The energy dissipation in a uniform vorticity flow, such as the flow in a precessing spheroid or the one associated with the earth’s free core nutation, is mainly confined to the boundary layers. However, the thinness of the boundary layer renders it difficult to study the energy dissipation in the turbulent regime, either in laboratory experiments or through direct numerical simulations. Here, we use a local Cartesian model to study the energy dissipation in the boundary layer of a precessing sphere when the flow becomes turbulent, contrasting it with the laminar case. We compute the evolution of the boundary layer over time at individual co-latitudes based on direct numerical simulations using the computational fluid dynamics solver Nek5000. We then estimate the total global dissipation by summing up individual contributions. A comparison with known analytical results in the laminar case validates this approach. We briefly discuss the applications to the lunar and the earth’s core cases.
I. INTRODUCTION
Free Core Nutation (FCN) is a rotational normal mode of the earth. Since its frequency is close to forced nutation, the retrograde annual nutation, some properties of the FCN can be determined by measuring the resonant amplitude and the phase lag of the forced nutation [e.g., Herring (1986)]. Inversions from the observed phase lag placed an upper bound for the fluid core kinematic viscosity of 0.54 m2 s−1 (Gwinn , 1986), about six orders larger than the value inferred from studies of liquid metal [see the review of Lums and Aldridge (1991)]; a low viscosity of ν = 10−6 m2 s−1 is still favored by recent studies based on ab initio calculations [e.g., Pozzo (2013) and Ichikawa and Tsuchiya (2015)]. Turbulence in the core’s boundary layers can lead, in principle, to an enhanced eddy viscosity. However, Ohmic dissipation in the core and mantle is a more natural way to explain the observations [e.g., the studies by Buffett (1992), Mathews and Guo (2005), Deleplace and Cardin (2006), and Buffett and Christensen (2007)]. However, accounting for the observed phase lag requires either a strong magnetic field at the core-mantle boundary (CMB) or a highly conductive layer at the base of the mantle (Mathews , 2002; Buffett , 2002). Although the flow near the CMB associated with the FCN motion is too weak to be turbulent, the flow associated with precession might be strong enough to be turbulent (Buffett, 2021; Triana , 2021). Thus, it is important to properly quantify the turbulent contribution to the energy dissipation from precessional motion to obtain tighter bounds on the Ohmic dissipation.
Instabilities in the boundary layer of a precession-driven flow have been previously identified in numerical simulations (Lorenzani and Tilgner, 2001); however, the thinness of the boundary layer makes it difficult to study experimentally [see the review of Tilgner (2015)]. That same reason, combined with current computational constraints, precludes numerical studies from reaching regimes of parameters appropriate for planetary interiors where the Coriolis force plays a central role. This limitation is reflected in the smallness of the Ekman number E, representing the ratio of viscous to Coriolis forces and controlling the thickness Lν of the boundary layer (also known as the Ekman layer), which scales as Lν ∝ E1/2. We can define a local Reynolds number Re to quantify the condition for turbulence in the boundary layer as Re = ULν/ν, where ν is the viscosity and U is the velocity. For the nominal value of the earth’s precession-driven flow, the Reynolds number is about Re = 500, higher than the critical value of Re ∼ 150 for the onset of turbulence proposed by Sous (2013). With a global spherical model, Cébron (2019) showed that the total viscous dissipation increases by a factor of 1.17 compared to the laminar solution when Re is around 230. Buffett (2021) restricted the modeling domain to the boundary layer and broke it into pieces according to the co-latitude θ. Using a local Cartesian model, similar to ours, he studied two representative cases at θ = 0 and 60° and showed that the friction velocity u*, a quantity proportional to the square root of the surface shear stress (see the definition below), increases by a factor of 1.2 at Re = 500 compared to the laminar value. Buffett (2021) demonstrated a promising way to study the turbulent effects in the boundary layer from a local point of view. However, the connection to the global picture is still missing. Dissipation depends not only on the surface shear stress but also on the veering angle β, the angle between the stress and the fluid velocity on the boundary. In the present work, we numerically investigate the local flow at different co-latitudes and integrate the results to obtain global estimates of the possible turbulent energy dissipation.
Energy dissipation in the lunar core due to precession is a key parameter in understanding the possible dynamo action of the past (Stys and Dumberry, 2020; Zhang and Dumberry, 2021). The Re for the boundary layer flow is about 104, much larger than earth’s. Therefore, it has been believed that the boundary layer is turbulent [e.g., Yoder (1981)]. The model by Buffett (2021) is, in principle, not applicable to the moon not only because of the much higher Reynolds number but also because the flattening in their model must be much larger than the ratio of precession to rotation (i.e., the Poincaré number Po). However, in the lunar case, the latter condition is not met as the CMB flattening of 10−4 (Viswanathan , 2019) is smaller than the Poincaré number, Po = 4 × 10−3. In order to extend the result achieved by Buffett (2021), we investigate the problem by considering the case where the flattening is zero, i.e., a precessing fluid in a spherical cavity.
The scaling of quantities in the boundary layer can be used to estimate the dissipation at large Re. The theory for the scaling of u* and β in the Ekman layer was proposed by Csanady (1967). It was later improved by Spalart (1989) by including higher order terms. Scaling has been studied via laboratory experiments (Caldwell , 1972; Sous , 2013) and numerical simulations (Coleman , 1990; Coleman, 1999; Shingai and Kawamura, 2004; Deusebio , 2014; and Braun , 2020). However, the Ekman layer in those studies is steady, i.e., it remains constant all the time, contrary to the one from precession-driven flow, which oscillates with diurnal frequency. The study of a Stokes boundary layer, an oscillating boundary layer but without Coriolis force, reveals complex behavior during the transition to turbulence: an instability growing in one oscillation cycle may eventually decay at intermediate Re (Ozdemir , 2014). Besides, analytical solutions show that the u* associated with an oscillatory Ekman layer is smaller than that of a steady Ekman layer by a factor of 21/4 (Buffett, 2021). Therefore, we want to know if scaling is applicable to the oscillatory Ekman layer and establish the difference with those previous studies.
In Sec. II, we present the Navier–Stokes equation in the fluid frame using local Cartesian coordinates and compare it to the one used by Buffett (2021). Next, we outline the numerical implementation in Sec. III. Numerical solutions are presented in Sec. IV. Finally, the discussion and conclusions appear in Sec. V.
II. PROBLEM FORMULATION
We consider a spherical cavity with radius R filled with an incompressible and viscous fluid, rotating with angular velocity Ωm and precessing with angular velocity Ωp. The angle between Ωm and Ωp is denoted as α, the precession angle. In the case of weak precession (Ωp ≪ Ωm), the leading order solution for the flow in the cavity may be expressed as a solid body rotation Ωf, which represents the mean rotation vector of the fluid. Since the dominant forcing in the system is solely related to precession and the fluid viscosity ν, the orientation of Ωf depends on their relative amplitudes. According to Busse’s theory (Busse, 1968), Ωf can be determined, given the Poincare number Po = Ωp/Ωm, the Ekman number E = ν/(ΩmR2), and the precession angle α. We choose our reference frame as the one rotating with the fluid’s mean angular velocity, and we refer to it as the fluid frame. In such a frame, the bulk of the fluid remains at rest, and the cavity rotates with . A boundary layer exists, the main focus of our study, with a nominal thickness .
To determine Ωf, we use the reduced model developed by Cébron (2019) [i.e., their Eqs. (5)–(7)]. Their solutions are obtained in the precessing frame where a Cartesian coordinate system is defined such that the -axis and Ωm are aligned. The spherical coordinates of Ωf are denoted as the co-latitude θf and the longitude ϕf. The coordinates are transformed to the local system by three steps: (1) rotation of the coordinate system to align the -axis with Ωf, (2) transformation from the precessing frame to the fluid frame, and (3) transformation from the global system to the local system. Details about the transformation are given in Appendix A. So far, we are ready to solve Eq. (2) numerically.
Parameter . | Definition . | Units . | Value . |
---|---|---|---|
ν | m2 s−1 | 10−6 | |
Ωm | rad s−1 | 2.66 × 10−6 | |
Ωf | Ωm cos θf | rad s−1 | 2.659 × 10−6 |
Ωp | rad s−1 | 1.07 × 10−8 | |
R | km | 380 | |
Lν | m | 0.61 | |
α | deg | 1.543 | |
θf | deg | 1.543 | |
ϕf | deg | 0.060 | |
γ | deg | 0.002 | |
U | RΩm sin θf | m s−1 | 2.72 × 10−2 |
Po | Ωp/Ωm | 4 × 10−3 | |
E | ν/(R2Ωm) | 3 × 10−12 | |
Re | ULν/ν | 1.67 × 104 |
Parameter . | Definition . | Units . | Value . |
---|---|---|---|
ν | m2 s−1 | 10−6 | |
Ωm | rad s−1 | 2.66 × 10−6 | |
Ωf | Ωm cos θf | rad s−1 | 2.659 × 10−6 |
Ωp | rad s−1 | 1.07 × 10−8 | |
R | km | 380 | |
Lν | m | 0.61 | |
α | deg | 1.543 | |
θf | deg | 1.543 | |
ϕf | deg | 0.060 | |
γ | deg | 0.002 | |
U | RΩm sin θf | m s−1 | 2.72 × 10−2 |
Po | Ωp/Ωm | 4 × 10−3 | |
E | ν/(R2Ωm) | 3 × 10−12 | |
Re | ULν/ν | 1.67 × 104 |
III. NUMERICAL IMPLEMENTATION
We use the spectral element code Nek5000 (Fischer , 2007) to solve Eq. (4) for an incompressible fluid in a Cartesian box. The domain is divided into elements, where solutions are represented by N-th order Lagrange interpolation polynomials (see Fig. 2). The typical resolution in the present study is and N = 9. The temporal discretization is based on a third-order backward-difference scheme for implicit terms and on an extrapolation scheme for explicit terms. We use dealiasing following the 3/2 rule.
The domain is extended downward from z = 0. The non-dimensional size is 30 × 30 × 30. The top boundary is a wall (representing the CMB), where the velocity is specified to mimic the precessing container (mantle), as seen from the fluid frame [see Eq. (3)]. The four lateral boundaries are periodic. The bottom boundary is stress-free (i.e., no tangential stress), and the normal velocity is zero. We have performed a height test to make sure that the domain height of 30Lν is large enough to accommodate the boundary layer flow (see Fig. S2 in the supplementary material). To resolve the boundary layer flow, we use a higher element density near the top boundary. We will explore the parameter space: Re = [5, 500] and θ = [0, 90°]. Simulations for Re = 700 are added to extend the Re range. We also simulate the case with Re = 600 at θ = 0. The range of Re covers both the laminar and turbulent regimes of the oscillatory Ekman layer.
The simulation always starts from the fluid at rest, where a random perturbation in velocity with root-mean-square amplitude of about 10−2 is imposed. In most cases, the flow will go through a transient state for a certain period and reach a statistically steady state (we shall refer to it as a steady state for the sake of brevity in the following sections), which means time-averaged quantities over the oscillation period(s) remain constant. The mean values presented below are obtained by taking time averages over oscillation periods. Due to the homogeneity on x- and y-directions, all the values are also averaged on the horizontal plane. The unit of time t⊕ presented below is 2π/Ωf, indicating the period of the oscillating boundary. In what follows, the superscript (+) denotes the quantities non-dimensionalized by the viscous wall scales, velocity u*, and length ν/u*. The notation is chosen to describe the flow near the boundary.
IV. RESULTS
We present now our numerical results, comparing them with analytical solutions, highlighting the difference between laminar and turbulent regimes. Then we use the numerical results in the turbulent regime together with the asymptotic theory to construct a turbulence model for the energy dissipation. Finally, we compute the total dissipation based on the local model and calibrate the turbulence model accordingly.
A. Friction velocity and kinetic energy
The friction velocity u* is a quantity commonly used to monitor the amplitude of surface shear stress |τ|: [e.g., in the study by Tennekes (1972)]. At θ = 0, Eqs. (B19) and (B20) give the laminar solution , which serves as the benchmark for low Reynolds number cases. Figure 3 shows the ratio as a function of time at θ = 0 for different values of Re. We see that the friction velocity deviates from its laminar solution when Re ≥ 300 and the deviation is larger for higher Re. Concerning the case of Re = 300, it appears that the deviation from the laminar solution by about 10% can be attributed to turbulence, as supported by the presence of a logarithmic region presented later in this study. We observe that the transient period happens before t⊕ = 3 (before t⊕ = 1 for Re = 400 and higher) and the friction velocity reaches a steady value. The result is consistent with the one from the study by Buffett (2021), where the solution at Re = 500 is about 20% larger than the laminar value after the onset of turbulence. To obtain the mean value, we take a time average over the interval 1 ≤ t⊕ ≤ 5, except for Re = 300 where we use the 3 < t⊕ < 5 interval (see Table II).
Re . | . | . | (deg) . | . |
---|---|---|---|---|
5 | 1.00 ± 0.00 | 0.447 192 ± 0.000 419 | 45.002 058 ± 0.095 827 | 0.96 |
100 | 1.00 ± 0.00 | 0.099 973 ± 0.000 224 | 45.010 85 ± 0.229 41 | 0.96 |
200 | 1.00 ± 0.00 | 0.070 693 ± 0.000 158 | 45.010 364 ± 0.229 03 | 0.97 |
300 | 1.13 ± 0.02 | 0.064 989 ± 0.001 234 | 29.350 199 ± 1.461 621 | 1.34 |
400 | 1.21 ± 0.02 | 0.060 584 ± 0.001 197 | 25.635 712 ± 1.273 489 | 1.61 |
500 | 1.29 ± 0.03 | 0.057 866 ± 0.001 131 | 22.978 039 ± 1.135 976 | 1.86 |
600 | 1.37 ± 0.03 | 0.055 831 ± 0.001 122 | 21.037 396 ± 1.317 849 | ⋯ |
700 | 1.43 ± 0.03 | 0.054 069 ± 0.001 123 | 20.081 535 ± 1.411 917 | 2.42 |
Re . | . | . | (deg) . | . |
---|---|---|---|---|
5 | 1.00 ± 0.00 | 0.447 192 ± 0.000 419 | 45.002 058 ± 0.095 827 | 0.96 |
100 | 1.00 ± 0.00 | 0.099 973 ± 0.000 224 | 45.010 85 ± 0.229 41 | 0.96 |
200 | 1.00 ± 0.00 | 0.070 693 ± 0.000 158 | 45.010 364 ± 0.229 03 | 0.97 |
300 | 1.13 ± 0.02 | 0.064 989 ± 0.001 234 | 29.350 199 ± 1.461 621 | 1.34 |
400 | 1.21 ± 0.02 | 0.060 584 ± 0.001 197 | 25.635 712 ± 1.273 489 | 1.61 |
500 | 1.29 ± 0.03 | 0.057 866 ± 0.001 131 | 22.978 039 ± 1.135 976 | 1.86 |
600 | 1.37 ± 0.03 | 0.055 831 ± 0.001 122 | 21.037 396 ± 1.317 849 | ⋯ |
700 | 1.43 ± 0.03 | 0.054 069 ± 0.001 123 | 20.081 535 ± 1.411 917 | 2.42 |
Figure 4 shows the total kinetic energy, Ekin, for Re = 500 at different values of θ. We observe that the total kinetic energies reach a steady state for most cases. It implies that the time-averaged change in the kinetic energy is zero, , and the system is in equilibrium state. According to Eq. (5), we can use the viscous power to approximate the dissipation. Although we can compute the dissipation directly from our model, the reason to use is two-fold: (1) it is consistent with the conventional torque approach to compute the dissipation, and (2) we want to apply the turbulence model that is associated with the surface shear stress (i.e., u*). A figure showing the time evolution of the three quantities (, , and ) for the case at Re = 500 and θ = 0 is given in the supplementary material (see Fig. S3), verifying the energy balance equation [Eq. (5)]. In Fig. 4, a longer transient period is observed at θ = 50°, but the flow eventually reaches a steady state. At θ = 60°, we see that Ekin increases without reaching saturation within the default simulation time t⊕ = 5. This phenomenon is directly related to the well-known anomalous behavior of the Ekman boundary layer thickness at the so-called critical latitude, which in our case is precisely at θ = 60° [see, e.g., Sec. 8.07.1.4 in the study by Tilgner (2015)].
Figures 5 and 6 show the viscous power per unit area for the cases θ = 50° and 60° with longer simulation time. The time-averaged value over four oscillation periods is plotted in dots. The orange dot indicates the value obtained in the default range, 1 ≤ t⊕ ≤ 5. As can be seen, for θ = 50°, the power reaches a steady value within t⊕ ≤ 5, suggesting that t⊕ = 5 is sufficient for most cases. On the other hand, θ = 60° requires a longer transient before reaching a steady value, which approximates the laminar solution. This finding suggests that we should replace the value obtained in the default range by its laminar solution when computing the total dissipation.
B. Mean velocity profiles
Figure 7 shows the horizontal velocity at different Reynolds numbers with θ = 0 and t⊕ = 5. The overbar indicates the spatial average on the horizontal plane, and the prime symbol denotes the difference with respect to the boundary velocity, i.e., . We adopt this notation in order to compare our results to the law of wall formalism [see the study by Tennekes (1972)]. According to the formalism, the velocity profile is divided into three parts: a viscous sublayer Q+ = |z+|, a buffer layer, and an inertial sublayer Q+ = (1/0.41) log|z+| + 5. The inertial sublayer is also known as the logarithmic region. As shown in Fig. 7, the velocity profiles approach the one predicted by the law of the wall when Re ≥ 300. A constant slope reveals the presence of a logarithmic region. The inset of Fig. 7 shows the Kármán measure (i.e., the reciprocal of the slope) for Re ≥ 300. A clear logarithmic region develops after Re ≈ 300. In the supplementary material, we show the velocity profiles at different time instants for the case Re = 300 and find that the appearance of a logarithmic layer coincides with an appreciable deviation of the dissipation from its laminar value, as expected from the presence of turbulence. Besides, the velocity profiles show a maximum above the logarithmic region, which is usually referred to as a low-level jet and is also observed for a steady Ekman layer (Deusebio , 2014). The fact that the maximum is not well represented in the case of Re = 700 implies that the domain height should be increased accordingly if we want to extend the Re range.
C. Similarity theory
D. Viscous dissipation
Figure 10 shows the viscous power per unit area at each co-latitude θ, which we compute by taking a time-average in the default range, 1 ≤ t⊕ ≤ 5, except for the case Re = 300 (see Sec. IV A). Note that we multiply the power by Re to remove the viscosity dependence and to see the effect of turbulence. As it can be seen, when Re ≥ 300, the power increases at almost every θ compared to the laminar solution, while the models with Re ≤ 200 remain close to the laminar solution. Given Re ≥ 300, the increments are not the same across θ. No increment is observed at the equator and is barely present near the critical latitude. As shown in Sec. IV A, the small increment at the critical latitude is the consequence of insufficient simulation time; the power will eventually reach the laminar solution, given long enough simulation time. This latitude variation can be attributed to the orientation of the rotation vector of the fluid, which goes from perpendicular to the flow at the pole to parallel at the equator. When parallel, there is no Coriolis force, and the Ekman boundary layer flow will reduce to the well-studied Stokes type, whose critical Reynolds number for the fully turbulent regime Re = 2500 [see, e.g., the study by Ozdemir (2014)] is higher than the present setting.
V. DISCUSSION AND CONCLUSIONS
We have derived the governing equation for the boundary layer generated by the precession-driven flow in a spherical cavity using a local Cartesian coordinate system. Both the order of magnitude estimation and the numerical solution suggest that the Poincaré term is negligible in the boundary layer if precession is weak. In order to solve the governing equation numerically, we constructed a local Cartesian box model to simulate the boundary layer flow. In the laminar regime, we found that the total energy dissipation, which is computed from the local contributions obtained from the individual box models, is consistent with the laminar value from previous studies based on a global model [e.g., Cébron (2019)]. In the turbulent regime, we have found that (1) the flow at certain latitudes (i.e., θ = 60° and 90°) may remain laminar and (2) the total energy dissipation deviates appreciably from the laminar value. For the flow at the pole, θ = 0, the energy dissipation begins to deviate from the analytical laminar expectation when Re ≥ 300, and the flow exhibits turbulent characteristics, such as a logarithmic layer region, allowing us to use an asymptotic similarity formalism. Using this, we have computed the asymptotic curves for both u* and β as a function of Re, which can be used to compute the total energy dissipation in the turbulent regime via turbulence models.
In the conventional turbulence model, the total turbulent dissipation is given as −(45π/32)kIΔΩ3, where parameter k is supposed to quantify the asymptotic behavior of the local dissipation in the turbulent regime at high Re. Note that the derivation of the model implicitly assumes that k is independent of latitude. Following the same approach as that followed by Cébron (2019) to estimate k via , we have found that both (u*/U) and β at high Re can be computed from the best fit asymptotic curve obtained at θ = 0. In order to recover the laminar solution [Eq. (6)] when Re goes into the laminar regime, we propose a turbulence model by writing the total dissipation as Eq. (7). The overestimation of the turbulence model, compared to the value obtained from the numerical solution, suggests that the assumption on k may be inappropriate. At least for the case of Re = 500, our numerical solution shows that the flow is not even turbulent at certain latitudes. A thorough investigation on the latitudinal variation on k combined with an investigation of the flow asymptotic behavior at each latitude would be helpful to establish a more reliable turbulence model. That aside, the uncertainty of the turbulence model given by Eq. (7) is about 15% in the Re ≤ 700 range.
The previous study by Deusebio (2014) shows that the asymptotic theory applied to the intermediate Re range (400 < Re < 775), where a logarithmic layer is observed, is sufficient to get a good prediction of u* and β at high Re (=1600). We adopt the theory for an oscillatory Ekman layer at θ = 0 not only because we see the logarithmic layer in the flow (Sec. IV B) but also because the theory should be generally unaffected by imposing a time varying boundary velocity. The theory developed by Csanady (1967) and Spalart (1989) is based on the assumption that there exists an overlapping region where the velocity profiles in both the inner layer (viscous sublayer) and the outer layer (free-stream region) match, a fact unaffected by imposing time dependence.
Application to the earth’s core is possible since the numerical results presented here are based on the reduced form of the governing equation [Eq. (4)], which is identical to the one derived for a spheroidal cavity such as Earth by Buffett (2021). We are able to reach the earth’s parameter regime with the local Cartesian box model, so we do not need to rely on any turbulence model, which will be discussed in the following. Our numerical result shows that the total energy dissipation in the turbulent regime at Re = 500 is 1.86 times of the laminar value. If the damping of FCN shares the same increasing factor, then the effective viscosity will increase by a factor of 3.5, which is still far below the required amount (6 orders of magnitude higher) inferred from the observation. It implies that the electromagnetic coupling may play a significant role in closing the gap.
An accurate calculation of the energy dissipation at the CMB in the moon is unfeasible since we not only lack a proper turbulence model but also because interior models obtained from seismic and gravity measurements are poorly constrained. There are three relevant parameters: (1) the angle between the rotation axes of the core and the mantle, (2) the lunar core radius, and (3) the moment of inertia of the fluid core. There is no direct measurement for the orientation of the core rotation axis. According to the dynamic argument that the lunar CMB flattening is smaller than its Poincaré number, the core is unlikely to share the precession of the mantle (Goldreich, 1967); the core rotation axis should thus be closely aligned with the ecliptic normal. We use the reduced model by Cébron (2019) to compute the orientation of Ωf (see Table I). The plausible range of the core radius obtained from gravity measurements is 200–380 km (Williams , 2014; Viswanathan , 2019). If we take the core radius as 380 km, then the relative velocity at CMB will be around 2.72 × 10−2 m s−1, and the Reynolds number will be 1.67 × 104. We obtain the moment of inertia of the fluid core by taking the moment ratio Cf/C = 7 × 10−4 (Williams , 2014). With our best fit parameters (A, B, C5) and the reduced turbulence model [Eq. (11)], we obtain or 84 MW for χ = 0.85 or 1.0, respectively. For a smaller core, say Re = 200 km, with the same moment ratio, the Reynolds number reduces to 8.79 × 103 while the dissipation increases by about 10%. For comparison, we also compute the dissipation with the parameters (A = 7.32, B = −2.67) (Sous , 2013, the conversion to the present notation is presented in Appendix C) and the conventional turbulence model, −(45π/32)kIΔΩ3 [see Sec. II; e.g., Cébron (2019)]; the obtained value is 126 MW, about 60% larger than our result. According to the analysis of lunar laser ranging (LLR) data, the dissipation at CMB is about 58 MW (Williams , 2001), 84 MW (Williams , 2014), and 73 MW (Williams and Boggs, 2015), respectively. Concerning the uncertainty of the interior model, it is still insufficient to say if our model matches the observation or not, but at least we provide a set of parameters with numerical support, which can be used for the turbulence model. A sensitivity test on the parameters (A, B, C5) shows that (1) β is sensitive to A and also sensitive to C5 when Re < 103 and (2) the sensitivity of u*/U to both A and B decreases as Re increases (see Figs. S7–S9 in the supplementary material).
The implication for planetary cores with turbulent boundary layers should be interpreted with care. In this study, the quoted Re for Earth and the Moon is obtained based on the laminar solution [e.g., Busse (1968) and Cébron (2019)], where the boundary layer is treated in the laminar regime, even though the obtained Re suggests the other way around. For Earth, the conclusion that the turbulence effect is unlikely to account for the observed FCN damping remains as Earth is at the margin of turbulence. For the Moon, our purpose is to demonstrate a way to forward-compute the dissipation, and a more comprehensive model is needed for future exploration.
We aim to demonstrate that through appropriate formulation, it is feasible to complement the results from global models. By narrowing our focus to the boundary layer exclusively, we can expand the parameter range and explore the turbulent regime. Potential applications of this approach include spin-up, libration, and precessing cylinders (Pizzi , 2021).
SUPPLEMENTARY MATERIAL
See the supplementary material for the figures of friction velocity showing the comparison between the solutions obtained from Eqs. (2) and (4), velocity profiles for the domain height test, energy balance for the case of Re = 500 and θ = 0, velocity profiles of Re = 300 and θ = 0 at different time instants, u* and β as a function of Re with the comparison to the result of Sous et al. (2013), and sensitivity test on (A, B, C5), β as a function of θ.
ACKNOWLEDGMENTS
We are grateful for the comments and suggestions by all the anonymous reviewers. S.-A.S. would like to thank Benjamin Favier and Philipp Schlatter for the helpful discussion on the numerical methods. The numerical solutions are obtained using NEK5000 Version 20.0-dev at Argonne National Laboratory, Illinois. (available at https://nek5000.mcs.anl.gov). S.-A.S. is a Research Fellow of the F.R.S.-FNRS. V.D., J.R., and S.A.T. acknowledge the funding from the European Research Council (ERC) under the European Union’s Horizon 2020 Research and Innovation Program (GRACEFUL Synergy Grant Agreement No. 855677).
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Sheng-An Shih (施勝安): Conceptualization (equal); Data curation (lead); Formal analysis (lead); Methodology (lead); Validation (lead); Visualization (lead); Writing – original draft (lead); Writing – review & editing (equal). Santiago Andrés Triana: Conceptualization (equal); Methodology (supporting); Supervision (equal); Writing – review & editing (equal). Jérémy Rekier: Conceptualization (equal); Writing – review & editing (equal). Véronique Dehant: Conceptualization (equal); Funding acquisition (equal); Supervision (equal); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.