<<

arXiv:astro-ph/0202236v1 12 Feb 2002 2 1 dmP Showman P. b-like” Adam Peg “51 of Planets and Circulation Atmospheric eecuae ytescessotie ntecs of case the in might obtained one successes measured, spec- the no been yet While by have 2001). encouraged al. objects be et these Barman of Gouken- 2000; tra 2000; al. 1998, Sasselov et (Bur- and leuque Seager growing 1997; their a al. modeling to et to rows led devoted has work of planets list extrasolar of discovery The Introduction 1. eec n hrceitcwn pesmyreach dif- may speeds temperature day-night characteristic the and ference photosphere, arguments the at- simple at homogeneous using that, a show assume we Nevertheless, planets torques mosphere. such of evo- fluid-dynamical and models radiative-transfer of lution Previous planets. because these within occur synchronous can from tidal departures rotation by small synchronized that these rapidly but that distances interactions, show are We orbital planets stars. b-like” their at Peg “51 from planets AU 0.1 giant than smaller extrasolar of spheres Abstract. 2002 January 18 Accepted 2001; March 23 Submitted topei yais neirsrcue in planets giant structure, interior dynamics, atmospheric words: Key these ideas. our from test radiation help measurements will thermally-emitted Future planets and planets. reflected these consequences of of sub- crucial evolution of with the deposition interior, for the the to in lead energy can stantial produc- the energy esti- in kinetic simpler tion that our show with we Furthermore, agree mates. generally and for circulation patterns plausible the three- indicate that preliminary HD209458b cir- of atmospheric present culation the of We simulations nonlinear planets. dimensional, properties b-like atmospheric Peg 51 there- the are of describing in conditions inadequate homogeneous Radiative-transfer fore assume nightside, regime. that or circulation dayside models the the on could on depending either dynamics; cov- predominantly the on exist The sensitively expected. depends are erage equilibrium chemical from and 08 are: codes thesaurus later) Your hand by inserted be (will no. manuscript A&A bevtied aCoedAu,LbrtieCsii CNRS Cassini, Laboratoire Cˆote d’Azur, la an de Sciences Observatoire Planetary of Department [email protected] Arizona, of University ∼ msec km 2 eeaietepoete fteatmo- the of properties the examine We xrslrpaes 1Pg D048 tides, HD209458, Peg, 51 planets, extrasolar − 1 epciey usata departures Substantial respectively. , 1 n rsa Guillot Tristan and 2 ∼ 0 K 500 M 59 60 ieCdx4 rne [email protected] France; 4, Cedex Nice 06304 6529, UMR ua n lntr aoaoy usn Z871 show- 85721; AZ Tucson, Laboratory, Planetary and Lunar d nPprIt vlaetednmclsaeo Pegasi- in- of may state dynamics dynamical how the discuss obtained evaluate and profiles to atmospheres, temperature planet I the use Paper we in down- paper, the this of en- by In flux kinetic a provided of with dissipation be ergy subsequent likely and transport most and ward HD209458b, would radius the this explain source that to additional needed an is Pe- that energy of of advocated atmo- also evolution We the the planets. how governs gasi showed Show- condition we and boundary I), (Guillot spheric Paper paper hereafter preceeding 2001, the man In Henry 2000). 2000; al. al. transit- et et the (Charbonneau of HD209458b discovery giant are gas the ing than by indicated method as transit also planets, the of other are by 27% characterized They constitute easily and planets. more orbit far giant they so extrasolar surveyed that currently-known Their stars fact of AU. the 0.1 1% by than nearly demonstrated less gi- at is gas stars as importance define solar-type and orbiting planets” ants “Pegasi which planets, dub b-like Peg henceforth 51 we important: most the is diation spectral planets. temperature, extrasolar composition, of evolution the and play for can appearance role it but Advec- major considered, planets. a been irradiated previously atmo- never intensely the has most of tion description the far adequate of thus an spheres calculated provide models to the fail may that also it so atmo- importantly, chemistry and most distribution spheric that temperature horizontal show the will Barman affects We 2000; 2001). al. (- al. et Goukenleuque et profile 1998; ef- temperature Sasselov profoundly same and vertical to ger the atmospheric shown with the been dwarfs has alter This brown temperature. isolated fective they of different that those significantly irradiation atmospheres than The their star: make a plan- 2000; a can only extrasolar al. to endure cite proximity of et their to feature is Liebert 2001 important ets al. 1997; an et al. However, Schweitzer et few). 2001; al. Allard et 1996; Geballe of al. case et the ley models in even ( theoretical well, objects which low-temperature observations the for reproduce dwarfs, now brown similar the ewl ou nteetaoa lnt o hc irra- which for planets extrasolar the on focus will We ∼ %o h bobdselrenergy. stellar absorbed the of 1% T eff ∼ ASTROPHYSICS 00Ko es eg Mar- (e.g. less) or K 1000 ASTRONOMY AND 2 A.P. Showman and T. Guillot: Dynamics of Pegasi-planet Atmospheres

fluence the cloud abundance, chemical composition, and thermal state (all of which will be amenable to observation in the near future). We then present preliminary numer- ical simulations of the circulation that indicate plausible circulation patterns and show how downward propagation of kinetic energy from the atmosphere to the interior can occur. After reviewing the expected interior structure (Sec- tion 2), we begin in Section 3 with the problem of tides: Pe- gasi planets have been predicted to rotate synchronously (Guillot et al. 1996), implying that they always present the same face toward the star. We argue however that dynamical torques may maintain the interior in a non- synchronous rotation state, which has important impli- cations for understanding atmospheric processes. In sec- tion 4, we discuss the probable wind speeds, day-night Fig. 1. Left: Temperature profiles for the “hot” model temperature differences, and flow geometries, including (thin line) and “cold” model (thick grey line) at times both order-of-magnitude arguments and our numerical (5.37 and 0.18 Ga, respectively) when the models match simulations. A summary of the results is provided in sec- HD209458b’s measured radius. The diamonds indicate tion 5. the radiative/convective boundary. The terms “hot” and “cold” are chosen to indicate the models’ relative tem- 2. Interior structure peratures in the region from ∼ 1–300 bars, which is the important region of the atmosphere for affecting the evo- Fully self-consistent models including both the atmo- lution. See Paper I for details. The discontinuity in the spheres and interiors of Pegasi planets do not yet exist. profiles’ slopes is unphysical; it results from the fact that To obtain first-cut estimates of the expected tempera- atmospheric models assume a much higher intrinsic lumi- ture profiles, we therefore use the same approach as in nosity than predicted by evolution models. Right: Dimen- Paper I: we choose two extreme evolution models (“hot” sionless mean specific heat per particle in the atmosphere. and “cold”) that match the radius of HD209458b (but at The local maxima in c are due to the dissociation of the different ages and hence intrinsic luminosities). These evo- p H molecule. lution calculations only pertain to pressures larger than 3 2 to 10 bars (depending on the boundary condition), and so we extend the profiles to lower pressure using radiative- transfer calculations for intensely-irradiated planets from the planet is about 10,000 times smaller than the energy Marley (personal communication) and Barman (2001). absorbed from the star. If energy dissipation in the outer The calculations are not strictly appropriate to the case of layers is large (see Paper I), another convective zone can HD209458b, but provide us with reasonable estimates of appear at the levels where dissipation is the largest. This the expected structure (and as we will see our qualitative however requires a very high dissipation of ∼ 10% of the conclusions are not sensitive to the type of profile chosen). incoming heat flux. We therefore chose to only consider The resulting “hot” and “cold” models that we will use the simple radiative/convective scenario, as depicted in hereafter are depicted in Figure 1. Interestingly, Paper I Fig. 2. shows that these two models are representative of a rel- We split the planet into an “atmosphere” dominated atively wide variety of models of HD209458b, including by stellar heating, with possible horizontal temperature those with energy dissipation. inhomogeneities, and an “interior” including the con- It should be noted that the high temperatures can trig- vective core, for which inhomogeneities should be much ger the dissociation of the hydrogen molecule, which is in- smaller. The crux of the problem is to understand how dicated in Fig. 1 by a local maximum of the heat capacity the heat absorbed on the day side and near the is cp. This effect is important for our purposes because it redistributed by and/or rotation to the night side implies that stellar heat can be stored in hot regions and and to the poles. reclaimed in colder regions by molecular recombination. Other molecules (e.g. H O) can undergo dissociation, but 2 3. Synchronization of Pegasi planets this will be neglected due to their small abundances. Evolution models of Pegasi planets show that two in- It has been shown that the tides raised by the star on terior regions exist: a radiative zone (including the atmo- Pegasi planets should rapidly drive them into synchronous sphere) that extends down to pressures of 100 to 800bar, rotation (Guillot et al. 1996; Marcy et al. 1997; Lubow and a deeper convective core. The intrinsic luminosity of et al. 1997). This can be seen by considering the time A.P. Showman and T. Guillot: Dynamics of Pegasi-planet Atmospheres 3

3.1. Spindown energies Angular momentum conservation requires that as the planet spins down, the orbit expands. The energy dissi- pated during the spindown process is the difference be- tween the loss in spin kinetic energy and the gain in orbital energy: d 1 1 E˙ = − k2MR2ω2 − Ma2ω2 , (2) dt  2 2 s 

where k is the dimensionless radius of gyration (k2 = I/MR2, I being the planet’s moment of inertia). The time derivative is negative, so E˙ , the energy dissipated, is positive. The orbital energy is the sum of the planet’s gravitational potential energy and orbital kinetic energy and is negative by convention. The conservation of angu- Fig. 2. Conjectured dynamical structure of Pegasi plan- lar momentum implies that the rate of change of ω is ets: At pressures larger than 100–800bar, the intrinsic s constrained by that on ω: heat flux must be transported by . The con- vective core is at or near synchronous rotation with the d Ma2ω + k2MR2ω =0. (3) star and has small latitudinal and longitudinal tempera- dt s ture variations. At lower pressures a radiative envelope is  The fact that the planetary radius changes with time may present. The top part of the atmosphere is penetrated by slightly affect the quantitative results. However, since τ the stellar light on the day side. The spatial variation in syn is so short, it can be safely neglected in this first-order insolation should drive winds that transport heat from the estimate. R being held constant, it is straightforward to day side to the night side (see text). show, using Kepler’s third law, that:

2 2 E˙ = −k MR (ω − ωs)ω. ˙ (4) scale to tidally despin the planet (Goldreich & Soter 1966; (Note thatω ˙ is negative, and so E˙ is positive.) 2 2 2 Hubbard 1984): The total energy dissipated is E ≈ k MR (ωs −ω) /2, neglecting variation of the orbital distance. Using the mo- R3 M 2 a 6 ment of inertia and rotation rate of (k2 = 0.26 τsyn ≈ Q (ω − ωs) , (1) −4 −1 GM  M⋆  R  and ω = 1.74 × 10 sec ), we obtain for HD209458b E ≈ 4 × 1041 erg. If this energy were dissipated evenly where Q, R, M, a, ω and ωs are the planet’s tidal dissi- throughout the planet, it would imply a global tempera- pation factor, radius, mass, orbital semi-major axis, rota- ture increase of 1400 K. tional angular velocity, and synchronous (or orbital) an- By definition of the synchronization timescale, the dis- gular velocity. M∗ is the star’s mass, and G is the gravi- sipation rate can be written: tational constant. Factors of order unity have been omit- k2MR2(ω − ω )2 ted. A numerical estimate for HD209458b (with ω equal E˙ = s . (5) to the current Jovian rotation rate) yields a spindown τsyn time τsyn ∼ 3Q years. Any reasonable dissipation factor With Q of 105, a value commonly used for Jupiter, Q (see Marcy et al. 1997; Lubow et al. 1997) shows that 5 τsyn ∼ 3 × 10 years and the dissipation rate is then HD209458b should be led to synchronous rotation in less 29 −1 than a few million years, i.e., on a time scale much shorter ∼ 10 ergsec , or 35,000 times Jupiter’s intrinsic lumi- than the evolution timescale. Like other Pegasi planets, nosity. Lubow et al. (1997) have suggested that dissipation HD209458b is therefore expected to be in synchronous ro- in the radiative zone could exceed this value by up to two tation with its 3.5-day orbital period. orders of magnitude, but this would last for only ∼ 100 years. Nevertheless, stellar heating drives the atmosphere away from synchronous rotation, raising the possibility The thermal pulse associated with the initial spin- that the interior’s rotation state is not fully synchronous. down is large enough that, if the energy is dissipated in Here, we discuss (1) the energies associated with the the planet’s interior, it may affect the planet’s radius. It planet’s initial transient spindown, and (2) the possible has previously been argued (Burrows et al. 2000) that equilibrium states that could exist at present. Pegasi planets must have migrated inward during their first 107 years of evolution; otherwise, they would have contracted too much to explain the observed radius of 4 A.P. Showman and T. Guillot: Dynamics of Pegasi-planet Atmospheres

HD209458b. But the thermal pulse associated with spin- two cases, depending on the physical mechanisms that de- down was not included in the calculation, and this extra termine the gravitational torque on the atmosphere. energy source may extend the time over which migration The first possibility is that the gravitational torque was possible. on the atmosphere pushes the atmosphere away from Nevertheless, it seems difficult to invoke tidal synchro- synchronous rotation (i.e., it increases the magnitude of nization as the missing heat source necessary to explain the atmosphere’s angular momentum measured in the HD209458b’s present (large) radius. High dissipation rates synchronously-rotating reference frame) as has been hy- are possible if τsyn is small, but in the absence of a mech- pothesized for Venus (Ingersoll and Dobrovolskis 1978, anism to prevent synchronization, E˙ would drop as soon Gold and Soter 1969). To balance the torque on the atmo- as t > τsyn. The most efficient way of slowing the planet’s sphere, the interior must have a net angular momentum 10 contraction is then to invoke τsyn ∼ 10 years. In that of the same sign as the atmosphere (so that both either case, the energy dissipated becomes E˙ ∼ 1024 erg sec−1, super- or subrotate). In Fig. 3, superrotation then corre- which is two orders of magnitude smaller than that neces- sponds to a clockwise flux of angular momentum, whereas sary to significantly affect the planet’s evolution (Paper I; subrotation corresponds to the anticlockwise scenario. Bodenheimer et al. 2001). For the present-day dissipation The second possibility is that the gravitational torque to be significant, an initial rotation rate 10 times that of on the atmosphere tends to synchronize the atmosphere. modern-day Jupiter would be needed. But the centripetal This possibility may be relevant because, on a gas-giant acceleration due to rotation exceeds the gravitational ac- planet, it is unclear that the high-temperature and high- ◦ celeration at the planet’s surface for rotation rates only pressure regions would be 90 out of phase, as is expected twice that of modern-day Jupiter, so this possibility is on a terrestrial planet. If the interior responds sufficiently ruled out. Furthermore, such long spindown times would to atmospheric perturbations to keep the deep isobars require a tidal Q of ∼ 109–1010, which is ∼ 104 times the Q independent of surface (an assumption that values inferred for Jupiter, Uranus, and Neptune from con- seems to work well in modeling Jupiter’s cloud-layer dy- straints on their satellites’ orbits (Peale 1999, Banfield and namics), then high-pressure and high-temperature regions Murray 1992, Tittemore and Wisdom 1989). Dissipation will be in phase, which would sweep the high-mass re- of the energy due to transient loss of the planet’s initial gions downwind and lead to a torque that synchronizes the spin energy therefore cannot provide the energy needed to atmosphere. Furthermore, if a resonance occurs between explain the radius of HD209458b. the tidal frequency and the atmosphere’s wave-oscillation Another possible source of energy is through circular- frequency, the resulting gravitational torques also act to ization of the orbit. Bodenheimer et al. (2001) show that synchronize the atmosphere (Lubow et al. 1997). In this the resulting energy dissipation could reach 1026 erg sec−1 case, the interior and atmosphere have net angular mo- if the planet’s tidal Q is 106 and if a hypothetical com- menta of opposite signs. Depending on the sign of the panion planet pumps HD209458b’s eccentricity to values atmosphere/interior momentum flux, this situation would near its current observational upper limit of 0.04. If such a correspond in Fig. 3 to either the clockwise or anticlock- companion is absent, however, the orbital circularization wise circulation of angular momentum cases. time is ∼ 108 years, so this source of heating would be neg- In both cases above, the gravitational torque on the at- ligible at present. Longer circularization times of 109–1010 mosphere arises because of spatial density variations asso- years would allow the heating to occur until the present- ciated with surface meteorology, which are probably con- day, but its magnitude is then reduced to 1025 erg sec−1 fined to pressures less than few hundred bars. Therefore, or lower, which is an order of magnitude smaller than the this torque acts on only a small fraction of the planet’s dissipation required. mass. In contrast, the gravitational torque on the gravi- tational tidal bulge affects a much larger fraction of the planet’s mass. It is not necessarily true that the torque on 3.2. The equilibrium state the atmosphere is negligible in comparison with the torque The existence of atmospheric winds implies that the atmo- on the interior, however, because the gravitational tidal sphere is not synchronously rotating. Because dynamics bulge is only slightly out of phase with the line-of-sight can transport angular momentum vertically and horizon- to the star (e.g., the angle is 10−5 radians for Jupiter). In tally (including the possibility of downward transport into contrast, the angle between the meteorologically-induced the interior), the interior may evolve to an equilibrium ro- density variations and the line of sight to the star could be tation state that is asynchronous. Here we examine the up to a radian. Detailed calculations of torque magnitudes possibilities. would be poorly constrained, however, and will be left for Let us split the planet into an “atmosphere”, a part of the future. small mass for which thermal effects are significant, and an Several mechanisms may act to transfer angular mo- “interior” encompassing most of the mass which has min- mentum between the atmosphere and interior. Kelvin- imal horizontal thermal contrasts. Suppose (since τsyn is Helmholtz shear instabilities, if present, would smooth short) that the system has reached steady state. Consider interior-atmosphere differential rotation. On the other A.P. Showman and T. Guillot: Dynamics of Pegasi-planet Atmospheres 5

This estimate is an upper limit for the globally averaged Orbit Atmosphere asynchronous rotation rate because it assumes that kinetic energy transported into the interior has angular momen- tum of a single sign. The global-average asynchronous ro- gravitational dynamical torques torques tation could be even lower if angular momenta transported downward in different regions have opposite signs. The spindown time is uncertain and depends on the Interior planet’s Q according to Eq. (1). To date, only one estimate for the Q of a Pegasi planet exists (Lubow et al. 1997), Fig. 3. Angular momentum flow between orbit, interior, which suggests Q ∼ 100, at least during the early stages and atmosphere for a Pegasi planet in steady state. Ar- of spindown. Orbital constraints on the natural satellites rows indicate flow of prograde angular momentum (i.e., of the giant planets imply that, at periods of a few days, that with the same sign as the orbital angular momen- the tidal Q of Uranus, Neptune, and Jupiter are of order 5 tum) for two cases: Anticlockwise: Gravitational torque ∼ 10 (Tittemore and Wisdom 1989, Banfield and Mur- on atmosphere is retrograde (i.e., adds westward angular ray 1992, Peale 1999), which suggests a spindown time of 5 momentum to atmosphere). For torque balance, the gravi- a few ×10 years (Eq. 1). Although the mechanisms that tational torque on the interior must be prograde (i.e., east- determine these planets’ Q values remain uncertain, likely ward). These gravitational torques must be balanced by possibilities include friction in the solid inner core (Der- fluid-dynamical torques that transport retrograde angular mott 1979) or overturning of tidally-forced waves in the momentum from atmosphere to interior. Clockwise: Grav- fluid envelope (Ioannou and Lindzen 1993, Houben et al. itational torque on atmosphere is prograde, implying a 2001). Both mechanisms are possible for Pegasi planets, retrograde torque on the interior and downward transport and the wave-dissipative mechanism should be more ef- of prograde angular momentum from atmosphere to inte- fective for Pegasi planets than for Jupiter because of the rior. Atmosphere will superrotate if gravitational torques thicker stable (radiative) layer in the former (Houben et push atmosphere away from synchronous (as on Venus). al. 2001). These calculations suggest that spindown times 5 6 It will subrotate if gravitational torques synchronize the of ∼ 10 –10 years are likely, although larger values can- atmosphere (e.g., gravity-wave resonance; cf. Lubow et al. not be ruled out. (1997)). As discussed in Section 4, experience with planets in our solar system suggests that atmospheric kinetic energy −2 is generated at a flux of 10 FH#, and if all of this energy enters the interior, then η ∼ 10−2. Using a spindown time 5 −5 −1 hand, waves transport angular momentum and, because of 3 × 10 years implies that ω − ωs ∼ 2 × 10 sec , they act nonlocally, often induce differential rotation which is comparable to the synchronous rotation fre- rather than removing it. The quasi-biennial oscillation on quency. The implied winds in the interior are then of or- is a classic example (see, e.g., Andrews et al. 1987, der ∼ 2000msec−1. Even if η is only 10−4, the interior’s Chapter 8). Atmospheric tides (large-scale waves forced winds would be 200 m sec−1. The implication is that the by the solar heating) are another example of such a wave; interior’s spin could be asynchronous by up to a factor on Venus, tides play a key role in increasing the rotation of two, depending on the efficiency of energy and momen- rate of the cloud-level winds. Finally, vertical advection tum transport into the interior. If spindown times of ∼ 106 may cause angular momentum exchange between the at- years turn out to be serious under- or overestimates, then mosphere and the interior of Pegasi planets, because the greater or lesser asynchronous rotation would occur, re- angular momentum of air in updrafts and downdrafts need spectively. not be the same. A simple estimate illustrates the extent of nonsyn- chronous rotation possible in the interior. Suppose that 4. : Possible regimes and influence on evolution the globally-averaged flux of absorbed starlight is FH#, 8 −1 −2 which is of order 10 ergs cm for Pegasi planets near 4.1. Basic Parameter Regime 0.05 AU, and that the globally-averaged flux of kinetic energy transported from the atmosphere to the interior is Rotation plays a central role in the atmospheric dynamics

ηFH#, where η is small and dimensionless. If this kinetic of Pegasi planets, and HD209458b in particular. The ratio energy flux is balanced by dissipation in the interior with of nonlinear advective accelerations to accelera- −1 −1 a spindown timescale of τsyn, then the deviation of the tions in the horizontal momentum equation is uf L rotation frequency from synchronous is (called the Rossby number), where u is the mean hori- zontal wind speed, f = 2Ω sin φ is the “Coriolis param- 1/2 4πηFH#τ eter” (e.g. Holton 1992, pp. 39-40), Ω is the rotational ω − ω = syn (6) s  k2M  angular velocity, φ is latitude, and L is a characteristic 6 A.P. Showman and T. Guillot: Dynamics of Pegasi-planet Atmospheres length scale. Rossby numbers of 0.03–0.3 are expected for when the Richardson number becomes smaller than 1/4 winds of planetary-scale and speeds ranging from 100– (cf. Chandrasekhar 1961), i.e. when −1 1000msec . In Section 3 we showed that modest asyn- 2 chronous rotation may occur, in which case the Rossby N 1 Ri = 2 < , (7) number could differ from this estimate by a factor up (du/dz) 4 to ∼ 2. These estimates suggest that nonlinear advective In the perfect gas approximation, assuming a uniform terms are small compared to the Coriolis accelerations, composition, which must then balance with the pressure-gradient accel- 2 g erations. (The Rossby number could be ∼ 1 if the winds N = (∇ad − ∇T ). (8) reach 2000–3000msec−1, which we show below is proba- H bly the maximum allowable wind speed.) where R is the universal gas constant divided by the mean molecular mass, cp is the specific heat, H = RT/g is the The zonality of the flow can be characterized by the pressure scale height, ∇ad = R/cp is the adiabatic gradi- 1/2 Rhines’ wavenumber, kβ ∼ (β/u) , where β is the ent, ∇T = d ln T/d ln p, T is temperature, and p is pres- derivative of f with northward distance (Rhines 1975). sure. The hypothesis of uniform composition is adequate The half-wavelength implied by this wavenumber, called despite hydrogen dissociation because the timescales in- the Rhines’ scale Lβ, provides a reasonable estimate for volved are generally much longer than the dissociation the jet widths on all four outer planets in our solar system timescales. It can be noted that a shear instability could (Cho and Polvani 1996; see Table 1). For HD209458b, the appear at smaller Richardson numbers in the presence Rhines’ scale is ∼ 1.5×1010(u/1000msec−1)1/2 cm, which of efficient radiative diffusion (Zahn 1992; Maeder 1995). exceeds the planetary radius if the wind speed exceeds This possibility will not be examined here. about 400msec−1. The maximal wind speed at which Kelvin-Helmholtz instabilities occur can then be derived by integration of Another measure of horizontal structure is the Rossby Eq. (7): deformation radius (Gill 1982, p. 205), L ∼ NH/f, D 1 P where H is the scale height and N is the Br¨unt-Vaisala u (P ) ∼ [RT (∇ − ∇ )]1/2 d ln p, (9) max 2 ad T frequency (i.e., the oscillation frequency for a vertically ZP0 displaced air parcel; Holton 1992, p. 54). At pressures of a The value of umax thus derived at P ∼ 1bar is of the or- few bars, the temperature profiles calculated for irradiated der of 3000ms−1, to be compared to the winds of Jupiter, extrasolar giant planets by Goukenlouque et al. (2000) Saturn, Uranus and Neptune which reach 100–500 m s−1 −1 suggest N ∼ 0.0015 sec . With a scale height of 700 km, (Ingersoll et al. 1995). Our estimate for Pegasi planets, the resulting deformation radius is 40,000 km. In contrast, assuming the winds are measured in the synchronously- the deformation radii near the of the Jupiter, rotating frame, may be uncertain by a factor of ∼ 2 due Saturn, Uranus, and Neptune are of order 2000 km (Table to the possibility of a non-synchronously-rotating inte- 1). rior. In comparison, the expected speed of sound at the tropopause of HD209458b is ∼ 2400msec−1. The estimated Rhines’ scale (for winds of > A characteristic timescale for zonal winds to redis- − 500msec 1, which we show later are plausible speeds) tribute temperature variations over scales similar to the and deformation radius of Pegasi planets are similar to the planetary radius R then stems from τzonal ∼> R/umax. planetary radius, and they are a larger fraction of the plan- The radiative heating timescale can be estimated by etary radius than is the case for Jupiter, Saturn, Uranus, a ratio between the thermal energy within a given layer and Neptune (Table 1). This fact suggests that eddies may and the layer’s net radiated flux. In the absence of dy- grow to hemispheric scale in the atmospheres of Pegasi namics, absorbed solar fluxes balance the radiated flux, planets and that, compared with the giant planets in our but dynamics perturbs the temperature profile away from solar system, the general circulation hot may be radiative equilibrium. Suppose the radiative equilibrium more global in character. Unless the winds are extremely temperature at a particular location is Trad and the actual weak, Pegasi planets are unlikely to have > 10 jets as do temperature is Trad + ∆T . The net flux radiated towards Jupiter and Saturn. 3 outer space is then 4σTrad∆T and the radiative timescale is An upper limit on the atmospheric wind speed can P c be derived from shear-instability considerations. We as- τ ∼ p , (10) rad g 4σT 3 sume that no zonal winds are present (u(P0) = 0) in the convective core, a consequence of synchronization by where σ is the Stefan-Boltzmann constant. tidal friction. The build-up of winds at higher altitudes in Figure 4 shows estimates of τzonal and τrad for the radiative envelope is suppressed by Kelvin-Helmholtz HD209458b calculated using the temperature profiles from instabilities if the shear becomes too large. This occurs the “hot” (thin line) and “cold” (thick grey line) models A.P. Showman and T. Guillot: Dynamics of Pegasi-planet Atmospheres 7

Table 1. Rhines’ and deformation lengths for giant planets

u R Ω Lβ LD Jet width − − − ( m sec 1) (107 m) (10 4 sec 1) (107 m) (107 m) (107 m) Jupiter 50 7.1 1.74 1.0 0.2 ∼ 1 Saturn 200 6.0 1.6 1.9 0.2 ∼ 2 Uranus 300 2.6 1.0 2.0 0.2 ∼ 2 Netpune 300 2.5 1.1 1.9 0.2 ∼ 2 − HD209458b ? 10 ∼ 0.2 15(u/1 km sec 1)1/2 ∼ 4 ? −1 −1 Note. LD calculated at the tropopause using N = 0.01 sec for Saturn, Uranus, and Neptune, and N = 0.02 sec for Jupiter. from Section 2. The zonal timescale is estimated by calcu- lating the maximum wind speed that can exist as a func- tion of pressure given the static stability associated with each model, while the radiative time is calculated using the temperatures and heat capacities shown in Fig. 1. At pressures exceeding 0.1 bar, radiation is slower than the maximal advection by zonal winds, but by less than one order of magnitude. The consequent day/night tempera- ture difference ∆Tday−night to be expected is:

∆T − − day night ∼ 1 − e τzonal/τrad . (11) ∆Trad where ∆Trad is the day-night difference in radiative equi- librium temperatures. Rough estimates from Fig. 4 sug- gest that τzonal/τrad ∼ 0.3 at 1 bar, implying that ∆Tday−night/∆Trad ∼0.3. If ∆Trad = 1000 K, this would Fig. 4. Left: Characteristic time scales as a function of imply day-night temperature differences of 300 K at 1 bar. pressure level. τzonal is the minimal horizontal advection Values of ∆Tday−night even closer to ∆Trad are likely given time (dashed). τ is the timescale necessary to cool/heat the fact that slower winds will lead to an even more ef- rad a layer of pressure P and temperature T by radiation alone fective cooling on the night side and heating on the day assuming the gas is optically thick (solid). (The plotted side. radiative timescale is an underestimate at pressures less The small radiative time scale implies that, for the than about 0.3 bars because the atmosphere becomes op- day-night temperature difference to be negligible near the tically thin at those pressures.) For each case, the thin planet’s photosphere, atmospheric winds would have to black and thick grey lines correspond to the hot and cold be larger than the maximum winds for the onset of shear models from Section 2 and Fig.1. Right: Approximate cool- instabilities. ing/heating rate as a function of pressure (see Eq. 18).

4.2. Nature of the circulation An understanding of the horizontal temperature difference Because the Rossby number is small, the dominant and mean wind speed is desirable. It is furthermore of in- balance in the horizontal momentum equation at mid- terest to clarify the possible geometries the flow may take. latitudes is between the and the pressure- For Pegasi planets, we envision that the dominant forcing gradient force (geostrophic balance). We adopt the prim- is the large-scale day-night heating contrast, with minimal itive equations, which are the standard set of large-scale role for moist convection. This situation differs from that dynamical equations in a stably-stratified planetary at- of Jupiter, where differential escape of the intrinsic heat mosphere. When differentiated with pressure (which is flux offsets the solar heating contrast and moist convection used here as a vertical coordinate), this balance leads to plays a key role. For Pegasi planets, such an offset between the well-known thermal wind equation (e.g. Holton 1992, intrinsic and stellar fluxes cannot occur, because the in- p. 75): 4 trinsic flux is 10 times less than the total flux. The fact ∂v that the day-night heating contrast occurs at hemispheric f = −Rk × ∇HT (12) ∂ ln p scale — and that the relevant dynamical length scales for Pegasi planets are also hemispheric (Section 4.1) — in- where v is the horizontal wind, p is pressure, ∇H is the creases our confidence that simple analysis, focusing on horizontal gradient, T is temperature, and k is the unit the hemispheric-scale circulation, can provide insight. upward vector. Assuming the interior winds are small 8 A.P. Showman and T. Guillot: Dynamics of Pegasi-planet Atmospheres compared to the atmospheric winds, this implies that, to q f 1/2 ∆Thoriz ∼ R (17) order-of-magnitude, cp R∆ ln p R A rough estimate of the heating rate, q/cp, results from |v|∼ ∆Thoriz∆ ln p (13) fR the analysis in Section 4.1:

3 where |v| and ∆Thoriz are the characteristic wind speed q 4σT ∆Tg and horizontal temperature difference in the atmosphere = (18) cp pcp and ∆ ln p is the difference in log-pressures from bottom to top of the atmosphere. Non-sychronous rotation of the where ∆T is the characteristic magnitude of the differ- interior would alter the relation by changing the value of ence between the actual and radiative equilibrium tem- f, but the uncertainty from this source is probably a factor peratures in the atmosphere. The heating rate depends on of two or less (see Eq. (6) and subsequent discussion). the dynamics through ∆T . We simply evaluate the heat- The thermodynamic energy equation is, using pressure ing rate using Eq. (18) with ∆T ≈ T/2. The results are as a vertical coordinate (e.g. Holton 1992, p. 60), shown in Fig. 4. The key difficulty in applying the equations to Pe- 2 2 ∂T H N q gasi planets is the fact that q/cp depends on pressure, + v · ∇HT − ω = (14) ∂t Rp cp and ∆Thoriz probably should too, but Eqs. (16) and (17) were derived assuming that ∆Thoriz is constant. We can where t is time, ω = dp/dt is vertical velocity, q is the still obtain rough estimates by inserting values of q/cp at −1 −1 −4 specific heating rate (ergg sec ), and we have assumed several pressures. At ∼ 50–100 bars, where q/cp ∼ 10 – an ideal gas. 10−3 Ksec−1, we obtain temperature differences and wind A priori, it is unclear whether the radiative heating speeds of 50–150K and 200–600msec−1. At 1 bar, where −2 −1 and cooling is dominantly balanced by horizontal advec- q/cp reaches 10 Ksec , the estimated temperature con- tion (second term on left of Eq. (14)) or vertical advec- trast and wind speed is ∼ 500K and ∼ 2000msec−1. tion (third term on the left). To illustrate the possibilities, The estimates all assume f ≈ 3 × 10−5 sec−1, R = we consider two endpoint scenarios corresponding to the 3500JkgK−1, and ∆ ln p ≈ 3. two limits. In the first scenario, the radiation is balanced Later we show that the equations successfully pre- purely by horizontal advection: zonal winds transport heat dict the mean wind speeds and temperature differences from dayside to nightside, and meridional winds transport obtained in numerical simulations of the circulation of heat from equator to pole. In the second scenario, the radi- HD209458b. This gives us confidence in the results. ation is balanced by vertical advection (ascent on dayside, Now consider the second scenario, where vertical ad- descent on nightside). vection (third term on the left of Eq. (14)) balances the ra- Consider the first scenario, where horizontal advection diative heating and cooling. The magnitude of ω can be es- dominates. Generally, we expect v and ∇HT to point in timated from the continuity equation. Purely geostrophic different directions, so to order of magnitude their dot flow has zero horizontal divergence, so ω is roughly product equals the product of their magnitudes. (If the |v| direction of ∇HT is independent of height and no deep ω ∼ Ro p (19) barotropic flow exists, then at mid-latitudes one could ar- R gue that winds and horizontal pressure gradients are per- Strictly speaking, this is an upper limit, because the non- pendicular to order Ro. However, the existence of either linear terms comprising the numerator of the Rossby num- asynchronous rotation or variations in the orientations of ber contain some terms (e.g., the centripetal accelera- ∇ T with height will imply that winds and ∇ T are not H H tion) that are divergence-free. Substituting this expres- perpendicular even if Ro << 1. Because asynchronous ro- sion into the energy equation, using Eq. (13), and setting tation and variation in the directions of ∇ T with height H Ro = |v|/fR implies that are likely, and because, in any case, v and ∇HT will not be perpendicular near the equator, we write the dot product 1/2 R q as the product of the magnitudes.) An order-of-magnitude |v|∼ fR (20) form of the energy equation is then NH cp 

2 1/2 |v|∆Thoriz q fR q f ∼ (15) ∆Thoriz ∼ (21) R cp NH∆ ln p cp R

The solutions are Numerical estimates using H = 700 km, N ∼ −1 −4 −3 −1 0.0015 sec , and q/cp = 10 –10 Ksec (appropriate q R∆ ln p 1/2 to 50–100 bars) yield temperature differences and speeds |v|∼ (16) −1  cp f  of 80–250K and 300–900msec , respectively. Here we A.P. Showman and T. Guillot: Dynamics of Pegasi-planet Atmospheres 9 have used the same values for R and f as before. A heating lie equatorward of the geostrophic flows that exist at more rate of 10−2 Ksec−1, appropriate at the 1 bar level, yields poleward latitudes. −1 |v|∼ 3000msec and ∆Thoriz ∼ 800K. These values are similar to those obtained when we balanced heating solely 4.3. Numerical Simulations of the Circulation against horizontal advection. We can estimate the vertical velocity under the as- To better constrain the nature of the circulation, we sumption that vertical advection balances the heating. Ex- performed preliminary three-dimensional, fully-nonlinear pressing the vertical velocity w as the time derivative of numerical simulations of the atmospheric circulation of an air parcel’s altitude, and using w = −ω/ρg, where ρ is HD209458b. For the calculations, we used the Explicit density, yields Planetary Isentropic Coordinate, or EPIC, model (Dowl- ing et al. 1998). The model solves the primitive equations q R in spherical geometry using finite-difference methods and w ∼ 2 (22) cp HN isentropic vertical coordinates. The equations are valid in stably-stratified atmospheres, and we solved the equations −2 −1 Using q/cp = 10 Ksec and H = 700 km implies that within the radiative layer from 0.01 to 100 bars assuming − w ∼ 20msec 1 near 1 bar. If this motion comprises the the planet’s interior is in synchronous rotation with the vertical branch of an overturning circulation that extends 3.5-day orbital period. The radius, surface gravity, and vertically over a scale height and horizontally over ∼ 1010 rotation rate of HD209458b were used (108 m, 10msec−2, cm (a planetary radius), the implied horizontal speed re- and Ω=2.1 × 10−5 sec−1, respectively). − quired to satisfy continuity is ∼ 3000msec 1, consistent The intense insolation was parameterized with a sim- with earlier estimates. ple Newtonian heating scheme, which relaxes the temper- The numerical estimates, while rough, suggest that ature toward an assumed radiative-equilibrium tempera- − winds could approach the upper limit of ∼ 3000msec 1 ture profile. The chosen radiative-equilibrium temperature implied by the shear instability criterion. This compari- profile was hottest at the substellar point (0◦ latitude, 0◦ son suggests that Kelvin-Helmholtz shear instabilities may longitude) and decreased toward the nightside. At the sub- play an important role in the dynamics. stellar point, the profile’s height-dependence was isother- The likelihood of strong temperature contrasts can mal at 550 K at pressures less than 0.03 bars and had also be seen from energetic considerations. The differ- constant Brunt-Vaisala frequency of 0.003 sec−1 at pres- ential stellar heating produces available potential energy sures exceeding 0.03 bars, implying that the temperature (i.e., potential energy that can be converted to kinetic increased with depth. The nightside radiative-equilibrium energy through rearrangement of the fluid; Peixoto and temperature profile was equal to the substellar profile mi- Oort 1992, pp. 365-370), and in steady-state, this po- nus 100 K; it is this day-night difference that drives all the tential energy must be converted to kinetic energy at dynamics in the simulation. The nightside profile was con- the rate it is produced. This requires pressure gradi- stant across the nightside, and the dayside profile varied ents, which only exist in the presence of lateral ther- as mal gradients. A crude estimate suggests that the rate of change of the difference in gravitational potential Φ Tdayside = Tnightside + ∆Trad cos α (23) between the dayside and nightside on isobars caused by the heating is ∼ R(q/cp)∆ ln p. The rate per mass at where Tdayside is the dayside radiative-equilibrium tem- which potential energy is converted to kinetic energy by perature at a given latitude and longitude, Tnightside is pressure-gradient work is v·∇H Φ, which is approximately the nightside radiative-equilibrium temperature, ∆Trad = |v|R∆Thoriz∆ ln p/R. Equating the two expressions sug- 100K, and α is the angle between local vertical and the gests, crudely, that ∆Thoriz ∼ (q/cp)(R/|v|), which implies line-of-sight to the star. For simplicity, the timescale over ∆Thoriz ∼ 500 K using the values of q/cp and |v| near 1 which the temperature relaxes to the radiative-equilibrium bar discussed earlier. temperature was assumed constant with depth with a One could wonder whether a possible flow geometry value of 3×105 sec. This is equal to the expected radiative consists of dayside , nightside , and timescale at a pressure of about 5 bars (Fig. 4, left). simple acceleration of the flow from dayside to nightside The Newtonian heating scheme described above is, elsewhere (leading to a flow symmetrical about the sub- of course, a simplification. The day-night difference in solar point). The low-Rossby-number considerations de- radiative-equilibrium temperature, 100 K, is smaller than scribed above argue against this scenario. A major com- the expected value. Furthermore, the radiative timescale ponent of the flow must be perpendicular to horizontal is too long in the upper (∼ 0.1–1bar) and too pressure gradients. Direct acceleration of wind from day- short at deeper pressures of ∼ 10–100 bars. Nevertheless, side to nightside is still possible near the equator, how- the net column-integrated heating per area produced by ever. This would essentially be an equatorially-confined the scheme ( W m−2) is similar to that expected to occur Walker-type circulation that, as described earlier, would in the deep troposphere of HD209458b and other Pegasi 10 A.P. Showman and T. Guillot: Dynamics of Pegasi-planet Atmospheres planets, and the simulation provides insight into the cir- culation patterns that can be expected. The temperature used as the initial condition was isothermal at 500 K at pressures less than 0.03 bars and had a constant Brunt-Vaisala frequency of 0.003 sec−1 at pressures exceeding 0.03 bars. There were no initial winds. The simulations were performed with a horizontal resolu- tion of 64×32 with 10 layers evenly spaced in log-pressure. Figure 5–8 show the results of such a simulation. De- spite the motionless initial condition, winds rapidly de- velop in response to the day-night heating contrast, reach- ing an approximate steady state after ∼ 400 Earth days. Snapshots at 42 and 466 days are shown in Figs. 5 and 6, respectively. In these figures, each panel shows pressure on an isentrope (greyscale) and winds (vectors) for three of the model layers corresponding to mean pressures of roughly 0.4, 6, and 100 bars (top to bottom, respectively). The greyscale is such that, on an isobar, light regions are hot and dark regions are cold. The simulation exhibits several interesting features. First, peak winds exceed 1kmsec−1, but despite these winds, a horizontal temperature contrast is maintained. The dayside (longitudes −90◦ to 90◦) is on average hotter than the nightside, but dynamics distorts the temperature pattern in a complicated manner. Second, an equatorial jet develops that contains most of the kinetic energy. The jet initially exhibits both eastward and westward branches (Fig. 5) but eventually becomes only eastward (Fig. 6), ◦ ◦ and extends from −30 to 30 in latitude. Third, away Fig. 5. EPIC simulation of atmospheric circulation on from the equator, winds develop that tend to skirt paral- HD209458b after 42 days. Panels depict pressure on an lel to the temperature contours. This is an indication that isentrope (greyscale) and winds (vectors) for three model geostrophic balance holds. (Because the thermal structure levels with mean pressures near 0.4, 6, and 100 bars (top to is independent of height to zeroeth order and we have as- bottom, respectively). From top to bottom, the maximum sumed no deep barotropic flow, horizontal pressure and wind speeds are 937, 688, and 224msec−1, respectively, temperature gradients are parallel.) Nevertheless, winds and the greyscales span (from dark to white) 0.31–0.49 are able to cross isotherms near the equator, and this is bars, 5.6–7.8 bars, and 92–113 bars, respectively. Substel- important in setting the day-night temperature contrast. lar point is at 0◦ latitude, 0◦ longitude. As expected from the order-of-magnitude arguments in Section 4.1, the jets and gyres that exist are broad in scale, with a characteristic width of the planetary radius. In the simulation, the day-night temperature differ- and in front of the star, respectively. On the other hand, ence (measured on isobars) is about 50K. This value de- if a broad westward jet existed instead, the maximum pends on the adopted heating rate, which was chosen to and minimum temperatures would face Earth after the be appropriate to the region where the pressure is tens transits. Therefore, an infrared lightcurve of the planet of bars. Simulations that accurately predict the day-night throughout its orbital cycle would help determine the di- temperature difference at 1 bar will require a more realis- rection and strength of the atmospheric winds. tic heating-rate scheme; we will present such simulations In the simulation, the intense heating and cooling in a future paper. causes air to change entropy and leads to vertical mo- The temperature patterns in Figs. 5 and 6 show that tion, as shown in Fig. 7. Light regions (Fig. 7) indicate Earth-based infrared measurements can shed light on the heating, which causes ascent, and dark regions indicate circulation of Pegasi planets. In Fig. 6, the superrotating cooling, which causes descent. Air is thus exchanged be- equatorial jet blows the high-temperature region down- tween model layers. This exchange also occurs across the wind. The highest-temperature region is thus not at the model’s lowermost isentrope, which separates the radia- substellar point but lies eastward by about 60◦ in longi- tive layer from the convective interior (Fig. 7, bottom). tude. The maximum and minimum temperatures would Because upgoing and downgoing air generally have differ- thus face Earth before the transit of the planet behind ent kinetic energies and momenta, energy and momentum A.P. Showman and T. Guillot: Dynamics of Pegasi-planet Atmospheres 11

Fig. 6. Simulation results for HD209458b at 466 days (af- Fig. 7. Additional simulation results at 466 days. Ar- ter a steady state has been reached). As in Fig. 5, panels rows are identical to those in Fig. 6, but here greyscale depict pressure on isentropes (greyscale) and winds (vec- is vertical velocity dθ/dt, where θ is potential tempera- tors) for three model levels with mean pressures near 0.4, ture. Light regions are heating (i.e., θ is increasing) and 6, and 100 bars (top to bottom, respectively). From top dark regions are cooling (θ is decreasing). From top to to bottom, the maximum wind speeds are 1541, 1223, and bottom, the greyscales span (from dark to white) −0.003 598msec−1, respectively, and the greyscales span (from to 0.003Ksec−1, −0.002 to 0.002Ksec−1, and −0.001 to dark to white) 0.32–0.51 bars, 5.6–8.1 bars, and 90–127 0.001Ksec−1, respectively. The lowermost panel separates bars, respectively. Substellar point is at 0◦ latitude, 0◦ the convective interior from the radiative region, and the longitude. implication is that mass exchange can happen across this interface.

can thus be exchanged between the atmosphere and inte- In the simulation, kinetic energy is transported from rior. the atmosphere into the interior at a rate that reaches Figure 8 indicates how the steady state is achieved. 2500Wm−2 (Fig. 8, bottom). This downward transport of The build-up of the mass-weighted mean speed (top panel) kinetic energy is ∼ 1% of the absorbed stellar flux and is involves two timescales. Over the first 10 days, the mean great enough to affect the radius of HD209458b, as shown winds accelerate to 110msec−1, and the peak speeds ap- in Paper I. Although the simulation described here does proach 1kmsec−1. This is the timescale for pressure gra- not determine the kinetic energy’s fate once it reaches the dients to accelerate the winds and force balances to be es- convective interior, we expect that tidal friction, Kelvin- tablished. The flow then undergoes an additional, slower Helmholtz instabilities, or other processes could convert (∼ 300 day) increase in mean speed to 300msec−1, with these winds to thermal energy. The exact kinetic energy peak speeds of ∼ 1.5kmsec−1. This timescale is that flux will depend on whether a deep barotropic flow ex- for exchange of momentum with the interior, across the ists. Furthermore, potential and thermal energy are also model’s lowermost isentrope, to reach steady state, as transported through the boundary, and pressure work is shown in Fig. 8, middle. done across it. Our aim here is not to present detailed di- 12 A.P. Showman and T. Guillot: Dynamics of Pegasi-planet Atmospheres agnostics of the energetics, but simply to point out that energy fluxes that are large enough to be important can occur. We are currently conducting more detailed simu- lations to determine the sensitivity of the simulations to a deep barotropic flow, and we will present the detailed energetics of these simulations in a future paper. The evolution of angular momentum (Fig. 8, middle) helps explain why the circulation changes from Fig. 5 to 6. After 42 days (Fig. 5), eastward midlatitude winds have already developed that balance the negative equator-to- pole temperature gradient. But at this time, the atmo- sphere’s angular momentum is still nearly zero relative to the synchronously-rotating state, so angular momen- tum balance requires westward winds along some parts of the equator. (Interestingly, in these regions, the temper- ature increases with latitude, as expected from geostro- phy.) The circulation involves upwelling on the dayside and equatorial flow — both east and west — to the night- side, where downwelling occurs. This circulation is similar in some respects to the in Earth’s at- mosphere. After 466 days, however, the atmosphere has gained enough angular momentum that the equatorial jet is fully superrotating. Some westward winds exist at high latitudes, but they are weak enough that the net angular momentum is still eastward. Application of the arguments from Section 4.2 pro- vides a consistency check. The mean heating rate in the simulations is about 1.7 × 10−4 Ksec−1. Inserting this value into Eqs. (16–17), we obtain ∆Thoriz = 70K and |v| = 240msec−1. These compare well with the day-night temperature difference and mass-weighted mean speed of Fig. 8. Top: Mass-weighted mean wind speed within 50K and 300msec−1 that are obtained in the simulation. the model domain (∼ 0.01–100 bar) during the sim- (For the estimate, we used f = 3 × 10−5 sec−1, R = ulation of HD209458b shown in Figs. 5–7. Middle: 3500Jkg−1 K−1, R = 108 m, and ∆p = 3.) If Eqs. (20– Globally-averaged zonal angular momentum per area mi- 21) are adopted instead, using values of N = 0.003 sec−1 nus that of the synchronously-rotating state. (That is, and H = 300 km, we obtain temperature differences and RA−1 u cos φ dm, where A is the planet’s area, u is east- − speeds of 130K and 470msec 1. Compared to the sim- ward speed,R φ is latitude, and dm is an atmospheric mass ulation, these estimates are too high by a factor of ∼ 2, element.) Bottom: Globally-averaged flux of kinetic en- but are still of the correct order of magnitude. An impor- ergy across the model’s bottom isentrope (at ∼ 100 bars), tant point is that the maximum speed in the simulation which is the interface between the radiative layer and the substantially exceeds that from the simple estimates; the convective interior in this model. estimates are most relevant for the mean speed.

4.4. Vertical motion and clouds ing and nightside cooling are balanced only by horizontal As shown in the appendix, hot and cold regions on the advection, air flowing from dayside to nightside remains planet may have distinct chemical equilibrium composi- at constant pressure but decreases in temperature (hence tions. Advection of air between these regions plays a key entropy) because of the radiative cooling. Nightside con- role for the (disequilibrium) chemistry in the atmosphere. densation may occur, and particle fallout limits the total Cloud formation critically depends on whether vertical mass of the trace constituent (vapor plus condensates) in motions dominate over horizontal motions. This in turns this air. As the air flows onto the dayside, its temperature affects the albedo and depth to which stellar light is ab- increases and any existing clouds will sublimate. This sce- sorbed. nario therefore implies that the dayside will be cloud-free. Clouds form when the temperature of moving air In an alternate scenario, dayside heating and nightside parcels decreases enough for the vapor pressure of con- cooling are balanced by vertical advection, at least near densible vapors to become supersaturated. If dayside heat- the sub- and antisolar points. Because entropy increases A.P. Showman and T. Guillot: Dynamics of Pegasi-planet Atmospheres 13 with altitude in a stably-stratified atmosphere, this sce- last throughout the star’s lifetime. Nevertheless, to fully nario implies that ascent occurs on the dayside and de- determine the feasibility of the mechanism, more detailed scent on the nightside. Clouds may therefore form on the numerical simulations are required (in particular, to test dayside, increasing the albedo and decreasing the pressure the dependence of the energetics on the possible existence at which stellar light is absorbed. of winds in the interior). We will present such simulations In reality, both horizontal and vertical advection are in a future paper. important, and the real issue is to determine the rela- Upcoming observations are likely to provide key in- tive contribution. We are pursuing more detailed numeri- formation within the decade. Several spacecraft missions cal simulations to address this issue. (either proposed or accepted) and dedicated ground pro- grams will observe extrasolar planets. Measurement of starlight reflected from these planets may allow the albedo 5. Conclusions to be estimated. Because the star-planet-Earth angle We examined the structure and dynamics of the atmo- changes throughout the planet’s orbit, crude informa- spheres of Pegasi planets and concluded that strong day- tion on the scattering properties of the atmosphere (e.g., night temperature contrasts (300K or more) are likely to isotropic versus forward scattering) may be obtainable. occur near the level where radiation is emitted to space Asymmetries in the reflected flux as the planet approaches (∼ 1 bar). These temperature variations should drive a and recedes from the transit could give information on rapid circulation with peak winds of 1 km sec−1 or more. the differences of albedo near the leading and trailing ter- Force-balance arguments suggest that the mean midlati- minators, which would help constrain the dynamics. Fi- tude winds are eastward, but the equatorial winds could nally, transit observations of Pegasi planets using high blow either east or west. Depending on the dynamical resolution spectroscopy should in the near future yield regime, the cloud coverage and consequently the radia- constraints on the atmospheric temperature, cloud/haze tive absorption of the incoming stellar flux will be very abundance, and winds (Brown 2001; see also Seager & different. Sasselov 2000; Hubbard et al. 2001). If these measure- Preliminary numerical simulations show that, while ments are possible during the ingress and egress, i.e., the the dayside is generally hotter than the nightside, the spe- phases during which the planets enters and leaves the stel- cific temperature distribution depends on the dynamics. lar limb, respectively, asymmetries of the planetary signal In our simulations, a broad superrotating jet develops that should be expected and would indicate zonal heat advec- sweeps the high-temperature regions downwind by about tion at the terminator. The duration of these phases being 1 radian. This implies that the greatest infrared flux would limited to less than 10 minutes, it is not clear that this ef- reach Earth ∼ 10–15 hours before the occultation of the fect is observable with current instruments. planet behind the star. On the other hand, if a subro- tating jet existed, the greatest infrared flux would reach Acknowledgements. We wish to thank F. Allard, P. Boden- heimer, H. Houben, S. Peale, D. Saumon, D.J. Stevenson, R.E. Earth after the occultation. Measurements of the infrared Young and K. Zahnle for a variety of useful contributions, and lightcurve of Pegasi planets will therefore constrain the T. Barman for sharing results in advance of publication. This direction and magnitude of the wind. research was supported by the French Programme National The simulations also produce a downward flux of ki- de Plan´etologie, Institute of Theoretical Physics (NSF PH94- netic energy across the ∼ 100 bar surface equal to ∼ 1% 07194), and National Research Council of the United States. of the absorbed stellar flux. Although the simulation did not explicitly include the convective interior, we surmise Appendix A: Day-night temperature variations that a substantial fraction of this kinetic energy would and chemical composition be converted to thermal energy by Kelvin-Helmholtz in- stabilities and other processes. As discussed in Paper I, Here we explore the implications of day-night temperature “standard” evolution models explain HD209458b’s radius variations on the abundances of a condensing species. Sup- only when the atmosphere is assumed to be unrealisti- pose that on the day side, a given condensate (e.g. iron) ⋆ cally hot, but addition of an internal energy source allows condenses at a pressure level Pday and a corresponding a more realistic model to match the observed radius. The temperature Tday. If the temperature on the night side ⋆ downward transport and subsequent dissipation of kinetic Tnight is smaller, at what pressure Pnight will condensa- energy described here is a promising candidate. Boden- tion take place? heimer et al. (2001) suggested an alternative — that the Assuming an ideal gas, we write the Clausius- internal heating could be provided by tidal circularization Clapeyron equation as of an initially eccentric orbit. The difficulty, as Boden- d ln p heimer et al. were careful to point out, is that the tidal = β (A.1) heating is a transient process in the absence of a detected d ln T close, massive companion capable of exciting the planet’s where p is the partial pressure of saturation of the con- eccentricity. In contrast, the mechanism proposed here can densing species and β = L/RT is the ratio of the latent 14 A.P. Showman and T. Guillot: Dynamics of Pegasi-planet Atmospheres heat of condensation L to the thermal energy RT . For Burrows, A., Marley, M., Hubbard, W.B., Lunine, J.I., Guillot, most condensing species of significance here, β ≈ 10. We T., Saumon, D., Freedman, R., Sudarsky, D., Sharp, C., furthermore assume that β is independent of T and P , 1997, ApJ, 491, 856. which introduces only a slight error in our estimates. Chandrasekhar, S. 1961, Hydrodynamic and Hydromagnetic By definition, on the day side, the saturation abun- Stability, Clarendon Press, Oxford, p. 491 dance of the condensing species, x = p/P is maximal and Charbonneau D., Brown T.M., Latham D.W., Mayor, M., 2000, ApJ 529, L45 equal to x⋆ at P = P ⋆ . On the other hand, the night day Cho J.Y-K., Polvani L.M., 1996, Science 273, 335 side temperature is lower and the abundance becomes: Dermott, S.F. Icarus, 37, 310 ln x(P ⋆ ) = ln x⋆ − β ln(T /T ). (A.2) Dowling, T.E., Fischer, A.S., Gierasch, P.J., Harrington, J., day day night LeBeau, R.P. Jr., Santori, C.M., 1998, Icarus 132, 221 In order to reach condensation, i.e. x = x⋆, one has to Geballe T.R., Saumon D., Leggett S.K., Knapp G.R., Marley penetrate deeper into the atmosphere. Eq. (A.1) implies M.S., Lodders K. 2001, ApJ, 556, 373 that Gill A.E., 1982, Atmosphere- Dynamics Academic Press, New York d ln x Gold T., Soter S. 1969, Icarus, 11, 356. = β∇ − 1, (A.3) d ln P T Goldreich P., Soter S. 1966, Icarus, 5, 375 Goukenleuque C., B´ezard B., Joguet B., Lellouch E., Freedman and hence, on the night side, R., 2000, Icarus 143, 308 ⋆ ⋆ Guillot T., Showman A.P. 2001, submitted to A&A (Paper I) ln x(P ) = ln x(P ) + (β∇T − 1) ln(P/P ), (A.4) day day Guillot T., Burrows A., Hubbard W.B., Lunine J.I., Saumon assuming that ∇T is constant. Using eq. (A.2), one obtains D., 1996, ApJ 459, L35 the condensation pressure on the night side: Henry G.W., Marcy G.W., Butler R.P., Vogt, S.S., 2000, ApJ 529, L41-L44 ∇ − P ⋆ T β/(β T 1) Holton J.R., 1992, An Introduction to Dynamic Meteorology, night = day . (A.5) P ⋆ T  Academic Press, New York. day night Houben H., Young R.E., Showman A., Mosqueira I. 2001, Jupiter: Planet, Satellites, & Magnetosphere [abstract for Using β ∼ 10, ∇T ∼ 0.15 and Tday/Tnight ∼ 1.2, one finds P ⋆ ∼ 38P ⋆ , a very significant variation of the conference held in Boulder, CO, June 25-30 2001], 51. night day Hubbard W.B. 1984. Planetary Interiors. New York: Van Nos- condensation pressure. This implies that air flowing on trand Reinhold Co., Inc. constant pressure levels around the planet would lead to Hubbard W.B., Fortney J.J., Lunine J.I., Burrows A., Su- a rapid depletion of any condensing species on the day darsky D., Pinto P.A. 2001, submitted to ApJ side, compared to what would be predicted from chemical Ingersoll A.P., Barnet C.D., Beebe R.F. et al., 1995, in: Nep- equilibrium calculations. This can potentially also remove tune and Triton, ed. D.P. Cruikshank, University of Ari- important absorbing gases from the day side, as in the case zona Press, Tucson, 613 of TiO, which can be removed by CaTiO3 condensation, Ingersoll A.P., Dobrovolskis A.R., 1978, Nature 275, 37 or Na, removed by Na2S condensation (Lodders 1999). Of Ioannou P.J., Lindzen R.S. 1993, ApJ, 406, 266 course, most of the variation depends on the exponential Liebert J., Reid I.N., Burrows A., Burgasser A.J., Kirkpatrick J.D., Gizis J.E. 2000, ApJL 533, L155 factor β/(β∇T − 1), which is infinite in the limit when the atmospheric temperature profile and the condensation Lodders K. 1999, ApJ, 519, 793 Lubow S.H., Tout C.A., Livio M. 1997, ApJ, 484, 866 profile are parallel to each other. Maeder A., 1995, A&A, 299, 84 In the discussion, we implicitly assumed β∇T − 1 > 0; Marcy G.W., Butler R.P., Williams E., Bildsten L., Graham however, when the atmosphere is close to an isotherm, this J.R., Ghez A.M., Jernigan J.G., 1997, ApJ, 481, 926 factor can become negative. In this case the day/night Marley M.S., Saumon D., Guillot T., et al., 1996, Science 272, effect is even more severe, as the condensing species is 1919 entirely removed from this quasi-isothermal region. Peale S.J., 1999, Ann. Rev. Astron. Astrophys., 37, 533 Peixoto J.P., Oort A.H., 1992, Physics of Climate, American Institute of Physics, New York. References Rhines P.B., 1975, J. Fluid. Mech. 69, 417 Allard F., Hauschildt P.H., Alexander D.R., Starrfield S. 1997, Schweitzer A., Gizis J.E., Hauschildt P.H., Allard F., Reid I.N. ARA&A, 35, 137 2001, ApJ, 555, 368 Andrews D.G., Holton J.R., Leovy C.B., 1987, Middle Atmo- Seager S., Sasselov D.D., 1998, ApJ 502, L157 sphere Dynamics Academic Press, New York Seager S., Sasselov D.D. 2000, ApJ, 537, 916 Banfield, D., Murray, N. 1992, Icarus, 99, 390 Tittemore, W.C., Wisdom, J. 1989, Icarus, 78, 63 Barman T., Hauschildt P.H., Allard F. 2001, ApJ, 556, 885 Zahn J.-P., 1992, A&A, 265, 115 Bodenheimer P., Lin D.N.C., Mardling R. 2001, ApJ, 548, 466 Brown T.M. 2001, ApJ, 553, 1006 Burrows A., Guillot T., Hubbard W.B., Marley M.S., Saumon D., Lunine J.I., Sudarsky D. 2000, ApJL, 534, L97