this bifurcation.}. What is more, large values of \rp are
873: becoming commonplace, and we can not eliminate this possibility
874: for $\upsilon$ And b. A large value of \rp increases the amplitude swing
875: from day to night. Therefore, many parameter combinations can fit these data.
876:
877: We summarize the lessons of Fig. \ref{fig6} as follows. All else remaining the same,
878: a change of P$_n$ from 0.0 to 0.3 results in an increase in the \rp required of
879: $\sim$0.3$-$0.5 \rj. Substituting $\kappa_{\rm e} = 0.2$ cm$^2$/g for $\kappa_{\rm e} = 0.0$
880: decreases the \rp necessary by $\sim$0.4$-$0.5 \rj. Replacing models with
881: $i$ = 45$^{\circ}$ by those with $i$ = 80$^{\circ}$ decreases the \rp required
882: by $\sim$0.2$-$0.3 \rj. Specifically, a (P$_n$, $\kappa_{\rm e}$, $i$)
883: = (0.0, 0, 80$^{\circ}$) model has a required radius of $\sim$2.0 \rj (top-left panel
884: of Fig. \ref{fig6}), while a (P$_n$, $\kappa_{\rm e}$, $i$) = (0.0, 0.2, 80$^{\circ}$) model
885: has a required radius of $\sim$1.5 \rj (top-right panel of Fig. \ref{fig6}).
886: And while a (P$_n$, $\kappa_{\rm e}$, $i$) = (0.3, 0, 80$^{\circ}$) model
887: requires an \rp of $\sim$2.5 \rj, one with $\kappa_{\rm e} = 0.2$ cm$^2$/g requires
888: a radius of ``only" $\sim$1.8 \rj. Clearly, introducing stratospheres into the mix allows
889: P$_n$ to assume a range of non-zero values for which heat redistribution would not be considered
890: small. However, the implied planetary radius would not be small either, though
891: it would still be within the currently measured range.
892:
893: Some of this degeneracy might be broken with light-curve measurements at many wavelengths
894: and with a more rapid cadence. Furthermore, astrometric measurements of the stellar wobble
895: can provide $\sin{(i)}$ and the planet's mass, eliminating one important ambiguity.
896: {\it Spitzer} can still be used to provide the former, while the latter is well within reach
897: of ground-based astronomy. Finally, we would be remiss if we did not mention that JWST
898: will inaugurate an era of stunning photometric improvement (by a factor of
899: at least $\sim$10$^2$) over current IR platforms for the study of the light
900: curves of both transiting and non-transiting EGPs.
901:
902:
903: \subsection{HD 179949b}
904:
905:
906: Cowan et al. (2007) obtained a light curve in IRAC channel 4 ($\sim$8 $\mu$m)
907: of another non-transiting giant planet, HD 179949b (Santos, Israelian \& Mayor
908: 2004; Wittenmyer, Endl \& Cochran 2007). It has an M$_{\rm p}$$\sin{(i)}$
909: of $\sim$0.95 \mj and a value of F$_p$ of $\sim$1.32$\times$10$^9$ erg cm$^{-2}$ s$^{-1}$).
910: This makes it similar to the $\upsilon$ And system, both in its general properties
911: and in the limitations on what can be uncovered.
912:
913: The lower right-hand panel of Fig. \ref{fig1} displays the dayside and
914: nightside $T/P$ profiles of the light-curve models used to model these data.
915: Figure \ref{fig7} portrays the corresponding eight-panel figure comparing
916: our theoretical models for various combinations of P$_n$, $\kappa_{\rm e}$, $\sin{(i)}$, and \rp
917: with the eight data points of Cowan et al. (2007). Figure \ref{fig7}
918: is similar to Fig. \ref{fig6}, but the thermal inversion models are for
919: $\kappa_{\rm e}$ = 0.08 cm$^2$/g and the range of model radii are different.
920: The data for HD 179949b are important, but no less ratty than those for $\upsilon$ And b.
921: Nevertheless, the values of \rp required to fit HD 179949b are systematically lower.
922:
923: We summarize our conclusions from Fig. \ref{fig7} as follows. A
924: change of P$_n$ from 0.0 to 0.3 requires \rp to increase by
925: $\sim$0.1$-$0.2 \rj. Replacing $\kappa_{\rm e} = 0.0$ cm$^2$/g by $\kappa_{\rm e} = 0.08$
926: decreases \rp by $\sim$0.2$-$0.4 \rj. Substituting models with
927: $i$ = 45$^{\circ}$ for those with $i$ = 80$^{\circ}$ decreases \rp
928: by $\sim$0.1$-$0.2 \rj. A (P$_n$, $\kappa_{\rm e}$, $i$)
929: = (0.0, 0, 80$^{\circ}$) model that fits the data requires a radius of $\sim$1.2 \rj (top-left panel
930: of Fig. \ref{fig7}), while a (P$_n$, $\kappa_{\rm e}$, $i$) = (0.0, 0.08, 80$^{\circ}$) model
931: requires a radius of $\sim$1.0 \rj (top-right panel of Fig. \ref{fig7}).
932: Whereas a (P$_n$, $\kappa_{\rm e}$, $i$) = (0.3, 0, 80$^{\circ}$) model
933: requires an \rp of $\sim$1.6 \rj, one with $\kappa_{\rm e} = 0.08$ cm$^2$/g requires
934: a much smaller radius, $\sim$1.1 \rj. These radii are unexceptional, and in general
935: the presence of a stratosphere substantially decreases the values required to fit
936: these data. Therefore, for HD 179949b we can fit the light-curve data with a quite
937: reasonable combination of parameters, though the various degeneracies
938: still need to be broken.
939:
940: Finally, Fig. \ref{fig7} suggests that
941: the shift between the transit ephemeris and the light curve phases
942: is not large, so we don't see any obvious advection downstream of the
943: hot spot. This is consistent with the interpretation by Harrington
944: et al. (2007) of the $\upsilon$ And b phase curve (Fig. \ref{fig6}) and
945: may be a feature of EGPs with stratospheres and/or hot upper atmospheres.
946: While very tentative, this suggestion is reinforced by the observation that
947: there is a definite displacement of the hot spot of HD 189733b, which
948: seems to have a cooler upper atmosphere (Fig. \ref{fig1}). However,
949: the perception of meaningful differences in the displacements of hot spots
950: could just as easily be false and be a consequence of having better
951: data for HD 189733b. Moreover, we don't yet have a good model
952: for the origin of such differences and possible correlations with P$_n$.
953: Clearly, better sampled phase curve data would be very useful.
954:
955:
956: \subsection{HD 189733b}
957:
958: Currently, the only light curve we have for a transiting EGP was obtained in IRAC 4 at
959: $\sim$8 $\mu$m by Knutson et al. (2007b)\footnote{However, this is only the first of many anticipated.}.
960: Not only do we have HD 189733b's radius (Table \ref{t1}), but these light-curve data have absolute calibrations.
961: In addition, there is dense coverage over a bit more than half the orbit, from just
962: before secondary eclipse to just after primary transit. Knutson et al. (2007b)
963: derive the longitudinal dependence of the surface brightness
964: and find a hot spot shifted by $16\pm 6^{\circ}$ east of the substellar point, while
965: the coolest region is shifted about 30$^{\circ}$ west of the anti-stellar point.
966: Curiously, both the hot spot and the coolest spot are in the same hemisphere.
967: Nevertheless, this is the first ``map" of the surface of an exoplanet (Burrows 2007).
968: The authors also found an indication of a nonzero, but small, eccentricity with
969: $e\, {\cos {\omega}}=0.0010\pm 0.0002$, where $\omega$ is the longitude of
970: periastron, a transit radius at 8 $\mu$m of $1.137\pm 0.006\, R_{J}$ (slightly
971: smaller than the optical radius), a stellar radius of $0.757\pm 0.003\, R_{\odot}$,
972: and an inclination of $85.61\pm 0.04$ degrees.
973:
974: These data are clearly the best of their kind and we have attempted to fit
975: them with our techniques and eq. \ref{phasel}. The results are displayed
976: in Fig. \ref{fig8}. The data are plotted as black hexagons, while the models
977: are for various values of P$_n$. One model (dashed green) assumes 10$\times$solar
978: metallicity. All these models, save one, assume (P$_0$: P$_1$) = (0.1, 1.0) bars,
979: not our default pair, but this makes little difference. As indicated in \S\ref{hd189},
980: we can fit the contrast ratio in IRAC 4 at secondary eclipse rather easily, with a slight preference
981: for a small non-zero $\kappa_{\rm e}$. As Fig. \ref{fig8} suggests, a super-solar
982: metallicity might also do the trick, but the metallicity dependence is
983: rather weak. Data in other bandpasses should break the degeneracy.
984:
985: However, we can not fit the small day/night difference with any of our models. The
986: data seem to imply a severe degree of heat redistribution, one that is still not
987: captured even with our P$_n = 0.5$ model. We note in passing that models with P$_n = 0.5$
988: do not imply that the dayside and nightside should look the same, only that the integral
989: fluxes over the entire spectrum should be comparable. Since the dayside is irradiated,
990: while the nightside emits into the blackness of space, as the upper right-hand panel
991: of Fig. \ref{fig1} indicates, the $T/P$ profiles at $\alpha = 0^{\circ}$ and $\alpha = 180^{\circ}$
992: are different. This translates quite naturally into different day-night contrast ratio
993: differences for different wavelengths, even for P$_n = 0.5$.
994:
995: What we seem to be seeing in the Knutson et al. (2007b) data are atmospheric inhomogeneities,
996: thermal structures (vortices?), on the surface of HD 189733b. That the hot spot and the coolest spot
997: are in the same hemisphere, separated by only $\sim$45$^{\circ}$, suggests our symmetric
998: models are inadequate to fit this phase curve. It is of paramount importance that
999: a {\it full} light curve over all phase angles be taken in a variety of wavebands.
1000: Well-sampled data at longer wavelengths would be particularly welcome.
1001: Moreover, model phase curves need to be sophisticated enough to incorporate temporal and
1002: 3D spatial variations. The light curves and night-side heating and
1003: thermal profiles depend centrally on jet streams, winds, and general thermal redistribution.
1004: This puts a premium on developing GCMs with reasonable global dynamics coupled to
1005: realistic radiative transfer models. Such models do not yet exist for the study of EGPs.
1006:
1007:
1008: \section{Discussion and Conclusions}
1009: \label{conclusions}
1010:
1011: In this paper, we have constructed atmosphere and spectral models for all the close-in
1012: extrasolar giant planets for which direct-detection data from {\it Spitzer} have been
1013: published (except for the ``Neptune" GJ 436b). These models incorporate the effects
1014: of external stellar irradiation, detailed atmospheres, heat redistribution, and,
1015: for some, a model for stratospheric heating. Comparing the resulting suite of models with
1016: the data for these six EGPs, we have derived constraints on their atmospheric properties.
1017: We find, as did Burrows et al. (2007b), that many severely irradiated EGPs can have thermal
1018: inversions at altitude which translate into qualitative changes in 1) the planet/star
1019: contrast ratios at secondary eclipse, 2) their wavelength dependences, and
1020: 3) day-night flux contrasts during a planetary orbit. Absorption features can flip
1021: into emission features, planetary fluxes at long wavelengths can be enhanced, and
1022: the secondary-eclipse spectra in the near-IR can be altered significantly. What is
1023: more, we find a correlation between the importance of such stratospheres and
1024: the flux at the substellar point on the planet.
1025:
1026: Hubeny, Burrows, \& Sudarsky (2003) and Burrows, Sudarsky, \& Hubeny
1027: (2006) showed that strongly irradiated atmospheres can experience
1028: a solution bifurcation to an atmosphere with an inversion for
1029: which water spectral features are reversed from troughs (absorption)
1030: to peaks (emission). This possibility is supported by the good fits
1031: obtained by Burrows et al. (2007b) to the HD 209458b IRAC data (Knutson
1032: et al. 2007c) and by our models in this paper for the subset of irradiated
1033: EGPs for which the presence of stratospheres is suggested (in particular
1034: HD 149026b and, perhaps, $\upsilon$ And b). In Hubeny, Burrows, \& Sudarsky (2003)
1035: and Burrows, Sudarsky, \& Hubeny (2006), as well as in the prescient paper
1036: by Fortney et al. (2006), the absorber was gas-phase TiO/VO,
1037: which for hot atmospheres in chemical equilibrium can exist at low pressures
1038: at altitude and not just at high temperatures at depth. The upper-atmosphere
1039: absorber that is producing stratospheres for the higher values of F$_p$ might indeed
1040: be TiO/VO, but a ``cold-trap" effect can operate to deplete the upper atmosphere
1041: of TiO/VO. However, when mass loss is ongoing, as we know to be
1042: the case for HD 209458b (Vidal-Madjar et al. 2003,2004), the atmosphere
1043: is constantly being replenished and TiO/VO at some non-zero abundance
1044: remains a viable option. Such vigorous mass loss is expected for
1045: those planets with the highest values of F$_p$, and in this paper we have
1046: discovered a possible correlation between F$_p$ and the existence of thermal
1047: inversions and stratospheres. Hence, the possible mass-loss/stratosphere
1048: connection may make for a compelling scientific narrative.
1049:
1050: The tholins, polyacetylenes, or various non-equilibrium compounds discussed in the
1051: context of solar-system bodies could also be the necessary upper-atmosphere optical absorber.
1052: Given the stellar UV and integral-flux regimes experienced by strongly-irradiated EGPs,
1053: such species might be photolytically produced with sufficient abundance (Burrows et al. 2007b;
1054: Marley et al. 2007). However, to study these molecules requires a full
1055: non-equilibrium chemical network and we won't attempt this here. Clearly, what
1056: the high-altitude absorber actually is, TiO/VO or some other compounds,
1057: awaits investigation and is the primary reason we parametrized its opacity
1058: with $\kappa_{\rm e}$\footnote{However, for some of the models in Burrows et al. (2007b)
1059: we used equilibrium TiO/VO abundances and the corresponding molecular opacities
1060: (Sharp \& Burrows 2007). These models reproduced the HD 209458b IRAC data reasonably well.}.
1061:
1062: The trend with F$_p$ we have uncovered suggests, however crudely, that those
1063: EGPs with values of F$_p$ higher than HD 209458b's ($\sim$$10^{9}$ erg cm$^{-2}$ s$^{-1}$)
1064: may well have stratospheres. Table \ref{t1} provides the needed numbers. What this table suggests
1065: is that TrES-2, TrES-3, TrES-4, HAT-P-2b, HAT-P-4b, HAT-P-5b, HAT-P-6b,
1066: OGLE-TR-10b, WASP-1b, WASP-3b, XO-3b, OGLE-TR-56b, OGLE-TR-211b, and OGLE-TR-132b,
1067: in addition to HD 149026b, are strong candidates for having stratospheres, with all the consequences
1068: implied for their spectra and light curves (\S\ref{planetstar}; \S\ref{curves}). Close-in,
1069: though non-transiting, EGPs with high values of F$_p$ (such as $\tau$ Boo b, to name
1070: only one of many) are also likely to have thermal inversions. They too should manifest the
1071: spectral discriminants identified in Figs. \ref{fig4} and \ref{fig5} and the
1072: changes in the phase curves discussed in \S\ref{curves} and suggested by Figs. \ref{fig6} and
1073: \ref{fig7}. Note that by including in the above list HAT-P-2b, which has $\sim$10$\times$
1074: the mass of the average close-in EGP, we have not addressed the possible role of
1075: gravity in these systematics. Though we suspect gravity is sub-dominant when compared
1076: with F$_p$, the dependence of upper-atmosphere physics and chemistry upon gravity should
1077: prove worth exploring.
1078:
1079: It is unlikely that we will soon obtain spectral data for the OGLE planets.
1080: However, it is distinctly possible that EGPs listed in Table \ref{t1} with
1081: F$_p$s slightly lower than HD 208459b's will have ``weak" stratospheres, as we
1082: speculated may be the case for HD 189733b. This could include XO-2b, HAT-P-1b,
1083: WASP-2b, and, perhaps, XO-1b. We note that F$_p$ for TrES-1 is lower still
1084: ($\sim$0.43$\times10^{9}$ erg cm$^{-2}$ s$^{-1}$) and that this planet shows good evidence
1085: for water in absorption and no appreciable stratosphere (\S\ref{tres1} and
1086: Burrows, Hubeny, \& Sudarsky 2005). Hence, we have a hint at a breakpoint
1087: between EGPs with and without significant stratospheres and thermal inversions.
1088:
1089: Despite speculation to the contrary, our models with abundant atmospheric water
1090: are fully consistent with all the {\it Spitzer} data for all the measured
1091: EGPs, though at times the water is in emission, not absorption. This conclusion
1092: is consistent with the possible identification of water in HD 189733b by
1093: Tinetti et al. (2007) (though see Ehrenreich et al. 2007) and in HD 209458b by Barman (2007).
1094:
1095: We find that the family of close-in EGPs probably boasts a range of values of P$_n$
1096: from $\sim$0.1 to $\sim$0.4. However, our constraints on this parameter are rather weak,
1097: particularly given the possible complicating effects in some EGP atmospheres of a stratospheric absorber.
1098: Currently, the presence of such an absorber makes it easier for values of P$_n$ that are not
1099: small to explain the data, even for $\upsilon$ And b. However, the magnitude of these
1100: effects is hard to pin down with rigor, other than to say they are in evidence $-$ there remains
1101: a slight degeneracy between P$_n$ and ``$\kappa_{\rm e}$." Without a more first-principles theory
1102: concerning the chemistry, spectroscopy, and abundance of this extra absorber at low
1103: pressures, and concerning the stellar environment in which it arises, usefully constraining
1104: P$_n$ may continue to be difficult. This is the case even if TiO and VO fit the bill,
1105: since their steady-state abundances would still be an issue. Moreover, much better models for redistribution
1106: are urgently needed. Lunine \& Lorenz (2002) have speculated that planetary atmospheres
1107: with jet streams to redistribute dayside heat to the nightside adjust their wind dynamics
1108: to maximize the rate of entropy generation. Their formalism suggests a value of P$_n$ of
1109: $\sim$0.2, which is not inconsistent with any of the currently known data on
1110: EGP atmospheres.
1111:
1112: The direct-detection data we have addressed in this paper are clearly only the
1113: first of many anticipated {\it Spitzer} contrast and light-curve measurements of
1114: strongly-irradiated EGPs. MOST will continue its campaign, future dedicated space
1115: missions will be proposed, and ground-based IR telescopes may have a role.
1116: Transiting EGPs are continuing to be discovered at an impressive rate that
1117: will not soon abate, providing an expanding catalog for follow-up and
1118: characterization. JWST is in the wings to revolutionize the field, and will
1119: come on line in the middle of the next decade. This paper is meant to provide
1120: a broad theoretical context for these initiatives and an interpretive vocabulary
1121: with which to address the ongoing study of extrasolar planets in tight orbits
1122: around their illuminating stars.
1123:
1124:
1125: %%GJ 436b: P/S $\sim$ 0.00057 (half the Rayleigh-Jeans value).
1126:
1127:
1128:
1129:
1130: \acknowledgments
1131:
1132: We thank Heather Knutson, Dave Charbonneau, Bill Hubbard,
1133: Mike Cushing, and Drew Milsom for helpful discussions and
1134: guidance and Drake Deming, Alex Sozzetti, and Jamie Matthews for
1135: the use of their data in advance of publication. This study was
1136: supported in part by NASA grants NNG04GL22G, NNX07AG80G, and NNG05GG05G
1137: and through the NASA Astrobiology Institute under Cooperative
1138: Agreement No. CAN-02-OSS-02 issued through the Office of Space
1139: Science. In addition, the first author thanks the Image Processing and Analysis Center (IPAC)
1140: and the Spitzer Science Center for hosting him during the preparation of this manuscript.
1141: Model data will be available at http://zenith.as.arizona.edu/\~{}burrows/.
1142:
1143:
1144: %\appendix
1145: \begin{appendix}
1146:
1147: \section{An Improved Treatment of the Redistribution of Stellar Irradiation from the Dayside to the Nightside}
1148: \label{redist}
1149:
1150: In this Appendix, we describe a slightly more physical
1151: version of our previous treatment (Burrows, Sudarsky, \& Hubeny 2006)
1152: of the day to night side redistribution.
1153: The total incident stellar flux (expressed as the $H$-moment) at the planetary surface is
1154: %
1155: \begin{equation}
1156: H_{\rm ext} = \frac{1}{2} \left(\frac{R_\ast}{a}\right)^2 \,
1157: \frac{\sigma}{4\pi}\, T_{\rm eff}^4 = \frac{1}{2}
1158: \left(\frac{R_\ast}{a}\right)^2 \, H_\ast\, ,
1159: \end{equation}
1160: %
1161: where $R_\ast$ is the stellar radius, $T_{\rm eff}$ is the effective
1162: temperature of the stellar surface, and $a$ is the planet-star distance.
1163: The basic feature of our model is an assumption that out of this
1164: total incident stellar flux, a fraction P$_n$ contributes an additional
1165: source of energy on the nightside, and is removed from the dayside.
1166: P$_n$ is bounded between 0.0 (no redistribution) and 0.5 (``complete" redistribution)
1167: and is conceptually the same as the redistribution parameter employed in Burrows, Sudarsky,
1168: \& Hubeny (2006), but is here implemented slightly differently. We introduce
1169: %
1170: \begin{equation}
1171: H_{\rm irr} \equiv {\rm P}_n\, H_{\rm ext} =
1172: \frac{{\rm P}_n}{2}\, \left(\frac{R_\ast}{a}\right)^2 \, H_\ast\, .
1173: \end{equation}
1174: %
1175: Formally, we take P$_n > 0$ at the nightside (signifying a gain in
1176: energy), and P$_n < 0$ at the dayside (signifying a sink of energy).
1177:
1178: We define a local gain/sink of energy, $D(m)$, such that
1179: %
1180: \begin{equation}
1181: \label{dint}
1182: \int_0^\infty D(m) \, dm = H_{\rm irr}\, .
1183: \end{equation}
1184: %
1185: We assume that $D(m)$ is non-zero only between column masses
1186: $m_0$ and $m_1$ (specified through limiting pressures $P_0$ and
1187: $P_1$, which for our default set of calculations are 0.05 and 0.5, bars, respectively).
1188: We have studied the model dependence on the values of $P_0$ and $P_1$
1189: around 0.1 to 1.0 bars. The results suggest only a modest effect on
1190: the overall spectra at secondary eclipse. However, as the top-left
1191: panels of Figs. \ref{fig1} and \ref{fig4} suggest, one can optimally fit the IRAC 2 to
1192: IRAC 1 flux ratio for HD 209458b by strategically placing the redistribution band between the
1193: corresponding photospheres, and thereby cool the IRAC 1 ``$\tau = 2/3$" surface relative
1194: to the IRAC 2 ``$\tau = 2/3$" surface.
1195: While this did not motivate our default values of $P_0$ and $P_1$ (used,
1196: one notes, for all objects in this study, not just HD 209458b, and motivated by the desire
1197: to redistribute heat near the $\tau_{\rm Rosseland} = 2/3$ level), the reader should
1198: be aware that detailed fits to the secondary eclipse data depend on the choice of $P_0$ and $P_1$.
1199: Clearly, better models of heat redistribution than we have employed here are called for.
1200:
1201: With this caveat in mind, we consider two models for $D(m)$. The first
1202: one assumes a constant $D(m)$ between $m_0$ and $m_1$:
1203: %
1204: \begin{equation}
1205: \label{dmod1}
1206: D(m) = \frac{H_{\rm irr}}{m_1-m_0}\, , \qquad {\rm (model\ 1)}\, .
1207: \end{equation}
1208: %
1209: The second model, model 2, assumes a linearly decreasing $D(m)$ between $m_0$
1210: and $m_1$, in such a way that $D(m)$ reaches $0$ at $m=m_1$. This
1211: is our default model for the calculations of this paper and its
1212: functional form is:
1213: %
1214: \begin{equation}
1215: \label{dmod2}
1216: D(m) = \frac{2 H_{\rm irr}}{m_1-m_0}\, \frac{m_1-m}{m_1-m_0}\, ,
1217: \qquad {\rm (model\ 2)}\, .
1218: \end{equation}
1219: %
1220: Therefore, $D(m)$ is non-negative on the nightside and
1221: is non-positive on the dayside.
1222:
1223: The first two moments of the transport equation read
1224: %
1225: \begin{equation}
1226: \label{hmomf}
1227: \frac{dH_\nu}{dm} = \kappa_\nu (J_\nu - B_\nu)\, ,
1228: \end{equation}
1229: %
1230: and
1231: %
1232: \begin{equation}
1233: \label{kmomf}
1234: \frac{dK_\nu}{dm} = \chi_\nu H_\nu\, ,
1235: \end{equation}
1236: %
1237: where $\kappa_\nu$ is the absorption coefficient per gram, $\chi_\nu$
1238: is the total extinction coefficient (absorption + scattering), and $m$
1239: is the column mass, related to the pressure by the relation $P=mg$, where $g$ is the
1240: gravity.
1241:
1242: Since we stipulate sinks or sources of energy at certain layers, the
1243: usual radiative equilibrium (or radiative+convective) equilibrium
1244: does not apply. Instead, it is replaced by the following energy equation,
1245: which can be written in two different ways:
1246:
1247: a) using the frequency-integrated first moment of the transfer equation,
1248: %
1249: \begin{equation}
1250: \label{ener1}
1251: \int_0^\infty\kappa_\nu (J_\nu - B_\nu)\, d\nu = -D(m)\, ,
1252: \end{equation}
1253: %
1254: because the energy gained per unit mass, $D(m)$, is balanced
1255: by the net radiation loss per unit mass, given by the integral on the
1256: left-hand-side of eq. (\ref{ener1}).
1257:
1258: b) using the equation for the frequency-integrated flux,
1259: $H \equiv \int_0^\infty H_\nu d\nu$. From eqs. (\ref{hmomf}) and (\ref{ener1}), we have
1260: %
1261: %
1262: \begin{equation}
1263: \label{ener2}
1264: \frac{dH}{dm} = - D(m)\, ,
1265: \end{equation}
1266: %
1267: which we rewrite as an equation for the integrated $H$ as
1268: %
1269: \begin{equation}
1270: \label{htot}
1271: H(m) = H_{\rm eff} + \int_m^{m_d} D(m^\prime)\, dm^\prime\, ,
1272: \end{equation}
1273: %
1274: where $m_d$ is a sufficiently large column mass at which one has
1275: $D(m_d)=0$ (that is, deeper than the region of the sources/sinks), and
1276: %
1277: \begin{equation}
1278: \label{heff}
1279: H_{\rm eff} \equiv \frac{\sigma}{4\pi} T_{\rm int}^4\,
1280: \end{equation}
1281: %
1282: is the nominal total flux deep in the atmosphere, expressed through an
1283: effective temperature, $T_{\rm int}$.
1284:
1285: In view of a simple linear form of $D(m)$, the integral in eq. (\ref{htot})
1286: can be evaluated analytically. We obtain for model 1:
1287: %
1288: \begin{eqnarray}
1289: \label{htotm1}
1290: H(m) & = & H_{\rm eff}
1291: \, , \qquad m \geq m_1\nonumber \\
1292: H(m) & = & H_{\rm eff} + H_{\rm irr}\frac{m_1-m}{m_1-m_0}
1293: \, , \qquad m_0 \leq m \leq m_1\nonumber \\
1294: H(m) & = & H_{\rm eff} + H_{\rm irr} \qquad m\leq m_0\, ,
1295: \end{eqnarray}
1296: %
1297: and for model 2:
1298: %
1299: \begin{eqnarray}
1300: \label{htotm2}
1301: H(m) & = & H_{\rm eff}
1302: \, , \qquad m \geq m_1\nonumber \\
1303: H(m) & = & H_{\rm eff} + H_{\rm irr}\left(\frac{m_1-m}{m_1-m_0}\right)^2
1304: \, , \qquad m_0 \leq m \leq m_1\nonumber \\
1305: H(m) & = & H_{\rm eff} + H_{\rm irr} \qquad m\leq m_0\, .
1306: \end{eqnarray}
1307: %
1308: As suggested by Hubeny \& Lanz (1995), it is numerically advantageous
1309: to represent the energy balance equation as a linear combination of
1310: eqs. (\ref{ener1}) and (\ref{htot}), where the $H$-moment is expressed as
1311: $H_\nu = dK_\nu/dm = d(f_\nu J_\nu)/dm$, where $f_\nu$ is the Eddington
1312: factor. Using the Eddington factor enables us to consider only one
1313: radiation moment, $J_\nu$, as an unknown quantity. The Eddington factor
1314: is not taken as an unknown; instead it is held fixed in linearization
1315: and is recalculated in the formal solution step.
1316:
1317: The above equations do not consider convection, so they apply in the
1318: radiative zone. In the convection zone, eqs. (\ref{htot}), (\ref{htotm1}),
1319: and (\ref{htotm2}) remain unchanged, provided we take
1320: %
1321: \begin{equation}
1322: H_{\rm eff} = \frac{\sigma}{4\pi} T_{\rm int}^4 - \frac{F^{\rm conv}}{4\pi}\, ,
1323: \end{equation}
1324: %
1325: where $F^{\rm conv}$ is the convective flux. Equation (\ref{ener1})
1326: is modified in the presence of convection to read
1327: %
1328: \begin{equation}
1329: \label{ener1c}
1330: \int_0^\infty\kappa_\nu (J_\nu - B_\nu)\, d\nu = -D(m)
1331: -\frac{1}{4\pi}\, \frac{d F^{\rm conv}}{dm}\, .
1332: \end{equation}
1333:
1334: {\sc CoolTLUSTY} solves the energy equation specified through
1335: eqs. (\ref{heff}) - (\ref{ener1c}) numerically and self-consistently
1336: with the set of equations of radiative transfer.
1337:
1338: However, one can gain deeper physical insight by
1339: developing a simplified gray model, in which we can actually derive analytic
1340: expressions for the local temperature. More importantly, such a model
1341: allows us to understand what values of the limiting pressures for
1342: the sink region on the dayside and the corresponding optical depths
1343: are physically acceptable. We develop these expressions in
1344: Appendices \S\ref{app2} \& \S\ref{app3} below.
1345:
1346: \section{A Semi-analytic, Gray model with Redistribution}
1347: \label{app2}
1348:
1349: In {\sc CoolTLUSTY}, the exact energy balance equation is solved self-consistently
1350: with the radiation transport equation. However, it is very useful to
1351: develop a simple gray model that allows us to study the conditions
1352: under which the structural equations have a solution at all.
1353: On the nightside, there is always a solution because we are adding energy
1354: at certain layers. However, on the dayside, we remove energy
1355: at certain layers. If we require that these layers are too deep in the
1356: optically-thick part of the atmosphere, the only way energy can
1357: be removed is to create a negative temperature gradient. If,
1358: moreover, we require the region with a negative gradient (the region of the
1359: energy sink) to continue to depth, we would eventually reach a negative temperature,
1360: which is clearly unphysical.
1361:
1362: To demonstrate this and to construct an analytic atmosphere model with
1363: redistribution, we closely follow the derivation of the analytic model given in Hubeny, Burrows, \&
1364: Sudarsky (2003), generalizing it to account for departures from
1365: radiative equilibrium due to the stipulated sources/sinks of energy.
1366: We write the frequency-integrated moment equations
1367: (\ref{hmomf}) and (\ref{kmomf}) using the mean opacities:
1368: %
1369: \begin{equation}
1370: \label{hmom}
1371: \frac{dH}{dm} = \kappa_J J - \kappa_B B\, ,
1372: \end{equation}
1373: %
1374: and
1375: %
1376: \begin{equation}
1377: \label{kmom}
1378: \frac{dK}{dm} = \chi_H H\, ,
1379: \end{equation}
1380: %
1381: where $\kappa_J$, $\kappa_B$, and $\chi_H$ are the absorption-mean,
1382: Planck-mean, and flux-mean opacities, respectively, and $J,H,K$ are
1383: the frequency-integrated moments.
1384:
1385: The energy balance equation reads (neglecting convection)
1386: %
1387: \begin{equation}
1388: \kappa_J J - \kappa_B B = -D(m)\, ,
1389: \end{equation}
1390: %
1391: which is just another form of eq. (\ref{ener1}).
1392:
1393: First, we obtain the solution for the second moment $K$. We write the second-moment
1394: equation as
1395: %
1396: \begin{equation}
1397: \frac{dK(\tau)}{d\tau} = H(\tau)\, ,
1398: \end{equation}
1399: %
1400: where $d\tau \equiv \chi_H dm \approx \chi_{\rm ross} dm$ is the
1401: flux-mean optical depth, which can be approximated as the Rosseland
1402: optical depth. In order to integrate the second-moment
1403: equation analytically, we introduce the limiting optical depths
1404: $\tau_0 \equiv \tau(m_0)$ and $\tau_1 \equiv \tau(m_1)$, and
1405: adopt the following approximation for $H(\tau)$:
1406: %
1407: \begin{equation}
1408: H(\tau) = H_{\rm eff} + H_{\rm irr}\left(\frac{\tau_1-\tau}{\tau_1-\tau_0}
1409: \right)^n\, ,
1410: \qquad {\rm for}\ \tau_0\leq \tau \leq \tau_1\, ,
1411: \end{equation}
1412: %
1413: together with the exact expressions $H(\tau)=H_{\rm eff}$ for $\tau\geq \tau_1$
1414: and $H(\tau) = H_{\rm eff} + H_{\rm irr}$ for $\tau\leq \tau_0$. Here,
1415: $n=1$ for model 1, and $n=2$ for model 2. This equation is easily
1416: solved and yields:
1417: %
1418: \begin{eqnarray}
1419: K(\tau) & = & K(0) + (H_{\rm eff} + H_{\rm irr})\, \tau\, ,
1420: \qquad \tau \leq \tau_0\, \nonumber \\
1421: K(\tau) & = & K(0) + H_{\rm eff}\, \tau + H_{\rm irr}\left\{\tau_0 +
1422: \frac{\tau_1-\tau_0}{n+1}
1423: \left[1-\left(\frac{\tau_1-\tau}{\tau_1-\tau_0}\right)^{n+1}
1424: \right]\right\}\, , \qquad \tau_0 \leq \tau \leq \tau_1\, , \nonumber \\
1425: K(\tau) & = & K(0) + H_{\rm eff}\,\tau + H_{\rm irr}
1426: \left(\frac{1}{n+1}\, \tau_1 + \frac{n}{n+1}\, \tau_0 \right)\, ,
1427: \qquad \tau \geq \tau_1\, .
1428: \end{eqnarray}
1429: %
1430: With the LTE-gray model, this equation is in fact an equation for
1431: the local temperature. We invoke the Eddington approximation,
1432: $K = J/3$, and use the energy balance equation
1433: %
1434: \begin{equation}
1435: \label{bloc0}
1436: \kappa_B B = \kappa_J J + D(m)\, ,
1437: \end{equation}
1438: %
1439: and the fact that $B=(\sigma/\pi) T^4$ to derive the
1440: local temperature. We introduce the following quantities:
1441: %
1442: \begin{equation}
1443: J(0)=3K(0)=\alpha J_{\rm ext}\, ,\quad H_{\rm ext} = f_H J_{\rm ext}\, ,\quad
1444: H_{\rm irr} = {\rm P}_n H_{\rm ext}, {\rm and} \quad w \equiv H_{\rm eff}/H_{\rm ext}\, ,
1445: \end{equation}
1446: %
1447: and we write $D(\tau)$ instead of $D(m)$ in eq. (\ref{bloc0}) as
1448: an approximate expression
1449: %
1450: \begin{equation}
1451: D(\tau) = \frac{H_{\rm irr}}{\tau_1-\tau_0}\, \frac{\bar\chi}{\bar\kappa_B}
1452: \equiv \frac{H_{\rm irr}}{\tau_1-\tau_0}\, \epsilon\, ,
1453: \end{equation}
1454: %
1455: where $\bar\chi$ and $\bar\kappa_B$ are the average values of the
1456: flux-mean and the Planck-mean opacities in the interval $(\tau_0, \tau_1)$.
1457: Note that in the strict gray model, $\epsilon=1$. Note also that the
1458: ratio $w$ is given by
1459: %
1460: \begin{equation}
1461: w = (T_{\rm int}/T_\ast)^4\, 2\,(R_\ast/a)^{-2}\, ,
1462: %w = \left(\frac{T_{\rm int}}{T_\ast}\right)^4\, 2\,
1463: %\left(\frac{R_\ast}{d}\right)^{-2}\, ,
1464: \end{equation}
1465: %
1466: so that for the case of strong irradiation, $w \ll 1$.
1467: Since here we are interested in deep layers, we use the approximation
1468: $\kappa_J = \kappa_B$, (which is, however, not valid at the surface layers).
1469: Using the above defined quantities, we can express the integrated Planck
1470: function (i.e., temperature) through $H_{\rm ext}$ only and derive:
1471: %
1472: \begin{equation}
1473: \label{btot}
1474: B=\left[\beta + w \tau +3{\rm P}_n q(\tau) +\frac{\epsilon {\rm P}_n}{\tau_1 - \tau_0}
1475: \right] H_{\rm ext}\, ,
1476: \end{equation}
1477: %
1478: where we denoted $\beta=\alpha/(3f_H)$ (which is a constant of order
1479: unity), and
1480: %
1481: \begin{equation}
1482: q(\tau) \equiv \tau_0 + \frac{1}{n+1}(\tau_1-\tau_0)\left[1 -
1483: \left(\frac{\tau_1-\tau}{\tau_1-\tau_0}\right)^{n+1}\right]\, ,
1484: \end{equation}
1485: %
1486: for $\tau_0 \leq \tau \leq \tau_1$; $q(\tau)=0$ for $\tau\leq\tau_0$,
1487: and $q(\tau)=q(\tau_1)$ for $\tau\geq \tau_1$.
1488: Notice also that since $B=(\sigma/\pi) T^4$, eq.(\ref{btot}) can be
1489: understood as an equation for the local temperature.
1490:
1491:
1492: \section{Condition for the Existence of the Solution on the Dayside}
1493: \label{app3}
1494:
1495: We now investigate the existence conditions for the solution on the dayside.
1496: We first make the following approximations: $\alpha=2$, $f_H=1/2$;
1497: thus, $\beta=4/3$, $\epsilon=1$.
1498: We assume that $\tau_0 \ll \tau_1$, so we
1499: neglect it in the expression for $q(\tau)$. We also introduce
1500: the notation $p_n = -{\rm P}_n$, noting that $p_n$ is a positive quantity.
1501: Let us first take model 1, in which $n=1$. Using all the above
1502: approximations, we obtain for the region between $\tau_0$ and $\tau_1$:
1503: %
1504: \begin{equation}
1505: B(\tau)/H_{\rm ext} = \beta-\frac{p_n}{\tau_1} + w\tau - p_n \tau\left(
1506: 1-\frac{\tau}{2\tau_1}\right)\, .
1507: \end{equation}
1508: %
1509: Since $\tau_1$ is typically larger than 1, we neglect
1510: the term $p_n/\tau_1$
1511: compared to $\beta$. This is not necessary for the formal development,
1512: but it simplifies the resulting expressions. The derivative
1513: $dB/d\tau = w-p_n(1-\tau/\tau_1)$; therefore, the local minimum of $B(\tau)$
1514: is at $\tau \equiv \tau_{\rm min} = \tau_1(1-w/p_n)$.
1515: Since the most interesting case is
1516: for strong irradiation where $w \ll 1$, we see that the minimum
1517: of $B$ is close to $\tau_1$. The value of $B(\tau)$ at the local minimum is
1518: %
1519: \begin{equation}
1520: B_{\rm min} = B(\tau_{\rm min}) = \beta - \frac{\tau_1 p_n}{2}
1521: \left(1 - \frac{w}{p_n}\right)^2\, .
1522: \end{equation}
1523: %
1524: The condition for the existence of the solution is that $B_{\rm min} > 0$,
1525: and, thus,
1526: %
1527: \begin{equation}
1528: \tau_1 < \frac{2\beta}{p_n}\, \left(1-\frac{w}{p_n}\right)^2\, .
1529: \end{equation}
1530: %
1531: Taking the most interesting case of strong irradiation, we can
1532: neglect the second factor, and write simply
1533: %
1534: \begin{equation}
1535: \tau_1 < \frac{2\beta}{p_n} \approx \frac{8}{3 p_n}\, .
1536: \end{equation}
1537: %
1538: An analogous analysis for model 2 gives a similar condition,
1539: %
1540: \begin{equation}
1541: \label{result}
1542: \tau_1 < \frac{4}{p_n}\, .
1543: \end{equation}
1544: %
1545: For instance, this demonstrates that for $p_n=0.5$, the case with the maximum energy sink on
1546: the dayside and the maximum degree of redistribution, the deeper (high-pressure) limit of the sink region must
1547: be at optical depths less than 8. Eq. \ref{result} indicates that this limit is larger
1548: for smaller $p_n$ (e.g., it is 40 for $p_n=0.1$). Hence, we have derived consistency conditions
1549: for our redistribution algorithm that have physical content and in our choices
1550: for $P_0$ and $P_1$, we are careful not to exceed this condition. Our default
1551: values of $P_0$ and $P_1$, 0.05 and 0.5 bars, respectively,
1552: translate into an optical depth range of a few$\times$0.1 to $\sim$a few, within the
1553: consistency constraints for all the P$_n$s. Indeed, if we were to exceed the consistency constraints
1554: in our atmosphere calculations, they would not converge and the simulations would crash numerically.
1555:
1556: \section{The Origin of the $\lowercase{f} = 2/3$ Term}
1557: \label{app5}
1558:
1559: To derive the proper $f$ factor, we now expand upon the formalism of Appendix \S\ref{redist}.
1560: The total energy flux received by a unit area on the planetary
1561: surface at angle $\theta_0$ from the substellar
1562: point is given by
1563: %
1564: \begin{equation}
1565: F(\mu_0) = 4\pi\left(\frac{R_\ast}{a}\right)^2\,H_\ast \mu_0\, ,
1566: \end{equation}
1567: %
1568: where $\mu_0 = \cos\theta_0$ and $\theta_0$ is the angle between
1569: the normal to the planetary surface and the direction toward the
1570: star. For simplicity, we assume that the angular diameter of
1571: the star is small. Consequently, all rays coming from the star
1572: are parallel. $H_\ast = F_\ast/4\pi$ and $F_\ast$ is
1573: the radiation flux at the surface of the star.
1574: The average flux received by the planet is then
1575: %
1576: \begin{equation}
1577: \label{fav}
1578: F_{\rm av} = \int_0^1 F(\mu_0)\, d\mu_0 = \frac{1}{2}\,
1579: 4\pi H_\ast \left(\frac{R_\ast}{a}\right)^2\, ,
1580: \end{equation}
1581: %
1582: which explains the origin of the $f=1/2$ ansatz.
1583:
1584: To improve upon this, we assume the Eddington approximation, $J=3K$, and
1585: ignore convection, which allows us to write down an analytic solution
1586: of eq. (B4) for $J$:
1587: %
1588: \begin{equation}
1589: \label{jeq}
1590: J(\tau) = J_0 + 3 H \tau\, ,
1591: \end{equation}
1592: where $4\pi H$ is the interior planetary flux.
1593: %
1594: To obtain the constant $J_0 = J(0)$, we employ the formal solution
1595: of the transfer equation for the specific intensity:
1596: %
1597: \begin{equation}
1598: \label{ifor}
1599: I(0, \mu) = \int_0^\infty (\kappa_J/\kappa_B)\, J(t)\, e^{-t/\mu}\, dt/\mu \, ,
1600: \end{equation}
1601: %
1602: since the source function, $S$, is given by $S=B=(\kappa_J/\kappa_B)J$.
1603: As shown by Hubeny et el. (2003), $\kappa_j/\kappa_B$ can differ
1604: significantly from unity, but only for low values of the flux-mean
1605: optical depth, $\tau \ll 1$. Therefore, we set $\kappa_J/\kappa_B=1$, and,
1606: using eq. (\ref{jeq}), we integrate eq. (\ref{ifor}) to obtain
1607: %
1608: \begin{equation}
1609: I(0,\mu) = J_0 + 3 H \mu\, ,
1610: \end{equation}
1611: which is the well-known Eddington-Barbier relation.
1612: The mean intensity on the planetary surface is given by
1613: %
1614: \begin{equation}
1615: J(0) = (1/2)\int_0^1 I(0, \mu)\, d\mu + (1/2)\int_{-1}^0
1616: I^{\rm ext}(\mu) d\mu\, ,
1617: \end{equation}
1618: %
1619: where
1620: %
1621: \begin{equation}
1622: I^{\rm ext}(\mu) = \delta(-\mu-\mu_0) F(\mu_0)/\pi\, ,
1623: \end{equation}
1624: %
1625: because we have assumed all incident rays from the star are parallel.
1626: $\delta()$ is the Dirac $\delta$-function.
1627: Therefore,
1628: %
1629: \begin{equation}
1630: J_0 = J(0) = \frac{1}{2} J_0 + \frac{3}{2} H + \frac{F(\mu_0)}{2\pi}\, ,
1631: \end{equation}
1632: %
1633: and, consequently
1634: %
1635: \begin{equation}
1636: J_0 = 3H + \frac{F(\mu_0)}{\pi}\, .
1637: \end{equation}
1638: %
1639: The specific intensity is then given by
1640: %
1641: \begin{equation}
1642: I(\mu, \mu_0) = \frac{F(\mu_0)}{\pi} + 3 H (\mu + 1)\, ,
1643: \end{equation}
1644: %
1645: where we have given the explicit dependence of the emergent specific
1646: intensity on $\mu_0$.
1647:
1648: For close-in planets, the irradiation flux is much larger than the
1649: intrinsic flux, $4\pi H$, so we neglect the second term.
1650: %
1651: The local atmosphere characterized by angle $\mu_0$ exhibits,
1652: within the present approximations, an essentially isotropic emergent radiation
1653: pattern, independent of the local polar angle $\mu$ and dependent only on the
1654: angular distance from the substellar point, $\mu_0$.
1655: %
1656: For the total planetary flux close to secondary eclipse received
1657: by a observer at a distance $D$, we have the expression
1658: %
1659: \begin{equation}
1660: \label{fobs1}
1661: (D/{\rm R}_p)^{2} F_{\rm obs} = \int_0^1 I(\mu_0, \mu_0) \mu_0\, d\mu_0 =
1662: \frac{1}{\pi} \int_0^1 F(\mu_0) \mu_0 d\mu_0 =
1663: \frac{4}{3} \left( \frac{R_\ast}{a}\right)^2 H_\ast \, .
1664: \end{equation}
1665: %
1666: For an average atmosphere characterized by the parameter, $f$,
1667: the external flux is given by eq. (\ref{fav}), where we replace $1/2$ by
1668: $f$. We obtain
1669: %
1670: \begin{equation}
1671: I(\mu, \mu_0) = \frac{F(\mu_0)}{\pi} = 4\left(\frac{R_\ast}{a}\right)^2
1672: H_{\ast}\, f\, ,
1673: \end{equation}
1674: %
1675: and, therefore,
1676: %
1677: \begin{equation}
1678: \label{fobs2}
1679: (D/{\rm R}_p)^{2} F_{\rm obs} = \int_0^1 I(\mu_0, \mu_0) \mu_0\, d\mu_0 =
1680: 4\left(\frac{R_\ast}{a}\right)^2 H_{\ast}\, f \int_0^1\mu_0\, d\mu_0 =
1681: \frac{4f}{2}\left(\frac{R_\ast}{a}\right)^2 H_{\ast}\, .
1682: \end{equation}
1683: %
1684: In order to get agreement between eqs. (\ref{fobs1}) and
1685: (\ref{fobs2}), we have to set $f=2/3$. This is the origin of our use of this value. %%%\hfill q.e.d.
1686:
1687:
1688:
1689:
1690: \end{appendix}
1691:
1692:
1693: \begin{thebibliography}{99}
1694:
1695: \bibitem[Alonso et al. 2004]{abt04} Alonso, R. et al. 2004, \apj, 613, L153 % TrES-1
1696:
1697: %\bibitem[Anderson et al. 2008]{anderson} Anderson, D.R. et al. 2008, submitted to \mnras, arXiv:0801.1658 % WASP-5b
1698:
1699: \bibitem[Asplund, Grevesse, \& Sauval 2006]{asplund} Asplund, M.,
1700: Grevesse, N., \& Sauval, A.J. 2006, Nucl. Phys. A, 777, 1
1701:
1702: %\bibitem[Atreya et al. 2003]{atreya} Atreya, S.K., Mahaffy, P.R., Niemann,
1703: %H.B., Wong, M.H., \& Owen, T.C. 2003, Planet. Space Sci., 51, 105 % Jupiter 3 times solar
1704:
1705: %\bibitem[Atreya 2006]{atreya2} Atreya, S.K. 2006, in Proceedings of the International Planetary
1706: %Probe Workshop, IPPW-4, Pasadena, CA, June 2006. % Jupiter and Saturn
1707:
1708: \bibitem[Bakos et al. 2006]{bkp06} Bakos, G.A. et al. 2006, \apj, 650, 1160 % (astro-ph/0603291)
1709: %refined parameters for HD 189733b: 1.154 Rj
1710:
1711: \bibitem[Bakos et al. 2007a]{bnk07a} Bakos, G.A. et al. 2007a, \apj, 656, 552 % astro-ph/0609369 % HAT-P-1b
1712:
1713: \bibitem[Bakos et al. 2007b]{bakos07b} Bakos, G.A. et al. 2007b, \apj, 670, 826 % (astro-ph/arXiv:0705.0126)
1714: % HAT-P-2b; HD 147506b
1715:
1716: \bibitem[Bakos et al. 2007c]{bakos07c} Bakos, G.A. et al. 2007c, \apj, 671, L173 %(astro-ph/arXiv:0710.1841)
1717: % HAT-P-5b
1718:
1719: \bibitem[Ballester, Sing, \& Herbert 2007]{ballester} Ballester, G.E.,
1720: Sing, D.K., \& Herbert, F. 2007, Nature, 445, 511 % H Balmer in winds: 1.33 vs. 1.32 R_j
1721:
1722: %\bibitem[Baraffe et al. 2003]{baraffe2} Baraffe, I.,
1723: %Chabrier, G., Barman, T.S., Allard, F., \& Hauschildt, P.H. 2003, \aap, 402, 701
1724: % Evolutionary models for cool brown dwarfs and extrasolar giant planets. The case of HD 209458
1725:
1726: %\bibitem[Baraffe et al. 2004]{baraffe4} Baraffe, I., Selsis, F., Chabrier, G.,
1727: %Barman, T.S., Allard, F., \& Hauschildt, P.H., \& Lammer, H. 2004, \aap, 419, L13 % , astro-ph/0404101 % evaporation
1728:
1729: %\bibitem[Baraffe et al. 2005]{baraffe5} Baraffe, I., Chabrier, G., Barman, T.S., Selsis, F., Allard, F., \&
1730: %Hauschildt, P.H. 2005, \aap, 436, L47 % evaporation, neptunes, paper 2
1731:
1732: \bibitem[Barbieri et al. 2007]{barb} Barbieri, M. et al. 2007, \aap, 476, L13
1733:
1734: \bibitem[Barman, Hauschildt, \& Allard 2005]{barman} Barman, T.S., Hauschildt, P.H., \& Allard, F.
1735: 2005, \apj, 632, 1132 % Phase-Dependent Properties of Extrasolar Planet Atmospheres
1736:
1737: \bibitem[Barman 2007]{barman2007} Barman, T. 2007, \apj, 661, 191 % astro-ph/0704.1114
1738:
1739: \bibitem[Beaulieu et al. 2007]{beaulieu} Beaulieu, J.P., Carey, S., Ribas, I., \& Tinetti,
1740: G. 2007, submitted to \apj (arXiv:0711.2142)
1741:
1742: %\bibitem[Ben-Jaffel 2007]{jaffel} Ben-Jaffel, L. 2007, \apj, 671, L61 % challenging Vidal-Madjar
1743:
1744: %\bibitem[Berdyugina et al. 2007]{berdy} Berdyugina, S.V., Berdyugin, A.V., Fluri, D.M., \& Pirola, V. 2007, accepted to \apjl, arXiv:0712.0193
1745: % polarization in the B band of HD 189 - A_g > 0.14 - \Omega = 16, 196 \pm 8^\circ
1746:
1747: %\bibitem[Bodenheimer, Lin, \& Mardling 2001]{boden} Bodenheimer, P.,
1748: %Lin, D.N.C., \& Mardling, R.A. 2001, \apj, 548, 466
1749:
1750: %\bibitem[Bodenheimer, Laughlin, \& Lin 2003]{boden2} Bodenheimer, P.,
1751: %Laughlin, G., \& Lin, D.N.C. 2003, \apj, 592, 555
1752:
1753: %\bibitem[Boss 1997]{boss1997} Boss, A. 1997, Science, 276, 1836
1754:
1755: %\bibitem[Boss 2001]{boss2001} Boss, A. 2001, \apj, 563, 367
1756:
1757: \bibitem[Bouchy et al. 2004]{bps04} Bouchy, F. et al. 2004, \aap, 421, L13
1758: % OGLE-TR-113 and OGLE-TR-132
1759:
1760: \bibitem[Bouchy et al. 2005]{bum05} Bouchy, F. et al. 2005, \aap, 444, L15
1761:
1762: %\bibitem[Brown et al. 2001]{brown01} Brown, T. M., Charbonneau, D.,
1763: %Gilliland, R.L., Noyes, R.W., \& Burrows, A. 2001, \apj, 552, 699
1764:
1765: \bibitem[Burke et al. 2007]{burke} Burke, C.J. et al. 2007, accepted to \apj\ (astro-ph/arXiv:0705.0003)
1766: % XO-2b !!
1767:
1768: %\bibitem[Burkert et al. 2005]{burkert} Burkert, A., Lin, D.N.C.,
1769: %Bodenheimer, P., Jones, C., \& Yorke, H. 2005, \apj, 618, 512
1770:
1771: %\bibitem[Burrows \& Lunine 1995]{burlun} Burrows, A. \& Lunine, J.I. 1995, Nature, 378, 333
1772:
1773: %\bibitem[Burrows \etal 1997]{Burrows97} Burrows, A., Marley, M.,
1774: %Hubbard, W. B., Lunine, J. I., Guillot, T., Saumon, D., Freedman, R.,
1775: %Sudarsky, D., \& Sharp, C. 1997, \apj, 491, 856
1776:
1777: \bibitem[Burrows \& Sharp 1999]{BurrowsSharp99} Burrows, A. \&
1778: Sharp, C. M. 1999, \apj, 512, 843
1779:
1780: \bibitem[Burrows \etal 2000]{Burrows00} Burrows, A., Guillot, T.,
1781: Hubbard, W. B., Marley, M. S., Saumon, D., Lunine, J. I.,
1782: \& Sudarsky, D. 2000, \apj, 534, 97
1783:
1784: \bibitem[Burrows \etal 2001]{bur01} Burrows, A., Hubbard, W.B.,
1785: Lunine, J.I., \& Liebert, J. 2001, Rev. Mod. Phys., 73, 719
1786:
1787: %\bibitem[Burrows, Sudarsky, \& Hubbard 2003]{Burrows03} Burrows, A.,
1788: %Sudarsky, D. \& Hubbard, W.B. 2003, \apj, 594, 545
1789:
1790: \bibitem[Burrows, Sudarsky, \& Hubeny 2003b]{burrows2003b} Burrows, A., Sudarsky, D., \& Hubeny, I. 2003,
1791: published in the proceedings of the 14th Annual Astrophysics Conference in Maryland ``The Search
1792: for Other Worlds," eds. S. Holt and D. Deming, (AIP Conference Proceedings),
1793: held in College Park, MD, October 13-14, 2003, p. 143.
1794:
1795: \bibitem[Burrows et al. 2004]{bur2004f} Burrows, A., Sudarsky, D., \& Hubeny, I.
1796: 2004 \apj, 609, 407 % Wide-Separation
1797:
1798: %\bibitem[Burrows et al. 2004b]{bur2004b} Burrows, A., Hubeny, I., Hubbard, W.B.,
1799: %Sudarsky, D., \& Fortney, J.J. 2004b \apj, 610, L53 % radius of OGLE-TR-56b
1800:
1801: \bibitem[Burrows 2005]{bur05} Burrows, A. 2005, Nature, 433, 261
1802:
1803: \bibitem[Burrows et al. 2005]{bur2005} Burrows, A., Hubeny, I., \& Sudarsky, D.,
1804: 2005 \apj, 625, L135 % second1
1805:
1806: \bibitem[Burrows, Sudarsky, \& Hubeny 2006]{bur06} Burrows, A., Sudarsky, D.
1807: \& Hubeny, I. 2006, \apj, 650, 1140 % (astro-ph/0607014) % second2
1808:
1809: \bibitem[Burrows 2007]{bnandv} Burrows, A. 2007, Nature, 447, 155, 2007
1810:
1811: \bibitem[Burrows, Hubeny, Budaj, \& Hubbard 2007a]{bur07} Burrows, A.,
1812: Hubeny, I., Budaj, J., \& Hubbard, W.B. 2007a, \apj, 661, 502 % Radii
1813:
1814: \bibitem[Burrows et al. 2007b]{bur07b} Burrows, A.,
1815: Hubeny, I., Budaj, J., Knutson, H.A., \& Charbonneau, D. 2007b, \apj, 668, L171 % inversion
1816:
1817: \bibitem[Butler et al. 1997]{butler} Butler, R.P., Marcy, G.W., Williams, E.,
1818: Hauser, H., \& Shirts, P. 1997, \apj, 474, L115 % Ups And b
1819:
1820: \bibitem[Cameron et al. 2007]{cbh07} Cameron, A.C. et al. 2007, \mnras, 375, 951
1821: % WASP 1b and WASP2b, astro-ph/0609688
1822:
1823: %\bibitem[Chabrier et al. 2004]{chab2004} Chabrier, G., Barman, T.,
1824: %Baraffe, I., Allard, F., \& Hauschildt, P.H. 2004, \apj, 603, L53 % the evolution of irradiated planets, transits
1825:
1826: \bibitem[Charbonneau \etal 2000]{Charbonneau00} Charbonneau, D.,
1827: Brown, T. M., Latham, D. W., \& Mayor, M. 2000, \apj, 529, L45
1828:
1829: \bibitem[Charbonneau \etal 2002]{Charbonneau02} Charbonneau, D.,
1830: Brown, T. M., Noyes, R. W., \& Gilliland, R. L. 2002, \apj, 568, 377 % sodium
1831:
1832: \bibitem[Charbonneau \etal 2005]{char05} Charbonneau, D. \etal 2005, \apj, 626, 523
1833: % Detection of Thermal Emission from an Extrasolar Planet
1834:
1835: %\bibitem[Charbonneau et al. 2006]{charb06} Charbonneau, D. et al. 2006, \apj, 636, 445 % HD 149026b radius, etc.
1836:
1837: %\bibitem[Charbonneau et al. 2007a]{ppv} Charbonneau, D., Brown, T.M., Burrows, A., \& Laughlin, G. 2007a,
1838: %in ``Protostars and Planets V," ed. B. Reipurth and D. Jewitt, p. 701
1839: %(University of Arizona Press), astro-ph/0603376
1840:
1841: \bibitem[Charbonneau et al. 2007b]{cwe07} Charbonneau, D., Winn, J.N.,
1842: Everett, M.E., Latham, D.W., Holman, M.J.,
1843: Esquerdo, G.A., \& O'Donovan, F.T. 2007, \apj, 658, 1322 % astro-ph/0610589 % WASPs
1844: % wasp1, wasp2
1845:
1846: \bibitem[Cho \etal 2003]{Cho02} Cho, J. Y-K., Menou, K., Hansen, B. M. S.,
1847: \& Seager, S. 2003, \apj, 587, L117
1848:
1849: %\bibitem[Cody \& Sasselov 2002]{CodySasselov02} Cody, A.M. \&
1850: %Sasselov, D.D. 2002, \apj, 569, 451
1851:
1852: \bibitem[Cooper \& Showman 2005]{cooper} Cooper, C.S. \& Showman, A.P. 2005,
1853: \apj, 629, L45 % (astro-ph/0502476)
1854:
1855: \bibitem[Cowan, Agol, \& Charbonneau 2007]{cowan2007} Cowan, N.B.,
1856: Agol, E., \& Charbonneau, D. 2007, \mnras, 379, 641 % light curves for HD 179949, HD 209458b, 51 Peg b
1857:
1858: \bibitem[Deming \etal 2005]{deming05} Deming, D., Seager, S., Richardson, L.J., \& Harrington, J.,
1859: 2005, Nature, 434, 740
1860: % Infrared radiation from an extrasolar planet, HD 209458b at 24 microns
1861:
1862: %\bibitem[Deming \etal 2005b]{deming05b} Deming, D., Brown, T.M.,
1863: %Charbonneau, D., Harrington, J., \& Richardson, L.J. 2005,
1864: %\apj, 622, 1149 % Search for CO
1865:
1866: \bibitem[Deming \etal(2006)]{deming06} Deming, D., Harrington, J.,
1867: Seager, S., Richardson, L.R. 2006, \apj, 644, 560 % HD 189733b at 16 microns
1868:
1869: \bibitem[Deming \etal(2007)]{deming07} Deming, D., Harrington, J.,
1870: Laughlin, G., Seager, S., Navarro, S.B., Bowman, W.C., \& Horning, K. 2007,
1871: submitted to \apj~ Letters (arXiv:0707.2778) % GJ436b at 8 microns, primary and secondary
1872:
1873: %\bibitem[Deming \etal(2007)]{deming07b} Deming, D. et al. 2007,
1874: %submitted to \apj % New 24-micron point for HD 209
1875:
1876: \bibitem[Demory et al. 2007]{demory} Demory, B.-O. et al. 2007, \aap, 475, 1125 % GJ 436b (arXiv:0707.3809)
1877:
1878: %\bibitem[Dyudina et al. 2005]{dyudina} Dyudina, U.A., Sackett, P.D., Bayliss, D.D.R., Seager,
1879: %S., Porco, C.C., Throop, H.B., \& Dones, L. 2005, \apj, 618, 973
1880:
1881: \bibitem[Ehrenreich et al. 2007]{Ehrenreich} Ehrenreich, D., H\'{e}brard,
1882: Lecavalier des Etangs, A., Sing, D., D\'{e}sert, J.-M., Bouchy, F.,
1883: Ferlet, R., \& Vidal-Madjar, A. 2007, \apj, 668, L179 % astro-ph/arXiv0709.0576
1884:
1885: %\bibitem[Fischer \& Valenti 2005]{fischer} Fischer, D.A. \& Valenti, J. 2005, \apj, 622, 1102
1886:
1887: \bibitem[Fischer et al. 2007]{fischer07} Fischer, D.A. et al. 2007, \apj, 669, 1336
1888:
1889: %\bibitem[Flasar et al. 2005]{flasar} Flasar, M. et al. 2005, Science, 307, 1247 % Saturn methane solar times 7
1890:
1891: \bibitem[Fortney et al. 2003]{fort03} Fortney, J.J., Sudarsky, D., Hubeny, I., Cooper, C.S.,
1892: Hubbard, W.B., Burrows, A., \& Lunine, J.I. 2003, \apj, 589, 615
1893:
1894: \bibitem[Fortney et al. 2005]{fort2005} Fortney, J.J., Marley, M.S., Lodders, K.,
1895: Saumon, D., \& Freedman, R.S. 2005, \apj, 627, L69 % HD 209458b and TrES-1 secondary eclipse, 3-5 solar ...
1896:
1897: \bibitem[Fortney et al. 2006]{fort2006} Fortney, J.J., Saumon, D., Marley, M.S., Lodders, K.,
1898: \& Freedman, R.S. 2006, \apj, 642, 495 % astro-ph/0507422 % HD 149026b and HD 189733b
1899:
1900: %\bibitem[Fortney \& Marley 2007]{fm07} Fortney, J.J. \& Marley, M.S. 2007, accepted to \apj\ (arXiv:0705.2457)
1901: % "Analysis of Spitzer Mid-Infrared Spectra of Irradiated Planets: Evidence for Water Vapor?"
1902:
1903: %\bibitem[Fressin et al. 2007]{fressin} Fressin, F., Guillot, T., Morello,
1904: %V., \& Pont, F. 2007, accepted to \aap (arXiv:0704.1919) % "...: Giant Planets in the OGLE fields"
1905:
1906: \bibitem[Gillon et al. 2006]{gpm06} Gillon, M., Pont, F., Moutou, C.,
1907: Bouchy, F., Courbin, F., Sohy, S., \& Magain, P. 2006, \aap, 459, 249
1908: %(astro-ph/0606395)
1909:
1910: \bibitem[Gillon et al. 2007a]{gpm07} Gillon, M., Pont, F., Moutou, C.,
1911: Santos, N.C., Bouchy, F., Hartman, J.D., Mayor, M., Melo, C., Queloz, D.,
1912: Udry, S., \& Magain, P. 2007a, \aap, 446, 743 % (astro-ph/0702192) % new OGLE-132b
1913:
1914: \bibitem[Gillon et al. 2007b]{gpm07b} Gillon, M., Pont, F., Demory,
1915: B.-O., Mallamn, F., Mayor, M., Mazeh, T., Queloz, D.,
1916: Shporer, A., Udry, S., \& Vuissoz, C. 2007b,
1917: \aap, 472, L13 % Letters % arXiv:0705.2219, GJ 426b Neptune transit
1918:
1919: \bibitem[Gillon et al. 2007c]{gpm07c} Gillon, M., Demory,
1920: B.-O., Barman, T., Bonfils, X., Mazeh, T., Pont, F., Udry, S., Mayor, M., \& Queloz, D.,
1921: 2007c, submitted to \aap (arXiv:0707.2261) % GJ 426b Neptune transit update
1922:
1923: \bibitem[Gillon et al. 2007d]{gpm07d} Gillon, M., Triaud, A.H.M.J., Mayor, M., Queloz, D.,
1924: Udry, S., \& North, P. 2007d, arXiv:0712.2073 % HD 17156b - e = 0.6717 (+0.0028-0.0027); R_p = 0.964 (+0.016-0.027) R_Jup; M_p = 3.111
1925: %% (+0.035-0.013) M_Jup
1926:
1927: \bibitem[Grillmair et al. 2007]{grillmair} Grillmair, C.J., Charbonneau, D., Burrows, A., Armus, L.,
1928: Stauffer, J., Meadows, V., Van Cleve, J., \& Levine, D. 2007, \apj, 658, L115
1929:
1930: \bibitem[Guillot \etal 1996]{Guillot96} Guillot, T., Burrows, A.,
1931: Hubbard, W. B., Lunine, J. I., \& Saumon, D. 1996, \apj, 459, 35
1932:
1933: \bibitem[Guillot \& Showman 2002]{GuillotShowman02} Guillot, T. \&
1934: Showman, A.P. 2002, \aap, 385, 156
1935:
1936: %\bibitem[Guillot \& Saumon 2004]{gsaumon} Guillot, T. \& Saumon, D. 2004, \apj, 609, 1170
1937:
1938: \bibitem[Guillot et al. 2006]{guillot06} Guillot, T.,
1939: Santos, N.C., Pont, F., Iro, N., Melo, C., \& Ribas, I. 2006,
1940: \aap, 453, L21 % correlation between heavy element content and stellar metallicity
1941:
1942: %\bibitem[guillot 2007]{guillot07} Guillot, T. 2007, in Physica Scripta -
1943: %Physics of Planetary Systems, Nobel Symposium 135, Stockholm : Sweden (2007),
1944: %arXiv:0712.2500 % misrepresents our "Possible ...." paper.
1945:
1946: \bibitem[Harrington et al. 2006]{harrington} Harrington, J.,
1947: Hansen, B., Luszcz, S., Seager, S., Deming, D., Menou, K.,
1948: Cho, J., \& Richardson, L. 2006, Science, 314, 623
1949: %Science{\bf{xpress}}, 12 October,
1950: %2006 (http://sciencenow.sciencemag.org/cgi/content/full/2006/1013/2).
1951: % Ups And b light curve
1952:
1953: \bibitem[Harrington et al. 2007]{harrington07} Harrington, J., Luszcz, S.,
1954: Seager, S., Deming, D., \& Richardson, L.J. 2007, Nature, 447, 691 %doi:10.1038/nature05863, to be published May 24
1955: % HD 149026b, 8 microns.
1956:
1957:
1958: \bibitem[Henry \etal 2000]{Henry00} Henry, G., Marcy, G. W., Butler, R. P.,
1959: \& Vogt, S. S. 2000, \apj, 529, L41
1960:
1961: \bibitem[Holman et al. 2006]{hwl06} Holman, M.J. et al. 2006, \apj, 652, 1715 % (astro-ph/0607571) % XO-1b
1962:
1963: \bibitem[Holman et al. 2007]{hws07} Holman, M.J., Winn, J.N., Stanek, K.Z., Torres, G., Sasselov, D.D.,
1964: Allen, R.L., \& Fraser, W. 2007, \apj, 655, 1103 % (astro-ph/0506569)
1965:
1966: %\bibitem[Hubbard \etal 2001]{Hubbard01} Hubbard, W.B., Fortney, J.F.,
1967: %Lunine, J.I., Burrows, A., Sudarsky, D., \& Pinto, P.A. 2001,
1968: %\apj, 560, 413
1969:
1970: %\bibitem[Hubbard et al. 2006]{hub06} Hubbard, W.B., Hattori, M.F.,
1971: %Burrows, A., Hubeny, I., \& Sudarsky, D. 2006,
1972: %accepted to Icarus (astro-ph/0508591)
1973:
1974: %\bibitem[Hubbard 1977]{hubbard77} Hubbard, W.B. 1977, Icarus, 29, 245
1975:
1976: \bibitem[Hubeny 1988]{hub88} Hubeny, I. 1988, Comput. Phys. Commun., 52, 103
1977:
1978: \bibitem[Hubeny \& Lanz 1995]{HubenyLanz95} Hubeny, I. \& Lanz, T. 1995,
1979: \apj, 439, 875
1980:
1981: \bibitem[Hubeny, Burrows, \& Sudarsky 2003]{Hubeny03} Hubeny, I., Burrows, A., \& Sudarsky, D. 2003,
1982: \apj, 594, 1011
1983:
1984: %\bibitem[Ida \& Lin 2004]{ida} Ida, S. \& Lin, D.N.C. 2004, \apj, 616, 567
1985:
1986: %\bibitem[Iro, B\'ezard, \& Guillot 2005]{niro} Iro, N., B\'ezard, B.,
1987: %\& Guillot, T. 2005, \aap, 436, 719
1988:
1989: %\bibitem[Ingersoll 1976]{inger} Ingersoll, A.P. 1976, Icarus, 29, 245
1990:
1991: %\bibitem[Ingersoll \& Porco 1978]{ingerporco} Ingersoll, A.P. \& Porco, C.C. 1978,
1992: %Icarus, 35, 27
1993:
1994: \bibitem[Johns-Krull et al. 2007]{jk07} Johns-Krull, C.M. et al. 2007, accepted to \apj, arXiv:0712.4283
1995: %B.A.A.S. 210 \#96.05 % XO-3b
1996:
1997: \bibitem[Knutson et al. 2007a]{kcn07} Knutson, H., Charbonneau, D., Noyes, R.W.,
1998: Brown, T.M., \& Gilliland, R.L. 2007a, \apj, 655, 564 % astro-ph/0603542 % HD 209458b radius
1999:
2000: \bibitem[Knutson et al. 2007b]{kcn07b} Knutson, H., Charbonneau, D., Allen, L.E., Fortney, J.J.,
2001: Agol, E., Cowan, N.B., Showman, A.P., Cooper, C.S., \& Megeath, S.T.
2002: 2007b, Nature, 447, 183 % HD 189733b, 8-micron light curve ramp, 8-micron
2003:
2004: \bibitem[Knutson et al. 2007c]{kcn07c} Knutson, H.A., Charbonneau, D., Allen, L.E.,
2005: Torres, G., Burrows, A., \& Megeath, S.T. 2007c, accepted to \apj
2006: % HD 209458b IRAC points
2007:
2008: %\bibitem[Konacki et al. 2003a]{konacki2003a} Konacki, M., Torres, G., Jha, S., \& Sasselov, D. 2003,
2009: %Nature, 421, 507 % OGLE-TR-56b
2010:
2011: %\bibitem[Konacki et al. 2003b]{konacki2003b} Konacki, M., Torres, G., Sasselov, D., \& Jha, S. 2003,
2012: %\apj, 597, 1076 % OGLE-TR-10
2013:
2014: \bibitem[Konacki et al. 2004]{konacki04} Konacki, M., et al. 2004, \apj, 609, L37 % (astro-ph/0404541) % OGLE-TR-113
2015:
2016: \bibitem[Kov\'{a}cs et al. 2007]{kovacs} Kov\'{a}cs, G. et al. 2007, \apj, 670, L41
2017: % (astro-ph/arXiv:0710.0602) % HAT-P-4b
2018:
2019: \bibitem[Kurucz 1994]{Kurucz94} Kurucz, R. 1994, {\it Kurucz CD-ROM
2020: No. 19}, (Cambridge: Smithsonian Astrophysical Observatory)
2021:
2022: %\bibitem[Laughlin et al. 2005]{laugh} Laughlin, G., Wolf, A., Vanmunster, T., Bodenheimer, P.,
2023: %Fischer, D., Marcy, G., Butler, P., \& Vogt, S. 2005, \apj, 621, 1072 % Models with cores, TrES-1 data
2024:
2025: %\bibitem[Lecavalier~des~Etangs et al. 2004]{lecav} Lecavelier des Etangs, A. et al. 2004, \aap, 418, L1
2026:
2027: %\bibitem[Lecavalier~des~Etangs 2006]{lecav06} Lecavelier des Etangs, A. 2006,
2028: %submitted to \aap (astro-ph/0609744) % evaporation
2029:
2030: \bibitem[Lunine \& Lorenz 2002]{luninelorenz} Lunine, J.I. \& Lorenz, R.D. 2002,
2031: in the proceedings of the 33rd Annual Lunar and Planetary Science Conference, held
2032: March 11-15, 2002, Houston, Texas, abstract no. 1429 (Lunar and Planetary Institute; Houston)
2033:
2034: \bibitem[Mandushev et al. 2007]{mandushev} Mandushev, G. et al. 2007, \apj, 667, 195 % TrES-4 (arXiv:07080834)
2035:
2036: \bibitem[Marley et al. 2007]{marley07} Marley, M.S., Fortney, J., Seager, S., \& Barman, T. 2007, in Protostars
2037: and Planets, eds. V, B. Reipurth, D. Jewitt, and K. Keil (University of Arizona Press; Tucson), pp. 733-747
2038:
2039: %\bibitem[Mayor \& Queloz 1995]{mayor} Mayor, M. \& Queloz, D. 1995, Nature, 378, 355
2040:
2041: %\bibitem[McArthur et al. 2007]{macarthur} McArthur, B.E. et al. 2007,
2042: %in preparation % astrometry on Ups And system: ~20%
2043:
2044: \bibitem[McCullough et al. 2006]{msv06} McCullough, P.R.,
2045: et al. 2006, \apj, 648, 1228 % XO-1b discovery
2046:
2047: \bibitem[Melo et al. 2006]{melo} Melo, C., et al. 2006, \aap, 460, 251 % astro-ph/0609259
2048:
2049: \bibitem[Menou \etal 2003]{Menou03} Menou, K., Cho, J. Y-K., Hansen, B. M. S.,
2050: \& Seager, S. 2003, \apj, 587, L113
2051:
2052: \bibitem[Moutou et al. 2004]{mpb04} Moutou, C., Pont, F., Bouchy, F., \& Mayor, M. 2004, \aap, 424, L31
2053:
2054: \bibitem[Murray et al. 1998]{murray} Murray, N., Hansen, B., Holman, M.J., \& Tremiane, S. 1998, Science, 279, 69
2055:
2056: \bibitem[Noyes et al. 2007]{noyes} Noyes, R.W. et al. 2007, submitted to \apj\ Letters, arXiv:0710.2894 % HAT-P-6b
2057:
2058: \bibitem[O'Donovan et al. 2006]{ocm06} O'Donovan, F.T.
2059: et al. 2006, \apj, 651, L61 % astro-ph/0609335 % TrES-2
2060:
2061: \bibitem[O'Donovan et al. 2007]{ocm07} O'Donovan, F.T.
2062: et al. 2007, \apj, 663, L37 % arXiv:0705.2004 % TrES-3
2063:
2064: \bibitem[Pollacco et al. 2007]{pollacco} Pollacco, D. et al. 2007, submitted to \mnras\ (arXiv:0711.0126) % WASP-3b
2065:
2066: %\bibitem[Pollack et al. 1996]{pollack} Pollack, J.B., Hubickyj, O., Bodenheimer, P., Lissauer, J.J.,
2067: %Podolak, M., \& Greenzweig, Y. 1996, Icarus, 124, 62
2068:
2069: %\bibitem[Pont et al. 2004]{pont2004} Pont, F., Bouchy, F., Queloz, D.,
2070: %Santos, N. C., Melo, C., Mayor, M., \& Udry, S. 2004, \aap, 426, L15 % OGLE-TR-111b
2071:
2072: \bibitem[Pont et al. 2007a]{pont07a} Pont, F. et al.
2073: 2007a, \aap, 465, 1069 % (astro-ph/0610827)
2074: % OGLE-TR-56b and OGLE-TR-10b
2075:
2076: \bibitem[Pont et al. 2007b]{pont07b} Pont, F. et al.
2077: 2007b, astro-ph/arXiv:0710.5278
2078: % OGLE-TR-182b
2079:
2080: %\bibitem[Pont et al. 2007c]{pont07c} Pont, F., Knutson, H., Gilliland,
2081: %R.L., Moutou, C., \& Charbonneau, D. 2007, accepted to \mnras,
2082: %arXiv:0712.1374 % Haze and HD 189?
2083:
2084: %\bibitem[Rauscher et al. 2007]{rauscher} Rauscher, E., Menou, K., Cho,
2085: %J.Y.-K., Seager, S., \& Hansen, B.M.S. 2007, arXiv:0712.2242 % "On the
2086: %Signatures of Atmospheric Features in Thermal Phase Curves of Hot Jupiters"
2087:
2088: %\bibitem[Redfield et al. 2007]{redfield} Redfield, S., Endl, M., Cochran, W.D., \& Koesterke, L. 2007, arXiv:0712.0761
2089: % Na I in HD 189: 3 times HD209
2090:
2091: \bibitem[Richardson, Deming, \& Seager 2003]{richard} Richardson, L.J., Deming, D.,
2092: \& Seager, S. 2003, \apj, 597, 581 % HD 209458b: Strong Limits on 2.2 micron
2093:
2094: %\bibitem[Richardson et al. 2006]{richard06} Richardson, L.J., Harrington, J.,
2095: %Seager, S., \& Deming, D. 2006, \apj. 649, 1043
2096: % radius of HD 209458b at 24 microns
2097:
2098: \bibitem[Richardson et al. 2007]{richard07} Richardson, L.J., Deming, D.,
2099: Horning, K., Seager, S., \& Harrington, J. 2007, Nature, 445, 892
2100: % HD 209458b IRS data, features at 7.78 and 9.67 microns, conflicts with Swain
2101:
2102: \bibitem[Rowe et al. 2006]{rowe} Rowe, J.F., Matthews, J.M., Seager, S., Kuschnig, R., Guenther, D.B.,
2103: Moffat, A.F.J., Rucinski, S.M., Sasselov, D., Walker, G.A.H., \& Weiss, W.W. 2006,
2104: \apj, 645, 1241 % (astro-ph/0603410)% A_B < 0.375, A_g < 0.25 "An Upper Limit on the Albedo of HD 209458b"
2105:
2106: \bibitem[Rowe et al. 2007]{rowe07} Rowe, J.F., et al. 2007, submitted to \apj
2107: (arXiv:0711.4111) % MOST albedo limit in the optical: 3.8% \pm 4.5 (1-sigma)
2108:
2109: \bibitem[Santos, Israelian, \& Mayor 2004]{sim04} Santos, N.C., Israelian, G.,
2110: \& Mayor, M. 2004, \aap, 415, 1153 % HD 179949b
2111:
2112: \bibitem[Santos et al. 2006a]{spm06} Santos, N.C. et al. 2006a, \aap, 450, 825
2113: % ogle 10, 56, 111, 113, tres1
2114:
2115: \bibitem[Santos et al. 2006b]{sei06} Santos, N.C. et al. 2006b, \aap, 458, 997 % astro-ph/0606758
2116: % ogle 10, 56, 111, 113, 132, tres1
2117:
2118: %\bibitem[Sasselov (2003)]{sasselov03} Sasselov, D. 2003, \apj, 596, 1327 % OGLE-TR56b
2119:
2120: \bibitem[Sato et al. 2005]{sfh05} Sato, B. et al. 2005, \apj, 633, 465
2121:
2122: %\bibitem[Saumon, Chabrier, \& Van Horn 1995]{sc95} Saumon, D., Chabrier, G., \& Van Horn, H. 1995, \apjs, 99, 713
2123:
2124: \bibitem[Seager et al. 2005]{seager2005} Seager, S., Richardson, L.J.,
2125: Hansen, B.M.S., Menou, K., Cho, J.Y.-K., \& Deming, D. 2005, \apj, 632, 1122
2126: % "On the Dayside Thermal Emission of Hot Jupiters"
2127:
2128: \bibitem[Sharp \& Burrows 2007]{sharp07} Sharp, C.M. \& Burrows, A. 2007, \apjs, 168, 140 % (astro-ph/0607211)
2129:
2130: \bibitem[Showman \& Guillot 2002]{ShowmanGuillot02} Showman, A. P.
2131: \& Guillot, T. 2002, \aap, 385, 166
2132:
2133: \bibitem[Shporer et al. 2007]{stzm07} Shporer, A., Tamuz, O., Zucker, S., \& Mazeh,
2134: T., 2007, \mnras, 376, 1296 % (astro-ph/0610556)
2135: % wasp1
2136:
2137: \bibitem[Schneider]{enc} J. Schneider's
2138: Extrasolar Planet Encyclopaedia at http://exoplanet.eu, the Geneva Search Programme at
2139: http://exoplanets.eu, and the Carnegie/California compilation at http://exoplanets.org
2140:
2141: %\bibitem[Sozzetti et al. 2004]{sozzetti} Sozzetti, A. et al. 2004, \apj, 616,
2142: %L167 (astro-ph/0410483) % High-res. TrES-1
2143:
2144: \bibitem[Sozzetti et al. 2007]{sozzetti07} Sozzetti, A., Torres, G.,
2145: Charbonneau, D., Latham, D.W., Holman, M.J., Winn, J.N., Laird, J.B.,
2146: \& O'Donovan, F.T. 2007, \apj, 664, 1190 % (arXiv:0704.2938) % TrES-2 data
2147:
2148: \bibitem[Stempels et al. 2007]{stempels} Stempels, H.C., Collier-Cameron,
2149: A., Hebb, L., Smalley, B., \& Frandsen, S. 2007,
2150: \mnras, 379, 773 %arXiv:0705.1677 ; WASP-1b
2151:
2152: \bibitem[Sudarsky \etal 2000]{Sudarsky00} Sudarsky, D., Burrows, A.,
2153: \& Pinto, P. 2000, \apj, 538, 885
2154:
2155: %\bibitem[Sudarsky et al. 2003]{sbh03} Sudarsky, D., Burrows, A., \& Hubeny, I.
2156: %2003, \apj, 588, 1121
2157:
2158: \bibitem[Sudarsky \etal 2005]{Sudarsky05} Sudarsky, D., Burrows, A.,
2159: Hubeny, I., \& Li, A. 2005, \apj, 627, 520 % Phase functions
2160:
2161: \bibitem[Swain et al. 2007]{swain} Swain, M., Bouwman, J., Akeson, R.,
2162: Lawler, S. \& Beichman, C. 2007, submitted to \apj (arXiv:astro-ph/0702593) % HD 209458b IRS data
2163:
2164: \bibitem[Tinetti et al. 2007]{tinetti} Tinetti, G. et al. 2007, Nature, 448, 169
2165: % "Water Vapour in the atmosphere of a transiting planet," HD 189733b
2166: % (primary eclipse, IRAC 1, 3, 4, and optical)
2167:
2168: %\bibitem[Thompson \& Lauson 1972]{thompsonl} Thompson, S.L. \& Lauson,
2169: %H.S. 1972, Sandia National Laboratory Report, SC-RR-71, 0714
2170:
2171: \bibitem[Torres et al. 2005]{torres} Torres, G., Konacki, M., Sasselov,
2172: D., \& Jha, S. 2005, \apj, 619, 558 % (astro-ph/0310114) % OGLE-TR56
2173:
2174: \bibitem[Torres et al. 2007]{torres07} Torres, G. et al. 2007, \apj, 666, L121 % Letters % HAT-P-3b, large core?
2175:
2176: \bibitem[Udalski et al.]{udalski} Udalski, A. et al. 2007, submitted to \aap (arXiv:0711.3978) % OGLE-TR-211b
2177:
2178: \bibitem[Vaccaro and Van Hamme 2005]{vv05} Vaccaro, T. \& Van Hamme,
2179: W. 2005, Astrophysics and Space Science, 296, Numbers 1-4, April, 2005
2180:
2181: %\bibitem[Valenti \& Fischer 2005]{valenti2} Valenti, J.A., Fischer, D.A. 2005, \apjs, 159, 141
2182:
2183: \bibitem[Vidal-Madjar et al. 2003]{vidal} Vidal-Madjar, A., Lecavelier des Etangs, A., D\'esert, J.-M.,
2184: Ballester, G.E., Ferlet, R., H\'ebrand, G., \& Mayor, M. 2003, Nature, 422, 143
2185:
2186: \bibitem[Vidal-Madjar et al. 2004]{vidal2} Vidal-Madjar, A., D\'esert, J.-M.,
2187: Lecavelier des Etangs, A., H\'ebrard, G., Ballester, G.E., Ehrenreich, D.,
2188: Ferlet, R., McConnell, J.C., Mayor, M., Parkinson, C. D. 2004, \apj, 604, L69
2189:
2190: \bibitem[Werner \& Fanson 1995]{werner} Werner, M.W. \& Fanson, J.L. 1995, Proc. SPIE, 2475, p. 418-427 ({\it Spitzer})
2191:
2192: \bibitem[Williams et al. 2006]{will 2006} Williams, P.K.G., Charbonneau, D., Cooper,
2193: C.S., Showman, A.P., \& Fortney, J.J. 2006, \apj, 649, 1020 % , astro-ph/0601092 % "Resolving the Surfaces..."
2194:
2195: %\bibitem[Winn \& Holman 2005]{winnh} Winn, J.N. \& Holman, M.J. 2005, \apj, 628, L159
2196:
2197: \bibitem[Winn, Holman, \& Fuentes 2007]{whf07a} Winn, J.N., Holman, M.J.,
2198: \& Fuentes, C.I. 2007, \aj, 133, 11 % astro-ph/0609471 % OGLE-TR-111b
2199:
2200: %\bibitem[Wilson et al. 2008]{wilson} Wilson, D.M. et al. 2008, accepted to \apjl, arXiv:0801.1509 % WASP-4b
2201:
2202: \bibitem[Winn, Holman, \& Roussanova (2007)]{whf07.2} Winn, J.N., Holman, M.J.,
2203: \& Roussanova, A. 2007, \apj, 657, 1098 % astro-ph/0611404 % TrES-1
2204:
2205: \bibitem[Winn et al. 2007]{whf07.3} Winn, J.N., Holman, M.J., Bakos, G.A.,
2206: Pal, A., Johnson, J.A., Williams, P.K.G., Shporer, A., Mazeh, T., Fernandez, J.,
2207: Latham, D.W., \& Gillon, M. 2007, \aj, 134, 1707 % (arXiv:0707.1908)
2208: % HAT-P-1b
2209:
2210: \bibitem[Wittenmyer, Endl, \& Cochran 2007]{witten} Wittenmyer, R.A., Endl, M., \& Cochran, W.D. 2007, \apj, 654, 625
2211:
2212: \end{thebibliography}{}
2213:
2214:
2215: \clearpage
2216:
2217: \begin{table*}
2218: \small
2219: \begin{center}
2220: \caption{Transiting Planet Data\tablenotemark{1}}
2221: %\tablewidth{19.0cm}
2222: \tablewidth{17.0cm}
2223: %\begin{tabular}{lllllll}
2224: \begin{tabular}{ccccccc}
2225: \hline\hline
2226: Planet& a & Period & M$_{p}$ & R$_{p}$ &F$_{p}$& Ref. \\
2227: & (AU) & (day) & ($M_{J}$) & ($R_{J}$) &(${\rm 10^9\ erg\ cm^{-2}\ s^{-1}}$)& \\
2228: \hline
2229: OGLE-TR-56b & 0.0225 & 1.2119 & $1.29\pm0.12 $ & $1.30\pm0.05 $& 5.912 & 1,2,3,4,5 \\
2230: % OGLE-TR-56b & 0.0225 & 1.2119 & $1.29\pm0.12 $ & $1.30\pm0.05 $& 4.112 (old) & 1,2,3,4,5 \\
2231: TrES-3 & 0.0226 & 1.3062 & $1.92\pm0.23 $ & $1.295\pm0.081 $& 1.567 & 31 \\
2232: OGLE-TR-113b & 0.0229 & 1.4325 & $1.32\pm0.19 $ & $1.09\pm0.03 $& 0.739 & 1,2,4,6,7,8 \\
2233: % WASP-4b & 0.0230 & 1.3382 & $1.27^{+0.08}_{-0.10}$ & $1.45^{+0.04}_{-0.08}$ & 1.928 & 48 \\
2234: CoRoT-Exo-1b & 0.026 & 1.5 & $1.3$ & $1.65-1.78$ & $\cdots$ & $\cdots$ \\
2235: % WASP-5b & 0.0268 & 1.6284 & $1.58^{+0.13}_{-0.08}$ & $1.09^{+0.094}_{-0.058}$ & 1.895 & 49 \\
2236: % GJ 436b & 0.0285 & 2.6438 & $0.071\pm0.006 $ & $0.352\pm0.03 $& 0.030 & 32 \\
2237: % New GJ 436b & 0.0285 & 2.6438 & $0.071\pm0.006 $ & $0.374^{+0.019}_{-0.14} $& 0.030 & 32, 35(8 $\mu$m) \\
2238: % Gillon 2007c
2239: GJ 436b & 0.0285 & 2.6438 & $0.071\pm0.006 $ & $0.386\pm{0.016} $& 0.044 & 32,35,36(8 $\mu$m) \\
2240: % Deming 2007c
2241: % OGLE-TR-132b & 0.0306 & 1.6899 & $1.19\pm0.13 $ & $1.13\pm0.08 $& 4.528 & 2,7,9,26 \\
2242: OGLE-TR-132b & 0.0306 & 1.6899 & $1.14\pm0.12 $ & $1.18\pm0.07 $& 3.500 & 26 \\
2243: WASP-2b & 0.0307 & 2.1522 & $0.88\pm0.11 $ & $1.04\pm0.06 $& 0.579 & 10,11 \\
2244: HD 189733b & 0.0313 & 2.2186 & $1.15\pm0.04 $ & $1.15\pm0.03 $& 0.468 & 4,12,13 \\
2245: WASP-3b & 0.0317 & 1.8463 & $1.76^{+0.08}_{-0.14}$ & $1.31^{+0.07}_{-0.14}$ & 3.520 & 44 \\
2246: % TrES-2 & 0.0367 & 2.4706 & $1.28^{+0.09}_{-0.04}$ & $1.24^{+0.09}_{-0.06} $& 1.150 & 14 \\
2247: TrES-2 & 0.0367 & 2.4706 & $1.28^{+0.09}_{-0.04}$ & $1.24^{+0.09}_{-0.06} $& 1.150 & 14,27 \\
2248: XO-2b & 0.0369 & 2.6158 & $0.57\pm0.06 $ & $0.973^{+0.03}_{-0.008} $& 0.759 & 29 \\
2249: WASP-1b & 0.0379 & 2.5199 & $0.79^{+0.13}_{-0.06}$ & $1.40\pm0.08 $& 2.488 & 10,15,28 \\
2250: HAT-P-3b & 0.0389 & 2.8997 & $0.599\pm0.026 $ & $0.89\pm{0.046} $& 0.395 & 37 \\
2251: TrES-1 & 0.0393 & 3.0301 & $0.75\pm0.07 $ & $1.08\pm0.3 $& 0.428 & 1,2,4,16,17 \\
2252: HAT-P-5b & 0.0408 & 2.7885 & $1.06\pm{0.11}$ & $1.26\pm{0.05}$ & 1.259 & 41 \\
2253: OGLE-TR-10b & 0.0416 & 3.1013 & $0.63\pm0.14 $ & $1.26\pm0.07 $& 1.344 & 1,2,4,5,18 \\
2254: HD 149026b & 0.042 & 2.8766 & $0.36\pm0.03$ & $0.73\pm0.03$ & 2.089 & 4,19 \\
2255: HAT-P-4b & 0.0446 & 3.0565 & $0.68\pm{0.04}$ & $1.27\pm{0.05}$ & 1.833 & 39 \\
2256: HD 209458b & 0.045 & 3.5247 & $0.64\pm0.06 $ & $1.32\pm0.03 $& 1.074 & 4,20,21 \\
2257: OGLE-TR-111b & 0.047 & 4.0144 & $0.52\pm0.13 $ & $1.07\pm0.05 $& 0.248 & 1,2,4,22 \\
2258: XO-3b & 0.0477 & 3.1915 & $13.24\pm0.64 $ & $1.92\pm{0.16} $& 4.156 & 33\\ % e = 0.219\pm0.037 % R_p ingress/egress fit = 1.32 R_J, M_p = 11.71 M_J !!!
2259: TrES-4 & 0.0488 & 3.5539 & $0.84\pm0.10 $ & $1.674\pm{0.094} $& 2.306 & 38\\ % e = 0.0 ?
2260: %% TrES-4 & 0.0488 & 3.5539 & $0.84\pm0.10 $ & bigger than (A. Sozzetti) $1.674\pm{0.094} $& 2.140 & 38 (Sozzetti et al. 2007)\\ % e = 0.0 ?
2261: XO-1b & 0.0488 & 3.9415 & $0.90\pm0.07 $ & $1.18^{+0.03}_{-0.02} $& 0.485 & 23,24 \\
2262:
2263: OGLE-TR-211b & 0.051 & 3.6772 & $1.03\pm0.20 $ & $1.36^{+0.18}_{-0.09} $& 2.034 & 45 \\
2264:
2265: OGLE-TR-182b & 0.051 & 3.9791 & $1.01\pm{0.15}$ & $1.13^{+0.24}_{-0.08}$ & 0.755 & 43 \\
2266:
2267: HAT-P-6b & 0.0526 & 3.8530 & $1.06\pm{0.12}$ & $1.33\pm{0.06}$ & 1.755 & 42 \\
2268: % old HAT--P-1b & 0.0551 & 4.4653 & $0.53\pm0.04 $ & $1.36^{+0.11}_{-0.09} $& 0.681 & 25 \\
2269: HAT-P-1b & 0.0551 & 4.4653 & $0.53\pm0.04 $ & $1.203\pm{0.051} $& 0.681 & 25,34 \\
2270: HAT-P-2b & 0.0685 & 5.6334 & $8.17\pm0.72 $ & $1.18\pm{0.16} $& 1.326 & 30 \\
2271: % HD 17156b & 0.15 & 21.2 & $3.08 $ & $1.15\pm{0.11} $& 0.161 & 40 \\ % e = 0.67 !
2272:
2273: HD 17156b & 0.1594 & 21.2173 & $3.11^{+0.035}_{-0.013} $ & $0.964^{+0.016}_{-0.027} $& 0.161 & 40,46,47 \\ % e = 0.67 !
2274:
2275:
2276: %$\upsilon$ And b& 0.059 & 4.6170 & 0.69 & - & 1.295 & 111 \\ %,117,110,112,116,114,118,109,119 \\
2277: % HD 179949 b & 0.045 & 3.0925 & 0.95 & - & 1.318 & 118, 122 \\ % 118,122,123 \\
2278:
2279:
2280:
2281: \hline
2282: \end{tabular}
2283: \tablenotetext{1}{Data, plus representative references, for 29 of the known transiting EGPs
2284: with measured M$_{\rm p}$ and R$_{\rm p}$. The list is in order of increasing semi-major axis. F$_p$ is the
2285: stellar flux at the planet's substellar point, given the stellar luminosities provided in Table \ref{t2}.
2286: }
2287: \bigskip
2288: \tablerefs{
2289: (1) Santos et al. (2006a),
2290: (2) Santos et al. (2006b),
2291: (3) Vaccaro \& Van Hamme (2005),
2292: (4) Melo et al. (2006),
2293: (5) Pont et al. (2007a),
2294: (6) Gillon et al. (2006),
2295: (7) Bouchy et al. (2004),
2296: (8) Konacki et al. (2004),
2297: (9) Moutou et al. (2004),
2298: (10) Cameron et al. (2007),
2299: (11) Charbonneau et al. (2007),
2300: (12) Bouchy et al. (2005),
2301: (13) Bakos et al. (2006),
2302: (14) O'Donovan et al. (2006),
2303: (15) Shporer et al. (2007),
2304: (16) Alonso et al. (2004),
2305: (17) Winn, Holman, \& Roussanova (2007),
2306: (18) Holman et al. (2007),
2307: (19) Sato et al. (2005),
2308: (20) Santos, Israelian, \& Mayor (2004),
2309: (21) Knutson et al. (2007a),
2310: (22) Winn, Holman, \& Fuentes (2007),
2311: (23) Holman et al. (2006),
2312: (24) McCullough et al. (2006),
2313: (25) Bakos et al. (2007a),
2314: (26) Gillon et al. (2007a), % OGLE-132b
2315: (27) Sozzetti et al. (2007), % TrES-2
2316: (28) Stempels et al. (2007),
2317: (29) Burke et al. (2007),
2318: (30) Bakos et al. (2007b),
2319: (31) O'Donovan et al. (2007),
2320: (32) Gillon et al. (2007b),
2321: (33) Johns-Krull et al. (2007),
2322: (34) Winn et al. (2007), % HAT-P-1b
2323: (35) Gillon et al. (2007c), % update GJ436b
2324: (36) Deming et al. (2007), % update 2 GJ436b
2325: (37) Torres et al. (2007), % HAT-P-3b
2326: (38) Mandushev et al. (2007), % TrES-4
2327: (39) Kov\'{a}cs et al. (2007), % HAT-P-4b
2328: (40) Barbieri et al. (2007), % HD 17156b
2329: (41) Bakos et al. (2007c), % HAT-P-5b
2330: (42) Noyes et al. (2007), % HAT-P-6b
2331: (43) Pont et al. (2007b), % OGLE-182b
2332: (44) Pollacco et al. (2007), % WASP-3b
2333: (45) Udalski et al. (2007), % OGLE-TR-211b
2334: (46) Gillon et al. (2007d), % HD 17156b
2335: (47) Fischer et al. (2007) % HD 17156b
2336: %(48) Wilson et al. (2008), % WASP-4b
2337: %(49) Anderson et al. (2008) % WASP-5b
2338: }
2339: %(109) Valenti \& Fischer (2005),
2340: %(110) Butler et al. (2006),
2341: %(111) Butler et al. (1997),
2342: %(112) Naef et al. (2004),
2343: %(114) Pasinetti Fracassini et al. (2001),
2344: %(116) Butler et al. (1999),
2345: %(117) Wittenmyer, Endl \& Cochran (2007),
2346: %(118) Santos, Israelian \& Mayor (2004)
2347: %(119) Harrington et al. (2006),
2348: %(122) Wittenmyer, Endl \& Cochran (2007) % HD 179949b
2349: %(123) Fischer \& Valenti (2005)
2350: \label{t1}
2351: \end{center}
2352: \end{table*}
2353:
2354: \clearpage
2355:
2356: \begin{table*}
2357: \small
2358: \begin{center}
2359: \caption{Data on Parent Stars\tablenotemark{1}}
2360: \tablewidth{17.0cm}
2361: \begin{tabular}{llllllllll}
2362: %\begin{tabular}{cccccccccc}
2363: %\tableline\tableline
2364: \hline\hline
2365: Star & Sp.T. & $R_*$ &$T_{\rm eff}$&$\log g$&[Fe/H]$_*$ & M$_*$ & L$_*$ & Age & Dist \\
2366: & & ($R_{\odot}$) & (K) & (cgs) & &($M_{\odot}$)&($L_{\odot}$) &(Gyr)&(pc) \\
2367: \hline
2368: OGLE-TR-56 & G & $1.32\pm0.06 $& 6119 & 4.21 & 0.25 & 1.04 & 2.20 &$2.5^{+1.5}_{-1.0} $& 1600 \\
2369: TrES-3 & G3V & $0.80\pm0.05 $& 5650 & 4.60 & $-0.19$ & 0.90 & 0.59 &$\cdots$& $\cdots$ \\
2370: OGLE-TR-113 & K & $0.77\pm0.02 $& 4804 & 4.52 & 0.15 & 0.78 & 0.29 &$5.35\pm{4.65} $& 550 \\
2371: % WASP-4 & G7V & $0.953^{+0.026}_{-0.046} $& 5500 & 4.3 & 0.0 & 0.90 & 0.750 & $\cdots$ & 300 \\
2372: CoRoT-Exo-1 & G & $1.2\pm0.2$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ &$>460$ \\
2373: % WASP-5 & G4V & $1.026^{+0.073}_{-0.044} $& 5700 & 4.3 & 0.0 & 0.972 & 1.00 & $\cdots$ & 300 \\
2374: GJ 436 & M2.5V & $0.44\pm0.04 $& 3500 & 4.5 & 0.0 & 0.44 & 0.026 &$> 3$ & 10.2 \\
2375: % New GJ 436 & M2.5V & $0.463^{+0.022}_{-0.17} $& 3500 & 4.5 & 0.0 & 0.44 & 0.026 &$> 3$ & 10.2 \\
2376: % Gillon 2007c
2377: % New 2 GJ 436 & M2.5V & $0.47\pm{0.02} $& 3500 & 4.5 & 0.0 & 0.47 & 0.026 &$> 3$ & 10.2 \\
2378: % Deming et al. 2007
2379:
2380: % OGLE-TR-132 & F & $1.43\pm0.10 $& 6411 & 4.86 & 0.43 & 1.35 & 3.12 &$1.25\pm{0.75} $& 2200 \\
2381: OGLE-TR-132 & F & $1.34\pm0.08 $& 6210 & 4.51 & 0.37 & 1.26 & 2.41 &$1.25\pm{0.75} $& 2200 \\ % Gillon et al. 2007
2382: WASP-2 & K1V & $0.81\pm0.03 $& 5200 & 4.50 & $\cdots$ & 0.79 & 0.44 & $\cdots$ & $\cdots$ \\
2383: HD 189733 & K1.5 & $0.76\pm0.02 $& 5050 & 4.53 & $-0.03$ & 0.82 & 0.34 &$5.25\pm{4.75} $& 19.3 \\
2384: WASP-3 & F7V & $1.31^{+0.05}_{-0.12} $& 6400 & 4.30 & $\cdots$ & 1.24 & 2.60 & $\cdots$ & 223 \\
2385: % TrES-2 & G0V & $1.00^{+0.06}_{-0.04} $& 5960 & 4.40 & $-0.15$ &1.08& 1.14 &$7.2^{+1.8}_{-7.1} $& $\cdots$ \\
2386: TrES-2 & G0V & $1.00^{+0.06}_{-0.04} $& 5960 & 4.40 & $-0.15$ &1.08& 1.14 &$4.9^{+2.9}_{-2.0} $& $\cdots$ \\
2387: XO-2 & K0V & $0.964^{+0.02}_{-0.009}$& 5400 & 4.62 & 0.45 & 0.98 & 0.76 &$5.0^{+1.0}_{-0.5}$ & 149 \\
2388: WASP-1 & F7V & $1.45\pm0.03 $& 6110 & 4.28 & 0.26 & 1.3 & 2.67 & $2.0\pm{1.0} $ & $\cdots$ \\
2389: HAT-P-3 & KV & $0.824^{+0.036}_{-0.062} $& 5185 & 4.61 & 0.27 & 0.94 & 0.44 & $0.4^{+6.5}_{-0.3}$ & 140 \\
2390: TrES-1 & K0V & $0.81\pm0.02 $& 5226 & 4.40 & 0.06 & 0.88 & 0.49 &$4.0\pm{2.0} $& 143 \\
2391: HAT-P-5 & G0 & $1.17\pm{0.05}$ & 5960 & 4.37 & 0.24 & 1.16 & 1.54 & $2.6\pm{1.8}$ & 340 \\
2392: OGLE-TR-10 & G & $1.16\pm0.06 $& 6075 & 4.54 & 0.28 & 1.02 & 1.65 &$2.0\pm1.0 $& 1300 \\
2393: HD 149026 & G0 IV & $1.45\pm0.10 $& 6147 & 4.26 & 0.36 & 1.3 & 2.71 &$2.0\pm{0.8} $& 78.9 \\
2394: HAT-P-4 & G0 & $1.59\pm{0.07}$ & 5860 & 4.14 & 0.24 & 1.26 & 2.68 & $4.2^{+2.6}_{-0.6}$ & 310 \\
2395: HD 209458 & G0 V & $1.13\pm0.02 $& 6117 & 4.48 & 0.02 & 1.10& 1.60 &$5.5\pm{1.5} $& 47 \\
2396: OGLE-TR-111 & G/K & $0.83\pm0.03 $& 5044 & 4.51 & 0.19 & 0.81 & 0.40 &$5.55\pm{4.45} $& 1000 \\
2397: XO-3 & F5V & $2.13\pm{0.21}$ & 6429 & 3.95 & $-0.18$ & 1.41 & 6.96 &$2.8\pm{0.15}$ & 260 \\ % e = 0.219\pm0.037
2398: TrES-4 & F & $1.74\pm{0.09}$ & 6200 & 4.05 & 0.14 & 1.22 & 4.04 &$4.7\pm{2.0}$ & 440 \\ % e = 0.0
2399: % TrES-4 & F & $1.74\pm{0.09}$ & 6100 & 4.05 & 0.0 & 1.22 & 3.74 &$4.7\pm{2.0}$ & 440 \\ % e = 0.0
2400: XO-1 & G2V & $0.93^{+0.02}_{-0.01}$& 5750 & 4.53 & 0.015 & 1.00 & 0.85 &$4.6\pm{2.3} $& 200 \\
2401:
2402: OGLE-TR-211 & F7V & $1.64^{+0.21}_{-0.07}$& 6325 & 4.22 & 0.11 & 1.33 & 3.88 &$\cdots$ & $\cdots$ \\
2403:
2404: OGLE-TR-182 & G0V & $1.14^{+0.23}_{-0.06}$& 5924 & 4.47 & 0.37 & 1.14 & 1.44 &$\cdots$ & $\cdots$ \\
2405:
2406: HAT-P-6 & F8 & $1.46\pm{0.06}$ & 6570 & 4.22 & $-0.13$ & 1.29 & 3.57 &$2.3^{+0.5}_{-0.7}$ & 260 \\
2407: HAT-P-1 & G0V & $1.15^{+0.10}_{-0.07} $& 5975 & 4.45 & 0.13 & 1.12 & 1.52 &$3.6\pm{1.0} $& 139 \\
2408: % HAT-P-1 & G0V & $1.115\pm{0.043} $& 5975 & 4.45 & 0.13 & 1.12 & 1.52 &$3.6\pm{1.0} $& 139 \\
2409: HAT-P-2 & F8V & $1.80\pm0.25 $& 6290 & 4.22 & 0.12 & 1.35 & 4.58 &$2.7^{+1.4}_{-0.6} $& 135 \\
2410: % HD 17156 & G0V & $1.47\pm0.13 $& 6079 & 4.18 & 0.24 & 1.20 & 2.66 &$5.7\pm{1.3}$& 78.24 \\
2411: HD 17156 & G0V & $1.354^{+0.012}_{-0.037} $& 6079 & 4.29 & 0.24 & 1.20 & 2.66 &$4.7^{1.3}_{1.9}$& 78.24 \\
2412:
2413: %$\upsilon$ And b& F8V & 1.57 & 6213 & 4.253& 0.153& 1.3 & 3.31 & &13.47 \\
2414: %HD 179949 b & F8V & 1.19 & 6260 & 4.43 & 0.22 & 1.28 & 1.96 & &27.05 \\
2415:
2416: \hline
2417: \end{tabular}
2418: \tablenotetext{1}{A compilation of the physical parameters derived for the parents of
2419: 29 of the known transiting EGPs. The error bars have been rounded from those found
2420: in the literature. The ages, the least well-known quantities, should be taken
2421: with caution. The stellar metallicities are given without
2422: error bars, which are assumed to be large. Due to their great distances
2423: (rightmost column), the stellar types of the OGLE objects are not well constrained.
2424: Refer to Table \ref{t1} for the corresponding references.}
2425: \label{t2}
2426: \end{center}
2427: \end{table*}
2428:
2429: \clearpage
2430: \begin{deluxetable}{lllllll}
2431: \tabletypesize{\tiny}
2432: \tablewidth{0pt}
2433: \tablecaption{Spectral, Photometric, and Composition Measurements of Extrasolar Giant Planets\label{t3}}
2434: \tablehead{
2435: \colhead{Planet} &
2436: \colhead{$\lambda$} &
2437: \colhead{Telescope/} &
2438: \colhead{${\rm F}_{p}/{\rm F}_{\star}$} &
2439: \colhead{$\Delta {\rm F}_{p}/{\rm F}_{\star}$} &
2440: \colhead{Comments} &
2441: \colhead{Reference} \\
2442: \colhead{ } &
2443: \colhead{$\mu$m} &
2444: \colhead{Instrument} &
2445: \colhead{Second. Eclipse} &
2446: \colhead{ } &
2447: \colhead{ } &
2448: \colhead{ }
2449: }
2450: \startdata
2451: HD 189733 b & 7.5-14.7 & Spitzer/IRS & Spectrum & & No $H_{2}O?,\,CH_{4}?$ & Grillmair et al. (2007)\tablenotemark{1} \\
2452: & 8.0 & Spitzer/IRAC4 & \bf 0.003392(55) & \bf 0.0012(2) & Light curve & Knutson et al. (2007b)\tablenotemark{2} \\
2453: & 16.0 & Spitzer/IRS & \bf 0.00551(30) & & Peak-up mode & Deming et al. (2006) \\
2454: & 3.6 & Spitzer/IRAC1 & 0.02356(020) & & Primary Transit Depth $-$ & Beaulieu et al. (2007) \\
2455: & 5.8 & Spitzer/IRAC3 & 0.02436(020) & & (Tinetti et al. 2007) & Beaulieu et al. (2007) \\
2456: & 8.0 & Spitzer/IRAC4 & 0.0239(02) & & " & Knutson et al. (2007b)\\
2457: TrES-1 & 4.5 & Spitzer/IRAC2 & \bf 0.00066(13) & & $H_2O$ identified $-$ & Charbonneau et al. (2005) \\
2458: & 8.0 & Spitzer/IRAC4 & \bf 0.00225(36) & & (Burrows et al. 2005) & Charbonneau et al. (2005) \\
2459: HD 209458 b &Ly $\alpha$& HST/STIS & & & HI, R$_{p}=4.3\,R_{J}$ & Vidal-Madjar et al. (2003)\tablenotemark{4}\\
2460: & 0.12-0.17& HST/STIS & & & C, O & Vidal-Madjar et al. (2004)\tablenotemark{5}\\
2461: & 0.3-0.5 & HST/STIS & & & R$_P=1.3300(6)\,R_{J}$ & Ballester et al. (2007)\tablenotemark{6} \\
2462: & & & & & at Balmer cont.? & \\
2463: & 0.58-0.64& HST/STIS & & & Na D (transit) & Charbonneau et al. (2002)\tablenotemark{7} \\
2464: & $\sim$0.94 & HST/STIS & & & $H_2O$ identified (transit) & Barman 2007 \\
2465: & 0.4-0.7 & MOST & $< 1.34\times 10^{-4}~(3\sigma)$ & &$A_{g}<0.68~(3\sigma)$ & Rowe et al. (2006)\tablenotemark{8} \\
2466: & & & $< 4.88\times 10^{-5}~(1\sigma)$ & &$A_{g}<0.25~(1\sigma)$ & Rowe et al. (2006) \\
2467: & 0.4-0.7 & MOST & $< 3.9\times 10^{-5}~(1\sigma)$ & &$A_{g}\sim 4.0\pm 4.0$\%& Rowe et al. (2007) \\
2468: & 2.2 & IRTF/SpeX & $< 0.0003~(1\sigma) $ & & & Richardson et al. (2003) \\
2469: & 3.6 & Spitzer/IRAC1 & & $< 0.015~(2\sigma)$ & Light curve & Cowan et al. (2007) \\
2470: & 8.0 & Spitzer/IRAC4 & & $< 0.0015~(2\sigma)$ & Light curve & Cowan et al. (2007) \\
2471: & & & & &$P_{n}\ \sgreat\ 0.32\,(1 \sigma)$& \\
2472: & 3.6 & Spitzer/IRAC1 & \bf 0.00094(9) & & Temperature Inversion $-$ & Knutson et al. (2007c)\\
2473: & 4.5 & Spitzer/IRAC2 & \bf 0.00213(15) & & $H_{2}O$ in emission $-$ & Knutson et al. (2007c)\\
2474: & 5.8 & Spitzer/IRAC3 & \bf 0.00301(43) & & (Burrows et al. (2007b) & Knutson et al. (2007c)\\
2475: & 8.0 & Spitzer/IRAC4 & \bf 0.00240(26) & & " & Knutson et al. (2007c)\\
2476: & 7.5-13.2 & Spitzer/IRS & Spectrum & & Features at & Richardson et al. (2007)\tablenotemark{9} \\
2477: & & & & & 7.78, 9.65 $\mu$m ? & \\
2478: & & & & & No $H_{2}O$, $CH_4$? & \\
2479: & 8.2-13.2 & Spitzer/IRS & Spectrum & &$F_{p}=0.40\pm 0.19$ & Swain et al. (2007)\tablenotemark{10} \\
2480: & & & & & mJy at 12 $\mu$m & \\
2481: & 24.0 & Spitzer/MIPS & \bf 0.00260(46) & & & Deming et al. (2005) \\
2482: & 24.0 & Spitzer/MIPS & \bf 0.0033(3) (?) & & & Deming (private communication) \\ % New Author?
2483: HD 149026 b & 8.0 & Spitzer/IRAC4 & \bf 0.00084(11) & & & Harrington et al. (2007) \\
2484: $\upsilon$ And b & 24.0 & Spitzer/MIPS & No transit & \bf 0.0029(7) & Light curve & Harrington et al. (2006) \\
2485: HD 179949 b & 3.6 & Spitzer/IRAC1 & No transit & $< 0.019~(2\sigma)$ & Light curve & Cowan et al. (2007) \\
2486: & 8.0 & Spitzer/IRAC4 & No transit & \bf 0.00141(33) & Light curve & Cowan et al. (2007)\tablenotemark{3} \\
2487: & & & & &$P_{n}\ \sles\ 0.30\,(1 \sigma)$& \\
2488: GJ 436b & 8.0 & Spitzer/IRAC4 & \bf 0.00057(8) & & & Deming et al. (2007) \\
2489: & & & & & & Demory et al. (2007) \\
2490: 51 Peg b & 4.5 & Spitzer/IRAC2 & No transit & $< 0.017~(2\sigma)$ & Light curve & Cowan et al. (2007) \\
2491: & 8.0 & Spitzer/IRAC4 & No transit & $< 0.0007~(2\sigma)$ & Light curve & Cowan et al. (2007) \\
2492: OGLE-113 b & 2.2 & NTT/SOFI & 0.0017(5)~$(3\sigma)$? & & & Snellen \& Covino (2007) \\
2493: \enddata
2494: \tablecomments{Notes $-$ This table contains the following columns:
2495: the name of the planet, spectral region of the observations,
2496: the telescope with the instrument used for the observations,
2497: the planet to star flux ratio during the secondary eclipse (${\rm F}_{p}/{\rm F}_{\star}$),
2498: the amplitude or peak to trough variations of the light curve ($\Delta {\rm F}_{p}/{\rm F}_{\star}$),
2499: brief comments, and the reference. Numbers in {\bf bold} are measurements (not upper limits)
2500: at secondary eclipse or actual light curve ($\Delta {\rm F}_{p}/{\rm F}_{\star}$) measurements.
2501: Models for all these bolded data (except the 8-$\mu$m point for GJ 436b at secondary
2502: eclipse) are presented in this paper. Numbers in parentheses are errors in the
2503: last digits. \tablenotetext{1} Grillmair et al. (2007) argue that a flat spectrum is caused by
2504: the lack of significant water or methane absorption. However, this is not
2505: consistent with the conclusions of Tinetti et al. (2007), using the primary
2506: eclipse data of Beaulieu et al. (2007) and Knutson et al. (2007b);
2507: \tablenotetext{2} Knutson et al. (2007b) also derived the longitudinal dependence of the surface brightness
2508: and found a hot spot shifted by $16\pm 6$ degrees east of the substellar point while
2509: the coolest region was shifted about 30 degrees west of the anti-stellar point.
2510: They also found an indication of nonzero eccentricity with $e{\cos {\omega}}=0.0010\pm 0.0002$,
2511: a transit radius at 8 $\mu$m of $1.137\pm 0.006\, R_{J}$, a stellar radius of
2512: $0.757\pm 0.003\, R_{\odot}$ and an inclination of $85.61\pm 0.04$ degrees,
2513: where $\omega$ is the longitude of periastron;
2514: \tablenotetext{3} Cowan et al. (2007) constrain P$_{n}$ for HD 209458b and HD 179949b to be
2515: P$_{n}>0.32~(1\sigma)$ and P$_{n}<0.30~(1\sigma)$, respectively;
2516: P$_{n}$ is the fraction of the total energy incident on the dayside of the planet
2517: which is transfered to and radiated out on the night side of the planet;
2518: \tablenotetext{4} Vidal-Madjar et al. (2003) detected atomic hydrogen in the planet's atmosphere
2519: with a transit absorption depth of $15\pm 4 (1\sigma)$\%,
2520: and evaporation at the rate of $\ge 10^{10}g\,s^{-1}$;
2521: \tablenotetext{5} Vidal-Madjar et al. (2004) detected oxygen and carbon in the planet's atmosphere;
2522: \tablenotetext{6} Ballester et al. (2007) claim to have identified HI absorption in the Balmer continuum,
2523: but Barman (2007) challenges this interpretation; \tablenotetext{7} Charbonneau et al.
2524: (2002) found that the transit depth at the Na D feature is deeper by
2525: $2.32\pm 0.57\times 10^{-4}$ than in the continuum,
2526: which is interpreted as a detection of Na in the planet's atmosphere;
2527: \tablenotetext{8} Rowe et al. (2006) constrained $A_{g}<0.25~(1\sigma)$, or $A_{g}<0.68~(3\sigma)$,
2528: but Rowe et al. (2007) constrained $A_{g}$ to be below 8\% to 1$-\sigma$. $A_{g}$ is
2529: the geometric albedo in the optical;
2530: \tablenotetext{9} Richardson et al. (2007) claim to have detected a broad emission peak centered near 9.65 $\mu$m
2531: which they attribute to the emission by silicate clouds, and a narrow unidentified
2532: emission feature at 7.78 $\mu$m. They say that models with water absorption fit the data poorly.
2533: However, Burrows et al. (2007b) conclude that water is in fact seen in emission, not absorption;
2534: \tablenotetext{10} Swain et al. (2007) determined the planet flux at 12 $\mu$m to be $0.40\pm 0.19$mJy
2535: and the normalized secondary eclipse depth to be $0.0046\pm 0.0006$. They are updating
2536: their absolute calibrations.}
2537: \end{deluxetable}
2538:
2539: \clearpage
2540: \thispagestyle{empty}
2541: \setlength{\voffset}{-10mm}
2542:
2543: % figure 1
2544: \begin{figure}
2545: \centerline{
2546: \includegraphics[width=6.cm,angle=-90,clip=]{f1a.eps}
2547: \includegraphics[width=6.cm,angle=-90,clip=]{f1b.eps}}
2548: \centerline{
2549: \includegraphics[width=6.cm,angle=-90,clip=]{f1c.eps}
2550: \includegraphics[width=6.cm,angle=-90,clip=]{f1d.eps}}
2551: \centerline{
2552: \includegraphics[width=6.cm,angle=-90,clip=]{f1e.eps}
2553: \includegraphics[width=6.cm,angle=-90,clip=]{f1f.eps}}
2554: \caption{
2555: Temperature-pressure profiles for the six close-in planets
2556: studied in this paper. Dayside profiles incorporate the external substellar
2557: irradiation/flux given in Table \ref{t1} and an internal
2558: flux for the planet corresponding to the temperature of 75 K at
2559: the lower boundary. A sink of energy corresponding to the particular
2560: value of P$_{n}$ was introduced between pressures of 0.05 and 0.5 bars.
2561: The nightside is calculated without irradiation, assuming
2562: an energy source corresponding to the same value of P$_{n}$
2563: at the same pressures employed for the dayside sink.
2564: Entropies at the bases of the convection zones on the day-
2565: and night-sides were approximately matched by adjusting the internal
2566: planetary flux at the bottom of the nightside model at the same gravity.
2567: Models with an extra upper-atmosphere absorber in the optical %%%%(\S\ref{app4})
2568: are included for HD 209458b, HD 189733b, HD 149026b, $\upsilon$ And b, and HD 179949b.
2569: See text in \S\ref{tp} for a discussion of these panels.
2570: }
2571: \label{fig1}
2572: \end{figure}
2573: \clearpage
2574: \setlength{\voffset}{0mm}
2575:
2576: % figure 2
2577: \begin{figure}
2578: \centerline{
2579: \includegraphics[width=14.cm,angle=-90,clip=]{f2.eps}
2580: }
2581: \caption{
2582: ``Brightness" or ``formation" temperature spectra on the dayside for three models (P$_n$ = 0.1, 0.3, and 0.5)
2583: of TrES-1. The formation temperature at the particular wavelength is the temperature
2584: where the optical depth at this wavelength reaches 2/3. See text for a discussion.
2585: }
2586: \label{fig2}
2587: \end{figure}
2588: \clearpage
2589:
2590: % figure 3
2591: \begin{figure}
2592: \centerline{
2593: \includegraphics[width=14.cm,angle=-90,clip=]{f3.eps}
2594: }
2595: \caption{
2596: Spectral energy distribution (SED) ($\lambda$F$_{\lambda}$ versus log$_{10}$($\lambda$))
2597: for the dayside of TrES-1 for three values of the day-night heat redistribution
2598: parameter P$_{n}$. Notice that observations in IRAC and MIPS bands cover
2599: only a small fraction of the SED of the planet.
2600: }
2601: \label{fig3}
2602: \end{figure}
2603: \clearpage
2604:
2605:
2606: % figure 4
2607: \begin{figure}
2608:
2609: \centerline{
2610: \includegraphics[width=6.cm,angle=-90,clip=]{f4a.eps}
2611: \includegraphics[width=6.cm,angle=-90,clip=]{f4b.eps}}
2612: \centerline{
2613: \includegraphics[width=6.cm,angle=-90,clip=]{f4c.eps}
2614: \includegraphics[width=6.cm,angle=-90,clip=]{f4d.eps}}
2615: \caption{
2616: The planet/star flux ratios versus wavelength from $\sim$1.5 $\mu$m
2617: to 30 $\mu$m for various models of four transiting
2618: EGPs measured by {\it Spitzer} at secondary eclipse.
2619: Notice the different scales employed in each panel. Models
2620: for different values of P$_n$ (\S\ref{redist}) and $\kappa_{\rm e}$
2621: %
2622: %(\S\ref{app4})
2623: %
2624: are provided where appropriate and the data
2625: from Table \ref{t3} for each planet are superposed. The plot legends
2626: indicate the color schemes used for the different EGPs. On the upper-left
2627: panel (HD 209458b), models with the lighter gray shade(s) are for
2628: the higher value(s) of P$_n$. Notice also that two different values
2629: for the flux at 24 $\mu$m (green) are shown on this same panel.
2630: The one with the question mark is a tentative update to
2631: the Deming et al. (2005) 24-$\mu$m measurement, kindly provided
2632: by Drake Deming (private communication). If the flux at
2633: 24 $\mu$m is indeed $\sim$0.0033$\pm{0.0003}$, then our model(s)
2634: with inversions provide the best fit at that wavelength as well.
2635: Note that the comparison between model and data should
2636: be made after the band-averaged flux-density ratios of the detected
2637: electrons are calculated. This not only incorporates the significant
2638: widths of the {\it Spitzer} bandpasses, but the fact that one should
2639: compare photon counts (or detected electrons), and not monochromatic
2640: fluxes. The result is that the theoretical IRAC predictions do not
2641: actually vary on the figures as much as do the plotted spectra and that the predicted
2642: contrasts, for instance between IRAC 1 and IRAC 2, are more muted, even
2643: when the pronounced bump at and near $\sim$3.6 $\mu$m obtains.
2644: However, to avoid the resultant clutter, we do not put these
2645: bandpass predictions on these spectral plots. See text for a discussion
2646: of each irradiated planet and the inferences drawn.
2647: }
2648: \label{fig4}
2649: \end{figure}
2650: \clearpage
2651:
2652: % figure 5
2653: \begin{figure}
2654:
2655: \centerline{
2656: \includegraphics[width=14.cm,angle=-90,clip=]{f5.eps}}
2657: \caption{
2658: The same as Fig. \ref{fig4}, but for various models of HD 209458b between 1 $\mu$m and 4 $\mu$m.
2659: The dependence on both P$_n$ and the presence and strength of a thermal inversion
2660: is greatest in this wavelength region. Note that a thermal inversion
2661: flips what would be water absorption features into emission features, altering
2662: the interpretation of any data in this spectral region significantly.
2663: Superposed is the Knutson et al. (2007c) data point for IRAC 1 and the
2664: Richardson et al. (2003) upper limit near 2.2 $\mu$m. See text
2665: for a discussion.
2666: }
2667: \label{fig5}
2668: \end{figure}
2669: \clearpage
2670: \thispagestyle{empty}
2671: \setlength{\voffset}{-25mm}
2672:
2673: % figure 6
2674: \begin{figure}
2675: \centerline{
2676: \includegraphics[width=4.7cm,angle=-90,clip=]{f6a.eps}
2677: \includegraphics[width=4.7cm,angle=-90,clip=]{f6b.eps}}
2678: \centerline{
2679: \includegraphics[width=4.7cm,angle=-90,clip=]{f6c.eps}
2680: \includegraphics[width=4.7cm,angle=-90,clip=]{f6d.eps}}
2681: \centerline{
2682: \includegraphics[width=4.7cm,angle=-90,clip=]{f6e.eps}
2683: \includegraphics[width=4.7cm,angle=-90,clip=]{f6f.eps}}
2684: \centerline{
2685: \includegraphics[width=4.7cm,angle=-90,clip=]{f6g.eps}
2686: \includegraphics[width=4.7cm,angle=-90,clip=]{f6h.eps}}
2687: \caption{
2688: Theoretical light curves for the non-transiting EGP $\upsilon$ And b in
2689: the MIPS 24-$\mu m$ band for different inclinations ($i$ = 45$^{\circ}$
2690: and 80$^{\circ}$), values of P$_{n}$ (0.0 and 0.3), values of $\kappa_{\rm e}$
2691: (0.0 and 0.2 cm$^2$/g), and a range of planetary
2692: radii. Superposed are the light-curve measurements of Harrington et al. (2006).
2693: Note that Harrington et al. (2006) obtain only relative contrast values, not
2694: absolute values. Therefore, the data on each panel are shifted in absolute
2695: contrast space, while maintaining the measured relative values. The day/night
2696: contrast difference is preserved. In this way, we find the corresponding best fits
2697: consistent with the observations, but for different inclinations, etc. Note that the scales
2698: for the different panels can be different, though the right panel that faces each left panel
2699: has the same scale as that left panel. The left panels depict models without
2700: a thermal inversion, while the right panels depict models that have thermal inversions
2701: created using $\kappa_{\rm e}$ = 0.2 cm$^2$/g.
2702: %
2703: %(\S\ref{app4}).
2704: %
2705: Together, this set of
2706: panels illustrates the dependence on the three most important
2707: free parameters: P$_{n}$, $i$, R$_{p}$, and $\kappa_{\rm e}$. See text for a discussion.
2708: }
2709: \label{fig6}
2710: \end{figure}
2711: \clearpage
2712: \setlength{\voffset}{0mm}
2713:
2714: % figure 7
2715: \begin{figure}
2716: \centerline{
2717: \includegraphics[width=5.cm,angle=-90,clip=]{f7a.eps}
2718: \includegraphics[width=5.cm,angle=-90,clip=]{f7b.eps}}
2719: \centerline{
2720: \includegraphics[width=5.cm,angle=-90,clip=]{f7c.eps}
2721: \includegraphics[width=5.cm,angle=-90,clip=]{f7d.eps}}
2722: \centerline{
2723: \includegraphics[width=5.cm,angle=-90,clip=]{f7e.eps}
2724: \includegraphics[width=5.cm,angle=-90,clip=]{f7f.eps}}
2725: \centerline{
2726: \includegraphics[width=5.cm,angle=-90,clip=]{f7g.eps}
2727: \includegraphics[width=5.cm,angle=-90,clip=]{f7h.eps}}
2728: \caption{
2729: The same as Fig. \ref{fig6}, but for HD 179949b in the IRAC 4 band
2730: for two different inclination angles (45$^{\circ}$ and 80$^{\circ}$), two values of
2731: P$_{n}$ (0.0 and 0.3), a range of planetary radii, and two values of
2732: $\kappa_{\rm e}$ (0.0 and 0.08 cm$^2$/g). The light curve data from
2733: Cowan et al. (2006) are superposed on each panel.
2734: See text for a discussion.
2735: }
2736: \label{fig7}
2737: \end{figure}
2738: \clearpage
2739:
2740: % figure 8
2741: \begin{figure}
2742: \centerline{
2743: \includegraphics[width=13.cm,height=18.cm,angle=-90,clip=]{f8.eps}}
2744: \caption{
2745: A comparison between the light-curve measurements of HD 189733b in the
2746: IRAC 4 band (8 $\mu$m) performed by Knutson et al. (2007b, hexagons)
2747: and our theoretical light curves for various values of P$_n$ (0.1, 0.3, 0.5).
2748: Most of these models employ values for the redistribution pressure range (P$_0$ and P$_1$)
2749: of 0.1 and 1.0 bars, but one model (dotted, and P$_n$ = 0.3) uses (P$_0$, P$_1$) = (0.05, 0.5) bars.
2750: See figure legend for model parameters. Also included is a model with 10$\times$solar metallicity
2751: (dashed). All the P$_n = 0.3$ models are in green.
2752: See text for a discussion.
2753: }
2754: \label{fig8}
2755: \end{figure}
2756:
2757:
2758: \end{document}
2759:
2760:
2761:
2762:
2763:
2764:
2765: (101) Winn, J.N., Noyes, R.W. et al. 2005, ApJ, 631, 1215
2766: (102) Deming, D., Seager, S., Richardson, L.J., Harrington, J. 2005,
2767: Nature, 434, 740
2768: (103) Benedict, G.F., McArthur, B.E., Forveille, T., Delfosse, X.,
2769: Nelan, E., Butler, R.P., Spiesman, W., Marcy, G., Goldman, B.,
2770: Perrier, C., Jefferys, W.H., Mayor, M. 2002, ApJ 581, L 115
2771: (104) Bean J.L., Benedict G.F., Endl M., 2006, ApJ 653, L65
2772: (105) Rivera, E.J., Lissauer, J.J., Butler, R.P.,
2773: Marcy, G.W., Vogt, S.S., Fischer, D.A., Brown, T.M.,
2774: Laughlin, G., Henry, G.W. 2005, ApJ 634, 625
2775: (106) Butler, R.P., Vogt, S.S., Marcy, G.W., Fischer, D.A.,
2776: Wright, J.T., Henry, G.W., Laughlin, G., Lissauer, J.J.
2777: 2004, ApJ 617, 580
2778: (107) Wright, J.T., Marcy, G.W., Fischer, D.A., Butler, R.P.,
2779: Vogt, S.S., Tinney, C.G., Jones, H.R.A., Carter, B.D.,
2780: Johnson, J.A., McCarthy, C., Apps, K. 2006, astro-ph/0611658
2781: (108) McArthur, B.E., Endl, M., Cochran, W.D.,
2782: Benedict, G.F., Fischer, D.A., Marcy, G.W., Butler, R.P.,
2783: Naef, D., Mayor, M., Queloz, D., Udry, S., Harrison, T.E.
2784: 2004, ApJ 614, L81
2785: (109) Valenti, J.A., Fischer, D.A. 2005 ApJS 159, 141
2786: (110) Butler R.P., Wright J.T., Marcy G.W., Fischer D.A., Vogt S.S.,
2787: Tinney C.G., Jones H.R.A., Carter B.D., Johnson J.A., McCarthy C.,
2788: Penny A.J. 2006 ApJ 646, 505 (Catalog of Nearby Exoplanets)
2789: (111) Butler, R.P., Marcy, G.W., Williams, E., Hauser, H., Shirts, P.
2790: 1997 ApJ 474, L115
2791: (112) Naef, D., Mayor, M., Beuzit, J.L., Perrier, C., Queloz, D.,
2792: Sivan, J.P., Udry, S. 2004, A\&A, 414 351
2793: (113) Mayor M., Queloz D. 1995, Nature 378, 355
2794: (114) Pasinetti Fracassini, L.E., Pastori, L., Covino, S., Pozzi, A. 2001,
2795: A\&A 367, 521
2796: (115) Santos, N.C., Israelian, G., Mayor, M., Rebolo, R., Udry, S.
2797: 2003 A\&A 398, 363
2798: (116) Butler, R.P., Marcy, G.W., Fischer, D.A., Brown, T.M.,
2799: Contos, A.R., Korzennik, S.G., Nisenson, P., Noyes, R.W. 1999, ApJ 526, 916
2800: (117) Wittenmyer R.A., Endl M., Cochran W.D., 2007, ApJ 654, 625
2801: (118) Santos N.C., Israelian G., Mayor M. 2004, A\&A 415,1153
2802: (119) Harrington, J., Hansen, B.M., Luszcz, S.H., Seager, S.,
2803: Deming, D., Menou, K., Cho, J.Y.-K., Richardson, L. J. 2006, Science 314, 623
2804: (120) Lovis, Ch., Mayor, M., Pepe, F., Alibert, Y.,
2805: Benz, W., Bouchy, F., Correia, A.C.M., Laskar, J.,
2806: Mordasini, Ch., Queloz, D., Santos, N.C., Udry, S.,
2807: Bertaux, J.-L., Sivan, J.-P. 2006, Nature 441, 305
2808: (121) Santos, N.C., Bouchy, F., Mayor, M., Pepe, F., Queloz, D.,
2809: Udry, S., Lovis, C., Bazot, M., Benz, W., Bertaux, J.-L., Lo Curto, G.,
2810: Delfosse, X., Mordasini, C., Naef, D., Sivan, J.-P., Vauclair, S.
2811: 2004, A\&A 426, L19
2812: (122) Wittenmyer, R.A., Endl, M., Cochran, W.D. 2007, ApJ 654 625
2813: (123) Fischer, D.A., Valenti, J. 2005, ApJ 622, 1102
2814:
2815: