tidal heating in icy satellite oceans€¦ · ixty1 2e xi ty 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p...

20
Tidal heating in icy satellite oceans E.M.A. Chen, F. Nimmo , G.A. Glatzmaier Dept. Earth and Planetary Sciences, University of California, Santa Cruz, 1156 High St., Santa Cruz, CA 95064, USA article info Article history: Received 14 May 2013 Revised 21 October 2013 Accepted 22 October 2013 Available online 30 October 2013 Keywords: Satellites, dynamics Tides, solid body Jupiter, satellites Saturn, satellites Triton abstract Tidal heating plays a significant role in the evolution of many satellites in the outer Solar System; how- ever, it is unclear whether tidal dissipation in a global liquid ocean can represent a significant additional heat source. Tyler (Tyler, R.H. [2008]. Nature 456, 770-772; Tyler, R.H. [2009]. Geophys. Res. Lett. 36, doi:10.1029/2009GL038300) suggested that obliquity tides could drive large-scale flow in the oceans of Europa and Enceladus, leading to significant heating. A critical unknown in this previous work is what the tidal quality factor, Q, of such an ocean should be. The corresponding tidal dissipation spans orders of magnitude depending on the value of Q assumed. To address this issue we adopt an approach employed in terrestrial ocean modeling, where a significant portion of tidal dissipation arises due to bottom drag, with the drag coefficient O (0.001) being relatively well-established. From numerical solutions to the shallow-water equations including nonlinear bottom drag, we obtain scalings for the equivalent value of Q as a function of this drag coefficient. In addition, we provide new scaling relations appropriate for the inclusion of ocean tidal heating in thermal–orbital evolution models. Our approach is appropriate for situations in which the ocean bottom topography is much smaller than the ocean thickness. Using these novel scalings, we calculate the ocean contribution to the overall thermal energy budgets for many of the outer Solar System satellites. Although uncertainties such as ocean thickness and satellite obliquity remain, we find that for most satellites it is unlikely that ocean tidal dissipation is important when compared to either radiogenic or solid-body tidal heating. Of known satellites, Triton is the most likely icy satellite to have ocean tidal heating play a role in its present day thermal budget and long-term thermal evolution. Ó 2013 Elsevier Inc. All rights reserved. 1. Introduction Tidal heating influences the present-day behavior of some plan- etary bodies, such as Io (Peale et al., 1979) and Enceladus (Spencer et al., 2006; Howett et al., 2011). It probably also played a role at earlier times elsewhere, including Europa (Hussmann and Spohn, 2004), Ganymede (Showman et al., 1997), Triton (Jankowski et al., 1989), and the Moon (Garrick-Bethell et al., 2010), and may be important in some super-Earth exoplanets (Henning et al., 2009). For solid bodies, the effects of tides and their associated dissipa- tion are typically calculated assuming a viscoelastic rheology, such as Maxwell, Andrade or Burgers (e.g. Ross and Schubert (1989); To- bie et al. (2005); Efroimsky and Williams (2009); Castillo-Rogez et al. (2011); Nimmo et al. (2012)), though other processes (such as frictional heating, e.g. Nimmo and Gaidos (2002)) may also play a role. For primarily fluid bodies, such as giant planets, a significant component of dissipation is likely to be due to the breaking of internal gravity waves (Ogilvie and Lin, 2004). Lastly, fluid layers on or within solid bodies may also be a source of dissipation. On the Earth it is well-known that tidal dissipation occurs mainly in the oceans (Munk and MacDonald, 1960; Egbert and Ray, 2000; Ray et al., 2001). Global subsurface oceans are thought to occur on at least Europa, Ganymede, Callisto, Titan and perhaps Encela- dus (Khurana et al., 1998; Kivelson et al., 2002; Bills and Nimmo, 2011; Iess et al., 2012; Postberg et al., 2011); our focus in this paper is to examine tidal dissipation within such oceans. In a prescient paper, Ross and Schubert (1989) discussed the pos- sibility of tidal heating on Enceladus arising from turbulent dissipa- tion in a subsurface ocean. More recently, Tyler (2011) expanded an analysis initially developed by Longuet-Higgins (1968) to investi- gate energy dissipation in tidally-driven satellite oceans. In Tyler (2011), a key free parameter is the linear drag constant a, which can be related to a tidal quality factor Q. It is worth noting that the value of a or Q is a priori very poorly known and the total energy dis- sipation scales linearly with the model’s prescribed value. In this paper, we follow an analysis similar to that of Tyler (2011). However, we depart from his approach in two important respects. First, we provide an estimate for Q using an approach and parameter values developed in studies of the Earth’s oceans. 0019-1035/$ - see front matter Ó 2013 Elsevier Inc. All rights reserved. http://dx.doi.org/10.1016/j.icarus.2013.10.024 Corresponding author. E-mail address: [email protected] (F. Nimmo). Icarus 229 (2014) 11–30 Contents lists available at ScienceDirect Icarus journal homepage: www.elsevier.com/locate/icarus

Upload: others

Post on 03-Jul-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

Icarus 229 (2014) 11–30

Contents lists available at ScienceDirect

Icarus

journal homepage: www.elsevier .com/locate / icarus

Tidal heating in icy satellite oceans

0019-1035/$ - see front matter � 2013 Elsevier Inc. All rights reserved.http://dx.doi.org/10.1016/j.icarus.2013.10.024

⇑ Corresponding author.E-mail address: [email protected] (F. Nimmo).

E.M.A. Chen, F. Nimmo ⇑, G.A. GlatzmaierDept. Earth and Planetary Sciences, University of California, Santa Cruz, 1156 High St., Santa Cruz, CA 95064, USA

a r t i c l e i n f o a b s t r a c t

Article history:Received 14 May 2013Revised 21 October 2013Accepted 22 October 2013Available online 30 October 2013

Keywords:Satellites, dynamicsTides, solid bodyJupiter, satellitesSaturn, satellitesTriton

Tidal heating plays a significant role in the evolution of many satellites in the outer Solar System; how-ever, it is unclear whether tidal dissipation in a global liquid ocean can represent a significant additionalheat source. Tyler (Tyler, R.H. [2008]. Nature 456, 770-772; Tyler, R.H. [2009]. Geophys. Res. Lett. 36,doi:10.1029/2009GL038300) suggested that obliquity tides could drive large-scale flow in the oceansof Europa and Enceladus, leading to significant heating. A critical unknown in this previous work is whatthe tidal quality factor, Q, of such an ocean should be. The corresponding tidal dissipation spans orders ofmagnitude depending on the value of Q assumed.

To address this issue we adopt an approach employed in terrestrial ocean modeling, where a significantportion of tidal dissipation arises due to bottom drag, with the drag coefficient O (0.001) being relativelywell-established. From numerical solutions to the shallow-water equations including nonlinear bottomdrag, we obtain scalings for the equivalent value of Q as a function of this drag coefficient. In addition,we provide new scaling relations appropriate for the inclusion of ocean tidal heating in thermal–orbitalevolution models. Our approach is appropriate for situations in which the ocean bottom topography ismuch smaller than the ocean thickness.

Using these novel scalings, we calculate the ocean contribution to the overall thermal energy budgetsfor many of the outer Solar System satellites. Although uncertainties such as ocean thickness and satelliteobliquity remain, we find that for most satellites it is unlikely that ocean tidal dissipation is importantwhen compared to either radiogenic or solid-body tidal heating. Of known satellites, Triton is the mostlikely icy satellite to have ocean tidal heating play a role in its present day thermal budget and long-termthermal evolution.

� 2013 Elsevier Inc. All rights reserved.

1. Introduction

Tidal heating influences the present-day behavior of some plan-etary bodies, such as Io (Peale et al., 1979) and Enceladus (Spenceret al., 2006; Howett et al., 2011). It probably also played a role atearlier times elsewhere, including Europa (Hussmann and Spohn,2004), Ganymede (Showman et al., 1997), Triton (Jankowski et al.,1989), and the Moon (Garrick-Bethell et al., 2010), and may beimportant in some super-Earth exoplanets (Henning et al., 2009).

For solid bodies, the effects of tides and their associated dissipa-tion are typically calculated assuming a viscoelastic rheology, suchas Maxwell, Andrade or Burgers (e.g. Ross and Schubert (1989); To-bie et al. (2005); Efroimsky and Williams (2009); Castillo-Rogezet al. (2011); Nimmo et al. (2012)), though other processes (suchas frictional heating, e.g. Nimmo and Gaidos (2002)) may also playa role. For primarily fluid bodies, such as giant planets, a significantcomponent of dissipation is likely to be due to the breaking ofinternal gravity waves (Ogilvie and Lin, 2004). Lastly, fluid layers

on or within solid bodies may also be a source of dissipation. Onthe Earth it is well-known that tidal dissipation occurs mainly inthe oceans (Munk and MacDonald, 1960; Egbert and Ray, 2000;Ray et al., 2001). Global subsurface oceans are thought to occuron at least Europa, Ganymede, Callisto, Titan and perhaps Encela-dus (Khurana et al., 1998; Kivelson et al., 2002; Bills and Nimmo,2011; Iess et al., 2012; Postberg et al., 2011); our focus in this paperis to examine tidal dissipation within such oceans.

In a prescient paper, Ross and Schubert (1989) discussed the pos-sibility of tidal heating on Enceladus arising from turbulent dissipa-tion in a subsurface ocean. More recently, Tyler (2011) expanded ananalysis initially developed by Longuet-Higgins (1968) to investi-gate energy dissipation in tidally-driven satellite oceans. In Tyler(2011), a key free parameter is the linear drag constant a, whichcan be related to a tidal quality factor Q. It is worth noting that thevalue of a or Q is a priori very poorly known and the total energy dis-sipation scales linearly with the model’s prescribed value.

In this paper, we follow an analysis similar to that of Tyler(2011). However, we depart from his approach in two importantrespects. First, we provide an estimate for Q using an approachand parameter values developed in studies of the Earth’s oceans.

Page 2: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

12 E.M.A. Chen et al. / Icarus 229 (2014) 11–30

Second, we present approximate scaling relationships which allowfluid dissipation to be calculated in a manner analogous to thewell-known equations for solid body dissipation (e.g. Segatz et al.(1988); Ross and Schubert (1989); Wisdom (2008)). This will facil-itate investigation of long-term satellite evolution, in which thethermal and orbital histories are coupled (e.g. Ojakangas and Ste-venson (1989); Hussmann and Spohn (2004); Bland et al. (2009);Meyer et al. (2010); Zhang and Nimmo (2012)).

The rest of this paper is organized as follows. Section 2 reviewsthe shallow water equations appropriate for flow in global fluid lay-ers on a rotating spherical shell. For clarity, we summarize a semi-analytic solution to these equations, similar to that adopted byLonguet-Higgins (1968) and Tyler (2011), and in addition, explicitlypresent the method and equations used to calculate quantities suchas the average kinetic energy and energy dissipation. Section 3 sim-plifies this system and carries out an analytical study of the re-sponse of a shallow global ocean to tidal forcing, building on themethod presented in Section 2. This novel analysis derives approx-imate scaling relationships for ocean tidal flow and the resultingdissipation under typical icy satellite parameters. The algebra in-volved can be tedious; to aid clarity, many details have been rele-gated to Appendices E and F, while Table 4 summarizes the keyresults. The advantage of these relationships is that they retainthe fundamental physical effects while being somewhat simplerto implement than the method presented in Section 2. The resultsof Section 3 are expressed in terms of an unknown effective (pre-sumably turbulent) viscosity. In Section 4, we present an estimatefor this viscosity using a numerical technique based on analogy tofrictional ocean dissipation on Earth. We discuss the applicationsand implications of these results in Section 5. In particular, oceandissipation is unlikely to be a significant heat source unless theorbital eccentricity is very small; Triton is thus the most likely can-didate for a satellite in which ocean tidal dissipation is significant.

2. Ocean tidal dissipation

Here we briefly review the equations of motion for a shallowglobal satellite ocean. These equations are equivalent to Eqs. (3)and (4) presented in Tyler (2011); we present fully dimensionalequations and explicitly expand these equations for the solutionsto unknown spherical harmonic coefficients.

The forced, dissipative, shallow water equations on a rotatingsphere are (cf. Longuet-Higgins (1968) Eqs. (13.1)–(13.3) or Tyler(2011) Eqs. (3) and (4), noting the sign difference in the tidal po-tential term)

@~u@tþ 2X cos hr �~u ¼ �g~rg� ~rU � a~uþ mr2~u ð1Þ

and

@g@tþ h~r �~u ¼ 0; ð2Þ

where~u is the radially-averaged, horizontal velocity vector, X is theconstant rotation rate, r is the unit vector in the radial direction, g isthe surface gravity, g is the vertical displacement of the surface, andh is the constant ocean depth. The dissipation can be represented aseither a linear process with a linear coefficient a or a Navier–Stokestype viscosity with a viscous diffusivity of m. We assume no radialgradients in our shallow-water model such that the Laplacian oper-ator, r2, has no radial contributions and thus, m is an effective hor-izontal diffusivity. U represents the forcing potential due to tides,either eccentricity- or obliquity-related. Eqs. (1) and (2) are validfor incompressible flow under the assumptions that the thicknessof the fluid layer is much smaller than the radius of the bodyðh� RÞ, the vertical displacement is much smaller than the layerthickness ðg� hÞ and fluid properties are constant (e.g. a and m).

These equations ignore ocean stratification, and thus do not includethe effects of internal tides. In addition, they do not include overly-ing ice shell rigidity, though this effect should be small (Matsuyama,2012).

2.1. Tidal potentials

For synchronously rotating satellites, such as the regular sat-ellites of Jupiter and Saturn, we are concerned with the oceanflow driven by the eccentricity of the orbit and the obliquity,the tilt of the rotational axis relative to the orbital axis. The forc-ing tidal potentials can be derived by assuming the planet is apoint mass and calculating the resulting gravitational potentialat every point on the satellite (cf. Kaula (1964); Murray and Der-mott (1999)).

2.1.1. Obliquity tidesThe obliquity tidal potential at a point of colatitude h and longi-

tude / on a synchronously rotating satellite with small obliquity h0

(in radians) is a standing wave and can be written as the sum of aneastward and a westward propagating potential (cf. Tyler (2011)Eq. (34))

Uobl ¼�32

X2R2h0 sin h cos hðcosð/�XtÞ þ cosð/þXtÞÞ: ð3Þ

We define Laplace spherical harmonics of degree l and order m, Yml ,

as

Yml ðh;/Þ �

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffið2lþ 1Þ

4pðl�mÞ!ðlþmÞ!

sPm

l ðcos hÞeim/ ð4Þ

employing a Condon–Shortley phase factor of ð�1Þm for m > 0.These spherical harmonics are orthogonal underZ 2p

0

Z p

0Ym

l Ym0�l0 sin hdhd/ ¼ dl;l0dm;m0 ð5Þ

where � denotes the complex conjugate and d is the Kronecker delta.The obliquity tidal potential thus can be expressed in spherical

harmonics as a westward propagating potential Uobl;W ,

Uobl;W ¼32

ffiffiffiffiffiffiffi2p15

rX2R2h0ðeiXtY1

2 � e�iXtY�12 Þ

¼ 232

ffiffiffiffiffiffiffi2p15

rX2R2h0

!RðeiXtY1

2Þ � 2U12;WR eiXtY1

2

� �; ð6Þ

and a symmetric eastward propagating potential Uobl;E for which theeiXt term is replaced by e�iXt and U1

2;E ¼ U12;W .

2.1.2. Eccentricity tidesThe eccentricity tidal potential can be expressed as (Kaula,

1964) (cf. Tyler (2011) Eq. (35))

Uecc ¼�34

X2R2e½�ð3 cos2 h� 1Þ cos Xt þ sin2 hð3

� cos 2/ cos Xt þ 4 sin 2/ sin XtÞ�: ð7Þ

For the subsequent analysis, the eccentricity tidal potential can besplit into three separate components. There is an axisymmetriccomponent, Uecc;rad, (Tyler (2011) calls this the ‘‘radial’’ component)

Uecc;rad ¼ 3ffiffiffiffip5

rX2R2e cos Xt

� �Y0

2

¼ 12

3ffiffiffiffip5

rX2R2e

� �ðeiXt þ e�iXtÞY0

2 � U02ðeiXt þ e�iXtÞY0

2 ð8Þ

There is also an asymmetric librational component, Uecc;lib, that canbe split into a westward propagating potential

Page 3: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

E.M.A. Chen et al. / Icarus 229 (2014) 11–30 13

Uecc;lib;W ¼34

ffiffiffiffiffiffiffi2p15

rX2R2e

!eiXtY2

2 þ e�iXtY�22

� �

¼ 234

ffiffiffiffiffiffiffi2p15

rX2R2e

!R eiXtY2

2

� �� 2U2

2;WR eiXtY22

� �ð9Þ

and an eastward propagating potential

Uecc;lib;E ¼�21

4

ffiffiffiffiffiffiffi2p15

rX2R2e

!e�iXtY2

2 þ eiXtY�22

� �

¼ 2�21

4

ffiffiffiffiffiffiffi2p15

rX2R2e

!R e�iXtY2

2

� �

� 2U22;ER e�iXtY2

2

� �: ð10Þ

2.2. Semi-analytic method of solution

The horizontal velocity can be expressed using a Helmholtzdecomposition as the sum of the gradient of a scalar potential Uand the curl of a streamfunction Wr where (cf. Tyler (2011) Eq. (5))

~u ¼ ~rUþ ~r�Wr ¼ uhhþ u//; ð11Þ

with the components of velocity given by

uh ¼1R@U@hþ 1

R sin h@W@/

ð12Þ

and

u/ ¼1

R sin h@U@/� 1

R@W@h

: ð13Þ

The divergence and radial component of the curl of (11) are

~r �~u ¼ r2U; ð14Þr � ð~r�~uÞ ¼ �r2W: ð15Þ

Taking the divergence of the momentum Eq. (1) and making thesubstitution from (14) results in

@

@tþ a� mr2

� �r2 þ 2X

R2

@

@/

� �Uþ 2X cos hr2 � sin h

R2

@

@h

� �W

¼ �gr2g�r2U: ð16Þ

Similarly, the radial component of the curl of the momentum equa-tion with the substitution (15) is

@

@tþa�mr2

� �r2þ2X

R2

@

@/

� �W�2X coshr2�sinh

R2

@

@h

� �U¼0: ð17Þ

Eqs. (16) and (17) are equivalent to Eq. (3.7) of Longuet-Higgins(1968), with the dissipative terms added. Eqs. (16) and (17) arealso equivalent to Eqs. (12) and (13) of Tyler (2011) if one ignoresthe residual terms (R1 and R2) and does not make the substitutionfor g using the continuity equation (see (18) below).

Substituting (14) into (2), the surface displacement is depen-dent only on U,

@g@tþ hr2U ¼ 0: ð18Þ

For the response to a westward propagating tidal potential, weseek wave solutions of the form

Wðh;/; tÞ ¼X1l¼1

Xl

m¼1

Wml eiXtYm

l þW�ml e�iXtY�m

l

¼ 2RX1l¼1

Xl

m¼1

Wml eiXtYm

l

!; ð19Þ

with equivalent expansions for U and g. Here the coefficients Wml

(and their equivalents Uml , and gm

l ) are complex, thereby permittinga phase lag between the tidal forcing and the response. The responsefrequency is assumed to be the same as the forcing frequency.

Using Eq. (18) and the wave solutions for W (19) and its equiv-alents, we have

gml ¼ �i

h

R2Xlðlþ 1ÞUm

l : ð20Þ

Plugging these wave solutions into Eqs. (16) and (17) and pro-jecting the equations onto each spherical harmonic, we obtainequations describing solutions for the coefficients Wm

l and Uml that

are coupled in l but separable in m,

iXþaþmlðlþ1ÞR2 � 2Xim

lðlþ1Þ

� �Wm

l �2Xl�1

lCm

l Uml�1þ

lþ2lþ1

Cmlþ1U

mlþ1

� �¼0; ð21Þ

iXþ aþ mlðlþ 1ÞR2 � 2Xim

lðlþ 1Þ �ighlðlþ 1Þ

XR2

� �Um

l

þ 2Xl� 1

lCm

l Wml�1 þ

lþ 2lþ 1

Cmlþ1W

mlþ1

� �

¼ �U12;Wdl;2dm;1;�U2

2;Wdl;2dm;2

n oð22Þ

with the term on the right hand side of (22) corresponding to theforcing potential of interest, either westward obliquity or westwardlibrational eccentricity. The constants Cm

l are due to differentialoperators of the spherical harmonics, and these are given by

Cml ¼

ðlþmÞðl�mÞð2lþ 1Þð2l� 1Þ

� �1=2

: ð23Þ

Eqs. (21) and (22) are effectively Eq. (3.19) of Longuet-Higgins (1968)with correction of the sign inconsistencies described in Appendix A.1of Tyler (2011). The differences in the constants between the equa-tions presented here and those of Longuet-Higgins (1968) and Tyler(2011) are primarily due to the use of normalized spherical harmon-ics throughout this work and a difference in the sign of the tidalpotentials. For helpful relations, similar to (3.17) and (3.18) of Long-uet-Higgins (1968), we refer the reader to Appendix A. The linear sys-tem defined by Eqs. (21) and (22) can be easily solved numerically forthe coupled coefficients Um

l and Wml , repeating for each potential of

interest. While the series is infinite, typically the coefficients dropoff rapidly with increasing l (Fig. 1). We exploit this fact in Section 3to derive analytical scalings that reduce the complexity of imple-menting ocean tidal effects in thermal evolution models.

A similar projection can be obtained for the eastward forcingpotentials by changing the direction of the response in (19) andits equivalents (see details in Appendix B). Additionally, in (19)we have neglected the contributions from the radial componentof the eccentricity tide ðm ¼ 0Þ. These coefficients must be realand thus, different wave solutions are required. We present thesolution method explicitly in Appendix C. We have presented thewestward response because for satellite oceans the dominant re-sponse is likely to be that to the westward propagating componentof the obliquity tide (Tyler, 2008); for further details, we direct thereader to Section 3.1.

Given the spherical harmonics coefficients for U and W, the aver-age kinetic energy and energy dissipation can then be calculated. Itcan be shown (see details in Appendix D) that the kinetic energyaveraged over the satellite surface and the orbital period is given by:

Etot;avg ¼ qhX1l¼1

lðlþ 1Þ jU0l;W j

2 þ jW0l;W j

2� �"

þX1l¼1

Xl

m¼1

lðlþ 1Þ jUml;W j

2 þ jUml;Ej

2 þ jWml;W j

2 þ jWml;Ej

2� �#

: ð24Þ

Page 4: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

Fig. 1. Power spectrum with varying spherical harmonic degree l of the full semi-analytic solution described in Section 2.2 using parameters relevant to an ocean on Europa(Tables 1 and 2), here with a viscosity of m ¼ 1:0� 102 m2=s for both the obliquity and eccentricity tides. The response of each tidal component is separated. Note thesignificant separation between degrees 1 and 2 in the case of the westward obliquity response and degrees 3 and 4 for all other tidal components, justifying the scalingarguments presented in Section 3 and Appendices E and F.

Table 1Constant parameters used in thermal energy calculations.

Parameter Value First reference

H 4.5 � 10�12 W kg�1 Section 5.1qsil 3500 kg m�3 Eq. (65)qice 950 kg m�3 Eq. (65)Qsol 100 Eq. (66)G 6.67 � 10�11 m3 kg�3 s�2 Eq. (66)l 4 � 109 N m�2 Eq. (67)c0 0.33 Eq. (68)q (ocean) 1000 kg m�3 Eq. (24)h 30 km Eq. (2)cD 0.002 Eq. (54)

14 E.M.A. Chen et al. / Icarus 229 (2014) 11–30

The corresponding average flow speed �u is then given by

�u ¼ Etot;avg

2pR2hq

!1=2

: ð25Þ

The energy dissipation rate similarly averaged (and taken to bepositive) is

_Elin;avg ¼ 2aEtot;avg ¼ 2aqhX1l¼1

lðlþ 1Þ jU0l;W j

2 þ jW0l;W j

2� �"

þX1l¼1

Xl

m¼1

lðlþ 1Þ jUml;W j

2 þ jUml;Ej

2 þ jWml;W j

2 þ jWml;Ej

2� �#

; ð26Þ

for linear dissipation and

_ENS;avg ¼2qhm

R2

X1l¼1

l2ðlþ 1Þ2 jU0l;W j

2 þ jW0l;W j

2� �"

þX1l¼1

Xl

m¼1

l2ðlþ 1Þ2 jUml;W j

2 þ jUml;Ej

2 þ jWml;W j

2 þ jWml;Ej

2� �#

:

ð27Þ

for Navier–Stokes type dissipation.An alternative method of thinking about tidal dissipation is that

in the time- and spatially-averaged sense it is equivalent to thework done by the tides. It can be shown that this work is relatedto distorting the surface (see details in Appendix D.2.2):

_Eavg ¼�2ð2Þð3Þqh U02R U0

2;W

� �þX2

m¼1

Um2;W RðUm

2;W ÞþUm2;ER Um

2;E

� �� �" #: ð28Þ

Note that dissipation depends only on the real part of U and is inde-pendent of W. Two physical consequences follow immediately. First,situations in which there is no surface displacement (g ¼ U ¼ 0 asin Tyler (2008)) will result in no dissipation. Second, while theamount of kinetic energy depends on the magnitude of U (see(24)), the energy dissipation rate depends on only the real compo-nent of U, and thus is sensitive to the phase lag between the tidalforcing and the fluid response, similar to viscoelastic tidal dissipa-tion in solid bodies (Kaula, 1964; Peale and Cassen, 1978; Segatz

et al., 1988). It will be shown in Section 3 and Appendices E and Fthat for a wide range of parameters, the primary role of the viscosityis to set the tidal phase lag without altering the amplitude of theresponse.

2.3. General features of the shallow-water tidally-forced system

In Fig. 1, we show a solution for the spectral coefficients fromthe semi-analytic method for a model global ocean on Europa(see parameters in Tables 1 and 2) including all tidal components.Even though in this case the eccentricity is ten times larger thanthe assumed obliquity, the dominant tidal component is the west-ward obliquity response ðl ¼ 1Þ, as pointed out in Tyler (2008). Thespherical harmonic modes are coupled in degree because of theCoriolis force. This coupling is fairly weak, and the response haspower primarily in the modes around the large scale tidal forcing.For the westward propagating obliquity response, the responses atdegrees 2 and 3 have much smaller amplitudes than that at degree1. Degrees 2 and 3 dominate the eccentricity response. For degreesl > 3, all responses are negligible. This separation between degrees3 and 4 allows us to carry out the scaling arguments presented inSection 3.

A key input parameter to the system is the value of the dissipa-tion coefficient, a or m. We present the average kinetic energy and

Page 5: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

Table 2Satellite specific parameters used in thermal energy calculations. The physical parameters on the left half of the table are taken from Schubert et al. (2004) for the Galileansatellites, Thomas (2010), Zebker et al. (2009) and Jacobson et al. (2006) for the saturnian satellites, Thomas (1988) and Jacobson et al. (1992) for the uranian satellites, andThomas (2000) and Jacobson (2009) for Triton. The orbital parameters are taken from JPL satellite ephemerides (http://ssd.jpl.nasa.gov/?sat_elem). Calculations for the quantitieson the right side of the table are described in Section 5.1; measured (as opposed to calculated) quantities are indicated with ⁄. The gravity coefficients for the Galiean satellites arefrom Galileo measurements (Schubert et al., 2004); others are calculated from a hydrostatic assumption. The obliquity for Titan is taken to be the observed value (Stiles et al.,2008) and is not a theoretical Cassini state value.

R (km) qsat (kg m�3) g (m/s2) X (rad/s) p e i (deg) k2 J2 C2;2 �h0 (deg)

Europa 1565.0 2989 1.31 2.05 � 10�5 �3.10 � 103 0.0094 0.466 2.08 � 10�1 4.36 � 10�4⁄ 1.32 � 10�4⁄ 0.053Ganymede 2631.2 1942 1.43 1.02 � 10�5 �6.77 � 103 0.0013 0.177 2.42 � 10�1 1.28 � 10�4⁄ 3.83 � 10�5⁄ 0.033Callisto 2410.3 1834 1.24 4.36 � 10�6 �7.42 � 103 0.0074 0.192 1.89 � 10�1 3.27 � 10�5⁄ 1.02 � 10�5⁄ 0.24

Mimas 198.2 1150 0.064 7.72 � 10�5 �3.82 � 102 0.0196 1.574 5.73 � 10�4 1.43 � 10�2 4.30 � 10�3 0.041Enceladus 252.1 1610 0.11 5.31 � 10�5 �6.29 � 102 0.0044 0.003 1.81 � 10�3 4.83 � 10�3 1.45 � 10�3 0.00014Tethys 531.0 985 0.15 3.85 � 10�5 �9.64 � 102 0.0001 1.091 3.01 � 10�3 4.17 � 10�3 1.25 � 10�3 0.039Dione 561.4 1478 0.23 2.66 � 10�5 �1.56 � 103 0.0022 0.028 7.56 � 10�3 1.32 � 10�3 3.97 � 10�4 0.0020Rhea 763.5 1237 0.26 1.61 � 10�5 �2.90 � 103 0.0002 0.333 9.78 � 10�3 5.73 � 10�4 1.74 � 10�4 0.030Titan 2574.73 1882 1.35 4.56 � 10�6 �1.61 � 104 0.0288 0.306 2.21 � 10�1 – – 0.32⁄

Miranda 235.8 1200 0.079 5.15 � 10�5 �4.58 � 103 0.0013 4.338 8.84 � 10�4 6.10 � 10�3 1.83 � 10�3 0.021Ariel 578.9 1665 0.27 2.89 � 10�5 �8.30 � 103 0.0012 0.041 1.02 � 10�2 1.39 � 10�3 4.16 � 10�4 0.00050Umbriel 584.7 1399 0.23 1.76 � 10�5 �1.12 � 104 0.0039 0.128 7.35 � 10�3 6.13 � 10�4 1.84 � 10�4 0.0026Titania 788.9 1714 0.38 8.35 � 10�6 �8.20 � 103 0.0011 0.079 1.99 � 10�2 1.13 � 10�4 3.38 � 10�5 0.014Oberon 761.4 1630 0.35 5.40 � 10�6 �5.30 � 103 0.0014 0.068 1.68 � 10�2 4.13 � 10�5 1.48 � 10�5 0.075

Triton 1353.4 2060 0.78 1.24 � 10�5 �4.27 � 104 0.000 156.87 8.11 � 10�2 2.07 � 10�4 6.20 � 10�5 0.35

E.M.A. Chen et al. / Icarus 229 (2014) 11–30 15

energy dissipation for a range of hypothetical viscosities in Figs. 2and 3. The value of the viscosity is important in this model for twomain reasons. First, the value of the viscosity determines thebehavior of the response to the westward propagating obliquitytide. For the tidal responses not due to the westward propagatingobliquity tide, the behavior is fairly straightforward. The averagekinetic energy is relatively independent of the value of the dissipa-tion coefficient prescribed for a large range of values (mK 106 m2=sin Fig. 2). For the westward obliquity response, the behavior is dif-ferent; there is a parameter regime where the amount of kineticenergy, and thus the flow velocities, are sensitive to the prescribedvalue of m. Additionally, in this regime, the obliquity response maybe less important than the eccentricity response. Second, and of

Fig. 2. Average kinetic energy associated with each of the tidal responses for an ocean oncalculated from scaling arguments (Section 3 and Appendices E and F) and the solid linshaded region indicates where the Reynolds number is less than 50 for this model oceansolution value (see (29)). The dashed line represents the effective viscosity for the respontide, both of which are numerically-derived (see Section 4.2) using the same coefficient ofobliquity tide the likely dominant contributor to the ocean energy dissipation (see Sect

more importance for models of thermal evolution, the energy dis-sipation rate depends on the viscosity. Thus, determination of a sa-tellite’s thermal and orbital history hinges on the prescribed valuefor the viscosity, which is in reality very poorly constrained. Weaddress this issue in Section 4.

3. Scaling laws for tidally-driven ocean flows

The shallow-water system presented in Section 2 can be solvedusing a semi-analytic method with computational ease. The chal-lenge in understanding the ocean response to the tidal forcing isdetermining how the various input parameters affect the resulting

Europa with parameters from (Tables 1 and 2), where symbols indicate the valueses indicate values calculated through the semi-analytic solution (Section 2.2). Theand the scaling argument predictions are no longer within 10% of the semi-analyticse to the eccentricity tide and the dot-dashed line represents that for the obliquitydrag ðcD ¼ 0:002Þ. Note that these effective viscosities are not the same, making the

ion 5.2 for further details).

Page 6: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

Fig. 3. Similar to Fig. 2 but for average energy dissipation.

16 E.M.A. Chen et al. / Icarus 229 (2014) 11–30

flow and choosing an appropriate value for m or a. We address thefirst issue here and discuss an estimate for the effective ocean vis-cosity in Section 4.

The shallow water system presented in Section 2.2 involves ser-ies solutions that are coupled in spherical harmonic degree. How-ever, as can be seen in Fig. 1, the power decreases quite rapidlywith increasing spherical harmonic degree l. The tidal responses,previously represented as infinite sums or for numeric purposesas a truncated series, can be reasonably estimated by using a verylimited number of modes (typically 2–3) without any significantloss of accuracy. Instead of solving the system of equations givenby (21) and (22) numerically, we now solve the systemalgebraically.

The benefits of these solutions are multifold. First, the simplifi-cations we present make the underlying physics more transparentthan the fully coupled problem. Second, these solutions providerelations for the kinetic energy and energy dissipation in termsof the various input parameters (e.g. R; g;h;X), instead of thespherical harmonic coefficients. These relations provide insightinto the sensitivity of the solutions to the various input parametersthat themselves may have a degree of uncertainty. Additionally,these relations are simple to implement in orbital evolution mod-els and are likely accurate in most scenarios, similar to the familiarsolid-body tidal heating equations (Segatz et al., 1988; Wisdom,2008). These scalings are limited in that they do not address the is-sue that the ocean’s effective viscosity is essentially unconstrained,but we will use these scalings to inform the estimates of the viscos-ity that we present in Section 4.

In addition to the tidal forcing, the ocean responds to threeforces (cf. momentum Eq. (1)): pressure gradients in the form ofsurface gravity waves, the Coriolis force, and the drag. For the sca-lings we suggest below, we need the ratios of these forces. We adopta Navier–Stokes viscosity such that the role of viscosity can be com-pared to that of rotation through a rotational Reynolds number Re,

Re � XR2

m;

where the characteristic velocity is taken to be the linear rotationalspeed, not the typical ocean flow velocity, and the characteristiclength scale is R. We are also concerned with the square of the ratio

of the linear rotational speed to the surface gravity wave speed, �,called the Lamb parameter in Longuet-Higgins (1968),

� � 4X2R2

gh:

Below, we present scaling arguments which allow us to analyt-ically calculate the ocean response, the average kinetic energy andaverage energy dissipation rate in the case of Navier–Stokes typedissipation. The algebra in the following section can be tedious;we present scalings for the response to the westward propagatingobliquity tidal potential here and direct the reader to Appendices Eand F for details associated with the other components of the tide.For those uninterested in the details, we summarize all of the scal-ing results in Table 4. These estimates are accurate to 10% of thefull solutions from Section 2.2 if the following limits are satisfied,high Reynolds number,

Re 50; ð29Þ

and small Lamb parameter,

� 64

15: ð30Þ

For typical satellite oceans, where the rotation rate and satellite ra-dius are well-known, these limits should be thought of as low effec-tive viscosity (mK 106 m2=s) and thick oceans (h J 10 km),respectively. Because the molecular viscosity of water is very small,the scaling constraint (29) should be appropriate for all satelliteoceans. Earth lies in a high Lamb parameter regime because theoceans on Earth are far shallower than those presumed to exist onthe icy satellites (Khurana et al., 1998). In the small Lamb parameterregime, the time lag between the forcing potential and the ocean’sresponse is small because the gravity wave speed is fast in compar-ison to the surface rotational speed; as the tidal potential changes,the gravity waves can rapidly propagate this disturbance.

Under these limits, we now present the analysis for the re-sponse to the westward propagating obliquity tide. This is likelyto be the dominant ocean response for many satellite parameters(as recognized in Tyler (2008)) but for completeness, we also directthe reader to Appendices E and F for details related to the othertidal components. As discussed below, the obliquity tide involvesa resonant response which can drive large flows. Other resonant

Page 7: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

E.M.A. Chen et al. / Icarus 229 (2014) 11–30 17

modes exist, but typically require very thin oceans (Tyler, 2011).We therefore ignore these modes, because of the expected thick-nesses (tens of km) of most icy satellite oceans (see Section 5.5).

3.1. Scalings for the response to the westward obliquity tidal potential

Looking at the response to the westward obliquity tide (Fig. 1),the dominant component of the power spectrum is the l ¼ 1 term(i.e. W1

1). This response is coupled through the Coriolis force to U12

and the forcing tide. We thus carry out this scaling analysis withthe assumption that the response is accurately captured by onlythe l ¼ 1 and l ¼ 2 terms. The vorticity Eq. (21) for W1

1 here writtenin terms of the Reynolds number and Lamb parameter is

iþ 2Re� i

2ð1Þð1Þð2Þ

� �W1

1 � 232

C12U

12

� �¼ 0: ð31Þ

The first imaginary term in (31) represents the ocean responsewhile the second imaginary term represents a Rossby–Haurwitzwave. In this instance, these two exactly cancel, and we obtain

W11 ¼

3C12

2ReU1

2; ð32Þ

where the amplitude of the response grows as the viscosity de-creases, and Reynolds number increases; this is the resonance thatTyler (2008) refers to. For other tidal responses (see Appendices Eand F), the imaginary terms do not disappear and in fact, are muchlarger than the Reynolds number term. These terms limit the re-sponse amplitude and, therefore, the overall behavior of the otherresponses differ dramatically from that of the westward obliquityresponse.

The amplitude of W11 is related to U1

2 by the horizontal diver-gence Eq. (22)

i23þ 6

Re� i

24�

� �U1

2 þ32ðC1

2Þ2ReU1

2 �U1

2

Xð33Þ

when the relation (32) is substituted and the contribution from W13

is neglected. In the appropriate scaling limits (large Reynolds num-ber and small Lamb parameter), the dominant terms of (33) are

32ðC1

2Þ2Re� i

24�

� �U1

2 �U1

2;W

X: ð34Þ

The total kinetic energy (24) of the response is dominated by W11

(Fig. 1) and can be approximated as

Etot 2qh W11

�� ��2 ¼ 9ðC12Þ

2

2qhRe2 U1

2

�� ��2; ð35Þ

when the relation (32) is included. Recasting (34), we have

jU12j

2 jU1

2;W j2

X2j 32 ðC12Þ

2Re� i 24

� j2

¼4jU1

2;W j2

X2Re2 9 C12

� �4þ 2304 1

�Re

� 2� � : ð36Þ

Substituting (6) and (36) into (35), the kinetic energy of the re-sponse is given as

Etot 18ðC1

2Þ2qhjU1

2;W j2

X2 9ðC12Þ

4þ 2304 1

�Re

� 2� �

¼27 C1

2

� �2p

5qhX8R12h2

0

9 C12

� �4X6R8 þ 144m2g2h2

� � : ð37Þ

In similar fashion, the average energy dissipation (27) can beapproximated as

_E 2qhmR2 ð1Þ

2ð2Þ2 W11

�� ��2 ¼ 18 C12

� �2qhXRe U1

2

�� ��2: ð38Þ

Substituting (6) and (36) into (38), the energy dissipated by the re-sponse is given as

_E 72 C1

2

� �2qh U1

2;W

��� ���2XRe 9 C1

2

� �4þ 2304 1

�Re

� 2� �

¼108 C1

2

� �2p

5qhmX8R10h2

0

9 C12

� �4X6R8 þ 144m2g2h2

� � : ð39Þ

We may also use Eq. (37) to derive an average flow speed �u:

�u ¼3 C1

2

� �U1

2;W

��� ���ffiffiffiffipp

XR 9 C12

� �4þ 2304 1

�Re

� 2� �1=2

¼ffiffiffiffiffiffi2750

rX4R5h0

9 C12

� �4X6R8 þ 144m2g2h2

� �1=2 : ð40Þ

The scalings (37), (39) and (40) are the general form for highReynolds number and small Lamb parameter as defined in (29)and (30). While the scalings are cumbersome, these values are sim-ple to calculate in a numerical code. As mentioned previously, it isclear from Figs. 2 and 3 that the response to the westward propa-gating obliquity tide can be split into two behavioral regimes;which regime applies depends on the product of the Reynoldsnumber and the Lamb parameter, 1

�Re. We present further simplifi-cations for these two regimes in order to highlight the underlyingforce balances and behavioral physics.

For high Reynolds numbers 1�Re K 1

80

� , the Rossby–Haurwitz

wave response dominates the bulk behavior. Ignoring the gravitywave contribution in (34), U1

2 is

U12

�U12;W

X 32 ðC

12Þ

2Re

� � ¼ �5

ffiffiffiffiffiffiffi2p15

rmh0; ð41Þ

which in turn yields

W11 �U1

2;W

ðC12ÞX¼ �3

2

ffiffiffiffiffiffiffi2p3

rXR2h0: ð42Þ

This solution asymptotes to the inviscid solution presented in Tyler(2008). In this regime, the kinetic energy is independent ofviscosity,

Etot 2qh W11

�� ��2 ¼ 3pqhX2R4h20; ð43Þ

and we can derive an average flow speed �u of

�u ¼ 32

� �1=2

XRh0: ð44Þ

This makes intuitive sense: if viscosity is not impeding flow signif-icantly, the flow speed is controlled by the rate at which the tidalbulge is moving around the body. From (27) or (28), the associatedenergy dissipation scaling scales linearly with the viscosity

_E 12pqhmX2R2h20 ¼ 8qhm�u2: ð45Þ

For lower Reynolds numbers 1�Re J 1

80

� , while still remaining in

the scaling limits defined by (29) and (30), the gravity waves dom-inate the response. Then,

Page 8: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

18 E.M.A. Chen et al. / Icarus 229 (2014) 11–30

U12

U12;W

X i24�

� ¼ �i4

ffiffiffiffiffiffiffi2p15

rX3R4h0

ghð46Þ

and therefore

W11

32

C12U1

2;W Re

X i24�

� ¼ �i340

ffiffiffiffiffiffiffi2p3

rX4R6h0

mgh: ð47Þ

In this regime, the viscosity has altered the natural frequency of thesystem, and thus the response amplitudes no longer behave reso-nantly. The kinetic energy now depends on the value of the Rey-nolds number,

Etot 2qh W11

�� ��2 C12

� �2p

320qhX2R4h2

0ð�ReÞ2 ¼ 3p400

qX8R12h20

m2g2hð48Þ

with an associated average flow speed of

�u ¼ 3800

� �1=2 X4R5h0

mgh: ð49Þ

In this limit flow speeds decrease as viscosity increases. The associ-ated energy dissipation is now inversely proportional to theviscosity,

_E 3p100

qX8R10h20

mg2h: ð50Þ

The shape of the energy dissipation curve (Fig. 3) can be thought ofas the product of two effects: the bulk flow velocity and the value ofthe viscosity. This peak in dissipation rate is similar to that seen forMaxwell viscoelastic dissipation in solid bodies (Ross and Schubert,1989). For the high Reynolds number regime, the velocities are highbut the viscosity is too low to generate significant dissipation. In thelower Reynolds number regime, the viscosity is large and thus candissipate energy. But in analogy with forced-damped harmonicoscillators, the viscosity has moved the natural frequency of thesystem sufficiently away from the forcing frequency that the re-sponse no longer has large resonant amplitudes.

3.2. Effective tidal quality factor Q in a satellite ocean

Throughout this section, we have considered viscous effectsthrough a Navier–Stokes viscosity, primarily because this allowsfor parameterization in the form of the Reynolds number. Thesescalings can also be expressed in terms of an effective tidal qualityfactor Q, which can in turn be related to the linear drag coefficienta mentioned in Section 2. Q is defined as (Murray and Dermott,1999),

Q � 2pEtot

_E 2pX

� ¼ XEtot

_Eð51Þ

and is related to the linear drag coefficient a by (Tyler, 2011),

a ¼ X2Q

: ð52Þ

For the westward propagating obliquity tide, by substituting (43),(45), (48) and (50), we find that the tidal quality factor is simply

Q obl;W 14

XR2

m

!¼ 1

4Re ð53Þ

for all regimes. We emphasize that this is an effective Q, in attempt-ing to maintain consistency with the terminology used in Tyler(2011). This Q is different from that typically defined for solid-bodytidal heating (Peale et al., 1979; Murray and Dermott, 1999) in thatthe dissipation is referenced to the kinetic energy of the ocean, in-stead of the stored elastic energy due to the tide.

The effective Q of the Earth is known (Q ¼ 12; Murray and Der-mott (1999)) from lunar laser ranging, and is due mainly to oceandissipation. Unfortunately, the Earth’s ocean is in many ways quitedifferent to that on an icy satellite: it spins non-synchronously, ithas bottom topography that is comparable to the ocean thickness,and it is stably stratified in some regions (see Section 5.5). As a result,it is hard to use this particular observational constraint to derive aneffective Q for icy satellites. Instead we adopt an approach based onan empirically-derived friction coefficient, as described next.

4. Effective viscosity in a global ocean

The scalings presented in Section 3 provide a description of glo-bal ocean behavior; however, to estimate the energy dissipation inthe ocean, these scalings require a prescribed value for the effec-tive viscosity. This viscosity is unlikely to be as small as the molec-ular viscosity, presumably due to turbulence (Pope, 2000), but it isunclear how large an effective viscosity one should adopt (Lumband Aldridge, 1991). While there is potential to infer from mea-surements what the effective viscosity is for a global liquid layer,such as for the liquid core of Earth (Smylie, 1999; Mound and Buff-ett, 2007; Buffett and Christensen, 2007) or Mercury, it is unlikelythat we can do so for icy satellite oceans. Also, while these mea-surements can give some insight into the effects of turbulence,the effective viscosity is likely to be time-variable, changing withflow conditions and/or degree of turbulence (Brito et al., 2004).

Because of the high degree of variability amongst satellites ofthe outer Solar System both in the past and at present, it is unlikelythat a single value for the effective ocean viscosity can be broadlyapplied. Thus, we approach the question of what the effective vis-cosity in a satellite ocean is by analogy with the Earth’s oceans. OnEarth, there is a long history of estimating dissipation of tides dueto bottom boundary friction (Taylor, 1920; Jeffreys, 1921; Munkand MacDonald, 1960). While the drag coefficient cD (see Eq. (54)below) was originally empirically-based (Taylor, 1920), a value cD

of O (0.001) is still commonly used, even in more modern numer-ical ocean models (Jayne and Laurent, 2001; Egbert and Ray, 2001).Sohl et al. (1995) discuss what value of cD to apply to Titan’s ocean.Taking the weak dependence of cD on Reynolds number into ac-count, these authors conclude that values in the range 0.002–0.01 are likely appropriate for liquid water.

We incorporate a cD-based parameterization into the shallowwater model for a global ocean. This approach is undeniably simplis-tic. However, when given the alternative of choosing a value for theeffective viscosity where the uncertainty is approximately twelve or-ders of magnitude (Lumb and Aldridge, 1991), we favor our approach,although there are caveats. We address these in Section 5.5. While weadopt the value cD ¼ 0:002 in this work, the results and scalings thatwe present in Section 4.2 are general (and linear) for cD, and thus, per-mit the effect of uncertainties in cD to be readily explored.

4.1. Numerical method

The momentum equation for an ocean subject to tidal forcingand bottom boundary friction is (cf. (Jayne and Laurent, 2001))

@~u@tþ 2~X�~u ¼ �g~rg� ~rU � cD

hj~uj~u: ð54Þ

Similar to the process in Section 2.2, we can employ a Helmholtzdecomposition and write the divergence and the radial componentof the curl of (54) as

@

@t

� �r2 þ 2X

R2

@

@/

� �Uþ 2X cos hr2 � sin h

R2

@

@h

� �W

¼ �gr2g�r2U � cD

h~r � ðj~uj~uÞ

� �; ð55Þ

Page 9: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

E.M.A. Chen et al. / Icarus 229 (2014) 11–30 19

@

@t

� �r2 þ 2X

R2

@

@/

� �W� 2X cos hr2 � sin h

R2

@

@h

� �U

¼ � cD

hr � ð~r� ðj~uj~uÞÞ

� �: ð56Þ

While the solutions to (55) and (56) are likely oscillatory, the non-linearity of the dissipative term requires that we integrate the sys-tem numerically. We seek solutions of the form

Wðh;/; tÞ ¼XNL

l¼1

Xl

m¼1

Wml Ym

l þW�ml Y�m

l

¼ 2RXNL

l¼1

Xl

m¼1

Wml Ym

l

!; ð57Þ

and similarly for U and g. Here the coefficients Wml (and their equiv-

alents Uml , and gm

l ) are still complex but are now functions of time,and NL is the series truncation level. Plugging in (57) and its equiv-alents into (55), (56) and (18), and projecting the resulting equa-tions onto individual spherical harmonics, the system of equationsto solve is

@

@t� 2Xim

lðlþ 1Þ

� �Wm

l � 2Xl� 1

lCm

l Uml�1 þ

lþ 2lþ 1

Cmlþ1U

mlþ1

� �

¼ R2

lðlþ 1ÞcD

hr � ð~r� ðj~uj~uÞÞ

� �m

lð58Þ

@

@t� 2Xim

lðlþ 1Þ

� �Um

l þ 2Xl� 1

lCm

l Wml�1 þ

lþ 2lþ 1

Cmlþ1W

mlþ1

� �

¼ �ggml þ

R2

lðlþ 1ÞcD

h~r � ðj~uj~uÞ

� �m

l� 2U0

2 cosðXtÞdl;2dm;0

� 2U12;W cosðXtÞdl;2dm;1

� U22;W þ U2

2;E

� �cosðXtÞ þ U2

2;W � U22;E

� �i sinðXtÞÞ

h idl;2dm;2;

ð59Þ

@gml

@t� hlðlþ 1Þ

R2 Uml ¼ 0: ð60Þ

These equations are coupled in both spherical harmonic degree land order m through the dissipative terms on the right hand sideof Eqs. (58) and (59). We include all components of the tide in(59) because nonlinearity has the ability to couple different spher-ical harmonic orders, while previously these were separable. Be-cause we have not assumed wave solutions, we now have to solveexplicitly for the spherical coefficients of g, whereas we previouslyassumed the relationship (20) between g and W.

Eqs. (58)–(60) are nonlinear and thus cannot be solved with thesemi-analytic method previously employed in Section 2.2. Insteadof assuming wave solutions as in Section 2.2, we employ a numer-ical integration method where the time derivatives are explicitlytreated using the Adams–Bashforth method. We calculate the non-linear dissipative terms using a spectral-transform method. Thehorizontal velocities are calculated on a spherical grid which con-sists of 100 evenly spaced grid points in longitude and 50 Gauss–Lobatto nodes in colatitude. The frictional dissipation contributionsare calculated on this grid, and subsequently, transformed backinto their spherical harmonic decomposition. To reduce the re-quired number of spectral coefficients, we added an enhanced Na-vier–Stokes dissipative term of the form mlðlþ 1Þ=R2 withm ¼ 1� 108 m2=s to (58) and (59) for all coefficients with l = NL(the series truncation level) and l = NL � 1. We adopt a series trun-cation level of NL = 30. The enhanced dissipative terms reduce therequired truncation level by an order of magnitude without affect-ing the results presented here; for example, the amount of energy

dissipated for our model Europan ocean is approximately 109 Wtotal, and the amount dissipated from the enhanced dissipationterm is less than 0.1 W.

We have run a series of numerical integrations varying the se-ven input parameters (i.e. R; g;h;X; h0; e, and cD) to understandthe sensitivity of the system to each. These numerical integrationsare initialized to either zero or the semi-analytic solution; the stea-dy-state solution is independent of these values. We evolve theequations in time until they produce a constant solution for the ki-netic energy and energy dissipation when averaged over an orbitalperiod, similar to the behavior of the linear solution (cf. (24) and(27)).

4.2. Scaling arguments

The shallow water system including nonlinear bottom drag set-tles on a wave solution. While the dissipation processes are not ex-actly alike, we can assume that the previous linear Navier–Stokesformulation is equivalent, in a time-averaged sense, to the nonlin-ear bottom drag with

� cD

hj~uj~u � mr2~u: ð61Þ

which implies that

m � cDR2j~ujlðlþ 1Þh ; ð62Þ

where the lengthscale associated with the Laplacian is taken to beR=

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffilðlþ 1Þ

p. Assuming that j~uj scales with the average velocity, we

find the scalings

meff ;obl ¼ koblcD

hX4R7h0

ð9ðC12Þ

4X6R8 þ 144m2

eff ;oblg2h2Þ

1=2 ð63Þ

by using the relation (40) for the obliquity-related response and

meff ;ecc ¼ kecccDX3R5e

gh2 ð64Þ

by using (F.10), (F.20) and (F.29) for the eccentricity response. Fromour numerical runs, the least squares fit for kobl is 0.40 and kecc is0.13.

The value of the scaling parameters in Eqs. (63) and (64) may beapproximated analytically by taking the flow speeds from Table 4and substituting into Eq. (62). For the eccentricity tide, this yieldskecc

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi57=112

p=lðlþ 1Þ 0:12, taking l ¼ 2. For the obliquity tide,

we obtain kobl ffiffiffiffiffiffiffiffiffiffiffiffiffiffi27=50

p=lðlþ 1Þ 0:37, taking l ¼ 1. These analyt-

ical approximations agree very well with the numerical results.The effective viscosity for obliquity (63) can be solved quadrat-

ically (see Table 4), and displays simpler behavior in the low andhigh Reynolds number limits discussed in Section 3.1. For a modelocean on Europa, the numerically-derived effective viscosity ismeff ;obl ¼ 3:3� 103 m2=s for the obliquity-related flow andmeff ;ecc ¼ 1:7� 102 m2=s for the eccentricity-related flow.

4.3. Summary

The numerically derived scalings (63) and (64) provide esti-mates for the effective viscosity given a set of input parameters.These viscosities can be input into either the semi-analytic system(Section 2.2) or the scaling laws (see Table 4) in order to determinethe flow response and energy dissipation rate of a satellite ocean. Itis worth noting that the effective viscosities for the obliquity andeccentricity responses scale differently with the input parametersbecause the velocities scale differently between the two responses.These values for a model Europan ocean are displayed in Figs. 2 and

Page 10: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

20 E.M.A. Chen et al. / Icarus 229 (2014) 11–30

3 for reference. In the context of thermal–orbital evolution models,where the input parameters change over time, the scalings for theviscosities are critical to better estimating the long-term impact ofocean tidal heating.

5. Implications for the outer Solar System satellites

With estimates of effective viscosity in hand, it is now possibleto complete a survey of the thermal heat budgets for various satel-lites of the outer Solar System. Specifically, we seek to answer twofairly broad questions: should ocean tidal heating be an importantheat source at present for any of the icy satellites; and could oceantidal heating have been a significant heat source in the past? Thesequestions are difficult to answer in that there are many uncertain-ties in the parameters that go into the following calculations. Be-cause our goal was to understand broad trends and not toexplain specific observations or quantities, we have used simplifiedtheoretical models in order to calculate the thermal budgets, withthe understanding that these results may diverge from what is ob-served, as for example in the long-recognized anomalous cases ofMimas or Enceladus (Squyres et al., 1983; Meyer and Wisdom,2007; Howett et al., 2011).

5.1. Calculating the available thermal energy

The primary heat sources available to a satellite are radiogenicheat and tidal dissipation, either in the solid ice and rock layers orin a liquid ocean layer. For the radiogenic component of the ther-mal budget, we assume a chondritic heating rate ofH ¼ 4:5� 10�12 W kg�1 (Ellsworth and Schubert, 1983). We calcu-late an equivalent silicate mass, Msil, assuming a two-componentsatellite

Msil ¼43qsil

qsat � qice

qsil � qice

� �pR3 ð65Þ

where the silicate density is taken as qsil ¼ 3500 kg m�3, the icedensity is taken as qice ¼ 950 kg m�3, and the satellite mean densityqsat as presented for various satellites in Table 2.

To estimate the amount of solid body tidal heating, we employthe standard viscoelastic model (Segatz et al., 1988; Ross and Schu-bert, 1989; Wisdom, 2004). In this model, the rate of solid bodydissipation is given by

_E ¼ 32

k2

Q sol

ðXRÞ5

Gð7e2 þ sin2 h0Þ; ð66Þ

where the satellite structure and dissipation characteristics areparameterized by the tidal Love number, k2, and the tidal qualityfactor, Qsol, and G is the gravitational constant. For the unknown sa-tellite parameters, k2 and Qsol, we adopt a value of Qsol of 100(Goldreich and Soter, 1966), and we estimate the Love numberusing Kelvin’s formula (Love, 1944),

k2 ¼3=2

1þ 19l2qsat gR

ð67Þ

for a homogeneous body with rigidity, l, assumed to bel ¼ 4:0� 109 N m�2. We also calculate the dissipation with ahomogeneous fluid k2 of 3=2 to understand the energy budget inthe limit of a fully deformable body, such as one with an ocean.

In previous works, the dissipation related to the obliquity tidehas typically been neglected, primarily because it is assumed thatthe obliquity is small and also, until recently, there have been nomeasurements of satellite obliquity. While for most satellites theobliquity solid body tidal dissipation is negligible when comparedto the eccentricity driven solid body dissipation, this relationship isreversed in the case of ocean tidal heating.

Titan is currently the only icy satellite with a measured obliq-uity (Stiles et al., 2008). In order to estimate the contributions toother satellites from the obliquity tide, we have assumed obliqui-ties appropriate for satellites in damped Cassini states with a singleorbit precession frequency; interactions with other satellites donot generally change the results significantly (Bills, 2005; Balandet al., 2012). In a Cassini state, the invariable pole, orbit normal,and spin pole remain coplanar as they precess (Colombo, 1966;Peale, 1969). For this coplanar precession to occur, the obliquitymust satisfy the relation (Peale, 1969; Ward, 1975; Bills and Nim-mo, 2008)

32½ðJ2 þ C2;2Þ cos h0 þ C2;2�p sin h0 ¼ c0 sinði� h0Þ ð68Þ

where J2 and C2;2 are the degree-2 gravity coefficients, i is the orbitalinclination, and p is the ratio of the orbital motion to the orbit planeprecession and is given by

p ¼ XdXorb=dt

ð69Þ

where Xorb is the longitude of the ascending node. The quantity c0 isthe normalized polar moment of inertia, with the prime denotingthe fact that the relevant moment of inertia may be that of a decou-pled shell rather than the entire body. We introduce this distinctionbecause of the case of Titan, where the moment of inertia c esti-mated from gravity measurements (see below) does not equal thatinferred from Titan’s obliquity (Bills and Nimmo, 2011).

While the orbital parameters are well-established, observationsof the gravity coefficients and moment of inertia for many satel-lites are limited. We estimate the unmeasured gravity coefficientsassuming a hydrostatic satellite using the Darwin–Radau relation(Darwin, 1899; Radau, 1885; Hubbard and Anderson, 1978)

C2;2 ¼14

X2

G 43 pqsat

� !

5

1þ 52� 15

4 c� 2 � 1

!ð70Þ

where c denotes the normalized polar moment of inertia of the en-tire body. For a hydrostatic body J2 ¼ 10C2;2=3.

For the ocean contributions to tidal heating, we estimate theeffective viscosities from (63) and (64) for the obliquity and eccen-tricity tides separately. We use values of h = 30 km and cD ¼ 0:002for all satellites. After calculating a satellite-specific effective vis-cosity, the associated tidal dissipation is calculated from the ana-lytical scalings summarized in Table 4; the semi-analyticapproach from Section 2.2 produces nearly identical results.

Table 1 summarizes all of the common parameters used in thisanalysis, while Table 2 shows the parameters used and calculatedquantities that are satellite specific. In preparing this Table, wetook c ¼ c0 ¼ 0:33, representative of a differentiated body with noshell decoupling. Increasing c0 to 0.67 (a decoupled shell) whilekeeping c ¼ 0:33 results in an increase in the predicted obliquityh0 by a factor of roughly 2, and the obliquity-related heating by afactor of roughly 4.

The various contributions to the thermal energy budget are pre-sented for each satellite in Table 3. In Figs. 4–6 we highlight key re-sults that are tabulated in Table 3. In Fig. 4, we present the relativecontributions to the ocean tidal dissipation from the eccentricityand obliquity tides. While the obliquity tidal response likely dom-inates ocean dissipation, there are examples of satellites where theeccentricity response could be more important. The full thermalbudgets of many satellites are presented in Figs. 5 and 6 wherethe solid-body tidal dissipation is calculated using illustrative val-ues of Qsol ¼ 100 and k2 corresponding to an elastic homogenousbody or a fully fluid body, respectively. The actual values for theseare likely to depend on the details of internal structure andrheology (Ross and Schubert, 1989), and in particular whether a

Page 11: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

Table 3Energy contributions calculated from parameters presented in Tables 1 and 2 as described in Section 5.1. Entries marked [] in final column have � > 4=15 for the nominalparameters.

Homogeneous k2 k2 ¼ 3=2

_EradðWÞ _Esolid;eccðWÞ _Esolid;oblðWÞ _Esolid;eccðWÞ _Esolid;oblðWÞ _Eocean;eccðWÞ _Eocean;oblðWÞ

Europa 2.02 � 1011 9.77 � 1011 1.34 � 109 7.05 � 1012 9.70 � 109 1.47 � 107 3.1 � 109

Ganymede 4.67 � 1011 8.79 � 109 2.49 � 108 5.46 � 1010 1.55 � 109 1.66 � 104 1.3 � 109

Callisto 3.20 � 1011 2.08 � 109 9.30 � 107 1.65 � 1010 7.40 � 108 8.80 � 102 3.8 � 109

Mimas 4.02 � 107 2.90 � 108 5.46 � 104 7.59 � 1011 1.40 � 108 2.38 � 107 [2.5 � 106]Enceladus 2.73 � 108 2.37 � 107 1.04 � 100 1.96 � 1010 8.64 � 102 2.32 � 104 1.2 � 10�1

Tethys 1.34 � 108 1.70 � 105 1.11 � 106 8.45 � 107 5.51 � 108 2.57 � 101 [3.7 � 107]Dione 2.42 � 109 4.25 � 107 1.52 � 103 8.44 � 109 3.02 � 105 4.46 � 103 2.2 � 103

Rhea 3.31 � 109 1.73 � 105 1.79 � 105 2.65 � 107 2.75 � 107 7.35 � 10�1 7.5 � 106

Titan 4.11 � 1011 6.43 � 1010 3.46 � 108 4.37 � 1011 2.35 � 109 1.23 � 105 1.1 � 1010

Miranda 8.50 � 107 6.17 � 105 7.10 � 103 1.05 � 109 1.21 � 107 6.37 � 102 2.6 � 105

Ariel 3.59 � 109 3.00 � 107 2.31 � 102 4.42 � 109 3.40 � 104 1.36 � 103 5.1 � 101

Umbriel 2.32 � 109 2.00 � 107 4.07 � 103 4.08 � 109 8.29 � 104 9.69 � 102 1.8 � 103

Titania 9.71 � 109 4.71 � 105 3.27 � 103 3.55 � 107 2.46 � 105 1.63 � 10�1 1.3 � 105

Oberon 7.76 � 109 6.10 � 104 7.81 � 103 5.44 � 106 6.98 � 105 5.82 � 10�3 3.3 � 106

Triton 7.12 � 1010 0 8.70 � 108 0 1.60 � 1010 0 6.6 � 1010

Fig. 4. Energy dissipated in various satellite oceans due to the obliquity andeccentricity tides calculated from the scalings summarized in Table 4. Theparameters used are h = 30 km, cD ¼ 0:002, and satellite parameters from Table 2.For satellites lying to the right of the dotted line, the obliquity response is thedominant contribution to dissipation. Here for plotting purposes, Triton has energydissipation associated with the eccentricity, but in reality, this value should be zero.

Fig. 5. Thermal heat budget for various outer Solar System satellites; solid bodytidal heating components are calculated assuming a homogeneous elastic structure(Eq. (67)). Details of the calculations are presented in Section 5.1, while thenumerical quantities are presented in Table 3. Pie charts indicate relativecontributions from radiogenic heating (blue), solid body eccentricity tidal heating(green), solid body obliquity tidal heating (yellow), ocean obliquity tidal heating(red), and ocean eccentricity tidal heating (gray). The size of the pie charts arescaled by logðRÞ. (For interpretation of the references to colour in this figure legend,the reader is referred to the web version of this article.)

Fig. 6. Similar to Fig. 5, here with the solid body tidal heating componentscalculated assuming a completely fluid response ðk2 ¼ 3=2Þ.

E.M.A. Chen et al. / Icarus 229 (2014) 11–30 21

decoupling ocean exists. In choosing the fluid and elastic limits, welikely bracket the actual value of the tidal dissipation, while pro-viding insight into the degree of uncertainty associated with the ti-dal model. It is clear that for most of the satellites the thermalbudgets are dominated by two components: radiogenic heatingand eccentricity solid-body tidal heating. Triton and Tethys arelikely the only satellites with large present-day heat contributionsfrom obliquity tides, because of their relatively small eccentricities.Triton might be primarily heated by ocean obliquity tides, whileCallisto and Titan could experience minor ocean heating.

5.2. Relative contributions from eccentricity and obliquity

While Tyler (2008, 2011) recognized that a satellite ocean re-sponse is typically dominated by the westward obliquity response,this is not always the case. Fig. 4 shows the energy dissipated in anocean calculated with the input parameters h = 30 km, cD ¼ 0:002,and the satellite-appropriate parameters from Table 2. For most sa-tellite oceans, the obliquity tidal heating contribution is larger thanthat from the eccentricity. The reason for this is twofold. First,

assuming the obliquity response is in the low viscosity (nearlyinviscid) regime, the average velocity will be much higher for theobliquity related flow due to the resonant response (Tyler, 2008),

Page 12: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

22 E.M.A. Chen et al. / Icarus 229 (2014) 11–30

if the obliquity and eccentricity are roughly comparable. Second,because higher velocities equate to higher effective viscosities(see Section 4.2), dissipation is further enhanced in the obliquityresponse.

Taking the scalings for the energy dissipation from Section 3.1and Appendix F (summarized in Table 4) and the numerical resultsfrom Section 4.2, it can be shown, here in the low viscosity regime,that the obliquity likely dominates unless there is a significant dis-parity between the eccentricity and obliquity:

_Eobl

_Eecc

44;800519

1�

� �3 kobl

kecc

� �h0

e

� �3

: ð71Þ

Recall that for these scalings to be applicable, the Lamb parameter,�, must be low (30), and additionally, kobl

keccis approximately 3 (cf. Sec-

tion 4.2). Oceanic energy dissipation is clearly biased towards theobliquity response. Because of this, oceans on satellites with loworbital inclination, and thus, low Cassini state obliquity, are unlikelyto experience significant dissipation (Chen and Nimmo, 2011).

5.3. Global tidal heating pattern

Thus far we have primarily discussed the time- and spatially-averaged energy dissipation rate. While this is an important quan-tity when thinking about the overall thermal and orbital evolution,the spatial distribution of tidal heat can also have implications forspacecraft observables. As suggested for solid-body tidal heating,spatial variations in the heat flux can generate lateral variationsin conductive ice shell thickness (Ojakangas and Stevenson,1989). These variations can potentially generate long-wavelengthtopography that can be detected through topography or gravity(Schenk and McKinnon, 2009; Nimmo and Bills, 2010). In addition,the spatial pattern of the heating at the base can affect convectionin the ice shell (Tobie et al., 2005; Roberts and Nimmo, 2008). Theeffects of this convection may potentially manifest as geologic fea-tures on the surface (Tackley et al., 2001; Besserer et al., 2013).

We present the time-averaged global heating patterns in theocean for the obliquity response and eccentricity response inFig. 7. The primary power for both responses occurs in degreesl = 0, 2, and 4, as would be expected from tidally-driven flow. How-ever the relative power in each degree and thus the location of hotand cold spots differ from the case for solid-body tidal heating(Ojakangas and Stevenson, 1989; Beuthe, 2013). In particular,ocean obliquity heating is maximized at the pole (as is solid-bodyeccentricity heating in a thin shell), while ocean eccentricity heat-ing, although much smaller, is maximized at the equator. Obliquitytidal heating will thus tend to reinforce shell thickness variationsarising because of solid-body eccentricity heating.

Table 4Summary of scalings presented in Section 3 and Appendices E and F. Here we define: u0 ¼ ðXfull semi-analytic solution of Section 2.2 when Re ¼ XR2

m > 50 and � ¼ 4X2 R2

gh K 415. The value

Tidal component Average velocity Kinetic energy

Obliquity-westward, 1�Re� 1

80

� ffiffi32

qXRh0

3pqhX2R4h20

Obliquity-westward, 1�Re J 1

80

� ffiffiffiffiffiffi3

800

qu0Reh0

3p400 E0Re2h2

0

Obliquity-eastwardffiffiffiffiffiffiffiffiffiffi2397

54880

qu0h0

2397p27;440 E0h2

0

Obliquity-generalffiffi32

qXRh0

1þ 400m2=u20 R2½ �ð Þ1=2

3pqhX2R4h20

ð1þ½400m2=u20 R2 �Þ

Eccentricity-radialffiffiffiffiffiffi57

560

qu0e

57p280 E0e2

Eccentricity-westward, librationalffiffiffiffiffiffi

3224

qu0e

3p112 E0e2

Eccentricity-eastward, librationalffiffiffiffiffiffi63

160

qu0e

63p80 E0e2

Eccentricity-total – 57p56 E0e2

5.4. Triton

The thermal history of Triton is a puzzle; if satellite ocean tidalheating is significant for any icy satellite, it is most likely to be Tri-ton. Observations of the surface features suggest that Triton hasbeen quite active in the past (Smith, 1989; Croft et al., 1995;McKinnon and Kirk, 2007), and this activity could have been as re-cent as 10 Ma (Schenk and Zahnle, 2007). Eccentricity tides havebeen evoked for heating mechanisms like frictional shear heating(Prockter et al., 2005) or ‘‘thermal blanketing’’ (Gaeman et al.,2012), but because the eccentricity at present is zero, there is noclear reason why geologic activity should be recent. A significantamount of tidal heating likely occurred during Triton’s captureand subsequent orbital circularization (Ross and Schubert, 1990);however, the likelihood that Triton’s capture is recent is very low(Agnor and Hamilton, 2006; Schenk and Zahnle, 2007) and addi-tionally, the circularization timescale, while dependent on theinterior structure, is likely to be less than 1 Gyr (Ross and Schubert,1990).

Prior to observations from Voyager 2, Jankowski et al. (1989)suggested that solid-body obliquity tidal heating could play a sig-nificant role in the thermal history of Triton, specifically if Tritonoccupied Cassini state 2. From imaging it was shown that this statewas unlikely (Smith, 1989); however, the role of the obliquity tideon the evolution of Triton should not be discounted. Ultimately, theobliquity is related to the inclination of the orbit (Peale, 1969), andunlike eccentricity, the inclination is not damped rapidly (Murrayand Dermott, 1999). It was recognized that the solid-body tidalheating contribution is small (Jankowski et al., 1989) if Triton isin Cassini state 1; however, the ocean can effectively contributeheat at a rate of approximately 66 GW if this is the case – compa-rable to radiogenic heat production (Figs. 5 and 6). Because theinclination has remained fairly constant, ocean tidal heating couldhave been involved throughout the thermal history of Triton, spe-cifically aiding in keeping an ocean from freezing and providing amoderate heat source for recent geologic activity. Coupled ther-mal–orbital models that include ocean tidal heating could providefurther insight into the long term evolution of Triton.

5.5. Caveats

In estimating the thermal budgets for the outer Solar Systemsatellites (Figs. 5 and 6), we have used fairly simplistic models inorder to calculate the amount of tidal dissipation both in the so-lid-body and ocean. These models are sufficient to capture andhighlight broad trends. Given the general lack of data and largeuncertainty in physical parameters, we do not attempt to carry

3R3Þ=ðghÞ; E0 ¼ qX6R8=ðg2hÞ and _E0 ¼ E0m=R2. These scalings are accurate to 10% of thes presented for the coefficients in m are numerically obtained (see Section 4.2).

_E m Q

12pqhX2R2mh20

– 14 Re

3p100

_E0Re2h20

9237p6860

_E0h20

– 12;316799 Re

12pqhX2R2mh20

ð1þ½400m2=u20 R2 �Þ X3R4

20ffiffi2p

gh�1þ

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1þ 200

30:40cD gh0

X2R

h i2r !1=2 1

4 Re

39p14

_Eoe2 19260 Re

69p140

_Eoe2 – 592 Re

231p20

_E0e2 - 344 Re

519p35

_E0e2 0:13 cDX3R5 egh2

951384 Re

Page 13: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

Fig. 7. Time-averaged energy dissipation pattern associated with (a) obliquity tidal response and (b) eccentricity tidal response for a model ocean on Europa. These arecalculated from the numerical model presented in Section 4 with cD ¼ 0:002 and h = 30 km.

E.M.A. Chen et al. / Icarus 229 (2014) 11–30 23

out models of specific bodies in this work. With regard to the solidbody dissipation calculated here, it is recognized that the Maxwellviscoelastic model is insufficient to explain the tidal dissipation ob-served, for example for Enceladus (Meyer and Wisdom, 2007;Howett et al., 2011).

In adopting a simple model for the ocean behavior by analogy toEarth’s oceans, we have neglected a significant dissipation mecha-nism, internal tides, which arise because of compositional stratifi-cation (Munk and MacDonald, 1960; Egbert and Ray, 2000). WhileEuropa’s ocean, at least, is electrically conductive, presumablydue to brine (Khurana et al., 1998; Kargel et al., 2000), it is unclearwhether satellite oceans are compositionally stratified (Goodmanet al., 2004). The major drivers of compositional variations inEarth’s oceans, such as evaporation or addition of fresh water, arelikely much less important on icy satellites. Because internal tidegeneration is largely due to ocean bathymetry (Baines, 1973; Jayneand Laurent, 2001), a property which is completely unconstrainedin the context of satellite oceans, we have ignored the effects of astratified ocean. Additionally, our shallow-water approach cannotcapture the effects of convection. Convection will alter the dynam-ics of the ocean flow to some extent (Chen et al., 2010; Soderlundet al., 2013). However, the magnitude of the tidally-driven veloci-ties are likely comparable to those generated through convectionand also, convection has a tendency to destroy stratification. Thus,the shallow-water model may be sufficient to predict the bulk sa-tellite ocean behavior. Unfortunately, due to computational limita-tions, three-dimensional models that can include convection maynot necessarily be more accurate than our shallow-water model,because these models are unable to resolve the processes that pre-sumably dissipate energy in the system (Glatzmaier, 2002).

As discussed in Section 4, the appropriate value of the drag coef-ficient cD is probably in the range 0.002–0.01. In using the lowerbound, we are being conservative in our estimate of ocean heating.Similarly, our approach to calculating the obliquity h0 might be toolow by a factor of 2 if decoupled shells are present (Section 5.1).Taken together, our ocean heating estimates could be low by atmost a factor of 20. In most cases, this uncertainty would notchange our conclusions – Mimas and Europa would join the exist-ing candidates (Triton, Titan and Callisto) as possibly having expe-rienced significant ocean obliquity heating.

Throughout this work, we assume a small Lamb parameter; theuncertainty in the Lamb parameter arises from the uncertainty inthe ocean depth. The scalings we present in Section 3 are likelynot applicable if the ocean depth is less than �10 km. However,if this is the case, the assumption of uniform ocean thickness in

Eqs. (1) and (2) may also not be valid. Long-wavelength topogra-phy of either the ice shell above the ocean or the silicate mantle be-low the ocean can have amplitudes on the kilometer scale (Schenkand McKinnon, 2009; Nimmo and Bills, 2010). The dynamics of theocean may be far more complex in the case of a very shallow ocean,and the shallow-water scalings we have described do not capturethis complexity.

6. Conclusion

It was recognized by Tyler (2008, 2009, 2011) that tidal dissipa-tion in a global ocean could provide a heat source that had not beenpreviously recognized; however, the amount of dissipation couldvary over orders of magnitude depending on the value of the effec-tive viscosity (parameterized as either m;a, or Q) implemented inthe model. We have expanded that work in two ways. First, weprovide scalings for the tidal responses, specifically the average ki-netic energy and energy dissipation, in terms of the model inputparameters. These can be easily implemented into thermal–orbitalevolution models for many satellites, similar to relations used forsolid-body tidal dissipation. Second and more importantly, we esti-mate the effective viscosity of a satellite global ocean based onanalogy to frictional dissipation in oceans on Earth. Both of theseresults are summarized in Table 4.

Based on these scaling relations and estimates for the effectiveviscosity, we find that for almost all satellites the thermal budget isdominated by the radiogenic and solid-body eccentricity tidalheating, as long recognized. For completeness, evolution modelscan include ocean tidal dissipation; however, this is unlikely to cre-ate large deviations from previous results, unless high obliquitiesexisted in the past. In the special case of Triton, ocean tidal heatingmay play a large role in keeping a primordial ocean warm, primar-ily because eccentricity tidal heating, either in the ocean or solid-body, is likely small after rapid circularization of the orbit post-capture. Because the inclination evolves on a much longer time-scale than the eccentricity, obliquity tidal heating in the oceancan be significant even up to the present and may provide a heatsource for the apparent recent geologic activity on the surface.

Acknowledgments

We thank Mikael Beuthe and Isamu Matsuyama for theirthoughtful and thorough reviews. Chris Edwards and Doug Lin pro-vided helpful insights and discussions. This work was supported by

Page 14: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

24 E.M.A. Chen et al. / Icarus 229 (2014) 11–30

NASA’s Planetary Geology and Geophysics Program(NNZ09AD89G) and a UCSC Dissertation-Year Fellowship.

Appendix A. Useful spherical harmonic relations

To obtain equations for the spherical harmonic coefficients Wml

and Uml (e.g. Eqs. (21) and (22)), we used the following spherical

harmonic relations:

Y�ml ¼ ð�1ÞmYm�

l ; ðA:1Þ

r2Yml ¼�lðlþ 1Þ

R2 Yml ; ðA:2Þ

cos hYml ¼ Cm

lþ1Ymlþ1 þ Cm

l Yml�1; ðA:3Þ

sin h@Ym

l

@h¼ lCm

lþ1Ymlþ1 � ðlþ 1ÞCm

l Yml�1: ðA:4Þ

These are similar to Eq. (3.17) of Longuet-Higgins (1968) and (23)–(25) of Tyler (2011) for normalized Laplace spherical harmonics. Weuse the convention that W�m

l ¼ ð�1ÞmWm�l and likewise with g�m

l

and U�ml . In addition, a useful relation when calculating spatially

averaged quantities, such as the kinetic energy, is

Z 2p

0

Z p

0

@Yml

@h@Ym0�

l0

@hþ 1

sin2 h

@Yml

@/@Ym0�

l0

@/

" #sin hdhd/

¼ lðlþ 1Þdl;l0dm;m0 : ðA:5Þ

Appendix B. Response to eastward propagating tidal potentials

In the linear system, we can separately solve for the response tothe eastward propagating potentials. In this case, we seek wavesolutions of the form:

Wðh;/; tÞ ¼Xl

m¼1

X1l¼1

Wml e�iXtYm

l þW�ml eiXtY�m

l

¼ 2RXl

m¼1

X1l¼1

Wml e�iXtYm

l

!ðB:1Þ

and similarly for U and g.The resulting coupled equations from substituting (B.1) and

similar expressions for U and g into (16)–(18) are

iX� a� mlðlþ 1ÞR2 þ 2Xim

lðlþ 1Þ

� �Wm

l

þ 2Xl� 1

lCm

l Uml�1 þ

lþ 2lþ 1

Cmlþ1U

mlþ1

� �¼ 0; ðB:2Þ

�iXþ aþ mlðlþ 1ÞR2 � 2Xim

lðlþ 1Þ þighlðlþ 1Þ

XR2

� �Um

l

þ 2Xl� 1

lCm

l Wml�1 þ

lþ 2lþ 1

Cmlþ1W

mlþ1

� �

¼ �U12;Edl;2dm;1 � U2

2;Edl;2dm;2 ðB:3Þ

with the terms on the right hand side of (B.3) corresponding to theeastward obliquity tidal potential and eastward librational eccen-tricity tidal potential.

Appendix C. Response to the radial eccentricity tide

For the prior wave solutions (e.g. (19) and its equivalents or(B.1)), we have not included the m ¼ 0 terms. The solutions forW;U, and g must be real, and because the spherical harmonicY0

l is real, the corresponding coefficients W0l ;U

0l , and g0

l must bereal. Because the linear system is separable in degree m, this onlyapplies specifically for the radial component of the eccentricityforcing. To solve for this response, we can break the tidal poten-tial (8) into a westward and eastward potential in a manner sim-ilar to that for the obliquity tide and librational component of theeccentricity tide with a constraint that the total response must bereal, i.e.

Wðh;/; tÞ ¼X1l¼1

W0l;W eiXt þW0

l;Ee�iXth i

Y0l 2 R; ðC:1Þ

and similarly for U and g. Here the coefficients are still complex andthe subscripts W and E denoting the westward and eastward re-sponse, respectively. It can be shown that for (C.1) to be true, therelation

W0l;W ¼ W0

l;E

� ��; ðC:2Þ

must be true, and similarly for U and g. We can solve either forthe westward or eastward response, by substituting �U0

2dl;2dm;0

for the right hand side of either (22) or (B.3) with the understand-ing that

Wðh;/; tÞ ¼X1l¼1

W0l;W eiXt þW0

l;Ee�iXth i

Y0l

¼ 2X1l¼1

R W0l;W eiXt

� �Y0

l ¼ 2X1l¼1

R W0l;Ee�iXt

� �Y0

l ðC:3Þ

and likewise for U and g.

Appendix D. Derived quantities

For ease of solution, the fluid dynamic equations are expressedin terms of the streamfunction W and scalar potential U. Howeverthe quantities of interest are those such as kinetic energy or energydissipation rate. Here we explicitly present the process to movefrom the spherical harmonic coefficients associated with the solu-tion of the fluid dynamic equations to physical quantities.

D.1. Kinetic energy

Once the coefficients W and U have been determined for theocean response, the kinetic energy of the ocean flow, Etot , can becalculated from

Etot ¼q2

ZV

~u �~u dV : ðD:1Þ

We ignore the potential energy contribution to the total energy be-cause it is negligible when compared to this. The kinetic energy ex-pressed in terms of W and U is

Etot ¼q

2R2

ZV

@U@hþ 1

sin h@W@/

� �2

þ 1sin h

@U@/� @W@h

� �2" #

dV : ðD:2Þ

Because the energy is a nonlinear product, we must include boththe westward and eastward responses, here denoted with the sub-scripts W and E, respectively. Expanding (D.2) in terms of the wavesolutions from (19), its equivalents and (B.1) results in

Page 15: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

E.M.A. Chen et al. / Icarus 229 (2014) 11–30 25

Etot¼qh2

Z 2p

0

Z p

0

Xl

m¼1

X1l¼1

Uml;W eiXt @Ym

l

@hþU�m

l;W e�iXt @Y�ml

@h

�"(

þUml;Ee�iXt @Ym

l

@hþU�m

l;E eiXt @Y�ml

@hþWm

l;W eiXt 1sinh

@Yml

@/

þW�ml;W e�iXt 1

sinh@Y�m

l

@/þWm

l;Ee�iXt 1sinh

@Yml

@/þW�m

l;E eiXt 1sinh

@Y�ml

@/

��2

þXl

m¼1

X1l¼1

Uml;W eiXt 1

sinh@Ym

l

@/þU�m

l;W e�iXt 1sinh

@Y�ml

@/

�"

þUml;Ee�iXt 1

sinh@Ym

l

@/þU�m

l;E eiXt 1sinh

@Y�ml

@/�Wm

l;W eiXt @Yml

@h

�W�ml;W e�iXt @Y�m

l

@h�Wm

l;Ee�iXt @Yml

@h�W�m

l;E eiXt @Y�ml

@h

��2)

sinhdhd/:ðD:3Þ

By orthogonality of the spherical harmonics (5), the non-zero termsare

Etot¼qhZ 2p

0

Z p

0

Xl

m¼1

X1l¼1

Uml;WU�m

l;W þUml;EU

�ml;E þWm

l;WW�ml;W þWm

l;EW�ml;E

�h(

þ Uml;WU�m

l;E þWml;WW�m

l;E

� �ei2Xtþ Um

l;EU�ml;W þWm

l;EW�ml;W

� �e�i2Xt

�� @Ym

l

@h@Y�m

l

@hþ 1

sin2 h

@Yml

@/@Y�m

l

@/

� ���sinhdhd/:ðD:4Þ

Substituting (A.5) into (D.4) and averaging over an orbital period,we find that the average kinetic energy is

Etot;avg ¼qhX1l¼1

Xl

m¼1

lðlþ1Þ Uml;W

��� ���2þ Uml;E

��� ���2þ Wml;W

��� ���2þ Wml;E

��� ���2� �; ðD:5Þ

where the time-dependent components of (D.4) integrate to zero.This expression makes physical sense: the kinetic energy of the flowdepends on the amplitude of the tidal response, but not on thephase.

While (D.1) is nonlinear, the solution is separable in m. For easein presentation, we calculate the m ¼ 0 component separately. Thekinetic energy Eq. (D.2) can be expanded further in terms of the ra-dial response from (C.3) resulting in

E0tot ¼

qh2

Z 2p

0

Z p

0

X1l¼1

U0l;W eiXt þU0�

l;W e�iXt� � @Y0

l

@h

" #28<:

þ �X1l¼1

W0l;W eiXt þW0�

l;W e�iXt� � @Y0

l

@h

" #29=; sin hdhd/: ðD:6Þ

From the relation (A.5), the non-trivial terms of (D.6) are

Erad;tot ¼qh2

X1l¼1

ðlÞðlþ 1Þ

U0l;W eiXt þU0�

l;W e�iXt� �2

þ W0l;W eiXt þW0�

l;W e�iXt� �2

� �ðD:7Þ

which include both time-dependent and time-independent terms.When averaged over the orbital period, the kinetic energy associ-ated with the radial response is

E0rad;tot;avg ¼ qh

X1l¼1

lðlþ 1Þ U0l;W

��� ���2 þ W0l;W

��� ���2� �: ðD:8Þ

The sum of (D.5) and (D.8) is presented in (24).

D.2. Energy dissipation

Additionally, the amount of energy dissipation is related to theflow response. Integrated over an orbital period, the amount ofwork done by the tides should be equivalent to the amount of fric-tional energy dissipated in the ocean. To calculate these quantities,

we can take the time-averaged volumetric integral of the scalarproduct of the momentum equation and the horizontal velocityvector. In this section, we present the calculation of both of thesequantities; the calculation for viscous dissipation is far morestraightforward, while the calculation of the tidal work providesmore physical insight. We consider the two linear models for vis-cous dissipation presented in (1).

D.2.1. Viscous dissipation in the oceanThe energy dissipated in the ocean is dependent on the viscous

term employed. Tyler (2011) employs a linear drag (Rayleigh dissi-pation) formulation, in which the global dissipation rate is given by

_Elin ¼ qZ

Vða~u �~uÞdV : ðD:9Þ

where here we take the dissipation to be positive. This essentiallyscales with the kinetic energy given by (D.1). Averaged over anorbital period (cf. (24)), the viscous dissipation in terms of W andU is

_Elin;avg ¼ 2aEtot;avg ¼ 2aqhX1l¼1

lðlþ 1Þ U0l;W

��� ���2 þ W0l;W

��� ���2� �"

þX1l¼1

Xl

m¼1

lðlþ 1Þ Uml;W

��� ���2 þ Uml;E

��� ���2 þ Wml;W

��� ���2 þ Wml;E

��� ���2� �#:

ðD:10Þ

For a Navier–Stokes type viscosity, the solution to the integral

_ENS ¼ �qZ

V½ðmr2~uÞ �~u�dV ðD:11Þ

is also similar to (D.1) with an additional prefactor of �lðlþ 1Þ=R2

introduced to the summation by the Laplacian operator. The associ-ated time-averaged energy dissipation is thus,

_ENS;avg ¼2qhm

R2

X1l¼1

l2ðlþ 1Þ2 U0l;W

��� ���2 þ W0l;W

��� ���2� �"

þX1l¼1

Xl

m¼1

l2ðlþ 1Þ2 Uml;W

��� ���2 þ Uml;E

��� ���2 þ Wml;W

��� ���2 þ Wml;E

��� ���2� �#:

ðD:12Þ

D.2.2. Work done by the tideThe work done by the tidal potential can be calculated from

_E ¼ �qZ

VðrU �~uÞdV ¼ �qh

ZSðrU �~uÞdS

¼ �qhR2Z 2p

0

�Z p

0

1R@U@h

� �uh þ

1R sin h

@U@/

� �u/

� �sin hdhd/; ðD:13Þ

The full tidal potential U is the sum of each of the components pre-sented in Section 2.1,

U ¼ U02ðeiXt þ e�iXtÞY0

2

þ U12;W eiXtY1

2 � e�iXtY�12 þ e�iXtY1

2 � eiXtY�12

� �þ U2

2;W eiXtY22 þ e�iXtY�2

2

� �þ U2

2;E e�iXtY22 þ eiXtY�2

2

� �: ðD:14Þ

Similar to (D.1), the integral (D.13) is separable in m. Therefore, forclarity, we present the dissipation associated with radial compo-nent ðm ¼ 0Þ separately.

Page 16: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

26 E.M.A. Chen et al. / Icarus 229 (2014) 11–30

For the obliquity tide and the librational components of theeccentricity tide, the integral (D.13) in terms of the tidal potentialsand response coefficients is

_E¼�qhZ 2p

0

Z p

0

X2

m¼1

Um2;W eiXt @Ym

2

@hþU�m

2;W e�iXt @Y�m2

@hþUm

2;Ee�iXt @Ym2

@hþU�m

2;E eiXt @Y�m2

@h

� �" #(

�X1l¼1

Xl

m¼1

Uml;W eiXt @Ym

l

@hþU�m

l;W e�iXt @Y�ml

@hþWm

l;W eiXt 1sinh

@Yml

@/þW�m

l;W e�iXt 1sinh

@Y�ml

@/

�"

þUml;Ee�iXt @Ym

l

@hþU�m

l;E eiXt @Y�ml

@hþWm

l;Ee�iXt 1sinh

@Yml

@/þW�m

l;E eiXt 1sinh

@Y�ml

@/

��

þX2

m¼1

Um2;W eiXt 1

sinh@Ym

2

@/þU�m

2;W e�iXt 1sinh

@Y�m2

@/þUm

2;Ee�iXt 1sinh

@Ym2

@/þU�m

2;E eiXt 1sinh

@Y�m2

@/

� �" #

�X1l¼1

Xl

m¼1

Uml;W eiXt 1

sinh@Ym

l

@/þU�m

l;W e�iXt 1sinh

@Y�ml

@/�Wm

l;W eiXt @Yml

@h�W�m

l;W e�iXt @Y�ml

@h

�"

þUml;Ee�iXt 1

sinh@Ym

l

@/þU�m

l;E eiXt 1sinh

@Y�ml

@/�Wm

l;Ee�iXt @Yml

@h�W�m

l;E eiXt @Y�ml

@h

���sinhdhd/:

ðD:15Þ

Using the relation (5), the non-trivial components of (D.15) are

_E ¼ �qhZ 2p

0

Z p

0

X2

m¼1

Um2;WU�m

2;W þ U�m2;WUm

2;W þ Um2;EU

�m2;E þ U�m

2;E Um2;E

�"

þ Um2;WU�m

2;E þ U�m2;E Um

2;W

� �ei2Xt þ Um

2;EU�m2;W þ U�m

2;WUm2;E

� �e�i2Xt

�� @Ym

l

@h@Y�m

l

@hþ 1

sin2 h

@Yml

@/@Y�m

l

@/

� ��sin hdhd/: ðD:16Þ

Making use of (A.5) and recalling that the coefficients describing thetides (e.g. U1

2;W ) are real, (D.16) simplifies to

_E ¼ �2ð2Þð3ÞqhX2

m¼1

Um2;WR Um

2;W

� �þ Um

2;ER Um2;E

� ��

þ Um2;WUm�

2;E þ Um2;EU

m2;W

� �ei2Xt þ Um

2;EUm�2;W þ Um

2;WUm2;E

� �e�i2Xt

�:

ðD:17Þ

Thus, the time-averaged tidal dissipation for the obliquity and libra-tional tides is

_Eavg ¼ �2ð2Þð3ÞqhX2

m¼1

Um2;WR Um

2;W

� �þ Um

2;ER Um2;E

� �� �: ðD:18Þ

The dissipation associated with the radial component is givenby

_Erad ¼ �qhZ 2p

0

Z p

0U0

2ðeiXt þ e�iXtÞ @Y02

@h

!(

X1l¼1

U0l;W eiXt þU0�

l;W e�iXt� � @Y0

l

@h

!" #)sin hdhd/: ðD:19Þ

The solution of (D.19) can be used to derive a time-averaged dissi-pation of

_Erad;avg ¼ �2ð3ÞqhU02 U0

2;W þU0�2;W

� �¼ �2ð2Þð3ÞqhU0

2R U02;W

� �: ðD:20Þ

The sum of (D.18) and (D.20) is presented in (28).

Appendix E. Scalings for the response to the eastward obliquitytide

While the magnitudes of the forcing potentials for the eastwardand westward propagating obliquity tides are the same, the re-sponse of the ocean is fundamentally different. For the westwardresponse, the kinetic energy and energy dissipation are dominatedby the lowest mode, W1

1

�� ��. For the eastward response, the kineticenergy is significant in the lowest three modes with the largestamplitude being U1

2

�� �� (see Fig. 1). There only exists a single scaling

regime, and the approximations made here break down at smallReynolds numbers (see criteria given in (29)).

The vorticity Eq. (B.2) provides the relationships betweenW1

1;W13 and U1

2, here assuming power existing only in the threelargest modes and expressed in terms of the Reynolds number:

W11 ¼

32

C12

1Re

� 2 þ 1� � 1

Reþ i

� �U1

2 ðE:1Þ

and

W13

43

C13

12Re

� 2 þ 4936

� � 12Reþ i

76

� �U1

2: ðE:2Þ

The response coefficients are related to the forcing potentialthrough the horizontal divergence of the momentum Eq. (B.3), herewritten in terms of Reynolds number and Lamb parameter,

�i43þ i

24�þ 6

Re

� �U1

2 þ C12W

11 þ

83

C13W

13 �U1

2;W

X: ðE:3Þ

For large Re, the Reynolds number terms in the denominator of (E.1)and (E.2) can be neglected. The dominant terms of (E.3) are

6Reþ 3

2

C12

� �2

Reþ 1536

49

C13

� �2

Reþ i

24�

264

375U1

2

¼ 9237686

1Re

� �þ i

24�

� �U1

2 �U1

2;W

X; ðE:4Þ

after making the substitutions for W11 and W1

3 and neglecting theirimaginary components for small Lamb parameter. The resultingflow is given by

U12

�� �� I U12

� ���� ��� U12;W�

24X¼ 1

4

ffiffiffiffiffiffiffi2p15

rX3R4h0

gh; ðE:5Þ

W11

�� �� I W11

� ���� ��� 32

C12 U1

2

�� �� ¼ 32ffiffiffi5p U1

2

�� ��; ðE:6Þ

W13

�� �� I W13

� ���� ��� 87

C13 U1

2

�� �� ¼ 87

ffiffiffiffiffiffi8

35

rU1

2

�� ��: ðE:7Þ

Combining (E.5) with the relations (E.6) and (E.7) and the ki-netic energy Eq. (24), we find that the kinetic energy is indepen-dent of viscosity,

Etot;obl;E qh 2 W11

�� ��2 þ 6 U12

�� ��2 þ 12 W13

�� ��2h i qh

910

U12

�� ��2 þ 6 U12

�� ��2 þ 61441715

U12

�� ��2� �

¼ qh7191686

� �U1

2

�� ��2 2397p27;440

qX6R8h20

g2h: ðE:8Þ

The kinetic energy in turn may be used to determine an averagedflow speed �u:

�u

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2397

54;880

sX3R3h0

gh: ðE:9Þ

The associated energy dissipation calculated from (28) and the solu-tion of (E.4) is given by

_E ¼ �12qhU12;WR U1

2;E

� �

�12qhU12;W

�3079131;712

U12;W�

2

XRe

! !

¼ 9237p6860

qX6R6mh20

g2h: ðE:10Þ

The tidal quality factor as defined by (51) is

Page 17: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

E.M.A. Chen et al. / Icarus 229 (2014) 11–30 27

Q obl;E ¼X _Eobl;E

Etot;obl;E

X 9237p6860

qX6R6mh20

g2h

� �2397p27;440

qX6R8h20

g2h

� � ¼ 12;316799

XR2

m

!

¼ 12;316799

Re: ðE:11Þ

E.1. Relative contribution from eastward obliquity

Throughout this paper, we have ignored the response to theeastward propagating obliquity tide. Under the assumptions ofhigh Reynolds number and low Lamb parameter, it can be formallyshown that the contribution from the eastward obliquity responseis negligible when compared to that of the westward propagatingresponse.

In the high Reynolds number regime, the kinetic energy ratio isindependent of the ocean viscosity. From Section 3.1 and AppendixE, the relative contributions to the kinetic energy from the west-ward and eastward obliquity responses are

Etot;W

Etot;E 3pqhX2R4h2

0

2397p27;440

qX6R8h20

g2h

¼ 439;040799

1�

� �2

J 7700: ðE:12Þ

In the low Reynolds number regime, this ratio is sensitive to theReynolds number with

Etot;W

Etot;E

3p400

qX8R12h20

m2g2h

2397p27;440

qX6R8h20

g2h

¼ 343495

Re2 J 1700: ðE:13Þ

Similarly, the ratio of energy dissipation is independent of viscosityin the high Reynolds number regime,

_Eobl;W

_Eobl;E

12pqhX2R2mh20

9237p6860

qX6R6mh20

g2h

¼ 439;0403079

1�

� �2

J 2000 ðE:14Þ

and scales inversely with the Reynolds number in the low Reynoldsnumber regime,

_Eobl;W

_Eobl;E

3p100

qX8R10h20

mg2h

9237p6860

qX6R6mh20

g2h

¼ 34315;395

Re2 J 55: ðE:15Þ

Typical satellite parameters (e.g. Tables 1 and 2) usually lie farfrom the scaling limits used to calculate these ratios, and the ratiospresented in Eqs. (E.12)–(E.14) and (E.15) are typically much lowerthan those for actual satellite parameters. Therefore, we reason-ably ignore the contribution of the eastward obliquity responsein all obliquity-related scalings.

Appendix F. Scalings for the response to the eccentricity tide

The responses to the eccentricity tide is similar to that of theeastward propagating obliquity tide, in that there is a single scalingregime characterized by a kinetic energy profile that is indepen-dent of viscosity and energy dissipation linearly varying with vis-cosity. Unlike the unequal responses of the eastward andwestward propagating obliquity tide, the magnitudes of the re-sponses to each of the three components of the eccentricity tideare comparable. We present scalings for each tidal component sep-arately, as well as including scalings for the total response.

F.1. Scalings for the response to the radial eccentricity tide

The response to the radial eccentricity tidal potential can beapproximated with the lowest three modes, W0

1;U02, W0

3. From

(21) and ignoring the contributions from the Reynolds numberterms in the denominator, similar to the process used in AppendixE, the relations between the three modes are

W01 3C0

22Re� i

� �U0

2 ðF:1Þ

and

W03

43

C03

12Re� i

� �U0

2 ðF:2Þ

These modes are related to the tidal forcing by

iþ 6Re� i24�þ2

C02

2

!3C0

22Re� i

� �� �þ2

4C03

3

!43

C03

12Re� i

� �� �" #U0

2

�U02

X;

ðF:3Þ

where the dominant real and imaginary terms end up being

1307

1Re

� �� i24�

� �U0

2 �U0

2

X: ðF:4Þ

The forced response jU02j is thus,

U02

�� �� I U02

� ���� ��� U02�

24X¼ 1

4

ffiffiffiffip5

rX3R4e

gh: ðF:5Þ

For large Reynolds numbers, the relations (F.1) and (F.2) aredominated by the imaginary term, such that

W01

�� �� I W01

� ���� ��� 3C02 U0

2

�� �� ¼ 3

ffiffiffiffiffiffi4

15

rU0

2

�� ��; ðF:6Þ

and

W03

�� �� IðW03Þ

�� �� 43

C03 U0

2

�� �� ¼ 4

ffiffiffiffiffiffi1

35

rU0

2

�� ��: ðF:7Þ

Using the lowest three modes, the average kinetic energy associatedwith the radial tide can be approximated from (24) as

Etot;rad qh 2 W01

�� ��2 þ 6 U02

�� ��2 þ 12 W03

�� ��2h i: ðF:8Þ

Combining the relations (F.1), (F.2) and (F.5) with (F.8), we find thatthe average kinetic energy is

Etot;rad qh 2 3C02 U0

2

�� ��� �2þ 6 U0

2

�� ��2 þ 1243

C03 U0

2

�� ��� �2" #

¼ qh114

7U0

2

�� ��2� � 57p

280qX6R8e2

g2h; ðF:9Þ

and thus the associated average velocity is given by

�u ffiffiffiffiffiffiffiffiffi57

560

rX3R3e

gh: ðF:10Þ

From (F.4), we can also calculate R U02

� �,

RðU02Þ

�U02

X 24�

� 2

1307

1Re

� �� �¼ �65

84

ffiffiffiffip5

rX4R4me

g2h2 : ðF:11Þ

The resulting dissipation from (D.20) is thus

_Erad �12qhU02R U0

2

� � 39p

14qX6R6me2

g2h: ðF:12Þ

Combining (F.9) and (F.12), we find that the tidal quality factorassociated with the radial response is

Page 18: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

28 E.M.A. Chen et al. / Icarus 229 (2014) 11–30

Q rad X 57p

280qX6R8e2

g2h

� �39p14

qX6R6me2

g2h

¼ 19260

XR2

m

!¼ 19

260Re: ðF:13Þ

F.2. Scalings for the response to the westward propagating componentof the librational eccentricity tide

The response of the ocean to the librational eccentricity tidalpotential is dominated by the lowest two modes, U2

2 and W23. To cal-

culate the westward response, we return to the vorticity equationgiven by (21). It can be shown that for large Reynolds numbers

W23 3C2

312Re� i

23

� �U2

2 ðF:14Þ

and

W23

�� �� IðW23Þ

�� �� 2C23 U2

2

�� �� ¼ 2

ffiffiffi17

rU2

2

�� ��: ðF:15Þ

From (22) and including the dominant real and imaginary terms asbefore, the tidal response is given by

6Reþ

96 C23

� �2

Re� i

24�

264

375U2

2¼138

71Re

� �� i

24�

� �U2

2�U2

2;W

X: ðF:16Þ

Similar to the response to the radial component, the kinetic energyis related to

U22

�� �� I U22

� ���� ��� U22;W

��� ����24X

¼ 18

ffiffiffiffiffiffiffi2p15

rX3R4e

ghðF:17Þ

while the dissipation is related to

R U22

� ��U2

2;W

X 24�

� 2

1387

1Re

� �� �¼ �23

56

ffiffiffiffiffiffiffi2p15

rX4R4me

g2h2

!: ðF:18Þ

The average kinetic energy associated with the response can beapproximated from (24), (F.15) and (F.17) as

Etot;lib;W qh 6 U22

�� ��2 þ 12 W23

�� ��2h i qh 6 U2

2

�� ��2 þ 12 2C23 U2

2

�� ��� �2� �

¼ qh907

U22

�� ��2� �

3p112

qX6R8e2

g2hðF:19Þ

with a corresponding average flow speed of

�u ffiffiffiffiffiffiffiffiffi

3224

rX3R3e

gh: ðF:20Þ

From the relations (28) and (F.18), the associated energy dissipationis

_Elib;W �12qhU22;WR U2

2;W

� � 69p

140qX6R6me2

g2h: ðF:21Þ

The associated Q is therefore

Q lib;W X 3p

112qX6R8e2

g2h

� �69p140

qX6R6me2

g2h

� � ¼ 592

XR2

m

!¼ 5

92Re: ðF:22Þ

F.3. Scalings for the response to the eastward propagating componentof the librational eccentricity tide

The process to calculate the response to the eastward propagat-ing librational component of the eccentricity tide is essentially the

same as that for the westward response. The relationship betweenU2

2 and W23 as given by (B.2) is

W23

3C23

412Reþ i

43

� �U2

2 ðF:23Þ

with

W23

�� �� IðW23Þ

�� �� C23 U2

2

�� �� ¼ffiffiffi17

rU2

2

�� ��: ðF:24Þ

The dominant terms in the horizontal divergence Eq. (B.3) are

6Reþ 24ðC2

3Þ2

Reþ i

24�

" #U2

2 ¼667

1Re

� �þ i

24�

� �U2

2 �U2

2;E

X: ðF:25Þ

The resulting flow is given by

U22

�� �� IðU22Þ

�� �� U22;E

��� ����24X

¼ 78

ffiffiffiffiffiffiffi2p15

rX3R4e

ghðF:26Þ

with

R U22

� ��U2

2;E

X 24�

� 2

667

1Re

� �� �¼ 11

8

ffiffiffiffiffiffiffi2p15

rX4R4me

g2h2

!: ðF:27Þ

Similar to (F.19), the average kinetic energy is

Etot;lib;E qh 6 U22

�� ��2 þ 12 C23 U2

2

�� ��� �2� �

¼ qh547

U22

�� ��2� �

63p80

qX6R8e2

g2hðF:28Þ

and the flow speed is

�u ffiffiffiffiffiffiffiffiffi63

160

rX3R3e

ghðF:29Þ

Note that this value exceeds the equivalent westward flow speed(Eq. (F.20)), as expected. The corresponding energy dissipation forthe eastward flow (28) is

_Elib;E �12qhU22;ER U2

2;E

� � 231p

20qX6R6me2

g2h: ðF:30Þ

The tidal quality factor is

Qlib;E X 63p

80qX6R8e2

g2h

� �231p

20qX6R6me2

g2h

� � ¼ 344

XR2

m

!¼ 3

44Re: ðF:31Þ

F.4. Scalings for the total response to the eccentricity tide

The three components of the eccentricity-driven flow behavequite similarly, and as seen in Appendices F.1, F.2 and F.3, the maindifference exists in the pre-factors defining each of the flows. From(F.9), (F.19) and (F.28), the ratios of the kinetic energies are

Etot;lib;E 14738

Etot;rad 147

5Etot;lib;W ðF:32Þ

with a total kinetic energy associated with the eccentricity tide of

Etot;ecc 57p56

qX6R8e2

g2h: ðF:33Þ

The energy dissipation behaves similarly with the eastward libra-tional response contributing a bulk of the dissipation with non-neg-ligible contributions from the other components. The relativedissipation rates from (F.12), (F.21) and (F.30) are

_Elib;E 539130

_Erad 53923

_Elib;W : ðF:34Þ

Page 19: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

E.M.A. Chen et al. / Icarus 229 (2014) 11–30 29

The total energy dissipation associated with the eccentricity tide is

_Eecc;tot 519p

35qX6R6me2

g2h: ðF:35Þ

Thus, an approximation for Q related to all components of theeccentricity tide is

Q ecc X 57p

56qX6R8e2

g2h

� �519p

35qX6R6me2

g2h

¼ 951384

XR2

m

!¼ 95

1384Re: ðF:36Þ

References

Agnor, C.B., Hamilton, D.P., 2006. Neptune’s capture of its moon Triton in a binary-planetary gravitational encounter. Nature 441, 192–194.

Baines, P., 1973. The generation of internal tides by flat-bump topography. Deep SeaRes. 20, 179–205.

Baland, R., Yseboodt, M., Hoolst, T.V., 2012. Obliquity of the Galilean satellites: Theinfluence of a global internal liquid layer. Icarus 220, 435–448.

Besserer, J., Nimmo, F., Roberts, J.H., Pappalardo, R.T., 2013. Convection-drivencompaction as a possible origin of Enceladus’s long wavelength topography. J.Geophys. Res. 118. http://dx.doi.org/10.1002/jgre.20079.

Beuthe, M., 2013. Spatial patterns of tidal heating. Icarus 223, 308–329.Bills, B.G., 2005. Free and forced obliquities of the Galilean satellites of Jupiter.

Icarus 175, 233–247.Bills, B.G., Nimmo, F., 2008. Forced obliquity and moments of inertia of Titan. Icarus

196, 293–297.Bills, B.G., Nimmo, F., 2011. Rotational dynamics and internal structure of Titan.

Icarus 214, 351–355.Bland, M.T., Showman, A.P., Tobie, G., 2009. The orbital–thermal evolution and

global expansion of Ganymede. Icarus 199, 207–221.Brito, D., Aurnou, J., Cardin, P., 2004. Turbulent viscosity measurements relevant to

planetary core-mantle dynamics. Phys. Earth Planet. Inter. 141, 3–8.Buffett, B.A., Christensen, U.R., 2007. Magnetic and viscous coupling at the core-

mantle boundary: Inferences from observations of Earth’s nutations. Geophys. J.Int. 171, 143–152.

Castillo-Rogez, J.C., Efroimsky, M., Lainey, V., 2011. The tidal history of Iapetus: Spindynamics in the light of a refined dissipation model. J. Geophys. Res. 116. http://dx.doi.org/10.1029/2010JE004664.

Chen, E.M.A., Nimmo, F., 2011. Obliquity tides do not significantly heat Enceladus.Icarus 214, 779–781.

Chen, E.M.A., Glatzmaier, G.A., Nimmo, F., 2010. Modeling the dynamics of icysatellite sub-surface oceans with focus on implications for spacecraftobservables. Lunar Planet. Sci. 41, 1454.

Colombo, G., 1966. Cassini’s second and third laws. Astron. J. 71, 891–900.Croft, S.K., Kargel, J.S., Kirk, R.L., Moore, J.M., Schenk, P.M., Strom, R.G., 1995. The

geology of Triton. In: Cruikshank, D.P. (Ed.), Neptune and Triton. Univ. ArizonaPress, pp. 879–947.

Darwin, G.H., 1899. The theory of the figure of the Earth carried to the second orderof small quantities. Mon. Not. R. Astron. Soc. 60, 82–124.

Efroimsky, M., Williams, J.G., 2009. Tidal torques: A critical review of sometechniques. Celest. Mech. Dyn. Astron. 104, 257–289.

Egbert, G.D., Ray, R.D., 2000. Significant dissipation of tidal energy in the deep oceaninferred from satellite altimeter data. Nature 405, 775–778.

Egbert, G.D., Ray, R.D., 2001. Estimates of M2 tidal energy dissipation from TOPEX/Poseidon altimeter data. J. Geophys. Res. 106, 22475–22502.

Ellsworth, K., Schubert, G., 1983. Saturn’s icy satellites: Thermal and structuralmodels. Icarus 54, 490–510.

Gaeman, J., Hier-Majumder, S., Roberts, J.H., 2012. Sustainability of a subsurfaceocean within Triton’s interior. Icarus 220, 339–347.

Garrick-Bethell, I., Nimmo, F., Wieczorek, M.A., 2010. Structure and formation of thelunar farside highlands. Science 330, 949–951.

Glatzmaier, G.A., 2002. Geodynamo simulations – How realistic are they? Annu.Rev. Earth Planet. Sci. 30, 237–257.

Goldreich, P., Soter, S., 1966. Q in the Solar System. Icarus 5, 375–389.Goodman, J.C., Collins, G.C., Marshall, J., Pierrehumbert, R.T., 2004. Hydrothermal

plume dynamics on Europa: Implications for chaos formation. J. Geophys. Res.109. http://dx.doi.org/10.1029/2003JE002073.

Henning, W.G., O’Connell, R.J., Sasselov, D.D., 2009. Tidally heated terrestrialexoplanets: Viscoelastic response models. Astrophys. J. 707, 1000–1015.

Howett, C., Spencer, J.R., Pearl, J., Segura, M., 2011. High heat flow from Enceladus’south polar region measured using 10–600/cm Cassini/CIRS data. J. Geophys.Res. 116. http://dx.doi.org/10.1029/2010JE003718.

Hubbard, W.B., Anderson, J.D., 1978. Possible flyby measurements of Galileansatellite interior structure. Icarus 33, 336–341.

Hussmann, H., Spohn, T., 2004. Thermal–orbital evolution of Io and Europa. Icarus171, 391–410.

Iess, L. et al., 2012. The tides of Titan. Science 337, 457–459.Jacobson, R.A. et al., 2006. The gravity field of the saturnian system from satellite

observations and spacecraft tracking data. Astron. J. 132, 2520–2526.Jacobson, R.A., 2009. The orbits of the Neptunian satellites and the orientation of the

pole of Neptune. Astron. J. 137, 4322–4329.

Jacobson, R.A., Campbell, J.K., Taylor, A., Synott, S.P., 1992. The masses of Uranus andits major satellites from Voyager tracking data and Earth-based Uraniansatellite data. Astron. J. 103, 2068–2078.

Jankowski, D.G., Chyba, C.F., Nicholson, P.D., 1989. On the obliquity and tidalheating of Triton. Icarus 80, 211–219.

Jayne, S.R., Laurent, L.C.S., 2001. Paramterizing tidal dissipation over roughtopography. Geophys. Res. Lett. 28, 211–219.

Jeffreys, H., 1921. Tidal friction in shallow seas. Phil. Trans. R. Soc. A 221, 239–264.Kargel, J.S., Kaye, J.Z., Head, J.W., Marion, G.M., Sassen, R., Crowley, J.K., Ballesteros,

O.P., Grant, S.A., Hogenboom, D.L., 2000. Europa’s crust and ocean: Origin,composition, and the prospects for life. Icarus 148, 226–265.

Kaula, W.M., 1964. Tidal dissipation by solid friction and the resulting orbitalevolution. Rev. Geophys. 2, 661–685.

Khurana, K.K. et al., 1998. Induced magnetic fields as evidence for subsurface oceansin Europa and Callisto. Nature 395, 777–780.

Kivelson, M.G., Khurana, K., Volwerk, M., 2002. The permanent and inductivemagnetic moments of Ganymede. Icarus 157, 507–522.

Longuet-Higgins, M., 1968. The eigenfunctions of Laplace’s tidal equations over asphere. Phil. Trans. R. Soc. A 262, 511–607.

Love, A.E.H., 1944. A Treatise on the Mathematical Theory of Elasticity. Dover.Lumb, L.I., Aldridge, K.D., 1991. On viscosity estimates for the Earth’s fluid outer

core and core-mantle coupling. J. Geomag. Geoelectr. 43, 93–110.Matsuyama, I., 2012. Tidal dissipation in the subsurface oceans of icy satellites.

Lunar Planet. Sci. 43, 2068.McKinnon, W.B., Kirk, R.L., 2007. Triton. In: McFadden, L.-A., Weissman, P.R.,

Johnson, T.V. (Eds.), Encyclopedia of the Solar System. Academic Press, pp. 483–502.

Meyer, J., Wisdom, J., 2007. Tidal heating in Enceladus. Icarus 188, 535–539.Meyer, J., Elkins-Tanton, L., Wisdom, J., 2010. Coupled thermal–orbital evolution of

the early Moon. Icarus 208, 1–10.Mound, J., Buffett, B., 2007. Viscosity of the Earth’s fluid core and torsional

oscillations. J. Geophys. Res. 112. http://dx.doi.org/10.1029/2006JB004426.Munk, W.H., MacDonald, G.J.F., 1960. The Rotation of the Earth: A Geophysical

Discussion. Cambridge Univ. Press.Murray, C.D., Dermott, S.F., 1999. Solar System Dynamics. Cambridge Univ. Press.Nimmo, F., Bills, B.G., 2010. Shell thickness variations and the long-wavelength

topography of Titan. Icarus 208, 896–904.Nimmo, F., Gaidos, E., 2002. Strike-slip motion and double ridge formation on

Europa. J. Geophys. Res. 107. http://dx.doi.org/10.1029/2000JE001476.Nimmo, F., Faul, U.H., Garnero, E.J., 2012. Dissipation at tidal and seismic

frequencies in a melt-free Moon. J. Geophys. Res. 117. http://dx.doi.org/10.1029/2012JE004160.

Ogilvie, G.I., Lin, D.N.C., 2004. Tidal dissipation in rotating giant planets. Astrophys.J. 610, 477–509.

Ojakangas, G.W., Stevenson, D.J., 1989. Thermal state of an ice shell on Europa.Icarus 81, 220–241.

Peale, S.J., 1969. Generalized Cassini’s laws. Astron. J. 74, 483–489.Peale, S.J., Cassen, P., 1978. Contributions of tidal dissipation to lunar thermal

history. Icarus 36, 245–269.Peale, S.J., Cassen, P., Reynolds, R.T., 1979. Melting of Io by tidal dissipation. Science

203, 892–894.Pope, S.B., 2000. Turbulent Flows. Cambridge Univ. Press.Postberg, F., Schmidt, J., Hillier, J., Kempf, S., Srama, R., 2011. A salt-water reservoir

as the source of a compositionally stratified plume on Enceladus. Nature 474,620–622.

Prockter, L.M., Nimmo, F., Pappalardo, R.T., 2005. A shear heating origin for ridges onTriton. Geophys. Res. Lett. 32. http://dx.doi.org/10.1029/2005GL0022832.

Radau, R., 1885. Sur la loi des densites a l’interieur de la terre. Compt. Rend. 100,972–974.

Ray, R., Eanes, R.J., Lemoine, F.G., 2001. Constraints on energy dissipation in theEarth’s body tide from satellite tracking and altimetry. Geophys. J. Int. 144, 471–480.

Roberts, J.H., Nimmo, F., 2008. Tidal heating and the long-term stability of asubsurface ocean on Enceladus. Icarus 194, 675–689.

Ross, M.N., Schubert, G., 1989. Viscoelastic models of tidal heating on Enceladus.Icarus 78, 90–101.

Ross, M.N., Schubert, G., 1990. The coupled orbital and thermal evolution of Triton.Geophys. Res. Lett. 17, 1749–1752.

Schenk, P.M., McKinnon, W.B., 2009. One-hundred-km-scale basins on Enceladus:Evidence for an active ice shell. Geophys. Res. Lett. 26. http://dx.doi.org/10.1029/2009GL039916.

Schenk, P.M., Zahnle, K., 2007. On the negligible surface age of Triton. Icarus 192,135–149.

Schubert, G., Anderson, J.D., Spohn, T., McKinnon, W.B., 2004. Interior composition,structure and dynamics of the Galilean satellites. In: Bagenal, F., Dowling, T.,McKinnon, W. (Eds.), Jupiter: The Planet, Satellites and Magnetosphere.Cambridge Univ. Press, pp. 281–306.

Segatz, M., Spohn, T., Ross, M., Schubert, G., 1988. Tidal dissipation, surface heatflow, and figure of viscoelastic models of Io. Icarus 75, 187–206.

Showman, A.P., Stevenson, D.J., Malhotra, R., 1997. Coupled orbital and thermalevolution of Ganymede. Icarus 129, 367–383.

Smith, B.A. et al., 1989. Voyager 2 at Neptune: Imaging science results. Science 246,1422–1449.

Smylie, D., 1999. Viscosity near Earth’s solid inner core. Science 284, 461–463.Soderlund, K.M., Schmidt, B.E., Blankenship, D.E., Wicht, J., 2013. Dynamics of

Europa’s ocean and sensitivity to water properties. Lunar Planet. Sci. 44, 3009.

Page 20: Tidal heating in icy satellite oceans€¦ · iXtY1 2e Xi tY 1Þ ¼ 2 3 2 ffiffiffiffiffiffiffi 2p 15 r X2R2h 0! RðeiXtY1 2Þ 2U 1 2;W R e iXtY1; ð6Þ and a symmetric eastward propagating

30 E.M.A. Chen et al. / Icarus 229 (2014) 11–30

Sohl, F., Sears, W.D., Lorenz, R.D., 1995. Tidal dissipation on Titan. Icarus 115, 278–294.

Spencer, J.R. et al., 2006. Cassini encounters Enceladus: Background and thediscovery of a south polar hot spot. Science 311, 1401–1405.

Squyres, S.W., Reynolds, R.T., Cassen, P.M., Peale, S.J., 1983. The evolution ofEnceladus. Icarus 53, 319–331.

Stiles, B.W. et al., 2008. Determining Titan’s spin state from Cassini RADAR images.Astron. J. 135, 1669–1680.

Tackley, P.J., Schubert, G., Glatzmaier, G.A., Schenk, P., Ratcliff, J.T., Matas, J.,2001. Three-dimensional simulations of mantle convection in Io. Icarus149, 79–93.

Taylor, G.I., 1920. Tidal friction in the Irish Sea. Phil. Trans. R. Soc. A 220, 1–33.Thomas, P.C., 1988. Radii, shapes and topography of the satellites of Uranus from

limb coordinates. Icarus 73, 427–441.Thomas, P.C., 2000. The shape of Triton from limb profiles. Icarus 148, 587–588.Thomas, P.C., 2010. Sizes, shapes, and derived properties of the saturnian satellites

after the Cassini nominal mission. Icarus 208, 395–401.

Tobie, G., Mocquet, A., Sotin, C., 2005. Tidal dissipation within large icy satellites:Applications to Europa and Titan. Icarus 177, 534–549.

Tyler, R.H., 2008. Strong ocean tidal flow and heating on moons of the outer planets.Nature 456, 770–772.

Tyler, R.H., 2009. Ocean tides heat Enceladus. Geophys. Res. Lett. 36. http://dx.doi.org/10.1029/2009GL038300.

Tyler, R., 2011. Tidal dynamical considerations constrain the state of an ocean onEnceladus. Icarus 211, 770–779.

Ward, W.R., 1975. Past orientation of the lunar spin axis. Science 189, 377–379.Wisdom, J., 2004. Spin–orbit secondar resonance dynamics of Enceladus. Astron. J.

128, 484–491.Wisdom, J., 2008. Tidal dissipation at arbitrary eccentricity and obliquity. Icarus

193, 637–640.Zebker, H.A., Stiles, B., Hensley, S., Lorenz, R., Kirk, R.L., Lunine, J., 2009. Size and

shape of Saturn’s moon Titan. Science 324, 921–923.Zhang, K., Nimmo, F., 2012. Late-stage impacts and the orbital and thermal

evolution of Tethys. Icarus 218, 348–355.