Skip to main content
Log in

Librations of a body composed of a deformable mantle and a fluid core

  • Original Article
  • Published:
Celestial Mechanics and Dynamical Astronomy Aims and scope Submit manuscript

Abstract

We present fully three-dimensional equations to describe the rotations of a body made of a deformable mantle and a fluid core. The model in its essence is similar to that used by INPOP19a (Integration Planétaire de l’Observatoire de Paris) Fienga et al. (INPOP19a planetary ephemerides. Notes Scientifiques et Techniques de l’Institut de Mécanique Céleste, vol 109, 2019), and by JPL (Jet Propulsion Laboratory) (Park et al. The JPL Planetary and Lunar Ephemerides DE440 and DE441. Astron J 161(3):105, 2021. doi:10.3847/1538-3881/abd414), to represent the Moon. The intended advantages of our model are: straightforward use of any linear-viscoelastic model for the rheology of the mantle; easy numerical implementation in time-domain (no time lags are necessary); all parameters, including those related to the “permanent deformation”, have a physical interpretation. The paper also contains: (1) A physical model to explain the usual lack of hydrostaticity of the mantle (permanent deformation). (2) Formulas for free librations of bodies in and out-of spin-orbit resonance that are valid for any linear viscoelastic rheology of the mantle. (3) Formulas for the offset between the mantle and the idealised rigid-body motion (Peale’s Cassini states). (4) Applications to the librations of Moon, Earth, and Mercury that are used for model validation.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14

Similar content being viewed by others

Notes

  1. Some expressions become simpler if we use the density tensor \(\varvec{M}_T\) Chandrasekhar (1969) defined as

    $$\begin{aligned} \begin{aligned}&M_{Tij}= \int \rho (\varvec{x})x_ix_jd\varvec{x}^3\,,\quad \text {with:}\quad \mathrm{I}_{\circ }=\frac{2}{3}\mathrm{\!\ Tr\!\ } \big ({\mathbf {M}}_T\big )\,,\quad {\mathbf {M}}_T=\frac{1}{2} \mathrm{I}_{\circ }{\mathbb {I}} + \mathrm{I}_{\circ }{\mathbf {B}}_T\,,\\&\varvec{M}_T=\frac{\mathrm{\!\ Tr\!\ } \varvec{I}_T}{2}{\mathbb {I}} -\varvec{I}_T\, \quad \text {and}\quad \varvec{I}_T=\mathrm{\!\ Tr\!\ } \big (\varvec{M}_T\big ){\mathbb {I}} -\varvec{M}_T\,. \end{aligned} \end{aligned}$$
    (2.10)

    For instance, \( \varvec{{\widehat{\pi }}}_T= \varvec{M}_T \, \varvec{{\widehat{\omega }}}_T + \varvec{\widehat{\omega }}_T\,\varvec{M}_T\).

  2. Most frames we have defined are similar to those used to describe the rotational motion of the Earth in the IERS2010, chapters 2 to 5 Petit and Luzum (2010). The correspondence is the following (the number in brackets refers to a section in the IERS 2010): “International Terrestrial Reference System (ITRS)[4.1.1]”\(\rightarrow {\mathrm {K}}_m\), “Terrestrial Intermediate Reference System (TIRS)[5.4.1]”\(\rightarrow {\mathrm {K}}_g\), “Celestial Intermediate Reference System (CIRS) [5.4.2 and 5.4.4]”\(\rightarrow {\mathrm {K}}_s\), and “Geocentric Celestial Reference System (GCRS) [5.4.4]”\(\rightarrow \kappa \). Our definition of \({\mathrm {K}}_m\) is different but related to that of the ITRS after the identification of \({\mathrm {K}}_m\) with the prestress frame. Our definitions of guiding frame and slow frame are conventional as well as those of TIRS and CIRS in the IERS2010 [5.3.2]. We decided to give different names to reference systems already defined in the IERS2010 because those in the later have precise definitions, which applies to the Earth, while ours \({\mathrm {K}}_g\) and \({\mathrm {K}}_s\) do not, since they are to be applied to any libration problem.

  3. In Eckhardt (1981), for instance, three small angles \((\sigma ,\rho ,\tau )\) describe the deviation of the lunar orientation from the ideal Cassini state, which is our guiding motion. A computation using Eckhardt’s parameterisation of the Moon’s body frame and the approximation \(\cos (I)=\cos (\iota _p)\approx 1\) shows that \(\tau \) is equal to our angle \(\alpha _{m3}\). The relation between \((\sigma ,\rho )\) to \((\alpha _{m1},\alpha _{m2})\) is not so simple and instead of these angles it is more convenient to use, as Eckhardt did, “the selenographic unit vector to the pole of the ecliptic” \(\varvec{p}=(p_1,p_2,p_3)\). In our notation, \(\varvec{p}= \varvec{R}_m^{-1} \varvec{e}_3= \varvec{R}_m^{-1}\varvec{R}_g\varvec{R}_g^{-1}\varvec{e}_3\approx ({\mathbb {I}} -\varvec{{\widehat{\alpha }}}_m)\varvec{R}_g^{-1}\varvec{e}_3.\) The same approximation used before, \(\cos (I)=\cos (\iota _p)\approx 1\), gives \(\varvec{R}_g^{-1}\varvec{e}_3=\varvec{e}_3\) and \(\varvec{p}\approx (-\alpha _{m2},\alpha _{m1},1)\) (we are assuming that the orientation of the Axis 1 of the guiding frame is positive towards the Earth). So, we get the correspondence

    $$\begin{aligned} (p_1,p_2,\tau )=(-\alpha _{m2},\alpha _{m1},\alpha _{m3}) \end{aligned}$$
    (2.18)

    between the libration elements used by Eckhardt and ours. The triple \((p_1,p_2,\tau )\) also appears in Eckhardt’s work, equation (5), where it is denoted as \(\varvec{X}\).

  4. There is a relation between the AP and the Correspondence Principle Efroimsky (2012a). The relation between both principles is addressed in Section 4 of (Correia et al. 2018). The main difference is that the AP is defined directly in the time domain while the correspondence principle is defined in the frequency domain.

  5. The gravitational modulus is related to the fluid Love number \(k_f\) by means of \(\frac{\omega ^2}{\gamma }=\frac{R^5\omega ^2}{3{\,\mathrm I}_\circ G}k_f\). This is the same relation that appears in Mathews et al. (2002) (paragraph [21]) after we replace \(\frac{\omega ^2}{\gamma }\) by a compliance coefficient. So, \(\gamma ^{-1}\) is a dimensional gravitational compliance similar to those in Mathews et al. (2002).

  6. In the particular problem treated in this paper we could have started directly with the effective parameters \(\mu _0,\eta _0,\ldots \) and avoided the previous definition of the “mantle parameters” \(\mu _{0m},\eta _{0m},\ldots \). We started with the mantle parameters for two reasons. At first the simplification is not possible for a body with more than one deformable layer. At second the rescaling in Eq. (3.9) shows that for a homogeneous mantle (see footnote 9) the coefficients \(\mu _0,\eta _0,\eta \) must change when the ratio \(\frac{\mathrm{I}_{\circ ,T}}{\mathrm{I}_{\circ ,m}}\) is varied while the material properties are preserved.

  7. There is a great uncertainty on the value of the viscosity (eddy) near the CMB of the Earth. Several values have been used in the literature: \(\nu =10^{-6}\) \(\hbox {m}^2\)/s (molecular viscosity of iron), \(\nu =3.5\times 10^{-2}\) \(\hbox {m}^2\)/s (\(E_k= 4\times 10^{-11}\)) Deleplace and Cardin (2006), \(\nu =883\) \(\hbox {m}^2\)/s (\(E_k=10^{-6}\)) Triana et al. (2019), \(\nu =3.5 \times 10^5\) \(\hbox {m}^2\)/s (\(E_k=4 \times 10^{-4}\)) Matsui and Buffett (2012), etc.

  8. At this point it would be interesting to relate the prestress frame to analogous frames used by other authors. We restrict the discussion to the case of the Moon. Eckhardt (1981) defines “selenographic coordinates whose axes are the same as those of the Moon’s principal moments of inertia in the absence of elastic deformation.” Viswanathan et al. (2019) use the same frame as Folkner et al. (2014): “The mantle coordinate system is defined by the principal axes of the undistorted mantle in which the moment of inertia matrix of the undistorted mantle is diagonal.” In both Eckhardt and Folkner approaches, the body or the mantle reference frames, respectively, is defined using an undistorted configuration on which the real distorted situation is described. In our approach the undistorted configuration is replaced by the prestress frame that is a Tisserand frame. It seems that in Eckhardt and Folkner the undistorted frame is implicitly assumed to be a Tisserand frame, since it is consistently used in this way.

  9. Associated with a nonhomogeneous body with an effective Kelvin-Voigt rheology and with parameters \({\,\mathrm I}_\circ \), m, \(\mu _0\), and \(\eta \) there is an equivalent homogeneous body with the same parameters and the same rheology such that the molecular shear modulus \(\mu _{0\,mol}\) and the molecular viscosity \(\eta _{0\,mol}\) of the later are given by Correia et al. (2018)

    $$\begin{aligned} \mu _{0\,mol} \ \text {(Pa)} =\frac{15}{152\pi }\frac{m}{R_{\,\mathrm I}}\, \mu _0\qquad \text {and}\qquad \eta _{mol} \ \text {(Pa sec)}=\frac{15}{152 \pi }\frac{m}{R_{\,\mathrm I}}\, \eta \,. \end{aligned}$$

    Under this correspondence we find the following values for the pair \((\mu _{0\,mol},\eta _{mol})\) for each one of the equivalent homogeneous bodies in Table 9 : Moon \(\big (\)62 GPa, 5.2\(\times 10^{14}\) Pa s\(\big )\), Mercury \(\big (\)8.9 GPa, 1.4\(\times 10^{14}\) Pa s\(\big )\), Earth \(\big (\)120 GPa, 1.7\(\times 10^{14}\) Pa s\(\big )\), and Mars \(\big (\)39 GPa, 6.5\(\times 10^{12}\) Pa s\(\big )\). Note that the molecular characteristic time \(\eta _{mol}/ \mu _{0\,mol}\) is equal to the characteristic time of the rheology \(\eta /\mu _0\).

  10. The terms \(\varvec{I}_c\varvec{\omega }_c\times \varvec{\omega }_c - k_c(\varvec{\omega }_m - \varvec{\omega }_c)-3 \sum _\beta \frac{\mathcal{G}m_\beta }{ \Vert \varvec{r}_\beta \Vert ^5}(\varvec{I}_c{\mathbf {r}}_\beta )\times {\mathbf {r}}_\beta \) in the first equation of system (3.19) represent the torque of the core upon the mantle. If the core is spherical, then \(\varvec{I}_c\varvec{\omega }_c\times \varvec{\omega }_c=-3 \sum _\beta \frac{\mathcal{G}m_\beta }{ \Vert \varvec{r}_\beta \Vert ^5}(\varvec{I}_c{\mathbf {r}}_\beta )\times {\mathbf {r}}_\beta =0\) and the torque of the core upon the mantle reduces to \(- k_c(\varvec{\omega }_m - \varvec{\omega }_c)\), which represents the shear-stress torque at the CMB. If the core is not spherical, then pressure can also produce torque. The term \(\varvec{I}_c\varvec{\omega }_c\times \varvec{\omega }_c\) can be interpreted as the pressure torque due to the motion and the inertia of the fluid. The term \(-3 \sum _\beta \frac{\mathcal{G}m_\beta }{ \Vert \varvec{r}_\beta \Vert ^5}(\varvec{I}_c{\mathbf {r}}_\beta )\times {\mathbf {r}}_\beta \) is the pressure torque from the core upon the mantle caused by the action of external gravity on the core. All the three terms \(\varvec{I}_c\varvec{\omega }_c\times \varvec{\omega }_c\), \( - k_c(\varvec{\omega }_m - \varvec{\omega }_c)\), and \(-3 \sum _\beta \frac{\mathcal{G}m_\beta }{ \Vert \varvec{r}_\beta \Vert ^5}(\varvec{I}_c{\mathbf {r}}_\beta )\times {\mathbf {r}}_\beta \) produce reactive counter-torques from the mantle upon the core. The first and second of these reactive counter-torques are present in the equation for \(\dot{\varvec{\pi }}_c\) in system (3.19). The third pressure-reactive term, \(3 \sum _\beta \frac{\mathcal{G}m_\beta }{ \Vert \varvec{r}_\beta \Vert ^5}(\varvec{I}_c{\mathbf {r}}_\beta )\times {\mathbf {r}}_\beta \), is absent because it is cancelled out by the external gravitational force that acts upon the core.

    Now, suppose the core is spherical, so that the torque of the core upon the mantle is \( - k_c(\varvec{\omega }_m - \varvec{\omega }_c)=- \dot{\varvec{\pi }}_c=-\frac{d}{dt}\varvec{I}_c \varvec{\omega }_c\). If we take the limit as \(k_c\rightarrow \infty \) while \(\varvec{\pi }_c\) remains bounded, then we obtain \(\varvec{\omega }_c\rightarrow \varvec{\omega }_m\) and the mantle and core move as they formed a rigid body. In this case the torque of the core upon the mantle becomes, as expected, \(-\frac{d}{dt}\varvec{I}_c \varvec{\omega }_m\).

  11. If we assume that the body is in a Cassini state in a s-to-2 spin-orbit resonance, s integer, then the coefficients \(c_1\) and \(c_2\) are explicitly given in terms of Hansen coefficients in Eqs.  (A.14) and (A.15). These expressions imply: for the Moon \(\xi _1=0.9976\), \(\xi _2=3.927\); for Enceladus \(\xi _1=1.00012\), \(\xi _2=3.99351\); and for Mercury \(\xi _1=1.27513\), \(\xi _2=2.14751\); and all these constants are of order of one. If there is no spin-orbit resonance then \(c_2=0\) and \(\xi _1=\xi _2\), for the Earth \(c_1=0.000027\) and \(\xi _1=1.000027\).

    For bodies that are out of spin-orbit resonance and are not close to massive bodies, the average external gravitational upon them is small and, so \(c_1\approx 0\) and \(\xi _1=\xi _2\approx 1\).

    For a body in 1:1 spin orbit resonance, if the inclination of the body spin axis to the normal to the orbital plane is small and the eccentricity of the orbit is small, then Eqs. (A.14), (A.15), and (A.11) give \(c_1\approx c_2 \approx \frac{3}{2} \frac{m_p}{m_p+m}\), where \(m_p\) is the mass of the point mass (the tidal raising body) and m is the mass of the extended body. In the case of the Moon or Enceladus \(c_1\approx c_2\approx 3/2\) and \((\xi _1,\xi _2)\approx (1,4)\).

  12. Our libration equations depend only on the time derivatives of the angles \(\alpha _{c1}\), \(\alpha _{c2}\), or \(\alpha _{c3}\) and not on the angles themselves. This gives rise to degenerated eigenmodes where all variables are zero but the angles \(\alpha _{c1},\alpha _{c2},\alpha _{c3}\). These degenerated eigenmodes could be easily removed if we had considered the angular velocities of the core as variables of the problem instead of the angles themselves. We decided to keep the angles because they are the variables which are more easily visualised.

  13. In the same way C(0) is related to an ideal flattening \(\alpha _{id}\) due to centrifugal forces, described before Eq. (6.6), \(C(0)\, (\xi _2-\xi _1)\) is related to an ideal ellipticity coefficient \(\gamma _{id}=\frac{ \overline{I}_{id2}-\overline{I}_{id1}}{\overline{I}_{id3}}\) due to tidal deformations. For simplicity suppose that the extended body is in 1-to-1 spin orbit resonance with an orbiting point mass with circular orbit. In this case \(c_1=c_2=3/2\), which implies \(\xi _2-\xi _1=3\) (see Footnote 11). As in the definition of \(\alpha _{id}\) we also assume that the body has no prestress \(\varvec{B}_{0,m}=0\). In this ideal situation the equilibrium Eqs. (4.2) and (4.12) and the expression for \(\overline{\gamma }\) in Table 4 imply that \(\gamma _{id}=C(0)\, (\xi _2-\xi _1)=3C(0)\). In the absence of elastic rigidity (\(\mu _0=0\)) \(\gamma _{id}\) would correspond to the hydrostatic equilibrium.

    It is of note that \(\overline{\gamma }=\gamma _{id}\) implies \(\sigma _{\ell o}=0\). This fact is well explained in Van Hoolst et al. (2013) Sects. 1 and 2. In a simplified way their explanation is the following. “If the tidal response for static tides were to be as for the short-periodic tides, the sum of all tides would be aligned with the satellite-planet axis\(\ldots \) Therefore, there would be no gravitational torque on the satellite, unless a frozen-in asymmetry unrelated to tides would be present.” The “frozen-in asymmetry” is what we called prestress. Since we are using the approximation \(C(0)\approx C(i\sigma _{\ell o})\) in Eq. (6.9), we are indeed assuming that the tidal response to static tides is the same as that at frequency \(\sigma _{\ell o}\). Equation (6.9) is equivalent to equation (18) in Van Hoolst et al. (2013).

  14. Our complex compliance \(C(i\sigma _w)\) is equivalent to the complex compliance \({{\tilde{\kappa }}}\) in equation (37) of Mathews et al. (2002). Our generalised Maxwell rheology aims to describe the rheology of the mantle in a generalised sense including oceans, atmosphere, and other effects, as far as these effects can be considered in a spherically average sense.

  15. The motion of the pole of the mantle (\(\varvec{e}_3\in {\mathrm {K}}_m\)) in the guiding frame is given by

    $$\begin{aligned} ({\mathbb {I}} + \varvec{{\widehat{\alpha }}}_m)\varvec{e}_3= \varvec{e}_3+ \varvec{\alpha }_m\times \varvec{e}_3\in {\mathrm {K}}_g \end{aligned}$$

    If \(\varvec{\alpha }_m\) oscillates as an eigenmode with eigenvalue \(\lambda = i\,\omega (1+x)\) and eigenvector \({\varvec{\alpha }}_m=\epsilon (1,i,0)\) (this is the case of the NDFW and the FLL modes) then

    $$\begin{aligned}{\varvec{\alpha }}_m=\epsilon \mathrm{Re}\left( \exp [t\lambda ] \begin{bmatrix}1 \\ i\\ 0 \end{bmatrix}\right) = \epsilon \left( \begin{matrix}\ \ \cos \big (t \omega (1+ x)\big ) \\ - \sin \big (t\omega (1+ x)\big )\\ 0 \end{matrix}\right) =\epsilon \varvec{R}_3^{-1}\big ( t\omega \left( 1+x\right) \big )\varvec{e}_1\,. \end{aligned}$$

    Since the transition from the guiding frame to the slow frame is given by \(\varvec{R}_3(\omega t):{\mathrm {K}}_g\rightarrow {\mathrm {K}}_s\) we obtain that the image of the pole of the mantle in \({\mathrm {K}}_s\) is

    $$\begin{aligned} \varvec{R}_3(\omega t)\Big ( \varvec{e}_3+ \varvec{\alpha }_m\times \varvec{e}_3\Big )= \varvec{e}_3+\epsilon \Big ( \varvec{R}_3^{-1}\big ( t\omega x\big )\varvec{e}_1\Big )\times \varvec{e}_3\,. \end{aligned}$$

    So, the period associated with \(\lambda \) in the slow frame is \(2 \pi /(\omega x)\) and the motion is retrograde if \(x>0\) and prograde if \(x<0\).

  16. Suppose the motion of the slow frame is a precession about the normal to the invariable plane (see Eqs. (A.3) and (A.4)) with \({\mathbf {R}}_s={\mathbf {R}}_{\mathbf {3}}(\psi _g)\mathbf {R_1}(\theta _g){\mathbf {R}}_{\mathbf {3}}(\zeta ):{\mathrm {K}}_s\rightarrow \kappa \), where: \(\psi _g={{\dot{\psi }}}_g\, t \), \(\zeta =-{{\dot{\psi }}}_g\cos \theta _g \,t\), and \({{\dot{\psi }}}_g\) and \(\theta _g\) are constants. From the last equation in Footnote 15 we obtain that the motion of the pole in the inertial frame is given by the sum of the usual precessing vector \(\varvec{R}_3({{\dot{\psi }}}_g t )\varvec{R}_1(\theta _g)\varvec{e}_3\) plus the small “physical libration”

    $$\begin{aligned} \epsilon \varvec{R}_3({{\dot{\psi }}}_g\, t)\varvec{R}_1(\theta _g)\Big \{\Big (\varvec{R}_3^{-1} \big ( t(-{{\dot{\psi }}}_g \cos \theta _g+\omega x)\big )\varvec{e}_1\Big )\times \varvec{e}_3\Big \}\,. \end{aligned}$$

    If \(\theta _g= 0\), then the libration of the spin axis in the inertial frame is equal to that in the slow frame. If \(\theta _g\ne 0\), then the three components of the physical libration of the spin axis are

    $$\begin{aligned}\epsilon \left[ \begin{array}{ll} + \cos ^2(\theta _g/2) \sin \Big (t \big (2 {{\dot{\psi }}} \sin ^2(\theta _g/2)-x \omega \big )\Big )&{} -\sin ^2(\theta _g/2) \sin \Big (t \big (2 {{\dot{\psi }}} \cos ^2(\theta _g/2)+x \omega \big )\Big ) \\ - \cos ^2(\theta _g/2) \cos \Big (t \big (2 {{\dot{\psi }}} \sin ^2(\theta _g/2)-x \omega \big )\Big )&{} + \sin ^2(\theta _g/2) \cos \Big (t \big (2 {{\dot{\psi }}} \cos ^2(\theta _g/2)+x \omega \big )\Big ) \\ &{} -\sin \theta _g \cos \big (t ({{\dot{\psi }}}_g\cos \theta _g+x \omega )\big ) \end{array}\right] \,.\end{aligned}$$
  17. Eq. (37, \(\sigma _2\)) in Mathews et al. (2002) has an additional term \({{\tilde{\beta }}}\) that is due to the deformability of the mantle and the core-mantle magnetic coupling. These two effects are known to be important in the dynamics of the NDFW. Since they were not taken into account in our modelling, our formula, with the accepted value of \(f_c\) for the Earth, does not give the observed value of \(\sigma _{dw}\).

  18. The transformation between the precessional frame and the orbital frame is

    $$ \varvec{R}^{-1}_{or} \varvec{R}_{pr}= \varvec{R}^{-1}_1(\iota _p)\varvec{R}_3(\psi _g-\varOmega _p)\varvec{R}_1(\theta _g)=\varvec{R}_1(\theta _g-\iota _p)\ \text {or} \ {\left( \begin{array}{ccc} -1 &{} 0 &{} 0 \\ 0 &{} -\cos \left( \theta _g+\iota _p\right) &{} \sin \left( \theta _g+\iota _p\right) \\ 0 &{} \ \ \sin \left( \theta _g+\iota _p\right) &{} \cos \left( \theta _g+\iota _p\right) \\ \end{array} \right) }\,, $$

    where we used \( \psi _g=\varOmega _p \), for Cassini state 1 (e.g. Mercury), or \(\psi _g=\varOmega _p+\pi \), for Cassini state 2 (e.g. Moon), see Eq. (A.8). Note: \( \varvec{R}^{-1}_{or} \varvec{R}_{pr}:{\mathrm {K}}_{pr}\rightarrow {\mathrm {K}}_{or}\) does not depend on time.

  19. For a point mass of mass \(m_\beta \), with a low eccentricity and low inclination (\(\iota _\beta \)) orbit of anomalistic mean motion \(n_\beta \), \( s_\beta \approx \frac{3}{2}( \frac{n_{\beta }}{\omega })^2 \frac{m_\beta }{m_\beta +m}\). If we assume that the orbit inclinations \(\iota _\beta \) of the tidal generating point masses satisfy \(0<\iota _\beta <\frac{\arccos (-1/3)}{2}\), then \(s_\beta >0\) and \(\dot{\psi }_g<0\).

  20. It is enough to make \(\mathrm{I}_{\circ c}=0\) in equation (7.24) to obtain the inertial offset of a deformable solid body, which is the case considered in Baland et al. (2017). Equation (7.24) with the parameters in ibid. gives \(\delta _{m1}=0.003''\) and \(\delta _{m2}=-0.0055'\). The values obtained in ibid. (see Figure 4 and Table C3) are: \(\delta _{m1B}=\epsilon _\zeta =0.995''\) and \(\delta _{m2B}=\varDelta \epsilon _\varOmega =-0.006'\). The signs of the angles agree, \(\delta _{m2}\approx 0.92 \, \delta _{m2B}\), but the value we obtained for \(\delta _{m1}\approx \delta _{m1B}/260\) is much smaller.

  21. In Nastula and Gross (2015), and in many other references, we find the value of the “quality factor” \(Q_w\) associated with the Chandler’s wobble and not \(\mathrm{Re}\, \lambda _w\). The two quantities are related by \(\mathrm{Re}\, \lambda _w=- \sigma _w/(2\times Q_w)\). Note: the quality factor \(Q_w\) is not \(Q(\sigma _w)\). The reason for this difference is the following. In Eq. (4.25) the quality factor is given by \(Q^{-1}(\sigma )=\sin (\delta (\sigma ))\), where \(\delta (\sigma )\) is the phase lag of the body response to a tidal force with frequency \(\sigma \). The Chandler’s wobble eigenvalue \(\lambda _w\) is related to a natural mode of oscillation of the body in the absence of oscillatory-external forces and so, it cannot be associated with a phase lag. The relation \(\mathrm{Re}\, \lambda _w=- \sigma _w/(2\times Q_w)\) comes from the definition \(Q_w^{-1}=\frac{1}{2\pi E}\int \dot{E} dt\), where E is the energy stored and \(\frac{1}{2\pi }\int \dot{E} dt\) is the energy dissipated per cycle, applied to the free oscillations of an underdamped harmonic oscillator. In Gross (2007) (Table 11) there is a review of several estimates of \(Q_w\), which are in the range (30, 1000), and in Vondrák et al. (2017)) we find the estimate \(Q_w=35\).

  22. $$\begin{aligned} \varvec{S}_g =\varvec{J}_g- \frac{{{\,\mathrm{Tr}\,}}{\mathbf {J}}}{3}{\mathbb {I}} = \omega ^2 \left( \begin{array}{ccc} -\frac{1}{2}c_{20}+c_{22} &{} s_{22} &{} c_{21} \\ s_{22} &{} -\frac{1}{2}c_{20}-c_{22} &{} s_{21} \\ \ \ \ c_{21} &{}\ \ \ s_{21} &{}\ \ \ c_{20} \end{array} \right) \,, \end{aligned}$$
    (9.1)

    Up to first order in the eccentricity:

    $$\begin{aligned} \begin{aligned} c_{20}&=- \frac{G m_p}{a_p^3 \omega ^2}\big (1+ e\cos (M_p))\\ s_{22}&=\frac{3}{2}\frac{G m_p}{a_p^3 \omega ^2}\bigg (\sin (2 M_p-2\omega t) +\frac{e}{2} \Big (-\sin ( M_p-2\omega t)+7\sin (3M_p-2\omega t)\Big )\bigg ).\end{aligned} \end{aligned}$$

    The notation follows that in “Appendix A”.

  23. We remark that the complex compliance C(0) in this equation appears because we used the approximation \(C(i\,\sigma _{\ell o})\approx C(0)\) in order to obtain an estimate of \(\sigma _{\ell o}\). It seems natural to replace C(0) by \(C(i\chi )\) in the case of a forced oscillation.

  24. For the libration in longitude of the Moon and for the Kelvin-Voigt rheology with the data in Table 9, \(| \beta _{3}| /| \alpha _{m3}|=3 \frac{\omega ^2}{\gamma +\mu _0}/\overline{\gamma }=2 \times 10^{-3}\). In the notation of Eckhardt (1981) and Folkner et al. (2014) \(\alpha _{m3}=\tau \) is the longitudinal angle between the ideal Cassini state and the principal axes of “ the undistorted mantle” and \(\beta _{3}\) is the longitudinal angle between the principal axes of “ the undistorted mantle” and the principal axes of the “distorted mantle”. From Rambaux and Williams (2011) Table 3 the angle of longitudinal libration is \(1.8^{\prime \prime }\), or 15.23 meters at the equator. According to equation \(|\beta _3/\alpha _{m3}|=\frac{C(0)}{\overline{\gamma }}(\xi _2-\xi _1)\), the part due to the Moon’s deformation is 0.03 meters, which is of the order of magnitude of the typical horizontal displacement of 0.05 meters due to tidal variations detected by the Lunar Laser Ranger Williams and Boggs (2008) (p.109).

  25. The inclination of the symmetry axis of the core cavity with respect to \({\mathrm {K}}_m\) implies that \(\varvec{I}_{c,m}\) is not diagonal in \({\mathrm {K}}_m\). If the density of the fluid is constant throughout the cavity, then \(\varvec{I}_{c,m}\) will be clearly constant in time because the geometry of the cavity is fixed in \({\mathrm {K}}_m\). In Appendix D, we show that under certain hypothesis \(\varvec{I}_{c,m}\) can be constant in time even when the density of the fluid inside the cavity is not constant.

  26. Data sharing not applicable to this article as no datasets were generated or analysed during the current study.

  27. In this paper we slightly misused the angle \(\theta _g\). For instance, in the case of the Earth \({{\tilde{\theta }}}_g=23.44^\circ \) is the inclination of the Earth equator (the plane that is perpendicular to the average Earth’s rotation axis) to the ecliptic and we use it as if it were \(\theta _g\), namely the inclination of the Earth conventional north pole (the axis of largest moment of inertia of a rigid Earth) to the pole of the ecliptic. For the guiding motions in this appendix \({{\tilde{\theta }}}_g=\theta _g-\frac{{{\dot{\psi }}}_g}{\omega }\sin \theta _g\) implies \(\frac{|{{\tilde{\theta }}}_g-\theta _g|}{\theta _g}<\frac{|{{\dot{\psi }}}_g|}{\omega }\), so the relative error between the two angles is small.

  28. This resonance condition is due to Peale (1969), here presented in a form given in Boué (2020). Notice that the sidereal rotation period (the spin period) of the guiding motion is \(2\pi /\omega \), where \(\omega = {{\dot{\phi }}}_g+\dot{\psi }_g\cos \theta _g\). As discussed in the cited references, the angular velocity \( {{\dot{\phi }}}_g\) that appears in the resonance condition is the spin of the extended body relative to the frame \({\mathrm {K}}_{or}\) that precesses with the orbit of the point mass, given by \({\mathbf {R}}_{\mathbf {3}}(\varOmega _p)\mathbf {R_1}(\iota _p):{\mathrm {K}}_{or}\rightarrow \kappa \).

  29. Assume \(\iota _p=\theta _g=0\). If the orbital mean motion is synchronous with the constant spin then \(s=2\) and it is clear that the resonance \(M_p+\omega _p =\phi _g\) has the meaning stated in the first law. If \(s\ne 2\) then the relation \(s M_p=2(\phi _g-\omega _p)\) means that after a time interval equal to the anomalistic period, which is \(T=2\pi /\dot{M}_p\), the orbit has an angular displacement of \(2\pi +\, T {{\dot{\omega }}}_p\) rad while the smallest axis of inertia of the body has an angular displacement of \(s\, \pi +\, T{{\dot{\omega }}}_p\) rad. So, if initially the periapsis occurs on the smallest axis of inertia, then all other periapsis will occur on the same axis (if s is even, then always at the same side of the extended body).

  30. In Eckhardt (1981), the orientation of the body frame of the rigid Moon \({\mathrm {K}}\) with respect to an inertial frame \(\kappa \) is given by \(\varvec{R}_3(\psi )\varvec{R}_1^{-1}(\theta )\varvec{R}_3(\phi ):{\mathrm {K}}\rightarrow \kappa \), where: \(\psi \) is the longitude of the descending node of the lunar equator, \(\theta \) is the inclination of the lunar equator to the ecliptic, and \(\phi \) is the angle between the descending node and the axis of smallest moment of inertia pointing towards the Earth. The Cassini’s laws are equivalent to \((\psi ,\theta ,\phi )= (\varOmega _p, \iota _p, \pi +M_p+\omega _p)\), so Eckhardt’s guiding frame is \(\varvec{R}_3(\varOmega _p)\varvec{R}_1^{-1}(\iota _p)\varvec{R}_3(\pi +M_p+\omega _p):{\mathrm {K}}_g\rightarrow \kappa \), which coincides with ours, namely \(\varvec{R}_3(\varOmega _p+\pi )\varvec{R}_1(\iota _p)\varvec{R}_3(M_p+\omega _p):{\mathrm {K}}_g\rightarrow \kappa \).

  31. A table of the Hansen coefficients \( X_s^{-3,2}(e)\), for \(-4\le s\le 8\) and eccentricity up to order \(e^6\), can be found in Correia et al. (2014). In this paper we use

    $$\begin{aligned} \begin{aligned} X_0^{-3,0}(e)&=(1-e^2)^{-3/2}\\ X_2^{-3,2}(e)&= 1 - \frac{5}{2} e^2 + \frac{13}{16} e^4 - \frac{35}{288} e^6\\ X_3^{-3,2}(e)&= \frac{7}{2} e - \frac{123}{16} e^3 + \frac{489}{128} e^5 \end{aligned}\,. \end{aligned}$$
    (A.12)
  32. Note that equation (D.18) implies the relation relation \(\varvec{B}=-\)Diagonal\(\{\epsilon _2+\epsilon _3,\epsilon _1+\epsilon _3, \epsilon _1+\epsilon _2\}\) between the inertial deformation matrix \(\varvec{B}\) associated with \( \varvec{I}_c(0)\) and the geometric quantities \(\epsilon _1,\epsilon _2,\epsilon _3\). This simple relation holds only because the density of the fluid is constant over homothetic ellipsoids.

  33. The coordinates in \({\mathrm {K}}_{pr}\) used in Stewartson and Roberts (1963) are related to ours by the map \((\varvec{e}_1, \varvec{e}_2)\rightarrow (-\varvec{e}_2, \varvec{e}_1)\) and their angle \(\alpha \) is equal to \(\pi -\theta _g\). In ibid. the motion of the mantle is the motion of our guiding frame \(\varvec{R}_g={\mathbf {R}}_{\mathbf {3}}( \psi _g )\mathbf {R_1}(\theta _g){\mathbf {R}}_{\mathbf {3}}(\phi _g) :{\mathrm {K}}_m\rightarrow \kappa \), which implies that the motion of \({\mathrm {K}}_m\) with respect to \({\mathrm {K}}_{pr}\) is \({\mathbf {R}}_{\mathbf {3}}(\phi _g):{\mathrm {K}}_m\rightarrow {\mathrm {K}}_{pr}\). Since the cavity is an ellipsoid of revolution it remains at rest in \({\mathrm {K}}_{pr}\) while its boundary rotates with angular velocity \({{\dot{\phi }}}_g\varvec{e}_3\) (what we call \({{\dot{\phi }}}_g\) they call \(\omega \)). So, if the precessional velocity would be zero, then \({\mathrm {K}}_{pr}\) would be an inertial frame and viscosity would eventually bring the fluid to rotate, at least in the average, as a rigid body. This explains the term \({{\dot{\phi }}}_g\varvec{e}_3\times \varvec{x}\) in the velocity field \(\varvec{u}\). The precessional angular velocity, which in \({\mathrm {K}}_{pr}\) is \({{\dot{\psi }}}_g(\sin \theta _g\varvec{e}_2+\cos \theta _g\varvec{e}_3)\), induces inertial forces upon the fluid that generate the additional non-rigid term to \(\varvec{u}\).

  34. The correspondence between the notation in Boué (2020) and ours is: \(\omega _p\rightarrow {{\dot{\phi }}}_g\), \(\alpha _c\rightarrow f_c\), \(g\rightarrow {{\dot{\psi }}}_g\), \(\alpha \rightarrow \overline{\alpha }_e\), \(\beta \rightarrow \overline{\gamma }\), \(C_c\rightarrow \overline{I}_{c3}\), \(C_m\rightarrow \overline{I}_{m3}\),\(C\rightarrow \overline{I}_3\), \(\theta _m^\prime \rightarrow {{{\chi }}}-\delta _m\), \(\theta _c^\prime \rightarrow {{{\chi }}}-\delta _c\), \(\theta _m^\prime -\iota \rightarrow \theta _g-\delta _m\), and \(\theta _c^\prime -\iota \rightarrow \theta _g-\delta _c\). These correspondences hold for both Cassini states 1 and 2.

References

  • Arnold, V.I.: Ordinary Differential Equations. Springer Textbook. Springer, Berlin (1992)

    Google Scholar 

  • Baland, Rose-Marie., Yseboodt, Marie, Rivoldini, Attilio, Van Hoolst, Tim: Obliquity of Mercury: Influence of the precession of the pericenter and of tides. Icarus 291, 136–159 (2017)

    Article  ADS  Google Scholar 

  • Batchelor, G.K.: An Introduction to Fluid Dynamics. Cambridge University Press, Cambridge (2000)

    Book  Google Scholar 

  • Beuthe, Mikael: Tidal Love numbers of membrane worlds: Europa, Titan, and Co. Icarus 258, 239–266 (2015)

    Article  ADS  Google Scholar 

  • Bland, D.R.: Linear Viscoelasticity. Pergamon Press, Oxford (1960)

    MATH  Google Scholar 

  • Boué, Gwenaël: Cassini states of a rigid body with a liquid core. Celest. Mech. Dyn. Astron. 132, 1–26 (2020)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Boué, Gwenaël, Rambaux, Nicolas, Richard, Andy: Rotation of a rigid satellite with a fluid component: a new light onto Titan’s obliquity. Celest. Mech. Dyn. Astron. 129(4), 449–485 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Běhounková, Marie, Tobie, Gabriel, Choblet, Gaël., Čadek, Ondřej: Coupling mantle convection and tidal dissipation: applications to Enceladus and Earth-like planets. J. Geophys. Res. 115(E9), E09011 (2010). https://doi.org/10.1029/2009JE003564

    Article  ADS  Google Scholar 

  • Běhounková, Marie, Souček, Ondřej, Hron, Jaroslav, Čadek, Ondřej: Plume activity and tidal deformation on enceladus influenced by faults and variable ice shell thickness. Astrobiology 17(9), 941–954 (2017). https://doi.org/10.1089/ast.2016.1629

    Article  ADS  Google Scholar 

  • Capitaine, Nicole, Guinot, B., Souchay, J.: A non-rotating origin on the instantaneous equator: definition, properties and use. Celest. Mech. 39(3), 283–307 (1986)

    Article  ADS  MATH  Google Scholar 

  • Chandrasekhar, Subrahmanyan: Ellipsoidal Figures of Equilibrium. Yale University Press, London (1969)

    MATH  Google Scholar 

  • Colombo, G.: Cassini’s second and third laws. Astron. J. 71, 891 (1966)

    Article  ADS  Google Scholar 

  • Correia, A.C.M., Ragazzo, C., Ruiz, L.S.: The effects of deformation inertia (kinetic energy) in the orbital and spin evolution of close-in bodies. Celest. Mech. Dyn. Astron. 130(8), 51 (2018)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Correia, Alexandre CM., Delisle, Jean-Baptiste.: Spin-orbit coupling for close-in planets. Astron. Astrophys. 630, A102 (2019)

    Article  Google Scholar 

  • Correia, Alexandre CM., Boué, Gwenaël, Laskar, Jacques, Rodríguez, Adrián: Deformation and tidal evolution of close-in planets and satellites using a Maxwell viscoelastic rheology. Astron. Astrophys. 571, A50 (2014)

    Article  ADS  Google Scholar 

  • Deleplace, Bérangère., Cardin, Philippe: Viscomagnetic torque at the core mantle boundary. Geophys. J. Int. 167(2), 557–566 (2006)

    Article  ADS  Google Scholar 

  • Eckhardt, Donald H.: Theory of the libration of the Moon. Moon Planets 25(1), 3–49 (1981)

    Article  ADS  MATH  Google Scholar 

  • Efroimsky, Michael: Bodily tides near spin-orbit resonances. Celest. Mech. Dyn. Astron. 112(3), 283–330 (2012)

    Article  ADS  MathSciNet  Google Scholar 

  • Efroimsky, Michael: Tidal dissipation compared to seismic dissipation: In small bodies, Earths, and super-Earths. Astrophys. J. 746(2), 150 (2012)

    Article  ADS  Google Scholar 

  • Ferraz-Mello, Sylvio, Beaugé, Cristian, Folonier, Hugo A., Gomes, Gabriel O.: Tidal friction in satellites and planets. The new version of the creep tide theory. Eur. Phys. J. Special Topics 229, 1441–1462 (2020)

    Article  ADS  Google Scholar 

  • Fienga, A., Deram, P., Viswanathan, V., Di Ruscio, A., Bernus, L., Durante, D., Gastineau, M., Laskar, J.: INPOP19a planetary ephemerides. Notes Scientifiques et Techniques de l’Institut de Mécanique Céleste, 109, (December 2019)

  • Folkner, W.M., Williams, J.G., Boggs, D.H., Park, R.S., Kuchynka, P.: The Planetary and Lunar Ephemerides DE430 and DE431. Interplanet. Netw. Progress Rep. 42–196, 1–81 (2014)

    Google Scholar 

  • Folonier, Hugo A., Ferraz-Mello, Sylvio: Tidal synchronization of an anelastic multi-layered body: Titan’s synchronous rotation. Celest. Mech. Dyn. Astron. 129(4), 359–396 (2017). https://doi.org/10.1007/s10569-017-9777-5

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Genova, Antonio, Goossens, Sander, Lemoine, Frank G., Mazarico, Erwan, Neumann, Gregory A., Smith, David E., Zuber, Maria T.: Seasonal and static gravity field of Mars from MGS, Mars Odyssey and MRO radio science. Icarus 272, 228–245 (2016)

    Article  ADS  Google Scholar 

  • Gevorgyan, Yeva: Homogeneous model for the TRAPPIST-1e planet with an icy layer. Astron. Astrophys. 650, A141 (2021). https://doi.org/10.1051/0004-6361/202140736

    Article  ADS  Google Scholar 

  • Gevorgyan, Yeva, Boué, Gwenaël, Ragazzo, Clodoaldo, Ruiz, Lucas S., Correia, Alexandre C.M.: Andrade rheology in time-domain. application to Enceladus’ dissipation of energy due to forced libration. Icarus, 343:113610, (2020). https://doi.org/10.1016/j.icarus.2019.113610. URL http://www.sciencedirect.com/science/article/pii/S0019103519305020

  • Glickman. Glossary of Meteorology. American Meteorological Society (2000)

  • Goldreich, Peter, Toomre, Alar: Some remarks on polar wandering. J. Geophys. Res. 74(10), 2555–2567 (1969)

    Article  ADS  Google Scholar 

  • Gross, Richard S.: Earth rotation variations-long period. Treatise Geophys. 3, 239–294 (2007)

    Article  Google Scholar 

  • Guinot, B.: Basic problems in the kinematics of the rotation of the Earth. In Symposium-International Astronomical Union, Vol. 82, pp. 7–18. Cambridge University Press, Cambridge (1979)

  • Henrard, Jacques: The rotation of Io with a liquid core. Celest. Mech. Dyn. Astron. 101(1–2), 1–12 (2008)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Holm, D.D., Schmah, T., Stoica, C.: Geometric Mechanics and Symmetry. Oxford University Press, New York (2009)

    MATH  Google Scholar 

  • Hough, Sydney Samuel: Xii. the oscillations of a rotating ellipsoidal shell containing fluid. Philosoph. Trans. R Soc. Lond. A 186, 469–506 (1895)

    Article  ADS  MATH  Google Scholar 

  • Iess, Luciano, Stevenson, D.J., Parisi, M., Hemingway, D., Jacobson, R.A., Lunine, J.I., Nimmo, F., Armstrong, J.W., Asmar, S.W., Ducci, M., et al.: The gravity field and interior structure of Enceladus. Science 344(6179), 78–80 (2014)

    Article  ADS  Google Scholar 

  • Jacobson, R.A., Lainey, V.: Martian satellite orbits and ephemerides. Planet. Space Sci. 102, 35–44 (2014)

    Article  ADS  Google Scholar 

  • Lainey, Valéry: Quantification of tidal parameters from Solar System data. Celest. Mech. Dyn. Astron. 126(1–3), 145–156 (2016)

    Article  ADS  Google Scholar 

  • Lamb, H.: Hydrodynamics, 6th edn. Cambridge Mathematical Library, Cambridge (1932)

    MATH  Google Scholar 

  • Lambeck, K.: The Earth’s Variable Rotation: Geophysical Causes and Consequences. Cambridge University Press, UK (1980)

    Book  Google Scholar 

  • Margot, J.-L., Hauck II, S.A., Mazarico, E., Peale, S.J., Padovan, S.: Mercury’s internal structure. Mercury-The view after MESSENGER, pp. 85–113, (2018)

  • Mathews, P.M., Herring, T.A., Buffett, B.A.: Modeling of nutation and precession: New nutation series for nonrigid Earth and insights into the Earth’s interior. J. Geophys. Res. Solid Earth, 107 (B4):ETG–3 (2002)

  • Matsui, Hiroaki, Buffett, Bruce A.: Large-eddy simulations of convection-driven dynamos using a dynamic scale-similarity model. Geophys. Astrophys. Fluid Dyn. 106(3), 250–276 (2012)

    Article  ADS  MathSciNet  Google Scholar 

  • Matsuyama, Isamu: Tidal dissipation in the oceans of icy satellites. Icarus 242, 11–18 (2014). https://doi.org/10.1016/j.icarus.2014.07.005

    Article  ADS  Google Scholar 

  • Matsuyama, Isamu, Beuthe, Mikael, Hay, Hamish C. F. C.., Nimmo, Francis, Kamata, Shunichi: Ocean tidal heating in icy satellites with solid shells. Icarus 312, 208–230 (2018). https://doi.org/10.1016/j.icarus.2018.04.013

    Article  ADS  Google Scholar 

  • Mazarico, Erwan, Genova, Antonio, Goossens, Sander, Lemoine, Frank G., Neumann, Gregory A., Zuber, Maria T., Smith, David E., Solomon, Sean C.: The gravity field, orientation, and ephemeris of Mercury from MESSENGER observations after three years in orbit. J. Geophys. Res. Planets 119(12), 2417–2436 (2014)

    Article  ADS  Google Scholar 

  • McKenzie, Dan P.: The viscosity of the lower mantle. J. Geophys. Res. 71(16), 3995–4010 (1966)

    Article  ADS  Google Scholar 

  • Munk, W.H., MacDonald, G.J.F.: The Rotation of the Earth. Cambridge University Press, New York (1961)

    MATH  Google Scholar 

  • Nastula, J., Gross, R.: Chandler wobble parameters from SLR and GRACE. J. Geophys. Res. Solid Earth 120(6), 4474–4483 (2015)

    Article  ADS  Google Scholar 

  • Nimmo, F., Pappalardo, R.T.: Ocean worlds in the outer solar system. J. Geophys. Res. 121(8), 1378–1399 (2016). https://doi.org/10.1002/2016JE005081

    Article  Google Scholar 

  • Noyelles, Benoît: Rotation of a synchronous viscoelastic shell. Mon. Not. R. Astron. Soc. 474(4), 5614–5644 (2018)

    Article  ADS  Google Scholar 

  • Park, Ryan S., Folkner, William M., Williams, James G., Boggs, Dale H.: The JPL Planetary and Lunar Ephemerides DE440 and DE441. Astron. J. 161(3), 105 (2021). https://doi.org/10.3847/1538-3881/abd414

    Article  ADS  Google Scholar 

  • Peale, Stanton J.: Generalized Cassini’s laws. Astron. J. 74, 483 (1969)

    Article  ADS  MATH  Google Scholar 

  • Peale, Stanton J.: Possible histories of the obliquity of Mercury. Astron. J. 79, 722 (1974)

    Article  ADS  Google Scholar 

  • Peale, Stanton J., Margot, Jean-Luc., Hauck, Steven A., Solomon, Sean C.: Effect of core-mantle and tidal torques on Mercury’s spin axis orientation. Icarus 231, 206–220 (2014). https://doi.org/10.1016/j.icarus.2013.12.007

    Article  ADS  Google Scholar 

  • Pedlosky, Joseph: Geophysical Fluid Dynamics. Springer Science & Business Media, New York (2013)

    MATH  Google Scholar 

  • Petit, G., Luzum, B.: IERS conventions (2010). Technical report, DTIC Document (2010)

  • Poincaré, Henri: Sur l’équilibre d’une masse fluide animée d’un mouvement de rotation. Acta Math. 7(1), 259–380 (1885)

    Article  MathSciNet  MATH  Google Scholar 

  • Poincaré, Henri: Sur la précession des corps déformables. Bulletin Astronomique, Serie I(27), 321–356 (1910)

    ADS  MATH  Google Scholar 

  • Ragazzo, C., Ruiz, L.S.: Viscoelastic tides: models for use in Celestial Mechanics. Celest. Mech. Dyn. Astron. 128(1), 19–59 (2017). https://doi.org/10.1007/s10569-016-9741-9

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Ragazzo, C., Ruiz, L.S.: Dynamics of an isolated, viscoelastic, self-gravitating body. Celest. Mech. Dyn. Astron. 122(4), 303–332 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Ragazzo, Clodoaldo: The theory of figures of Clairaut with focus on the gravitational modulus: inequalities and an improvement in the Darwin-Radau equation. São Paulo J. Math. Sci. 14, 1–48 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  • Ragazzo, Clodoaldo, Ruiz, L.S.: Viscoelastic tides: models for use in Celestial Mechanics. Celest. Mech. Dyn. Astron. 128(1), 19–59 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Rambaux, N., Williams, J.G.: The Moon’s physical librations and determination of their free modes. Celest. Mech. Dyn. Astron. 109(1), 85–100 (2011)

    Article  ADS  MATH  Google Scholar 

  • Roberts, PH, Stewartson, K: On the motion of a liquid in a spheroidal cavity of a precessing rigid body. II. In Mathematical Proceedings of the Cambridge Philosophical Society, Vol. 61, pp. 279–288. Cambridge University Press, (1965)

  • Rochester, M.G., Smylie, D.E.: On changes in the trace of the Earth’s inertia tensor. J. Geophys. Res. 79(32), 4948–4951 (1974)

    Article  ADS  Google Scholar 

  • Stark, Alexander, Oberst, Jürgen., Preusker, Frank, Peale, Stanton J., Margot, Jean-Luc., Phillips, Roger J., Neumann, Gregory A., Smith, David E., Zuber, Maria T., Solomon, Sean C.: First MESSENGER orbital observations of Mercury’s librations. Geophys. Res. Lett. 42(19), 7881–7889 (2015). https://doi.org/10.1002/2015GL065152

    Article  ADS  Google Scholar 

  • Steinbrügge, G., Padovan, S., Hussmann, H., Steinke, T., Stark, A., Oberst, J.: Viscoelastic tides of mercury and the determination of its inner core size. J. Geophys. Res. Planets 123(10), 2760–2772 (2018)

    Article  ADS  Google Scholar 

  • Stewartson, K., Roberts, P.H.: On the motion of liquid in a spheroidal cavity of a precessing rigid body. J. Fluid Mech. 17(1), 1–20 (1963)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Thomas, P.C., Tajeddine, R., Tiscareno, M.S., Burns, J.A., Joseph, J., Loredo, T.J., Helfenstein, P., Porco, C.: Enceladus’s measured physical libration requires a global subsurface ocean. Icarus 264, 37–47 (2016). https://doi.org/10.1016/j.icarus.2015.08.037

    Article  ADS  Google Scholar 

  • Tilgner, Andreas, Busse, F.H.: Fluid flows in precessing spherical shells. J. Fluid Mech. 426, 387 (2001)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Triana, Santiago Andrés, Rekier, Jérémy., Trinh, Antony, Dehant, Veronique: The coupling between inertial and rotational eigenmodes in planets with liquid cores. Geophys. J. Int. 218(2), 1071–1086 (2019)

    Article  ADS  Google Scholar 

  • Van Hoolst, T., Baland, R., Trinh, A.: On the librations and tides of large icy satellites. Icarus 226(1), 299–315 (2013). https://doi.org/10.1016/j.icarus.2013.05.036. URL http://www.sciencedirect.com/science/article/pii/S0019103513002364

  • Viswanathan, V., Rambaux, N., Fienga, A., Laskar, J., Gastineau, M.: Observational constraint on the radius and oblateness of the Lunar Core-Mantle boundary. Geophys. Res. Lett. 46(13), 7295–7303 (2019). https://doi.org/10.1029/2019GL082677

    Article  ADS  Google Scholar 

  • Vondrák, Jan, Ron, Cyril, Chapanov, Ya.: New determination of period and quality factor of Chandler wobble, considering geophysical excitations. Adv. Space Res. 59(5), 1395–1407 (2017)

    Article  ADS  Google Scholar 

  • Williams, James G.: Contributions to the Earth’s obliquity rate, precession, and nutation. Astron. J. 108, 711–724 (1994)

    Article  ADS  Google Scholar 

  • Williams, J.G., Boggs, D.H.: Lunar core and mantle. what does LLR see. In Proceedings of the 16th International Workshop on Laser Ranging, Poznan, Poland, Vol. 1317, (2008)

  • Williams, James G., Boggs, Dale H., Yoder, Charles F., Todd Ratcliff, J., Dickey, Jean O.: Lunar rotational dissipation in solid body and molten core. J. Geophys. Res. Planets 106(E11), 27933–27968 (2001)

    Article  ADS  Google Scholar 

  • Williams, James G., Konopliv, Alexander S., Boggs, Dale H., Park, Ryan S., Yuan, Dah-Ning., Lemoine, Frank G., Goossens, Sander, Mazarico, Erwan, Nimmo, Francis, Weber, Renee C., et al.: Lunar interior properties from the GRAIL mission. J. Geophys. Res. Planets 119(7), 1546–1578 (2014)

    Article  ADS  Google Scholar 

  • Yan, Jianguo, Goossens, Sander, Matsumoto, Koji, Ping, Jinsong, Harada, Yuji, Iwata, Takahiro, Namiki, Noriyuki, Li, Fei, Tang, Geshi, Cao, Jianfeng, et al.: CEGM02: An improved lunar gravity model using Chang’E-1 orbital tracking data. Planet. Space Sci. 62(1), 1–9 (2012)

    Article  ADS  Google Scholar 

  • Yoder, Charles F.: Astrometric and geodetic properties of Earth and the Solar System. Glob. Earth Phys. A Handbook Phys. Const. 1, 1–31 (1995)

    ADS  Google Scholar 

  • Zanazzi, J.J., Lai, Dong: Triaxial deformation and asynchronous rotation of rocky planets in the habitable zone of low-mass stars. Mon. Not. R. Astron. Soc. 469(3), 2879–2885 (2017)

    Article  ADS  Google Scholar 

  • Zhang, Wenying, Shen, Wenbin: New estimation of triaxial three-layered Earth’s inertia tensor and solutions of Earth rotation normal modes. Geodesy Geodyn. 11(5), 307–315 (2020)

    Article  Google Scholar 

Download references

Acknowledgements

We thank Prof. Hervé Manche (Observatoire de Paris, IMCCE) for the data used to generate Fig. 13-Left and also for helping with the interpretation of the results. We also thank Prof. V. Viswanathan (NASA Goddard Space Flight Center; University of Maryland) for the discussions about the INPOP19a model and about the offset of the centre of libration in Fig.  13-Left. CR is partially supported by FAPESP Grant 2016/25053-8. YG is partially supported by FAPESP Grants 2015/26253-8 and 2018/02905-4. LR was partially supported by CFisUC projects (UIDB/04564/2020 and UIDP/04564/2020), and ENGAGE SKA (POCI-01-0145-FEDER-022217), funded by COMPETE 2020 and FCT, Portugal.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Clodoaldo Ragazzo.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

The value of the coefficients \(c_1,c_2,c_3\) in Eq.  (2.23) for a body in a Cassini state

The theoretical determination of the almost equilibrium orientation of the spin axis of an extended body, or the guiding motion as described in Sect. 2, is a difficult task that requires some knowledge about the rheological behaviour of the body. An important situation, which will be the one considered in this appendix, is that of a rigid body under the influence of a single point mass that moves according to a Keplerian precessing orbit. In this case the guiding motion is called “a Cassini state” Peale (1969). We remark that different Cassini states can be obtained under different rheological hypotheses (see, for instance, Boué (2020) for bodies with a rigid mantle and a fluid core).

Let \(\kappa \) be the inertial frame and \({\mathrm {K}}_g\) be the guiding frame described in Sect. 2. We recall that the common origin of both frames is the centre of mass of the extended body. The position of the point mass with respect to the extended body in the frame \(\kappa \) is given by

$$\begin{aligned} {\mathbf {r}}={\mathbf {R}}_{\mathbf {3}}(\varOmega _p)\mathbf {R_1}(\iota _p){\mathbf {R}}_{\mathbf {3}}(\omega _p) \begin{bmatrix} r \cos (f_p) \\ r \sin (f_p) \\ 0 \end{bmatrix}, \end{aligned}$$
(A.1)

where: \(\iota _p\), \(f_p\), \(\omega _p\), and \(\varOmega _p\) are the inclination with respect to an invariable plane (the “Laplace plane”), the true anomaly, the argument of the periapsis and the longitude of the ascending node, respectively, of a classical Keplerian orbit that has semi-major axis \(a_p\), eccentricity e, and mean anomaly \(M_p\) (the index p stands for point-mass).

We recall that the three axes of \({\mathrm {K}}_g\) are those of the body’s principal moments of inertia, \(\overline{I}_{1}< \overline{I}_{2}< \overline{I}_{3}\), in the absence of deformation. The orientation of the guiding frame \(\mathbf {R_g}:{\mathrm {K}}_g\rightarrow \kappa \) is given by three Euler angles

$$\begin{aligned} \mathbf {R_g}={\mathbf {R}}_{\mathbf {3}}(\psi _g)\mathbf {R_1}(\theta _g){\mathbf {R}}_{\mathbf {3}}(\phi _g), \end{aligned}$$
(A.2)

where: \(\theta _g\) is the inclination of the body equator to the Laplace planeFootnote 27, \(\psi _g\) is the longitude of the ascending node of the body equator, and \(\phi _g\) is the angle between the ascending node and Axis 1. There are two possible orientations of Axis 1, so \(\varvec{R}_g\) is not uniquely defined. In the case of the Moon it is usual to choose the positive orientation of Axis 1 as that pointing towards the Earth (see, e.g. Eckhardt 1981). To change from one orientation to the other it is enough to change the signs of both unit vectors \(\varvec{e}_1\) and \(\varvec{e}_2\).

The orientation of the slow frame is given by \({\mathbf {R}}_s(t)= {\mathbf {R}}_g(t){\mathbf {R}}_{\mathbf {3}}^{-1}(\omega \, t):{\mathrm {K}}_s\rightarrow \kappa \), Eq. (2.14), that in the present case becomes

$$\begin{aligned} {\mathbf {R}}_s={\mathbf {R}}_{\mathbf {3}}(\psi _g)\mathbf {R_1}(\theta _g){\mathbf {R}}_{\mathbf {3}}(\zeta )\quad \text {where}\quad \zeta =\phi _g-\omega t\,. \end{aligned}$$
(A.3)

The identity \(\langle \varvec{\omega }_{s,g}\,,\varvec{e}_3\rangle =0\) implies

$$\begin{aligned} {{\dot{\zeta }}}=- {{\dot{\psi }}}_g\cos \theta _g\,. \end{aligned}$$
(A.4)

The position of the point mass in the guiding frame \({\mathrm {K}}_g\) is given by \(\mathbf {r_g}=\mathbf {R_g}^{-1}{\mathbf {r}}\) and Eqs. (A.1) and (A.2) imply

$$\begin{aligned} \mathbf {r_g}=r \Big ( \cos \big (f_p\big ) \varvec{u} + \sin \big (f_p\big ) \varvec{v} \Big ), \end{aligned}$$
(A.5)

where \(\varvec{u}\) and \(\varvec{v}\) are the first and second columns, respectively, of the matrix \(\mathbf {R_g}^{-1}{\mathbf {R}}_{\mathbf {3}}(\varOmega _p)\mathbf {R_1}(\iota _p){\mathbf {R}}_{\mathbf {3}}(\omega _p)\). The force matrix in the guiding frame \({\mathbf {J}}_{g}\), as defined in Sect. 2, is given by

$$\begin{aligned} \begin{aligned}&J_{gij}=\bigg (\frac{3\mathcal{G}m_p}{|\mathbf {r_g}|^5}\mathbf {r_g}\otimes \mathbf {r_g} \bigg )_{ij}\\&\ \quad = \frac{3\mathcal{G}m_p}{2r^3}\bigg (\big (u_iu_j+v_iv_j\big )+ \cos \big (2f_p\big )\big (u_iu_j-v_iv_j\big )+ \sin \big (2f_p\big )\big (u_iv_j+v_iu_j\big )\bigg )\, .\end{aligned} \end{aligned}$$
(A.6)

Notice that the vectors \(\varvec{u}\) and \(\varvec{v}\) in Eq. (A.5) are the rows of an orthogonal matrix, which implies \(|\varvec{u}|=|\varvec{v}|=1\) and \(\langle \varvec{u},\varvec{v}\rangle =0\). So,

$$\begin{aligned} {{\,\mathrm{Tr}\,}}\varvec{J}_g=\frac{3\mathcal{G}m_p}{r^3}\, . \end{aligned}$$
(A.7)

Now we impose that the extended body is in a Cassini state that is characterised by three laws Colombo (1966), Peale (1969), Peale (1974), which are generalisations of the original laws of Cassini (1693) (aimed to describe the motion of the Moon, see Eckhardt (1981)).

  1. 1-

    The body spin is in s-to-2 resonance with the orbital mean motion, where s is a positive integer (for the Moon \(s=2\) and for Mercury \(s=3\));

  2. 2-

    The inclination of the body equator \(\theta _g\) with respect to the Laplace plane is constant;

  3. 3-

    Either the ascending node (state 1) or the descending node (state 2) of the body equator on the Laplace plane precesses in coincidence with the ascending node of the orbit on the Laplace plane (the spin axis, orbit normal, and Laplace plane normal are coplanar).

The third law implies that either

$$\begin{aligned} \begin{array}{rll} \psi _g&{}=\varOmega _p \quad &{}\text {state 1 (e.g. Mercury) or}\\ \psi _g&{}=\varOmega _p+\pi \quad &{}\text {state 2 (e.g. Moon),} \end{array} \end{aligned}$$
(A.8)

then the second law implies that the inclination of the body spin axis with respect to the normal to the orbital plane, given by (\(\theta _g\ge 0\), \(\iota _p\ge 0\))

$$\begin{aligned} \begin{array}{rll} \chi _p&{}=\theta _g-\iota _p\ge 0\quad &{} \text {state 1}\\ \chi _p&{}=\theta _g+\iota _p\quad &{}\text {state 2}\end{array} \end{aligned}$$
(A.9)

is constant (the inequality in state 1 is of dynamical origin, see Peale (1969) paragraph Eq. (18)); and then the first law implies that \(s M_p=2(\phi _g-\omega _p)\)Footnote 28\(^{,}\)Footnote 29.

For a body in a Cassini state 2 (e.g. Moon) Eq. (A.5) becomesFootnote 30

$$\begin{aligned} \mathbf {r_g}=r \cos f_p\underbrace{\begin{bmatrix} -\cos \phi _g \cos \omega _p - \cos \chi _p \sin \phi _g \sin \omega _p\\ \cos \omega _p \sin \phi _g - \cos \phi _g \cos \chi _p \sin \omega _p\\ \sin \chi _p \sin \omega _p \end{bmatrix}}_{ \varvec{u}}+ r \sin f_p\underbrace{\begin{bmatrix} -\cos \chi _p \cos \omega _p \sin \phi _g + \cos \phi _g \sin \omega _p\\ -\cos \phi _g \cos \chi _p \cos \omega _p - \sin \phi _g \sin \omega _p\\ \cos \omega _p \sin \chi _p \end{bmatrix}}_{\varvec{v}} \qquad \end{aligned}$$
(A.10)

and for a body in a Cassini state 1 we must replace \(\varvec{u}\rightarrow -\varvec{u}\) and \(\varvec{v}\rightarrow -\varvec{v}\). Since \(J_{gij}\) given in Eq. (A.6) depends only on the products of components of \(\varvec{u}\) and \(\varvec{v}\), the matrix \(\varvec{J}_g\) has the same expression for both states 1 and 2. Therefore, the following computation of \(c_1\), \(c_2\) and \(c_3\) holds in both cases.

The coefficients \(c_1,c_2,\) and \(c_3\) we aim to compute are determined by the constant term in the Fourier expansion of \(J_{gij}\) given in Eq. (A.6). In order to compute the Fourier expansion of the factors involving r and \(f_p\) we use the Hansen coefficients defined byFootnote 31

$$\begin{aligned} \bigg (\frac{r}{a_p}\bigg )^n\mathrm{e} ^{i m f}=\sum _{k=-\infty }^\infty X^{n,m}_k(e) \mathrm{e} ^{i k M}. \end{aligned}$$
(A.13)

The computation of the coefficients \(c_1,c_2,\) and \(c_3\) in Eq. (2.23) can be done in the following way. At first we use Eq. (A.7) to obtain

$$\begin{aligned} c_3=\frac{\mathcal{G}m_p}{\omega ^2 a_p^3} X_0^{-3,0}(e)\, . \end{aligned}$$
(A.14)

The term \( \left( \frac{3\mathcal{G}m_p}{\omega ^2|\mathbf {r_g}|^5}\mathbf {r_g}\otimes \mathbf {r_g}\right) _{33}\) can be easily computed using Eqs. (A.6) and (A.12) and the result is \(\frac{3\mathcal{G}m_p}{2 \omega ^2 a_p^3} X_0^{-3,0}(e)\sin ^2(\chi _p)=c_3-2c_1/3\). So, we obtain that

$$\begin{aligned} c_1=\frac{3\,\mathcal{G}m_p}{2\,\omega ^2 a_p^3} X_0^{-3,0}(e)\bigg (1-\frac{3}{2}\sin ^2(\chi _p)\bigg )\,. \end{aligned}$$
(A.15)

In order to compute the constant \(c_2\) associated with an \(s-to-2\) spin-orbit resonance it is enough to compute the constant term in the Fourier expansion of \(\left( \frac{3\mathcal{G}m_p}{\omega ^2|\mathbf {r_g}|^5}\mathbf {r_g}\otimes \mathbf {r_g}\right) _{11}\) and then to use that this term is equal to \(c_1/3+c_2+c_3\), Eq. (2.23). Assuming that the precession of the periapsis is different from zero a computation gives

$$\begin{aligned} c_2=\frac{3\,\mathcal{G}m_p}{2\,\omega ^2 a_p^3} X_s^{-3,2}(e) \cos ^4(\chi _p/2). \end{aligned}$$
(A.11)

Another case of interest is that of a point mass that moves in a precessing Keplerian orbit, as that parameterised in Eq. (A.1), with constant inclination, \(\iota _p=\)constant, and with no spin-orbit resonance. By no spin-orbit resonance we mean that the constant frequencies \({{\dot{\psi }}}_g\), \({{\dot{\varOmega }}}_p\), \(\omega \), \({{\dot{\omega }}}_p\), and \(\dot{M}\) are noncomensurable. Equation (A.7) implies that \(c_3\) is again given by equation (A.13) and the condition of no spin-orbit resonance implies \(c_2=0\). As before, to compute \(c_1\) we look for a term in \(J_{g33}\) that is constant in time. Using that \(X^{-3,\pm 2}_0=0\) and the no spin-orbit hypothesis we obtain that

$$\begin{aligned} \frac{3\mathcal{G}m_p}{2r^3}\bigg (\cos \big (2f_p\big )\big (u_3^2-v_3^2\big )+ \sin \big (2f_p\big )2 u_3v_3\bigg ) \end{aligned}$$

does not contain any term that is constant. Using that \(-\dot{\psi }_g+{{\dot{\varOmega }}}_p\ne 0\), an analysis of the remaining term, \(\frac{3\mathcal{G}m_p}{2r^3}\big (u_3^2+v_3^2\big )\), shows that the constant term of \(J_{g33}\) is

$$\begin{aligned} \frac{3\mathcal{G}m_p}{2 a_p^3} X^{-3,0}_0 \frac{1}{8} \Big (-\cos (2 \theta _g)-(3 \cos (2 \theta _g)+1) \cos (2 \iota _p)+5\Big )= \omega ^2\left( c_3-\frac{2}{3}c_1\right) . \end{aligned}$$

So, using Eq. (A.13) for \(c_3\) we obtain that for no spin-orbit resonance:

$$\begin{aligned} \begin{aligned} c_1&=\frac{3\mathcal{G}m_p}{\omega ^2 a_p^3} X^{-3,0}_0 \frac{1}{32} \Big (3 \cos (2 \theta _g)+1\Big ) \Big (3 \cos (2 \iota _p)+1\Big )\\ c_2&=0 \end{aligned}. \end{aligned}$$
(A.16)

If there are several point masses, \(\beta =1,2,\ldots \), orbiting the extended body, each one in a precessing Keplerian orbit, with constant inclination, \(\iota _{\beta }=\)constant, and with no spin-orbit resonance; as in the case where the Earth is the extended body and the Moon and the Sun represent the point masses; then \(c_2=0\) and (the index p was omitted in several constants)

$$\begin{aligned} c_1\!=\!\bigg (\sum _\beta s_\beta \bigg )\frac{1+3 \cos (2 \theta _g)}{4}\quad \text {where}\quad s_\beta \!=\! \frac{3\mathcal{G}m_\beta }{\omega ^2 a_\beta ^3} (1-e_\beta ^2)^{-3/2}\, \frac{1+3 \cos (2 \iota _\beta )}{8}\,. \qquad \end{aligned}$$
(A.17)

In this case the average rate of precession \({{\dot{\psi }}}_g\) of the spin axis of the extended body is Williams (1994):

$$\begin{aligned} \frac{{{\dot{\psi }}}_g}{\omega }=-\big (\sum _\beta s_\beta \big ) \frac{\overline{I}_3-\overline{I}_e}{\overline{I}_3} \cos \theta _g\approx -\big (\sum _\beta s_\beta \big ) \overline{\alpha }_e\, \cos \theta _g\,; \end{aligned}$$
(A.18)

and the motion of the slow frame is determined by \(\varvec{R}_s\) as given in Eqs. (A.3) and (A.4) with \(\theta _g=\)constant.

A simple model: Considerations about the guiding frame

In this appendix, we show by means of an example how the inertial forces that appear due to the non-inertial character of the guiding frame are mostly cancelled out by external torques. This example will be used in Appendix E.

The problem is to describe the axial precession of an extended rigid body of mass m and moment of inertia \(\varvec{I}\) under the gravitational force of a point of mass \(m_p\) that moves in a circular orbit of radius \(a_p\) and with angular frequency n. We assume that the body is axisymmetric with \(\overline{I}_1=\overline{I}_2=\overline{I}_e<\overline{I}_3\). We define a body frame \({\mathrm {K}}\) such that the \(\varvec{e}_3\)-axis is the \(\overline{I}_3\)-principal axis and an inertial frame \(\kappa \) such that the orbit of the point mass has components \((a_p\cos (nt),a_p\sin (nt),0)\).

Let \(\omega >0\) be the initial spin angular speed of the body that is defined as the projection of the initial angular velocity vector \(\varvec{\omega }\) on the \(I_3\)-principal axis. We assume that the angular velocity is almost aligned with the \(I_3\)-principal axis and that inertial forces prevail over the gravitational torque, namely

$$\begin{aligned} \omega /\Vert \varvec{\omega }\Vert \approx 1\quad \text {and}\quad \frac{3}{2} \frac{G m_p}{\omega ^2 a^3}=s \ll 1. \end{aligned}$$
(B.1)

In this case the torque-free inertial motion dominates and the body rotates almost steadily about the spin-axis. Note that s is the quantity that we denoted as \(s_\beta \) in Eq. (A.17).

The tidal-force operator \(\varvec{J}= \frac{3\mathcal{G}m_p}{r^5}{\mathbf {r}}\otimes {\mathbf {r}}\) can be decomposed into two terms

$$\begin{aligned} \varvec{J}= \underbrace{s \, \omega ^2\left( \begin{array}{ccc} 1 &{} 0 &{} 0 \\ 0 &{} 1 &{} 0 \\ 0 &{} 0 &{} 0 \\ \end{array} \right) }_{= \varvec{J}_0} +\underbrace{s\,\omega ^2 \left( \begin{array}{ccc} \cos (2 n t) &{} \sin (2 n t) &{} 0 \\ \sin (2 n t) &{} -\cos (2 n t) &{} 0 \\ 0 &{} 0 &{} 0 \\ \end{array} \right) }_{=\varvec{J}_1}\,. \end{aligned}$$
(B.2)

The axial precession is determined by the constant part \(\varvec{J}_0\), so in the following we consider the problem of the motion of the extended body only under tidal-force operator \(\varvec{J}_0\), namely

$$\begin{aligned} \begin{aligned}&\dot{\varvec{{\widehat{\pi }}}} =[\varvec{I}\, ,{\mathbf {J}}_0] \quad \text {where}\quad \varvec{R}:{\mathrm {K}}\rightarrow \kappa \\&\varvec{{\widehat{\pi }}}=\mathrm{\!\ Tr\!\ } \big (\varvec{I}\big ) \varvec{\widehat{\omega }}- \varvec{I} \varvec{{\widehat{\omega }}}- \varvec{\widehat{\omega }}\varvec{I}\,,\quad \varvec{{\widehat{\omega }}}= \dot{\varvec{R}}\varvec{R}^{-1}\,.\\ \end{aligned} \end{aligned}$$
(B.3)

A computation shows that \( \varvec{R}=\varvec{R}_3({{\dot{\psi }}} \, t)\varvec{R}_1(\theta ) \varvec{R}_3(-{{\dot{\psi }}} \, t\cos \theta )\varvec{R}_3(\omega \, t)\) is a solution to this equation, with \({{\dot{\theta }}}=\ddot{\psi }=0\), if

$$\begin{aligned} \left( s\frac{I_3-I_e}{I_3} -\frac{I_e}{I_3} \frac{{{\dot{\psi }}}^2}{\omega ^2}\right) \cos \theta +\frac{\dot{\psi }}{\omega }=0. \end{aligned}$$
(B.4)

This equation has two solutions: one with \({{\dot{\psi }}}/\omega >0\) (prograde) and another with \({{\dot{\psi }}}/\omega <0\) (retrograde). Up to leading order in the small parameter s the prograde solution

$$\begin{aligned} \frac{{{\dot{\psi }}}}{\omega }=\frac{I_3}{I_e} \sec \theta \end{aligned}$$
(B.5)

is the fast torque-free precession and the retrograde solution

$$\begin{aligned} \frac{{{\dot{\psi }}}}{\omega }= -s\frac{I_3-I_e}{I_3} \cos \theta \approx -s\overline{\alpha }_e \cos \theta \end{aligned}$$
(B.6)

is the slow precession that exists due to the gravitational interaction.

We will use the retrograde solution to define the guiding-frame:

$$\begin{aligned} \varvec{R}_g=\underbrace{\varvec{R}_3({{\dot{\psi }}}_g \, t)\varvec{R}_1(\theta _g) \varvec{R}_3(-{{\dot{\psi }}}_g \, t\cos \theta _g)}_{=\varvec{R}_s}\varvec{R}_3(\omega \, t)\,, \end{aligned}$$
(B.7)

where \({{\dot{\psi }}}_g<0\) is the solution to Eq. (B.6) with \(\theta =\theta _g\). Notice that if \(\theta _g=0\) then \(\varvec{R}_s=\)Identity.

The angular velocities of the guiding frame \(\varvec{\widehat{\omega }}_{g,g}=\varvec{R}_g^{-1} \dot{\varvec{R}}_g:{\mathrm {K}}_g\rightarrow {\mathrm {K}}_g\) and of the slow frame \(\varvec{{\widehat{\omega }}}_{s,s}=\varvec{R}_s^{-1} \dot{\varvec{R}}_s:{\mathrm {K}}_s\rightarrow {\mathrm {K}}_s\) are:

$$\begin{aligned} \varvec{\omega }_{g,g}=\omega \varvec{e}_3+\underbrace{\varvec{R}_3^{-1}(\omega t) \varvec{\omega }_{s,s}}_{{\varvec{\omega }}_{s,g}}\,,\quad \varvec{\omega }_{s,s}= {{\dot{\psi }}}_g \sin \theta _g\left( \begin{array}{c} - \sin (t{{\dot{\psi }}}_g \cos \theta _g) \\ \ \ \cos (t{{\dot{\psi }}}_g \cos \theta _g)\\ 0 \end{array} \right) . \end{aligned}$$
(B.8)

Since for the problem considered in this section the guiding motion is a solution to the equations of motion, the linearised equations about the guiding motion are just the ordinary linearised equations about a solution. These equations can be obtained directly from Eq. 5.16), for the motion of the mantle, with the following simplifications: \(\varvec{\alpha }_m\rightarrow \varvec{a}\) (\({\mathbb {I}} +\varvec{{{\widehat{a}}}}:{\mathrm {K}}\rightarrow {\mathrm {K}}_g\)), \(\varvec{I}_m=\varvec{I}\) and \(\varvec{I}_c=0\) (there is no fluid core), \(\varvec{\delta }\varvec{B}_T=0\) (the body is rigid), and \(c_2=0\Longrightarrow \xi _1=\xi _2=1=c_1\) (there is no spin-orbit resonance) with

$$\begin{aligned} c_1= s\frac{1+3 \cos (2 \theta _g)}{4} \quad \Big (\text {consequence of Eq.} (A.17)\Big )\,. \end{aligned}$$
(B.9)

There are two forcing terms that appear in the right-hand side of Eq. (5.16): the “true-torque” that comes from \(\varvec{J}_0\) and the “fictitious-torque” that comes from the non-inertial nature of the guiding frame. The true-torque term is given by:

$$\begin{aligned}{}[\overline{\varvec{I}},\varvec{R}_g^{-1}\varvec{J}_0\varvec{R}_g]^\vee =[\overline{\varvec{I}},{\varvec{\delta }}\varvec{J}_{0,g}]^\vee = s\,\omega ^2\,(\overline{I}_3-\overline{I}_e)\cos \theta _g\sin \theta _g \begin{bmatrix} -\cos \phi _g\\ \ \ \, \sin \phi _g\\ 0 \end{bmatrix}\,, \end{aligned}$$
(B.10)

where we used that: \([\overline{\varvec{I}},\overline{\varvec{J}}]=0\), the check map \(^\vee \) is the inverse of the hat map, and \(\phi _g=t(\omega -{{\dot{\psi }}}_g\cos \theta _g)\) (see Eqs. (A.3) and (A.4)). The fictitious torque is

$$\begin{aligned} \begin{bmatrix} \overline{I}_{e}\dot{{\omega }}_{s,g1} +\omega \big (\overline{I}_3-\overline{I}_e\big ){\omega }_{s,g2} \\ \overline{I}_{e}\dot{{\omega }}_{s,g2} -\omega \big (\overline{I}_3-\overline{I}_e\big ){\omega }_{s,g1} \\ 0 \end{bmatrix}=\sin \theta _g\Big (\overline{I}_e(\omega -{{\dot{\psi }}}_g\cos \theta _g){{\dot{\psi }}}_g+\omega (\overline{I}_3-\overline{I}_e) {{\dot{\psi }}}_g\Big ) \begin{bmatrix} -\cos \phi _g\\ \ \ \, \sin \phi _g\\ 0 \end{bmatrix}\,,\nonumber \\ \end{aligned}$$
(B.11)

where we used equation \({\varvec{\omega }}_{s,g}\) as given in Eq. (B.8). If we use that \({{\dot{\psi }}}_g\) is a solution to Eq. (B.4), then we obtain that the fictitious torque is equal to minus the true torque. So, the true torque cancels out the fictitious torque.

In general the guiding motion is not a particular solution of the dynamical equations and so, we cannot expect the full cancellation of the fictitious torque. If the guiding motion is a good approximation for the real motion, then the residue after the partial cancellation of the fictitious torque must be of the order of the small terms in the equation.

In conclusion, for Eq. (B.3) the linearised equation about the guiding motion is

$$\begin{aligned} \begin{aligned}&\overline{I}_{e}\ddot{a}_{1}-\omega (2\overline{I}_{e}-\overline{I}_{3}){\dot{a}}_{2} +\omega ^2\xi _1(\overline{I}_3-\overline{I}_e)a_{1}=0\\&\overline{I}_{e}\ddot{a}_{2} +\omega (2\overline{I}_e-\overline{I}_{3}){\dot{a}}_{1} +\omega ^2\xi _2(\overline{I}_3-\overline{I}_e)a_{2}=0\\&\overline{I}_{3}\ddot{a}_{3}=0 \,. \end{aligned} \end{aligned}$$
(B.12)

The classification of the roots of Eq. (6.21) into NDFW and FLL eigenvalues

In this Appendix, we classify the two roots of the equation (6.21), i.e. \( x^2-x(f_0+1) (y+z )+(f_0+1)z\,y=0\).

If the discriminant \(\varDelta \) of Eq. (6.21) is different from zero, then the equation has two solutions. If the core is negligible \(f_0=\frac{\mathrm{I}_{\circ c}}{\mathrm{I}_{\circ m}}=0\), then one of the root is \(x_{dw}=y\) and is related to the Nearly Diurnal Free Wobble (NDFW) and the other \(x_{\ell a}=z\) is related to the Free Libration in Latitude (FLL). In order to classify a given solution one must deform \(f_0>0\) from its current value to \(f_0=0\) keeping (yx) constant. If the function \(f_0\rightarrow \varDelta \) remains different from zero during the deformation, then the root will move in the complex plane as a smooth function of \(f_0\) and it will eventually become either y, and the root will be classified as \(x_{nd}\), or z, and the root will be classified as \(x_{\ell a}\).

The discriminant of Eq. (6.21) becomes simpler if we use the variables

$$\begin{aligned}&p:=\frac{1}{2}(y+ z)\,,\quad q:=\frac{1}{2}( y-z)\Longrightarrow y=p+q\,,\quad z=p-q\,,\nonumber \\&\text {and}\quad \varDelta =4 (f_0+1) \left( f_0 \,p^2+q^2\right) \,. \end{aligned}$$
(C.1)

The two solutions to Eq. (6.21) are \( (f_0+1)p\pm \sqrt{1+f_0}\sqrt{f_0 \,p^2+q^2}\), where denotes the principal value of the square root, which is not continuous on the non-positive real axis and satisfies Re \(\sqrt{z}\ge 0\) with \(\sqrt{-1}=i\). The lack of continuity of the principal value of the square root leads to some difficulties in the classification of the roots because \(f_0\rightarrow \varDelta \) can cross the non-positive real axis during the variation of \(f_0\). The eigenvalues are given by the following algorithm obtained essentially from Fig. 15.

Notation: (1) Given (yz) we define \(q:=q_r+iq_i=(y_r-z_r)/2+i(y_i-z_i)/2\) and \(p:=p_r+ip_i=(y_r+z_r)/2+i(y_i+z_i)/2\), as in Eq. (C.1), and \(p^2:=p_{2r}+i p_{2i}=(p^2_r-p^2_i)+i2 p_rp_i \) and \(=q^2:=q_{2r}+i q_{2i}=(q^2_r-q^2_i)+i2 q_rq_i \). (2) We define the set of inequalities

$$\begin{aligned} V=\bigg \{q_{2r}<0\,,q_{2i}<0\,, \frac{q_{2i}}{q_{2r}}<\frac{p_{2i}}{p_{2r}}\,, \ \text {and} \ f_0>-\frac{q_{2r}}{p_{2r}}\bigg \}\,. \end{aligned}$$
(C.2)

All the inequalities in V hold if \(p^2\), \(q^2\) and \(f_0\) are arranged as illustrated in Fig. 15.

Assumptions: (1) To simplify the analysis we assume \(p_i/p_r<1\), which implies that \(p^2\) is in the positive quadrant of the complex plane, \(p_{2r}>0\) and \(p_{2i}>0\) (the analysis could also be made without this hypothesis). (2) We also assume that \(|y|\ne |z|\) in order to avoid \(\varDelta =0\) during a variation of \(f_0\).

  • If \( y_r>z_r\), then either all inequalities in V are true and Eq. (C.4) holds or at least one inequality in V is not true and Eq. (C.3) holds.

  • If \( y_r<z_r\), then either all inequalities in V are true and Eq. (C.3) holds or at least one inequality in V is not true and Eq. (C.4) holds.

$$\begin{aligned}&x_{dw}=(1+f_0)p+\frac{\sqrt{\varDelta }}{2}\,,\quad x_{\ell a}=(1+f_0)p-\frac{\sqrt{\varDelta }}{2}\,; \end{aligned}$$
(C.3)
$$\begin{aligned}&x_{dw}=(1+f_0)p-\frac{\sqrt{\varDelta }}{2}\,,\quad x_{\ell a}=(1+f_0)p+\frac{\sqrt{\varDelta }}{2}\,; \end{aligned}$$
(C.4)

where \(\varDelta =4 (f_0+1) \left( f_0 \,p^2+q^2\right) \).

Fig. 15
figure 15

Diagram used in the classification of the eigenvalues of NDFW and FLL. Note that \(y_r>z_r\) ( \(y_r<z_r\)) implies: \(q_r>0\) (\(q_r<0\)), \(\sqrt{q^2}=q\) (\(\sqrt{q^2}=-q\)). The inequality \(\{q_{2r}<0\,,q_{2i}<0\,, \frac{q_{2i}}{q_{2r}}<\frac{p_{2i}}{p_{2r}}\}\) is represented by the shaded areas in the figure and \(f_0>-\frac{q_{2r}}{p_{2r}}\) implies \(\mathrm{Re}\,(f_0p^2+q^2)>0\)

The two eigenfrequencies coincide, \(x_{dw}=x_{\ell a}\), when \(\varDelta =0\), what happens for \((y,z,f_0)\) in the set

$$\begin{aligned} \bigg \{ |z|=|y|\,, f_0=\tan ^2(\psi /2)\bigg \}\,, \end{aligned}$$
(C.5)

where \(\psi \) is the angle between y and z, \(y=\mathrm{e} ^{i\psi }z\).

The NDFW and the FLL eigenfrequencies (this property does not extend to the eigenvectors) are dual to each other in the sense that if two bodies have the same \(f_0\) but one has \((y,z)=(a_1,a_2)\) while the second \((y,z)=(a_2,a_1)\), then the value of \(x_{dw}\) (\(x_{\ell a}\)) of the first is equal to the value of \(x_{\ell a}\) (\(x_{dw}\)) of the second. This is a consequence of the symmetry of Eq. (6.21) with respect to the permutation of y and z.

In the limit as the mantle becomes negligible, i.e. \(f_0=\frac{\mathrm{I}_{\circ c}}{\mathrm{I}_{\circ m}}\rightarrow \infty \), we find two solutions to Eq. (6.21)

$$\begin{aligned} x_c=\frac{yz}{y+z} \end{aligned}$$
(C.6)

and \((1+f_0)(y+z)\). These solutions represent the limit of \(x_{dw}(f_0)\) and \(x_{\ell a}(f_0)\) as \(f_0\rightarrow \infty \). In order to decide whether \(x_{dw}\) or \(x_{\ell a}\) is asymptotic to \(x_c\), we solve Eq. (6.21) with \(y=0\), the solutions being \(x_{dw}(f_0)=0=x_c\) and \(x_{\ell a}(f_0)=(f_0+1)z\), and with \(z=0\), the solutions being \(x_{\ell a}(f_0)=0=x_c\) and \(x_{dw}(f_0)=(f_0+1)y\). Since \(\varDelta \ne 0\), by continuity the same result hold for |y| small in the first case and |z| small in the second. This fact, the duality of \(x_{dw}\) and \(x_{\ell a}\) discussed in the previous paragraph, and the invariance of \(x_c\) and \((1+f_0)(y+z)\) with respect to the permutation \(y\leftrightarrow z\), lead us to:

$$\begin{aligned} \begin{aligned}&|y|<|z| \Longrightarrow x_{dw}(f_0)\rightarrow x_c\ \ \text {and} \ \ x_{\ell a}(f_0)\rightarrow (f_0+1)z\ \ \text {as}\ \ f_0\rightarrow \infty \\&|z|<|y| \Longrightarrow x_{dw}(f_0)\rightarrow (f_0+1)z \ \ \text {and} \ \ x_{\ell a}(f_0)\rightarrow x_c \ \ \text {as}\ \ f_0\rightarrow \infty \,.\end{aligned} \end{aligned}$$
(C.7)

The Poincaré-Hough flow

Following Poincaré (see Lamb (1932) paragraphs 146 and 384), let \(\varvec{x}_0\in \kappa \) be the position at time \(t=0\) of a fluid particle inside a triaxial ellipsoid with semi-axes \(A_1>A_2>A_3\). The mean radius of the ellipsoid is \(R_c=(A_1A_2A_3)^{1/3}\). For any time the center of the ellipsoid coincides with the origin of \(\kappa \) and at \(t=0\) the axes of the ellipsoid are aligned with the axes of \(\kappa \). At \(t=0\) the operator \(\varvec{A}:=\)Diagonal\(\{A_1,A_2,A_3\}\) maps the ball of radius unity onto the ellipsoid. The ellipsoid is fixed in the frame of the mantle \({\mathrm {K}}_m\) and rotates inside the inertial frame \(\kappa \) according to \(\varvec{R}_m(t):{\mathrm {K}}_m\rightarrow \kappa \), with \(\varvec{R}_m(0)={\mathbb {I}} \). The motion of the fluid particle initially at \(\varvec{x}_0\) is assumed to be:

$$\begin{aligned} \varvec{x}(t)=\varvec{R}_m(t)\varvec{A}\varvec{R}_f(t)\varvec{A}^{-1}\varvec{x}_0\,, \end{aligned}$$
(D.1)

where \(t\rightarrow \varvec{R}_m(t)\) is known. The unknown rotation matrix \(\varvec{R}_f(t)\) must be determined from the fluid dynamic equations.

The velocity field associated with the motion in Eq. (D.1) is

$$\begin{aligned} \varvec{v}(\varvec{x}, t)=\dot{\varvec{x}}=\underbrace{\big (\varvec{{\widehat{\omega }}}_m+ \varvec{R}_m\varvec{A}\varvec{{\widehat{\omega }}}_f\varvec{A}^{-1} \varvec{R}_m^{-1}\big )}_{:=\varvec{P}}\varvec{x}=\varvec{P} \varvec{x}. \end{aligned}$$
(D.2)

It is convenient to split the operator \(\varvec{A}\varvec{\widehat{\omega }}_f\varvec{A}^{-1}\) into symmetric and anti-symmetric parts

$$\begin{aligned} \varvec{A}\varvec{{\widehat{\omega }}}_f\varvec{A}^{-1}=\underbrace{\frac{1}{2}\Big ( \varvec{A}\varvec{\widehat{\omega }}_f\varvec{A}^{-1}+ \big (\varvec{A}\varvec{{\widehat{\omega }}}_f\varvec{A}^{-1}\big )^T\Big )}_{:=\varvec{S}_A}+ \underbrace{\frac{1}{2}\Big ( \varvec{A}\varvec{{\widehat{\omega }}}_f\varvec{A}^{-1}- \big (\varvec{A}\varvec{{\widehat{\omega }}}_f\varvec{A}^{-1}\big )^T\Big )}_{:=\varvec{{\widehat{\tau }}}}\,, \end{aligned}$$
(D.3)

such that

$$\begin{aligned} \varvec{P}= \varvec{{\widehat{\omega }}}_m+ \varvec{R}_m\varvec{{\widehat{\tau }}}\varvec{R}_m^{-1}+ \varvec{R}_m\varvec{S}_A \varvec{R}_m^{-1}\,. \end{aligned}$$
(D.4)

The expressions for \(\varvec{S}_A\) and \(\varvec{\tau }\) are

$$\begin{aligned} \varvec{S}_A=\frac{1}{2}\left( \begin{array}{ccc} 0 &{} \left( \frac{A_2}{A_1}-\frac{A_1}{A_2}\right) \omega _{f3} &{} \left( \frac{A_1}{A_3}-\frac{A_3}{A_1}\right) \omega _{f2} \\ \left( \frac{A_2}{A_1}-\frac{A_1}{A_2}\right) \omega _{f3} &{} 0 &{} \left( \frac{A_3}{A_2}-\frac{A_2}{A_3}\right) \omega _{f1} \\ \left( \frac{A_1}{A_3}-\frac{A_3}{A_1}\right) \omega _{f2} &{} \left( \frac{A_3}{A_2}-\frac{A_2}{A_3}\right) \omega _{f1} &{} 0 \\ \end{array} \right) \end{aligned}$$
(D.5)

and

$$\begin{aligned} \varvec{{\widehat{\tau }}}=\frac{1}{2}\left( \begin{array}{ccc} 0 &{} -\left( \frac{A_2}{A_1}+\frac{A_1}{A_2}\right) \omega _{f3} &{} \left( \frac{A_1}{A_3}+\frac{A_3}{A_1}\right) \omega _{f2} \\ \left( \frac{A_2}{A_1}+\frac{A_1}{A_2}\right) \omega _{f3} &{} 0 &{}- \left( \frac{A_3}{A_2}+\frac{A_2}{A_3}\right) \omega _{f1} \\ - \left( \frac{A_1}{A_3}+\frac{A_3}{A_1}\right) \omega _{f2} &{} \left( \frac{A_3}{A_2}+\frac{A_2}{A_3}\right) \omega _{f1} &{} 0 \\ \end{array} \right) . \end{aligned}$$
(D.6)

Using that \(\varvec{v}=\varvec{P} x= \varvec{ \omega }_m\times \varvec{x} + (\varvec{R}_m\varvec{\tau })\times \varvec{x} + \varvec{R}_m\varvec{S}_A \varvec{R}_m^{-1}\varvec{x}\) we obtain the vorticity field

$$\begin{aligned} \varvec{w}:=\varvec{curl}\, \varvec{v}= 2 \big (\varvec{ \omega }_m+\varvec{R}_m\varvec{\tau }\big )\,, \end{aligned}$$
(D.7)

which implies

$$\begin{aligned} \varvec{P}= \frac{\varvec{{{\widehat{w}}}}}{2}+ \varvec{R}_m\varvec{S}_A \varvec{R}_m^{-1}\quad \text {and}\quad \varvec{v}= \frac{\varvec{w}}{2}\times \varvec{x} + \varvec{R}_m\varvec{S}_A \varvec{R}_m^{-1}\varvec{x}\,. \end{aligned}$$
(D.8)

The equations for the motion of an inviscid, volume preserving, isentropic fluid are (Euler):

$$\begin{aligned} \begin{array}{lll} &{} \partial _t \varvec{v}+\varvec{v}\cdot \nabla \varvec{v}=-\nabla h-\nabla \varPhi \quad &{} (\text {dynamical equation})\,,\\ &{} \varvec{div}\, \varvec{v}=0\quad &{}(\text {incompressibility})\,,\\ &{}\partial _t \rho + \varvec{div}\,(\rho \varvec{v})=0\quad &{}(\text {conservation of mass})\,,\end{array} \end{aligned}$$
(D.9)

where: \(\rho \) is the density, h is the enthalpy (\(dh=\frac{dp}{\rho }\) where p is the pressure), and \(\varPhi \) is the external gravitational potential.

The divergence of \(\varvec{v}=\varvec{P} x\) is zero because \({{\,\mathrm{Tr}\,}}(\varvec{P})=0\). In order to fulfil the equation of continuity we impose that,

$$\begin{aligned} \rho (\varvec{x},t)=\rho _\circ (r)\,,\ \ \text {where}\ \ r= \Vert \varvec{A}^{-1} \varvec{R}_m^{-1}(t) \varvec{x}\Vert \,, 0\le r\le R_c. \end{aligned}$$
(D.10)

If the fluid cavity would be slowly deformed to a round shape, then the density of the fluid inside the round core would be the radial function \(\rho _\circ (r)\). Using that \(\varvec{div}\, \varvec{v}=0\) the equation for conservation of mass can be written as \(\partial _t \rho + \varvec{v}\cdot \nabla \rho =\frac{d}{dt}\rho (\varvec{x}(t),t)=0\). Equation (D.1) implies \(\rho (\varvec{x}(t),t)=\rho _\circ ( \Vert \varvec{A}^{-1} \varvec{R}_m^{-1}(t) \varvec{x}(t)\Vert )= \rho _\circ ( \Vert \varvec{A}^{-1} \varvec{x}_0\Vert )\), so \(\frac{d}{dt}\rho (\varvec{x}(t),t)=0\) and the equation for conservation of mass is verified.

The last equation in (D.9) to be verified is the dynamical one. If we take the curl of this equation, then we obtain \(\partial _t \varvec{w}+\varvec{v}\cdot \nabla \varvec{w}-\varvec{w}\cdot \nabla \varvec{v}=0\), which is the dynamical equation for the vorticity. If the equation for the vorticity is verified, then the enthalpy can be determined integrating the dynamical equation for \(\varvec{v}\). Since \(\varvec{w}\) does not depend on \(\varvec{x}\), \(\varvec{v}\cdot \nabla \varvec{w}=0\) and, since \(\varvec{v}=\varvec{P} x\) depends linearly on \(\varvec{x}\), \(\varvec{w}\cdot \nabla \varvec{v}=\varvec{P}\varvec{w}\), so the equation for the vorticity reduces to

$$\begin{aligned} \dot{\varvec{w}}=\varvec{P} \varvec{w}\,. \end{aligned}$$
(D.11)

This equation yields a simple ordinary differential equation for the unknown function \(\varvec{\omega }_f\) from which we determine the Poincaré-Hough flow.

Equation (D.8) implies that the Tisserand angular velocity \({\varvec{\omega }}_c\) associated with the Poincaré-Hough flow satisfies

$$\begin{aligned} {\varvec{\pi }}_c= \varvec{I}_c {\varvec{\omega }}_c=\int _{\mathcal B(t)}\rho (\varvec{x}\times \varvec{v})d x^3= \varvec{I}_c \frac{\varvec{w}}{2} + \int _{{\mathcal {B}}(t)}\rho \, \varvec{x}\times \big (\varvec{R}_m\varvec{S}_A \varvec{R}_m^{-1}\varvec{x})dx^3,\quad \end{aligned}$$
(D.12)

where \({\mathcal {B}}(t)\) is the set of points inside the cavity at time t. Since the density function is carried by the flow, \(\varvec{I}_c(t)=\varvec{R}_m(t) \varvec{I}_c(0)\varvec{R}_m^{-1}(t)\).

In order to compute the moment of inertia of the fluid at \(t=0\) we first compute the components of the density tensor \(\varvec{M}(0)\)

$$\begin{aligned} M_{ij}(0)=\int _{\Vert \varvec{A}^{-1}\varvec{x}\Vert \le 1}x_ix_j\rho _\circ (\Vert \varvec{A}^{-1}\varvec{x}\Vert )dx^3= \frac{ A_iA_j}{R_c^2}\int _{\Vert \varvec{y}\Vert \le R_c}y_iy_j\rho _\circ (\Vert \varvec{y}\Vert )dx^3\,, \end{aligned}$$
(D.13)

where we did the change of variables \(x_i=(A_i/R_c) y_i\), \(i=1,2,3\). Parity arguments imply that \(M_{ij}(0)=0\) if \(i\ne j\). Now, consider a round cavity of radius \(R_c\) filled with a fluid with spherical density \(\rho _\circ (r)\). The moment of inertia about any axis, say \(\varvec{e}_3\), of the spherical mass of fluid is

$$\begin{aligned} \mathrm{I}_{\circ c}=\int _{\Vert \varvec{y}\Vert \le R_c}(y_1^2+y_2^2)\rho _\circ (\Vert \varvec{y}\Vert )dx^3 \end{aligned}$$
(D.14)

So, by symmetry \(\int _{\Vert \varvec{y}\Vert \le R_c}y_i^2\rho _\circ (\Vert \varvec{y}\Vert )dx^3= \mathrm{I}_{\circ c}/2\) for any \(i=1,2,3\) and

$$\begin{aligned} M_{ii}(0)= \frac{ A_i^2}{R_c^2}\frac{\mathrm{I}_{\circ c}}{2}\,, \quad i=1,2,3\,. \end{aligned}$$
(D.15)

Using that \(\varvec{I}_c={{\,\mathrm{Tr}\,}}\big (\varvec{M}\big ){\mathbb {I}} -\varvec{M}\) we obtain

$$\begin{aligned} \varvec{I}_c(0)=\frac{\mathrm{I}_{\circ c}}{2 R_c^2}\left( \begin{array}{ccc} A_2^2+A_3^2&{}0&{}0\\ 0&{} A_1^2+A_3^2&{}0\\ 0&{}0&{} A_1^2+A_2^2\end{array}\right) . \end{aligned}$$
(D.16)

Note that \({{\,\mathrm{Tr}\,}}\varvec{I}_c(0)= \mathrm{I}_{\circ c}\frac{A_1^2+A_2^2+A_3^2}{R_c^2}\), so, \(\mathrm{I}_{\circ c}\) may not have the usual meaning \(\frac{1}{3}{{\,\mathrm{Tr}\,}}\varvec{I}_c(0)\).

In order to compare the results obtained from the Poincaré-Hough flow with ours we have to assume that the fluid cavity is slightly aspherical, namely

$$\begin{aligned} A_1=R_c(1+\epsilon _1)\quad A_2=R_c(1+\epsilon _2)\quad A_3=R_c(1+\epsilon _3)\,, \end{aligned}$$
(D.17)

where \(|\epsilon _1|,|\epsilon _2|,\) and \(|\epsilon _3|\) are small. Since \(A_1A_2A_3=R_c^3\), \(\epsilon _1+\epsilon _2+\epsilon _3=0\) and equation (D.16) implies

$$\begin{aligned} \varvec{I}_c(0)=\mathrm{I}_{\circ c} {\mathbb {I}} + \mathrm{I}_{\circ c}\left( \begin{array}{ccc} \epsilon _2+\epsilon _3&{}0&{}0\\ 0&{} \epsilon _1+\epsilon _3&{}0\\ 0&{}0&{} \epsilon _1+\epsilon _2\end{array}\right) +{\mathcal {O}}(|\epsilon |^2)\,, \end{aligned}$$
(D.18)

so, \(\frac{1}{3}{{\,\mathrm{Tr}\,}}\varvec{I}_c(0)=\mathrm{I}_{\circ c}\) holds up to second order in the ellipticityFootnote 32. Equation (D.5) implies

$$\begin{aligned} \varvec{S}_A=\left( \begin{array}{ccc} 0 &{} \left( \epsilon _2-\epsilon _1\right) \omega _{f3} &{} \left( \epsilon _1-\epsilon _3\right) \omega _{f2} \\ \left( \epsilon _2-\epsilon _1\right) \omega _{f3} &{} 0 &{} \left( \epsilon _3-\epsilon _2\right) \omega _{f1} \\ \left( \epsilon _1-\epsilon _3\right) \omega _{f2} &{} \left( \epsilon _3-\epsilon _2\right) \omega _{f1} &{} 0 \\ \end{array} \right) +{\mathcal {O}}(|\epsilon |^3)\,, \end{aligned}$$
(D.19)

and equation (D.6) implies \( \varvec{ \tau }={\varvec{\omega }}_f+{\mathcal {O}}(|\epsilon |^2)\).

A computation shows that the last term in the right-hand side of equation (D.12) is of the order \({\mathcal {O}}(|\epsilon |^2)\) and, therefore

$$\begin{aligned} {\varvec{\pi }}_c\!= \!\varvec{I}_c {\varvec{\omega }}_c= \varvec{I}_c \frac{\varvec{w}}{2} \!+\!{\mathcal {O}}(|\epsilon |^2) \Longrightarrow {\varvec{\omega }}_c\!=\! \frac{\varvec{w}}{2} +{\mathcal {O}}(|\epsilon |^2)= \varvec{\omega }_m+\varvec{R}_m \varvec{\omega }_f +{\mathcal {O}}(|\epsilon |^2) \end{aligned}$$
(D.20)

The fact that \(\varvec{\omega }_c\approx \varvec{w}/2\) for small ellipticities has been noted in Henrard (2008).

For small ellipticities we can write \(\varvec{\omega }_c= \varvec{w}/2\) and equation (D.11) implies

$$\begin{aligned} \dot{\varvec{\omega }}_c=\varvec{P} \varvec{\omega }_c\,. \end{aligned}$$
(D.21)

Therefore

$$\begin{aligned} \dot{{\varvec{\pi }}}_c=\dot{\varvec{I}}_c\varvec{\omega }_c+{\varvec{I}}_c\dot{\varvec{\omega }}_c= {\varvec{\omega }}_m\times \varvec{I}_c \varvec{\omega }_c+ \varvec{I}_c(-\varvec{{\widehat{\omega }}}_m+\varvec{P}) \varvec{\omega }_c. \end{aligned}$$
(D.22)

Equation (D.2) implies that the last term in the right-hand side of this equation can be written as

$$\begin{aligned} \varvec{I}_c(-\varvec{{\widehat{\omega }}}_m+\varvec{P}) \varvec{\omega }_c=\varvec{R}_m \varvec{I}_c(0)\varvec{A}\varvec{\widehat{\omega }}_f\varvec{A}^{-1} \varvec{R}_m^{-1} \varvec{\omega }_c\,. \end{aligned}$$

A computation using equation (D.18) shows that \(\varvec{I}_c(0)\varvec{A}\varvec{{\widehat{\omega }}}_f\varvec{A}^{-1}= \varvec{{\widehat{\omega }}}_f \varvec{I}_c(0)+ {\mathcal {O}}(|\epsilon |^2)\), so up to second order in the ellipticities

$$\begin{aligned} \dot{{\varvec{\pi }}}_c= {\varvec{\omega }}_m\times \varvec{I}_c \varvec{\omega }_c+ \varvec{R}_m \varvec{{\widehat{\omega }}}_f \varvec{I}_c(0) \varvec{R}_m^{-1} \varvec{\omega }_c= {\varvec{\omega }}_m\times \varvec{I}_c \varvec{\omega }_c+ \varvec{R}_m \varvec{{\widehat{\omega }}}_f \varvec{R}_m^{-1} \varvec{I}_c \varvec{\omega }_c \end{aligned}$$

and using equation (D.20) we obtain

$$\begin{aligned} \dot{{\varvec{\pi }}}_c= {\varvec{\omega }}_c\times \varvec{I}_c \varvec{\omega }_c, \end{aligned}$$
(D.23)

that is exactly the equation for the angular momentum of the core obtained in (5.1) with \(k_c=0\).

We remark that the hypotheses we used in this Appendix to obtain the general results in Eq. (5.1) from the Poincaré-Hough flow, namely: small asphericity, density stratification along concentric ellipsoidal shells and the volume preserving property of the fluid flow; are the same hypotheses we have assumed since Ragazzo and Ruiz (2017). In this Appendix, any rotational motion \(\varvec{R}_m(t)\) of the ellipsoid is allowed and there is no reason to assume that the principal axes of the mantle are aligned with those of the fluid cavity (a hypothesis commonly assumed). Finally, it is crucial for the results in this Appendix that the centre of mass of the mantle coincides with that of the fluid core for all time.

The inertial offset of the core rotation axis

In this appendix, we analyse the effect of the inertial torque studied in Sect. 7 upon the fluid core. Two different situations will be considered: one without spin-orbit resonance and another with spin-orbit resonance.

1.1 Precession under no spin-orbit resonance, the Tisserand angular velocity of the fluid core, and the vorticity of Robert and Stewartson

In this section, we assume that both the mantle and the core are axisymetric with \(\overline{\gamma }=0\), \(\overline{\alpha }=\overline{\beta }=\overline{\alpha }_e\), and \(I_{c1}=I_{c2}=I_{c3}(1-f_c)\), the mantle is rigid, and the precession rate is retrograde, \({{\dot{\psi }}}_g<0\). In this section, we further assume that \(\eta _c/(\omega f_c)\ll 1\). The goal is to study the angular velocity of the core produced exclusively by the inertial torque. The axisymmetry of the problem implies that in the precessional frame (equation (7.2)) both \({\varvec{\omega }}_{m,pr}\) and \({\varvec{\omega }}_{c,pr}\) are stationary, and according to equations (7.14) and (7.17) are given by

$$\begin{aligned} \begin{aligned} {\varvec{\omega }}_{m,pr}&= \omega \varvec{e}_3+{{\dot{\psi }}}_g\sin \theta _g\varvec{e}_2+ \omega \, {\varvec{\delta }}_m\\ {\varvec{\omega }}_{c,pr}&= \underbrace{ \omega \varvec{e}_3+{{\dot{\psi }}}_g\sin \theta _g\varvec{e}_2}_{\varvec{\omega }_{g,pr}}+ \omega \, {\varvec{\delta }}_c\end{aligned}. \end{aligned}$$
(E.1)

In order to write \( {\varvec{\delta }}_c\) we follow the same steps we did to obtain \({\varvec{\delta }}_m\) in equation (7.22). After doing this and using that the complex compliance is null (the mantle is rigid) and \(\eta _c/(\omega f_c)\ll 1\) (\(\Rightarrow y\approx f_c\) in equation (7.22)) we obtain

$$\begin{aligned} \begin{aligned} \varvec{\delta }_m&= \frac{\mathrm{I_{\circ c}}}{{\,\mathrm I}_\circ }\frac{ {{\dot{\psi }}}_g}{\omega f_c} \frac{2 \cos ^2 \theta _g}{ \sin \theta _g} \left( \frac{ \mathrm{I}_{\circ m}}{{\,\mathrm I}_\circ } \frac{\eta _c}{f_c\,\omega }\, \varvec{e}_1-\varvec{e}_2 \right) \\ \varvec{\delta }_c&=\frac{{{\dot{\psi }}}_g}{\omega f_c} \sin \theta _g \left( 1- 2 \frac{\mathrm{I}_{\circ c}}{{\,\mathrm I}_\circ }\cot ^2\theta _g\right) \left( \varvec{e}_2- \frac{ \mathrm{I}_{\circ m}}{{\,\mathrm I}_\circ } \frac{\eta _c}{f_c\,\omega }\varvec{e}_1\right) \\&=\varvec{\delta }_m+ \frac{ {{\dot{\psi }}}_g\,\sin \theta _g}{\omega f_c}\left( \varvec{e}_2- \frac{ \mathrm{I}_{\circ m}}{{\,\mathrm I}_\circ } \frac{\eta _c}{f_c\,\omega }\varvec{e}_1\right) \end{aligned} \end{aligned}$$
(E.2)

Roberts and Stewartson studied the motion of an incompressible fluid of constant density inside an ellipsoidal shell of revolution that precesses in the same way as the guiding motion used to obtain the angular velocities in equations (E.1) and (E.2). In Stewartson and Roberts (1963) and Roberts and Stewartson (1965) the authors solved the Navier-Stokes equations by perturbation methods (for a numerical study in the case of a spherical shell, see Tilgner and Busse (2001)) assuming the two hypotheses

$$\begin{aligned} \frac{f_c \,\omega }{|{{\dot{\psi }}}_g|}\gg 1\quad \text { and}\quad f_c \,\omega \frac{R_c^2}{\nu }=f_c\,\frac{\omega }{\eta _c}\gg 1\,, \end{aligned}$$
(E.3)

where: \(R_c\) is the core mean radius, \(\nu \) is kinematic viscosity of the fluid, and \(\eta _c= R_c^2/\nu \) is the viscosity coefficient defined in equation (3.22). These are the same hypotheses we assumed to obtain equation (E.2). In the following we show that the difference \({\varvec{\omega }}_{c,pr}-{\varvec{\omega }}_{m,pr}=\omega (\varvec{\delta }_c-\varvec{\delta }_m)\) is equal to the average vorticity of the flow inside the cavity, computed by Robert and Stewartson, minus the vorticity of the flow induced by a rigid rotation.

The velocity field of the fluid inside the cavity with respect to the precessional frame \({\mathrm {K}}_{pr}\) is Stewartson and Roberts (1963)

$$\begin{aligned} \varvec{u}={{\dot{\phi }}}_g\varvec{e}_3\times \varvec{x}+ 2{{\dot{\psi }}}_g\sin \theta _g \frac{a^2}{a^2-b^2}\left( x_3\varvec{e}_1-\frac{b^2}{a^2} x_1\varvec{e}_3\right) , \end{aligned}$$
(E.4)

where \(b<a\) are the semi-axis of the cavityFootnote 33. The vorticity associated with this velocity field is

$$\begin{aligned} \mathrm{curl}\,\varvec{u}=2{{\dot{\phi }}}\varvec{e}_3+ 2{{\dot{\psi }}}_g\sin \theta _g \frac{a^2+b^2}{a^2-b^2}\varvec{e}_2. \end{aligned}$$
(E.5)

If the fluid were moving as a rigid body attached to the mantle, then its vorticity would be \(\mathrm{curl} {{\dot{\phi }}}_g\varvec{e}_3\times \varvec{x}= 2{{\dot{\phi }}}\varvec{e}_3\). Recalling that vorticity is a measure of rotation of the vector field, as we have already mentioned in Appendix D,

$$\begin{aligned} 2{{\dot{\psi }}}_g\sin \theta _g \frac{a^2+b^2}{a^2-b^2}\varvec{e}_2= \underbrace{\mathrm{curl}\, \varvec{u}}_{rot. fluid}- \underbrace{2{{\dot{\phi }}}\varvec{e}_3}_{rot. mantle}. \end{aligned}$$
(E.6)

In order to relate equation (E.6) to the difference \( {\varvec{\omega }}_{c,pr}-{\varvec{\omega }}_{m,pr}\) we use that the moments of inertia of an ellipsoid of revolution of constant density \(\rho \) are \(\overline{I}_{ce}=\frac{4\pi }{15}\rho (a^2+b^2)a^2b\) and \(\overline{I}_{c3}=\frac{8\pi }{15}\rho a^4b\) that implies

$$\begin{aligned} f_c= \frac{\overline{I}_{c3}-\overline{I}_{ce}}{\overline{I}_{c3}}\approx \frac{\overline{I}_3-\overline{I}_{ce}}{\overline{I}_{ce}}=\frac{a^2-b^2}{a^2+b^2}. \end{aligned}$$
(E.7)

Therefore, using the second hypothesis of Robert and Stewartson \(\frac{\eta _c}{f_c\omega }\ll 1\) (equation (E.3)) equations (E.1), (E.2), (E.6), and (E.7) imply the claimed result:

$$\begin{aligned} {\varvec{\omega }}_{c,pr}-{\varvec{\omega }}_{m,pr}\approx \frac{1}{2}\bigg (\mathrm{curl}\, \varvec{u}- 2{{\dot{\phi }}}\varvec{e}_3\bigg )= \frac{{{\dot{\psi }}}_g}{f_c}\sin \theta _g\varvec{e}_2\,. \end{aligned}$$
(E.8)

We could also have computed the total angular momentum of the fluid with respect to the precessional frame and arrived at the same result.

1.2 The resonant case and the Cassini states of G. Boué

The Cassini states of a body made of a rigid mantle and a fluid core with no viscous coupling (\(k_c=\eta _c=0\)) were computed in Boué (2020). In ibid, the difference between the angular velocities of mantle and core was not supposed small and many possible Cassini states were found. Our goal in the rest of this Appendix is to compare the inertial offsets \({\varvec{\delta }}_m\) and \({\varvec{\delta }}_c\) with some of the Cassini states found in Boué (2020).

The expressions for \({\varvec{\delta }}_m\) and \({\varvec{\delta }}_c\) in equations (7.24) simplify when we assume that the moment of inertia of the core is not much smaller than that of the mantle is rigid and \(\eta _c=0\):

$$\begin{aligned} \begin{aligned}&\delta _m= \delta _{m2}= - \frac{\frac{\mathrm{I}_{\circ c}}{\mathrm{I}_{\circ m}} \frac{{{\dot{\psi }}}_g}{\omega } \cos \theta _g}{\left( \frac{{{\dot{\psi }}}_g}{\omega } \cos \theta _g\right) ^2 +\frac{{{\dot{\psi }}}_g}{\omega } \cos \theta _g\frac{{\,\mathrm I}_\circ }{\mathrm{I}_{\circ m}}(y+z)+\frac{{\,\mathrm I}_\circ }{\mathrm{I}_{\circ m}} y z } \frac{{{\dot{\psi }}}_g}{\omega }\sin \theta _g\,, \\&\delta _c=\delta _{c2} = \frac{\frac{{\,\mathrm I}_\circ }{\mathrm{I}_{\circ m}} z+ \frac{{{\dot{\psi }}}_g}{\omega } \cos \theta _g}{\left( \frac{{{\dot{\psi }}}_g}{\omega } \cos \theta _g\right) ^2 +\frac{{{\dot{\psi }}}_g}{\omega } \cos \theta _g\frac{{\,\mathrm I}_\circ }{\mathrm{I}_{\circ m}}(y+z)+\frac{{\,\mathrm I}_\circ }{\mathrm{I}_{\circ m}} y z }\frac{{{\dot{\psi }}}_g}{\omega }\sin \theta _g\,, \end{aligned} \end{aligned}$$
(E.9)

where \(z=c_1 \overline{\alpha }_e +\frac{c_2}{2}\overline{\gamma }\) and \(y=f_c\). According to equation (7.17) all the three vectors: the normal to the invariable plane \(\varvec{e}_3\in \kappa \), the angular velocity of the mantle \(\langle {\varvec{\omega }}_{m}\rangle _{pr}\in \kappa \), and the angular velocity of the core \(\langle {\varvec{\omega }}_{c}\rangle _{pr}\in \kappa \), both averaged in the precessional frame; are contained in the same plane, which is denoted by \(\varSigma \) in Fig. 8 LEFT.

The Cassini states in Boué (2020) are determined by the following two equations (Equations (34a) and (34b) in ibid.)Footnote 34

$$\begin{aligned} \begin{aligned}&f_c{{\dot{\phi }}}_g\cos (\delta _m-\delta _c)\sin (\delta _m-\delta _c)+ {{\dot{\psi }}}_g\sin ({{\tilde{\theta }}}_g-\delta _c)=0 \\&\overline{I}_{c3}\Big ({{\dot{\psi }}}_g\sin ({{\tilde{\theta }}}_g-\delta _m)+f_c{{\dot{\phi }}}_g \cos (\delta _m-\delta _c)\sin (\delta _m-\delta _c)\Big )-\frac{\overline{I}_3\omega ^2}{{{\dot{\phi }}}_g}P_e(\delta _m)=0 \end{aligned} \end{aligned}$$
(E.10)

where

$$\begin{aligned} \begin{aligned} P_e(\delta _m)=&\frac{3}{2}\frac{G m_p}{a_p^3\omega ^2}\bigg (\overline{\alpha }_e X^{-3,0}_0\cos ({{\tilde{\chi }}}_p-\delta _m)\sin ({{\tilde{\chi }}}_p-\delta _m)\\ {}&+\frac{\overline{\gamma }}{4}X^{-3,2}_s \big (1+\cos ({{\tilde{\chi }}}_p-\delta _m)\big )\sin ({{\tilde{\chi }}}_p-\delta _m)\bigg ) +\frac{{{\dot{\psi }}}_g{{\dot{\phi }}}_g}{\omega ^2}\sin ({{\tilde{\theta }}}_g-\delta _m) \,.\end{aligned} \end{aligned}$$
(E.11)

The identity \(P_e(0)=0\) holds due to Peale’s equation (7.6).

Since our inertial offset was computed by means of a perturbation of a rigid-body motion, for which \(\delta _m=\delta _c=0\), it is natural to look for solutions \((\delta _m,\delta _c)\approx (0,0)\) to equations (E.10). For \((\delta _m,\delta _c)=(0,0)\), the first equation in (E.10) implies \({{\dot{\psi }}}_g\sin ({{\tilde{\theta }}}_g)=0\) and then the second equation implies

$$\begin{aligned} P_e(0)=0\Longrightarrow \bigg (\overline{\alpha }_e X^{-3,0}_0\cos ({{\tilde{\chi }}}_p) +\frac{\overline{\gamma }}{4}X^{-3,2}_s \big (1+\cos ({{\tilde{\chi }}}_p)\big )\bigg )\sin ({{\tilde{\chi }}}_p)=0\,. \end{aligned}$$
(E.12)

This equation has several solutions \({{\tilde{\chi }}}_p=0\), \({{\tilde{\chi }}}_p=\pi \), and \(\cos {{\tilde{\chi }}}_p= -\frac{\overline{\gamma }X^{-3,2}_s}{4 \overline{\alpha }_e X^{-3,0}_0+\overline{\gamma }X^{-3,2}_s}\), where \({{\tilde{\chi }}}_p\) is the obliquity of the body spin to the orbital plane. Among these solutions only \({{\tilde{\chi }}}_p=0\) has obliquity smaller then \(\pi /2\) (except \(\cos {{\tilde{\chi }}}_p= -\frac{\overline{\gamma }X^{-3,2}_s}{4 \overline{\alpha }_e X^{-3,0}_0+\overline{\gamma }X^{-3,2}_s}\) with s=1, for which the obliquity is smaller but close to \(\pi /2\)). We will only consider solutions to Eq. (E.10) with obliquities smaller than \(\pi /2\) that originate from \(\tilde{\chi }_p=0\).

In order to show that the small solutions \((\delta _m,\delta _c)\) to Eqs. (E.10) are given by Eq. (E.9), it is enough to expand the functions in the left-hand side of equations (E.10) up to first order in \((\delta _m,\delta _c)\) and then to solve the linear system. This was done using the algebraic manipulator Mathematica.

The Hessian determinant of equation (E.10) is proportional to \((x_{\ell a}+\frac{{{\dot{\psi }}}_g}{\omega } \cos \theta _g) (x_{dw}+\frac{{{\dot{\psi }}}_g}{\omega } \cos \theta _g)\), where \(\omega x_{dw}\) and \(\omega x_{\ell a}\) are the eigenfrequencies of the nearly diurnal free wobble (NDFW) and the free libration in latitude (FLL) in the inertial space. This shows that the critical flattening of the core \(f_c\approx 0.005 \overline{\alpha }_c\) found for Mercury in Section 6.1 of Boué (2020) (see also Fig. 3 in ibid.) is indeed a resonance of the precession angular speed \(-\dot{\psi }_d\cos \theta _g\approx \)330 kyr with the NDFW eigenfrequency, which decreases to zero as \(f_c\rightarrow 0\). The presence of dissipation attenuates the singularity, as illustrated in Fig. 9.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ragazzo, C., Boué, G., Gevorgyan, Y. et al. Librations of a body composed of a deformable mantle and a fluid core. Celest Mech Dyn Astr 134, 10 (2022). https://doi.org/10.1007/s10569-021-10055-3

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10569-021-10055-3

Keywords

Navigation