Skip to main content
Log in

The effects of deformation inertia (kinetic energy) in the orbital and spin evolution of close-in bodies

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

Abstract

The purpose of this work is to evaluate the effect of deformation inertia on tide dynamics, particularly within the context of the tide response equations proposed independently by Boué et al. (Celest Mech Dyn Astron 126:31–60, 2016) and Ragazzo and Ruiz (Celest Mech Dyn Astron 128(1):19–59, 2017). The singular limit as the inertia tends to zero is analyzed, and equations for the small inertia regime are proposed. The analysis of Love numbers shows that, independently of the rheology, deformation inertia can be neglected if the tide-forcing frequency is much smaller than the frequency of small oscillations of an ideal body made of a perfect (inviscid) fluid with the same inertial and gravitational properties of the original body. Finally, numerical integration of the full set of equations, which couples tide, spin and orbit, is used to evaluate the effect of inertia on the overall motion. The results are consistent with those obtained from the Love number analysis. The conclusion is that, from the point of view of orbital evolution of celestial bodies, deformation inertia can be safely neglected. (Exceptions may occur when a higher-order harmonic of the tide forcing has a high amplitude.)

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
Fig. 15
Fig. 16
Fig. 17

Similar content being viewed by others

Notes

  1. In order to understand this claim, it is enough to analyze the one degree of freedom harmonic oscillator \(\mu \ddot{x}+\eta \dot{x}+\gamma x=\cos (\omega t)\), with solution \(x(t)=a\cos (\omega t)-b\sin (\omega t)\), where \(a+ib\) plays the role of the Love number. Multiplying both sides of the equation by x and time-averaging gives \(\lim _{T\rightarrow \infty }\frac{1}{T}\int _0^T \cos (\omega t) x(t)\hbox {d}t=\frac{a}{2} =\lim _{T\rightarrow \infty }\frac{1}{T}\int _0^T [-\mu \dot{x}^2(t)+\gamma x^2(t)]\hbox {d}t\) that is the time-average balance of kinetic and potential energy. This same reasoning applied to the linear equations of tide dynamics, equation (47) in Ragazzo and Ruiz (2017), shows that the time-average tidal balance of energy is proportional to the real part of the Love number.

  2. The value of \(\gamma /\gamma _{\,\mathrm I}\) for the Sun given in Table 1 does not coincide with that given in Ragazzo and Ruiz (2015). The value \({\mathrm{I}_\circ }/mR^2=0.059\) used in Ragazzo and Ruiz (2015) was taken from Yoder (1995, p. 26). The value \({\mathrm{I}_\circ }/mR^2=0.070\) used here is taken from https://nssdc.gsfc.nasa.gov/planetary/factsheet/sunfact.html. Using \({\mathrm{I}_\circ }/mR^2=0.070\), the new value of \(\gamma \) for the Sun is \(\gamma = 3.668\times 10^{-6}\,\mathrm{s}\).

  3. Notice that the \(\varepsilon \) defined in Eq. (23) can also be written as \(\varepsilon =\tau _0^2/\tau _2^2\), since for the Kelvin–Voigt rheology \(\tau _1=0\) and \(\tau _2=\tau _1+\tau _3=\tau _3\). So, there is no ambiguity in using the same letter in both Eqs. (23) and (26).

  4. These high-frequency oscillations are the source of difficulty for the numerical integration of the full equations of motion. The small inertia approximation in “Appendix” smooths out these oscillations by means of an averaging.

References

  • Antognini, F., Biasco, L., Chierchia, L.: The spin–orbit resonances of the Solar system: a mathematical treatment matching physical data. J Nonlinear Sci. 24, 473–492 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Bambusi, D., Haus, E.: Asymptotic behavior of an elastic satellite with internal friction. Math. Phys. Anal. Geom. 18(1), 1–18 (2015)

    Article  MathSciNet  MATH  Google Scholar 

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

    MATH  Google Scholar 

  • Boué, G., Correia, A.C.M., Laskar, J.: Complete spin and orbital evolution of close-in bodies using a Maxwell viscoelastic rheology. Celest. Mech. Dyn. Astron. 126, 31–60 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Celletti, A.: Analysis of resonances in the spin–orbit problem in celestial mechanics: the synchronous resonance (Part I). J. Appl. Math. Phys. (ZAMP) 41, 174–204 (1990)

    Article  MATH  Google Scholar 

  • Chandrasekhar, S.: Ellipsoidal Figures of Equilibrium. Dover, New York (1987)

    MATH  Google Scholar 

  • Correia, A.C.M., Boué, G., Laskar, J., Rodríguez, A.: Deformation and tidal evolution of close-in planets and satellites using a Maxwell viscoelastic rheology. Astron. Astrophys. 571, A50 (2014)

    Article  Google Scholar 

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

    Article  ADS  MathSciNet  Google Scholar 

  • Efroimsky, M., Williams, J.G.: Tidal torques: a critical review of some techniques. Celest. Mech. Dyn. Astron. 104, 257–289 (2009)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Ferraz-Mello, S.: Tidal synchronization of close-in satellites and exoplanets. A rhephysical approach. Celest. Mech. Dyn. Astron. 116, 109–140 (2013)

    Article  ADS  Google Scholar 

  • Hale, J.K.: Ordinary Differential Equations, 2nd edn. Robert E. Krieger Publishing Company, New York (1980)

    MATH  Google Scholar 

  • Kaper, T.J.: Systems theory for singular perturbation problems. In: Analyzing Multiscale Phenomena Using Singular Perturbation Methods: American Mathematical Society Short Course, vol. 56, Baltimore, MD, 5–6 Jan 1998 (1999)

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

    MATH  Google Scholar 

  • Love, A.E.H.: Some Problems of Geodynamics: Being an Essay to which the Adams Prize in the University of Cambridge was Adjudged in 1911. CUP Archive (1911)

  • Mignard, F.: The evolution of the lunar orbit revisited I. Moon Planets 20, 301–315 (1979)

    Article  ADS  MATH  Google Scholar 

  • Munk, W., MacDonald, G.: The Rotation of the Earth. Cambridge University Press, New York (1960)

    MATH  Google Scholar 

  • Nishida, K.: Earth’s background free oscillations. Annu. Rev. Earth Planet. Sci. 41, 719–740 (2013)

    Article  ADS  Google Scholar 

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

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

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Ragazzo, C., Ruiz, L.: Viscoelastic tides: models for use in celestial mechanics. Celest. Mech. Dyn. Astron. 128(1), 19–59 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Ray, R.D., Eanes, R.J., Lemoine, F.G.: Constraints on energy dissipation in the Earth’s body tide from satellite tracking and altimetry. Geophys. J Int. 144, 471–480 (2001)

    Article  ADS  Google Scholar 

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

    Article  ADS  Google Scholar 

  • Sanders, J.A., Verhulst, F., Murdock, J.A.: Averaging Methods in Nonlinear Dynamical Systems, vol. 59. Springer, New York (2007)

    MATH  Google Scholar 

  • Williams, J.G., Boggs, D.H., Yoder, C.F., Todd Ratcliff, J., Dickey, J.O.: Lunar rotational dissipation in solid body and molten core. J. Geophys. Res. 106, 933–968 (2001)

    Google Scholar 

  • Wisdom, J., Meyer, J.: Dynamic elastic tides. Celest. Mech. Dyn. Astron. 126, 1–30 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Yoder, C.: Astrometric and geodetic properties of Earth and the Solar System. In: Ahrens, T.J. (ed.) Global Earth Physics: A Handbook of Physical Constants, vol. 1, pp. 1–31. American Geophysical Union, Washington (1995)

    Google Scholar 

  • Zlenko, A.A.: A celestial-mechanical model for the tidal evolution of the Earth–Moon system treated as a double planet. Astron. Rep. 59, 72–87 (2014)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

We are grateful to Sylvio Ferraz Mello for having called our attention to the work of Love about the effect of inertia on tides. We also thank Yeva Gevorgyan who first pointed out the equivalence between the Association Principle and the Correspondence Principle in the frequency domain. C.R. is partially supported by FAPESP 2016/25053-8. A.C. is partially supported by CIDMA strategic project (UID/MAT/04106/2013), ENGAGE SKA (POCI-01-0145-FEDER-022217), and PHOBOS (POCI-01-0145-FEDER-029932), funded by COMPETE 2020 and FCT, Portugal.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to A. C. M. Correia.

Additional information

This article is part of the topical collection on Recent advances in the study of the dynamics of N-body problem.

Guest Editors: Giovanni Federico Gronchi, Ugo Locatelli, Giuseppe Pucacco and Alessandra Celletti.

Appendix: The small inertia limit for the Maxwell oscillator

Appendix: The small inertia limit for the Maxwell oscillator

In this appendix, we apply three steps of averaging to the Maxwell oscillator in Eq. (3) that can be written as (see Eq. 49):

$$\begin{aligned} \tau _1\tau _0^2\dddot{x}+\tau _0^2\ddot{x}+ \tau _2\dot{x} +x=\frac{F(t)}{\gamma }+\tau _1 \frac{\dot{F}(t)}{\gamma } \end{aligned}$$
(61)

or

$$\begin{aligned} \frac{\mu \eta }{\gamma \alpha }\dddot{x}+\frac{\mu }{\gamma }\ddot{x}+ \eta \frac{\alpha +\gamma }{\alpha \gamma }\dot{x} +x=\frac{F(t)}{\gamma }+\frac{\eta }{\alpha } \frac{\dot{F}(t)}{\gamma }. \end{aligned}$$

Let \(t=s\tau _4\) define a nondimensional time s, where \(\tau _4=\sqrt{\mu /(\alpha +\gamma )}\), and \(x^\prime (s)=\frac{\hbox {d}}{\hbox {d}s}x(s)\). Then in the nondimensional time scale the equation for the Maxwell oscillator becomes

$$\begin{aligned} x^{\prime \prime \prime }+x^\prime +\sqrt{\varepsilon \frac{\tau _1}{\tau _2}} \left( \frac{\tau _2}{\tau _1}x^{\prime \prime }+x-H(t)\right) =0, \end{aligned}$$

where

$$\begin{aligned} H(t)=\frac{F(t)}{\gamma }+\tau _1 \frac{\dot{F}(t)}{\gamma },\quad \sqrt{\varepsilon }=\frac{\tau _0}{\tau _2}, \quad \tau _1=\frac{\eta }{\alpha },\quad \tau _2=\eta \frac{\gamma +\alpha }{\gamma \alpha }. \end{aligned}$$

Using the definitions

$$\begin{aligned} x^\prime =u,\quad x^{\prime \prime }=v=u^\prime ,\quad w=x+v, \end{aligned}$$

and \(\tau _3=\eta /\gamma \), the third-order equation can be written as

$$\begin{aligned} \begin{aligned} u^\prime =&v\\ v^\prime =&-u-\sqrt{\varepsilon \frac{\tau _1}{\tau _2}}\left( \frac{\alpha }{\gamma }v+w-H(t)\right) \\ w^\prime =&-\sqrt{\varepsilon \frac{\tau _1}{\tau _2}}\left( \frac{\alpha }{\gamma }v+w-H(t)\right) \\ t^\prime =&\sqrt{\varepsilon \tau _1\tau _2}. \end{aligned} \end{aligned}$$

This equation can be explicitly solved when \(\varepsilon =0\). The solution motivates the change of variables \((u,v,w,t)\rightarrow (p,q,w,t)\) where

$$\begin{aligned} \left( \begin{matrix} p \\ q\end{matrix}\right) = \left( \begin{matrix} \cos s &{} - \sin s \\ \sin s &{} \cos s\end{matrix}\right) \left( \begin{matrix} u \\ v\end{matrix}\right) . \end{aligned}$$

In the new variables, the equation becomes

$$\begin{aligned} \begin{aligned} p^\prime =&\sqrt{\varepsilon \frac{\tau _1}{\tau _2}}\sin (s) \left( \frac{\alpha }{\gamma }v+w-H(t)\right) \\ q^\prime =&-\sqrt{\varepsilon \frac{\tau _1}{\tau _2}}\cos (s)\left( \frac{\alpha }{\gamma }v+w-H(t)\right) \\ w^\prime =&-\sqrt{\varepsilon \frac{\tau _1}{\tau _2}}\left( \frac{\alpha }{\gamma }v+w-H(t)\right) \\ t^\prime =&\sqrt{\varepsilon \tau _1\tau _2}, \quad \text {where}\\ v(s)=&-\sin (s)p+\cos (s)q. \end{aligned} \end{aligned}$$

Using the definitions

$$\begin{aligned} z=\left( \begin{matrix} p \\ q\end{matrix}\right) ,\quad A(s)=\frac{1}{4}\left( \begin{matrix} \sin 2s &{} - \cos 2s \\ -\cos 2s &{} -\sin 2s\end{matrix}\right) , \quad \xi (s)=\left( \begin{matrix} \cos s \\ \sin s\end{matrix}\right) ,\quad \varGamma =\frac{\gamma }{\alpha }(-w+H(t)), \end{aligned}$$

this equation can be written as

$$\begin{aligned} \begin{aligned} z^\prime =&\sqrt{\varepsilon }\frac{\tau _3}{\sqrt{\tau _1\tau _2}}\left( -\frac{z}{2}+A^\prime z+\varGamma \xi ^\prime \right) \\ w^\prime =&\sqrt{\varepsilon }\frac{\tau _3}{\sqrt{\tau _1\tau _2}}\left( -\xi ^\prime \cdot z+\varGamma \right) \\ t^\prime =&\sqrt{\varepsilon \tau _1\tau _2}, \quad \text {where}\\ \frac{\hbox {d}}{\hbox {d}s}A(s)=&A^\prime (s)=\frac{1}{2}\left( \begin{matrix} \cos 2s &{} \sin 2s \\ \sin 2s &{} -\cos 2s\end{matrix}\right) , \quad \frac{\hbox {d}}{\hbox {d}s}\xi (s)= \xi ^\prime (s)=\left( \begin{matrix} -\sin s \\ \cos s\end{matrix}\right) . \end{aligned} \end{aligned}$$

The following change of variables \((z,w,t)\rightarrow (y,w,t)\) was obtained after two steps of averaging Hale (1980),

$$\begin{aligned} \begin{aligned}&z=y+(Ay+\varGamma \xi )\sqrt{\varepsilon }\frac{\tau _3}{\sqrt{\tau _1\tau _2}}+(cA^\prime y+\varLambda \xi ^\prime )\varepsilon \frac{\tau ^2_3}{\tau _1\tau _2},\quad \text {where}\\&c=\frac{1}{8}-\frac{\gamma }{2\alpha },\quad \varLambda =\left( -\varGamma +\tau _2\frac{\gamma }{\alpha }\dot{H}\right) \frac{\gamma }{\alpha }. \end{aligned} \end{aligned}$$
(62)

Applying this change of variables, using \(\varGamma ^\prime =\sqrt{\varepsilon \frac{\tau _1}{\tau _2}}(\xi ^\prime \cdot z+\frac{\alpha }{\gamma }\varLambda )\) and \(\varLambda ^\prime =\{-\varGamma ^\prime +\sqrt{\varepsilon \frac{\tau _1}{\tau _2}}\tau _2^2\frac{\gamma }{\alpha }\ddot{H}\}\frac{\gamma }{\alpha }\), and averaging the resulting equations we obtain, after a long computation,

$$\begin{aligned} \begin{aligned} y^\prime&=\sqrt{\varepsilon }\frac{\tau _3}{\sqrt{\tau _1\tau _2}} \left( -\left\{ \frac{1}{2}+\varepsilon \frac{\tau _3}{\tau _2}c \right\} y+ \sqrt{\varepsilon }\frac{\tau _3}{\sqrt{\tau _1\tau _2}} \left( \frac{1}{8}+\frac{\gamma }{2\alpha }\right) \left( \begin{matrix} 0 &{} -1 \\ 1 &{} 0\end{matrix}\right) y\right) +\mathscr {O}(\varepsilon ^2)\\ w^\prime&=\sqrt{\varepsilon }\frac{\tau _3}{\sqrt{\tau _1\tau _2}} \left( \varGamma -\varepsilon \frac{\tau _3^2}{\tau _1\tau _2} \varLambda \right) +\mathscr {O}(\varepsilon ^2). \end{aligned} \end{aligned}$$

In the original time t, and neglecting the correction term, these equations become

$$\begin{aligned} \begin{aligned} \tau _2 \dot{y}&=\frac{\tau _3}{\tau _1}\left( -\left\{ \frac{1}{2}+ \varepsilon \frac{\tau _3}{\tau _2}c \right\} y+\sqrt{\varepsilon }\frac{\tau _3}{\sqrt{\tau _1\tau _2}} \left( \frac{1}{8}+\frac{\gamma }{2\alpha }\right) \left( \begin{matrix} 0 &{} -1 \\ 1 &{} 0\end{matrix}\right) y\right) \\ \tau _2 \dot{w}&=\frac{\tau _3}{\tau _1} \left( \varGamma -\varepsilon \frac{\tau _3^2}{\tau _1\tau _2} \varLambda \right) . \end{aligned} \end{aligned}$$

The second of these equations implies the first equation in (27).

Notice that \(|y(t)|\rightarrow 0\) as \(t\rightarrow \infty \), so after some transient where |y(t)| decay as \(\exp [-t\tau _3/(2\tau _1\tau _2)]\) Eq. (62) implies

$$\begin{aligned} z=\varGamma \xi \sqrt{\varepsilon }\frac{\tau _3}{\sqrt{\tau _1\tau _2}}+ \varLambda \xi ^\prime \varepsilon \frac{\tau _3^2}{\tau _1\tau _2}. \end{aligned}$$

This and the relation \(x=w-v=w+\sin (s)p-\cos (s) q=w-\xi ^\prime (s)\cdot z\) imply

$$\begin{aligned} x=w-\xi ^\prime \cdot z=w-\varGamma \underbrace{\xi ^\prime \cdot \xi }_{=0} \sqrt{\varepsilon }\frac{\tau _3}{\sqrt{\tau _1\tau _2}}- \varLambda \underbrace{\xi ^\prime \cdot \xi ^\prime }_{=1} \varepsilon \frac{\tau _3^2}{\tau _1\tau _2} =w-\varepsilon \frac{\tau _3^2}{\tau _1\tau _2}\varLambda . \end{aligned}$$

This equation plus the definition of \(\varLambda \) in Eq. (62) implies the second equation in (27).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

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 Astr 130, 51 (2018). https://doi.org/10.1007/s10569-018-9847-3

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10569-018-9847-3

Keywords

Navigation