arxiv: v1 [astro-ph.ep] 23 Nov 2011

Similar documents
arxiv: v1 [astro-ph.ep] 11 Aug 2016

Dynamical properties of the Solar System. Second Kepler s Law. Dynamics of planetary orbits. ν: true anomaly

Observational Cosmology Journal Club

arxiv: v3 [astro-ph.ep] 20 Aug 2015

Secular Orbital Evolution of Compact Planet Systems

TIDAL EVOLUTION OF CLOSE-IN SATELLITES AND EXOPLANETS Lecture 1. Darwin theory

Eccentricity pumping of a planet on an inclined orbit by a disc

arxiv: v1 [astro-ph.ep] 3 Apr 2018

Secular Planetary Dynamics: Kozai, Spin Dynamics and Chaos

arxiv: v1 [astro-ph.ep] 8 Feb 2017

Tides and Lagrange Points

Kozai-Lidov oscillations

The rotation of extra-solar planets

Tidal evolution of the Kepler-10 system

DYNAMICS OF THE LUNAR SPIN AXIS

Lecture 13. Gravity in the Solar System

Supplementary Materials for

TIDAL EVOLUTION OF PLANETARY SYSTEMS

Dynamical Stability of Terrestrial and Giant Planets in the HD Planetary System

Possible commensurabilities among pairs of extrasolar planets

arxiv: v1 [astro-ph.ep] 12 Dec 2016

Dynamics of the conservative and dissipative. spin orbit problem

Evolution of Mercury s obliquity

AST111, Lecture 1b. Measurements of bodies in the solar system (overview continued) Orbital elements

Planetary system dynamics. Planetary migration Kozai resonance Apsidal resonance and secular theories Mean motion resonances Gravitational scattering

arxiv:astro-ph/ v1 17 Nov 2004

arxiv: v1 [astro-ph.ep] 22 Nov 2018

What can be learned from the dynamics of packed planetary systems?

Equation of orbital velocity: v 2 =GM(2/r 1/a) where: G is the gravitational constant (G=6.67x10 11 N/m 3 kg), M is the mass of the sun (or central

EART162: PLANETARY INTERIORS

Symplectic Correctors for Canonical Heliocentric N-Body Maps

Accretion of Phobos and Deimos in an extended debris disc stirred by transient moons

Theory of mean motion resonances.

Tidal Dissipation in Binaries

1 General Definitions and Numerical Standards

Resonance Capture. Alice Quillen. University of Rochester. Isaac Newton Institute Dec 2009

Can we see evidence of post-glacial geoidal adjustment in the current slowing rate of rotation of the Earth?

arxiv: v1 [astro-ph] 30 May 2007

arxiv: v1 [astro-ph] 22 Feb 2008

Rotation and Interior of Terrestrial Planets

THE PLANE OF THE KUIPER BELT

Orbital Evolution in Extra-solar systems

Data from: The Extrasolar Planet Encyclopaedia.

Short-period planetary systems and their mysteries

Why is the Moon synchronously rotating?

arxiv: v2 [astro-ph.ep] 28 Sep 2017

arxiv: v1 [astro-ph.ep] 30 Aug 2018

Dynamical Tides in Binaries

Geodesy and interior structure of Mercury

Astronomy 6570 Physics of the Planets

Exoplanets: a dynamic field

ORBITS WRITTEN Q.E. (June 2012) Each of the five problems is valued at 20 points. (Total for exam: 100 points)

arxiv:astro-ph/ v2 5 Aug 1997

Pervasive Orbital Eccentricities Dictate the Habitability of Extrasolar Earths

Dynamic Exoplanets. Alexander James Mustill

Basics of Kepler and Newton. Orbits of the planets, moons,

Formation and evolution of the two 4/3 resonant giants planets in HD

Third Body Perturbation

A = 6561 times greater. B. 81 times greater. C. equally strong. D. 1/81 as great. E. (1/81) 2 = 1/6561 as great Pearson Education, Inc.

Observational Astronomy - Lecture 4 Orbits, Motions, Kepler s and Newton s Laws

arxiv: v1 [astro-ph.ep] 7 Feb 2017

Chaotic diffusion in the Solar System

Habitability in the Upsilon Andromedae System

arxiv: v1 [astro-ph.ep] 20 Nov 2018

arxiv: v1 [astro-ph] 24 Dec 2008

The dynamics of Plutinos

Physics Mechanics Lecture 30 Gravitational Energy

Information on internal structure from shape, gravity field and rotation

arxiv: v1 [astro-ph.ep] 23 Aug 2012

arxiv: v1 [astro-ph] 23 May 2007

Gravitation and the Motion of the Planets

UNIVERSITY of CALIFORNIA SANTA CRUZ

arxiv: v1 [astro-ph.ep] 20 Mar 2016

Is Solar Activity Modulated by Astronomical Cycles? Leif Svalgaard HEPL, Stanford University SH34B-08, Dec. 7, 2011

ANNEX 1. DEFINITION OF ORBITAL PARAMETERS AND IMPORTANT CONCEPTS OF CELESTIAL MECHANICS

Celestial Mechanics of Asteroid Systems

Date of delivery: 29 June 2011 Journal and vol/article ref: IAU Number of pages (not including this page): 5

Resonance and chaos ASTRONOMY AND ASTROPHYSICS. II. Exterior resonances and asymmetric libration

Tidal dissipation in rotating low-mass stars hosting planetary or stellar companions. From BinaMIcS to SPIRou

arxiv: v2 [astro-ph.ep] 22 Jul 2013

Dynamics of the Earth

Introduction To Modern Astronomy I

What is it like? When did it form? How did it form. The Solar System. Fall, 2005 Astronomy 110 1

Tides, planetary companions, and habitability: habitability in the habitable zone of low-mass stars

A REGION VOID OF IRREGULAR SATELLITES AROUND JUPITER

arxiv: v1 [astro-ph.ep] 18 Mar 2010

A dwarf planet is a celestial body that (a) is in orbit around the Sun, (b) has sufficient mass for its selfgravity

Frequency map analysis of the 3/1 resonance between planets b and c in the 55 Cancri system ABSTRACT

APPENDIX B SUMMARY OF ORBITAL MECHANICS RELEVANT TO REMOTE SENSING

Gravitation and the Waltz of the Planets

Gravitation and the Waltz of the Planets. Chapter Four

Predictions for a planet just inside Fomalhaut s eccentric ring

Dynamically Unstable Planetary Systems Emerging Out of Gas Disks

arxiv:astro-ph/ v1 17 Jul 2003

arxiv: v1 [astro-ph.ep] 6 Mar 2018

EPM the high-precision planetary ephemerides of IAA RAS for scientific research, astronavigation on the Earth and space

FORMING DIFFERENT PLANETARY ARCHITECTURES. I. FORMATION EFFICIENCY OF HOT JUPITES FROM HIGH-ECCENTRICITY MECHANISMS

The primordial excitation and clearing of the asteroid belt Revisited

ECCENTRICITY EXCITATION AND APSIDAL RESONANCE CAPTURE IN THE PLANETARY SYSTEM ANDROMEDAE E. I. Chiang 1 and N. Murray 2

Probing planetary interiors by spacecraft orbital observations

Transcription:

Draft version September 3, 18 Preprint typeset using L A TEX style emulateapj v. 11/1/9 PUMPING THE ECCENTRICITY OF EXOPLANETS BY TIDAL EFFECT Alexandre C.M. Correia Department of Physics, I3N, University of Aveiro, Campus de Santiago, 381-193 Aveiro, Portugal; and Astronomie et Systèmes Dynamiques, IMCCE-CNRS UMR88, 77 Av. Denfert-Rochereau, 7514 Paris, France arxiv:1111.5486v1 [astro-ph.ep] 3 Nov 11 Gwenaël Boué Centro de Astrofísica, Universidade do Porto, Rua das Estrelas, 415-76 Porto, Portugal; and Astronomie et Systèmes Dynamiques, IMCCE-CNRS UMR88, 77 Av. Denfert-Rochereau, 7514 Paris, France and Jacques Laskar Astronomie et Systèmes Dynamiques, IMCCE-CNRS UMR88, 77 Av. Denfert-Rochereau, 7514 Paris, France Draft version September 3, 18 ABSTRACT Planets close to their host stars are believed to undergo significant tidal interactions, leading to a progressive damping of the orbital eccentricity. Here we show that, when the orbit of the planet is excited by an outer companion, tidal effects combined with gravitational interactions may give rise to a secular increasing drift on the eccentricity. As long as this secular drift counterbalances the damping effect, the eccentricity can increase to high values. This mechanism may explain why some of the moderate close-in exoplanets are observed with substantial eccentricity values. Subject headings: celestial mechanics planetary systems planets and satellites: general 1. INTRODUCTION Close-in exoplanets, as for Mercury, Venus and the majority of the natural satellites in the Solar System, are supposed to undergo significant tidal interactions, resulting that their spins and orbits are slowly modified. The ultimate stage for tidal evolution is the synchronization of the spin and the circularization of the orbit (e.g. Correia 9). The spin evolves in a shorter time-scale than the orbit, so long-term studies on the tidal evolution of exoplanets usually assume that their rotation is synchronously locked, and therefore limit the evolution to the orbits. However, these two kinds of evolution cannot be dissociated because the total angular momentum must be conserved. Synchronous rotation can only occur when the eccentricity is very close to zero. Otherwise, the rotation rate tends to be locked with the orbital speed at the periapsis, because tidal effects are stronger when the two bodies are closer to each other. In addition, in presence of a companion body, the eccentricity undergoes oscillations (e.g. Mardling 7), and the rotation rate of the planet shows variations that follow the eccentricity (Correia & Laskar 4). As a consequence, some unexpected behaviors can be observed, such as a secular increase of the eccentricity. In this Letter we provide a simple averaged model for the orbital and spin evolution of an exoplanet with a companion (Sect. ), and apply it to the HD 117618 planetary system (Sect. 3). We then give an explanation for the eccentricity pumping (Sect. 4), and derive some conclusions (Sect. 5).. THE MODEL We consider here a system consisting of a central star of mass m, an inner planet of mass, and an outer companion of mass m. We use Jacobi canonical coordicorreia@ua.pt nates, with r 1 being the position of relative to m, and r the position of m relative to the center of mass of and m. We further assume that r 1 r, that the system is coplanar, and that the obliquity of the planet is zero. The inner planet is considered an oblate ellipsoid with gravity field coefficients given by J, rotating about the axis of maximal inertia, with rotation rate ω, such that (e.g. Lambeck 1988) J = k ω R 3 3G. (1) G is the gravitational constant, R is the radius of the planet, and k is the second Love number for potential. Since we are interested in the secular behavior of the system, we average the motion equations over the mean anomalies of both orbits. The averaged potential, quadrupole-level for the spin (e.g. Correia & Laskar 1a), octopole-level for the orbits (e.g. Lee & Peale 3; Laskar & Boué 1), and with general relativity corrections (e.g. Touma et al. 9) is given by: where U = C ( 1) 1/ C 1 ( 1) 3/ (1 + 3 C e 1) ( + C e 1 e (1 + 3 4 e 1) )3/ 3 ( cos ϖ, () )5/ C = 3β 1G (m + ) a 1 c, C 1 = Gm J R a 3 1 C = Gβ 1m a 1 4a 3, C 3 = 15Gβ 1m a 3 1 16a 4, (3) (m ) m +. (4) a i is the semi-major axis (that can also be expressed using the mean motion n i ), e i is the eccentricity, and

A.C.M. Correia et al. ϖ = ϖ 1 ϖ is the difference between the longitudes of the periastron, ϖ i. We also have β 1 = m /(m + ), and β = (m + )m /(m + + m ). The contributions to the orbits are easily obtained using the Lagrange planetary equations (e.g. Murray & Dermott 1999): Thus, and ė i = i β i n i a i e i U ϖ i, ϖ i = i β i n i a i e i U e i. (5) ė 1 = ν 31 e (1 + 3/4e 1) 1 ( )5/ sin ϖ, (6) ė = ν 3 e 1 (1 + 3/4e 1) ( ) sin ϖ, (7) ϖ = ν ( 1 ) + ν 1 x ( 1 ) + ν 1 1 ( ν (1 + 3 e 1) )3/ ( ) ν 31 e 1 (1 + 9 4 e 1) e 1 ( )5/ cos ϖ + ν 3 e 1 (1 + 3 4 e 1)(1 + 4e ) e ( )3 cos ϖ, (8) where x = ω/, and the constant frequencies ( n1 a ) 1 ν = 3 c, (9) ( ) 5 k m + R ν 1 =, (1) ( ) 3 3 m a1 ν 1 = 4 m + a, (11) ( ) 3 m a1 ν = n 4 (m + ) a, (1) ( ) 4 15 m m a1 ν 31 = 16 m + m + a, (13) ( ) 3 15 m m a1 ν 3 = n 16 (m + ) m + a. (14) Notice that the variations in e 1 and e (Eqs. 6, 7) are related by the conservation of the total angular momentum (after dividing by n a ): ξ ω n ( R a ) + β 1 n ( a1 a ) 1 + β = Cte, (15) where ξ is a structure coefficient. The conservative system (Eq. ) can thus be reduced to one degree of freedom. In our model, we additionally consider tidal dissipation raised by the central star on the inner planet. The dissipation of the mechanical energy of tides in the planet s interior is responsible for a time delay t between the initial perturbation and the maximal deformation. As the rheology of planets is badly known, the exact dependence of t on the tidal frequency is unknown. Several models exist (for a review see Correia et al. 3; Efroimsky & Williams 9), but for simplicity we adopt here a model with constant t, which can be made linear (Singer 1968; Mignard 1979). The contributions to the equations of motion are given by (e.g. Correia 9): where K = 3k ξq ω = K (f 1 (e 1 )x f (e 1 )), (16) ȧ 1 = K (f (e 1 )x f 3 (e 1 )), (17) ė 1 = 9K ( 11 18 f 4(e 1 )x f 5 (e 1 ) m β 1 m 1 ( ) 3 R, K = K 1/ξ ) e 1, (18) β 1 ( ) R, (19) Q 1 = t, and f 1 (e) = (1 + 3e + 3e 4 /8)/( ) 9/, f (e) = (1+15e /+45e 4 /8+5e 6 /16)/(1 e ) 6, f 3 (e) = (1+31e /+55e 4 /8+185e 6 /16+5e 8 /64)/(1 e ) 15/, f 4 (e) = (1+3e /+e 4 /8)/(1 e ) 5, f 5 (e) = (1+15e /4+ 15e 4 /8 + 5e 6 /64)/( ) 13/. We neglect the effect of tides over the longitude of the periastron, as well as the flatenning of the central star. Their effect is only to add a small supplementary frequency to ϖ 1, similar to the contributions from the general relativity (for a complete model see Correia et al. 11). Under the effect of tides alone, the equilibrium rotation rate, obtained when ω =, is attained for (Eq. 16): ω = f(e 1 ) = f (e 1 ) f 1 (e 1 ) = 1 + 6e 1 + O(e 4 1). () Usually K K, so tidal effects modify the rotation rate much faster than the orbit. It is thus tempting to replace the equilibrium rotation in expressions (17) and (18). With this simplification, one obtains always negative contributions for ȧ 1 and ė 1 (Correia 9), ȧ 1 = 7K f 6 (e 1 )e 1, (1) ė 1 = 7 K f 6 (e 1 )( 1)e 1, () with f 6 (e) = (1+45e /14+8e 4 +685e 6 /4+55e 8 /448+ 5e 1 /179)( ) 15/ /(1 + 3e + 3e 4 /8). Thus, the semi-major axis and the eccentricity can only decrease until the orbit of the planet becomes circular (Fig. 1a). However, planet-planet interactions can produce eccentricity oscillations with a period shorter, or comparable to the damping timescale of the spin. In that case, the expression () is not satisfied and multi-planetary systems may show non-intuitive eccentricity evolutions, such as eccentricity pumping of the inner orbit (e 1 increases while e decreases).

Pumping the eccentricity by tidal effect 3 TABLE 1 Single planetary systems with < <.3 and e 1 >.3. Star e 1 m Age τ (name) (AU) (M J ) (M ) (Gyr) (Gyr) HD 18147 3 1.19..1 CoRoT-1 5 3.75.89 3. HD 3383 45 8.33 1.4 3..34 HD 17156 63 8 3.19 1.8 3.4 4 HIP 575 64.31.3.34 39.4 HD 117618 76 8 1.5 3.9.6 HD 4565.8.38 7.83 93.3 HD 9156.31.6.84 4.4 35.8 HD 3765.74.84.8 1.7 1 HD 3651.84 3..79 5.1 15.5 Note. Data is taken from http://exoplanet.eu/ τ = a 8 1 (1 e 1 )8 t1k Gm R5, with t = 1 s (Correia & Laskar 1b). 3. APPLICATION TO EXOPLANETS As an illustration of the eccentricity pumping, we apply our model to the HD 117618 system. This Sun-like star (m M ) has been reported to host a single Saturnlike planet on a eccentric orbit (Butler et al. 6). The residuals of the best fitted solution to the observational data are 5.5 m/s, so we assume that any additional companion with a doppler shift semi-amplitude smaller than this value is presently undetected, that is, any planet with m <. M J and a > 1.4 AU. In our simulations we adopt for the observed planet the same geophysical parameters as for Saturn, R = R Sat, k = 1/, ξ = 1/5, and a dissipation time lag t = s (which is equivalent to Q 3 1 4 ). Since the semimajor axis of the planet undergoes tidal dissipation, its value was certainly larger when the system formed. We then adopt = AU as initial value for all simulations. The initial eccentricity is chosen such that e 1 when = 75, the present observed values (Tab. 1). We further assume initial ϖ = 18, and π/ω = 5 day, that quickly evolves near the equilibrium rotation (Eq. ). In absence of a companion, the eccentricity and the semi-major axis are damped following an exponential decay (Correia & Laskar 1b), and the present configuration is attained after 1 Gyr (Fig. 1a). At present the observed eccentricity would be around, and we still needed to explain the high initial value near.7. We now add a companion to the system with m =. M J, a = 1.8 AU, and e =, and set e 1 =.3. At first, we only consider dissipation in the spin (Eq.16) and neglect its effect on the orbit (Eqs. 17, 18), in order to highlight the eccentricity pumping (Fig. 1b). We then clearly observe this effect, the eccentricity of the inner planet rising up to.7. We also observe that the eccentricity of the outer planet is simultaneously damped, because of the conservation of the total angular momentum (Eq. 15). Orbital and spin evolution cannot be dissociated, so we then integrate the full system (Fig. 1c). We observe that the initial behavior of the system is identical to the situation without dissipation on the orbit (Fig. 1b). However, as the eccentricity increases, the inner planet comes closer to the star at periastron, and tidal effects on the orbit become stronger. As a consequence, the semi-major axis decreases and the damping effect on the eccentricity (Eq. 18) overrides the pumping drift. The system ultimately evolves into a circular orbit. The present configuration is attained around 4 Gyr of evolution, which is compatible with the present estimated age of the star. The pumping effect is then responsible for a delay in the final evolution of planetary systems and may explain the high values observed for some of them (Tab. 1). Finally, we repeat the integration of the full system, but with a smaller-mass companion m = 5 M J at a = 1.4 AU. The companion eccentricity is still e =, but the inner planet now begins with e 1 = 5 (Fig. 1d). The initial evolution is still similar to the previous simulation (Fig. 1c), except that the eccentricity oscillations of both planets are higher, because the orbits are closer. Since the companion mass is smaller, its eccentricity also decreases more than before, and reaches zero around 3 Gyr. At this stage, the angle ϖ stops circulating, and begins librating around 18. The two orbits are then tightly coupled and evolve together, showing an identical behavior to close-in planets with a few days of orbital period (e.g. Mardling 7; Laskar et al. 11). As a consequence, the evolution time-scale is much longer, allowing the inner planet to maintain high eccentricity for longer periods of time (Fig. 1d). The present eccentricity is only observed after 6 Gyr. 4. ECCENTRICITY PUMPING In order to understand the unexpected behavior of the eccentricity during the initial stages of the evolution, we can perform some simplifications in the equations of motion without loss of generality (Sect. ). We can neglect tidal effects on orbital quantities (Eqs. 17, 18), which is justified since K K (Eq.19). The only contribution of tides is then on the rotation rate (Eq.16). The semimajor axis and the mean motion are thus constant, and the eccentricity only varies due to the gravitational perturbations (Eq. 6). In addition, we linearize the set of equations of motion in the vicinity of the averaged values of x, e 1, and e. Let x = x + δx, where x is the solution of (), e 1 = e 1 +δe 1, and e = e +δe. In the following, δe is expressed as a function of δx and δe 1 using the conservation of the angular momentum (Eq.15). Then, the set of equations of motion (6, 8, 16) reduces to: δė 1 = A sin ϖ, (3) with g = ϖ = g + g x δx + g e δe 1, (4) δẋ = ν x δx + ν e δe 1, (5) A = ν 31 e (1 + 3/4e 1) 1 ( )5/, (6) ν ( 1 ) + ν 1x ( 1 ) +ν 1 1 ( ν (1 + 3e 1/) )3/ (, (7) ) x g x = ν 1 (, (8) 1 )

4 A.C.M. Correia et al..8.7 (a).8.7 (b).3.3.. 1 3 4 5 6 1 3 4 5 6 (c) (d).3.3.. 1 3 4 5 6 1 3 4 5 6 time (Myr) time (Myr) Fig. 1. Long-term evolution of the HD 117618 system in different situations. We show the semi-major axis in AU (blue), and the eccentricities e 1 (red) and e (green). (a) without a companion; (b) with a m =. M J companion at a = 1.8 AU, but without dissipation on the orbit; (c) same as (b), but with a full model; (d) with a m = 5 M J companion at a = 1.4 AU, with a full model. The dashed line gives the slope of the linear approximation (Eq. 37). e 1 g e = ν ( + ν 4x e 1 1 ) 1 ( ν 3e 1 1 )3 ( ) e 1 ν 1, (9) 1 ( )3/ ν x = Kf 1 (e 1 ), (3) ν e = K(f 1(e 1 )x f (e 1 )), (31) where f 1(e) = 15(e + 3e 3 / + e 5 /8)/( ) 11/, and f (e) = 3(9e + 65e 3 / + 15e 5 /8 + 5e 7 /8)/( ) 7. We neglected the octupole terms since ν 3i ν i, the contributions from δe, and assumed that e i. At first order, the precession of the periastron is constant ϖ g, and the eccentricity is simply given from expression (3) as δe 1 = e cos(gt + ϖ ), (3) where e = A/g. That is, the eccentricity e 1 presents periodic variations around an equilibrium value e 1, with amplitude e and frequency g. Since g x δx, g e δe 1 g, the above solution for the eccentricity can be adopted as the zeroth order solution of the system of equations (3 5). With this approximation, the equation of motion of δx (5) becomes that of a driven harmonic oscillator whose the steady state solution is δx = x cos(gt + ϖ φ), (33) with x = ν e e/ ν x + g, and sin φ = g/ ν x + g. The rotation rate thus presents an oscillation identical to the eccentricity (Eq.3), but with smaller amplitude and delayed by an angle φ (see Correia 11). Using the above expression in equation (4) and integrating, gives for the periastron: ϖ = gt+ϖ + g x g x sin(gt+ϖ φ)+ g e g e sin(gt+ϖ ). (34) Finally, substituting in expression (3) and using the approximation g x x, g e e g gives δė 1 A sin(gt + ϖ ) g ea g e sin(gt + ϖ ) cos(gt + ϖ ) g xa g x sin(gt + ϖ φ) cos(gt + ϖ ), (35)

Pumping the eccentricity by tidal effect 5 or, combining the two products of periodic functions, δė 1 = A sin(gt + ϖ ) g xa g x sin(gt + ϖ φ) g ea g e sin(gt + ϖ ) + g xa x sin φ. (36) g The two middle terms in the above equation can be neglected since they are periodic and have a very small amplitude (g x x, g e e g). However, the last term in sin φ is constant and it adds a small drift to the eccentricity, < e 1 >= ν eg x A g(ν x + g ) t. (37) The drift is maximized for g ν x, it vanishes for weak dissipation ( x ), but also for strong dissipation (φ ). Note that the phase lag φ between the eccentricity (Eq. 3) and the rotation variations (Eq. 33) is essential to get a drift on the eccentricity. That is why the eccentricity pumping was never observed in previous studies that did not take into account the spin evolution. The major difference when we consider the full nonlinearized problem is that the drift (Eq. 37) cannot grow indefinitely. Indeed, when the eccentricity reaches high values, the drift vanishes (Fig. 1b). Moreover, tidal effects are also enhanced for high eccentricities and counterbalance the drift (Eq. 18). As a consequence, the drift on the eccentricity is never permanent, although it can last for the age of the system (Fig. 1c,d). In order to observe the pumping effect, the eccentricity should not be damped, while the damping timescale of the spin of the planet should be of the order of the period of eccentricity oscillation. This is valid for gaseous planets roughly within < <.3. About half of the planets in this range present eccentricities higher than.3 (Tab. 1). This is somehow unexpected, since the time-scale for damping the eccentricity (τ 1/K ) is shorter than the age of those systems (Tab. 1, Fig. 1a). Thus, unless the initial eccentricity of those planets was extremely high, one may suspect of the existence of undetected companions that help the eccentricity to maintain the present values. 5. CONCLUSION In this Letter we have shown that, under some particular initial conditions, orbital and spin evolution cannot be dissociated. Indeed, counterintuitive behaviors can be observed, such as the secular augmentation of the eccentricity. This effect can last over long time-scales and may explain the high eccentricities observed for moderate close-in planets. The variations in the flattening of the inner planet due to rotation (Eq. 1) is a key element to pump the eccentricity by means of g x (Eq. 37). We have considered an instantaneous response to the rotation in J, but a time delay until the maximum deformation is reached is to be expected. This may increase the phase lag between the eccentricity forcing (Eq. 3) and the precession angle oscillations (Eq. 34). It results that the drift effect on the eccentricity can be even more pronounced than the one presented here. We acknowledge support from the PNP-CNRS, the FCT (grant PTDC/CTE-AST/9858/8), and the European Research Council. Butler, R. P., et al. 6, Astrophys. J., 646, 55 Correia, A. C. M. 9, Astrophys. J., 74, L1 Correia, A. C. M. 11, in IAU Symposium, Vol. 76, IAU Symposium, ed. A. Sozzetti, M. G. Lattanzi, & A. P. Boss, 87 Correia, A. C. M., & Laskar, J. 4, Nature, 49, 848 Correia, A. C. M., & Laskar, J. 1a, Icarus, 5, 338 Correia, A. C. M., & Laskar, J. 1b, in Exoplanets (University of Arizona Press), 534 Correia, A. C. M., Laskar, J., Farago, F., & Boué, G. 11, Celestial Mechanics and Dynamical Astronomy, 111, 15 Correia, A. C. M., Laskar, J., & Néron de Surgy, O. 3, Icarus, 163, 1 Efroimsky, M., & Williams, J. G. 9, Celestial Mechanics and Dynamical Astronomy, 14, 57 Lambeck, K. 1988, Geophysical geodesy : the slow deformations of the earth Lambeck. (Oxford [England] : Clarendon Press ; New York : Oxford University Press, 1988.) REFERENCES Laskar, J., & Boué, G. 1, Astron. Astrophys., 5, A6 Laskar, J., Boué, G., & Correia, A. C. M. 11, arxiv:111565 Lee, M. H., & Peale, S. J. 3, Astrophys. J., 59, 11 Mardling, R. A. 7, Mon. Not. R. Astron. Soc., 38, 1768 Mignard, F. 1979, Moon and Planets,, 31 Murray, C. D., & Dermott, S. F. 1999, Solar System Dynamics (Cambridge University Press) Singer, S. F. 1968, Geophys. J. R. Astron. Soc., 15, 5 Touma, J. R., Tremaine, S., & Kazandjian, M. V. 9, Mon. Not. R. Astron. Soc., 394, 185