Free Access
Issue
A&A
Volume 603, July 2017
Article Number A108
Number of page(s) 5
Section Planets and planetary systems
DOI https://doi.org/10.1051/0004-6361/201628701
Published online 19 July 2017

© ESO, 2017

1. Introduction

Twenty years after the discovery of the first exoplanet around a Solar-type star (Mayor & Queloz 1995), the population of detected low-mass extra-solar planets has grown rapidly. Their number is now considerable and will keep increasing. This generates a strong need for theoretical modelling and predictions. Because the mass of such planets is not large enough to allow them to accrete a voluminous gaseous envelope, they are mainly composed of a telluric core. In many cases (e.g. 55 Cnc e, see Demory et al. 2016), this core is covered by a thin atmospheric layer as is observed on the Earth, Venus and Mars. The rotation of these bodies strongly affects their surface temperature equilibrium and atmospheric global circulation (Forget & Leconte 2014). Therefore, it is a key quantity to understand their climate, particularly in the so-called habitable zone (Kasting et al. 1993) where some of them have been found (Kopparapu et al. 2014).

The rotational dynamics of super-Venus and Venus-like planets is driven by the tidal torques exerted both on the rocky and atmospheric layers (see Correia et al. 2008). The solid torque, which is induced by the gravitational tidal forcing of the host star, tends to despin the planet to pull it back to synchronization. The atmospheric torque is the sum of two contributions. The first one, caused by the gravitational tidal potential of the star, acts on the spin in a similar way to the solid tide. The second one, called the “thermal tide”, results from the perturbation due to the heating of the atmosphere by the stellar bolometric flux. The torque induced by this component is in opposition to those gravitationally generated. Therefore, it pushes the angular velocity of the planet away from synchronization. Although the mass of the atmosphere often represents a negligible fraction of the mass of the planet (denoting fA this fraction, we have fA ~ 10-4 for Venus), thermal tides can be of the same order of magnitude and even stronger than solid tides (Dobrovolskis & Ingersoll 1980).

This competition naturally gives rise to prograde and retrograde rotation equilibria in the semi-major axis range defined by the habitable zone, in which Venus-like planets are submitted to gravitational and thermal tides of comparable intensities (Gold & Soter 1969; Dobrovolskis & Ingersoll 1980; Correia & Laskar 2001). Early studies of this effect were based on analytical models developed for the Earth (e.g. Chapman & Lindzen 1970) that show a singularity at synchronisation, while Correia & Laskar (2001) avoid this weak point by a smooth interpolation annulling the torque at synchronization. Only recently, the atmospheric tidal perturbation has been computed numerically with global circulation models (GCM; Leconte et al. 2015), and analytically by Auclair-Desrotour et al. (2017, hereafter referred to as P1), who generalized the reference work of Chapman & Lindzen (1970) by including the dissipative processes (radiation, thermal diffusion) that regularize the behaviour of the atmospheric tidal torque at the synchronization.

Here, we revisit the equilibrium rotation of super-Earth planets based on the atmospheric tides model presented in P1. For the solid torque, we use the simplest physical prescription, a Maxwell model (Remus et al. 2012; Correia et al. 2014), because the rheology of these planets is unknown.

2. Tidal torques

2.1. Physical set-up

For simplicity, we considered a planet in a circular orbit of radius a and mean motion n around a star of mass M and luminosity L exerting on the planet thermal and gravitational tidal forcing (Fig. 1). The planet, of mass MP and spin Ω, has zero obliquity so that the tidal frequency is σ = 2ω, where ω = Ω−n. It is composed of a telluric core of radius R covered by a thin atmospheric layer of mass MA = fAMP and pressure height scale H such that HR. This fluid layer is assumed to be homogeneous in composition, of specific gas constant A = ℛGP/m (GP and m being the perfect gas constant and the mean molar mass respectively), in hydrostatic equilibrium and subject to convective instability, that is, N ≈ 0, with N designating the Brunt-Väisälä frequency, as observed on the surface of Venus (Seiff et al. 1980). Hence, the pressure height scale is H=AT0g,\begin{equation} H = \frac{\mathcal{R}_{\rm A} T_0}{g}, \end{equation}(1)where T0 is the equilibrium surface temperature of the atmosphere and g the surface gravity which are related to the equilibrium radial distributions of density ρ0 and pressure p0 as p0 = ρ0gH. We introduce the first adiabatic exponent of the gas Γ1 = (lnp0/lnρ0)S (the subscript S being the specific macroscopic entropy) and the parameter κ = 1−1/Γ1. The radiative losses of the atmosphere, treated as a Newtonian cooling, lead us to define a radiative frequency σ0, given by Jrad=Cpσ0δT,\begin{equation} J_{\rm rad} = C_p \sigma_0 \delta T, \end{equation}(2)where Cp = ℛA/κ is the thermal capacity of the atmosphere per unit mass and Jrad is the radiated power per unit mass caused by the temperature variation δT around the equilibrium state. We note here that the Newtonian cooling describes the radiative losses of optically thin atmospheres, where the radiative transfers between layers can be ignored. We applied this modelling to optically thick atmosphere, such as Venus’, because the numerical simulations by Leconte et al. (2015) show that it can also describe well tidal dissipation in these cases, with an effective Newtonian cooling frequency. For more details about this physical set-up, we refer the reader to P1.

2.2. Atmospheric tidal torque

For null obliquity and eccentricity, the tidal gravitational potential is reduced to the quadrupolar term of the multipole expansion (Kaula 1964), U=U2r)(P22(cosθ)ei(σt+2ϕ)\hbox{$U = U_2 \left( r \right) P_2^2 \left( \cos \theta \right) e^{i \left( \sigma t + 2 \varphi \right)}$}, where t is the time, ϕ the longitude, θ the colatitude, r the radial coordinate, P22\hbox{$P_2^2$} the normalized associated Legendre polynomial of order (l,m) = (2,2) and U2 its radial function. Since HR, U2 is assumed to be constant. The thick atmosphere of a Venus-like planet absorbs most of the stellar flux in its upper regions, which are, as a consequence, strongly thermally forced; only 3% of the flux reaches the surface (Pollack & Young 1975). However the tidal effects resulting from the heating by the ground determine the tidal mass redistribution since the atmosphere is far denser near the surface than in upper regions (Dobrovolskis & Ingersoll 1980; Shen & Zhang 1990). These layers can also be considered in solid rotation with the solid part as a first approximation because the velocity of horizontal winds is less than 5 m s-1 below an altitude of 10 km (Marov et al. 1973). So, introducing the mean power per unit mass J2, we choose for the thermal forcing the heating at the ground distribution J=J2τJeτJxP22(cosθ)ei(σt+2ϕ)\hbox{$J = J_2 \tau_{J} e^{- \tau_{J} x} P_2^2 \left( \cos \theta \right) e^{i \left( \sigma t + 2 \varphi \right)}$}, where τJ ≫ 1 represents the damping rate of the heating with altitude and depends on the vertical thermal diffusion of the atmosphere at the surface (e.g. Chapman & Lindzen 1970). This allows us to establish for the atmospheric tidal torque the expression (see P1, Eq. (174)) 𝒯̂A=2πκR2ρ0(0)gU2J2σσ2+σ02,\begin{equation} \hat{\mathcal{T}}_{\rm A} = - 2 \pi \kappa \frac{R^2 \rho_0 \left( 0 \right)}{g} U_2 J_2 \frac{\sigma}{\sigma^2 + \sigma_0^2}, \label{Torque_a1} \end{equation}(3)where ρ00)(\hbox{$\rho_0 \left( 0 \right)$} is the density at the ground. This function of σ is of the same form as that given by Ingersoll & Dobrovolskis (1978, Eq. (4)). The quadrupolar tidal gravitational potential and heat power are given by (P1, Eq. (74)) U2=1232R2𝒢Ma3andJ2=1232αεR2LMAa2,\begin{equation} \begin{array}{rcl} \displaystyle U_2 = -\frac{1}{2} \sqrt{\frac{3}{2}} \frac{R^2 \mathcal{G} M_*}{a^3} & \mbox{and} & \displaystyle J_2 = \frac{1}{2} \sqrt{\frac{3}{2}} \frac{\alpha \varepsilon R^2 L_*}{M_{\rm A} a^2}, \end{array} \end{equation}(4)where \hbox{$\mathcal{G}$} designates the gravitational constant, ε the effective fraction of power absorbed by the atmosphere and α a shape factor depending on the spatial distribution of tidal heat sources. Denoting Ψ)(\hbox{$\mathcal{F} \left( \Psi \right)$} the distribution of tidal heat sources as a function of the stellar zenith angle Ψ, the parameter α is defined as α=0π(Ψ)P2(cosΨ)sinΨdΨ\hbox{$\alpha = \int_0^{\pi} \mathcal{F} ( \Psi ) P_2 ( \cos \Psi ) \sin \Psi d \Psi$}, where P2(X)=5/8(3X21)\hbox{$P_2(X)=\sqrt{5/8} ( 3 X^2 - 1 )$} is the normalized Legendre polynomial of order n = 2. If we assume a heat source of the form Ψ)(=cosΨ\hbox{$\mathcal{F} \left( \Psi \right) = \cos \Psi$} if Ψ[0,π2]\hbox{$\Psi \in [ 0,\frac{\pi}{2} ]$}, and Ψ)(=0\hbox{$\mathcal{F} \left( \Psi \right) = 0$} else, we get the shape factor α=1/85/20.19\hbox{$\alpha = 1/8\sqrt{5/2} \approx 0.19$}. The tidal torque exerted on the atmosphere \hbox{$\hat{\mathcal{T}}_{\rm A}$} is partly transmitted to the telluric core. The efficiency of this dynamical (viscous) coupling between the two layers is weighted by a parameter β (0 ≤ β ≤ 1). Hence, with ω0 = σ0/ 2, the transmitted torque \hbox{${\mathcal{T}}_{\rm A} = \beta\hat{\mathcal{T}}_{\rm A}$} (Eq. (3)) becomes 𝒯A=332κβR4𝒢MαεLAT0a5ωω2+ω02·\begin{equation} \mathcal{T}_{\rm A} = \frac{3}{32} \frac{\kappa \beta R^4 \mathcal{G} M_* \alpha \varepsilon L_* }{\mathcal{R}_{\rm A} T_0 \, a^5} \frac{\omega}{\omega^2 + \omega_0^2}\cdot \label{tideA} \end{equation}(5)

thumbnail Fig. 1

Tidal elongation of a Venus-like rotating planet, composed of a solid core (brown) and a gaseous atmosphere (blue), and submitted to gravitational and thermal forcings.

2.3. Solid tidal torque

For simplicity and because of the large variety of possible rheologies, we have assumed that the telluric core behaves like a damped oscillator. In this framework, called the Maxwell model, the tidal torque exerted on an homogeneous body can be expressed (e.g. Remus et al. 2012) 𝒯S=34𝒢M2R5a6{k2},with{k2}=3K2σMσσM2+(1+K)2σ2,\begin{equation} \mathcal{T}_{S} \!=\! \frac{3 }{4} \frac{\mathcal{G} M_*^2 R^5}{a^6} \Im \left\{ k_2 \right\} ,\, \mbox{with} \ \Im \left\{ k_2 \right\} \!=\! - \frac{3 K}{2 } \frac{ \sigma_{\rm M} \sigma }{ \sigma_{\rm M}^2 + \left( 1 + K \right)^2 \sigma^2}, \end{equation}(6)the imaginary part of the second order Love number (k2), K a non-dimensional rheological parameter, σM the relaxation frequency of the material composing the body, with K=38π3R4G𝒢MP2andσM=Gη,\begin{equation} \displaystyle K = \frac{38 \pi}{3} \frac{ R^4 G}{\mathcal{G} M_{\rm P}^2} \quad \mbox{and} \quad \displaystyle \sigma_{\rm M} = \frac{G}{\eta}, \end{equation}(7)where G and η are the effective shear modulus and viscosity of the telluric core. Finally, introducing the frequency ωM = σM/(2 + 2K), we obtain 𝒯S=98𝒢M2R5KωM(1+K)a6ωω2+ωM2·\begin{equation} \mathcal{T}_{S} = - \frac{9}{8} \frac{\mathcal{G} M_*^2 R^5 K \omega_{\rm M}}{ \left( 1 + K \right) a^6} \frac{\omega}{\omega^2 + \omega_{\rm M}^2}\cdot \label{tideS} \end{equation}(8)

3. Rotational equilibrium states

3.1. Theory

The total torque exerted on the planet is the sum of the two previous contributions: \hbox{$\mathcal{T}_{\rm S+A} = \mathcal{T}_{\rm S} + \mathcal{T}_{\rm A}$} . When the atmospheric and solid torques are of the same order of magnitude, several equilibria can exist, corresponding to \hbox{$\mathcal{T}_{\rm S+A} = 0$} (Fig. 2). The synchronization is given by Ω0 = n and non-synchronized retrograde and prograde states of equilibrium, denoted Ω and Ω+ respectively, are expressed as functions of a (Eqs. (5) and (8)) Ω(a)=nωeq(a)andΩ+(a)=n+ωeq(a),\begin{equation} \Omega_{- } \left( a \right) = n - \omega_{\rm eq} \left( a \right) \quad \mbox{and} \quad \Omega_{+} \left( a \right) = n + \omega_{\rm eq} \left( a \right), \label{Omega_pm} \end{equation}(9)where the difference to synchronization, ωeq, is given by ωeq(a)=ωMaAω02/ωMAωMawithA=12AT0MRKαεLκβ(1+K)·\begin{equation} \displaystyle \omega_{\rm eq} \left( a \right) = \omega_{\rm M} \sqrt{ \frac{a - A \omega_0^2/\omega_{\rm M}}{A \omega_{\rm M} - a} } ~ \mbox{with} ~ \displaystyle A = 12 \frac{ \mathcal{R}_{\rm A} T_0 M_* R K}{\alpha \varepsilon L_* \kappa \beta \left( 1 + K \right)}\cdot \label{weq} \end{equation}(10)

thumbnail Fig. 2

Normalized total tidal torque exerted on a Venus-like planet (in red) and its solid (in green) and atmospheric (in blue) components as functions of the forcing frequency ω/n with ω0/n = 2 and ωM/n = 6.

As a first approximation, we consider that parameters A, ω0 and ωM do not vary with the star-planet distance. In this framework, the equilibrium of synchronization is defined for all orbital radii, while those associated with Eq. (10) only exist in a zone delimited by ainf<a<asup, the boundaries being ainf=Aω0λ-1andasup=Aω0λ,\begin{equation} a_{\rm inf} = A \omega_0 \lambda^{-1} \quad \mbox{and} \quad a_{\rm sup} = A \omega_0 \lambda, \label{abd} \end{equation}(11)with the ratio λ=maxω0(/ωM,ωM/ω0)\hbox{$\lambda = \max \left( \omega_0/\omega_{\rm M}, \omega_{\rm M}/\omega_0 \right)$}. In the particular case where ω0 = ωM, the distance aeq = Aω0 corresponds to an orbit for which the atmospheric and solid tidal torques counterbalance each other whatever the angular velocity (Ω) of the planet. Studying the derivative of ωeq (Eq. (10)), we note that when the star-planet distance increases, the pseudo-synchronized states of equilibrium get closer to the synchronization (ωM<ω0) or move away from it (ωM>ω0), depending on the hierarchy of frequencies associated with dissipative processes. To determine the stability of the identified equilibria, we introduce the first order variation δω such that ω = ωeq + δω and study the sign of \hbox{$\mathcal{T}_{\rm S+A}$} in the vicinity of an equilibrium position for a given a. We first treat the synchronization, for which 𝒯S+Asign(aaeq)δωwithaeq=Aω02ωM·\begin{equation} \mathcal{T}_{\rm S+A} \, \propto \, {\rm sign} \left( a - a_{\rm eq} \right) \delta \omega \quad \mbox{with} \quad \displaystyle a_{\rm eq} = A \frac{\omega_0^2}{\omega_{\rm M}}\cdot \label{stab1} \end{equation}(12)We note that aeq = ainf if ω0<ωM and aeq = asup otherwise. In the vicinity of non-synchronized equilibria, we have 𝒯S+An4ωMωeq2(ω02ωM2)(ωeq2+ωM2)2(ωeq2+ω02)δω.\begin{equation} \mathcal{T}_{\rm S+A} \, \propto \, n^4 \omega_{\rm M} \frac{\omega_{\rm eq}^2 \left( \omega_0^2 - \omega_{\rm M}^2 \right)}{\left( \omega_{\rm eq}^2 + \omega_{\rm M}^2 \right)^2 \left( \omega_{\rm eq}^2 + \omega_0^2 \right) } \delta \omega. \label{stab2} \end{equation}(13)Therefore, within the interval aainf],asup[\hbox{$a \in \left] a_{\rm inf} , a_{\rm sup} \right[$}, the stability of states of equilibrium also depends on the hierarchy of frequencies. If ωM<ω0 (ωM>ω0), he synchronized state of equilibrium is stable (unstable) and the pseudo-synchronized ones are unstable (stable). For a<ainf the gravitational tide predominates and the equilibrium at Ω = n is stable. It becomes unstable for a>asup, because the torque is driven by the thermal tide.

thumbnail Fig. 3

Top: solid (left panel), atmospheric (middle panel) and total tidal torque (right panel) exerted on a Venus-like planet as functions of the reduced forcing frequency ω/n (horizontal axis) and orbital radius a in logarithmic scale (vertical axis). The colour level corresponds to the torque in logarithmic scale with isolines at \hbox{$\log_{10} \left( \mathcal{T} \right) = 8, 10, 12, \ldots, 26$} . Bottom: sign of the solid (left panel), atmospheric (middle panel) and total (right panel) tidal torque as functions of the same parameters: white (blue) areas are associated with positive (negative) torques. Stable (unstable) states of equilibrium are designated by blue (red) lines (with A = 1.88 × 1019 m s). The pink band corresponds to the habitable zone for the black body equilibrium temperature Teq = 288 K ± 20% for a 1 M Solar-type star at the age of the Sun.

3.2. Comparison with previous works

In many early studies dealing with the spin rotation of Venus, the effect of the gravitational component on the atmosphere is disregarded so that the atmospheric tide corresponds to a pure thermal tide (e.g. Dobrovolskis & Ingersoll 1980; Correia & Laskar 2001; Correia et al. 2003, 2008; Cunha et al. 2015). Moreover, the torque induced by the tidal elongation of the solid core is generally assumed to be linear in these works. This amounts to considering that |ω| ≪ ωM and ω0ωM. Hence, the relation Eq. (10) giving the positions of non-synchronized equilibria can be simplified: ωeq=ωMAaω02.\begin{equation} \displaystyle \omega_{\rm eq} = \sqrt{ \frac{\omega_{\rm M}}{A} a - \omega_0^2}. \end{equation}(14)Following Correia et al. (2008), we take ω0 = 0. We then obtain ωeqa\hbox{$\sqrt{a}$} while Correia et al. (2008) found ωeqa. This difference can be explained by the linear approximation of the sine of the phase lag done in this previous work. Given that the condition ω0<ωM is satisfied, the non-synchronized states of equilibrium are stable in this case.

By using a GCM, Leconte et al. (2015) obtained numerically for the atmospheric tidal torque an expression similar to the one given by Eq. (3). They computed with this expression the possible states of equilibrium of Venus-like planets by applying to the telluric core the Andrade and constant-Q models and showed that for both models, up to five equilibrium positions could appear. This is not the case when the Maxwell model is used, as shown by the present work. Hence, although these studies agree well on the existence of several non-synchronized states of equilibrium, they also remind us that the theoretical number and positions of these equilibria depend on the models used to compute tidal torques and more specifically on the rheology chosen for the solid part.

Table 1

Numerical values used in the case of Venus-like planets.

4. The case of Venus-like planets

We now illustrate the previous results for Venus-like planets (Table 1). The frequency ωM is adjusted so that the present angular velocity of Venus corresponds to the retrograde state of equilibrium identified in the case where the condition ωM>ω0 is satisfied (Ω in Eq. (9)).In Fig. 3, we plot the resulting tidal torque and its components, as well as their signs, as functions of the tidal frequency and orbital radius. These maps show that the torque varies over a very large range, particularly the solid component (\hbox{$\mathcal{T}_{\rm S}$} a-6 while \hbox{$\mathcal{T}_{\rm A}$} a-5). The combination of the solid and atmospheric responses generates the non-synchronized states of equilibrium observed on the bottom left panel, which are located in the interval ainf],asup[\hbox{$\left] a_{\rm inf} , a_{\rm sup} \right[$} (see Eq. (11)) and move away from the synchronization when a increases (see Eq. (10)), as predicted analytically.

For illustration, in Fig. 4, we show the outcome of a different value of ωM, with ω0>ωM, contrary to Fig. 3 where ωM>ω0. We observe the behaviour predicted analytically in Sect. 3. In Fig. 3, the non-synchronized equilibria are stable and move away from the synchronization when a increases, but they are unstable in Fig. 4.

We note that the value of the solid Maxwell frequency obtained for stable non-synchronized states of equilibrium, ωM = 1.075 × 10-4 s-1, is far higher than those of typical solid bodies (ωM ~ 10-10 s-1, see Efroimsky 2012). The Maxwell model, because of its decreasing rate as a function of ω (i.e. ω-1), underestimates tidal torque for tidal frequencies that are greater than ωM, which leads to overestimate the Maxwell frequency when equalizing atmospheric and solid torques. Using the Andrade model for the solid part could give more realistic values of ωM, as proved numerically by Leconte et al. (2015), because the decreasing rate of the torque is lower in the Andrade model than in the Maxwell model (i.e. ωα with α = 0.2−0.3).

Finally, we must discuss the assumption we made when we supposed that the parameters of the system did not depend on the star-planet distance. The surface temperature of the planet and the radiative frequency actually vary with a. If we consider that T0 is determined by the balance between the heating of the star and the black body radiation of the planet, then T0a− 1/2. As σ0T03\hbox{$T_0^3$}, we have σ0a− 3/2. These new dependences modify neither the expressions of ωeq (Eq. (10)), nor the stability conditions of the states of equilibrium (Eqs. (12) and (13)). However, they have repercussions on the boundaries of the region where non-synchronized states exist. This changes are illustrated by Fig. 4 (bottom panel), which shows the stability diagram of Fig. 3 (bottom left panel) computed with the functions T0(a) = T0;(a/a)−1/2 and σ0(a) = σ0;(a/a)−3/2 (a, T0; and σ0; being the semi-major axis of Venus and the constant temperature and radiative frequency of Table 1 respectively).

thumbnail Fig. 4

Sign of the total tidal torque as a function of the tidal frequency ω/n (horizontal axis) and orbital radius (a) (vertical axis) for two different cases. Top: with ω0 = 3.77 × 10-7 s-1 and ωM = 3.7 × 10-8 s-1 (ω0>ωM). Bottom: with T0 and ω0 depending on the star-planet distance and ωM>ω0. White (blue) areas correspond to positive (negative) torque. Blue (red) lines designate stable (unstable) states of equilibrium.

5. Discussion

A physical model for solid and atmospheric tides allows us to determine analytically the possible rotation equilibria of Venus-like planets and their stability. Two regimes exist depending on the hierarchy of the characteristic frequencies associated with dissipative processes (viscous friction in the solid layer and thermal radiation in the atmosphere). These dissipative mechanisms have an impact on the stability of non-synchronized equilibria, which can appear in the habitable zone since they are caused by solid and atmospheric tidal torques of comparable intensities. This study can be used to constrain the equilibrium rotation of observed super-Earths and therefore to infer the possible climates of such planets. Also here the Maxwell rheology wasused. This work can be directly applied to alternative rheologies such as the Andrade model (Efroimsky 2012). The modelling used here is based on an approach where the atmosphere is assumed to rotate uniformly with the solid part of the planet (Auclair-Desrotour et al. 2017). As general circulation is likely to play an important role in the atmospheric tidal response, we envisage examining the effect of differential rotation and corresponding zonal winds of the tidal torque in future studies.


Acknowledgments

P. Auclair-Desrotour and S. Mathis acknowledge funding by the European Research Council through ERC grant SPIRE 647383. This work was also supported by the “exoplanètes” action from Paris Observatory and by the CoRoT/Kepler and PLATO CNES grant at CEA-Saclay. A. C. M. Correia acknowledges support from CIDMA strategic project UID/MAT/04106/2013.

References

  1. Auclair-Desrotour, P., Laskar, J., & Mathis, S. 2017, A&A, 603, A107 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  2. Chapman, S., & Lindzen, R. 1970, Atmospheric tides: Thermal and gravitational (Dordrecht: Reidel) [Google Scholar]
  3. Correia, A. C. M., & Laskar, J. 2001, Nature, 411, 767 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
  4. Correia, A. C. M., Laskar, J., & de Surgy, O. N. 2003, Icarus, 163, 1 [NASA ADS] [CrossRef] [Google Scholar]
  5. Correia, A. C. M., Levrard, B., & Laskar, J. 2008, A&A, 488, L63 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  6. Correia, A. C. M., Boué, G., Laskar, J., & Rodríguez, A. 2014, A&A, 571, A50 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  7. Cunha, D., Correia, A. C., & Laskar, J. 2015, Int. J. Astrobiol., 14, 233 [CrossRef] [Google Scholar]
  8. Demory, B.-O., Gillon, M., Madhusudhan, N., & Queloz, D. 2016, MNRAS, 455, 2018 [NASA ADS] [CrossRef] [Google Scholar]
  9. Dobrovolskis, A. R., & Ingersoll, A. P. 1980, Icarus, 41, 1 [NASA ADS] [CrossRef] [Google Scholar]
  10. Efroimsky, M. 2012, ApJ, 746, 150 [NASA ADS] [CrossRef] [Google Scholar]
  11. Forget, F., & Leconte, J. 2014, Phil. Trans. R. Soc. Lond. Ser. A, 372, 20130084 [Google Scholar]
  12. Gold, T., & Soter, S. 1969, Icarus, 11, 356 [NASA ADS] [CrossRef] [Google Scholar]
  13. Ingersoll, A. P., & Dobrovolskis, A. R. 1978, Nature, 275, 37 [NASA ADS] [CrossRef] [Google Scholar]
  14. Kasting, J. F., Whitmire, D. P., & Reynolds, R. T. 1993, Icarus, 101, 108 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
  15. Kaula, W. M. 1964, Reviews of Geophysics and Space Physics, 2, 661 [NASA ADS] [CrossRef] [Google Scholar]
  16. Kopparapu, R. K., Ramirez, R. M., SchottelKotte, J., et al. 2014, ApJ, 787, L29 [Google Scholar]
  17. Leconte, J., Wu, H., Menou, K., & Murray, N. 2015, Science, 347, 632 [NASA ADS] [CrossRef] [Google Scholar]
  18. Marov, M. Y., Avduevskij, V. S., Kerzhanovich, V. V., et al. 1973, J. Atmos. Sci., 30, 1210 [NASA ADS] [CrossRef] [Google Scholar]
  19. Mayor, M., & Queloz, D. 1995, Nature, 378, 355 [Google Scholar]
  20. Mohr, P. J., Taylor, B. N., & Newell, D. B. 2012, J. Phys. Chem. Ref. Data, 41, 043109 [NASA ADS] [CrossRef] [Google Scholar]
  21. Pollack, J. B., & Young, R. 1975, J. Atmos. Sci., 32, 1025 [NASA ADS] [CrossRef] [Google Scholar]
  22. Remus, F., Mathis, S., Zahn, J.-P., & Lainey, V. 2012, A&A, 541, A165 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  23. Seiff, A., Kirk, D. B., Young, R. E., et al. 1980, J. Geophys. Res., 85, 7903 [Google Scholar]
  24. Shen, M., & Zhang, C. Z. 1990, Icarus, 85, 129 [NASA ADS] [CrossRef] [Google Scholar]

All Tables

Table 1

Numerical values used in the case of Venus-like planets.

All Figures

thumbnail Fig. 1

Tidal elongation of a Venus-like rotating planet, composed of a solid core (brown) and a gaseous atmosphere (blue), and submitted to gravitational and thermal forcings.

In the text
thumbnail Fig. 2

Normalized total tidal torque exerted on a Venus-like planet (in red) and its solid (in green) and atmospheric (in blue) components as functions of the forcing frequency ω/n with ω0/n = 2 and ωM/n = 6.

In the text
thumbnail Fig. 3

Top: solid (left panel), atmospheric (middle panel) and total tidal torque (right panel) exerted on a Venus-like planet as functions of the reduced forcing frequency ω/n (horizontal axis) and orbital radius a in logarithmic scale (vertical axis). The colour level corresponds to the torque in logarithmic scale with isolines at \hbox{$\log_{10} \left( \mathcal{T} \right) = 8, 10, 12, \ldots, 26$} . Bottom: sign of the solid (left panel), atmospheric (middle panel) and total (right panel) tidal torque as functions of the same parameters: white (blue) areas are associated with positive (negative) torques. Stable (unstable) states of equilibrium are designated by blue (red) lines (with A = 1.88 × 1019 m s). The pink band corresponds to the habitable zone for the black body equilibrium temperature Teq = 288 K ± 20% for a 1 M Solar-type star at the age of the Sun.

In the text
thumbnail Fig. 4

Sign of the total tidal torque as a function of the tidal frequency ω/n (horizontal axis) and orbital radius (a) (vertical axis) for two different cases. Top: with ω0 = 3.77 × 10-7 s-1 and ωM = 3.7 × 10-8 s-1 (ω0>ωM). Bottom: with T0 and ω0 depending on the star-planet distance and ωM>ω0. White (blue) areas correspond to positive (negative) torque. Blue (red) lines designate stable (unstable) states of equilibrium.

In the text

Current usage metrics show cumulative count of Article Views (full-text article views including HTML views, PDF and ePub downloads, according to the available data) and Abstracts Views on Vision4Press platform.

Data correspond to usage on the plateform after 2015. The current usage metrics is available 48-96 hours after online publication and is updated daily on week days.

Initial download of the metrics may take a while.