Skip to main content
Log in

Analysis of Schwarz waveform relaxation for the coupled Ekman boundary layer problem with continuously variable coefficients

  • Original Paper
  • Published:
Numerical Algorithms Aims and scope Submit manuscript

Abstract

In this paper, we present a global-in-time non-overlapping Schwarz method applied to the Ekman boundary layer problem. Such a coupled problem is representative of large-scale atmospheric and oceanic flows in the vicinity of the air-sea interface. Schwarz waveform relaxation (SWR) algorithms provide attractive methods for ensuring a “tight coupling” between the ocean and the atmosphere. However, the convergence study of such algorithms in this context raises a number of challenges. Numerous convergence studies of Schwarz methods have been carried out in idealized settings, but the underlying assumptions to make these studies tractable may prohibit them to be directly extended to the complexity of climate models. We illustrate this aspect on the coupled Ekman problem, which includes several essential features inherent to climate modeling while being simple enough for analytical results to be derived. We investigate its well-posedness and derive an appropriate SWR algorithm. Sufficient conditions for ensuring its convergence for different viscosity profiles are then established. Finally, we illustrate the relevance of our theoretical analysis with numerical results and suggest ways to improve the computational cost of the coupling. Our study emphasizes the fact that the convergence properties can be highly sensitive to some model characteristics such as the geometry of the problem and the use of continuously variable viscosity coefficients.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

References

  1. Brassington, G.B., Martin, M.J., Tolman, H.L., Akella, S., Balmeseda, M., Chambers, C.R.S., Chassignet, E., Cummings, J.A., Drillet, Y., Jansen, P.AEM, Laloyaux, P., Lea, D., Mehra, A., Mirouze, I., Ritchie, H., Samson, G., Sandery, P.A., Smith, G.C., Suarez, M., Todling, R.: Progress and challenges in short- to medium-range coupled prediction. J. Oper. Oceanogr. 8, s239–s258 (2015)

    Google Scholar 

  2. Smith, G.C., Blanger, J.-M., Roy, F., Pellerin, P., Ritchie, H., Onu, K., Roch, M., Zadra, A., Colan, D.S, Winter, B, Fontecilla, J.-S., Deacu, D.: Impact of coupling with an ice-ocean model on global medium-range NWP forecast skill. Mon. Weather Rev. 146(4), 1157–1180 (2018)

    Article  Google Scholar 

  3. Lemarié, F., Blayo, E., Debreu, L.: Analysis of ocean-atmosphere coupling algorithms: Consistency and stability. Procedia Comput. Sci. 51(Supplement C), 2066–2075 (2015). https://doi.org/10.1016/j.procs.2015.05.473. iCCS 2015

    Article  Google Scholar 

  4. Beljaars, A., Dutra, E., Balsamo, G., Lemarié, F.: On the numerical stability of surface–atmosphere coupling in weather and climate models. Geosci. Model Dev. 10(2), 977–989 (2017)

    Article  Google Scholar 

  5. Keyes, D.E., et al.: Multiphysics simulations: challenges and opportunities. Int. J. High Perform. Comput. Appl. 27(1), 4–83 (2013)

    Article  Google Scholar 

  6. Gander, M.J., Halpern, L.: Optimized Schwarz waveform relaxation methods for advection reaction diffusion problems. SIAM J. Numer. Anal. 45(2) (2007)

  7. Blayo, E., Halpern, L., Japhet, C.: Optimized Schwarz waveform relaxation algorithms with nonconforming time discretization for coupling convection-diffusion problems with discontinuous coefficients. In: Domain decomposition methods in science and engineering XVI, Lect. Notes Comput. Sci. Eng., vol 55, pp. 267–274. Springer, Berlin (2007)

  8. Gander, M.J., Jiang, Y.-L., Li, R.-J.: Parareal Schwarz waveform relaxation methods. In: Bank, R., Holst, M., Widlund, O., Xu, J. (eds.) Domain Decomposition Methods in Science and Engineering XX, pp 451–458. Springer, Berlin (2013)

  9. Qaddouri, A., Laayouni, L., Loisel, S., Côté, J., Gander, M.J.: Optimized Schwarz methods with an overset grid for the shallow-water equations: preliminary results. Appl. Numer. Math. 58(4), 459–471 (2008). https://doi.org/10.1016/j.apnum.2007.01.015

    Article  MathSciNet  Google Scholar 

  10. Martin, V.: Schwarz waveform relaxation algorithms for the linear viscous equatorial shallow water equations. SIAM J. Sci. Comput. 31(5), 3595–3625 (2009). https://doi.org/10.1137/070691450

    Article  MathSciNet  Google Scholar 

  11. Audusse, E., Dreyfuss, P., Merlet, B.: Optimized Schwarz waveform relaxation for the primitive equations of the ocean. SIAM J. Sci. Comput. 32(5), 2908–2936 (January 2010). https://doi.org/10.1137/090770059

    Article  MathSciNet  Google Scholar 

  12. Qaddouri, A., Lee, V.: The Canadian global environmental multiscale model on the Yin-Yang grid system. Quart. J. Roy. Meteorol. Soc. 137(660), 1913–1926 (2011). https://doi.org/10.1002/qj.873

    Article  Google Scholar 

  13. Marti, O., Nguyen, S., Braconnot, P., Valcke, S., Lemarié, F., Blayo, E.: A Schwarz iterative method to evaluate ocean- atmosphere coupling schemes. Implementation and diagnostics in IPSL-CM6-SW-VLR. Geosc. Model Dev. Disc., 1–19. https://doi.org/10.5194/gmd-2020-307 (2020)

  14. Monin, A.S., Obukhov, A.M.: Basic laws of turbulent mixing in the surface layer of the atmosphere. Trudy Akademii Nauk SSSR Geofizicheskogo Instituta 24, 163–187 (1954)

    Google Scholar 

  15. O’Brien, J.J.: A note on the vertical structure of the eddy exchange coefficient in the planetary boundary layer. J. Atmos. Sci. 27(8), 1213–1215 (1970)

    Article  Google Scholar 

  16. Ekman, V.W.: On the influence of the Earth’s rotation in ocean-currents. Arch. Math. Astron. Phys. 2, 1–52 (1905)

    MATH  Google Scholar 

  17. Grisogono, B.: A generalized Ekman layer profile with gradually varying eddy diffusivities. Quart. J. Roy. Meteorol. Soc. 121(522), 445–453 (1995)

    Article  Google Scholar 

  18. Bye, J.A.: Inertially coupled Ekman layers. Dynam. Atmos. Oceans 35(1), 27–39 (2002)

    Article  Google Scholar 

  19. Lewis, D.M., Belcher, S.E.: Time-dependent, coupled, Ekman boundary layer solutions incorporating Stokes drift. Dynam. Atmos. Oceans 37(4), 313–351 (2004)

    Article  Google Scholar 

  20. Vallis, G.K.: Atmospheric and oceanic fluid dynamics: fundamentals and large-scale circulation. Cambridge University Press (2006)

  21. Lemarié, F., Debreu, L., Madec, G., Demange, J., Molines, J.M., Honnorat, M.: Stability constraints for oceanic numerical models: implications for the formulation of time and space discretizations. Ocean Modell. 92, 124–148 (2015)

    Article  Google Scholar 

  22. Gander, M.J., Xu, Y.: Optimized Schwarz methods for model problems with continuously variable coefficients. SIAM J. Sci. Comput. 38, A2964–A2986 (2016)

    Article  MathSciNet  Google Scholar 

  23. Lemarié, F., Debreu, L., Blayo, E.: Toward an optimized global-in-time Schwarz algorithm for diffusion equations with discontinuous and spatially variable coefficients, part 2: the variable coefficients case. Electron. Trans. Numer. Anal. 40, 170–186 (2013)

    MathSciNet  MATH  Google Scholar 

  24. Lions, P.-L.: On the Schwarz alternating method. III. A variant for nonoverlapping subdomains. In: Third International Symposium on Domain Decomposition Methods for Partial Differential Equations (Houston,TX, 1989), pp 202–223. SIAM, Philadelphia (1990)

  25. Madsen, O.S.: A realistic model of the wind-induced Ekman boundary layer. J. Phys. Oceanogr. 7(2), 248–255 (1977). https://doi.org/10.1175/1520-0485(1977)007∖(〈∖)0248:ARMOTW∖(〉∖)2.0.CO;2

    Article  Google Scholar 

  26. Lions, JL, Magenes, E: Problèmes aux limites non homogènes et applications, vol. 1. Dunod (1968)

  27. Gander, M.J., Halpern, L., Kern, M.: A Schwarz waveform relaxation method for advection—diffusion—reaction problems with discontinuous coefficients and non-matching grids. In: Widlund, O.B., Keyes, D.E. (eds.) Domain Decomposition Methods in Science and Engineering XVI, pp 283–290. Springer, Berlin (2007)

  28. Lions, JL, Magenes, E: Problèmes aux limites non homogènes et applications, vol. 2. Dunod (1968)

  29. Gander, M.J., Halpern, L., Nataf, F.: Optimal convergence for overlapping and non-overlapping Schwarz waveform relaxation. In: Lai, C.-H., Bjørstad, P., Cross, M., Widlund, O. (eds.) Proceedings of the 11th International Conference on Domain Decomposition Methods (1999)

  30. Gander, M.J., Kwok, F., Mandal, B.C.: Dirichlet-Neumann and Neumann-Neumann waveform relaxation for the wave equation. In: Dickopf, T., Gander, M.J., Halpern, L., Krause, R., Pavarino, L.F. (eds.) Domain Decomposition Methods in Science and Engineering XXII, pp 501–509. Springer International Publishing, Cham (2016)

  31. Bennequin, D., Gander, M.J., Halpern, L.: Optimized Schwarz waveform relaxation methods for convection reaction diffusion problems. Technical Report, 2004-24, LAGA, Université Paris 13 (2004)

  32. Thery, S.: Etude numérique des algorithmes de couplage océan-atmosphère avec prise en compte des paramv́etrisations physiques des couches limites. Ph.D. Thesis, Université Grenoble-Alpes. http://www.theses.fr/s184608 (2021)

  33. Olver, F.W.J: Hypergeometric function. In: NIST handbook of mathematical functions hardback and CD-ROM. Cambridge University Press. http://dlmf.nist.gov/15(2010)

  34. Gander, M.J., Halpern, L.: Méthodes de relaxation d’ondes (SWR) pour l’équation de la chaleur en dimension 1. Compt. Rend. Math. 336(6), 519–524 (2003)

    Article  Google Scholar 

  35. Lemarié, F., Debreu, L., Blayo, E.: Toward an optimized global-in-time Schwarz algorithm for diffusion equations with discontinuous and spatially variable coefficients, part 1: the constant coefficients case. Electron. Trans. Numer. Anal. 40, 148–169 (2013)

    MathSciNet  MATH  Google Scholar 

  36. Gander, M.J., Kwok, F., Mandal, B.C.: Dirichlet-neumann and neumann-neumann waveform relaxation algorithms for parabolic problems. Electron. Trans. Numer. Anal. 45, 424–456 (2016)

    MathSciNet  MATH  Google Scholar 

  37. Gander, M.J., Kwok, F., Mandal, B.C.: Dirichlet-Neumann waveform relaxation methods for parabolic and hyperbolic problems in multiple subdomains. BIT Numer. Math. https://doi.org/10.1007/s10543-020-00823-2 (2020)

  38. Dunster, T.M.: Legendre and related functions. In: NIST handbook of mathematical functions hardback and CD-ROM. In: Cambridge University Press. http://dlmf.nist.gov/14 (2010)

  39. Dubois, O., Gander, M.J., Loisel, S., A., A. S.-C., Daniel, B.: The optimized Schwarz method with a coarse grid correction. SIAM J. Sci. Comput. 34, A421–A458 (2012)

    Article  MathSciNet  Google Scholar 

Download references

Funding

The authors acknowledge the support of the French national research agency (ANR) through contract ANR-16-CE01-0007.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sophie Thery.

Additional information

Publisher’s note

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

Appendices

Appendix:

Quadratic viscosity basis functions

Equation (12a) using the quadratic viscosity ν(z) = c + bz + az2 for zΩ, can be reduced to the Legendre ODE \(\displaystyle \left (1 - \eta ^{2} \right ) \frac {\textup {d}^{2} y}{\textup {d} \eta ^{2}} - 2 \eta \frac {\textup {d} y }{\textup {d} \eta } + \xi \left (\xi + 1 \right ) y = 0 \) with

$$ \eta\left( z \right) = \frac{2a}{\sqrt{b^{2}-4ac}} \left( z+\frac{b}{2a} \right) \quad\text{ and}\quad \xi = -\frac{1}{2} \left( 1 \pm \sqrt{1+ \frac{4 \imath \left( f + \omega \right)}{a}} \right), $$
(64)

where the choice of the sign in (64) does not bear any consequence. In particular, one can show that using relevant quadratic viscosity functions, (e.g. satisfying ν(z) > 0 for all zΩ and \((a,b,c) \in \mathbb {R}_{-}^{*} \times \mathbb {R}^{2}\)), \(\eta \in \mathbb {R}\) and satisfies − 1 < η(z) < 1 for zΩ. The functions given by (32a) are then numerically satisfactory solutions to the Legendre ODE [meaning that computing a reasonable amount of sum terms leads to accurate values, [38]], and thus to (12a). In our numerical results, Legendre functions have been computed using the hypergeometric function 2F1 [33] through: \(\displaystyle P_{\xi }^{0} \left (\eta \right ) = {}_{2} F_{1}\left (\xi +1, -\xi , 1; \eta \right )\).

On the observed discretized convergence factor

Let us consider the numerical solution of the coupled problem (3a) on a given time window \(\mathcal {T}_{n}\) using a finite difference discretization in time. The observed error can be written as (with Uj = (uj,vj)):

$$ \mathbf{E}_{j}^{k,m}(z) = \mathbf{U}_{j}^{k,m}(z) - \mathbf{U}_{j}(z,{t^{m}_{j}}) 1 \leq m \leq M_{j}, j \in \{1,2\}, k \in \mathbb{N}^{*}, $$
(65)

where m is an index for the time step, \({t^{m}_{j}}\) is the physical time corresponding to time step m, and Mj is the total number of time steps for the numerical approximation of Uj in \(\mathcal {T}_{n}\). A convergence rate characterizing the behaviour of \(\mathbf {E}_{j}^{k,m} = (E_{u,j}^{k,m},E_{v,j}^{k,m} )\) as a function of k is:

$$ \mathcal{R}_{\mathbf{U}_{j}}^{k} = \frac{ \left\| \left( \mathbf{E}_{j}^{k,m}(0^{\mp})\right)_{m} \right \|_{2}} { \left\| \left( \mathbf{E}_{j}^{k-1,m}(0^{\mp})\right)_{m} \right \|_{2}} = \frac{\left[ \underset{m=1}{\overset{M_{j}}{\sum}} \left( \left( {E}_{u,j}^{k,m}(0^{\mp})\right)^{2} + \left( {E}_{v,j}^{k,m}(0^{\mp})\right)^{2} \right) \right]^{1/2}} {\left[ \underset{m=1}{\overset{M_{j}}{\sum}} \left( \left( {E}_{u,j}^{k-1,m}(0^{\mp})\right)^{2} + \left( {E}_{v,j}^{k-1,m}(0^{\mp})\right)^{2} \right) \right]^{1/2}}, k \in \mathbb{N}^{*}, j \in \{1,2\}. $$
(66)

Also, the discrete errors on the complex variables \(\left (\varphi ,\overline {\varphi }\right )\) are:

$$ \mathbf{e}_{j}^{k,m}(z) = \left( \begin{array}{c} e_{\varphi,j}^{k,m}(z) \\ e_{\overline{\varphi},j}^{k,m}(z) \end{array} \right) = \left( \begin{array}{c} \varphi_{j}^{k,m}(z) - \varphi(z,{t_{m}^{j}})\\ \overline{\varphi}_{j}^{k,m}(z) - \overline{\varphi}(z,{t_{m}^{j}}) \end{array} \right) , $$
(67)

where \(\varphi _{j}^{k,m}\) and \(\overline {\varphi }_{j}^{k,m}\) are defined by the Schwarz algorithm (2.3). \(\mathbf {E}_{j}^{k,m}\) is linked to \(\mathbf {E}_{j}^{k,m}\) by:

$$ \textbf{E}^{k,m}_{j} = \mathbf{P} \mathbf{e}_{j}^{k,m} = \frac{1}{\sqrt{2}}\begin{pmatrix} e_{\varphi,j}^{k,m} + e_{\overline{\varphi},j}^{k,m} \\ - \imath e_{\varphi,j}^{k,m} + \imath e_{\overline{\varphi},j}^{k,m} \end{pmatrix}, $$
(68)

where P is given in Section 2.2. Since P is a unitary matrix (i.e. \(\mathbf {P}^{-1}=\overline {\mathbf {P}}^{T}\)),

$$ \left\| \textbf{E}^{k,m}_{j} \right\|_{2}^{2} = \left( \overline{\mathbf{e}}_{j}^{k,m}\right)^{T} \overline{\mathbf{P}}^{T} \mathbf{P} \mathbf{e}_{j}^{k,m} = \left( \overline{\mathbf{e}}_{j}^{k,m}\right)^{T} \mathbf{e}_{j}^{k,m} = \left\| \mathbf{e}_{j}^{k,m} \right\|_{2}^{2}, $$
(69)

meaning that (66) becomes

$$ \mathcal{R}_{\mathbf{U}_{j}}^{k} = \frac{\left[ \underset{m=1}{\overset{M_{j}}{\sum}} \left( \left( {e}_{\varphi,j}^{k,m}(0^{\mp})\right)^{2} + \left( {e}_{\overline{\varphi},j}^{k,m}(0^{\mp})\right)^{2} \right) \right]^{1/2}} {\left[ \underset{m=1}{\overset{M_{j}}{\sum}} \left( \left( {e}_{\varphi,j}^{k-1,m}(0^{\mp})\right)^{2} + \left( {e}_{\overline{\varphi},j}^{k-1,m}(0^{\mp})\right)^{2} \right) \right]^{1/2}}, k \in \mathbb{N}^{*}, j \in \{1,2\}. $$
(70)

Let us now define \( \widetilde {\rho }_{\varphi ,j}^{k,M_{j}} = \frac { \left \| \left (e_{\varphi ,j}^{k,m}(z=0^{\mp })\right ) \right \|_{2}} { \left \| \left (e_{\varphi ,j}^{k-1,m}(z=0^{\mp })\right ) \right \|_{2}} = \frac {\left [ \underset {m=1}{\overset {M_{j}}{\sum }} \left ({e}_{\varphi ,j}^{k,m}(0^{\mp })\right )^{2} \right ]^{1/2}} {\left [ \underset {m=1}{\overset {M_{j}}{\sum }} \left ({e}_{\varphi ,j}^{k-1,m}(0^{\mp })\right )^{2} \right ]^{1/2}}\) for \(k \in \mathbb {N}^{*}, j \in \{1,2\}\). When \(M_{j}\rightarrow +\infty \) (i.e. simulating (2.3) on a infinite time window), the temporal grid can only generate modes which frequencies lie in \(I_{\omega } := \left [ -\frac {\pi }{ \underset {j \in \{1,2\}}{\textup {min}} \left \{ \delta t_{j} \right \}} ;~ \frac {\pi }{ \underset {j \in \{1,2\}}{\textup {min}} \left \{ \delta t_{j} \right \}} \right ]\). This is due to the Nyquist-Shannon sampling theorem [39]. \(\widehat {\left ({e}_{\varphi ,j}^{k,m}(0^{\mp }) \right )}\) corresponds to the Fourier transform of the continuous error \(\hat {e}_{\varphi }^{k}\) on Iω. The convergence factor denoted \(\widetilde {\rho }_{\varphi ,j}^{k,\infty }\) can be linked to the errors in the Fourier space by the Parseval’s theorem:

$$ \widetilde{\rho}_{\varphi,j}^{k,\infty} = \frac{ \left[ \underset{m=1}{\overset{\infty}{\sum}} ({e}_{\varphi,j}^{k,m}(0^{\mp})^{2} \right]^{1/2} } { \left[ \underset{m=1}{\overset{\infty}{\sum}} ({e}_{\varphi,j}^{k-1,m}(0^{\mp})^{2} \right]^{1/2} } = \frac{\left( {\int}_{\omega\in I} \left( \rho_{\varphi,j}^{k}(\omega)\right)^{2} \left|\hat{e}_{\varphi,j}^{k-1}(\omega)\right|^{2} d\omega\right)^{1/2}}{\left( {\int}_{\omega\in I} \left|\hat{e}_{\varphi,j}^{k-1}(\omega)\right|^{2} d\omega\right)^{1/2}}, $$

which implies that \(\underset {|\omega | \leq \pi / \underset {j \in \{1,2\}}{\textup {min}} \{\delta t_{j}\}} {\textup {inf}}~\rho _{\varphi ,j}^{k}(\omega ), \le \widetilde {\rho }_{\varphi ,j}^{k,\infty } \le \underset {|\omega | \leq \pi / \underset {j \in \{1,2\}}{\textup {min}} \{\delta t_{j}\}} {\textup {sup}}~\rho _{\varphi ,j}^{k}(\omega )\). Moreover, using the identity \(\displaystyle \underset {m=1}{\overset {M_{j}}{\sum }} x_{m} = \left (\underset {m=1}{\overset {\infty }{\sum }} x_{m} \right ) \left (1 - \left \{ \underset {m=M_{j}+1}{\overset {\infty }{\sum }}x_{m} \right \} / \left \{ \underset {m=1}{\overset {\infty }{\sum }}x_{m} \right \} \right )\), \(\widetilde {\rho }_{\varphi ,j}^{k,M_{j}}\) can easily be linked to \(\widetilde {\rho }_{\varphi ,j}^{k,\infty }\):

$$ \left( \widetilde{\rho}_{\varphi,j}^{k,M_{j}} \right)^{2} = \left( \widetilde{\rho}_{\varphi,j}^{k,\infty} \right)^{2} \frac{1-{{{\varGamma}}}_{\varphi,j}^{k,M_{j}}}{1-{{{\varGamma}}}_{\varphi,j}^{k-1,M_{j}}} \qquad\text{with } {{{\varGamma}}}_{\varphi,j}^{k,M_{j}}=\frac{\underset{m=M_{j}+1}{\overset{\infty}{\sum}} \left( e_{\varphi,j}^{k,m}(0^{\mp})\right)^{2}}{\underset{m=1}{\overset{\infty}{\sum}}\left( e_{\varphi,j}^{k,m}(0^{\mp})\right)^{2}}. $$
(71)

Since \({{{\varGamma }}}_{\varphi ,j}^{k,M_{j}}\longrightarrow 0\) as \(M_{j}\rightarrow \infty \) (thanks to Section 2.1), (71) can be rewritten as \(\displaystyle \widetilde {\rho }_{\varphi ,j}^{k,M_{j}} = \widetilde {\rho }_{\varphi ,j}^{k,\infty } + \mathcal {O}\left ({{{\varGamma }}}_{\varphi ,j}^{k,M_{j}}, {{{\varGamma }}}_{\varphi ,j}^{k-1,M_{j}}\right ) \). Finally, the same relationships being true for \(e_{\overline {\varphi },j}^{k,m}\), we end up with:

$$ \mathcal{R}_{\mathbf{U}_{j}}^{k} \leq \max \left\{ \underset{|\omega| \leq \pi / \underset{j \in \{1,2\}}{\textup{min}} \{\delta t_{j}\}} {\textup{sup}}~\rho_{\varphi,j}^{k}(\omega), ~\underset{|\omega| \leq \pi / \underset{j \in \{1,2\}}{\textup{min}} \{\delta t_{j}\}} {\textup{sup}}~\rho_{\overline{\varphi},j}^{k}(\omega) \right \} + \epsilon . $$
(72)

Proof for the upper bound (47)

In this Appendix, we provide the proof for the upper bound (47) on \(\rho _{\text {DN}}^{\text {cst}}\) for constant viscosity and Dirichlet-Neumann interface conditions given in Section 4.3. We first reformulate \(\rho _{\text {DN}}^{\text {cst}}\) as \(\rho _{\text {DN}}^{\text {cst}} = \sqrt {\lambda }\sqrt {\frac { Q(\sqrt {2\text {Fo}_{2}})}{ Q(\sqrt {2\text {Fo}_{1}})}}\) with

$$ \mathcal{Q}(x) = \frac{ \cosh x -\cos x }{ \cosh x + \cos x }. $$

4.1 Proof for the upper bound (47) for Fo2 < Fo1

The derivative of \(\mathcal {Q}(x)\) cancels for x such that \(\tanh (x^{\star }) + \tan (x^{\star }) = 0\). This transcendental equation has no root on [0,π/2] (and \(0 \le \mathcal {Q}(x) \le 1\) on that interval) and then one root x(⋆,k) per interval [π/2 + kπ, 3π/2 + kπ] (k ≥ 0). For even values of k, \(\mathcal {Q}(x^{(\star ,k)})\) is a local maximum and \(\mathcal {Q}(x) \ge 1\), while for odd values of k, \(\mathcal {Q}(x^{(\star ,k)})\) is a local minimum and \(\mathcal {Q}(x) \le 1\). Moreover, it can be shown that \(\mathcal {Q}(x^{(\star ,0)}) > \mathcal {Q}(x^{(\star ,2)}) > \mathcal {Q}(x^{(\star ,4)}) > \hdots > 1\) and \(\mathcal {Q}(x^{(\star ,1)}) < \mathcal {Q}(x^{(\star ,3)}) < \mathcal { Q}(x^{(\star ,5)}) < \hdots < 1\). This implies that the maximum of \(\mathcal {Q}(x)\) is \(\mathcal {Q}(x^{(\star ,0)})\) and thus that

$$ \mathcal{Q}(x) \le \mathcal{Q}(x^{(\star,0)}), \qquad x \ge 0. $$

If we now introduce \(\alpha = \sqrt {\text {Fo}_{2} / \text {Fo}_{1}}\) and consider \(x=\sqrt {2\text {Fo}_{1}}\), our problem is to find the upper bound of \(\frac {\mathcal {Q}(\alpha x)}{ \mathcal {Q}(x) }\) with α ≤ 1 (because we consider the case Fo2 < Fo1 here). It is straightforward to prove that \(\displaystyle \mathcal {Q}(\alpha x) \le \mathcal {Q}(x^{(\star ,0)})\), for x ≥ 0,α ≤ 1 and \(\mathcal {Q}(\alpha x)\) behaves like \(\mathcal {Q}(x)\) on the subintervals \([\frac {\pi }{2\alpha } + k \pi , \frac {3 \pi }{2\alpha } + k \pi ]\). We can then proceed sub-interval by sub-interval: (i) on [0,π/2] \(\mathcal {Q}(x)\) increases more rapidly than \(\mathcal {Q}(\alpha x)\). We thus have \(\frac {\mathcal {Q}(\alpha x)}{ \mathcal {Q}(x) } \le 1\) as soon as α ≤ 1; (ii) on [π/2, 3π/2] \(\mathcal {Q}(x)\) is larger than 1, meaning that \(\frac {1}{\mathcal {Q}(x)} \le 1\). We thus have \(\frac {\mathcal {Q}(\alpha x)}{ \mathcal {Q}(x) } \le \mathcal {Q}(x^{(\star ,0)})\) whatever the value of α; (iii) on [3π/2, 5π/2] \(\mathcal {Q}(x)\) has a local minimum \(\mathcal {Q}(x^{(\star ,1)})\) which is smaller than 1. We thus have \(\frac {1}{\mathcal {Q}(x)} \le \frac {1}{\mathcal {Q}(x^{(\star ,1)})}\) and the worst case scenario is when this coincides with values of α such that the maximum of \(\mathcal {Q}(\alpha x)\) occurs for x ∈ [3π/2, 5π/2]. We thus obtain:

$$ \frac{\mathcal{Q}(\alpha x) }{\mathcal{Q}(x) } \le \frac{\mathcal{Q}(x^{(\star,0)})}{\mathcal{Q}(x^{(\star,1)})}. $$

Because of the ordering of the values \(\mathcal {Q}(x^{(\star ,k)})\) given earlier (i.e. \(\mathcal {Q}(x^{(\star ,k)})\) gets closer and closer to 1 as k increases), it is not needed to go beyond x = 5π/2 since the upper bound could not be larger for x > 5π/2.

4.2 Proof for the upper bound (47) for Fo2 ≥Fo1

For this proof, we first study the behaviour of \(\frac {\mathcal {Q}(x)}{x^{2}}\). The sign of the derivative of \(\frac {\mathcal {Q}(x) }{x^{2}}\) with respect to x is the same as

$$ \cos^{2}(x)-\cosh^{2}(x)+x(\cosh(x)\sin(x)+\sinh(x)\cos(x)), $$

which is smaller or equal to zero for x ≥ 0, \(\frac {\mathcal {Q}(x) }{x^{2}}\) is thus a decreasing function of x for x ≥ 0. In the case α ≥ 1 (i.e. Fo2 ≥Fo1) we have αxx, hence \(\displaystyle \frac {\mathcal {Q}(\alpha x) }{\alpha ^{2}x^{2}}\le \frac {\mathcal {Q}(x) }{x^{2}}\) which leads to \(\displaystyle \frac {\mathcal {Q}(\alpha x) }{\mathcal {Q}(x)}\le \alpha ^{2} \) where \(x =\sqrt {2\text {Fo}_{1}}\) and \(\alpha =\sqrt {\text {Fo}_{2}/\text {Fo}_{1}}\). We can rewrite this last inequality in terms of Foj to obtain

$$ \rho_{\text{DN}}^{\text{cst}}\le \sqrt{\lambda\frac{\text{Fo}_{2}}{\text{Fo}_{1}}}. $$
(73)

Note that this also proves that \(\left |\mathcal {S}_{2}^{\text {cst}}(\text {Fo})\right |\) decreases, and thus \(\left |\mathcal {S}_{2}^{\text {cst}}(\text {Fo})\right |\le \left |\mathcal {S}_{2}^{\text {sta, cst}}\right |=|h_{2}|\). Similarly, \(\left |\mathcal {S}_{1}^{\text {cst}}(\text {Fo})\right |\) increases and \(\left |\mathcal {S}_{1}^{\text {cst}}(\text {Fo})\right |\ge \left |\mathcal {S}_{1}^{\text {sta, cst}}\right |=|h_{1}|^{-1}\).

Details on sufficient conditions for RR-based SWR algorithms

5.1 Proofs of Props. 2 and 3

These proofs are based upon the following lemma:

Lemma 1

If there exists \((\gamma ^{\nu },~\sigma _{1}^{\nu },~{\Sigma }_{2}^{\nu }) \in \mathbb {R}_{+}^{3}\) such that, \(\forall \omega \in \mathbb {R}\):

$$ \begin{array}{@{}rcl@{}} \forall j \in \{1,2\},~ \left|\text{Im}\left( \mathcal{S}_{j}(\omega)\right)\right| &\le& \gamma^{\nu} \left|\text{Re}\left( \mathcal{S}_{j}(\omega)\right)\right| , \end{array} $$
(74a)
$$ \begin{array}{@{}rcl@{}} \left|\text{Re}\left( \mathcal{S}_{1}(\omega)\right)\right|&\ge&\sigma_{1}^{\nu} , \end{array} $$
(74b)
$$ \begin{array}{@{}rcl@{}} \left|\text{Re}\left( \mathcal{S}_{2}(\omega)\right)\right|&\le&{\Sigma}_{2}^{\nu} , \end{array} $$
(74c)

then the following combination of conditions on \((p,q) \in \mathbb {R}_{-}^{*} \times \mathbb {R}_{+}^{*}\) guarantees the convergence of the RR-based SWR algorithm, for all \(\omega \in \mathbb {R}\):

$$ -p \leq 2 \lambda \sigma_{1}^{\nu} \qquad \text{ and } \qquad q \geq \frac{1+(\gamma^{\nu})^{2}}{2}{\Sigma}_{2}^{\nu}. $$
(75)

Proof of Lemma 1

By reorganising (56), ρRR(p,q,ω) < 1 is equivalent to :

$$ \underbrace{pq\rho_{\text{DN}}^{2}-1 +2p\vartheta_{1}}_{\vartheta_{3}} +\underbrace{2q\text{Re}\left( \mathcal{S}_{2}\right)\left|\lambda\mathcal{S}_{1}\right|^{2}+\rho_{\text{DN}}^{2}}_{\vartheta_{4}} -q\underbrace{(p+2\text{Re}\left( \lambda\mathcal{S}_{1}\right))}_{\vartheta_{5}}<0 . $$
(76)

Since pq < 0 and per hypothesis (57), we have 𝜗3 < 0. Let us now exploit the conditions (1) to obtain (76). Using \(\rho _{DN}^{2}=|\lambda \mathcal {S}_{1}||\mathcal {S}_{2}|\), and recalling that \(\text {Re}(\mathcal {S}_{2})<0\) and q > 0, then 𝜗4 < 0 is equivalent to \(-2q|\text {Re}(\mathcal {S}_{2})|+|\mathcal {S}_{2}|^{2}<0\). (74c) implies \(|\mathcal {S}_{2}|^{2}\leq (1+(\gamma ^{\nu })^{2})|\text {Re}(\mathcal {S}_{2})|^{2}\). Then, \(q>\frac {{{{{\varGamma }}}_{0}^{2}}}{2}|\text {Re}(\mathcal {S}_{2})|\) with \({{{\varGamma }}}_{0}=\sqrt {1+(\gamma ^{\nu })^{2}}\) implies 𝜗4 < 0. In the same way, because q > 0 and \(\text {Re}(\mathcal {S}_{1})>0\), \(-p<2\lambda |\text {Re}(\mathcal {S}_{1})|\) is a sufficient condition for 𝜗5 > 0. □

Proof of Prop. 2

As previously mentioned, we rely on Lemma 1. In the case of constant viscosities, then, using (28) and extensive reformulation (not detailed, analoguous to (46)) yields:

$$ \left| \frac{ \textup{Im}(\mathcal{S}_{j}^{\text{cst}}(\omega)) }{ \textup{Re}(\mathcal{S}_{j}^{\text{cst}}(\omega)) } \right| = \mathcal{V}(y_{j}) := \left|\frac{\sinh(y_{j})-\sin(y_{j})}{\sinh(y_{j})+\sin(y_{j})} \right| \qquad \text{with } \qquad y_{j} = \sqrt{2\textup{Fo}_{j}(\omega)}. $$
(77)

Analogously to what is done in App. 3, it can then be proved that:

$$ \forall \omega \in \mathbb{R},~ \left| \frac{ \textup{Im}(\mathcal{S}_{j}^{\text{cst}}(\omega)) }{ \textup{Re}(\mathcal{S}_{j}^{\text{cst}}(\omega)) } \right| \leq \gamma^{\text{cst}} := \mathcal{V}(y^{\star}), $$
(78)

where y is the smallest positive root of \(\tanh (y) - \tan (y) = 0\). Hence, (74a) is satisfied. Moreover, still in App. 3, it is proved that \(\forall \omega \in \mathbb {R}\), \(|\mathcal {S}_{1}^{\text {cst}}(\omega ) | \geq |\mathcal {S}^{\text {sta,cst}}_{1,f=0} | = 1/|h_{1}|\) and that \(|\mathcal {S}_{2}^{\text {cst}}(\omega ) | \leq |\mathcal {S}^{\text {sta,cst}}_{2,f=0} | = |h_{2}|\). Hence, (74b) and (74c) can be obtained by defining:

$$ \sigma_{1}^{\text{cst}} := {\left( \sqrt{1+(\gamma^{\text{cst}})^{2}} |h_{1}| \right)}^{-1} \quad \text{ and } \quad {\Sigma}_{2}^{\text{cst}} := |h_{2}|. $$
(79)

Injecting (79) into (75), and defining \({{{\varGamma }}}_{0} := \sqrt {1+(\gamma ^{\text {cst}})^{2}}\), lead to (59). □□□

Proof Proof of Prop. 3

The proof is analogous to that of Prop. 2; hence, extensive detail is not provided. (78) still holds in that case with the same γcst. Moreover, since here \(\omega \in I_{\omega } \subsetneq \mathbb {R}\), we know that there exists \(({\Sigma }_{1}^{\text {cst}}, \sigma _{2}^{\text {cst}}) \in \mathbb {R}_{+}^{2}\) such that:

$$ \forall \omega \in I_{\omega}, \quad {\Sigma}_{1}^{\text{cst}} \geq \left|\text{Re}\left( \mathcal{S}_{1}(\omega)\right)\right| \quad \text{ and } \quad \sigma_{2}^{\text{cst}} \leq \left|\text{Re}\left( \mathcal{S}_{2}(\omega)\right)\right|. $$
(80)

Analogously to the proof of Lemma 1, and using (80), it can then be shown that (76), and thus SWR convergence, are satisfied as soon as:

$$ -p \geq \frac{1 + (\gamma^{\text{cst}})^{2}}{2} \lambda {\Sigma}_{1}^{\text{cst}} \qquad \text{ and } \qquad q \leq 2 \sigma_{2}^{\text{cst}}. $$
(81)

Moreover, since \(|\text {Re}(\mathcal {S}_{1}^{\text {cst}}(\omega ))|\le |\mathcal {S}_{1}^{\text {cst}}(\omega )|\) and \(\omega \mapsto |\mathcal {S}_{1}^{\text {cst}}(\omega )|\) is increasing (see App. 3), then \({\Sigma }_{1}^{\text {cst}} := |\mathcal {S}_{1}^{\text {cst}}(\omega _{\max \limits }) |\) is a suitable choice in (80). In the same way, since \({{{\varGamma }}}_{0}|\text {Re}(\mathcal {S}_{2}^{\text {cst}}(\omega ))|\ge |\mathcal {S}_{2}^{\text {cst}}(\omega )|\) and \(\omega \mapsto |\mathcal {S}_{2}^{\text {cst}}(\omega )|\) is decreasing, then \(\sigma _{2}^{\text {cst}} = |\mathcal {S}_{2}^{\text {cst}}(\omega _{\max \limits })| / {{{\varGamma }}}_{0}\) satisfies (80). Reinjecting this into (81) and relying on \({{{\varGamma }}}_{0} := \sqrt {1 + (\gamma ^{\text {cst}})^{2}}\) lead to (60), which, as previously mentioned, is a sufficient condition for SWR convergence with RR transmission conditions for ωIω. □

5.2 Discussion on Conjectures 1, 2 and 3

As soon as the viscosity profiles are no longer constant, the validity of (57) and (1) can only be numerically assessed. From Fig. 3 we can conjecture that \(\text {Re}(\mathcal {S}_{1}^{\text {aff}} ) \ge 1\) (resp. \(\text {Re}(\mathcal {S}_{1}^{\text {par}} ) \ge 1\)), which suggests that (57) and (74b) are satisfied on \(\mathcal {S}_{1}\). Since \(\mathcal {S}_{2}(\mu ,\text {Fo}) = - |h_{2}| / (|h_{1}| \mathcal {S}_{1}(\mu ,\text {Fo}))\), this would imply that (57) and (74c) are also satisfied for \(\mathcal {S}_{2}\). Figure 9 shows numerical values of \(\log _{10}(\gamma )\), with \(\gamma = | \textup {Im} (\mathcal {S}_{1}^{\nu }) | / | \textup {Re} (\mathcal {S}_{1}^{\nu })|\) in the cases where ν is affine and parabolic. It supports the idea that this ratio does not significantly depart from its constant viscosity value, which has been analytically found before. Then, using result from App. 3 and (4.3), we find a lower bound to \(|\text {Re}(\mathcal {S}^{\nu }_{1}(\omega ))|\):

$$ |\text{Re}(\mathcal{S}^{\nu}_{1}(\omega))|\ge|\mathcal{S}^{\nu}_{1}(\omega)|{{{\varGamma}}}_{0}^{-1}=|\mathcal{R}^{\nu}_{1}(\omega)|{{{\varGamma}}}_{0}^{-1}|\mathcal{S}^{\text{cst}}_{1}(\omega)|\ge\left|\mathcal{S}^{\text{cst, sta}}_{1,f=0}\right|{{{\varGamma}}}_{0}^{-1}. $$
(82)

Analogously, an upper bound to \(|\text {Re}(\mathcal {S}^{\nu }_{2}(\omega ))|\) can be found. Both bounds justify the first part of Conj. 1.

Fig. 9
figure 9

Left: \(\log _{10}\left [\left |\text {Im}\left (\mathcal {S}_{1}^{\text {aff}}\right )\right | \left /\left |\text {Re}\left (\mathcal {S}_{1}^{\text {aff}}\right )\right |\right ]\right .\); right: \(\log _{10}\left [ \left |\text {Im}\left (\mathcal {S}_{1}^{\text {par}}\right )\right | \left /\left |\text {Re}\left (\mathcal {S}_{1}^{\text {par}}\right )\right .\right |\right ]\) with respect to Fo and μ (both axes are in logarithmic scale). The grey area corresponds to values of (Fo,μ) for which the computations of \(\mathcal {S}_{1}^{aff}\) suffer from numerical instability (in the Bessel function evaluation)

If ωIω, then the bounds on \(|\text {Re}(\mathcal {S}^{\nu }_{1}(\omega ))|\) can be adjusted:

$$ \left|\mathcal{R}^{\nu}_{1}(\omega_{\max})\right| \left|\mathcal{S}^{\text{cst,sta}}_{1,f=0}\right|{{{\varGamma}}}_{0}^{-1} \le|\text{Re}(\mathcal{S}^{\nu}_{1}(\omega))| \le|\mathcal{R}^{\nu_{1},\text{sta}}_{1, f=0}| \left|\mathcal{S}^{\text{cst}}_{1}(\omega_{\max})\right|, $$
(83)

with analogous results for \(|\text {Re}(\mathcal {S}^{\nu }_{2}(\omega ))|\), hence Conj. 2.

Finally, assuming that \(\omega \mapsto |\text {Re}(\mathcal {S}^{\nu }_{1}(\omega ))|\) is increasing and \(\omega \mapsto |\text {Re}(\mathcal {S}^{\nu }_{2}(\omega ))|\) is decreasing, these new bounds justifying Conj. 3 can be established:

$$ \begin{array}{@{}rcl@{}} |\mathcal{S}^{\nu,sta}_{1,f=0}|{{{\varGamma}}}_{0}^{-1}\le |\text{Re}(\mathcal{S}^{\nu,sta}_{1,f=0})| \le|\text{Re}(\mathcal{S}^{\nu}_{1}(\omega))| \le|\text{Re}(\mathcal{S}^{\nu}_{1}(\omega_{\max}))| \le|\mathcal{S}^{\nu}_{1}(\omega_{\max})|, \end{array} $$
(84a)
$$ \begin{array}{@{}rcl@{}} |\mathcal{S}^{\nu}_{2}(\omega_{\max})|{{{\varGamma}}}_{0}^{-1} \le|\text{Re}(\mathcal{S}^{\nu}_{2}(\omega_{\max}))| \le |\text{Re}(\mathcal{S}^{\nu}_{2}(\omega))| \le|\text{Re}(\mathcal{S}^{\nu,sta}_{2,f=0})| \le |\mathcal{S}^{\nu,sta}_{2,f=0}|. \end{array} $$
(84b)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Thery, S., Pelletier, C., Lemarié, F. et al. Analysis of Schwarz waveform relaxation for the coupled Ekman boundary layer problem with continuously variable coefficients. Numer Algor 89, 1145–1181 (2022). https://doi.org/10.1007/s11075-021-01149-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11075-021-01149-y

Keywords

Navigation